precursor chemistry, thin film deposition, mechanistic studies

207
Rational Development of Precursors for MOCVD of TiO 2 : Precursor Chemistry, Thin Film Deposition, Mechanistic Studies Raghunandan Krishna Bhakta, M.Sc. 2005

Upload: others

Post on 25-Jan-2022

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

Rational Development of Precursors for MOCVD of TiO2:

Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

Raghunandan Krishna Bhakta, M.Sc.

2005

Page 2: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

A dissertation submitted for the degree of Dr. rer. nat. (Doctor rerum naturalium)

in the Faculty of Chemistry at Ruhr- University Bochum,

Germany

Raghunandan Krishna Bhakta, M. Sc.

2005

Page 3: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

Rational Development of Precursors for MOCVD of TiO2:

Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

Dissertation Zur Erlangung des Grades eines

Doktors der Naturwissenschaften

der Fakultät für Chemie

an der Ruhr-Universität Bochum,

Deutschland

vorgelegt von

Raghunandan Krishna Bhakta, M. Sc.

2005

Page 4: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

This dissertation is based on the experimental work carried out during the period from

April 2001 to September 2004 at the chair of Inorganic Chemistry II, Ruhr University

Bochum, Germany. Herewith I declare that the following work has been carried out

independently by me and all the sources of help and services used during this work have

been reported in the acknowledgement section.

Research Supervisor: Prof. Dr. Roland A. Fischer

Co-supervisor: Jun. Prof. Dr. Anjana Devi

2nd examiner: Prof. Dr. M. Epple

3rd examiner: Prof. Dr. W. S. Sheldrick

Day of examination 02.02.2005

Page 5: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

Acknowledgements

I am indebted to many people who contributed in several ways to this work, and

supported me with their cooperation and timely help.

In particular, I wish to express my sincere gratitude to my research supervisor Prof. Dr.

Roland A. Fischer, for providing me an opportunity to work in his research group, for

introducing to an interesting research theme for the present work and for closely

following the work, for his support throughout this work with fruitful suggestions and for

reviewing this thesis.

I equally express my sincere gratitude to my co-supervisor, Jun. Prof. Dr. Anjana Devi,

who directly supervised this work. I thank her for unique personal support in every aspect

of the experimental work, from the precursor synthesis to the CVD, for all the stimulating

ideas, all the deep discussions, for providing support in preparation of various

manuscripts for publications, presentations, posters and for the time she spent in

reviewing this thesis.

This work forms a part of the project funded by Deutsche Forschungs Gemeinschaft

(DFG). I gratefully acknowledge DFG, who funded the project under CVD- SPP-1119.

I thank all my colleagues Andreas Kempter, Andrian Milanov, Jun.-Prof. Dr. Anjana

Devi, Arne Baunemann, Beatrice Buchin, Dr. Christian Gemel, Daniel Rische, Eliza

Gemel, Eva Maile, Eun Jeong Kim, Felicitas Schröder, Dr. Frank Hipler, Dr. Harish

Parala, Heike Kampschulte, Jayaprakash Khanderi, Manuela Winter, Marie-Katrin

Schoeter, Dr. Maxim Tafipolski, Mirza Cokoja, Dr. Rochus Schmid, Sabine Bendix,

Sabine Masukowitz, Stephan Hermes, Stephan Spöllmann, Thomas Kadenbach, Tobias

Steinke, Todor Hikov, Urmila Patil, Ursula Fischer, Ursula Herrmann, Dr. Younsoo Kim,

Dr. Wenhua Zhang for friendly and encouraging work atmosphere, help and cooperation.

I also thank and cherish the support and help received from my former colleagues Dr.

Andreas Wohlfart, Dr. Dana Weiß, Dr. Jurij Weiß, Prof. Dr. Jens Müller, Dr. Julia

Page 6: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

Hambrock, Dr. Lianhai Lu, Dr. Nicola Oberbeckmann, Dr. Oliver Segnitz, Dr. Oliver

Stark, Dr. Pia Wennek, Dr. Ralf Becker, and Dr. Ulrike Weckenmann.

I gratefully acknowledge the following colleagues for their help and assistance as noted:

Stephan Hermes, Jutta Schäfer (GC, MS); Karin Bartholomäus (elemental analysis); Dr.

Christian Gemel (hydrolysis studies, scientific discussions); Manuela Winter, Dr. Iris M.

Müller, Dr. Klaus Merz (single crystal X-ray diffraction); Urmila Patil, Dr. Harish Parala

(TGA/DTA); Dr.Andreas Wohlfart, Dr. Harish Parala, Eva Maile, (XRD analysis); Dr.

Hans-Werner Becker (RBS analyses); Dr. Rolf Neuser (SEM); Dr. Frank Hipler (XPS

analysis); Prof. Dr. Jens Müller, Dr. Frank Hipler, Dr. Holger F. Bettinger, Eliza Gemel,

Sabine Bendix (helping in matrix isolation experiments).

My sincere thanks to Dr. Peter Ehrhart, Dr. Reji Thomas, Dr. Stephan Regnery, Prof. Dr.

Rainer Waser (Forschungszentrum Jülich) for their help in thin film depositions using an

industrial tool reactor, for various analyses of thin films at their facility and for a fruitful

collaboration.

I am extremely thankful to Sabine Masukowitz, and Heike Kampschulte for helping me

with administrative and official procedures. I gratefully acknowledge the help extended

by library personnel, mechanical workshop personnel, glass blowers, electrical workshop

personnel and chemical store personnel at the faculty of chemistry, which was

instrumental during my research efforts.

I would like to thank my former employer Rallis India Ltd. for hosting my career for one

year at their facilities. My sincere thanks are due to Dr. G. Ananda Rao, CEDT, Indian

Institute of Science, Bangalore, for his encouragement and support for my research

activities.

I am thankful to my friends, Aravinda, Eva, Ganesh, Janardana, JP, Keshav, Koushik,

Sasi, Shakila, Stephan, Suresh, Uday, Urmila, Vijay, Yathish, and my well-wishers

whose emotional support, encouragement and inspiration were instrumental during all

these years.

Page 7: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

I dedicate this work to my parents M. Krishna Bhakta and Radha Bhakta, my wife Sarita,

my sisters, Sucheta, Suman, and Sahana and their families.

Page 8: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

List of abbreviations and Acronyms AFM Atomic Force Microscopy

CVD Chemical Vapor Deposition

CFD Computational Fluid Dynamics

DFT Density Functional Theory

DRAM Dynamic Random Access Memory

DTA Differential Thermal Analysis

EDX Energy Dispersive X-ray

IR Infra Red

LPCVD Low Pressure Chemical Vapor Deposition

MBE Molecular Beam Epitaxy

Meaoac Methylacetoacetate

MIS Metal-Insulator-Semiconductor

MO Metal Organic

MOCVD Metalorganic Chemical Vapor Deposition

MS Mass Spectrometry

NMR Nuclear Magnetic Resonance

OEt Ethoxy group

OPri Isopropoxy group

PACVD Plasma Assisted Chemical Vapor Deposition

PECVD Plasma Enhanced Chemical Vapor Deposition

RT Room Temperature

SEM Scanning Electron Microscopy

Tbaoac Tertiary Butyl Acetoacetate

TGA Thermogravimetric analysis

TTIP Titanium Tetraisopropoxide

UV Ultraviolet

XPS X-ray Photoelectron Spectroscopy

XRD X-ray Diffraction

XRF X-ray fluorescence

Page 9: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

Contents

Chapter 1

1.0 Introduction 1

1.1 Titanium dioxide 2

1.2 An overview of various thin film deposition technologies 5

1.3 An overview of CVD 7

1.4 CVD kinetics 11

1.5 Growth rate 14

1.6 Precursor for CVD 16

1.7 The liquid injection CVD 17

1.8 Gas-phase chemical species measurements 19

1.8.1 Mass spectroscopy 20

1.8.2 Gas chromatography 21

1.8.3 IR absorption spectroscopy 22

1.8.4 Matrix isolation FTIR spectroscopy 23

1.9 MOCVD of TiO2 thin films 25

1.10 Scope of the present work 26

1.11 References 30

Chapter 2

2.1 Introduction 34

2.2 Experimental section 36

2.3 Results and Discussion 39

2.3.1 Objectives of ligand design 39

2.3.2 Precursor synthesis and properties 43

2.3.3 NMR studies 48

2.3.4 Single crystal X-ray diffraction analysis 49

2.4 References 56

Chapter 3

3.1 Introduction 58

3.2 Experimental section 60

Page 10: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

3.2.1 Experiments using home built CVD reactor 60

3.2.2 Experiments using industrial tool CVD reactor 63

3.3 Results and Discussion 65

3.3.1 Deposition of TiO2 thin films using home built CVD reactor 65

3.3.2 Deposition parameters 66

3.3.3 Crystal structure of the films 66

3.3.4 Effect of substrate temperature on growth rate 68

3.3.5 Film composition 70

3.3.6 Film microstructure 72

3.4 Deposition of TiO2 thin films using liquid injection industrial tool 75

3.4.1 Susceptor temperature dependent growth rate, surface roughness

structural and electrical properties 75

3.4.2 Surface roughness 76

3.4.3 Crystal structure 78

3.4.4 Effect of post deposition annealing on the structure and morphology 80

3.4.5 Electrical properties 82

3.5 Deposition of SrTiO3 thin films 85

3.6 References 90

Chapter 4

4.1 Introduction 93

4.2 Experimental section 99

4.2.1 Description of matrix isolation apparatus 101

4.2.2 Photolysis experiments using ultraviolet source 103

4.2.3 Preparation of gaseous mixture of argon and iso-propanol 103

4.3 Results and Discussion 104

4.3.1 Fundamental aspects of the matrix isolation technique 104

4.3.2 Spectroscopic methods 106

4.3.3 Vibrational spectroscopy in the infrared region 106

4.3.4 Sample preparation and concerns 107

4.3.5 In situ generation 108

4.3.6 Chemistry with matrices 109

Page 11: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

4.3.7 Rigidity and mobility of matrix material 109

4.3.8 Ultraviolet photolysis of the decomposition products in the matrix 109

4.3.9 The FTIR spectra of [Ti(OPri)4] and iso-propanol 110

4.3.10 Thermolysis of the TTIP 114

4.3.11 Matrix isolation studies on mixed alkoxide complexes of titanium 121

4.3.12 The FTIR spectrum of the ligand tert. butylacetoacetate 122

4.3.13 Thermolysis of tert. butylacetoacetate 123

4.3.14 Thermolysis of [Ti(OPri)2(tbaoac)2] 127

4.3.15 The FTIR spectrum of the ligand 2,2,6,6,-teramethyl-3,5

-heptane dione ligand (Hthd) 136

4.3.16 Thermolysis of the ligand Hthd 137

4.3.17 Thermolysis of [Ti(OPri)2(thd)2] 139

4.4 References 144

Chapter 5

5.1 Introduction 146

5.1.1 Volatility 147

5.1.2 Long term stability 148

5.1.3 Hydrolytic stability 150

5.1.4 Thermal analyses 150

5.2 Experimental section 152

5.3 Results and Discussion 153

5.3.1 Sublimation studies 160

5.3.2 Shelf life 163

5.3.3 Hydrolysis studies 165

5.4 References 166

Page 12: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 1 -

Chapter 1

Introduction

Sustainable technological development is strongly dependent on new materials with

particular mechanical, chemical, electrical, magnetic, or optical properties. In order to

address this challenge, interdisciplinary research bridging science and technology to

develop new materials, to impart new functional properties, and to provide new

processing methods for the formation of useful objects is under intense focus. It is hard to

achieve the functionality of materials in a macroscopic form employing basic chemical

compounds off the shelf. In such a situation the role of a chemist as a manipulator of the

compounds at a molecular level to design and develop new materials with desired

properties is gaining critical importance.

Of all the functional materials, those under intense investigation can be broadly classified

in to following different types: namely, semiconductors, metals, alloys, ceramics, and

biomaterials.[1] Ceramics and semiconductors are often prepared as surface layers of

different composition from that of the bulk and some times also to impart a specific

functionality to the surface or to act as a protective layer for the bulk material. In

particular, simple inorganic metal oxides, silicates, nitride compounds have been used in

a variety of applications that take advantage of their optical properties, chemical

resistance, high thermal stability, and resistance to environmental degradation. In many

of these applications a ceramic is also coated onto other material, such as plastics,

semiconductors, or metals thus forming a heterogeneous interface. With the general trend

to miniaturize devices, the interest in the interface layers between surface and bulk

material having vast technological significance is gaining momentum.[2]

Within the class of inorganic materials, oxides display perhaps the most diverse range of

functionality. The most commonly observed electrical property observed in oxides is

insulating, and the number of dielectric and ferroelectric oxides that have been realized as

thin films is quite large.[3] The oxide ceramics are the most used materials for technical

applications, particularly in electronic and structural areas. The purity of these materials

is extremely important, especially for electronic and high temperature applications.

Page 13: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 2 -

Among the oxide class MO2, Titanium dioxide (TiO2) is one of the most widely studied

oxide over the years and the following section gives a brief summary of its properties and

applications.

1.1 Titanium dioxide (TiO2)

Titanium dioxide is perhaps one of the most studied oxides both in bulk and thin film

forms. Pure titanium dioxide is extracted from ilmenite or leuxocene ores. Rutile beach

sand is another source for pure rutile form of TiO2. Titanium dioxide mainly exists in

three different allotropic forms, each with distinguishing electronic, optical, and

structural properties. The crystalline phases of TiO2 are rutile (tetragonal, a = 4.5845 ? , c

= 2.9533 Å), anatase (tetragonal, a = 3.7842 Å, c = 9.5146 Å), and brookite

(orthorhombic, a = 9.184 Å, b = 5.447 Å, c = 5.145 Å).[4]

Other forms exist as well, for example, cotunnite TiO2 has been synthesized under high

pressure and is reported to be the hardest known oxide material.[5] However, mostly rutile

Table 1.1: Important properties of three different phases of TiO2

Phase Refractive

index

Density

(g.cm-3) Crystal structure

Dielectric

constant

Band gap

(eV)

Brookite 2.58 4.23

Orthorhombic

a = 9.184 Å

b = 5.447 Å

c = 5.145 Å

14 2.2

Anatase 2.49 3.79

Tetragonal

a = 3.7842 Å,

c = 9.5146 Å

31 3.2

Rutile 2.903 4.26

Tetragonal

a = 4.5845 ? ,

c = 2.9533 Å

86 3

Page 14: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 3 -

and anatase phases play important roles in many applications of TiO2. Table 1.1

summarizes some important properties of three different phases of TiO2.

Each titanium ion in anatase is coordinated by six oxygens, and each oxygen ion by three

titaniums. The TiO6 octahedra share edges with four adjacent octahedra – refer to the

central octahedron in the Fig. 1.1 which shows this arrangement. The rutile structure

contains a tetragonal close packing. The structure consists of TiO6 octahedra, with the O

atoms shared by neighboring Ti atoms. Each Ti atom is surrounded by six O atoms and

each O atom is surrounded by three Ti atoms. Brookite phase contains cubic close

packing of the oxygen atoms. Titanium atoms occupy the vacancies of the oxygen-

octahedra in such a way that one octahedron shares three edges with neighboring TiO6

octahedra. Fig. 1.1 shows the atomic arrangement of three TiO2 structures.

Brookite Anatase Rutile

Fig. 1.1: The crystalline phases of TiO2: Brookite (Orthorhombic), Anatase (tetragonal)

and Rutile (tetragonal).

Page 15: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 4 -

Rutile represents the stable phase at high temperatures and is the easiest to realize as

phase pure crystals or as thin films. As the more stable phase, rutile is easily realized as

an epitaxial film using most film growth techniques.[6-8] Anatase is only metastable in

bulk, but despite this fact, it has been realized as an epitaxial film using different

techniques. Epitaxial anatase films can be grown between 500 and 700 °C over a wide

range of processing pressures.[9-11] These phases possess different electro-optical

properties. For instance, although uniform anatase films can be grown with a quite high

permittivity value (70),[12] the anatase is still characterized by a lower dielectric

constant,[13] higher leakage current,[14] and lower breakdown field strength,[13] than rutile.

At the same time, pure anatase is optically less absorbent than rutile.[15] The crystal

structures of three different phases are shown in Fig. 1.1.

Though the largest commercial application of TiO2 remains as an additive to pigments for

imparting white color to paints, it finds numerous other applications as well. Even in

mildly reducing atmospheres, TiO2 tends to lose oxygen and become sub stoichiometric.

This form of TiO2 acts as a semiconductor and its electrical resistivity can be correlated

to the oxygen content of the atmosphere to which it is exposed.

Hence TiO2 is often used to sense the amount of oxygen (or reducing species) present in

an atmosphere. Over the past four decades remarkable progress has been made in to the

research related to superconductors, photovoltaic and silicon based semiconductors.

Table 1.2: Proposed materials for the high-k applications and their dielectric constants.[16]

Material Bandgap

(eV)

Relative dielectric

constant

SiO2 9 3.9

Al2O3 8.8 9.5-12

ZrO2 5.7-5.8 12-16

HfO2 4.5-6 16-30

La2O3 ~ 6 20.8

Ta2O5 4.4 25

TiO2 3.05 80

Page 16: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 5 -

Titanium dioxide has been under focus because of its higher dielectric constant which has

been extensively utilized in semiconductor and related technologies. The dielectric

constants of different materials under consideration are given in table 1.2. The

opportunities afforded by titanium dioxide based materials in many applications have

driven significant efforts exploring their formation as thin films. Titanium dioxide thin

films are widely used in construction applications because of their chemical durability

and mechanical resistance.[17]

It is well known that thin films of TiO2 can modify the optical and electrical properties of

glass. TiO2 has been successfully used for optical wave guides and antireflection (AR)

coatings in silicon solar cells, due to its high refractive index and high transparency in the

visible and near IR range.[18] In addition, TiO2 forms one of the main components of high-

? and ferroelectric materials such as SrTiO3, (Ba,Sr)TiO3 and, Pb(Zr,Ti)O3, which are

considered for the fabrication of ultra high density DRAMs, non-volatile computer

memories, sensors, IR detectors etc.[19, 20] Since the research efforts to develop a method

for depositing TiO2 thin films spans over five decades and detailed descriptions are

unwarranted for the present study, only a selected number of notable techniques that have

been used for TiO2 thin film depositions are noted in following section.

1.2 An overview of various thin film deposition technologies

The technology of the film deposition has advanced dramatically during the past 30

years. This advancement was driven primarily by the need for new products and devices

in the electronic and optical industries. The rapid progress in solid-state electronic

devices would not have been possible without the development of new type of film

deposition processes, improved film characteristics and superior film qualities. Thin films

are applied to surfaces using vapor deposition to many types of work pieces (substrates)

materials. There are two major vapor deposition process categories, they are: PVD

(Physical Vapor Deposition) and CVD (Chemical Vapor Deposition). Each of these two

processes does essentially same thing that bring vapor deposition. Basic difference

between PVD and CVD is how the material is transported in vapor form to the substrate

where it is deposited.

Page 17: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 6 -

Deposition techniques

Physical methods

Sputtering

Co-evaporation

Ion implantation

Molecular beam epitaxy

Pulsed laser deposition

Chemical methods

Solution deposition

Sol-gel

Spin coating

Spray pyrolysis

Chemical vapor deposition

Fig. 1.2: Broad classification of different processes used for the deposition of TiO2

thin films.

In PVD, the particles to be deposited are carried by a physical means to the substrate,

whereas in CVD, the particles are carried through a chemical reaction. PVD involves the

atomic vapor deposition of materials onto a substrate, through physical transport. The

PVD technologies available today are electron beam evaporation, arc vapor deposition,

sputtering, molecular beam epitaxy and pulsed laser deposition.

The thickness of a surface layer can vary from less than a µm in case of ion implantation

or chemical vapor deposition to few mm in case of electrochemical methods.

Correspondingly, process temperatures range from ambient to several hundred degrees

Celsius depending on the process and material to be deposited. The choice of a process is

therefore limited to a range of substrate materials depending on their temperature

sensitivity and the thickness requirement. In case of thin film requirements, the vapor

deposition methods are considered as a better choice over other deposition processes. A

plethora of deposition processes have been utilized for the growth of titanium dioxide

thin films and still new techniques are being introduced. The different deposition

techniques are broadly classified in to two types (Fig. 1.2).

Page 18: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 7 -

Physical vapor deposition or PVD consists of atom/ion beam techniques. A wide range of

techniques is available, wherein the main goal is to attract metal ions to the surface of a

substrate under the influence of an electrical bias and in the presence of a reactive gas

under low pressures.[21] These are sometimes assisted by sub processes involving

generation and action of a plasma. Broadly the generation of vapors is done in four

different ways. They are evaporation, sputtering, ion-plating and reactive ion-plating.

PVD is a line–of–sight process.[22] In case of highly uniform deposition requirements

over large areas having non planar surfaces, vias, trenches, PVD suffers from this draw

back.[20] As this work is based on the research in to the field of CVD, the details

regarding PVD are kept at a minimum.

Chemical vapor deposition or CVD is a common term used for a set of processes that

involve depositing a solid material from a gaseous phase. There are distinct differences

between CVD and PVD although some similarities do exist. In a PVD process the

precursors are solid, with the material to be deposited being vaporized from a solid target

and deposited onto the substrate. But in case of a CVD process, the precursor for the

material can be a solid, liquid or a gas and/or a combination of these.[20]

Metal organic chemical vapor deposition (MOCVD) which uses vapor phases of different

metal organic compounds as precursors and is the most versatile and promising

deposition technique. It offers a possibility for large area deposition with good

composition control, high degree of uniformity and excellent conformal step coverage

over non planar structures.[23] During the present study, MOCVD was used as a

deposition process for TiO2 thin films and hence will be discussed in detail in the

following sections.

1.3 An overview of Chemical Vapor Deposition

Methods of film formation by purely chemical processes in the gas or vapor phases

include chemical vapor deposition (CVD) and thermal oxidation. CVD is a materials

synthesis process whereby constituents of the vapor phase react chemically near or on a

substrate surface to form a solid product. The deposition technology has become one of

Page 19: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 8 -

the most important means of creating thin films and coatings of a very large variety of

materials essential to advanced technology particularly solid state electronics where some

of the most sophisticated purity and composition requirements must be met. The main

feature of CVD is its versatility for synthesizing both simple and complex compounds

with relative ease at generally low temperatures.

Both chemical composition and physical structure can be tailored by control of the

reaction chemistry and deposition conditions. Fundamental principles of CVD encompass

an interdisciplinary range of gas phase reaction chemistry, thermodynamics, kinetics,

transport mechanisms, film growth phenomena and reactor engineering.

Chemical reaction types basic to CVD include pyrolysis (thermal decomposition),

oxidation, reduction, hydrolysis, nitride and carbide formation, synthesis reactions,

disproportionation and chemical transport. A sequence of several reaction types may be

involved in more complex situations to create a particular end product. Deposition

variables such as temperature, pressure, input concentrations, gas flow rates and reactor

geometry and operating principle determine the deposition rate and the properties of the

film deposit.

The use of modern chemical vapor deposition technique can be dated back to the year

1880 where the technique was used to strengthen the filaments in incandescent lamps

using carbon or metal depositions on them.[24] The CVD related activities remained more

or less passive for the first half of the last century, but for the last 45 years the growth has

been tremendous. The term chemical vapor deposition was first used in 1960.[25] The

increasing requirements of semiconductor industry have been the driving force in the

development of CVD techniques and very importantly, the intensified efforts to

understand the fundamentals of the CVD processes.

Most CVD processes are chosen to be heterogeneous. That is, they take place at the

substrate surface rather than in the gas phase. Undesirable homogenous reactions in the

gas phase nucleate particles that may form powdery deposits and lead to particle

contamination instead of clean and uniform coatings. The reaction feasibility (other than

reaction rate) of a CVD process under specified conditions can be predicted by

thermodynamic calculations provided reliable thermodynamic data (especially the free

energy of formation) are available. Kinetics control and the rate of reactions depend on

Page 20: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 9 -

temperature and factors such as substrate orientation. Considerations related to heat, mass

and momentum transport phenomenon are especially important in designing CVD

reactors of maximum efficiency. Since the important physical properties of a given

material are critically influenced by the structure (such as crystallinity) control of the

factors governing the nucleation and structure of growing film is necessary.

CVD has become an important process technology in several industrial fields. As

mentioned earlier, applications in solid-state microelectronics are of prime importance.

Thin films of insulators, dielectrics (oxides, silicates, nitrides), elemental and compound

semiconductors (silicon, gallium arsenide etc.) and conductors (tungsten, molybdenum,

aluminum, refractory metal silicides) are extensively utilized in the formation of solid

state devices. Hard and wear resistance coatings of materials such as boron, diamond-like

carbon, borides, carbides and nitrides, are used for metal protection in metallurgical

applications. Numerous other types of materials, including vitreous graphite and

refractory metals, have been deposited mainly in bulk form or as thick coatings. Many of

these CVD reactions have long been used for coating of substrates at reduced pressure,

often at high temperatures.

Reactors

The reactor system (comparing the reaction chamber and all associated equipment) for

carrying out CVD process must provide several basic functions common to all types of

systems. It must allow transport of the reactant and diluent gasses to the reaction site,

provide activation energy to the reactants (heat, radiation, plasma), maintain a specific

system pressure and temperature allow the chemical processes for film deposition and

proceed optimally, and remove the by-product gasses and vapors. These functions must

be implemented with adequate control, maximal effectiveness and complete safety.

The most sophisticated CVD reactors are those used for the deposition of electronic

materials. Low temperature (below 600 °C) production reactors for normal or

atmospheric CVD (APCVD) include rotary vertical flow reactors and continuous in line

conveyorized reactors with various gas distribution features. They are used primarily for

depositing oxides and binary and ternary silicate glass coatings for solid-state devices.

Page 21: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 10 -

Reactors for mid-temperature (600 °C – 900 °C) and high temperature (900 °C – 1300

°C) operation are either hot-wall or cold wall types constructed of fused quartz.

Hot wall reactors, usually tubular in shape are used for exothermic processes and have

the advantage of close temperature control. They have been used for synthesizing

complex layer structures of compound semiconductors for microelectronic devices. A

disadvantage is that deposition occurs everywhere on the part as well as on the walls of

the reactors, which requires periodic cleaning or the use of disposable lines. Cold wall

reactors, usually bell-jar shaped, are used for endothermic processes, such as the

deposition of silicon from the halides or the hydrides. Heating is accomplished by RF

induction or by high intensity radiation lamps. Substrate susceptors of silicon carbide

coated graphite slabs are used for RF heated systems.

Reactors operating at low pressures (typically 0.1 – 10 torr) for low pressure CVD

(LPCVD) is the low, mid or high temperature range are resistance heated hot-wall

reactors of tubular, bell-jar, or close spaced design. In the horizontal tubular design, the

substrate slices (silicon device wafers) stand up in a carrier sled and gas flow is

horizontal. The reduced operating pressure increases the mean free path of the reactant

molecules, which allows a closely spaced wafer stacking. The very high packing density

achieved (typically 100-200 wafers per tube) allows a greatly increased throughput,

hence substantially lower product cost. In the vertical bell-jar design, the gas is

distributed over the stand-up wafers, hence there is much gas depletion and generation of

few particles, but the wafer load is smaller (50 to 100 wafers per chamber). Finally, the

close spaced design developed most recently processes each wafer in its own separate,

close space chamber with the gas flowing across the wafer surface to achieve maximum

uniformity. In LPCVD, no carrier gasses are required, particle contamination is reduced

and film uniformity and conformality are better than in conventional APCVD reactor

systems. It is for these reasons that low-pressure CVD is widely used in the high cost-

competitive semiconductor industry for depositing films of insulators, amorphous and

polycrystalline silicon, refractory metals and silicides.

Current developments in CVD focus on low temperature forms of CVD, such as

MOCVD, plasma-CVD, and photo-CVD and atomic layer deposition ALD.[26] These are

assisted processes and are employed widely in the fields of semiconductor industry and

Page 22: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 11 -

microelectronics, as well as in hard coatings corrosion and wear resistance applications as

lower deposition temperatures now permit the use of a broader spectrum of substrates. In

addition, the need for the better understanding of the process in order to control the

process in a better way has led to increased activities to monitor the deposition

parameters through in situ observations.[23]

CVD - Chemical vapor deposition is a materials-synthesis process in which one or more

vapor-phase chemical components are transported into a reaction chamber. They are

activated thermally or by several other methods, such as plasma or laser stimulation in

the vicinity of the substrate to chemically react at the substrate surface. The reaction

taking place at the vicinity of substrate surface lead to a solid film. The unreacted vapors

and volatile byproducts of decomposition reaction are transported away from the surface.

The CVD process belongs to those vapor transfer processes that are atomistic in nature.

The main features of CVD are its versatility for synthesizing both simple and complex

compounds with relative ease. Both chemical composition and physical structure can be

tailored by control of the reaction chemistry and deposition conditions.[23]

1.4 CVD Kinetics

In CVD, film growth takes place following complex set of chemical reactions taking

place sequentially and continuously. In order to get optimal quality films one needs to

have a better understanding of the theory of CVD. A theoretical analysis is in most cases,

an essential step, which if properly carried out should predict any one for the following

1) Chemistry of the reaction (intermediate steps, by products)

2) Reaction mechanism

3) Composition of the deposit (i.e. stoichiometry)

4) Structure of deposit (i.e. the geometric arrangement of its atoms).

This analysis may then provide guidance for an experimental program and considerably

reduce its scope and save a great deal of time and effort. Such an analysis requires a clear

understanding of the CVD process and a series of several fundamental considerations in

the disciplines of thermodynamics, kinetics, and chemistry is in order. Critical to CVD

Page 23: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 12 -

Impinging reactant atoms

Nucleation Surface reactions

Surface migration

Condensation

Desorption of volatile reaction products Gas phase reactions

theory are chemical kinetics, fluid mechanics, chemical engineering principles as well as

an understanding of growth mechanisms.

Any vapor deposition technique is based on the principles of mass transfer from one

source to another. Macroscopically following three fundamental steps play critical role in

the growth of a film.

• Transfer of the precursor to the gas phase

• Transport of gas phase to the substrate

• Deposition onto substrate and film growth.

These three steps are either separated in space and time or superimpose with each other,

depending on process requirements.[23]

The individual process steps are displayed in Fig. 1.3. There are efforts to simulate the

situation in CVD reactors from predictions of thermodynamical equilibrium calculations.

But the deposition reaction is almost a heterogeneous reaction. The sequence of events in

the usual heterogeneous process can be described as follows.

• mass transport of the gas phase in to the deposition zone;

Fig. 1.3: The representation of the individual process steps involved in chemical vapor

deposition

Page 24: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 13 -

• gas phase reactions leading to the formation of film precursors and byproducts;

• mass transport of film precursors on the growth surface;

• surface diffusion of film precursors to growth sites;

• incorporation of film constituents into growing film;

• desorption of byproducts of the surface reactions; and

• mass transport of byproducts in the bulk gas flow region away from the

deposition zone towards the reactor exhaust.

The individual steps take place simultaneously and can not be independently followed.

The slowest of these process steps determines the deposition rate. The close control of

these steps through the vital CVD parameters - substrate temperature, reactor pressure

gas flow rates and gas-phase composition (precursor concentration) determines the film

growth rate and enables the deposition of a wide variety of coatings. The strict quality

requirements (e.g., thickness and compositional uniformity, crystallinity, electrical

properties) imposed on CVD deposits imply the need for stringent control of the CVD

processes. In particular following parameters are controlled.

1. Temperature in the reactor (one or more zones)

2. The quantities and compositions of all gasses or vapors entering the reactor.

3. The time sequencing of variables mentioned in (1) and (2)

4. The pressure in the case of low pressure CVD

Most chemical reactions in CVD are thermodynamically endothermic and/or have a

kinetic energy of activation associated with them. Generally this is an advantage since the

reactions can be controlled by regulating the energy input. However, it does mean that

energy has to be supplied to the reacting system, and traditionally CVD processes have

been initiated and controlled by the input of thermal energy to the substrate. Based on the

energy input, three different methods of energy input in CVD processes are practiced.

a) Thermal CVD

Thermal CVD requires high temperature, generally from 800-2000 °C, which can be

generated by resistance heating, high frequency induction, radiant heating etc. The choice

of heating methods depends largely on factors such as the type of deposition process and

the shape, size, and composition of the substrate material as well as economics.

b) Plasma CVD

Page 25: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 14 -

In plasma CVD, also known as plasma enhanced CVD (PECVD) or plasma assisted CVD

(PACVD), the reaction is activated by plasma and the deposition temperature is

substantially lower. Plasmas are extremely complex chemical soups and deposition

characteristics can depend markedly on system variables such as gas pressure, flow rate,

RF power and frequency, and reactor geometry and substrate temperature.

c) Laser and photo CVD

Two methods based on photo activation have recently been developed: A laser produces

a coherent monochromatic high energy beam of photons, which can be used effectively to

activate a CVD reaction. Laser CVD occurs as a result of thermal energy from the laser

coming in contact with and heating an absorbing substrate. The wavelength of the laser is

such that little or no energy is absorbed by the gas molecules.

In photo CVD, the chemical reaction is activated by the action of photons, specifically

UV radiation; which have sufficient energy to break the chemical bonds in the reactant

molecules. No heat is required and the deposition may occur essentially at room

temperature. More over, there is no constraint on the type of substrate which can be

opaque, absorbent or transparent. A limitation of photo CVD is the slow rate of

deposition which has so far restricted its applications.

1.5 Growth rate and modeling principles for CVD

The growth rate is primarily determined by the substrate temperature, reactor pressure

and gas-phase composition. When CVD growth rates are plotted against reciprocal of

deposition temperature, the following regimes can be distinguished.[23]

• The growth rate is controlled by surface reaction kinetics at the substrate at lower

temperatures.

• At intermediate temperatures, mass-transfer-kinetics through the stagnant

boundary layer limits the growth rate

• Reduced growth rate at high temperatures due to parasitic reactions, such as

increased decomposition rate on hot reactor walls or through pronounced gas-

phase reactions

The growth rate is primarily determined by the substrate temperature, reactor

pressure, and gas-phase composition. At low temperatures, the growth rate is limited

Page 26: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 15 -

by chemical kinetics and increases exponentially with temperature according to the

Arrhenius expression,

G = A exp (EA/RT)

Where, G = growth rate

EA = the apparent activation energy of the rate determining step,

R = universal gas constant

T = temperature in K.

Since the rate is limited by chemical kinetics, uniform film thickness can be achieved by

minimizing temperature variations. This is the regime desired by hot wall low pressure

CVD reactors. The growth rate is nearly independent of temperature in the intermediate

temperature regime where mass transport to the surface controls the rate. This

temperature independence is particularly advantageous in cold wall reactors, where it is

often difficult to obtain completely uniform substrate heating. The growth rate may

decrease at high temperatures because of an increased desorption rate and depletion of

reactants on reactor walls. The emergence of an alternative reaction pathway may also

lead to a decline in the temperature-dependent growth rate.

Because of the complexity of transport phenomena and chemical reactions underlying

CVD, models of the process are required to identify rate controlling steps and to link

1/T (K)

Log growth rate (µm/min)

Fig. 1.4: Typical growth rate dependence on substrate temperature

Page 27: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 16 -

growth uniformity performance to process conditions precursor chemistry. In addition,

accurate models play a significant role in the design of reactors capable of producing high

deposition rates and composition uniformity over large areas. Thermodynamic analysis

has been the traditional approach to CVD process modeling because of relative ease of

determining the system state by equilibrium calculations relative to experiments or

detailed kinetic models. Equilibrium composition at constant temperature and pressure is

generally computed in two ways; by direct minimization of the Gibbs free energy of the

system, subject to elemental abundance and mole number non-negativity constraints; or

by transforming the species mole number variables into a new set of reaction variables

and then minimizing the Gibbs free energy in terms of these new variables.[23]

In the solution of CVD reactor models it is often possible to separate the flow and energy

solution from the mass transfer analysis since CVD reactants often are used in low

concentration in some inert carrier gases. The continuum approach is based upon

Knudsen diffusion concepts and has the advantage of being simple. The Monte-Carlo

approach is versatile, allowing the inclusion of surface transport and gas-phase reactions

besides the surface reactions with variable sticking coefficients.[23]

Because of the low pressures used in LPCVD systems, the fluid flow can be in either the

usual continuum regime or in the transition regime, depending on the relative magnitude

of the mean free path of the reactant molecules and the characteristic dimension of the

reactors, as reflected by the Knudsen number, Kn = ? / d, where ? is the mean free path

of gas molecules and d is the characteristic reactor dimension. For Kn = 0.01 the flow is

dominated by gas molecule collisions and the continuum models apply to such systems.

When Kn > 10, the molecules primarily collide with solid surfaces and the flow is

described as `free-molecular`. Deposition in this regime can not be modeled by the

classical continuum equations but must be described through view factor, computations

similar to those used for the radiation heat transfer or by Monte-Carlo simulations. The

intermediate range of Knudsen numbers, 0.01 < Kn < 10, corresponds to the so-called

transition flow regime where both gas-phase and surface interactions are important. In

this case, solution of the Boltzmann equation or use of specialized Monte-Carlo

simulation techniques must be applied in order to model LPCVD systems accurately.[23]

Page 28: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 17 -

1.6 Precursors for CVD

During initial stages of CVD process development, the materials deposited employing the

process were simple and the proper reactants referred to as precursors were just taken

“off the shelf”. With increasing demand for new materials as thin films, having two or

more elemental components, efforts have been made to search new precursors depending

on requirements.

Precursors for CVD can be broadly classified into three types:

1. Inorganic precursors: which do not contain any carbon

2. Metal-organic precursors: which possess organic ligands but without metal to

carbon bond.

3. Organometallic precursors: which contain organic ligands wherein metal to

carbon bond exists.[27]

CVD processes using metal organic or organometallic precursors are generally referred to

as metal-organic chemical vapor deposition (MOCVD). The highest potential for future

materials and process development is associated with the use of the metal organic (MO)

compounds. Through molecular engineering, the selected ligands influence the precursor

properties such as thermodynamic stability, kinetic lability, solubility and volatility of the

compound. The ligands have a major influence on the quality of the deposited product, as

ideally, they do control the chemical decomposition reaction taking place.[28]

Inorganic precursors are kinetically or even thermodynamically stable compounds that

need high activation energies to decompose. Decomposition takes place near

thermodynamic equilibrium and only thermodynamically stable phases can be formed.

Inorganic precursors are often prone to aggregation and as a result exhibit lower vapor

pressure. Although the vapor pressure can be increased by increasing the temperature,

care must be taken to avoid premature reaction in the gas phase before reaching the hot

substrate surface. This can lead to irreproducible growth rates and the transport of

unknown species. Particle formation resulting in clogged delivery lines and formation of

powdery products interfere with the growth of clean smooth films.[27]

Page 29: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 18 -

Metalorganic compounds suitable as precursors for most main-group and transition-metal

elements are now available from a number of commercial sources. In case of specific

requirements with stringent restrictions, the multitude of potential ligands available for

MO compounds offers the possibility to design the precursor and to engineer the

molecular decomposition pathway.

The choice of a precursor is governed by certain general characteristics which can be summarized as follows:

• Good volatility to achieve high transport rates and to accommodate high growth

rates by minimizing intermolecular forces in the condensed state and suppression

of molecular aggregation.

• High purity of precursor to reduce contamination

• Good thermal stability during evaporation and transport in the gas phase to avoid

premature decomposition

• Clean decomposition on pyrolysis to give desired material with minimum

contamination.

• Long term stability for storage

• Non-toxic, non-pyrophoric, non corrosive

• Inexpensive and simple to synthesize.

Precursors should be synthesized through easy routes to achieve high yield using

inexpensive, readily available chemicals. Although significant progress has been made on

synthesis of new compounds, there have been no source compounds satisfying all the

above criteria.[29] Thus, a great deal of work remains to be carried out before all the

challenges are addressed.

1.7 The liquid injection CVD

The lack of vapor pressure and in some cases high thermal stability of the precursors has

driven the development of the liquid injection technique in which reactant gas

composition is set by volumetric metering of liquids followed by flash evaporation.

Accurate and precise mass transport from the source in to the reaction zone is a salient

Page 30: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 19 -

feature of a CVD process. It is difficult to get an ideal precursor stable at room

temperature with high vapor pressure. There are numerous cases where the precursors are

solids. It has been long established that solid source delivery using conventional bubblers

is mass transfer limited.[30]

Consequently the effective transport rate drops as the precursor is consumed. This is a

non linear phenomenon because the smallest particles have the highest surface area to

volume ratio and they vaporize rapidly. Also, changing the carrier gas flow affects the

transport rate because to a first approximation the boundary layer thickness is inversely

proportional to carrier gas velocity. Finally, several of the precursors are non-ideal and

sublime as clusters of molecules and the average cluster size decreases with decreasing

pressure. Hence the volatility of the precursor varies with pressure. These effects make

precise delivery of low vapor pressure materials using conventional bubblers extremely

difficult. Continuous feed methods such as that developed by Hiskes et al. attempt to

compensate for these limitations.[31]

The liquid delivery technique relies on the flash vaporization of liquid solutions and over

comes the limitations of bubbling. Neat liquids as well as liquid solutions comprised of

solids dissolved in organic or inorganic media can be used with this technique. The liquid

Fig. 1.5: The industrial tool liquid injection reactor with five six inch wafers

having gas foil rotation.

Page 31: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 20 -

precursor solutions are maintained at room temperature and the composition of the inlet

to the CVD chamber is controlled by one of several methods. In the preferred

embodiment reactant gas composition is controlled through real time volumetric mixing

of the individual precursor solutions. The liquid mixture is then flash vaporized to

generate a homogenous gas at the inlet to the CVD tool. This method ensures process

reproducibility as variations in delivery rate give rise to variations in the over-all

deposition rate but do not impact film composition.[31]

Variations of this approach utilize mixtures of precursors or a single liquid solution of the

desired composition which are metered and then vaporized as a single liquid to deposit an

oxide film by MOCVD. This technique requires the solvents and solutes to be un-reactive

at the vaporization temperature which is a necessary but not sufficient criterion to prevent

homogenous nucleation inside the reaction chamber. An alternative approach to liquid

delivery uses a metering device and individual vaporizer for each liquid stream and

mixing occurs inside the shower head or process chamber. This method suffers from the

limitation that variation in the delivery rate of any of the individual precursor solutions

impact composition control. The industrial tool liquid injection reactor with five six inch

wafers having gas foil rotation used in this study is shown in Fig. 1.5.

1.8 Gas-phase chemical species measurements

Thermal decomposition of the precursor leads to the formation of thin films in a typical

CVD process. Adequate molecular stability of the precursor is required to prevent

premature reaction or decomposition of the precursor during vapor phase transport. The

volatility of a metal compound is a complex function of intermolecular forces (van der

Waals interactions, hydrogen bonds etc.). Precursor designing is a complex issue for

which reliable analytical feed back from different stages of development is necessary.

One such important stage where analysis is essential is, thermolysis of precursor. It is

interesting to know how a metal complex decomposes by the application of thermal

energy. Preferably there should be simple steps leading to formation of required

stoichiometry. The ideal case should be the decomposition of the precursor at the

required temperature process zone without any prior decomposition and leading to highly

Page 32: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 21 -

oriented growth of the film. But most of the precursors do not meet the ideal conditions.

There are deviations from ideal cases when inter and intra molecular reactions and

rearrangements taking place during thermolysis of the precursor. A detailed feed back is

very helpful in improving design and inclusion/exclusion of moieties suiting the

requirement. Typical study of precursor decomposition has to be done on gas phase of the

molecules. Gas phase studies are generally difficult and require a highly sensitive method

with low response times. This is because, the interaction between different species in the

gas phase leads to a constant change in the composition and species under investigation

during the period of measurement.

The often neglected by-products of the CVD process are volatile gases. Analysis of the

gas phase can also lead to a better understanding of the CVD reaction mechanisms and

the information can be used to refine the process. The spatial distribution of gas-phase

chemical species in a CVD reactor is controlled by chemical reactions, which are coupled

to heat and mass transfer through the strong temperature dependence of chemical reaction

rates.[23] In this section we discuss various techniques that have been used to study gas-

phase chemical species in CVD systems.

Information on the identity and concentrations of chemical species in the gas-phase can

yield significant insights into the mechanism of a process. Monitoring the depletion of the

reactant can give information on overall process efficiency, as well as the gas-phase

decomposition of the reactant. Understanding gas-phase reactions is crucial to identifying

the chemical species incident on the deposition surface, which in turn is important in

understanding how process parameters affect film properties. Measurements of gas phase

species also provide data for extensive testing of CVD models. The ideal probe for

chemical species analysis would be non intrusive, capable of selective detection of a

desired species in the presence of other species, sensitive, quantitative, and capable of

high spatial resolution in the presence of rapidly changing chemical and temperature

fields. While no single technique meets all these requirements, a combination of the

methods discussed below comes very close to this ideal situation.

Page 33: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 22 -

1.8.1 Mass spectrometry

Mass spectrometry (spectroscopy) is a well-developed analytical technique in which the

gases to be analysed are joined by electron bombardment. The resulting ions are

separated either by time-of-flight, quadrupole or magnetic methods and are counted as a

function of their charge/mass ratio. When a molecule is ionized, it often `cracks´,

producing a number of fragments of different masses with relative intensities that are

characteristic of the molecule and ionization energy.

The major advantages of mass spectrometry are its generality and ability to yield

quantitative measurements. Any chemical system can be studied because all molecules

have mass spectra. Highly reactive intermediates, as well as stable molecules can be

studied. Mass spectrometry can yield accurate numbers for the relative abundance of

different species and, with appropriate calibrations, can give absolute partial pressures

(number densities) as well.[23]

The major disadvantages of mass spectrometry for studying CVD processes are the need

to use sampling probes, which can perturb the system, and the need for independent data

on molecular cracking patterns. Sampling probes are required because mass

spectrometric analysis must be carried out at pressures of 10-5 Torr or less, much lower

than CVD processing pressures. Probes must be carefully designed to ensure

representative sampling of the gas and minimal perturbation of the CVD process. This is

particularly important when studying reactive intermediate species, as they are likely to

undergo further chemical reactions in the sampling probe. In CVD systems, probes are

particularly susceptible to clogging by deposited material. For CVD reactions that are

primarily heterogeneous, the presence of the probe can induce extraneous chemical

reactions in the sampling region and produce misleading results; in addition mass

spectrometry gives little information on the structure of a molecule.

1.8.2 Gas chromatography

Gas chromatography (GC) (vapor-phase chromatography) is a well established technique

for separating and identifying mixtures of chemicals. To analyze a chemical mixture by

GC, part of the sample is injected onto a column that has a carrier gas flowing through it.

Page 34: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 23 -

Chemical species are separated by differential adsorption on the column (generally a

high-molecular-weight liquid on a solid support or a solid adsorbent) which is chosen for

a particular chemical system. With the proper choice of column conditions, the different

species emerge from the column at different times and are detected, generally by means

of a thermal conductivity or flame ionization detector. A particular chemical is then

identified by its elusion time from the column.

The advantages of using GC technique include its generality as most stable molecules can

be analysed. GC also gives quantitative measurements of species concentrations in a

straight forward manner. With use of the proper calibrations and standards, the identity

and amount of each species in a mixture can be determined.[23] The major disadvantages

of using GC for CVD studies are that sampling is required and that it is not a real-time

analysis technique. Reactive intermediate can not be observed directly because they

undergo reactions during the analysis procedure. The dynamic range of a GC analysis can

be limited and it can be difficult to separate and detect the minor components in a

mixture. In addition, developing an analysis scheme i.e. determining the optimum column

materials and analysis conditions often must be done by trial and error.

A number of optical techniques have been applied to the analysis of deposition systems.

One of the major advantages of using photons to probe CVD processes is that they are

much less intrusive than a physical sampling probe. In addition, optical methods can

provide unambiguous species identification quantitative measurements. The

spectroscopic probes discussed in this section can be divided into those involving

vibrational excitation of the molecules and those that involve electronic excitation of the

molecules. Infrared (IR) absorption spectroscopy, laser Raman spectroscopy and coherent

anti-Stokes Raman spectroscopy (CARS) fall in to the first category while optical

emission spectroscopy, UV/visible absorption spectroscopy and laser-induced

fluorescence (LIF) spectroscopy fall into the second category.[23] During the course of

this work IR spectroscopy was extensively used for the gas phase analysis and hence a

brief description is given below.

Page 35: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 24 -

1.8.3 IR absorption spectroscopy

In IR absorption spectroscopy, a sample is exposed to a beam of IR radiation, and the

transmission is monitored as a function of the wavelength. This can be done either with

traditional dispersive techniques, Fourier transform methods, or tunable lasers. Molecules

are identified by the characteristic frequencies of IR radiation that they absorb, which

correspond to the excitation of vibrations in the molecule. This technique is widely used

in analytical chemistry, particularly for the identification of organic molecules; and

commercial instruments are readily available. Generality is one of the advantages of IR

absorption technique. With the exception of a few highly symmetric molecules, most

molecules have electric-dipole allowed transitions in the IR region. For studies of CVD

mechanisms, standard IR spectroscopy can be used to analyze gas samples extracted from

the reactor.

IR absorption spectroscopy can also be used for in situ analysis to monitor reactive

intermediate species. In this case, the IR beam passes through the CVD reactor and the

absorption spectrum of the hot, reacting gas is obtained. The disadvantage of this

technique is poor spatial resolution. The measurement is averaged over total path length

and the size of the IR beam can be quite large if conventional incoherent sources are

used. Although this advantage can be mitigated via careful reactor design, this spatial

averaging makes quantitative measurements difficult because of the steep temperature

and concentration gradients present in most CVD reactors.[23]

1.8.4 Matrix isolation FTIR spectroscopy

CVD is a process where the gas phase is highly dynamic and plays a crucial role and has

a bearing impact on the deposited material. As discussed above, during a CVD process

several chemical reactions occur, which rely on the heat and mass transfer through the

strong temperature dependence of chemical reaction rates. To study a system undergoing

constant physico-chemical changes, is an extremely challenging task and need a

combination of analytical methods. CVD processes often use metalorganic precursors

Page 36: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 25 -

designed for specific purposes. To get an understanding of the process and optimizing the

conditions for depositions mechanistic studies are sought. But decomposition of

metalorganic complexes pose a challenge as the resulting reactive intermediates

sometimes can not be detected simply because the analytical methods lack such

capability. In order to detect reactive intermediates during a chemical reaction, either the

detection method should be fast (as well as sensitive) enough (of the order of few nano

seconds or faster) or there should be a method to preserve the reactive intermediates for

sufficiently long time so that they can be analyzed and characterized.

Fig. 1.6: Schematic diagram of the matrix isolation unit.

In matrix isolation techniques, individual molecules resulting from a chemical reaction

are trapped and isolated from one another in a solid, inert matrix at low temperature while

Vacuum pump

Expander Radiant heat shield

Sample holder

Temperature controller

Compressor

Cooling water supply and drain

Vacuum shroud

Gas lines Expander electrical power cable

Vacuum valve

Instrumentation skirt

Pressure gauge P

Page 37: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 26 -

their spectrum is measured. According to IUPAC Compendium of Chemical

Terminology, Matrix isolation is “A term which refers to the isolation of a reactive or

unstable species by dilution in an inert matrix (argon, nitrogen, etc.), usually condensed

on a window or in an optical cell at low temperature, to preserve its structure for

identification by spectroscopic or other means.”[32, 33]

Most of the chemical processes result in atoms or molecules which undergo further

reactions. In such a case isolating the primary reaction product suppresses successive

reactions. In matrix isolation technique, reactive intermediates and compounds which are

unstable under normal conditions are frozen in a rigid matrix at about 4 to 40 K mainly

consisting of non-reactive media such as noble gases (Argon, Neon, etc.) or nitrogen.[34]

Conventional techniques fail to detect the reactive intermediates because of their

insensitivity towards low concentrations of the species. Matrix isolation method provides

a possibility to accumulate the reactive intermediates over a period of time so that

spectroscopically detectable concentrations can be achieved.[35] Infrared and visible-

ultraviolet (electronic) spectroscopies are the most widely used tools for detecting and

studying reaction intermediates. Infrared absorptions of molecules trapped in neon or

argon matrices are sharp, with typical band widths (full width at half maximum, FWHM)

of the order of 1 cm-1. Earlier experiments have also demonstrated that at 20 K and below

solid nitrogen and argon are sufficiently rigid so that molecular diffusion and subsequent

chemical reactions are effectively inhibited. Additional advantage is that, nitrogen and

the noble gases are transparent through the entire far infrared spectral region to the

vacuum ultraviolet region. Hence when coupled with FTIR/UV-Visible spectroscopy,

matrix isolation measurements provide a potentially valuable analytical tool.[34] A

schematic diagram of basic matrix isolation unit is shown in Fig. 1.6.

This suits very well with the requirement for studying the decomposition of metal organic

precursors. With the increased life time of reactive intermediates, using different optical

spectroscopic tools, analyses can be done which helps in improving the basic

understanding of decomposition process. In addition, the designer of the precursors for

CVD gets much needed feed back from the decomposition studies and which helps to

include/exclude design changes in the precursor molecule.

Page 38: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 27 -

1.9 MOCVD of TiO2 thin films

The CVD of TiO2 thin films spans over five decades and a variety of precursors used for

thin film depositions. Mainly the simple halides like [TiCl4], [TiI4][49] used in early

studies have been replaced with alkoxides like [Ti(OiPr)4], [Ti(OEt)4],[13] etc. Simple

compounds like titanium nitrate [Ti(NO3)4], have also been investigated for the

deposition of TiO2 thin films using CVD technique.

The air and moisture sensitivity of halides and alkoxides of titanium has made their

handling difficult during a CVD process. In order to overcome the sensitivity problem,

several chelating ligands were tried in combination with alkoxides of titanium. Table 1.3

lists notable precursors used for the deposition of TiO2 thin films. Though there are

several precursors which can be used for the deposition of TiO2 thin films, improvements

are possible in terms of tuning of thermal stability, chemical reactivity and orderly

decomposition of available precursors by the inclusion/exclusion of moieties in the ligand

sphere. During the course of this work, the effect of fine tuning the ligand sphere on the

Table 1.3: The list of notable precursors reported for the deposition of TiO2 thin films using CVD

Precursor Tb (°C) Minimum Td

(°C)

References

[TiCl4] Not reported 200 36-38

[Ti(OiPr)4] 80-120 250 39-41

[Ti(NO3)4] 120 300 39, 42

[Ti(µ-ONep)(ONep)3]2 22-40 230 43

[{HB(pz)3}Ti(OPri)3] 115 450 44

[TiOPri)2(dmae)2] Liquid injection 400 45

[TiOPri)2(thd)2] Liquid injection 350 46, 47

[Ti(2meip)2] Liquid injection 400 48

Tb = bubbler temperature, Td = Deposition temperature

Page 39: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 28 -

chemical and physical properties of the precursors were investigated. Following section

provides the details about the scope of the present work.

1.10 Scope of the present work

The growth of TiO2 and related oxides has been successfully demonstrated using the

above mentioned titanium sources with a varying degree of success. From the point of

view of precursor chemistry, the interactions at the molecular level which influence the

precursor purity, volatility and decomposition kinetics needs to be thoroughly

investigated. Due to the complexity involved with respect to the influence of process

parameters, it is worthwhile to retain the established concepts of precursor design rather

than to explore totally new ideas. There is still scope for improvement with respect to the

stability and thermal decomposition of titanium precursors. The use of metal alkoxides as

a synthetic platform for molecular design of suitable precursors for TiO2 thin films has

been demonstrated. The main objective of this work was to engineer the ligand

framework of well established key structures by introducing small and distinct changes in

a systematic manner. Chelating ligands such as ß-ketoesters, malonates and ß-ketoamides

were tested in combination with alkoxides of titanium to yield mixed alkoxide

complexes.

The resulting compounds were characterized for their chemical and physical properties

by NMR, IR, melting point, mass spectrometry, elemental analysis and single crystal X-

ray diffraction. Single crystal X-ray diffraction was employed as a tool to probe and

engineer the structure and function of the novel precursors. The objective in this study

was to systematically study the properties of the building blocks such as size, shape, and

directionality of functional groups with a view to determine how these parameters control

and influence the packing in the crystal lattice. The main focus was to design and

synthesize coordinatively saturated metal complexes with appropriate ligands and to

correlate the design changes to changes in the thermal properties of the precursors. It is

important to control the nuclearity of metal complexes by the use of efficient designing.

This is because nuclearity of the precursor has a significant influence on the thermal

properties of the precursor. The determination of the thermal properties of the compounds

Page 40: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 29 -

which is a key figure of merit for CVD precursors was carried out in detail using

thermogravimetry (TG) and differential thermal analysis (DTA).

The rationally developed compounds of titanium were then screened for CVD

applications using a homebuilt cold wall CVD reactor. It was demonstrated that by

introducing small and distinct changes in the ligand system, which act as targeted

cleavage points, it was possible to grow TiO2 thin films at low deposition temperatures.

Therefore the scaling up precursor synthesis was taken up (~ 25 g). As these compounds

showed high degree of solubility in common organic solvents, they were dissolved in n-

butyl acetate and used as precursors for liquid injection MOCVD of TiO2 and SrTiO3

employing an industrial tool reactor (MOCVD facilities of Forschungszentrum Jülich).

The benchmark precursor [Ti(OPri)2(thd)2] which is commercially available was used to

grow TiO2 and SrTiO3 to compare the results to those obtained from the engineered

precursors.

1.10.1 Understanding the precursor fragmentation

The main concern of the research community involved in the field of CVD is to prepare

films with the most satisfying properties for the desired application. However, a very

good control of the CVD process is generally difficult as many complex phenomena such

as mass transport, gas phase and surface reactions etc. are involved during deposition.

Among these phenomena, gas phase reactions are particularly important in determining

the material characteristics and properties. A very good understanding of the reaction

pathways and kinetics is thus required for optimization of the process. A powerful

method to analyze relevant molecular decomposition mechanisms, which has not yet

been widely recognized in the CVD community, is to analyze the reaction intermediates

formed using the matrix-isolation–FTIR spectroscopy techniques. Using this technique,

the reaction intermediates formed by the thermolysis of CVD precursors in the gas phase

can be preserved by shock-like quenching of the gas phase onto a cooled matrix window.

(e.g. CsI maintained at 10 K) and are subsequently analyzed by FTIR spectroscopy. This

elegant technique gives insights into the CVD process on a molecular level and thus

provides a rational basis for further precursor development. For this purpose, a matrix

Page 41: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 30 -

apparatus consisting of a cooling system (Helium closed cycle), vacuum parts and a FTIR

spectrometer was fabricated and used for investigating the molecular mechanisms

involved during precursor decomposition. The most widely used alkoxide of titanium,

[Ti(OPri)4] and its ß-ketoester derivative, [Ti(OPri)2(tbaoac)2] and the commercially

available precursor [Ti(OPri)2(thd)2] were studied for thermal decomposition using matrix

isolation FTIR spectroscopy. By employing the present set up of matrix isolation unit

with a themolysis oven, there is always a possibility of coupling between the

homogeneous gas phase reactions and heterogeneous gas phase reactions occurring on

the surface of the oven. However, the present set up has advantages over other methods

and can be effectively be used in combination with other gas phase or surface studies for

a deeper understanding of the reaction mechanisms. There are relatively very few studies

on the gas phase decomposition studies for organometallic precursors when compared to

a large number of solution based studies. The results obtained from matrix isolation

studies were successful to identify clearly the organic intermediates involved in the

process. Use of FTIR in combination with matrix isolation techniques could provide

insights in to mechanism not considered earlier in such studies.

The work presented here focuses on,

a) the very important rational basis for further precursor development which includes

mechanistic studies on the fragmentation and transformation of the precursors to the final

material and

b) development of the CVD process for technologically useful oxide materials such as

TiO2 and SrTiO3 and the study of these materials for their device related characteristics.

The nature of the work in this field, i.e. bridging precursor chemistry and thin film

processes with mechanistic studies is somewhat unique in the field of CVD and paves

way to give scientific knowledge on the relationship between the molecular structure of

the precursors and the nature of the films obtained.

Although MOCVD growth TiO2 and related oxide materials has been successfully

accomplished with varying degrees of success using above mentioned titanium

complexes, current generation precursors exhibit significant deficiencies. The use of

metal alkoxides as synthetic platform, molecular design of the useful precursors for TiO2

can be achieved. The main goal of the present work was to introduce small changes in the

Page 42: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 31 -

ligand sphere of the well known ß-diketonate systems and use them for the MOCVD of

TiO2.

First part of the work was to explore new compounds by molecular engineering, a

systematic approach by varying the ligands by introducing small changes. As a result ß-

ketoesters, malonate and ß-ketoamide ligands were tested in combination with alkoxides

and amides of titanium. These precursors were characterized for their chemical and

physical properties by NMR, IR, Mass spectrometry, elemental analysis, and single

crystal X-ray diffraction. After clear confirmation for the formation of the compounds, a

detailed analysis of the thermal properties of the precursors for their suitability for CVD

was carried out using thermo gravimetric and differential thermal analysis.

The second part of the work was to deposit TiO2 thin films using a home built low

pressure horizontal CVD reactor. The available choice of precursors with small variations

in the ligand sphere offered a large possibility for depositing TiO2 thin films. Selected

precursors were tried taking thermal properties as merit and TiO2 thin films were grown

with various temperature and pressure series. Low temperature depositions were possible

with designed precursors without the use of additional oxygen. However, higher carbon

content in the film was noticed.

High solubility of these designed precursors makes them candidates for liquid injection

CVD. One of the precursor [Ti(OPri)2(tbaoac)2], was tried for deposition of TiO2 and

complex oxide SrTiO3 using solution based liquid injection CVD. These experiments

were carried out using MOCVD facilities at Forschungszentrum Jülich. Using the

commercially available precursor, [Ti(OPri)2(thd)2] comparative studies were performed

for depositing TiO2 and SrTiO3 thin films. The second part of the thesis provides details

of the CVD experiments carried out using home built reactor as well as the industrial tool

reactor.

For the macroscopic understanding of the CVD process it s essential that individual steps

involved must be studied carefully. Gas phase of a precursor is one such quite complex

systems for investigation and vitally influential of all processes in a CVD process. Any

design change involved during precursor synthesis has to provide detailed logic for

design inclusion in the molecule. Mechanistic details are often sought in such cases for

fundamental understanding of the functionality of the precursor. Most widely used

Page 43: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 32 -

alkoxide of titanium, [Ti(OPri)4] and its ß-ketoester derivative, [Ti(OPri)2(tbaoac)2] and

commercially available precursor [Ti(OPri)2(thd)2] were studied for thermal

decomposition using matrix isolation FTIR spectroscopy. The third part of the work

extensively reports the details of the decomposition studies carried out on the above

mentioned TiO2 precursors using MI-IR. There are relatively very few studies on the gas

phase decomposition studies for organometallic precursors when compared to a large

number of solution based studies. The results obtained from matrix isolation studies were

successful to identify clearly the organic intermediates involved in the process. Use of

FTIR in combination with matrix isolation techniques could provide insights in to

mechanism not considered earlier in such studies.

Design changes in the ligand sphere influence the resulting physical properties of the

precursors. Fourth part of the work describes the thermal properties of the newly

developed precursors relevant for the CVD purposes. Simultaneous TG-DTA were

carried out on all of the newly developed complexes to evaluate their suitability as

precursors. A comparison to the thermal properties of the parent alkoxides and the

recently developed titanium precursors were done. Selected precursors were screened

with isothermal studies at different temperatures and sublimation rates were determined.

The inclusion of ester moiety in the ligand sphere induced low decomposition

temperature. In order to evaluate the hydrolytic stability, a comparative study was carried

out with structurally similar ß-diketonate complexes using NMR as an analytical tool.

Page 44: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 33 -

1.11 References

[1] S. I. Stupp, P. V. Braun, Science 1997, 277, 1242.

[2] S. B. Sinnotta, E. C. Dickey, Materials Science and Engineering 2003, R 43, 1.

[3] D. P. Norton, Materials Science and Engineering 2004, R 43, 139.

[4] U. Diebold, Surface science reports 2003, 48, 53.

[5] L. S. Dubrovinsky, N. A. Dubrovinskaia, V. Swamy, J. Muscat, N. M. Harrison,

R. Ahuja, B. Holm, B. Johansonn, Nature 2001, 410, 654.

[6] S. A. Chambers, Y. Gao, S. Thevuthasan, Y. Liang, R. J. Shivaparan, J. Smith, J.

Vac. Sci. Technol. A, Part2 1996, 14(3), 1387.

[7] P. A. Morris Hotsenpiller, A. Roshko, J. B. Lowekamp, G. S. Rohrer, J. Cryst.

Growth 1997, 174 (1-4), 424.

[8] S. Yamamoto, T. Sumita, T. Yamaki, A. Miyashita, H. Naramoto, J. Cryst.

Growth 1997, 237-239, 569.

[9] B.-S. Jeong, J. D. Budai, D. P. Norton, Thin Solid Films 2002, 422 (1–2), 166.

[10] M. Murakami, Y. Matsumoto, K. Nakajima, T. Makino, Y. Segawa, T. Chikyow,

P. Ahmet, M. Kawasaki, H. Koinuma, Appl. Phys. Lett. 2001, 78 (18), 2664.

[11] S. A. Chambers, C. M. Wang, S. Thevuthasan, I. Droubay, D. E. McCready, A. S.

Lea, V. Shutthanandan, C. F. Windisch Jr., Thin Solid Films 2002, 418, 197.

[12] M. L. Hitchman, J. Zhao, J. Phys. IV 1999, 9, Pr8-357.

[13] H.-K. Ha, M.Yoshimoto, H. Koinuma, B.-K. Moon, H. Ishiwara, Appl. Phys. Lett.

1996, 68, 2965.

[14] J. V. Grahn, M. Linder, E. Fredriksson, J. Vac. Sci. Technol. A 1998, 16, 2495.

[15] J.-S. Chen, S. Chao, J.-S. Kao, G.-R. Lai, W.-H. Wang, Appl. Opt. 1997, 36,

4403.

[16] The National Technology Roadmap for Semiconductors, 3rd edition,

Semiconductor Industry Association, San Jose, CA, 1997.

[17] G. H. Johnson, J. Am. Ceram. Soc. 1953, 36, 97.

[18] K. S. Yeung, Y. W. Lam, Thin solid films 1989, 109, 169.

Page 45: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 34 -

[19] A. J. Moulson, J. M. Herbert, Electroceramics : Materials, Properties and

Applications, Chapman and Hall, London, 1990.

[20] W. C. Hendricks, S. B. Desu, C. H. Peng, Chem. Mater. 1994, 6, 1955.

[21] R. F. Bunshah, PVD and CVD coatings, ASM international, Ohio, 1992.

[22] J. E. Mahan, Physical vapor deposition, Wiley Interscience, 2000.

[23] M. L. Hitchman, K. F. Jensen, Chemical vapor deposition-Principles and

applications, Academic Press, 1992.

[24] W. E. Sawyer, A. Mann, US. Patent 229,335, 1880.

[25] J. M. Blocher. Jr., J. Electrochem. Soc. 1960, 107, 117C.

[26] P. D. Agnello, IBM J. Res. & Dev. 2002, 46, 317.

[27] T. Kodas, M. J. Hampden-Smith, The chemistry of metal CVD, VCH Publishers,

Weiheim, Germany, 1994.

[28] R. A. Fischer, Chemie i. u. Zeit, 1995, 29, 141.

[29] A. C. Jones, J. Mater. Chem. 2002, 12, 2576.

[30] P. C. van Buskirk, J. Zhang, P. S. Kirlin, Ferroelectric thin film memories,

Science Forum, Tokyo, 1995.

[31] R. Hiskes, S. A. DiCarollis, J. L. Young, S. S. Laderman, R. D. Jacowitz, R. C.

Taber, Appl. Phys. Lett. 1991, 59, 607.

[32] IUPAC, Compendium of Chemical Terminology 1990, 62, 2200.

[33] IUPAC, Compendium of Chemical Terminology 1994, 66, 1138.

[34] M. E. Jacox, Chem. Soc. Rev. 2002, 31, 108.

[35] E. Whittle, D. A. Dows, G. C. Pimentel, J. Chem. Phys. 1954, 22.

[36] G. Hass, Vacuum 1952, 2, 331.

[37] R. N. Ghoshtagore, J. Electrochem. Soc. 1970, 117, 1310.

[38] R. N. Ghoshtagore, J. Electrochem. Soc. 1970, 117, 529.

[39] C. J. Taylor, D. C. Gilmer, D. G. Colombo, G. D. Wilk, S. A. Campbell, J.

Roberts, W. L. Gladfelter, J. Am. Chem. Soc. 1999, 121, 5220.

[40] T. W. Kim, M. Jung, H. J. Kim, T. H. Park, Y. S. Yoon, W. N. Kang, S. S. Yom,

H. K. Na, Appl. Phys. Lett. 1994, 64, 1407.

[41] A. C. Jones, T. J. Leedham, J. Brooks, H. O. Davies, J. Phys. IV 1999, 9, Pr8-

553.

Page 46: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 35 -

[42] D. C. Gilmer, C. J. Taylor, D. G. Colombo, J. Roberts, G. Haugstad, S. A.

Campbell, H.-S. Kim, G. D. Wilk, M. A. Gribelyuk, W. L. Gladfelter, Chem. Vap.

Deposition 1998, 4, 9.

[43] J. J. Gallegos, T. L. Ward, T. J. Boyle, M. A. Rodriguez, L. P. Francisco, Chem.

Vap. Deposition 2000, 6, 21.

[44] E.-C. Plappert, K.-H. Dahmen, R. Hauert, K.-H. Ernst, Chem. Vap. Deposition

1999, 5, 79.

[45] A. C. Jones, T. J. Leedham, P. J. Wright, M. J. Crosbie, K. A. Fleeting, D. J.

Otway, P. O’Brien, M. E. Pemble, J. Mater. Chem. 1998, 8, 1773.

[46] M. Balog, M. Schieber, J. Cryst. Growth 1972, 17, 298.

[47] D. C. Bradley, R. C. Mehrotra, D. P. Gaur, Metal Alkoxides, Academic Press,

New York, 1978.

[48] Y.-S. Min, Y.-J. Cho, D. Kim, J.-H. Lee, B.-M. Kim, S.-K. Lim, I.-M. Lee, W.-I.

Lee, Chem. Vap. Deposit. 2001, 7, 146.

[49] M. Schuisky, A. Hårsta, J. Phys. IV 1999, 9, Pr8-381.

Page 47: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 36 -

Chapter 2

Rational development of titanium precursors used for MOCVD of titanium containing oxide thin films

Abstract

The concept of introducing small variations in the ligand sphere of well known structures

for the synthesis of improved metal organic chemical vapor deposition (MOCVD)

precursors for titanium have been studied in detail. The reaction of titanium alkoxides

with ß-ketoester, ß-ketoamide and malonate ligands resulted in complexes having

potential to serve as precursors for MOCVD of TiO2 thin films. Titanium

bis(isopropoxide) bis(methylacetoacetate) (1), titanium bis(ethoxide)

bis(methylacetoacetate) (2), titanium bis(isopropoxide) bis(tert-Butylacetoacetate) (3),

titanium bis(ethoxide) bis(tert-Butylacetoacetate) (4) and titanium bis(isopropoxide)

bis(N,N-diethylacetoacetamide) (5) and bis-[(di-ethylmalonato) tetra(isopropoxy)-µ-

ethoxy-titanium(IV)] (6) Titanium bis(isopropoxide) bis(diethylaminoethoxide) (7),

Titanium bis(ethoxide) bis(diethylaminoethoxide) (8) have been synthesized and

characterized by elemental analysis, NMR, and mass spectrometry. While the

aminoalkoxide complexes were viscous liquids, ß-ketoester/malonate and ß-ketoamide

complexes were crystalline solids. The molecular structure of the compounds as

determined by single crystal X-ray diffraction revealed that complexes with ß-ketoester

and ß-ketoamide ligands exist as monomers while complex with malonate ligand

undergoes trans-esterification reaction which exhibits a centro-symmetric dimeric

structure. This study assumes importance as there are seldom any reports about solid state

structures of ß-ketoesters with titanium alkoxides in monomeric forms. It is shown that

by introducing small changes in the established key structures of the existing ligands, it is

possible to tailor the physical properties of the resulting titanium complexes.

Page 48: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 37 -

2.1 Introduction

MOCVD offers many attractions for titanium dioxide film growth, including conformal

deposition on a variety of complex substrates, low equipment costs, easy scale-up, lower

growth temperatures, and higher growth rates. However, the success of an MOCVD

process depends critically on the availability of volatile, thermally stable precursors

which exhibit constant vapor pressure and the capacity to selectively form the desired

phase at the substrate surface. Generally low-melting precursors are preferred because on

the one hand solid compounds can be handled with ease at room temperature; on the

other, the liquid form at reservoir operating temperatures affords constant surface area for

stable vapor delivery to the reactor (as discussed in section 1.7). The versatility of

precursor chemistry derives largely from variety of molecular precursors available as

sources.

The chemical vapor deposition of titanium dioxide has been practiced for over five

decades.[1, 2] [TiCl4] has been used as precursor for TiO2 thin films extensively. Using O2

and H2O as reactants thin films of TiO2 have been obtained. However, [TiCl4] is toxic

and requires special equipment and safety installation for handling.[3] The pioneering

work by Bradley and co-workers[4] explored the relationship between the molecular

structure of metal alkoxides and their physical properties such as degree of association,

i.e. nuclearity and volatility. The volatility of metal alkoxides is strongly influenced by

their tendency to form oligomers and clusters. This is related to positive charge on the

metal center, which has a tendency to form bonds with negatively charged oxygen of the

neighboring M(OR)n molecules. It was concluded that, in order to inhibit oligomerization

in metal alkoxides containing large, highly positively-charged metal atoms, bulky

sterically demanding ligands such as iso-propoxide, tert-Butoxide must be employed. So

volatile titanium alkoxides complexes were synthesized and used as precursors for

deposition of TiO2 thin films and one of the most widely used titanium precursor is

titanium tetraisoproxide (TTIP). However, the alkoxide precursors contain unsaturated

four-coordinate metal centers and the alkoxy ligands undergo a facile catalytic hydrolytic

decomposition reaction in the presence of trace water.[5] These complexes are therefore

extremely air and moisture-sensitive, which limits their shelf-life and makes them

Page 49: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 38 -

difficult to handle and use in MOCVD, especially in the case of solution-based liquid

injection MOCVD applications.

In order to reduce the moisture sensitivity of the Ti-alkoxide precursors, chelating ß-

diketonate groups have been inserted to increase the saturation of the Ti(IV) coordination

sphere. [Ti(OPri)2(acac)2][6] and [Ti(OPri)2(thd)2][7] have been used for the deposition of

TiO2 thin films. [Ti(OPri)2(thd)2] is a monomeric six coordinated complex in solution and

stable in solution than its parent alkoxide [Ti(OPri)]4.[8] Presence of ß-diketonate group

with bulky side chain such as Hthd results in higher thermal stability of the precursor and

[Ti(OPri)2(thd)2] is perhaps most widely used mixed alkoxide-ß-diketonate complex of

titanium.

Another approach to increasing the coordinative saturation of the metal center is to

introduce bidentate donor functionalized ligands such as 2-dimethylaminoethanolate or

diolates such as 2-methylpentane-2, 4-diolate. There are few examples where complexes

other than ß-diketonates or alkoxides have been used for MOCVD of TiO2 and related

oxides; namely monomeric ß-ketiminate complex [Ti(2meip)2] 2meip = 4-(2-

methylethoxy)imino-2-pentonate[9]and [Ti(NO3)]4[10-11] are the notable ones.

The objective of this work was to vary the terminal groups of the well established ß-

diketonate complexes of titanium alkoxides and study the effect on the volatility and

decomposition pathways. In addition, mixed alkoxides of titanium with amino alcohols

were also studied by inclusion of small variations in the ligand periphery. With relative

ease in preparation of metal alkoxides after Bradley[4], the efforts aimed at the precursor

development were turned towards mixed alkoxide-chelating ligand complexes, mainly ß-

diketonates. In a deliberate effort to include targeted “cleavage points” to bring about

orderly thermolytic decomposition, we tried ß-ketoesters, ß-ketoamides and malonates, as

chelating ligands. In addition, the use of aminoalcohols in combination with titanium

alkoxides was expected to act as donor functionalized chelating ligand. The ligands used

for complexation were, methylacetoacetate (meaoac), tert-Butylacetoacetate (tbaoac), N,

N-diethylacetoacetamide, (deacam) di-ethylmalonate (deml) and diethylaminoethanol

(deae). Inclusion of ester/malonate/amide moiety in the side chain of the ß-keto structure

is expected to weaken the metal-oxygen bond of the chelating ring thus leading to facile

thermal fragmentation of the complexes. It was expected that these complexes would

Page 50: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 39 -

combine the chemical stability derived from ß-keto structure and facile fragmentation

pattern induced by ester moiety embedded therein and volatility derived from alkoxy

ligands.

2.2 Experimental Section

For precursor synthesis all manipulations were performed utilizing oven-dried reaction

vessels and Schlenk techniques under inert atmosphere of purified argon or in a glove

box. Solvents were dried under N2 by standard methods and stored over 4 Å molecular

sieves. Proton- 1H and carbon- 13C NMR spectra were recorded for all synthesized

compounds. The spectra were referenced to residual protic impurities of the internal

solvent and corrected to tetramethylsilane. The integration of peaks and peak intensity

analyses were done using Mestrec® software version 2.30. Bruker Advance DPX 200 and

Bruker Advance DPX 250 spectrometers were used.

Elemental analysis (Elemental, CHNSO Vario EL, Hanau) and mass spectra were

provided by the Spectroscopy and Chromatography Analytical section of the Faculty of

Chemistry at the Ruhr-University of Bochum. Electron Ionization (EI) mass spectra were

recorded using ionization energies between 24 eV to 70 eV using CHS-Mass

spectrometer “Varian MAT” (Bremen). Output spectra was given as specific masses

(m/z) based on abundant isotopes, 1 1H, 12

6C 14 7N 16 8O, 48

22Ti. Infrared data were

collected on a 1650 Perkin-Elmer spectrometer. The alkoxides of titanium namely,

Titanium(IV) isopropoxide (Aldrich) and Titanium (IV) ethoxide (Fluka) were purchased

and used as such without further purification, but stored in a glove box excluding air and

moisture during storage. Ligands used in the study were purchased from Aldrich and

distilled prior to use.

Synthesis of [Ti(OPri)2(meaoac)2] (1). About 2.32 ml (0.02 mol) of methyl acetoacetate

ester was added to 2.97 ml (0.01 mol) of Ti(OPri)4, the resulting yellowish clear solution

was stirred for 2 hours and kept in a refrigerator at -20 °C. Colourless needles were

crystallised after 24 hours. The crystals were washed with cold pentane and dried in

vacuum. Yield 87%. C16H28O8Ti calculated C, 48.50, H, 7.12; Found C, 47.87, H, 7.35.

Page 51: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 40 -

1H NMR ( 200 MHz, C6D6, 25 °C) δ 1.4 (12H, d, OCH(CH3)2) ; δ 1.9 (6H, s, CO(CH3));

δ 3.6 (6H, s, OCH3); δ 5.1 (2H, s, COCHCO ); 5.3 (2H, hept, CH(CH3)2); 13C{1H}

NMR 200 MHz C6D6 25 °C. 186(CH3CO), 173(COO), 89(COCHCO), 79(CH -OPri),

51(OCH3), 29(CH3 lig), and 26(CH3 -OPri). Mass spectrometry. (EI+) m/z 337, 25% [1 -

isopropyl group]; 281, 100% [Ti(meaoac)2]; 211, 35% [Ti2(meaoac)].

Synthesis of [Ti(OEt)2(meaoac)2] (2). About 2.32 ml (0.02 mol) of methyl acetoacetate

ester was added to 2.09 ml (0.01 mol) of Ti(OEt)4, and the resulting clear solution was

kept in a refrigerator at -20 °C. Slightly yellow colored crystalline solid was formed

which was filtered and dried in vacuum. Yield 80%. Elemental analysis (%):C14H24O8Ti,

calculated C, 45.65, H, 6.57.; Found C, 45.69, H, 6.49. 1H NMR (200 MHz, C6D6, 25

°C) δ 1.4 (6H, t, OCH2CH3) ; δ 1.8 ( H, s, CO(CH3)) ; δ 3.5 (6H, s, OCH3) ; δ 4.7 (4H, q,

OCH2CH3) ; δ 5.1 (2H, s, COCHCO); 13C{1H}-NMR (200 MHz C6D6 25 °C) δ[ppm] =

185 (CH3CO), 173 (COO), 88(COCHCO), 73 (CH2 –ethoxy), 51 (OCH3 -lig), 29(CH3 -

lig), 18(CH3-Ethoxy). Mass spectrometry: (EI+) m/z 323, 20% [2 – Ethoxy group]; 278,

25% [Ti(meaoac)2]; 211, 15% [Ti2(meaoac)]; 162, 100% [Ti(meaoac)].

Synthesis of [Ti(OPri)2(tbaoac)2] (3). About 3.3 ml (0.02 mol) of the tert. Butyl

acetoacetate ester was added to 2.97 ml (0.01 mol) of Ti(OPri)4. The resulting yellowish

clear solution was stirred for 2 hours and stored at -20 °C. Pale yellow needles were

crystallized after 24 hours, and were washed with cold pentane and dried in vacuum.

Yield, 84%. Elemental analysis (%):C22H40O8Ti calculated C, 55,0, H, 8.39. Found C,

53.99, H, 8.39. 1H-NMR (200 MHz, C6D6, 25 °C) δ[ppm] = 1.5 (12H, d, OCH(CH3)2) ;

1.7 (18H, s, C(CH3)3); 1.9 (6H, s, CO(CH3)); 5.2 (2H, hept, CH(CH3)2); 5.3 (2H, s,

OCCHCO); 13C{1H}-NMR (200 MHz C6D6 25 °C) δ[ppm] = 187 (CH3CO), 172 (COO),

91 (CHCO), 81 (C tBut), 79 (CH -OPri), 29 (CH3), 26 (CH3 -OPri) and 25 ( CH3 -tBut).

Mass spectrometry: (EI+) m/z 422, 15% [3- tBut]; 362, 15% [Ti(tbaoac)2]; 267, 100%

[Ti(OPri)(tbaoac)]; 225, 20%[Ti(OPri)3]; 165, 20% [Ti(OPri)2]; 57, 20% C(CH3)3.

Synthesis of [Ti(OEt)2(tbaoac)2] (4). About 3.3 ml (0.02 mol) of tert. Butyl acetoacetate

ester was added to 2.09 ml (0.01 mol) of Ti(OEt)4, the resulting yellowish clear solution

Page 52: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 41 -

was stirred for 2 hours and kept in a refrigerator at -20 °C. Off white solid was

crystallised after 24 hours as pale yellow needles. Washed with cold pentane and dried in

vacuum. Yield, 83%. C20H36O8Ti, Calculated C, 53,10, H, 8.02. Found C, 53.14, H, 8.22. 1H NMR ( 200 MHz, C6D6, 25 °C) δ 1.5 ( 6H, t, OCH2CH3 ) ; δ 1.7 ( 18H, s, C(CH3)3 );

δ 1.9 ( 6H, s, CO(CH3) ); δ 4.8 ( 4H, q, OCH2CH3 ) ; δ 5.2 ( 2H, s, COCHCO ); 13C{1H}

NMR (200 MHz C6D6 25°C). 186 (CH3CO), 173(COO), 90(CHCO), 81 (C tBut), 73

(CH2 Ethoxide), 29(CH3), 28(CH3 tBut), and 19(CH3 Ethoxide). Mass spectrometry. (EI+)

m/z 08, 3% [4- CO2]; 362, 3% [Ti(tbaoac)2]; 239, 10% [Ti3(OPri)2]; 57 100% tBut.

Synthesis of [Ti(OPri)2(deacam)2] (5). A diluted solution of 3.14 ml (0.02 mol) of

diethyl acetamide in 20 ml of hexane was added to a diluted solution of Ti(OPri)4 [2.97

ml (0.01 mol) in hexane (20 ml)]. The mixture was refluxed for 12 hrs at 68 °C and then

the resulting mixture was stored in the refrigerator at -20 °C for 24 hours. Brown colored

solid crystallized which was washed repeatedly with cold hexane which resulted in pale

yellow crystalline product. Dissolved in hot hexane and was allowed to recrystallize

slowly at -20 °C to yield off white crystals. Yield, 43%. C22H42O6N2Ti, Calculated C,

55,19, H, 8.78, N, 5.85. Found C, 55.14, H, 8.82, N, 5.78. 1H-NMR (250 MHz, C6D6, 25

°C) d 1.90 (6H, s, CH3 deacam), 1.45 (6H, d, CH3 OPri, 1J ~ 5.97 Hz), 0.7 (6H, t,

NCH2CH3a,), 0.95 (6H, t, NCH2CH3b,), 2.64 (4H, q, NCH2a CH3), 3.07 (4H, q, NCH2b

CH3), 4.66 (2H, s, CH OPri), 4.75 (2H, s, CH deacam) 13C{1H} NMR (200 MHz C6D6 25

°C). d 41 (NCH2aCH3), 42 (NCH2bCH3), 27 (CH3 OPri), 26 (CH3 deacam), 13.02

(NCH2CH3a), 13.07 (NCH2CH3b), 86 (CH deacam), 70 (CH OPri), 185 (COCH3

deacam), 168 (CO deacam). Mass spectrometry. (EI+) m/z 419, 23% [5- OPri]; 362,

15% [Ti(deacam)2]; 239, 10% [Ti3(OPri)2]; 85 100% [deacam-NEt2]

Synthesis of [Ti2(µ-OEt)2(OPri)4(deml)2] (6). About 3.1 ml (0.02 mol) of diethyl

malonate was added to a solution of 2.97 ml (0.01 mol) Ti(OPri)4 in 10 ml hexane. The

colourless solution turned slightly yellow and warm. After two hours of stirring, the

mixture was cooled down to -25 °C for 24 hours. Slightly yellow, cubic crystals were

formed and dried in vacuum after the removal of the solvent. Yield, 88 %. 1H-NMR

(250MHz, Toluene, 25 °C) Unidentified peaks (refer results and discussion part): d

Page 53: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 42 -

[ppm] = 5.6 – 4.0 (multiplet), 3.866 (q), 3.046 (s), 3.036 (s), 3.026 (s), 1.7-1.0 (multiplet),

0.971 (q), 0.859 (multiplet); 13C-NMR (250 MHz, C6D6, RT, 20000 scans): can not be

interpreted.

Synthesis of [Ti(OPri)2(deae)2] (7). A diluted solution of 2.66 ml (0.02 mol) of diethyl

aminoethanol in 20 ml of hexane was added to a diluted solution of titanium

tetraisopropoxide 2.97 ml (0.01 mol) in hexane (20 ml). The resulting solution was

refluxed at 70 °C for two hours. Solvent and volatile products were removed under low

pressure resulting in a clear, yellow oily liquid as product. Yield, 94 %. 1H NMR ( 200

MHz, C6D6, 25 °C) δ δ 1.0 (12H, t, NCH2CH3 ) δ 1.5 ( 12H, d, OCH(CH3)2), δ 2.8 (12H,

q + t, NCH2CH3+ OCH2CH2NEt2), δ 4.6( 4H, t, O-CH2CH2N), δ 5.0 (2H, sep,

OCH(CH3)2). 13C{1H}-NMR (200 MHz C6D6 25 °C) δ[ppm] 11.5 (N-CH2-CH3)2, 26.67

(OCH(CH3)2), 47.51 (NCH2CH3), 57.23 (OCH2CH2NEt2), 68.04 (OCH2CH2NEt2),

77.13 (OCH(CH3)2). Mass spectrometry. (EI+) m/z Expected mass 398.8. 340 (2%) [7-

2Et], 279 (4%) [7 – lig. deae], 86 (100%) remains speculative.

Synthesis of [Ti(OEt)2(deae)2] (8). A diluted solution of 2.09 ml (0.01 mol) of titanium

tetraethoxide in 20 ml of hexane was added to a diluted solution of 2.66 ml (0.02 mol) in

hexane (20 ml). Resulting solution was refluxed at 70 °C for two hours. The volatile

products were removed under low pressure resulting in a clear, yellow oily liquid as

product. Yield, 96 %. TiC16H38O4N2 Calculated C, 51,89, H, 10.34, N, 7.56. Found C,

51.56, H, 10.39, N, 7.96. 1H NMR ( 200 MHz, C6D6, 25 °C) δ 0.9 ( 12H, t, NCH2CH3 ),

δ 1.3 ( 6H, t, OCH2CH3 ), δ 2.5 ( 8H, q, NCH2CH3 ), δ 2.8 ( 4H, t, OCH2CH2NEt2 ) δ

4.7 (8H, q+q, OCH2CH3+ OCH2CH2NEt2). 13C{1H}-NMR (200 MHz C6D6 25 °C)

δ[ppm] 12.16 (N-CH2-CH3)2, 20.23 (OCH2CH3), 48.38 (N-CH2-CH3)2, 57.9

(OCH2CH2NEt2), 70.06 (OCH2CH2NEt2), 72.37 (OCH2CH3). Mass spectrometry. (EI+)

m/z remains speculative with 100% peak at 86 and highest m/z peak ( 2%) at 326.

Expected mass 370.25.

Page 54: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 43 -

2.3 Results and Discussion

2.3.1 Objectives of ligand design

The design and realization of new molecular MOCVD precursors offers a considerable

synthetic challenge. It is a challenging task in terms of complex chemistry but it provides

an inspiration for the construction of advanced materials, as well as it is a fascinating

research topic in its own right. Precursor design and synthesis involves new molecules or

known molecules with inclusion of design changes characterized by intricate structures,

optimized to their function for material applications.The most important precursor design

requirements that ideally must be satisfied are volatility, thermal stability, clean

decomposition, which are discussed in detail in section 1.6. It is highly unlikely that a

single precursor meets all the requirements for an ideal precursor. So designing a

precursor by including rational variations in precursor molecule according to process

requirements helps to a considerable extent in optimizing the process.

Volatility of a chemical precursor is based on weak attractive interaction between

molecules. In an ideal case monomeric, neutrally charged, complex is preferred. This is

because, higher molecular masses tend to reduce volatility; similarly polar groups and

polarizable groups also affect volatility.[12] The ideal ligand system for early transition

metals in a high oxidation state should have optimal steric bulk for stabilizing the metal

center. The selection of ligands with electron donor properties and optimal bulk not only

avoids oligomerization but also helps in stabilizing the electrophilic metal center. It is

well known that the volatility of a precursor is a complex function of intermolecular

forces (van der Waals interactions, p-stacking or hydrogen bonds) which not only depend

on the molecular weight and geometry but also (for solids) on the lattice structure.

Control of the polymerization in the solid state and in the vapor by optimizing the steric

bulk of the ligands, manipulation of ligands, represent the main directions for tailoring

volatility.

Indeed, the thermal behavior of these complexes is extensively influenced by ligand

dissociation reactions taking place during the transport into the vapor phase. The extent

of dissociation during or before evaporation depends on the stability of the complexes.

Page 55: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 44 -

The complexes based on monodentate donors such as alkoxides dissociate completely.

The partial substitution of alkoxide ligands by ß-diketonates, the latter acting essentially

as chelating ligands, could be a means to reduce polymerization and thus to enhance the

stability. Decreased susceptibility towards air and moisture could be an added advantage.

2.3.2 Precursor synthesis and properties

The most common precursors for metal oxide are compounds which already contain an

M-O linkage, namely alkoxides [M(OR)n]m, carboxylates [M(O2CR)n]m or ß-diketonates

[M(OCRCHCR´O)n]m where n = oxidation state and m = molecular complexity.[4] These

three ligand types are considered versatile because of the backbone flexibility and the

variations possible in substituents. During the course of this study alkoxides and ß-

Fig 2.1: Molecular structure of [Ti(OPri)2(meaoac)2]. Important bond lengths [? ] and bond angles [deg] are listed here.

Bond lengths (Å) Bond Angles (deg.)

Ti-O(11) 2.126(4) O(16)-Ti-O(26) 100.83(17) Ti-O(25) 1.970(4) O(15)-Ti-O(25) 161.11(15) Ti-O(16) 1.775(4) O(21)-Ti-O(11) 78.99(14) Ti-O(21) 2.114(4) O(26)-Ti-O(21) 90.79(16) Ti-O(26) 1.782(4) C(24)-O(25)-Ti 133.5(4) O(11)-C(12) 1.224(6) O(16)-Ti-O(15) 98.33(18) C(12)-C(13) 1.412(8) O(15)-Ti-O(11) 82.72(15)

Page 56: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 45 -

ketoesters were employed for synthesizing the titanium complexes, hence details of

carboxylates are not discussed.

Metal alkoxides contain at least one M-O-C structural unit. Titanium with a Pauling

electronegativity of 1.5 does not have strong polar bonds between metal and oxygen. As a

result, alkoxides of titanium have limited nuclearities and simple titanium tetraalkoxides

are all volatile, including the titanium tetramethoxide, which is a solid at room

temperature that sublimes at 190 °C.[13] The demands for large ligands for minimizing the

nuclearity and low molecular weight are conflicting to each other and not easily achieved.

Titanium tetraisopropoxide is the most volatile of the titanium alkoxides a property

attributed to its low molecular weight and the steric effect of the ligand which restricts

nuclearity (molecular complexity) of the complex.[14] It is accepted that titanium

Fig. 2.2: Molecular structure of [Ti(OPri)2(tbaoac)2]. Important bond lengths [? ] and bond angles [deg] are as follows

Bond lengths (Å) Bond Angles (deg.)

Ti-O(16) 1.784(2) O(15)-Ti-O(25) 161.85(10) Ti-O(21) 2.118(2) O(26)-Ti-O(11) 167.78(10) Ti-O(26) 1.787(3) O(16)-Ti-O(26) 101.41(11) Ti-O(15) 1.976(3) O(26)-Ti-O(25) 100.07(12) C(14)-O(15) 1.286(4) O(25)-Ti-O(11) 83.29(10) O(11)-C(12) 1.249(4) O(11)-Ti-O(21) 78.68(9) C(12)-O(121) 1.337(4) C(12)-O(121)-C(121) 122.1(3) O(121)-C(121) 1.486(4) C(14)-C(13)-C(12) 122.3(4) C(22)-C(23) 1.418(5) C(14)-O(15)-Ti 134.4(2) C(23)-C(24) 1.358(5)

Page 57: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 46 -

tetraisopropoxide has a nuclearity of approximately 1.2 by earlier studies.[15] Titanium

tetraethoxide is a tetramer in solid state and a trimer in solution.[16] These two alkoxides

are used in the present study for inducing optimum volatility to the resulting precursor

complexes.

Fig. 2.3: Molecular structure of [Ti(OEt)2(tbaoac)2]. The dotted line shows the bond for

the disordered methyl group. Important bond lengths [? ] and bond angles [deg]

are listed here.

Bond lengths (Å) Bond Angles (deg.)

Ti-O(5) 2.130(4) O(3)-Ti-O(6) 160.98(17)

Ti-O(52) 1.788(4) O(52)-Ti-O(2) 88.77(17)

Ti-O(51) 1.797(4) C(2)-O(2)-Ti 128.7(4)

Ti-O(6) 1.966(4) O(51)-Ti-O(2) 170.85(19)

Ti-O(3) 1.966(4) O(3)-Ti-O(2) 82.68(17)

O(4)-C(7) 1.342(7) C(4)-O(3)-Ti 131.9(4)

O(4)-C(6) 1.481(7) O(2)-Ti-O(5) 81.49(15)

C(7)-C(8) 1.406(8) O(6)-Ti-O(2) 84.26(17)

C(51)-C(52) 1.351(18)

C(51)-C(52') 1.29(2)

Page 58: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 47 -

The ß-ketoesters always exist in keto and enol forms in equilibrium with each other [17] as

shown in the scheme 2.1.

O O

R1

R2

OR3

O O

R1

R2

OR3

H O O

R1

R2

O

H R3

Scheme 2.1.

The methane proton in the keto form and the hydroxyl proton in the enol form of ß-

ketoesters are acidic and their removal generates 1,3-ketoesterate anions which are source

of an extremely broad class of coordination compounds. 1,3-ketoesterate anions are

powerful chelating species and form complexes with virtually every transition and main

group element. There has been a large scope for this chemistry and has been reviewed

several times.[18-21]

The reaction between a ß-ketoester and tetraalkoxy-titanium dates back to more than half

a century.[22,23] The first two studies on such type of reaction could only predict the

formation of a desired compound but failed to either isolate the pure compound or to

provide convincing analytical data. But these studies could shed some light on the

possibility and access to such compounds. Subsequently, during the same decade the first

convincing structural study based on IR spectroscopy and cryoscopy could predict the

structure to be monomeric and octahedral coordination around titanium metal center.[24]

It remained unclear whether the complex has cis-substituted or trans-substituted

structure. However, Bradley and Holloway conclusively ruled out the possibility of the

trans configuration in favor of cis configuration. Based on proton NMR studies they

found that all of the derivatives of [Ti(ß-diketonate)2(OR)2] type complexes existed only

in the cis (optically active ) form over a wide range of temperatures. Activation energies

for intramolecular exchange of ligands in these fluxional molecules showed that steric

hindrance of the alkoxy group increased the energy of activation but did not promote the

trans- form [25] In the case of reactions of ß-diketonates or ß-ketoesters with titanium

alkoxides, bis-ß-diketonate or ketoester derivatives were the final products even when the

excess of these ligands were used. The non-replaceability of the third or fourth alkoxy

Page 59: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 48 -

groups with ß-diketonates or ketoesters may most probably due to the preferred

coordination number of 6 for titanium in the bis –derivatives and has been reported

earlier [24]

Titanium prefers a coordination number six to form stable complexes. Complexes having

coordination number four are sensitive towards hydrolysis by moisture. This is evident in

most of the tetra-alkoxy compounds of titanium. Alkoxy ligands being monodentate,

bond through oxygen atom to the metal center. ß-ketoester and ß-ketoamides act as

bidentate ligands. Reactions of tetra alkoxy titanium with two equivalents of ß-ketoesters/

ß-ketoamides resulted in stable six coordinated complexes. The representative scheme of

such a reaction is given in schemes 2.2 and 2.3 respectively.

Fig. 2.4: Molecular structure of [Ti(OPri)2(deacam)2].

Important bond lengths [? ] and bond angles [deg] are listed here.

Bond lengths (Å) Bond Angles (deg.)

Ti-O(115) 1.8086(16) O(111)-Ti-O(110) 100.84(7) Ti-O(111) 1.8178(16) O(15)-Ti-O(11) 82.58(6) Ti-O(16) 2.0394(15) O(115)-Ti-O(16) 169.97(7) Ti-O(11) 2.0737(16) O(115)-Ti-O(111) 99.32(7) C(17)-N(17) 1.345(3) C(123)-N(12)-C(121) 117.29(19) N(17)-C(171) 1.464(3) O(11)-C(12) 1.271(3) C(14)-O(15) 1.299(3)

Page 60: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 49 -

O O

N

Ti(OPri)4 + 2 O

O

N

Ti

O

O NO O

+ 2 HOPri

Complex 5 Scheme 2.3

Comparison of the pKa values of ß-ketoester with a series of substituted ß-diketonates

shows that the Lewis acidity at the metal center increases, moving from ß-diketonates to

ß-ketoesters and malonates.[26] In such a situation, it is likely that ß-ketoester can undergo

a facile trans-esterification reaction in the presence of catalytically active Lewis acidic

centers, such as titanium metal. Interestingly during our studies the use of malonic ester

as ligand led to trans-esterification reaction and ligand moieties were exchanged between

two types of ligands surrounding the metal. On the contrary when ß-ketoesters were used

as ligands we did not encounter such type of reactions. Interestingly, the reaction with

malonic esters follows a different path altogether. It was hoped that use of malonates,

being bulky ligands, stabilizes the metal center by restricting the nuclearity to monomer.

So the ratio of metal to ligand was altered to favor monomeric products, but interestingly

these efforts resulted in the same dimeric complex 6. The reaction of malonates with

titanium alkoxides is shown in the reaction scheme 2.4.

O O

R2 OR3

Ti(OR1)4 + 2 + 2 R1OH

R1=iPr R2, R3 = CH3 Complex 1

O

O

R2

OR3

O

O

R2

OR3Ti

R1O OR1

R1=Et R2, R3 = CH3 Complex 2

R1=iPr R2, R3 = tBut Complex 3

R1= Et R2, R3 = tBut Complex 4

Scheme 2.2

Page 61: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 50 -

O O

O OTi

O

O

O

O O

O

O

O

O

O

O

OO

O

2Ti(OPri)4 +2 + 2 EtOH

Complex 6 Scheme 2.4

Fig. 2.5: Molecular structure of [Ti2(µ-OEt)2(OPri)4(deml)2].

Important bond lengths [? ] and bond angles [deg] are listed here.

Bond lengths (Å) Bond Angles (deg.)

Ti(1)-O(6) 1.791(3) O(6)-Ti(1)-O(5) 97.09(13) Ti(1)-O(7) 1.802(3) O(6)-Ti(1)-O(7) 96.44(14) Ti(1)-O(5) 2.027(3) O(6)-Ti(1)-O(1) 86.54(12) Ti(1)-O(1) 2.109(3) O(5)-Ti(1)-O(1) 81.38(12) Ti(1)-O(8) 2.080(3) O(7)-Ti(1)-O(1) 172.53(12) Ti(1)-Ti(1) 3.2113(16) O(8)-Ti(1)-O(5) 157.89(12) O(41)-C(41) 1.450(6) O(8)-Ti(1)-O(1) 84.55(11) C(2)-C(3) 1.391(6) O(7)-Ti(1)-O(8) 101.69(12) O(1)-C(2) 1.250(5) C(4)-C(3)-C(2) 120.3(4) C(4)-O(5) 1.262(5)

Page 62: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 51 -

Due to the presence of catalytically active Lewis acidic metal center Ti, the trans-

esterification reaction was observed. Logic for this can be derived from the fact that

malonate ligands have higher pKa values compared to ß-ketoesters and thus render the

metal center more Lewis acidic. The octahedral coordination of titanium results from

bridging of alkoxide ligands of low steric hindrance, i.e. in this case, ethoxide moieties

drawn from the periphery of the malonate ligand used in the reaction. The reaction is

facile and slightly exothermic at room temperature.

The compounds with ß-ketoesters/malonates/amides were off white to pale yellow in

appearance and were crystalline solids. The purified products were sublimed at moderate

temperatures and the melting points were below 90 °C. In addition the complexes were

highly soluble in organic solvents. All these factors make these compounds attractive for

MOCVD applications. Donor functionalized ligands of the type amionoalkoxides, where

the amino group acts as a donor is used widely for sol gel processes.[12] Donor

functionalized ligand are assumed to be less easily separated from the metal center by

hydrolysis than monodentate ligands because of chelate effect. Scheme 2.5

O

N O

N

RO OR

HO

NTi(OR)4 + 2

-2HOR

Complex 7 R = OPri

Complex 8 R= OEt Scheme 2.5

The use of such type of ligands to synthesize titanium complexes has been demonstrated

earlier and employed for TiO2 deposition.[27] The synthesis of aminoalkoxide classes of

compounds was simple and was reproducible in high yields. The complexes of

aminoalkoxides were viscous liquids which form glassy solid at -77 °C. The higher

temperatures required (>150 °C) for vaporizing these complexes deprived their use as

MOCVD precursors but perhaps suitable for liquid injection CVD. Nevertheless, these

Page 63: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 52 -

complexes have been used to synthesize TiO2 nano particles using a non hydrolytic

approach.[28]

2.3.3 NMR studies

The room temperature 1H and 13C NMR spectra of compounds 1-8 except 6 in benzene-d6

are structurally significant without any dynamic solution behavior. Sharp singlets were

observed for the protons of methyl, tert. Butyl and methyne of the tbaoac ligand. A septet

(2H) as well as a doublet (12H) refers to the isopropoxide ligands. Similarly a quartet

(4H) and a triplet (6H) were clearly visible for ethoxide complexes. The 13C NMR was in

good agreement with the proposed monomeric structures as well. As a typical example

the 1H NMR spectrum for the compound [Ti(OEt)2(tbaoac)2] is depicted in Fig. 2.6.

Complex 6 showed very complex NMR behaviour. 1H and 13C-NMR spectra, taken from

the crystals in benzene-d6 or toluene-d8 are complex and could be hardly interpreted. The

reason is that 6 undergoes highly fluxional exchange reactions of the alkoxy groups in

solution and probably forms species with different nuclearities. 1H spectra of 6 in

toluene-d8 at temperatures between -40 °C and 65 °C failed to provide the solution for the

ppm (t1)1.02.03.04.05.06.07.08.0

-CH-

-CH2

-

-CH3

tBu

-CH3

Fig. 2.6: 1H NMR spectrum of the compound [Ti(OEt)2(tbaoac)2] in benzene-d6

Page 64: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 53 -

complexity. Only four peaks could be identified which were readily assigned to

isopropoxy group and ethoxy group in an agreeable resolution.

2.3.4 Single crystal X-ray diffraction analysis

Data collections for compounds 1, 3, 4, 5 and 6 were performed on a Bruker AXS CCD

1000 diffractometer, equipped with a cryogenic nitrogen cold stream to prevent loss of

solvent and using graphite monochromated Mo-Ka radiation (0.71073 ? ). The crystals

were mounted on glass capillaries. The overall molecular geometries with atomic labeling

are illustrated in Figs. 2.1, 2.2, 2.3, 2.4 and 2.5 respectively. Relevant crystallographic

details are summarized in Table 2.1. The structures were solved by direct method using

SHELXL-97[29] software package and refined by full matrix least-squares methods based

on F2 with all observed reflections. Empirical absorption corrections were applied to the

data by SADABS (version 2.03).[30] All non hydrogen atoms were refined anisotropically

and all the hydrogen atoms were placed in calculated positions.

Single-crystal X-ray diffraction experiments reveal that the titanium complexes (1, 3, 4

and 5) formed with the ß ketoesters/ ß-ketoamides as ligands, are monomeric, a property

which is preferred for better volatility. Crystals of complex 2 failed to give enough

reflections for determining the structure. The structures of the four compounds 1, 3, 4 and

5 carry similar features. The mononuclear crystal structures with ß-diketo systems are

rarely reported in the literature. Though there are several studies wherein based on NMR

studies the structure was predicted to be monomeric. Because of similarity in structural

features, the general observations will be discussed.

Page 65: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 54 -

Compounds are 1, 3, 4 and 5 are mononuclear with octahedral geometry surrounding the

metal center. While compounds 1, 3 and 5 crystallize in the monoclinic space group

P2(1/n), the compound 4 crystallizes in orthorhombic, P2(1)2(1)2(1) space group and

form a distorted six coordinated environment at the titanium center (Fig. 2.3). The two ß-

ketoester ligands as well as the two alkoxy ligands are arranged cis to each other, with

both ester moieties (amide moiety in case of ß-ketoamides) arranged trans to the alkoxy

ligands. In all complexes it was observed that one of the oxygen to metal bond is shorter

by about 0.2 Å. This is due to trans effect induced by alkoxy ligands on these bonds. So

the bond angles O-Ti-O always deviate from the ideal octahedral angles of 90°. The

deviation varies between 80° to 100° in different complexes. In all cases, it was observed

that ß-ketoesters/ß-ketoamides tend a small bite angle (~ 82°) and these result in

expansion of external angle O-Ti-O subtended by oxygens of alkoxy groups (~100°).

Fig. 2.7: Distorted octahedral geometry surrounding metal center. Bite angles

subtended by ß-ketoesters are smaller than 90°; in typical case (~83°) and

one of the metal-oxygen bonds from ß-ketoesters are smaller by 0.2 Å.

The cis arrangement of ligands around metal center was observed with

alkyl groups of side chains pointing away from each other.

Page 66: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 55 -

As a typical example, octahedral geometry surrounding the titanium center is depicted in

Fig. 2.7.

In the solid state, compound 6 (Fig. 2.5) exists as a dimer and crystallizes in space group

P2(1)/c. Ti is six coordinated in which each titanium atom is surrounded by two oxygen

atoms of chelating ß-ketoester group in an ?2 fashion, oxygen of two terminal

isopropoxide groups and by two µ-oxo atoms of bridging ethoxide groups. The chelating

ligands on each Ti atom are trans to each other. The geometry around each titanium

center is distorted octahedral (Fig. 2.8). A distance of 3.2113(16) Å exists between two

metal centers which are non bonding.

Bridging position is occupied by ethoxy moieties for the simple reason that they are

sterically least demanding. One more reason is that the bridging position, aside from

preferring the less-hindered ligand, also demands the more Lewis basic oxygen. In this

regard, oxygen of ethoxy moiety is readily available for saturation and thus preferred to

oxygen from malonates. Selected bond lengths and angles are reported in the figure

captions.

Fig.2.8: The distorted octahedral geometry around titanium centers. Centro

symmetric structure has malonate groups pointed in opposite directions.

Page 67: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 56 -

Molecular packing in the unit cell is an important factor which influences the thermal

property. Minimum molecular interactions in the unit cell are contributive to the volatility

of the complex. The ß-ketoesterate complexes reported above show similar packing

arrangements. As a representatitive example crystal packing for complex 3 is shown in

Fig. 2.9. It shows the projections of the three-dimensional packing arrangement of 3

along the b axis. It appears that the bending of the basal skeleton is related to the bulky

tert. Butyl group attached to the ß-keto functional group. And the bulky tert. Butyl groups

are found to point in the same direction and are away from the basal plane and always

away from each other. As a result, each individual molecule is packed in an alternated

layer structure, composed of a parallel arrangement of individual units. Moreover, the

observed curvature is a natural consequence of the face-to-back stacking of the

molecules, wherein, tert. Butyl groups point into sterically unencumbered regions.

Fig. 2.9: Crystal packing for 3, four units are packed in a cell. Crystals are arranged

with face to back type stacking.

Page 68: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 57 -

Table 2.1: Crystal data and refinement details of compounds 1 and 3.

Compound 1 3

Empirical formula C16 H28 O8 Ti C22 H40 O8 Ti

Formula weight 396.28 480.44

Crystal size [mm] 0.28 x 0.15 x 0.10 0.30 x 0.18 x 0.15

T,K 213(2) 213(2)

Crystal system Monoclinic Monoclinic

Space group P2(1)/n P2(1)/n

a, Å 7.753(5) 10.646(6)

b, Å 19.451(13) 16.019(6)

c, Å 13.751(6) 15.992(6)

alpha 90 90

beta 93.13(5) 90.44

gamma 90 90

V, Å3 2071(2) 2727(2)

Z 4 4

Calc. Density [g/cm3] 1.271 1.170

m (Mo-Ka) mm-1 0.448 0.352

F (100) 840 1032

Theta range [°] 2.83 to 25.19 1.80 to 25.04

Index range -9<=h<=9

-22<=k<=23

-11<=l<=16

-7<=h<=12

-18<=k<=19

-19<=l<=18

No. of reflections 11608 12176

Unique reflections 3675 4783

Observed reflections

[I>2 sigma (I)]

1763 2168

No. of parameters 239 292

Largest diff. peak/hole

[e Å-3 ]

0.411 and -0.448 0.286 and -0.329

R1, wR2 [I >2 (I)] 0.0708 / 0.1634 0.0632 / 0.1076

Goodness of fit 0.965 0.865

Page 69: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 58 -

Table 2.2: Crystal data and refinement details of compounds 4, 5 and 6.

Compound 4 5 6

Empirical formula C20 H36 O8 Ti C22 H42 N2 O6 Ti C34 H68 O14 Ti2

Formula weight 452.39 478.47 796.68

Crystal size [mm] 0.25 x 0.15 x 0.10 0.30 x 0.25 x 0.22 0.40 x 0.34 x 0.09

T,K 203(2) 100(2) 208(2)

Crystal system Orthorhombic Monoclinic Monoclinic

Space group P2(1)2(1)2(1) P2(1)/n P2(1)/c

a, Å 8.095(5) 16.7363(16) 11.971(4)

b, Å 15.964(7) 17.6227(13) 16.674(4)

c, Å 19.022(10) 18.3725(15) 11.029(5)

alpha 90 90 90

beta 90 102.085(8) 103.50(3)

gamma 90 90 90

V, Å3 2458(2) 5298.6(8) 2140.5(14)

Z 4 8 2

Calc. Density [g/cm3] 1.222 1.200 1.236

m (Mo-Ka) mm-1 0.386 0.358 0.431

F (100) 968 2064 856

Theta range [°] 2.73 to 25.07 2.75 to 25.00 2.13 to 27.50

Index range -9<=h<=9

-12<=k<=18

-13<=l<=22

-19<=h<=19

-20<=k<=20

-21<=l<=21

-15<=h<=15

-21<=k<=19

-13<=l<=14

No. of reflections 6112 77867 5662

Unique reflections 4285 9310 4607

Observed reflections

[I>2 sigma (I)]

2387 7652 2489

No. of parameters 283 579 248

Largest diff. peak/hole

[e Å-3 ]

0.321 and -0.307 1.565 and -0.441 0.295 and -0.363

R1, wR2 [I >2 (I)] 0.0653 / 0.1204 0.0578 / 0.1546 0.0749 / 0.1795

Goodness of fit 0.963 1.102 1.028

Page 70: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 59 -

2.4 Summary

The rational design, synthesis and characterization of several MOCVD precursors for

TiO2 have been carried out. Keeping in view of increasing the sublimation rates and

lowering the decomposition temperature, several ligands were used for synthesizing the

titanium complexes. Variations in the ligand sphere on the side chain of the well known

ß-keto systems have been carried out. ß-ketoesters, ß-ketoamides, malonates and

aminoalkoxides have been used as ligands for synthesizing the mixed alkoxide

complexes of titanium.

Titanium bis(isopropoxide) bis(methylacetoacetate) (1), titanium bis(ethoxide)

bis(methylacetoacetate) (2), titanium bis(isopropoxide) bis(tert-Butylacetoacetate) (3),

titanium bis(ethoxide) bis(tert-Butylacetoacetate) (4) and titanium bis(isopropoxide)

bis(N, N-diethylacetoacetamide) (5) and bis-[(di-ethylmalonato) tetra(isopropoxy)-µ-

ethoxy-titanium(IV)] (6) Titanium bis(isopropoxide) bis(diethylaminoethoxide) (7),

titanium bis(ethoxide) bis(diethylaminoethoxide) (8) were synthesized and characterized

by elemental analysis , NMR and mass spectrometry.

The use of ß-ketoesters, and ß-ketoamide ligands led to six coordinated complexes,

namely titanium bis(isopropoxide) bis(methylacetoacetate) (1), titanium

bis(isopropoxide) bis(tert-Butylacetoacetate) (3), titanium bis(ethoxide) bis(tert-

Butylacetoacetate) (4) and titanium bis(isopropoxide) bis(N, N-diethylacetoacetamide)

(5), as determined by single crystal X-ray diffraction. The six coordinated complexes

showed distorted octahedral geometry around titanium center. In such an arrangement,

the alkoxy groups positioned cis to each other and so are the ß-keto ester/ß-ketoamide

groups.

With the use of malonate ligands, the trans-esterification reaction was observed. This led

to the exchange of alkoxy groups from the malonate ligand periphery to that of parent

alkoxide. The resulting complex bis-[(di-ethylmalonato) tetra(isopropoxy)-µ-ethoxy-

titanium(IV)] (6) showed a dynamic behavior in NMR studies and could not be

understood even by variable temperature NMR studies. The single crystal X-ray

diffraction studies showed the formation of dimeric centrosymmetric complex. The

bridging positions were occupied by ethoxy moieties which are sterically least

Page 71: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 60 -

demanding. The complexes 1-6 were all solids and showed high solubility (~ 0.8 g/ml) in

common organic solvents. These complexes could be sublimed at moderate temperatures

and the melting points were below 90 °C.

Donor functionalized ligands of the type aminoalcohols, where amino group acts as a

donor, resulted in complexes titanium bis(isopropoxide) bis(diethylaminoethoxide) (7),

titanium bis(ethoxide) bis(diethylaminoethoxide) (8). The synthesis of these class of

compounds was simple and was reproducible in high yields. These complexes were

viscous liquids and attempts to measure X-ray diffraction at low temperatures (-77 °C)

resulted in glassy solids. It is demonstrated that by introducing small variations in the

ligand sphere, it is possible to change the physical and chemical properties of the

resulting complexes.

Page 72: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 61 -

2.5 References

[1] G. Hass, Vacuum 1952, 2, 331.

[2] R. N. Ghoshtagore, A. J. Norieka, J. Electrochem. Soc. 1970, 117, 1310.

[3] R. N. Ghoshtagore, J. Electrochem. Soc. 1970, 117, 529.

[4] D. C. Bradley, R. C. Mehrotra, D. P. Gaur, Metal Alkoxides, Academic Press,

New York, 1978.

[5] D. C. Bradley, Chem. Rev. 1989, 89, 1317.

[6] M. Balog, M. Schieber, J. Cryst. Growth 1972, 17, 298.

[7] H. Yamazaki, T. Tsuyama, I. Kobayashi, Y. Sugimori, Jpn. J. Appl. Phys. 1992,

31, 2995.

[8] P. Comba, H. Jakob, B. Nuber, B. K. Keppler, Inorg. Chem. 1994, 33, 3396.

[9] Y.-S. Min, Y.-J. Cho, D. Kim, J.-H. Lee, B.-M. Kim, S.-K. Lim, I.-M. Lee, W.-I.

Lee, Chem. Vap. Deposit. 2001, 7, 146.

[10] R. C. Smith, T. Ma, N. Hoilien, L. Y. Tsung, M. J. Bevan, L. Colombo, J.

Roberts, S. A. Campbell, W. L. Gladfelter, Adv. Mater. Opt. Electron. 2000, 10,

105.

[11] D. G. Colombo, D. C. Gilmer, V. G. J. Young, S. A. Campbell, W. L. Gladfelter,

Chem. Vap. Deposit. 1998, 4, 220.

[12] W. A. Herrmann, N. W. Huber, O. Runte, Angew. Chem. Int. Ed. Engl. 1995, 34,

2187.

[13] I. D. Varma, R. C. Mehrotra, J. Chem. Soc. (London) 1960, 2966.

[14] S. M. Damo, K. -C. Lam, A. Rheingold, M. A. Walters, Inorg. Chem. 2000, 39,

1635.

[15] D. C. Bradley, R. C. Mehrotra, W. Wardlaw, J. Chem. Soc. 1952, 5020.

[16] W. R. Russo, W. H. Nelson, J. Am. Chem. Soc. 1970, 92, 1521.

[17] R. L. Toung, C. Wentrup, Tetrahedron 1992, 48, 7641.

[18] R. C. Mehrotra, R. Bohra, D. P. Gaur, Metal ß-diketonates and allied derivatives,

Academic publishers New York, 1978.

[19] K. C. Joshi, V. N. Pathak, Coord. Chem. Rev. 1977, 22, 37.

Page 73: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 62 -

[20] D. P. Graddon, Coord. Chem. Rev. 1969, 81, 79.

[21] B. Bock, K. Flatau, H. Junge, M. Kuhr, H. Musso, Angw. Chem.Int. Ed. Engl.

1971, 10, 225.

[22] F. Schmidt, Angew. Chem. 1952, 64, 536.

[23] R. E. Reeves, L. W. Mazzeno, Jr., J. Am. Chem. Soc. 1954, 76, 2533.

[24] A. Yamamoto, S. Kambara, J. Am. Chem. Soc. 1957, 79, 4344.

[25] D. C. Bradley, C. E. Holloway, J. Chem. Soc., Chem. Comm. 1965, 13, 284.

[26] J. W. Bunting, J. P. Kanter, J. Am. Chem. Soc. 1993, 115, 11705.

[27] A. C. Jones, T. J. Leedham, P. J. Wright, M. J. Crosbie, K. A. Fleeting, D. J.

Otway, P. O´Brien, M. E. Pemble, J. Mater. Chem. 1998, 8, 1773.

[28] H. Parala, A. Devi, R. Bhakta, . J.Mater. Chem., 2002, 12, 1625.

[29] G. M. Sheldrick,, Program for crystal structure analysis ed., University of

Göttingen, Germany, 1997.

[30] G. M. Sheldrick, SADABS, Program for area detector adsorption correction,

Institute for Inorganic Chemistry, University of Göttingen, Germany, 1997.

Page 74: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 63 -

Chapter 3

Chemical vapor deposition of TiO2 an SrTiO3 thin films using

rationally developed titanium precursor

Abstract

The MOCVD of TiO2 thin films using newly developed precursors namely, Titanium

bis(isopropoxide) bis(tert-Butylacetoacetate) [Ti(OPri)2(tbaoac)2] (A), titanium

bis(ethoxide) bis(tert-Butylacetoacetate) [Ti(OEt)2(tbaoac)2] (B), and Titanium

bis(isopropoxide) bis(methylacetoacetate) [Ti(OPri)2(meaoac)2] (C), is discussed in

detail. Initial screening of the precursors was carried out using a homebuilt horizontal

cold wall CVD reactor. The films were analyzed by XRD, SEM, XPS and RBS. The

growth rates were determined based on weight gain method considering the bulk density

of anatase phase of TiO2. Comparison of growth rates at different deposition

temperatures were done under similar deposition conditions. One of the precursors,

[Ti(OPri)2(tbaoac)2] (A), which showed high solubility in common organic solvents was

tested in a liquid injection industrial MOCVD tool. The film depositions were compared

to the films obtained using a standard commercially available Ti precursor,

[Ti(OPri)2(thd)2]. Films were analyzed for crystallinity, surface roughness, surface

morphology, and electrical characteristics. It was found that new precursor efficiently

incorporates Ti into growing films at lower temperatures compared to standard precursor.

The surface roughness was lower in comparison with standard precursor. Equivalent

oxide thickness (EOT) and dielectric constant of the films were measured and it was

found that EOT of ~ 2 nm and dielectric constant of ~35 were obtained. After successful

testing for TiO2 depositions, the precursor [Ti(OPri)2(tbaoac)2] A, was tested for complex

oxide SrTiO3 depositions using standard [Sr(thd)2] precursor. Polycrystalline, off

stoichiometric films were obtained at a deposition temperature of 500 °C. Low deposition

temperatures (<600 °C) resulted in Sr rich films. At deposition temperatures above 600

°C stoichiometric Sr:Ti ratio could be observed resulting in crystalline films with (200)

and (110) orientation, exhibiting required electrical properties.

Page 75: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 64 -

3.1 Introduction

In an effort to explore new and improved titanium precursors by molecular engineering, a

systematic approach by varying the ligands by introducing small changes was pursued in

our laboratory. Use of ß-ketoester instead of ß-diketonate ligands in combination with

alkoxides of titanium resulted in mononuclear complexes (refer chapter 2). Titanium

bis(isopropoxide) bis(tert-Butylacetoacetate) [Ti(OPri)2(tbaoac)2] (A), titanium

bis(ethoxide) bis(tert-Butylacetoacetate) [Ti(OEt)2(tbaoac)2] (B), Titanium

bis(isopropoxide) bis(methylacetoacetate) [Ti(OPri)2(meaoac)2] (C), are three of the

newly developed precursors. The thermal properties of these compounds were promising

enough for MOCVD applications. All the three are low melting compounds; possess a

sufficient temperature window between volatility and decomposition. The onsets of

volatilization and decomposition temperatures were lower when compared to the ß-

diketonate derivatives. The compounds can be sublimed at low temperatures (< 100 °C)

under mild conditions (3.2 x 10-2 torr) which shows promise for MOCVD applications.

These compounds showed extremely high solubility in common organic solvents making

them suitable for liquid injection as CVD as well.

In the case of MOCVD processes, optimization of deposition conditions can be difficult

and time consuming given the large number of process variables involved.[1-4] In such a

situation, it is worthwhile to explore the possibility of depositions using lab scale reactors

and when the results are promising enough, the precursor can be scaled up and tested in

industrial scale reactors. Following this line the above mentioned precursors were tested

for TiO2 depositions using the homebuilt CVD reactor and results are reported in

following sections. Among the newly developed precursor complexes, titanium

bis(isopropoxide) bis(tert-Butylacetoacetate), [Ti(OPri)2(tbaoac)2] (A), showed the most

promising deposition behavior using the home built CVD reactor. Therefore the precursor

was scaled up to large batches (~ 25 g.) and was tested for TiO2 thin film deposition

using an industrial tool reactor. Complex oxides of titanium are important in many

ternary perovskites such as BaTiO3, SrTiO3 and the quaternary materials [Pb(Zr,Ti)O3]

(PZT),[(Pb,La)(Zr,Ti)O3] (PLZT) which are predicted to be useful in next generation

Page 76: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 65 -

non-volatile computer memories.[5-9] So after successful TiO2 depositions, the precursor

was tested for SrTiO3 depositions using standard Sr precursor [Sr(thd)2]. In an effort to

evaluate the performance of the new precursor A, the depositions were compared with

deposition results obtained using a commercially available bench mark titanium

precursor, namely titanium bis(isopropoxide) bis(tetramethylheptanedionate)

[Ti(OPri)2(thd)2]. The thin films were tested for their morphology and electrical

properties, keeping in view, their use for device applications.[10-17]

3.2 Experimental Section

3.2.1 Experiments using a home built CVD reactor

Thin films of TiO2 were generally grown on Si(100) substrates (1 cm x 1 cm). Prior to

film deposition the substrates were degreased in trichloroethylene and rinsed with

deionized water followed by rinsing in hot acetone and iso-propanol. They were dried

using pressurized air. The substrates were weighed before and after the depositions in

order to determine the weight gained due to film deposition and subsequently to calculate

the thickness of the film. A home built horizontal, cold wall, low pressure reactor was

used to deposit titanium dioxide (TiO2) thin films. A schematic diagram of the reactor is

shown in Fig. 3.1. The CVD reactor consisted of a quartz tube, at the center of which the

substrates are placed on a SiC coated graphite susceptor. Substrate heating was

accomplished by an inductive heating arrangement attached to a radiation pyrometer. The

quartz tube was attached to a glass vaporizer (bubbler) by means of O-ring joints, and the

vaporizer was placed in an air bath that can be heated up to 150 °C. The precursor was

filled into the vaporizer in a glove box and generally about 200 mg. of freshly filled

precursor was used for each deposition. The path between the precursor evaporator and

deposition chamber is maintained as short as possible to avoid any decomposition in the

reactor lines or on the walls. This path is maintained at precursor evaporator temperature

by using an air bath. The reactor chamber is surrounded by water jacket operating

approximately at 90 °C to avoid the condensation of the precursor. Thin film depositions

were generally carried out at reduced pressure, and therefore a turbo molecular pump was

used for this purpose. The reactor pressure was controlled using a motor driven throttle

Page 77: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 66 -

valve and the deposition pressure was varied between 0.1 -100 mbar. Typically, a base

pressure of 2.8 x 10-5 mbar could be achieved. The gas flow in the reactor was monitored

and controlled by mass flow controllers (MFC). Prior to film deposition, the substrates

were cleaned by heating the substrate to 800 °C for about 30 minutes under N2 using a

bypass line. Then temperature was lowered to the required value. The onset of deposition

was marked by opening the vaporizer valves and bubbling nitrogen as the carrier gas. A

typical deposition lasted for 90 minutes. At the end of 90 minutes, the valves (inlet and

outlet) were closed and substrates were slowly cooled to room temperature under flowing

nitrogen (using bypass line).

The substrates were removed carefully and stored for analyses. The substrates were

weighed again to estimate the weight again due to film deposition. A microbalance

delivering up to five decimal values of a gram was used for weight measurements. Bulk

density ? = 3.94 g/cm-3 was assumed to determine average thickness of the TiO2 thin

films using the equation

AWt ρ/∆= (equation 3.1)

MFC MFC

Turbomolecular pump

Cooling trap

Valve for pressure control

Air bath

SiC coated graphite susceptor

Inductive heating

Substrate

Water cooling

Reactive gas Carrier gas

Capacitance manometer

Pressure gauge

Precursorreservoir

MFC MFC

Turbomolecular pump

Cooling trap

Valve for pressure control

Air bath

SiC coated graphite susceptor

Inductive heating

Substrate

Water cooling

Reactive gas Carrier gas

Capacitance manometer

Pressure gauge

Precursorreservoir

Fig. 3.1: Schematic diagram of the homebuilt CVD reactor used for TiO2 thin film

deposition

Page 78: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 67 -

where ? W is the weight of the deposited film in grams and A is the surface area of the

substrate.

The reactor was cleaned with acetone and iso-propanol and loaded with a new set of

substrates and the vaporizer was charged with fresh precursor and the reactor was

evacuated.

Table 3.1: Deposition parameters for precursors A, B and C using home built CVD

reactor.

X-ray diffraction analyses were carried out on all of the deposited films, employing a D8-

Advance Bruker axs diffractometer. CuKa radiation (? = 1.5418 Å) with Nickel filter was

used as the source. High angle XRD measurements were carried out with ?-2? geometry

in the range 20-60° using a position sensitive detector. The film composition was

analysed by energy dispersive analysis of X-rays (EDAX) and Rutherford back-scattering

(RBS) and X-ray photoelectron spectroscopy (XPS). The RBS spectra were measured

with the 2 MeV single charged He-beam of the 4MV Dynamitron-Tandem laboratory in

Bochum. A beam intensity of about 20 nA incident to the sample perpendicular to the

surface was used. The back-scattered particles were measured at an angle of 170° by a Si-

detector with a resolution of 16 keV. The film thickness was also determined using the

RBS data. The surface morphology and composition of the films was studied using a

scanning electron microscope attached with EDX system. This facility was provided by

faculty of Geology, Ruhr University, Bochum, employing LEO Gemini SEM 1530. X-

ray photoelectron spectroscopy analyses were carried out with a modified Fisons X-ray

Precursor

Vaporizer

temperature

(°C)

Deposition

temperature

(°C)

Deposition

time

(min)

Reactor

pressure

(mbar)

A [Ti(OPri)2(tbaoac)2] 85-90 350-800 90 10 & 50

B [Ti(OEt)2(tbaoac)2] 85-90 350-800 90 10 & 50

C [Ti(OPri)2(meaoac)2] 85-90 350-800 90 10 & 50

Carrier gas N2 = 100 sccm, substrate = Si(100)

Page 79: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 68 -

photoelectron spectrometer equipped with an Al Ka X-ray source and a CLAM3 electron

energy analyzer. Substrates were loaded into the chamber and evacuated overnight for 12

hours at 10-9 mbar. Survey X-ray photoelectron spectra and high resolution XPS spectra

were collected for desired elements.

3.2.2 Experiments using an industrial tool CVD reactor

System description

Most of the experimental reactors currently used for the development of mass production

tools use the conventional single wafer showerhead designs.[18-19] In contrast to the

current trend towards single substrate reactors, which can be integrated with cluster tools

along with other manufacturing process steps, present AIXTRON systems use the

Planetary Reactor®. These planetary systems are capable of offering extremely high

throughput due to possible batch mode processing resulting in low cost of ownership.

This reactor type was already well established, because it offers good homogeneity, high

efficiency of the precursors. For the deposition of complex-oxide films like SrTiO3, the

reactor is combined with a liquid delivery system from ATMI (LDS-300B).

The vaporizer is placed in the immediate vicinity of the reactor. The line to the reactor

was kept as short as possible and held at the evaporation temperature in order to avoid

condensation of the precursors which results into particle generation and subsequent yield

loss. As shown in Fig. 1.5 the wafers are placed on a coated graphite susceptor that

rotates typically at 8 rpm and carries five smaller plates (satellites) which rotate at 40 rpm

by gas foil rotation. The gas inlet is placed central above the reactor providing a pure

horizontal gas flow direction which makes this reactor a radial flow system. The oxidizer

gases enter the reactor just below the nozzle separate from the precursors in order to

avoid premature reactions. The homogenous heating of the graphite susceptor is achieved

by infrared lamps positioned below the rotating disk.

The desorbed products as well as the non reacted molecules are transferred through a ring

line, into a cold trap. The reactor operates under low pressure at ~2 mbar in order to

increase the gas diffusivity and prevent pre-reactions. The pumping system consists of a

Page 80: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 69 -

A, [Ti(OPri)2 (tbaoac)2] 0.05 mol in butyl acetate

Precursor solutions

[Sr(thd)2] 0.05 mol in butyl acetate

Injection pulse length and period

0.8 ms and 0.32 s

Argon carrier flow 1000 sccm

Vaporization temperature 170-240°C

Process pressure 1-1.5 mbar

Susceptor temperature 350-800°C

O2 flow 200 sccm

root or booster pump and a rotary pump for pumping larger gas loads like the carrier gas.

The length of the injection pulses was kept at 0.8 ms, which corresponds to 5 µL/ pulse

and the delay between pulses was 160 ms. The deposition characteristics of the new

precursor were compared to those of a commercial [Ti(OPri)2(thd)2] precursor.

[Ti(OPri)2(tbaoac)2] was dissolved in n-butyl acetate (0.05 molar solution) and oxygen

(200 sccm) was used for the deposition of TiO2 films. Deposition pressure was in the

range 1 to 1.5 mbar. Films were deposited on Pt/ZrOx/SiOx/Si and on SiOx/p-Si

substrates. The process conditions are summarized in table 3.2.

Table 3.2: Standard MOCVD deposition conditions used for deposition of TiO2 and

SrTiO3 thin film depositions.

X-ray fluorescence (XRF, RIGAKU ZSX-100e) was used for the determination of molar

amount of the individual element in the deposited films. The Ti incorporation in the films

was determined by X-ray fluorescence (XRF) and film thickness was deduced from the

measured areal mass density of Ti atoms by assuming the density of TiO2 anatase phase,

3.8 g/cm3. Films were deposited on Pt/ZrOx/SiOx/Si substrates. Crystal structure of the

films was studied with the aid of X-ray diffraction (Philips Analytical) employing grazing

angle and Bragg-Brentano geometry with Cu-Kα radiation. Surface morphology of the

Page 81: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 70 -

films was studied with AFM (SIS pico station). Electrical properties of the metal-

insulator-semiconductor (MIS) structures, Capacitance-Voltage (C-V) characteristics

were obtained using HP4284 LCR meter by sweeping the voltage from inversion to

accumulation and back.

3.3 Results and Discussion

3.3.1 Deposition of TiO2 thin films using homebuilt CVD reactor

The rationally developed precursors were screened for MOCVD applications using a

home built cold wall CVD reactor. Inclusion of ester moiety in the side chain of the ß-

keto structure is expected to act as a cleavage point and facilitate lower deposition

temperatures (refer chapter 2). During the course of this work, it was found that the

complexes with small variations in the ligand sphere showed significant differences in the

hydrolytic stability and thermal behavior. These observations drove us to test a series of

precursors with small variations for CVD of titanium dioxide thin films using the home

built reactor.

Fig. 3.2 shows the TG curves for three different precursors A, B and C used in this study

and is compared with standard precursors. It can be seen from the TG curves, that the

onset of volatilization for the rationally developed precursors (A,B,C) occur at lower

temperatures compared to the bench mark precursor [Ti(OPri)2(thd)2]. The vaporization

temperature is lowered by about 100 °C and a monotonic weight loss is observed. The

residue left behind in the case of precursor A is negligible (<2%) indicating a clear

volatilization of the precursor. (More details on the thermal properties of the rationally

developed precursors are given in chapter 5).

Page 82: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 71 -

100 200 300 400

0

20

40

60

80

100

CBA

Wei

ght [

%]

Temperature [°C]

[Ti(OPri)2(tbaoac)

2]

[Ti(OEt)2(tbaoac)

2]

[Ti(OPri)2(meaoac)

2]

[Ti(OPri)2(thd)

2]

[Ti(OPri)4]

Fig. 3.2: Thermogravimetric analysis of three different precursors A, [Ti(OPri)2(tbaoac)2],

B, [Ti(OEt)2(tbaoac)2], and C, [Ti(OPri)2(meaoac)2] compared with parent

alkoxide [Ti(OPri)4] and bench mark titanium precursor [Ti(OPri)2(thd)2].

3.3.2 Deposition parameters

Thin films of TiO2 were deposited in the temperature range 350-800 °C and at two

different reactor pressures (10 and 50 mbar). The deposition conditions employed for

TiO2 film deposition are listed in table 3.1. No additional reactive gas (oxygen) or water

vapor used during film deposition. The vaporizing temperatures for all the three

precursors were maintained at similar temperature range, so we expect the uniformity in

the transport rate to the reactor zone. The films grown at 350 °C were dark in color and

dull in appearance whereas films grown at temperatures above 400 °C had a shiny

appearance with a bluish green color. The highest growth rate was observed in case of

precursor A [Ti(OPri)2(tbaoac)2], at 500 °C where the growth rate was of the order of 0.6

µm/hr.

Page 83: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 72 -

3.3.3 Crystal structure of the films

The crystalline properties of the film were examined by X-ray diffraction (XRD). The

XRD analyses of TiO2 films grown at different temperatures are shown in Figs. 3.3 to

3.5. Films grown at 350 °C did not show any reflexes indicating that the films were

amorphous in nature. The temperature onset for crystallization was 400 °C as seen the

appearance of (111) reflection corresponding to the anatase phase of TiO2. With increase

in substrate temperatures, the films were more crystalline and the anatase phase was

predominant till 600 °C. The appearance of rutile phase in addition to anatase phase was

evident at temperatures above 700 °C. It should be noted that oriented and textured films

of TiO2 with the rutile phase (110) were observed at 800 °C. A complete phase transition

Fig. 3.3: X-ray diffraction patterns of TiO2 films with [Ti(OPri)2(tbaoac)2] precursor on

Si (100) substrates deposited under various susceptor temperatures.

20 30 40 50 60

Inte

nsity

(ar

b. u

nits

)

2 theta [deg]

400 °C

rutil

e (2

11)

rutil

e (2

20)

anat

ase

(200

)

Si(2

11)

anat

ase

(200

)

rutil

e (1

10)

anat

ase

(101

)

Si(2

11)

800 °C

500 °C

700 °C

600 °C

Page 84: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 73 -

to a more stable phase occurs at this temperature. The reflexes remain broad along the

series probably due to the small crystallite size.

Precursors B and C were also screened for the TiO2 film depositions. The XRD plots of

TiO2 films grown, using precursors B and C are shown in Figs. 3.4 and 3.5 respectively.

As in the case of precursor A, the TiO2 films were amorphous below 400 °C. Above 400

°C, the films were polycrystalline with anatase phase dominant till about 600 °C in case

of precursor B (Fig. 3.4). The rutile phase begins to appear and at 700 °C and 800 °C

rutile phase is predominant with some traces of anatase with lowered intensity.

However, in contrast to precursor A and B, the films grown using precursor C consisted

of brookite phase (201) and (111) at 400 °C. At higher temperature the reflexes of anatase

and rutile were observed consistent to those obtained from precursors A and B.

20 30 40 50 60

400 °C

500 °C

600 °C

Inte

nsity

(ar

b. U

nits

)

2 theta [deg]

rutil

e (2

11)

anat

ase

(101

)

rutil

e (1

01)

rutil

e (1

10)

anat

ase

(004

) rutil

e (1

11)

700 °C

800 °C

Fig. 3.4: X-ray diffraction patterns of TiO2 films with [Ti(OEt)2(tbaoac)2] precursor on Si

(100) substrates deposited under various susceptor temperatures.

Page 85: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 74 -

3.3.4 Effect of substrate temperature on growth rate

Film growth by CVD is strongly dependent on substrate temperature, not only because

surface reactions are temperature-activated, but also because reaction pathways could be

altered by temperature. Therefore, the growth rate of TiO2 films studied as a function of

substrate temperatures for A [Ti(OPri)2(tbaoac)2], B [Ti(OEt)2(tbaoac)2] and C

[Ti(OPri)2(meaoac)2]. Fig. 3.6 shows the Arrhenius plot of growth rate with different

precursors.

Three precursors behave differently along the series. It was observed that precursor A

[Ti(OPri)2(tbaoac)2] has maximum growth rate at 500 °C (~ 0.6 µm/hour) and thereafter

the growth rate decreases to ~ 0.3 µm/hour at 800 °C. Precursors B [Ti(OEt)2(tbaoac)2]

and C [Ti(OPri)2(meaoac)2] have maximum growth rates at 600 °C, 0.45 µm/hour and

20 30 40 50 60

broo

kite

(201

)

anat

ase(

101)

rutil

e(22

0)

anat

ase(

112)

Si(1

11)

Si(2

11)

rutil

e(11

0)

anat

ase(

200)

broo

kite

(111

)800 °C

400 °C

600 °C

2 theta [deg]

Inte

nsity

[arb

. Uni

ts]

700 °C

500 °C

Fig. 3.5: X-ray diffraction patterns of TiO2 films with [Ti(OPri)2(meaoac)2] precursor on

Si (100) substrates deposited under various susceptor temperatures.

Page 86: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 75 -

0.25 µm/hour respectively. While the growth rate of precursor B [Ti(OEt)2(tbaoac)2]

decreases rapidly to 0.11 µm/hour at 800 °C, that of precursor C [Ti(OPri)2(meaoac)2] is

not so rapid and has a growth rate of 0.25 µm/hour. It has to be noted that growth rates

were calculated based on density of anatase phase of the films and assuming that the

depositions were uniform over the surface of the substrate. Since the films contain rutile

phase and some carbon impurities, the growth rates determined by other methods may

deviate from these results.

0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7

0.1

0.2

0.3

0.4

0.5

0.6

Gro

wth

rat

e [µ

m/h

our]

1000 /T [K-1]

[Ti(OPri)2(tbaoac)

2]

[Ti(Et)2(tbaoac)

2]

[Ti(OPri)2(meaoac)

2]

Fig. 3.6: Arrhenius plot of growth rates with precursors, A [Ti(OPri)2(tbaoac)2],

B [Ti(OEt)2(tbaoac)2] and C [Ti(OPri)2(meaoac)2].

Page 87: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 76 -

1200 1000 800 600 400 200 00

10000

20000

30000

40000 Ti LMM

O KLL

O 2S

Ti 3STi 3P

C 1S

Ti 2P

O 1S

Ti 2S

Inte

nsity

[cou

nt/S

ec]

Binding Energy [eV]

XPS Survey spectrum

470 465 460 455 450

1500

2000

2500

3000

Ti 2p(3/2)

Ti 2p(1/2)

Inte

nsity

[Cou

nts/

Sec

]

Binding Energy [eV]

XPS Ti 2p

Fig. 3.7: A XPS survey and high resolution spectra of TiO2 thin film grown on Si(100) at

700 °C using precursor A [Ti(OPri)2(tbaoac)2] (b) Titanium Ti2p region (c) O1s

region

536 534 532 530 528 526 524

2000

2500

3000

3500

4000

Inte

nsity

[Cou

nts/

Sec

]

Binding Energy [eV]

XPS O 1S

(a)

(b) (c)

Page 88: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 77 -

3.3.5 Film composition

Routine film composition analyses could not be done for all films and hence only results

from representative samples are reported here. Fig. 3.7 (a) shows a XPS survey spectrum

of a sputtered-cleaned TiO2 film grown using precursor A [Ti(OPri)2(tbaoac)2], at 700 °C

where the signals can be attributed to titanium, oxygen and some traces of carbon.

Detailed spectra for the titanium Ti 2p region and for the oxygen O1s region are shown in

Fig. 3.7 (b) and 3.7 (c) respectively. A gauss peak-fit gives binding energies of 530.4 eV

for O1s and 458.9 eV for Ti 2p, in agreement with the values reported for TiO2 in the

literature.[ 21]

0 200 400 600 800 1000-500

0

500

1000

1500

2000

2500

3000

3500

4000

O - edge at interface and Si - edge (substrate)below threshold

Ti - edge (surface)

O - edge

Ti - edge (interface)

Pulser

Experimental curve Simulated curve

Yie

ld

Channel Number

Fig. 3.8: RBS spectrum of TiO2 thin film grown using precursor A

[Ti(OPri)2(tbaoac)2], the red line represents the values simulated for

TiO2 and black line represents the experimentally obtained spectra.

Page 89: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 78 -

In the RBS spectrum of the same film, the signals from Ti as well as O in the layer could

be clearly identified as depicted in Fig. 3.8. The comparison with the simulation gave no

hint to a sizeable additional contamination and supports the stoichiometric atomic ratio of

Ti to O as ½. The thickness of the film was found to be of the order 3.6 µm.

In addition to XPS and RBS, EDX analyses were carried out on selected films. Only

peaks belonging to titanium and oxygen could be detected in high intensity. No sizeable

carbon contamination could be detected. Fig. 3.9 shows the EDX spectrum obtained for

the TiO2 film deposited using the precursor A [Ti(OPri)2(tbaoac)2].

3.3.6 Microstructure of the films

The film morphology is determined by the relative rates of precursor vapor transport,

decomposition reaction, surface diffusion, and lattice incorporation.[22] The growth of a

thin film is a non equilibrium phenomenon where kinetics and thermodynamics play

essential role in determining the microstructure of the film. The morphology of the films

was analyzed by scanning electron microscopy, SEM. As a typical case the films grown

between temperatures 400 and 800 °C using precursor A [Ti(OPri)2(tbaoac)2] at a reactor

Fig. 3.9: The EDX spectrum of the TiO2 films deposited using the precursor A [Ti(OPri)2(tbaoac)2]

Page 90: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 79 -

pressure of 10 mbar is discussed here. At low temperatures, the microstructure is nearly

equiaxed and the texture shows very little preferred orientation. However, as the

temperature increases, the energy of the surface diffusion increases, so are the growth

rate, resulting in more rapid lattice incorporation on low energy crystal planes. Anatase

films were observed in this temperature regime (XRD) and granular structure was

observed with dense packing. At temperatures above 700 °C high degree of orientation

was observed with rutile crystalline forms. The film morphology shows densely packed

bigger crystallites grown at the cost of smaller crystallites as depicted in Fig. 3.10.

Fig. 3.10: SEM micrographas of films obtained using precursor A

[Ti(OPri)2 (tbaoac)2], at different susceptor temperatures and at a reactor

pressure of 10 mbar.

These initial experiments with new set of precursors were performed in order to test the

ability of precursors to deposit TiO2 thin films. The preliminary experiments conducted

using the homebuilt reactor had varying degree of success and paved the way for their

detailed study using commercial scale reactor. The precursor A [Ti(OPri)2(tbaoac)2],

showed promise for titanium dioxide films over a wide range of temperatures. As shown

in TGA, (Fig. 3.2) this precursor sublimes cleanly with negligible residue under normal

500 °C 600 °C

700 °C 800 °C

500 °C

Page 91: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 80 -

pressure and has high solubility (0.8 g/ml) in most of the common organic solvents hence

it was decided to scale up the synthesis of this precursor and use for liquid injection CVD

using commercial scale reactor.

3.4 Deposition of TiO2 thin films using a liquid injection industrial tool

3.4.1 Susceptor temperature dependent Growth rate, Surface roughness, Structural

and Electrical properties

The efficiency of a precursor is defined as the ratio of the amount (moles) of the

respective elements (titanium in present case) in the films deposited on all the five

wafers, (as determined by XRF analyses), to the amount (moles) of the element in the

consumed precursor. The XRF results indicate almost the same maximum efficiency in

the case of both precursors, but the temperature range for this was different in both cases

300 350 400 450 500 550 600 650 700 750 8000

10

20

30

40

50

2

4

6

8

10

12

Effi

cien

cy [%

]

Susceptor Temperature [oC]

[Ti(OPri)2(tbaoac)

2] on Si

[Ti(OPri)2(tbaoac)

2] on Pt

[Ti(OPri)2(thd)

2] on Si

[Ti(OPri)2(thd)

2] on Pt

Gro

wth

rat

e (n

m/m

in)

Fig. 3.11: Precursor efficiency of commercially available [Ti(OPri)2(thd)2]

precursor and newly developed [Ti(OPri)2(tbaoac)2] precursor on

Silicon and Pt/Zr/Si substrates under identical conditions. (as the

period is fixed the growth rate scales with the efficiency)

Page 92: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 81 -

( as shown in Fig. 3.11). [Ti(OPri)2(tbaoac)2] showed maximum value between 450 °C to

550 °C, where as [Ti(OPri)2(thd)2] showed this behaviour above 600 °C. This lower

growth temperature of TiO2 films with [Ti(OPri)2(tbaoac)2] compared to [Ti(OPri)2(thd)2]

precursor, making it useful for the technologically important low temperature

depositions. The [Ti(OPri)2(tbaoac)2] precursor has a lower efficiency at high deposition

temperatures probably due to the decomposition at elevated temperatures (>700 °C).

Efficiency of the precursors on Si and Pt/Zr/Si substrates prepared under identical

conditions were at variance (Pt showed higher efficiency) at lower deposition

temperatures than at higher temperature. This may be due to the temperature difference

on the Pt and Si substrates. We expect Pt substrates are at a higher temperature than the

Si substrates, and hence in the kinetically controlled low temperature region (exponential

growth with temperature) higher deposition was expected on the Pt coated silicon

substrates. In other words, this difference was due to the variation of reactant

concentration on the substrate surface due to difference in temperature on the Si and

Pt/Zr/Si substrates. Convergence of the efficiencies on both substrates at high

temperatures supports this argument, as at high reactor temperatures, both substrates may

be at the same temperature.

3.4.2 Surface roughness

Surface roughness of the films deposited at various reactor temperatures were studied

with AFM for both type of precursors. The average roughness showed strong dependence

on the susceptor temperature and the substrate. Films deposited on Si substrates had

lower surface roughness compared to those deposited on platinized substrates.

In the case of films prepared with [Ti(OPri)2(thd)2] precursor, this dependency was more

drastic compared to the films deposited with [Ti(OPri)2(tbaoac)2] precursor. For films

from [Ti(OPri)2(thd)2] precursor, RMS roughness variation from 0.35 nm to 7 nm for Si,

and 1.14 nm to 8 nm for Pt/Zr/Si substrates were observed. Fig. 3.12 shows the AFM

micrographs of the films deposited using two precursors at 600 °C on Si substrates. In

this case, for films deposited at low temperature (<550 °C) and high temperature (700-

Page 93: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 82 -

750 °C), the RMS roughness were found to be lower than those deposited in the medium

temperature range (550-650 °C).

Fig. 3.12: AFM micrographs of TiO2 films grown using [Ti(OPri)2(tbaoac)2]

and [Ti(OPri)2(thd)2] precursors on Si substrates deposited at 600 °C.

The probable reason for this behaviour may be attributed to the crystalline grain growth

as a function of temperature which can also be correlated with the XRD results as

crystalline phase starts at a temperature of 550 °C. As the susceptor temperature

increased grain growth increased and reached a maximum value in the medium range and

then again reduced due to the fine-grained structure at elevated temperatures. This

argument also supported with the XRD results. In the low temperature region lower

roughness is due to the amorphous nature of the film and high temperature this behaviour

may be due the effect of growth rate/efficiency on the grain growth. High growth rate

(10.22 nm/min on Pt at 600 °C) may result in the formation of localized clusters of grains

or the coalition of these clusters to form bigger grains in the case of medium

temperatures. We observed a slightly low growth rate (~ 9.67 nm/min on Pt at 750 °C)

for films at high temperature. This low growth rate along with the high mobility of

adatoms at high temperature might have influenced the fine-grained growth and hence the

lower roughness was observed in the crystalline films at high temperatures. On the other

hand, roughness of the films deposited with [Ti(OPri)2(tbaoac)2] precursor didn’t show

large dependence on the susceptor temperature. The roughness slightly increased with the

[Ti(OPri)2(tbaoac)2] Si- 600 °C

[Ti(OPri)2(thd)2] Si - 600 °C

Page 94: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 83 -

temperature and which can be explained due to the growth of crystalline phase and the

subsequent grain growth. The difference in the surface morphology of the TiO2 films

with two different precursors may be due to the difference in the thermal characteristics

and reaction steps leading to the formation dense films on the substrate surface.

3.4.3 Crystal structure

X-ray diffraction studies were performed to understand the onset of crystallization

temperatures on different substrates and with both precursors and the results are depicted

in Fig. 3.13 (a) and 3.13 (b).The reflections observed at 2θ = 28.3° and 29.95° for films

on Pt/Zr/Si are due to the Zr adhesion layer (ZrOx) as it is present in the amorphous case

also and absent in the films on Si substrates. Films deposited at 450 and 500 °C were

amorphous for both type of substrates and crystallization in the tetragonal ‘anatase’ phase

starts at a deposition temperature of 550 °C. Another low temperature ‘metastable’

orthorhombic phase, brookite was not observed in the crystallinity evolution. So in the

thin film forms, amorphous and anatase phases were stabilized in the present deposition

process. As the deposition temperature increased intensity of the most prominent peak

(2θ = 25.28°) corresponding to the (101) phase increased and other reflection (200),

(105) and (211) were also started to appear at 2θ = 48.05°, 53.89° and 55.06°,

respectively; suggesting polycrystalline nature of the films. The films on Si showed an

additional peak around 37.87° corresponding to the anatase (004) reflection. In the case

platinized substrate a weak reflection corresponding to the rutile form of the TiO2 (110)

was observed at 2θ = 27.3° and this was not present for the films on Si substrates. In

both form of substrates intensity of reflections increased with susceptor temperature.

Similar XRD patterns of TiO2 films on Pt substrates can be seen for the films deposited

with the [Ti(OPri)2(tbaoac)2] precursor, but at high deposition temperatures, contrary to

the [Ti(OPri)2(thd)2] case, anatase (101) peak intensity reduced. This may be due to the

low efficiency of the [Ti(OPri)2(tbaoac)2] precursor at high deposition temperatures (refer

to Fig.3.11), which results in low thickness as the films were deposited under identical

condition with the same number of pulses.

Page 95: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 84 -

Fig. 3.13: (a) X-ray diffraction patterns of TiO2 films with [Ti(OPri)2(thd)2] precursor on

Pt/Zr/Si and (b) with [Ti(OPri)2(tbaoac)2] precursor on Pt/Zr/Si deposited under various

susceptor temperatures.

20 30 40 50 60

(b)A

(211

)

A(2

00)R(1

10)

A(1

01)

Pt (

200)

Pt (111)ZrOx

350oC

400oC

450oC

500oC

550oC

650oC

750oC

TiO2/Pt/Zr/Si(100)

700oC

600oC

Inte

nsity

(ar

b.un

its)

2θ (Deg.)

20 30 40 50 60

0

50

100

150

200

250(a)

R(1

10)

A(2

11)

A(1

05)

A(2

00)

A(1

01)

Pt (200)Pt (111)ZrO

x 450oC

500oC

550oC

650oC

750oC

TiO2/Pt/Zr/Si(100)

700oC

600oC

Inte

nsity

(arb

.uni

ts)

2θ (Deg.)

Page 96: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 85 -

Also, in the case of films with [Ti(OPri)2(tbaoac)2] precursor crystallization started in the

anatase phase at 500 °C, which is 50 °C lower than [Ti(OPri)2(thd)2] precursor. Here, in

the case of films on Si substrates at higher temperature (>700 °C) films contains rutile

phase also, as the small peak at 2θ = 27.4° is corresponds to the rutile (110) reflection

and hence the films have two phases. On Pt substrates this trend is weakly seen at 750 °C

depositions. Compared to the films on Si from [Ti(OPri)2(thd)2] precursor,

[Ti(OPri)2(tbaoac)2] showed a different pattern. First, the appearance of rutile phase at

and above 700 °C of deposition. Anatase (004) reflection was absent in the case of

[Ti(OPri)2(tbaoac)2] precursor. Also (105) & (211) reflections were also not so well

resolved as in the case of [Ti(OPri)2(thd)2]; suggesting a lower grain size in the case of

films with [Ti(OPri)2(tbaoac)2] precursor. This lower grain size may be the reason for the

smaller surface roughness in the case of films with [Ti(OPri)2(tbaoac)2] precursor as

compared [Ti(OPri)2(thd)2] precursor.

Fig. 3.14: X-ray diffraction patterns of TiO2 films with [Ti(OPri)2(tbaoac)2] precursor on

Pt/Zr/Si substrates deposited at 550°C and then annealed (room temperature

annealing for 20 min in O2) at various temperatures.

20 30 40 50 60

A(004) A(112)

R(2

20)R

(211

)

R(2

10)

R(1

11)

R(2

00)

Pt

r

r

n

r

n

R(1

01)

A(211)A(200)

R(1

10)

A(101)

ZrOx

as.depo.

800oC

550oC

650oC

750oC

TiO2/Pt/Zr/Si(100)

700oC

600oCInte

nsity

(arb

.uni

ts)

2θ (Deg.)

Page 97: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 86 -

3.4.4 Effect of post deposition annealing on the structure and morphology

Rutile, technologically preferred form has higher refractive index and dielectric constant [33] than anatase could not be obtained in the single phase (in situ) due to the process

limitation of MOCVD system and low efficiency of the precursors at high temperatures.

This can be achieved by thermally induced transformation of anatase phase at

temperature higher than 800 °C as can be seen from the X-ray diffraction depicted in the

Fig. 3.14. X-ray diffraction shows as deposited film (550 °C) and the films annealed at

550 °C have anatase as crystalline phase. As the annealing temperature is increased to

600°C reflexes for rutile phase starts to grow at the cost of anatase phase. This trend was

observed till 750 °C and at 800 °C films were converted to rutile phase structure. Surface

morphology reveals grain growth with annealing temperature as expected due to the

growth of rutile phase. Larger grains were obtained for the films annealed at 800 °C

which corresponds to the tetragonal rutile single phase.

Fig. 3.15: SEM cross sectional view of the films grown at 550 °C and then annealed at

different temperatures.

SEM cross sectional view of the films annealed at different temperatures also showed

similar behaviour as can be seen from the images shown in Fig. 3.15. Closely packed

550 °C as deposited

annealed at 650°C annealed at 750°C

annealed at 550°C

Page 98: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 87 -

grains with columnar growth were observed in the cross-sectional view of the SEM

micrographs. Furnace annealing even at 850 °C resulted in the mixed phase where as

room temperature annealing at 800 °C resulted in only rutile phase as observed in the

XRD analyses.

3.4.5 Electrical properties

Electrical measurements in terms of C-V and I-V were done on the films in the MIS

configuration. C-V characteristics of the TiO2 films deposited at various susceptor

temperatures with [Ti(OPri)2(tbaoac)2] precursors were done first to understand the effect

of EOT (equivalent oxide thickness)and effective dielectric constant with the deposition

temperature. Fig. 3.16 shows the high frequency C-V measurements performed on 49.1

nm2 circular capacitor patterns. The capacitance was measured at 100 kHz as a function

of gate voltage and capacitor was swept from inversion to accumulation and back to

inversion to check the amount of hysteresis.As is evident from Fig.3.16, there was no

significant hysteresis as the crystallization of initial amorphous TiO2 films took place.

This suggests that introduction of grain boundaries or crystal/amorphous boundaries do

not significantly increase the density of carrier trapping defects sites that contribute to C-

V hysteresis at these frequencies. The films deposited at 500 and 600 °C showed almost

same characteristics, except an increase in the accumulation capacitance. This is due to

increase in the crystallinity in the films with deposition temperature temperatures as

observed with the XRD. In contrast, the films deposited at 650 °C showed a reduction in

the accumulation capacitance preferably due to the amorphous interface oxide growth at

higher temperatures. Also the shape of this curve and the one deposited at 450 °C was

found to be different compared to others, stretched along the X-axis, suggesting a large

interface traps, in it.

Page 99: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 88 -

-2 -1 0 1 2

0.00E+000

1.00E-010

2.00E-010

3.00E-010

4.00E-010

5.00E-010

6.00E-010

7.00E-010

Cap

acita

nce

(F)

Bias Voltage (V)

450oC 500oC 600oC 650oC

-2 -1 0 1 2

0.00E+000

2.50E-010

5.00E-010

Cap

acita

nce

[F]

Bias Voltage [V]

450oC 500oC 600oC 650oC

Fig.3.16: C-V characteristics of TiO2/ Si as a function of deposition temperatures (a) as-

deposited (b) post deposition top Pt annealing at 550 °C for 15 min, only for the

films which did not show the saturation.

(a)

(b)

Page 100: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 89 -

450 500 550 600 6500

2

4

6

8

10

Target requirement

Oxide thickness tox

[nm]

24.7 27.6 18.7 21.7

EOT ε

r

Susceptor temperature [°C]

EO

T [n

m]

0

10

20

30

40

50

Die

lect

ric c

onst

ant [

ε r]

Fig.3.17: Equivalent oxide thickness (EOT) and dielectric constant of films deposited at

various temperatures. Dotted line represents the requirement for EOT.

For a gate dielectric of thickness Td and relative dielectric constant k, EOT (equivalent

oxide thickness) is defined by EOT = Td/k/3.9 where 3.9 is the relative dielectric constant

of thermal SiO2. The ideal gate capacitance per unit area of the gate dielectric of

thickness Td is the same as that of a gate dielectric made up of thermal SiO2 with a

thickness of EOT. The equivalent oxide thickness and dielectric constant of the TiO2 thin

films deposited at various temperatures were measured. Particularly, the electrical studies

of ultra-thin films on Si substrates showed promising properties; EOT of ~ 2 nm and

dielectric constant ~ 35 were obtained. The variation of EOT and dielectric constant as a

function of deposition temperature is depicted in the Fig. 3.17.

Room temperature leakage current behaviour was measured at various stages of micro

structural evolutions and the representative results are shown in Fig. 3.18 for TiO2 films

deposited onto Si(100) substrates at different susceptor temperatures.

Page 101: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 90 -

-2 -1 0 1 2

1E-9

1E-8

1E-7

1E-6

1E-5

1E-4

1E-3

0.01

0.1Le

akag

e cu

rren

t (A

/cm

2 )

Applied Voltage (V)

Deposition temperatures 450oC 500oC 600oC 650oC

Fig.3.18: Leakage current characteristics of TiO2 deposited on Si structures

3.6 Deposition of SrTiO3 thin films

The compatibility of the Ti precursor with the standard thd-precursor for the group-II

metals was investigated by depositing SrTiO3 thin films. In addition, it was of special

interest to investigate whether the advantageous low temperature deposition behaviour of

the new Ti precursor could be preserved during SrTiO3 depositions. Deposition

conditions are listed in table 3.2.

Fig. 3.19 summarizes the efficiency of the metal incorporation, (the ratio of total amount

of Ti or Sr in the film to the amount of Ti or Sr in the injected precursor), into the films at

different temperatures. The precursors were injected alternatively with a period of 0.45 s

and an offset between Sr and Ti of 0.225 s in order to reduce gas phase reactions.

Page 102: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 91 -

Fig. 3.19: Variation of efficiency of the precursor and stoichiometry of the film with

susceptor temperature on Pt substrates (Period = 0.45 s and delay between Sr

and Ti injections was 0.225 s)

It can be seen from the figure that efficiency of both precursors was improved as the

susceptor temperature increased and from the trend high temperature deposition yields

stoichiometric films of SrTiO3 with the same injection rate for Sr and Ti. But at low

temperature especially the efficiency of the Ti precursor was drastically decreased

compared to the deposition of pure TiO2.

The behavior of the Sr precursor in combination with [Ti(OPri)2(thd)2] is given as a

reference line and there is only a small decrease. Hence, the drastic reduction in the Ti

incorporation in the films resulted in Sr rich films. There seems to be some interaction

and some hindering of the Ti incorporation at low temperatures in the presence of the Sr

precursor. In this temperature region reactions occur near the substrate or on the substrate

450 500 550 600 650 700 7500

5

10

15

20

25

30

35

40

45

Ti-Eff Sr-Eff Sr-Eff with [Ti(OPri)

2(thd)

2]

Sr/Ti ratio

Susceptor Temperature (°C)

Eff

icie

ncy

(%)

1

3

5

7

9

11

13

15

SrTiO3/Pt/ZrO

x/SiO

x/Si

Sto

ichi

omet

ry (S

r/Ti

)

Page 103: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 92 -

in such a way that kinetically controlled adsorption of the precursor molecule and the

subsequent desorption of the organic ligand.

0 2 4 6 8 10 120

10

20

30

40

50 Ti-Eff Sr-Eff Sr/Ti ratio

Delay between Sr and Ti pulses (s)

Eff

icie

ncy

(%)

1.0

3.0

5.0

7.0

9.0(a)

Sto

ichi

omet

ry (S

r/Ti

)

0 2 4 6 8 10 120

10

20

30

40

50

Ti-Eff Sr-Eff Sr/Ti ratio

Delay between Sr and Ti pulses (s)

Eff

icie

ncy

(%)

1

3

5

7

9

11

13

15(b)

Sto

ichi

omet

ry (S

r/Ti

)

Fig. 3.20: Efficiency and Sr/Ti ratio as function delay time for the films deposited at

(a) 450 °C, (b) 500 °C on Pt substrates. (Ti:Sr pulse ratio = 1:1).

Page 104: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 93 -

0.0 0.2 0.4 0.6 0.8 1.00

5

10

15

20

25

30

35

40

45

50(a)

B C

Delay between Sr and Ti pulses (s)

Eff

icie

ncy

(%)

0.00

0.25

0.50

0.75

1.00

1.25

1.50

1.75

2.00

Sr/Ti

Sto

ichi

omet

ry (S

r/Ti

)

1.0 1.1 1.2 1.3 1.4 1.50.01

0.1

1

10

(b)

EA~ 0.9 eV

1/de

lay

time

(s-1)

1000/T(K-1)

Fig. 3.21: Efficiency and Sr/Ti ratio as function delay time for the films deposited at

(a) 650 °C on Pt substrates. (Ti:Sr pulse ratio = 1:1) and (b) Arrhenius plot

of desorption rate.

Page 105: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 94 -

Hence the presence of each precursor affect the deposition of the other, with one more

sensitive compared to the other. Since the layer growth at 400 °C to 500 °C is of high

interest for technological applications like the manufacture of high-k embedded

capacitors and as it is shown possible from the efficiency of the individual precursors,

depositions were tried to obtain stoichiometric films at these temperatures. The delay

between the Sr and Ti injection were increased, while maintaining the same injection

rate, towards the direction of individual deposition.

The results for depositions at 450 °C, 500 °C and 650 °C are depicted in the Figs. 3.20

(a), 3.20 (b) and 3.21 (a). As can be seen from the figure, 650 °C deposition resulted in

stable efficiencies and stoichiometric STO deposition with a very small delay (0.2 s)

between the Sr and Ti injection into the vaporizer. At 500 °C a plateau in the Sr/Ti curve

was observed after approximately 5 s delay, however, an additional change of the 1:1

ratio of the Sr and Ti pulses would be necessary for obtaining exactly stoichiometric

films. At 450 °C this saturation could be observed only after a delay above 10 s. As gas

phase reactions can be excluded due to the pulse separation, this behavior suggests that

adsorption and desorption reactions on the substrate surface affect the incorporation of

the cations in the growing films. The corresponding cleaning rate of the surface, 1/delay

time, is plotted in Fig. 3.21 (b) in an Arrhenius diagram.

A straight line with activation energy of about 0.9 eV could be observed. This effective

activation energy is less compared with the activation energy for the TiO2 deposition of

~1eV discussed above, and therefore this desorption process becomes the rate limiting

reaction, resulting in increase in Sr content at low temperatures. Although there was some

success in the deposition of stoichiometric films by additional correction of the Sr/Ti

injection rates, the low temperature deposition (T < 450 °C) demands long deposition

times, which are undesirable for industrial throughput. But a more compatible Sr

precursor with newly developed Ti-precursor may result in stoichiometric deposition of

SrTiO3 at low temperatures.

X-ray diffraction of the film deposited at three different temperatures, 450 °C, 500 °C

and 650 °C is shown in Fig. 3.22. Grazing angle XRD patterns (Fig.3.22 (a)) reveal an

amorphous nature for the film deposited at 450 °C. At 500 °C we observe a

polycrystalline phase although this film is strongly off-stoichiometric, Sr/Ti ratio of about

Page 106: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 95 -

1.7. Peaks at 2θ = 22.78°, 32.42°, 46.48° and 57.79° are corresponding to the (100),

(110), (200) and (211) reflections of the cubic SrTiO3 phase. Additional peaks, apart from

the substrate peak, in the XRD pattern of this film could be attributed to SrOx phases.

Stoichiometric films deposited at 650 °C didn’t show any reflection corresponding to a

crystalline phase, due to the texturing and hence Bragg-Brentano XRD was taken for this

film and is shown in Fig. 3.22 (b). θ-2θ scan shows crystalline peaks with some texturing

along the (200)- and (110)-direction. In this case the (200)-reflection can be distinguished

from Pt and attributed to SrTiO3 considering the ratio of the 100 and 200 reflections.

Fig. 3.22: X-ray diffraction of SrTiO3 on Pt/ZrOx/SiOx/Si substrates: (a) grazing angle

XRD at three different temperatures (b) θ-2θ scan on the film deposited at

650 °C.

20 30 40 50 60

(a)SrO

xt

tt (211

)

(110

)

(100

)

(200

)

Pt (111)ZrOx 450oC

500oC

650oC

SrTiO3/Pt/Zr/Si(100)

Inte

nsity

(arb

.uni

ts)

2θ (deg.)

20 30 40 50 600

200

(b) θ-2θ scan

Pt (111)ZrOx

(200

)

(110

)

(100

)

Inte

nsity

(arb

.uni

ts)

2θ (deg.)

Page 107: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 96 -

3.6 Summary

The newly developed titanium complexes namely, titanium bis(isopropoxide) bis(tert-

Butylacetoacetate) [Ti(OPri)2(tbaoac)2] (A) titanium bis(ethoxide) bis(tert-

Butylacetoacetate) [Ti(OEt)2(tbaoac)2] (B), Titanium bis(isopropoxide)

bis(methylacetoacetate) [Ti(OPri)2(meaoac)2] (C), were tested for TiO2 depositions using

home built horizontal cold wall reactor. No additional reactive gases like oxygen or water

vapor were used for depositions. Thin film depositions were carried out in the

temperature range 350 °C to 800 °C and reactor pressures of 10 mbar and 50 mbar were

used. Nitrogen was used as carrier gas (100 sccm) and depositions were carried out on

Si(100) substrates. The deposited TiO2 films were analyzed by XRD, SEM, EDX, XPS

and RBS for crystal structure, microstructure, composition and thickness. Weight gain of

the substrate was used to determine the thickness of the films assuming the density of

bulk anatase phase.

Films grown at 350 °C did not show any XRD reflexes because of amorphous nature of

the films. The temperature onset of crystallization was 400 °C with all the three

precursors. At 400 °C of deposition temperature, anatase phase was observed in case of

precursors [Ti(OPri)2(tbaoac)2] (A) and [Ti(OEt)2(tbaoac)2] (B), while depositions with

precursor [Ti(OPri)2(meaoac)2] (C) showed only brookite phase at this temperature. With

increasing temperature the transition from anatase phase to rutile phase was observed.

The rutile phase was predominant from 700 °C up to 800 °C but some traces of anatase

phase with lower (XRD) intensity was observed.

Growth rates as determined by weight gain of the substrates was highest (~ 0.6 µm/hour)

in case of precursor [Ti(OPri)2(tbaoac)2] (A) at a temperature of 500 °C. Precursors

[Ti(OEt)2(tbaoac)2] (B) and [Ti(OPri)2(meaoac)2] (C) showed highest growth rates of 0.5

µm/hour and 0.3 µm/hour at 600 °C respectively. Since the films contain rutile phase and

some carbon impurities, the results may deviate from the results obtained by other

methods. Film composition of the films grown at 700 °C using [Ti(OPri)2(tbaoac)2] (A)

as determined by XPS showed the presence of titanium, oxygen and some traces of

carbon. In RBS spectrum of the same film, showed the signals from Ti as well as O.

Simulated curves for TiO2 were compared with experimentally observed curves and it

Page 108: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 97 -

was found that there was no hint for sizeable carbon contamination and the stoichiometric

atomic ratio of Ti to O was found to be ½ . Microstructure of the films as observed in

SEM shows microcrystalline morphology at low temperatures. At high temperatures

granular structure was observed with dense packing.

With the initial testing of the precursors it was concluded that the precursors can be used

for the deposition of TiO2 thin films. The precursor [Ti(OPri)2(tbaoac)2] (A) was highly

soluble in most of the organic solvents and showed promise for deposition over wide

range of temperatures. Hence it was decided to test this precursor using liquid injection

industrial scale reactor. The TiO2 depositions using industrial scale reactor were carried

out at MOCVD facilities at Forschungszentrum Jülich.

The comparison between the performances of the commercially available titanium

precursor [Ti(OPri)2(thd)2] and newly developed precursor [Ti(OPri)2(tbaoac)2] (A) were

made. Precursors were dissolved in butyl acetate and solutions of 0.05 M concentration

were prepared. Depositions were carried out from 350 °C to 800 °C using oxygen as

reactive gas. Films were deposited on Pt/ZrOx/SiOx/Si and on SiOx/p-Si substrates.

Deposition pressure of 1-1.5 mbar was employed. In addition to TiO2 thin films, complex

oxide SrTiO3 depositions were investigated using standard Sr precursor [Sr(thd)2].

Composition of the films determined by XRF results indicated that [Ti(OPri)2(tbaoac)2]

(A) showed maximum efficiency of titanium incorporation in to growing films between

450 °C to 550 °C, where as the [Ti(OPri)2(thd)2] precursor showed the this behavior

above 600 °C. Surface roughness as determined by AFM showed that at in the

temperature range of 550 °C-650 °C the films had higher roughness compared to films

deposited at higher or lower temperatures. The average roughness showed strong

dependency on the susceptor temperature and the substrate. Onset of crystallization

temperature using precursor [Ti(OPri)2(thd)2] was found to be 550 °C. But the newly

developed precursor [Ti(OPri)2(tbaoac)2] showed the onset to be at 500 °C .

Crystallization of anatase films starts at 500 °C with prominent reflex (2 ? = 25.28°). At

higher temperature (>700 °C) reflexes due to rutile (110) (2 ? = 27.4°) were also

detected.

Electrical measurements in terms of C-V and I-V were done on the films in the MIS

configuration. No significant hysteresis was observed as the crystallization of initial

Page 109: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 98 -

amorphous phase took place. Films deposited at 650 °C showed a reduction in the

accumulation capacitance. Electrical studies of ultra-thin films on Si substrates showed

promising properties; EOT of ~ 2 nm and dielectric constant of ~ 35 were obtained.

The compatibility of newly developed precursor [Ti(OPri)2(tbaoac)2] (A) and

[Ti(OPri)2(thd)2] with [Sr(thd)2] were investigated for the deposition of SrTiO3 thin films.

The [Sr(thd)2] and titanium precursors were injected alternatively with an offset period

of 0.225 s. Efficiency of both precursors to incorporate titanium in to growing films

increased with increasing temperature. At high deposition temperature (>600 °C)

stoichiometry between Sr:Ti in the 1:1 ratio could be achieved. At lower temperatures

stoichiometric ratio could not be achieved and films had titanium deficiency. In order to

reduce the interaction between the two precursors, the delay between injections were

increased up to 5 s. It was observed that stoichiometric films are possible at 500 °C with

approximately a 5 s delay between injections.

Page 110: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 99 -

3.7 References

[1] R. N. Ghoshtagore, A. J. Norieka, J. Electrochem. Soc. 1970, 117, 1310.

[2] A. C. Jones, Materials Science in Semiconductor Processing 1999, 2, 165.

[3] T. Kodas, M. J. Hampden-Smith, The chemistry of metal CVD, VCH Publishers,

Weiheim, Germany, 1994.

[4] M. L. Hitchman, K. F. Jensen, Chemical vapor deposition-Principles and

applications, Academic Press, 1992.

[5] A. R. Teren, J. A. Belot, N. L. Edleman, T. J. Marks, B. W. Wessels, Chem. vap.

deposition. 2000, 6, 175.

[6] C. S. Hwang, Mater. Sci. Eng. 1998, B, 56, 178.

[7] C. S. Hwang, S. O. Park, H. J. Cho, C. S. Kang, H. K. Kang, S. I. Lee, M. Y. Lee,

Appl. Phys. Lett. 1995, 67, 2819.

[8] M. Shimizu, M. Fujimoto, T. Katayama, T. Shiosaki, K. Nakaya, M. Fukagawa,

E. Tanikawa, Mater. Res. Soc. Symp. Proc. 1993, 310, 255.

[9] K. Tominaga, A. Shirayanagi, T. Takagi, M. Okada, Jpn. J. Appl. Phys. 1993, 32,

4082.

[10] T. Carlson, G. L. Giffin, J. Phys. Chem. 1986, 90, 5896.

[11] B. E. Yoldas, T. W. O´Keeffe, Appl. Opt. 1979, 18, 3133.

[12] M. A. Butler, D. S. Ginley, J. Mater. Sci. 1980, 15, 19.

[13] T. Matsunaga, R. Tomoda, T. Nakajima, T. Komine, Appl. Environ. Microbiol.

1988, 54, 330.

[14] S. I. Borenstain, U. Arad, I. Lyubina, A. Segal, Y. Warschawer, Thin Solid Films

1999, 75, 2659.

[15] P. S. Peercy, Nature 2000, 406, 1023.

[16] D. Wang, Y. Masuda, W. S. Seo, K. Koumoto, Key Eng. Mater. 2002, 214, 163.

[17] K. Abe, S. Komatsu, Jpn. J. Appl. Phys. 1992, 31 Part 1., 2985.

[18] C. S. Kang, H.-J. Cho, C. S. Hwang, B. T. Lee, K.-H. Lee, H. Horii, W. D. Kim,

S. I. Lee, M. Y. Lee, Jpn. J. Appl. Phys. 1997, 36, 6946.

Page 111: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 100 -

[19] T. Horikawa, M. Tarutani, T. Kawahara, M. Yamamuka, N. Hirano, T. Sato, S.

Matsuno, T. Shibano, F. Uchikawa, K. Ono, T. Oomori, MRS Symp. Proc. 1999,

3, 541.

[20] A. Sherman, Chemical Vapor Deposition For Microelectronics, Noyes

Publications, Park Ridge, 1987.

[21] D. Briggs, M. P. Seah, Practical Surface Analysis, Vol. 1, 2 ed., John Wiley &

Sons, Chichester, 1996.

[22] J. A. Venables, G. L. Price, Epitaxial growth, Academic Press, NY, 1975.

Page 112: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 101 -

Chapter 4

Investigations into thermal decomposition of the precursors

using matrix isolation – FTIR techniques

Abstract

The thermal decomposition of precursors used for CVD plays an important role in the

process. The thermal decomposition of the precursors is studied using several techniques.

Mostly decomposition studies are tried either in situ using mass spectroscopy or by the

use of combinational studies like GC-MS, temperature desorption studies etc. Techniques

employed so far to study mechanism of decomposition of Titanium tetraisopropoxide

[Ti(OPri)4], (TTIP) were either in the bulk phase or on a surface. In a MOCVD process,

gas phase of a precursor plays an important role which determines final film quality and

stoichiometry. Present work describes the investigation of the gas phase decomposition of

CVD precursors using matrix isolation (MI) technique coupled with FTIR spectroscopy.

TTIP the most volatile alkoxide of titanium has been studied for thermal decomposition

using MI–IR technique. As a supplementary experiment, the matrix isolation of iso-

propanol was carried out. During the course of this work new class of mixed alkoxide

titanium complexes having ß-ketoester were synthesized. It was speculated that, an ester

moiety on the side chain of the ß-diketonate system will act as a “cleavage point”

facilitating easy decomposition of the precursor complex. In order to ascertain these facts

the mixed ß-diketonate/ß-ketoester complexes like [Ti(OPri)2(thd)2], [Ti(OPri)2(tbaoac)2],

were studied for gas phase decomposition behavior, so that the difference between

decomposition behavior of the two complexes could be understood. In addition, the

thermal decomposition studies of ligands Htbaoaoc (tert.-Butyl acetoacetate) and Hthd

(2,2,6,6- tetramethyl-3,5- heptanedione) were also studied to gain insights into their

decomposition mechanism.

The mixed alkoxide-ß-diketonate/ketoester complexes decompose through the formation

of acylketenes as intermediates. Matrix isolation studies coupled with FT-IR

Page 113: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 102 -

spectroscopy techniques were very useful to analyze these short-lived intermediates.

Efforts were made to suggest corresponding gas phase decomposition mechanisms based

on these studies. With the help of matrix isolation investigations it was able to reveal new

decomposition pathways not previously considered by other researchers. In addition this

chapter presents shortly the details of the matrix isolation technique and the issue

concerning the use of thermolysis oven used during the study.

Page 114: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 103 -

4.1 Introduction

Precursor designing is a complex issue for which reliable analytical feedback from

different stages of development is necessary.[1] One such important stage where analysis

is essential is thermolysis of the precursor. It is interesting to know how a metal complex

decomposes by the application of thermal energy. Preferably there should be simple steps

leading to formation of required stoichiometry. The ideal case should be the

decomposition of the precursor at the required temperature without any prior

decomposition and leading to the formation of high quality films. There are deviations

from the ideal case where intra and inter molecular reactions and rearrangements taking

place during thermolysis of the precursors. A detailed analytical feedback is very helpful

in improving design and inclusion/exclusion of ligand moieties suiting the requirement.

Therefore mechanistic studies on precursor decomposition are crucial to understand a

CVD process. Typically study of precursor decomposition mechanism has to be done on

gas phase of the molecules. Gas-phase studies are generally difficult due to

multidimensional problems encountered during these studies.[2] For e.g. the molecules in

the `gas phase´ require sensitive method with low response times. Reason being,

interactions between different species with gas phase leads to constant change in the

composition and species under investigation during the period of measurement.[3]

Most of the mechanistic studies of the decomposition of organometallic precursors have

been carried out in solutions in organic solvents.[4] In addition, the efforts are often

directed towards understanding the precursor behavior using surface

adsorption/desorption during a CVD process.[5-6] In such cases the role of the gas phase

of a precursor is often neglected. There are several empirical methods, which can be used

to study the decomposition mechanisms. But cautious approach has to be used, as CVD

process itself is often associated with extremely complex physical and chemical growth

kinetics. Also the amount/concentration of gas-phase species required for detailed

analysis has to be high or method of detection should be highly sensitive. The most

commonly used methods are in situ molecular beam mass spectrometry and in situ

Fourier transform infrared spectroscopy.[7-8] Though these analytical methods are

Page 115: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 104 -

successful to certain extent, they still lack qualitative measurement of transient species

involved in gas phase decomposition of a MOCVD precursor.

The matrix isolation (MI) technique is particularly well suited for the study of such

transient species, which react rapidly or decompose under normal conditions. In matrix

isolation technique, the molecules or reactive intermediates are shock-like frozen in an

inert gas matrix at low temperatures (5-20 K).[9] Since these frozen matrices are

accessible in principle, for spectrometric analysis for indefinite length of period, different

analytical tools can be used according to the convenience and the need to analyze the

matrices. Specifically, for the gas phase generated by a MOCVD precursor, which is one

of the highly dynamic systems, matrix isolation techniques are very helpful for retaining

the transient species. Matrix isolation technique in combination with vibrational

spectroscopy in the infrared region (IR) is reasonably well suited combination, wherein

the qualitative analysis of the reactive intermediates can be performed.

Basically the MI apparatus consists of a cooling system (having cryostat with closed

Helium cycle) to cool the matrix deposition window (CsI and associated accessories)

maintained at 10 K and a high vacuum system capable of attaining a vacuum of the order

10 -7 mbar. Inert gas systems form a matrix, act as a trap for the molecules and ions

resulting from the thermolysis of the precursors. Analytical part mainly consists of FTIR,

ESR, and laser induced fluorescence etc. In MI experiments, the precursor is vaporized

by using different methods. The vapors are passed by the action of carrier gas (inert)

through an oven of aluminum oxide where they get thermolyzed over a wide range of

temperatures. Resulting fragments are trapped in a highly dilute matrix of inert gases at

low temperatures (15-20 K). The whole system is maintained at reduced pressure within a

range of 10-5 to 10-7 mbar. So the matrix isolation set up strongly resembles a hot wall

CVD reactor. The themolysis oven of the matrix isolation set up basically acting as a hot

zone of the reactor and the exhaust gases are trapped inside the matrix. By using

absorption spectroscopy, a broad spectral range from the far infrared to the UV can be

explored in a single series of experiments. The results are compared with well established

results from the small molecules and fragments and trapped ions and identified in most

cases.

Page 116: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 105 -

By changing the experimental conditions like temperature of deposition window

(annealing), photolysis in addition to thermolysis, helps to determine isomers of the same

species. Intensity ratios of different species represent the concentration of species in the

matrix. Valuable information about possible reaction path can be deduced from detailed

analysis of the intensity ratios. In addition to the MI experiments, quantum mechanical

calculations can help in determining the possible equilibrium structures corresponding to

a particular atomic composition and to simulate their IR spectra. Relevant advantages of

MI over other analytical methods: highly reactive intermediates can be detected as they

are frozen in a dilute matrix of inert gases; Identification of fragments with short life

time; Use of optimal sensitive methods for detection; Use of low pressure conditions

more close to a real time CVD process; Use of broad range of thermolysis temperatures

to get an insight in to the process, possibility to get an idea of photolysis in addition to

thermolysis.

4.1.1 Thermolysis oven of the matrix isolation apparatus and its limitations

The oven used for thermolyses experiments in matrix isolation apparatus consists of an

Al2O3 tube having twin channels of 1 mm diameter. One of the channels is used for the

flow of the gaseous mixture and the other is used for inserting thermocouple to measure

the temperature of the oven without being in contact with the gaseous mixture. The rear

end of the Al2O3 tube is attached to the vacuum line and front end is directed towards the

matrix cooled window. The last 15 mm of the Al2O3 tube is heated using resistive

heating. A heat shield is provided to minimize the convection of heat of the oven to the

matrix window. The whole set up is enclosed in a stainless steel case to be attached to

matrix vacuum lines. This assembly can be effectively used to thermolyze the precursors

and carry out matrix isolation of the resulting thermolysis products. But at the same time,

the Al2O3 tube used in the oven acts as a heated substrate and depositions of materials are

observed inside the oven surface. In order to ascertain the fact that there is indeed

coupling of gasphase and surface reactions taking place certain parameters like residence

time of the gaseous molecules inside the oven, mean free path of molecules and Knudsen

number need to be specified. Following sections describe these issues in detail.

Page 117: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 106 -

4.1.2 Residence time of the gas phase molecules in the channel of the matrix oven

The oven used in the matrix apparatus is a narrow Al2O3 tube with an orifice having a

diameter of 1 mm. The inner surface of the Al2O3 tube coming in contact with gas phase

of the precursors is potentially active surface upon which reactions can take place. In

order to minimize the surface reactions inside the oven, it is important for the gas phase

molecules to have minimum possible residence time. The mean residence time (t) of the

molecules inside the matrix oven can be defined by,

(equation 4.1)

Where Vr is the volume of the channel of the matrix oven and v is the volumetric flow

rate of gases through it.

In all the experiments the argon flow maintained through the container into the oven at

the rate of 1.25 sccm and operating pressure is of the order of 10-6 mbar. A flow of 1.25

sccm at atmospheric pressure corresponds to 1.25 x 109 cm3/minute of volumetric gas

flow into the oven, which is maintained at 10-6 mbar (calculated at constant temperature).

The diameter of the oven channel is 1 mm (dimension provided by the supplier) and total

length of the oven is 35 mm of which the last 15 mm are heated to required temperature.

The total volume of the channel of the oven therefore is 0.0275 cm3. The total length of

the oven is considered to be the critical dimension for all calculative purposes. The

residence time of the molecules in the channel of the oven therefore can be calculated

from the above mentioned equation 4.1 as,

0.0275/1.25 x 109 = 3.66 x 10-13 sec.

The residence time also depends on various parameters like, temperature, pressure,

sticking coefficient of the precursor molecules, surface diffusion etc, which are not

considered in the calculations above.

ντ rV

=

Page 118: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 107 -

4.1.3 Coupling of gas-phase and surface/wall reactions

Because of the low pressures used in matrix isolation systems, the gas flow can be in

either the usual continuum regime or in the transition regime, depending on the relative

magnitude of the mean free path of the reactant molecules and the characteristic

dimension of the system, as reflected by the Knudsen number. The Knudsen number is

proportional to {length of mean free path (?) / characteristic dimension (L)} and is used

in momentum and mass transfer in general and very low pressure gas flow calculations in

particular. It is normally defined in the following form:

L

Knλ

= (equation 4.2)

For Kn = 0.01 the flow is dominated by gas molecule collisions and the continuum

models apply to such systems. When Kn > 10, the molecules primarily collide with solid

surfaces and the flow is described as `free-molecular`. Gas flows in this regime can not

be modeled by the classical continuum equations but must be described through view

factor, computations similar to those used for the radiation heat transfer or by Monte-

Carlo simulations. The intermediate range of Knudsen numbers, 0.01 < Kn < 10,

corresponds to the so-called transition flow regime where both gas-phase and surface

interactions are important. In this case, solution of the Boltzmann equation or use of

specialized Monte-Carlo simulation techniques must be applied in order to model low

pressure systems accurately.

The diameter of the argon is 4.17 Å and the mean free path ? of the argon atoms inside

the oven are given by the equation,

PNdRT

A22π

λ = (equation 4.3)

where d is critical dimension of the oven, P is the operating pressure of the matrix

isolation unit and T is the temperature. At 27 °C and at 10-6 mbar of operating pressure

the mean free path of the argon is found to be 53.63 m (using equation 4.3). For argon

atoms experiencing a mean free path of 53.63 m at 10-6 mbar, the Knudsen number, Kn is

found to be ~ 1532.28 (using equation 4.2). This high Kn is in accordance with studies

Page 119: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 108 -

which deal with gas transport characteristics through microchannels operating at low

pressures. It was reported that these microchannel gas flows are characterized by high

Knudsen numbers. High Knudsen number gas flows are characterized by ‘rarified’ or

‘microscale’ behavior; wherein significant gas-phase as well as surface/wall reactions are

possible. Because of significant non-continuum effect, traditional CFD (computational

fluid dynamics) techniques are often inaccurate for analyzing high Kn gas flows. The

direct simulation Monte Carlo (DSMC) method offers an alternative to traditional CFD

which retains its validity in slip and transition flow regimes.[10]

This clearly indicates that during the process of matrix isolation, the gas phase of the

precursors not only undergoes collisions in the gas phase but also with the inner surface

of the oven. Reactions that take place at the surface of the oven are known as

heterogeneous reactions. Reactions that take place in the gas phase are known as

homogeneous reactions. In short, though the homogeneous reactions are much more

desirable than heterogeneous reactions occurring at the surface, during matrix isolation

the coupling between these two types of reactions is unavoidable using present set up and

a complete gas phase analyses are thus hindered. The lower residence time of the

molecules on the surface of the oven may de-couple these heterogeneous reactions to

some extent but the surface reactions do occur in considerable number. Nevertheless, the

technique of matrix isolation can be used as a complementary to other gas phase studies,

surface decomposition techniques and a useful picture of the reaction mechanism can be

derived.

In the course of this study the mechanistic studies on selected precursors of titanium were

carried out using MI-IR techniques. The most commonly used precursors for deposition

of TiO2 thin films using MOCVD are titanium alkoxides. Titanium alkoxide complexes

have been prepared from a variety of alcohols.[20] The simple titanium tetra alkoxides are

all volatile, but among them titanium tetraisopropoxide [Ti(OPri)4] is the most volatile of

the titanium alkoxides, a property that it owes to its low molecular weight and the steric

effects of the ligand which restricts the nuclearity of the complex.[20] Titanium

tetraisopropoxide [Ti(OPri)4] is a liquid at or above room temperature and its nuclearity

has long been accepted as being approximately 1.4.[20] Titanium tetraisopropoxide is the

most widely used precursor for TiO2 thin film depositions using MOCVD. Therefore the

Page 120: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 109 -

gas phase decomposition of the [Ti(OPri)4] is of interest to understand the CVD processes

involving this precursor. Though there are several studies[4,7,11] on the decomposition of

the [Ti(OPri)4] complex on surface and in the bulk phase, the gas phase studies have not

been reported so far. So it was appropriate to study the gas phase decomposition of

[Ti(OPri)4] using matrix isolation technique. MI technique can be a powerful method to

analyze relevant molecular decomposition mechanism and this methodology has not yet

been widely recognized within CVD community.

Titanium tetraisopropoxide [Ti(OPri)4], has Ti(IV) center which is coordinatively

unsaturated. Therefore, [Ti(OPri)4] is susceptible for attack of air or moisture resulting in

facile hydrolysis. This sensitivity towards air and moisture is reduced by reacting

[Ti(OPri)4] with chelating ligands such as ß-diketonates wherein two of the alkoxy

ligands are replaced by bidentate chelating ligands resulting in full coordinatively

saturated titanium complexes. The most widely used chelating ligand is thd (2,2,6,6,

tetramethyl-3,5-heptane-dione) and resulting complex, [Ti(OPri)2(thd)2] is reported to be

six coordinated and stable complex based on NMR studies.[12] It is one of the most widely

used standard precursors for the deposition of titanium containing oxide thin films using

MOCVD.[13] During our efforts to synthesize new precursors for MOCVD of titanium

dioxide thin films, ß-ketoester, tert. Butylacetoacetate (tbaoac) was used as chelating

ligand in combination with titanium tetraisopropoxide [Ti(OPri)4], resulting in

monomeric six coordinated stable complex [Ti(OPri)2(tbaoac)2]. Thin films of TiO2 were

deposited using this precursor and it was found that the new precursor was able to

efficiently incorporate titanium into growing films at temperatures as low as 350 °C.

(refer chapter 2 and 3 for details)

So it was thought that gas phase decomposition of these precursors would be helpful in

understanding the difference in behavior. Though there are reports about decomposition

mechanism involving the decomposition of the standard complex [Ti(OPri)2(thd)2], there

is a lack of clear understanding of the decomposition mechanism involving the reactive

intermediates in the process.[15] The main aim of the MI-IR in present case was to provide

qualitative information about the gas phase during the thermal decomposition of the

titanium MOCVD precursors under isolated conditions.

Page 121: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 110 -

During this study the complexes were investigated for decomposition in the gas phase

using MI-IR techniques. As a preliminary set of the experiments, the individual ligands

used in the study were studied for their thermal decomposition behavior using MI-IR

techniques. A series of thermolysis experiments were conducted for these ligands and the

starting complex [Ti(OPri)4] starting from ambient till the decomposition temperature.

Most of the IR bands resulting from the thermal decomposition of the ligands were

assigned to different species based on literature database available over last forty years.

The decomposition of [Ti(OPri)4] was found to occur through the formation of iso-

propanol and propene as intermediates but the variation was observed depending on

temperature wherein acetone and water were also observed as decomposition products.

The ligands having ß-keto structure were found to decompose through the formation of

Acylketene intermediates. While the complex containing tert. Butylacetoacetate

decomposed through the formation of acetylketene intermediate, the complex containing

ligand thd, decomposed through the formation of pivaloyl ketene. The matrices

containing these intermediates resulted in complex spectra because of various conformers

possible for these species. Bands due to these conformers could be assigned with the help

of photolysis studies. In addition, DFT calculations were carried out and experimental

spectra were compared with simulated spectra of the species which helped to a great

extent to identify the species. In later stage the metal complexes having these ligands

were subjected to a series of thermolysis experiments and assignments of the IR bands

were done based on preliminary studies on each ligand set. A tentative mechanism was

proposed for the decomposition of the titanium complexes in the gas phase based on

these observations. This study throws some light on the reaction mechanism of the gas

phase decomposition of above mentioned precursors through the formation of acylketene

intermediates which were not considered by other reports so far.

Page 122: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 111 -

4.2 Experimental section

To investigate the molecular mechanism involved during precursor decomposition a

matrix isolation apparatus was designed and fabricated. (Fig. 4.1) The following section

describes the details of the apparatus.

4.2.1 Description of matrix isolation apparatus

The matrix isolation apparatus consists of a vacuum line (Pfeiffer TMH 261; Pfeiffer

DUO 5) and an ARS 8200 cryogenic closed cycle system (ARS cryogenics Inc.). The

starting compound can be kept at constant temperature in a small stainless steel vaporizer

connected to high vacuum line through Swagelok® fittings.

Fig. 4.1: The matrix isolation unit used during the present study.

Page 123: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 112 -

Fig. 4.2: (a) Schematic diagram of the refrigeration unit (b) Arrangement of spectroscopic

and photolysis windows attached to a vacuum shroud. (Top view)

Helium pressure connections

Rotatable seals

Heater lead through Thermocouple lead through

Valve power supply

Radiation shield (80 K)

Very cold station (8 K)

Cooled window

Spectroscopic window

Warm flange (a)

Cold station (77 K)

Thermolysis oven

Spectrometer source Spectrometer detector

Photolysis window

Cooled window

(b)

Page 124: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 113 -

A small (~25 mm) window (optically polished cesium iodide suitable for infrared work)

was suspended at the tip of the cryostat within the vacuum shroud (Figs. 4.2 (a) and (b))

and can be cooled to temperatures as low as 9 K. Vacuum windows (CsI 40 mm

diameter, 3 mm thick) on the chamber permit spectroscopic measurements of samples

prepared within. Additional ports permit the admission of the inert gaseous matrix

material (usually argon for IR work) and vacuum ultraviolet light for sample photolysis.

The thermolysis oven consists of an Al2O3 tube with two parallel, inner channels (outer

diameter of 4 mm; inner diameter of 1 mm each) where one of the inner channels is

equipped with a thermocouple (Thermocox: NiCrSi/NiSi) and the other channel is for the

argon/compound mixture. With this set up, reliable thermolysis temperatures could be

measured during the experiment without a contact between the gaseous mixture and the

thermocouple. The oven is constructed in such a way that the heat radiation from the

Al2O3 tube is not conducted to the matrix by having a metal shield surrounding the Al2O3

tube wherein only one small orifice is provided for the passage of gaseous mixture from

the oven and directed to the cold window.

Argon (Linde 6.0) was used as the carrier gas and passed over the compound using a

mass flow controller (flow =1.25 sccm) and the gaseous mixture was passed through an

Al2O3 tube (inner diameter of 1mm; heated by tungsten wire coiled around the last 15

mm). The hot end of the pyrolysis oven was 25 mm away from the cooled CsI window to

assure that a maximum amount of volatile fragments emerging from the oven were

trapped in the matrix. The cold-trap, designed both to prevent back-diffusion of vapors

from the pumps to the very cold sample mount and to protect the pumps from reactive

vapors released when the matrix evaporates. The cold trap was, `flow-through` type with

no abrupt changes of direction or constrictions was attached between vacuum pump and

the cold head. The steps in a typical experiment were; first, the sample window was

cooled to 8-9 K and positioned to face the beam axis of the spectrometer where a

background spectrum of the bare cooled window was recorded.

Afterwards, the window was rotated to face the sample deposition ports. Precursor vapor

was generated by sublimation of precursor sample which was placed in a stainless steel

vaporizer and attached to the system through standard stainless steel Swagelok® vacuum

fittings. The inert gas flows from a cylinder, through a length of stainless steel tubing,

Page 125: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 114 -

(gas flow controlled by mass flow controller) through a mass flow controller and into the

precursor container. The two vapor streams coalesce and are carried through the oven and

then freeze at the surface of the cold window. Use of a great excess of inert gas ensures

that in the resulting solid solution, the sample molecules are effectively isolated in the

inert matrix. Once a suitable amount of sample has been deposited (generally 45 minutes

to one hour at a flow rate of 1.25 sccm), the cold window was returned to the first

position and its spectrum was recorded and difference spectrum against background

spectrum was recorded. With the operating pressure of the order of the 10-6 mbar, and the

flow rates of 1.25 sccm, the residence times of the gas-phase molecules are calculated to

be around 3.66 x 10-13 sec. The short residence time and the high dilution in the argon gas

flow are expected to limit the surface reactions inside the oven surface, though these

reactions can not be avoided to a certain extent. Oven temperatures ranging from room

temperature (RT) to 1000 °C were used. The IR spectra of the matrices, cooled down to

10 K, were recorded on a Bruker EQUINOX 55 with a KBr beam splitter in the range of

400 to 4000 cm-1 with a resolution of 1.0 cm-1.

4.2.2 Photolysis experiments using an ultraviolet source

The experimental arrangement for photolysis is rather simple. An aperture in the various

heat shields surrounding the matrix is provided, together with a window (quartz, optically

polished 40 mm diameter, 3 mm thick) sealed to the outer vacuum shroud. The ultraviolet

lamp is mounted so that radiation is directed on to the matrix through the window and the

aperture. In an ideal case, one may monitor the disappearance of the precursor (indicated

by the disappearance of corresponding IR bands) or the growth of the product without

interrupting the photolysis. But in the present system employed in our laboratory, the

photolysis window is at right angles to the spectroscopic windows in which case the

matrix and cooling system must be rotated for photolysis and then back again before the

progress of photolysis can be monitored spectroscopically.

4.2.3 Preparation of gaseous mixture of argon and iso-propanol

A 2L-glass balloon having two flanges, one attached to the pressure gauge and the other

to the inlet valve for gases, was dried for 12 hours at 150 °C and evacuated under high

Page 126: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 115 -

vacuum to remove traces of volatiles that could interfere with the measurements. With

the help of a pressure controller, a gaseous mixture of highly volatile sample (iso-

propanol, propionic anhydride, diisopropyl ether etc.) and argon in the ratio 1:1200 was

filled into the glass balloon which was then attached to the matrix isolation apparatus.

The drop in pressure was taken as a guide to control the flow of the gaseous mixture in to

the matrix isolation instrument.

4.3 Results and Discussion

4.3.1 Fundamental aspects of the matrix isolation technique

It is important to know the structure and other properties of individual molecules,

although matter is rarely found in the form of isolated molecules. Intermolecular

interactions dominate the physical nature of the matter in the solid and liquid phases, and

are experimentally observable even in gases, where they are smallest. In general, the

`molecular` properties of a substance can only be deduced from gas-phase studies.[15] The

intermolecular interactions are strongest between chemically reactive species such as

most atoms, free radicals and “high temperature monomers”, all of which can be studied

in the gas phase only at low concentrations and high effective temperatures. Even under

such extreme conditions some species are so reactive that they exist for only a few micro-

or milliseconds after they are formed, so that the study of their molecular properties is a

difficult matter.[15]

The technique of matrix isolation is one result of attempts to overcome some of the

difficulties associated with the study of very reactive molecules. Matrix isolation is a

technique developed over the last five decades which enables reactive, short-lived

molecules to be trapped in a solid matrix (of inert gases) and studied spectroscopically.

In essence, the method involves the trapping of the molecule in a rigid cage of a

chemically inert substance (the matrix) at a low temperature. The rigidity of the cage

prevents diffusion of reactive molecules, which would lead to reaction with other such

species. The inertness of the matrix material prevents loss of reactive molecules by

reaction with their environment. The low temperature, besides contributing to the rigidity

Page 127: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 116 -

of the cage, serves to reduce that rate of possible internal rearrangements that require any

activation energy. Under such conditions molecules that normally have very short

lifetimes can be preserved indefinitely and studied at leisure.[9]

In practice, few materials other than the rare gases and molecular nitrogen are chemically

inert enough to serve as matrices for most reactive species. The formation of a rigid

matrix implies the use of temperatures not exceeding above one-third of the melting point

of the solid, i.e. temperatures of 9 K for neon, 29 K for argon, 40 K for krypton, 55 K for

xenon or 26 K for nitrogen. As the lowest temperature attainable using liquid nitrogen as

coolant is 63 K, the triple point of nitrogen, the most inert materials available can only be

used as matrices if colder refrigerants are employed. Only liquid hydrogen and liquid

helium are suitable; they are usable over the ranges 12-13 K and 2-5 K under `boil off`

pressures that can be controlled to adjust the temperature of the liquid. The necessity for

the use of such low temperatures has controlled the development of the technique of

matrix isolation.[15]

In the early 1950s Broida in Washington and Pimental in Berkeley began to use matrix

isolation technique in the study of atoms and reactive molecules, but the method spread

only slowly until the wider availability of liquid helium in the early 1960s and the advent

of microrefrigerators in the late 1960s made it possible for matrix isolation experiments

to be performed widely.[15]

The matrix isolation technique necessarily involves a combination of several distinct

technologies, each of which interacts with the others. The most basic factor, the low

temperature needed to give a rigid matrix, implies cryogenic technology, and in turn

requires the use of high vacuum techniques without which low temperatures can not be

maintained conveniently. The nature of the matrix, the low temperature and the need to

isolate the sample in a vacuum all imply that only spectroscopic methods can be used to

study matrix-isolated species, and the experimental technique is to a large extent

dominated by the need to expose the sample to the spectrometer at the same time as

cooling it in a high vacuum.

A matrix-isolated species is indeed isolated, in the sense that one can only operate on it

under conditions where the matrix is not disrupted. This essentially limits means of

characterization and study to non-destructive spectroscopic methods. These may be

Page 128: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 117 -

further limited by the matrix itself, which must not interfere with the spectrum of the

species under investigation.

4.3.2 Spectroscopic methods

The essence of most spectroscopic techniques is the absorption or emission of

electromagnetic radiation in resonance with a transition between two discrete states of a

molecule or atom. The two states may, for example, be different in the electronic, the

vibrational or the rotational parts of the molecular wavefunction; for atoms only the

electronic wavefunction is relevant.

The main spectroscopic methods used to study matrix-isolated species are electronic

absorption and emission spectroscopy in the visible and ultraviolet regions, vibrational

absorptions spectroscopy in the infrared region and electron spin resonance (e.s.r)

spectroscopy. Electronic and vibrational absorption studies are usually carried out on

samples deposited on cooled windows transparent to the corresponding radiation. During

the course of this study vibrational spectroscopy in the infrared region was studied and

hence will be detailed in following section.

4.3.3 Vibrational spectroscopy in the infrared region

There are only two classes of substances (isolated atoms and homonuclear diatomic

molecules) which have no vibrational spectrum in `the infrared` region. Commercial

spectrometers are available covering the range containing all known stretching vibrations

(~ 4000 to 100 cm-1); most deformation vibrations fall in the same range, though they are

lower in energy than the stretching vibrations involving the same atoms. In practice the

region in which KBr is transparent and can be used as a window material for the matrix

and within the spectrometer (? > 400 cm-1) has been far more extensively studied than the

`far infrared` below 400 cm-1, but even the latter region is strictly available for study and

is becoming more thoroughly investigated.[9]

The ideal matrix material is rather more limited for infrared region than for ultraviolet

studies, only nitrogen, oxygen and the rare gases being completely transparent over the

normal range. Window materials pose little problem in the normal infrared region.

Potassium bromide, which is cheap and readily available, is used down to 400 cm-1, while

Page 129: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 118 -

cesium iodide or bromide may be used to lower frequencies (down to 250 or 200 cm-1).

The cesium salts are much more expensive but mechanically rather stronger than

potassium bromide. Interestingly cesium iodide transmits radiation down to 120 cm-1

when cooled to matrix temperatures. At wavenumbers below about 400 cm-1 polythene

can be used, though it suffers from several disadvantages. It has an absorption band at 72

cm-1, conducts heat very poorly, and is mechanically weak. Silicon and Germanium

windows may be used below 200 cm-1, and quartz is also useful, transmitting far-infrared

radiation below about 200 cm-1. Most of these `far-infrared` windows suffer the great

disadvantage that they do not transmit in the normal infrared region, so that the

comparison of the two region must be made using two separate samples. For this reason

potassium bromide or cesium iodide windows are preferred wherever possible. Then it is

assured that at least one infrared band from any matrix-isolated species other than an

atom or a homonuclear diatomic molecule.[15]

4.3.4 Sample preparation and concerns

The sample itself must be carefully prepared for high temperature studies. In particular it

must be free from more volatile impurities, since these would vaporize preferentially and

might mask the spectrum of the sample or prevent the formation of the desired species.

The sample container and all hot parts of the oven must be similarly clean; even so, traces

of water, carbon dioxide and so on arise from fingerprints and similar minor

contamination of surfaces during sample preparation and loading the container. These,

and carbon monoxide released during the heating of metal parts of the oven, can

effectively be removed only by prolonged degassing under high vacuum at a temperature

just below that required for sample evaporation before cooling down the apparatus and

beginning the deposition of the matrix.

Wherever possible the sample is arranged so that the evaporated molecules have a

straight-line path to the cold window. If this is not done, a considerable amount of

material will condense at the obstruction, unless this is heated to a temperature above that

of the precursor container.

Any method used for heating a precursor must overcome a number of problems. These

include:

Page 130: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 119 -

• The sample container must not itself vaporize or react with the sample or any

other material in contact with it;

• The amount of radiation reaching the cold window from the hot parts of the oven

must be minimized;

• The transfer of heat from the oven itself to its surroundings by any method

(radiation or conduction) must be minimized, not only to prevent damage to

vacuum seals and the like but also to maintain the efficiency of the oven and

maximize the sample temperature.

An even simpler experiment than that using an oven to generate reactive species is

practicable when the desired species can be made by the vaporizing a volatile precursor

that can be handled at room temperature. In this case, a mixture of matrix gas and the

precursor is prepared and passed through a heated tube (oven) before condensation of the

matrix. The presence of the large excess of the matrix gas is usually helpful, as it reduces

the mean free path of the unstable species and thus reduces the probability of wall

reactions and aggregation before deposition.[15]

4.3.5 In situ generation

In many ways it is easier to prepare a matrix containing the more-or-less stable precursor,

and decompose the precursor in the matrix. The problems of heat transfer from an oven to

the matrix and of loss of reactive species by wall collisions and aggregation before and

during deposition are thereby avoided, and it is much easier to ensure thorough mixing of

matrix gas with a stable precursor than with a rapidly flowing gas stream containing

reactive species.

This last consideration is critical, as the importance of interactions with nearest neighbors

is bound to be increased by any uneven distribution of matrix-isolated species. It is

common-place of elementary physics that a gas expands to fill the volume available to it,

and it is usually assumed that this process is an instantaneous one and is independent of

the presence of another gas. It is found, on the contrary, that the rate of the mixing

process depends on the gases involved, and diffusion through heavy gases like xenon is

particularly slow. Adequate mixing requires either a long time or a very turbulent gas

Page 131: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 120 -

flow if the diffusion is slow, and it is more satisfactory to allow the matrix gas and stable

precursor to mix in a large bulb than simply to inject one gas into a stream of the other.

Uniformity is easier to achieve in a spherical bulb than in a cylindrical vessel.

During our experiments we had to handle at least three of such volatile precursors. When

cooling down of the precursor container was ineffective for maintaining proper dilution

of the matrix, we had to resort to glass balloon preparative methods, which resulted in

nice spectra with proper dilutions.

4.3.6 Chemistry with matrices

Once the precursor and thermolysis products are isolated in the matrix it is quite difficult

to decompose them by any means other than photolysis, unless the matrix also contains

some reactive intermediate that can react chemically with the precursor after diffusion.

Other methods involving bombarding the matrix with high energy particles are best

thought of as special types of photolysis. These two issues are discussed briefly in

following sections.

4.3.7 Rigidity and mobility of matrix material

In a matrix isolation experiment it is important that no changes should take place while

spectra are being recorded, and this implies that the matrix must remain rigid and prevent

any diffusion of reactive species. A useful rule –of –thumb is that a matrix may be taken

as rigid below 30 % of its melting point. Below this temperature essentially no

rearrangement of matrix atoms or diffusion of isolated species is expected. In the range

from 30 % to 50 % of the melting point, the process of annealing of the matrix may

occur. This is basically the rearrangement of the matrix structure at the atomic level

towards the most stable crystal structure. Thus grain growth begins in this temperature

range, while large trapped species will cause a local rearrangement to give the most

stable possible cage. Small trapped species such as atoms and diatomics may be able to

take part in the lattice rearrangement to such an extent that segregation at grain

boundaries or reaction with neighboring trapped species occurs.

Above 50 % of the melting point, the matrix must be regarded as non-rigid. Diffusion of

trapped species now begins, and the ultimate effect is that all impurities will be

Page 132: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 121 -

segregated at grain boundaries and reactive species will be lost, as diffusion will occur

until reaction results in formation of a large and stable species. This ultimate stage can

rarely be achieved in practice because the vapor pressure of the matrix solid becomes too

high and the matrix evaporates.[15]

4.3.8 Ultraviolet photolysis of the precursor in the matrix

The photolysis of a precursor in a matrix can only be efficiently carried out using

radiation that is strongly absorbed by the molecule, and that has sufficient energy to

break chemical bonds. In a few cases it may be possible to photolyze a colored precursor

using visible radiation, but in generally ultraviolet radiation is required. A wide variety of

ultraviolet lamps is available, and most useful ones are discussed here.

All ultraviolet lamps depend on an electrical discharge through a vapor to excite the

ultraviolet radiation. The characteristics of the radiation emitted are controlled by the

nature of the vapor and its pressure. Low pressure lamps tend to produce predominantly

atomic emission lines, because the vapor is largely behaving as isolated atoms under the

high temperature / low pressure conditions. Thus, if the lamp is filled with hydrogen at

low pressure, the emission is largely the atomic hydrogen line at 121.6 nm, while if it is

filled with mercury vapor, the emission lines at 184.9 nm and 253.7 nm are excited. High

pressure lamps on the other hand give emission spectra consisting of broader bands,

which may be continuous or may be made up of many lines. In either case the effect is to

give a broad spectrum of emission. The high pressure hydrogen lamp gives a many-line

spectrum below 160 nm and above 500 nm, as well as a true continuum between 160 nm

and 500 nm. The high pressure mercury lamp gives broad bands over the range 250 nm to

470 nm.

The windows through which the radiation passes between the discharge and the matrix

(including the lamp envelope) are of great importance in governing the transmission of

various wavelengths. Glass transmits only a little way in to the ultraviolet region and is

generally useless for photolysis experiments. Sodium chloride and potassium bromide

both transmit radiation down to 200 nm wavelength and make cheap windows for

irradiation, but they fog easily in damp conditions and are too weak mechanically to be

used for lamp envelopes or windows for matrix vacuum shroud. Quartz is preferred for

Page 133: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 122 -

this purpose as its high strength and chemical and mechanical stability suits it for use in

discharge tubes as well as for windows.[9]

4.3.9 Limitations of matrix isolation techniques for studying gas phase

decomposition of a CVD precursor

Matrix isolation studies can be performed on relatively simple molecules. This is mainly

because interpretation of the data resulting from complex molecules is difficult. Even

with the best efforts to solve the spectra, understanding of the decomposition mechanism

has to be speculated based on intermediates observed during the decomposition reactions.

Added to these problems, most of the CVD processes use reactive gases or more than one

molecular precursor for depositing required material. These factors result in highly

complex reactions taking place either in gas phase or on the substrates. Matrix isolation

technique can not be effectively used for the purpose of studying with additional gases or

with precursors containing two different ligand systems. The best way of using this

technique is in combination with other in-situ methods like in situ MS or GC-MS which

provide complementary data in order to get better understanding.

4.3.10 The FTIR spectra of [Ti(OPri)4] and iso-propanol

Vibrational spectroscopy is an excellent method for structural analysis and for the

determination of molecular interactions. [Ti(OPri)4] (TTIP) is one of the most widely

studied transition metal oxide precursors. The vibrational spectrum is under detailed

investigation with new assignments which are not in agreement with each other.[16-18]

Titanium alkoxides are known to form dimeric, trimeric and higher molecularity species,

which is a response to the electron deficient nature of the transition metal in these

tetracoordinated compounds.[19] This is achieved through the formation of alkoxide

bridges, where the degree of association or “molecular complexity” is dependent on the

steric constraints of the ligand. Employing ebullioscopic measurements it was found that,

[Ti(OPri)4] has a molecular association of 1.4.[20]

The IR spectra of the matrix isolated TTIP is shown in the figure 4.3. And corresponding

assignments for observed bands are given in table 4.1.

Page 134: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 123 -

Given the presence of four isopropoxy ligands per TTIP molecule, the likelihood of

extensive coupling of the C-O and C-C stretching modes of each isopropoxy ligand, and

the possibility of molecular association, results in a complex spectrum. To reduce the

complexity of the spectra it was necessary to compare the vibrational spectra of TTIP

with that of corresponding alcohol, i.e. iso-propanol. Any differences between these two

spectra can then be correlated with structural differences between the molecules, to help

in understanding of the vibrational spectrum of the TTIP.

Surprisingly there are no reports about the spectra of matrix isolated iso-propanol in the

range 400 to 400 cm-1. So as a supplementary experiment, highly diluted iso-propanol in

argon (1:1000) was matrix isolated using glass balloon preparative method (see

experimental section). The assignment of observed bands is given in table 4.2. The most

intense band in case of matrix isolated iso-propanol appears at 1253.1 cm-1 which could

be assigned to ?s(CCC) mode. The ?(O-H) mode is detected at 3639.4 cm-1. Despite good

dilution, we observed two bands 3482.6 and 3507.1 cm-1, which we assume are due to

molecular associations of iso-propanol.

4000 3500 3000 1500 1000 500

0.0

0.4

0.8

1.2

1.6

2.0

A

Wavenumber [ cm-1]

Fig. 4.3: FT-IR spectrum of matrix-isolated [Ti(OPri)4] (TTIP).

Page 135: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 124 -

Increased drying time of the glass balloon, and dilution to 1:1200 failed to avoid these

associations of iso-propanol observed in the matrix and still we could observed the bands

at 3482.6 and 3507.1 cm-1. Nevertheless, the IR Spectrum of matrix isolated iso-propanol

was effectively used for the comparison of the bands in the complex TTIP. The IR

spectrum of matrix isolated TTIP, showed presence of small amounts of iso-propanol.

These bands showed typical half band widths at 1253.1, 1077.2, 948.9 and 3639.4 cm-1

which were readily assigned to that of iso-propanol.

Table 4.1: IR frequencies for (TTIP) in the bulk phase (literature reported) and matrix-isolated in solid argon.

bulk phase

frequencies

[cm-1][21]

assignment observed

frequencies

[cm-1]

relative

intensity†

430 (CCC) sym. def. 412 vw

509 (Ti-O) str. as. 509 vw

557 (Ti-O) str. as. 557 w

583 (Ti-O) str. as. 580 sh

611 (Ti-O) str. as. 616 s,b

849 (CCC), (Ti-O) str. 853 s

940 (C-O), (Ti-O) str. 947 m

988 (C-O), (Ti-O) str. 1008 vs, b

1115 (CH3), (Ti-O) str. 1132 vvs

1161 (CH3) str. 1165 mw

1331 (C-H) str. 1334 m,b

1362 (CH3) str. s. 1365 m

1376 (CH3) str. s. 1378 m

1450 (CH3) bend. as. 1451 vw

1463 (CH3) bend. as. 1464 vw

2866 (CH3), (C-H) str. 2871 w

2912 (CH3) str. 2916 w,sh

2930 (CH3) str. as. 2937 m

2968 (CH3) str. as. 2980 m

s = strong, vs = very strong, m = medium, b = broad, sh = shoulder, w = weak, vw = very weak, vvw = very very weak

Page 136: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 125 -

Other bands listed in the table 4.1 showed broad band widths and most intense one

among these bands could be detected at 1115 cm-1. Tentative assignments of these bands

were made based on bulk phase studies of this molecule reported in the literature.[21]

There are significant shifts in the wavenumbers assigned in the bulk phase studies to the

one observed in the gas phase using matrix isolation technique. The shifts observed were

in the range 800 to 1200 cm-1, particularly ?(C-O), ?as(Ti-O) and ?r(CH3),?as(Ti-O)

modes showed shifts of around 15 to 20 cm-1.

4000 3500 3000 2500 2000 1500 1000 500

0.00

0.25

0.50

0.75

1.00

1.25

1.50

A

Wave number [cm-1] Fig. 4.4: Matrix isolated iso-propanol. Sample prepared by glass balloon

preparative method, having a 1:1000 dilution in argon.

Page 137: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 126 -

Table 4.2: IR frequencies of matrix-isolated iso-propanol in solid argon and

corresponding assignments.

observed

frequencies

[cm-1]

assignments relative

intensities†

414.1 ds(CCC) bend. m

813.7 ?s(CCC) str.s. m

948.9 (CH3) vvs

1077.2 (C-O), str. m

1128.3 (C-O), str. vvw

1134.7 (CCC) str.as. w

1165.3 (CH3) m

1253.1 (CCC) str.s. vvs

1380.4 (CH3) bend.s. w

1461.4 (CH3) bend. as. vw

1469.5 (CH3) bend. as. w

2921.3 (CH3), (C-H) str. m

2937.4 (CH3) str.s m

2968.2 (CH3) str. as. m

2988.0 (CH3) str. as. s

3482.6 association m

3507.1 association m,b

3639.4 (O-H) str. s

s = strong, vs = very strong, m = medium, b = broad, sh = shoulder, w = weak,

vw = very weak, vvw = very very weak

Page 138: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 127 -

4.3.11 Thermolysis experiments on TTIP

The high vacuum thermolysis of TTIP, with matrix isolation techniques is described.

Since the experiments were conducted under dynamic conditions where in the TTIP is

highly diluted in the argon gas, it was expected that the secondary reactions are unlikely

to occur. The container for TTIP had to be cooled to 0 °C in order to suppress the high

vapor pressure and maintain a proper dilution in the matrix. Initially the intact TTIP

molecule was matrix isolated and a spectrum was recorded in order to ascertain the

existence of intact molecule and then compare it with the spectra of thermolysis products.

Initial thermolysis experiments were carried out in steps of 100 °C to note the dynamic

region where most of the decomposition reactions were taking place. Under matrix

conditions, in the temperature range of 300 to 400 °C there were distinct developments in

terms of appearance/disappearance of different species. Hence, this region was studied

closely with increase in temperature steps of 20 °C till 400 °C.

While first new IR bands started to appear at oven temperatures as low as 100 °C, the

bands belonging to starting material TTIP completely disappear at 390 °C indicating the

completion of thermolysis. From 100 to 300 °C of oven temperatures it was observed that

bands assigned to iso-propanol were growing in intensity along with bands of the starting

complex TTIP. The formation of iso-propanol was indicated by the appearance of strong

bands at 3639.4, 948.9 and 1253.1 cm-1 which were assigned to ?(O-H), ?r(CH3) and

?s(CCC) modes, respectively. The strong bands observed which are in good agreement

with spectra of matrix-isolated iso-propanol. These bands along with 14 other bands are

listed in table 4.2. It must be noted that even though the iso-propanol was being observed,

the broad bands belonging to intact complex molecule TTIP still existed.

Up to 310 °C of oven temperature, bands belonging to propene could be observed only in

low intensities. At 330 °C, all the bands belonging to propene appear in very good

intensity. And thereafter, propene bands persist and are observed till 400 °C with good

intensity. The major bands for propene were observed at 1453.3 and 908.8 cm-1 which

were assigned to das(CH3) and ? (CH2) modes respectively. A list of frequencies observed

for propene and assignments have been given in table 4.3 which are compared to the

matrix-isolation work of Barnes et el. Furthermore we could reproduce all the bands,

except the band at 1212 cm-1, assigned to propene in this temperature range.

Page 139: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 128 -

One point to be noted here is that, though the concentration of iso-propanol was lowered

at temperatures higher than 390 °C, (which was noted by comparison of band intensities),

but that of propene are slightly increased. In addition there was an increase in the

intensities of bands due to water at 1623.8 and 1608.0 cm-1. The bands belonging to

acetone start appearing at a temperature of 320 ° C and these bands were in very less

intensity till 390 °C. The most intense bands appeared at 1721.5 and 1216.4 cm-1 which

are in good agreement with bands for matrix isolated acetone.[23]

Table 4.3. FT-IR frequencies of matrix-isolated propene in solid argon and corresponding assignments.

literature reported frequencies

[cm-1][22]

assignment observed frequencies?

[cm-1]

relative intensities

3091 (CH2) str. as. 3091.8 s 3036 (CH) str.s. 3036.7 s 2983 (CH2) str.s. 2983.8 s 2941 (CH3) str.as. 2940.8 s 2923 (CH3) str.as. 2922.7 s 2859 (CH3) str.s. 2858.8 m 1650 (C=C) str. 1650.4 s 1453 (CH3)

bend.as. 1453.3 vs

1439 (CH3) bend.as.

1438.8 s

1415 (CH) bend. 1415.3 m 1373 (CH3) str.s. 1373.2 w 1212 (CH2) wag. Not observed 1043 ( CH3) tor. 1043.4 s 998 (CH) bend. 998.1 s 932 (CC) str. 932.5 m 908 (CH2) wag. 908.8 vs 578 ( CH2) tor. 578.4 s

s = strong, vs = very strong, m = medium, b = broad, sh = shoulder, w = weak,

vw = very weak, vvw = very very weak

Page 140: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 129 -

A couple of bands could not be assigned to any species. The intense one is at 1820.1

accompanied by weak bands at 916.4, 1893.0 and 1768.30 cm-1. These bands showed no

group behavior and even with thorough observation of these bands we failed to assign

them to any of the species.

Though earlier flash vacuum pyrolysis studies accompanied by NMR and GC-MS

analysis showed the presence of diisopropyl ether in 2% concentration, in our studies we

failed to observe this species.[4] The bands of TiO2, TiOx or that of TiOxHy were not

observed in the matrices. We speculate that any oxides of titanium formed during the

decomposition might have deposited on the inner surface of the Al2O3 oven tube which is

Table 4.4: FT-IR frequencies of matrix-isolated acetone and corresponding assignments.

literature reported

frequencies [cm-1][23]

assignment observed frequencies [cm-1]

relative intensities

529.0 (C-O) bend. 528.9 w 882.3 (CH3) wag. 882.3 w 1091.6 (CH3) wag. 1091.6 w 1216.4 (C-C) str.as. 1216.5 s 1354.0 (CH3) bend.s. 1353.9 s 1361.8 (CH3) bend.s. 1361.6 s 1406.8 (CH3) bend.as. 1406.8 w 1429.4 (CH3) bend.as. 1429.4 w 1721.5 (C-O) str. 1721.8 vs

s = strong, vs = very strong, m = medium, b = broad, sh = shoulder, w = weak, vw = very weak, vvw = very very weak

Ti(OC3H7)4 TiO2 + 4C3H6 + 2H2O (> 330 °C)

Ti(OC3H7)4 TiO2 + 2C3H6 + 2HOC3H7 (<330 °C)

Scheme 4.1

Scheme 4.2

Page 141: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 130 -

hottest part of the oven and effectively failed to travel the length of the oven tube to be

trapped by the matrix. The reaction mechanism for the formation of TiO2 films using

TTIP with the help of MI-IR technique can be classified in to two different temperature

regimes.

These mechanisms are based on observation of different species in high and low

temperature region. In low temperature region the bands due to TTIP started to decrease,

and those of iso-propanol, propene and water started to increase with the temperature.

Also during our experiments iso-propanol was the first product to be detected at

temperatures as low as 100 °C. With increasing steric crowding around a-carbon the

formation of corresponding alcohol is more likely to occur as depicted in reaction scheme

4.1.

Ti

PriO O

CH3C

H

C

H

HH

O

Ti

C

C

H H

H CH3

Ti

PriO OH Ti

O

H

Scheme 4.4

C OH

H3C

H3C

Ti O

C

CH3

HCH

H H C OH

H3C

H3C H

Ti

O

C

C

HH

HH3C

(C3H7O) (OC3H7)

Scheme 4.3

H3C

CH

H3C

OH CH3C CH3

O

+ H2

Page 142: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 131 -

This reaction (path 1 in scheme 4.6) possibly was dominating at low temperature region

up to 300 °C where predominantly iso-propanol could be observed in matrices. And

propene bands are weak in this temperature range. At temperatures higher than 320 °C,

both reactions (1 and 2 in scheme 4.6) might likely to have occurred as, we could observe

the bands due to propene with growing intensity in addition to bands due to iso-propanol.

It is also possible that TTIP decomposes straight to propene and iso-propanol in the

temperature range 100 to 320 °C according to reaction scheme shown in scheme 4.8.

Formation of acetone can be reasoned out in two different ways. One possibility is

through the dehydration of iso-propanol at 320 °C (scheme 4.3) leading to the formation

of acetone and hydrogen. Second possibility is the decomposition of TTIP directly into

acetone, propene and hydrogen. (scheme 4.9) During temperature range 320-390 °C

bands due to propene are the most intense ones and bands of iso-propanol were observed

to be decreased in favour of increasing intensity for water bands. The strong presence of

propene bands above 320 °C supports the fact that straight decomposition of TTIP in

toTiO2, propene and water is predominant. (scheme 4.2). One more possibility would be

that the water released in the reaction shown in reaction schemes 4.5 and 4.2 could then

react with TTIP forming iso-propanol, as shown in reaction scheme 4.7

Ti O

C

CH3

HC

HH

H

OH

HO

H

Ti O

C

C

HH

HH3C

Scheme 4.5

Ti O CH

OC 3H7 CH 3

(C3H7O)2

CH 2 H

Ti O(C3H7O)2 C3H7OH H2CHC CH 3+ +

1

2

Scheme 4. 6

Page 143: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 132 -

Ti(OC3H7)4 + 2H2O TiO2 + 4 (CH3)2CHOH

Scheme 4.7

Ti(OC3H7)4 TiO2 + 2 (CH3)2CHOH + 2 C3H6

Scheme 4.8

Ti(OC3H7)4 TiO2 + CH3COCH3 + 2H2 + C3H6

Scheme 4.9

1800 1600 1400 1200 1000 800

Wavenumber [cm-1]

100 °C

A 200°C

Oven temperature

400°C

Fig. 4.5: An overview spectra of thermolysis of [Ti(OC3H7)4] (TTIP). Bands

due to TTIP disappear and product bands could be seen at higher

temperatures.

Page 144: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 133 -

An important point to be considered here is the role of gas phase collisions proposed by

Griffin.[24-26] The gas phase collisions could have an important role in the reaction

mechanism. It was proposed that due to collision of TTIP molecule with another gas

phase TTIP molecule or carrier gas (argon) molecule, it gets activated. This activated

molecule can undergo any of the above mentioned reactions. During our studies we could

detect only stable species and no hints were found for existence of radicals in the

reaction. Also dilution of the precursor molecule in argon increases the mean free path

which generally hinders collision to a large extent.

1800 1600 1400 1200 1000 800

0.0

0.1

0.2

0.3

A

IP

PI

HH

P

I

P

p

IPP

HH

I

I

I I

I

A

A

AA

XX

Wave number [cm-1]

Figure 4.6: Spectrum of thermolysed [Ti(OC3H7)4] at 400 °C of oven temperature. Most

informative range 1900-800 cm-1 is represented. A: acetone, H: water, P:

propene, I: isopropanol, X: Unidentified bands.

Page 145: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 134 -

4.3.12 Matrix isolation studies on mixed alkoxide complexes of titanium

The experiments to study molecular fragmentations of [Ti(OPri)2(tbaoac)2] were carried

out in the above mentioned matrix-isolation apparatus; and the matrix-isolated species

were analysed by FIIR spectroscopy. Fig. 1.6 shows the schematic of the instrument

employed in this study. The matrix-isolation apparatus can be envisioned as a small CVD

model reactor, where the Al2O3 tube of the thermolysis oven acts as a substrate. In such a

set up the possibility of coupling between the gas phase and surface reactions, which are

certainly important, cannot be ruled out. Titanium tetra isopropoxide (TTIP) and Htbaoac

are the starting compounds used for the synthesis of the precursor [Ti(OPri)2(tbaoac)2]

(refer chapter 2 for details). In order simplify the process of identifying the fragments

generated by the decomposition of the precursor, independent studies on the

fragmentation of TTIP and Htbaoac were carried out. The results of molecular

mechanism involved in the fragmentation of TTIP, which is as such one of the widely

used precursor for deposition of TiO2, are reported in detail in a previous section 4.3.1. In

this section, the fragmentation of the ligand Htbaoac and [Ti(OPri)2(tbaoac)2] precursor

are discussed.

4.3.13 The FTIR spectra of ligand tert. Butylacetoacetate

In general, ß-dicarbonyl compounds, which include diketones, ketoaldehydes, ketoacids

and their esters and amides, may exist in five tautomeric forms: the diketo form (A), two

cis-enolic forms (B and C) and two trans-enolic forms (D and E in Scheme 4.10).[27] The

interconversion of structures B and C does not require the dissociation of the

intramolecular hydrogen bond and it reduces to the migration of a proton between two

oxygen atoms. Therefore it is very fast. All other tautomeric transformations in this

system as a rule are slow processes. For ß-diketones, the non-bonded van der Waals

interactions between X and Y becomes important. The tautomeric equilibria for the ß-

diketones then favor the enol tautomers.[27] In case of these compounds, it must be also

considered whether there is an enol form present or not. There are experimental and

theoretical evidence for existence of enol form. A carbonyl group is conjugated to the

ester and the hydrogen is bonded in such a way that resonance leads to a negative charge

on the carbonyl oxygen and to a positive charge on the atom which carries the bonding

Page 146: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 135 -

proton. Because of hydrogen bonding, resonance is increased, resulting in weakening of

the C=O and C=C bonds. The characteristic C=O stretching bands appear at 1750 and

1728 cm-1 for keto tautomer. The band at 2988 cm-1 originated by the CH stretching

vibration of the enol form indicates the presence of this structure. The observed

frequencies in main spectral region, 1800 to 800 cm-1 together with band assignments are

given in table 4.5.

C

O

X CH

C

O

Y

Z

O

CC

O

YX

Z

HO

CC

O

YX

Z

H

O

X

CC

Y

OH

Z

HOC

X

C O

Y

Z

A B C

D E

Scheme 4.10

4000 3500 3000 2500 2000 1500 1000 500

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

A

Wavenumber [cm-1]

Fig. 4.7: FT-IR spectrum of matrix isolated tert. Butylacetoacetate (Htbaoac).

Page 147: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 136 -

4.3.14 Thermolysis of Htbaoac

Gas-phase pyrolysis of ß-ketoesters using MI-IR spectroscopy had been studied in detail

by Wentrup et al.[28-30] It was shown that ß-ketoesters fragment to give α-oxoketenes

(acylketenes) (scheme 4.11). The thermolysis reaction proceeds through the enol

tautomer b, and presumably through the intermediate c (scheme 4.11). Based on these

known results it is expected that Htbaoac fragments to give a mixture of s-Z and s-E

isomers of acetylketene and tert-butanol (R1 = But, R2 = H, R3 = Me in Scheme 4.11).

High vacuum thermolyses of Htbaoac were carried out in the temperature range of

ambient to 750 °C. While the first new IR bands were seen at temperatures as low as 150

°C, the bands belonging to starting material completely disappear at 650 °C indicating the

completion of the thermolysis. Figure 4.8 shows a representative IR spectrum of matrix-

isolated gas-phase species from a thermolysis of Htbaoac at 650 °C. The most intense IR

bands appear in the range of 2100 to 2200 cm-1 indicating the presence of ketene

intermediates. From the enlargement of the ketene CO stretching region in Fig. 4.8 it can

Fig. 4.8: Thermolysis spectrum of Htbaoac at an oven temperature of 650 °C.

A: acetone, H: water, 2mp: 2-methylpropene: P2O: 1-propene-2-ol,

gBu:gauche 1-butene, cBu: cis 1-butene, sEa: s-E-acetylketene,

sZa: s-Z-acetylketene, ack: acetylketene, tba: tert-butanol.

2000 1600 1200 800 400

0.0

0.1

0.2

ack

ack

tba

tba

P2O

tba

P2O

P2O

tba

A

A

ack

P2O

wavenumbers [cm-1]

4000 3600 3200 2800 2400

0.0

0.1

0.2

sEa

gBu

2mp

2mp

cBu

sEasE

a

2mp

2mp

H

sZa

sZa

+ tb

asZa

sZa

+ sE

a

CO

2

2mp

tba

+ P

2O

gBu

H HH

A

Wavenumber [cm-1]

Page 148: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 137 -

be seen that the most intense bands are found at 2132.7 and 2142.8 cm-1. These values

match well with those published for acetylketene.[30] In accordance with the literature, the

two main IR frequencies are always accompanied by two minor IR bands at 2137.1 and

2147.6 cm-1. Based on experimental and theoretical data, Wentrup et al. could assign the

two IR bands at 2143 and 2148 cm-1 to s-Z-acetylketene and those at 2133 and 2137 cm-1

to s-E-acetylketene.[30]

Each acetylketene isomer shows a distinct pattern of IR bands. Besides the ketene CO

stretches discussed so far, the acetyl group discloses its identity by intense carbonyl

bands at 1681 cm-1 (s-Z-acetylketene) and 1698 cm-1 (s-E-acetylketene), again in perfect

agreement with the published values.[30] Furthermore, we could reproduce all published

IR frequencies of the two acetylketenes in our thermolysis experiments (Fig. 4.8 and

Table 4.6). There are frequencies between 750 cm-1and 400 cm-1 which show similar

behavior with acetylketene bands. On photolysis, the bands belonging to acetylketene

disappear. This characteristic photolysis behavior of acetylketene was helpful in

assigning these bands which were not assigned in earlier studies. The frequencies of

2160 2150 2140 2130 2120

0.2

0.4

0.6

0.821

47

2142 21

32

2137

2149

A

Wavenumber [cm-1] Fig. 4.9: Ketene region of the spectrum. Two strong bands of acetyl ketene

showing two conformers along with CO band at 2149 cm-1.

Page 149: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 138 -

newly assigned bands are listed in the table 4.6. In addition to the ketene IR bands, Fig.

4.9 shows a shoulder at 2149 cm-1. Depending on the thermolysis conditions, a second

weak IR band at 2139 cm-1 appears in this region. These two IR bands are indicative for

CO trapped in argon matrices.[32] CO2 is also one of the reaction products; IR bands at

2345 and 2339 cm-1 unequivocally show its presence (Fig. 4.8). Based on the known

fragmentation of ß-ketoesters (Scheme 4.11), tert-butanol should be formed by the

fragmentation of Htbaoac. With the aid of published matrix IR data[31] we could assign

several IR bands to tert-butanol. The most intense band appears at 1213 cm-1 followed by

a band at 1139 cm-1; a complete list of assigned IR bands is given in table 4.7. Our data

match very well with those reported in the literature.[31]

Two different C4 alkenes could be identified among the thermolysis products, 2-

methylpropene and 1-butene. In accordance with the literature, the most intense IR band

of 2-methylpropene appears at 887 cm-1.[22] Other less intense IR bands could be assigned

to this alkene and they match well with reported values (see experimental section for

details).[22] 1-Butene is known to exist as cis and gauche conformers.[22] Bands due to the

gauche conformer (strong bands at 796, 631, 998, cm-1) were intense compared to the

bands of the cis conformer (554, 1444 cm-1). At oven temperature of 300 °C, bands

belonging to acetone and 1-propen-2-ol were detected. Acetone has an intense band at

O O

R3

R2OR1

O O

R2R3 OR1

O O

OR3

R2

R1HH

O

R2

CO

R3

O

R2

C

O

R3

a b c

+ + R1OH

s-Z-acylketene s-E-acylketene Scheme 4.11: Thermal decomposition of a ß-ketoester leading to two conformers of

ketenes.

Page 150: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 139 -

1721 cm-1 while 1-propen-2-ol show strong absorptions at 1181 and 1002 cm-1 which

match very well with reported values.[33] While acetone bands are weak at all

thermolysis temperatures, those of 1-propen-2-ol are higher in intensity over the entire

temperature range. Water is formed during the course of the thermolysis of Htbaoac. The

intensities of IR bands indicative of matrix-isolated H2O increase with increasing

thermolysis temperatures. The intensities of those IR bands are far more intense than

those of H2O of the ‘background’. Scheme 4.12 summarises the identified species of the

vacuum thermolysis of Htbaoac.

Table 4.5: FT-IR frequencies of matrix isolated ligand Htbaoac and assignments.

observed

frequencies [cm-1]

relative

intensity†

assignment

1750 vs C=O str.

1728 vs C=O str.

1652 m

1635 m OCC enol str.

1453 s CH3 scissors

1414 m (H3)C-O str.

1394 s

1369 s CH3 in phase def.

1326 s CH2 wagging

1266 s C-O(O) str.

1206 w CH3 twisting

1180 s CH2 twisting

1148 s CCC str.

1028 m (H3)C-O str.

985 m C-C(O) str.

940 m C-CH3 str.

852 m CCO def.

801 m CCO def.

s = strong, vs = very strong, m = medium, b = broad, sh = shoulder, w = weak,

vw = very weak, vvw = very very weak

Page 151: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 140 -

Table 4.6: Literature reported and experimental IR frequencies of s-Z- acetylketene and

s-E- acetylketene between 4000 and 400 cm-1

s-Z-acetylketene (sZa) s-E-acetylketene (sEa)

reported frequencies

[cm-1]

observed

frequencies

[cm-1]

intensities† reported frequencies

[cm-1]

observed

frequencies

[cm-1]

intensities†

3095 3095.4 m 3084 3084.8 m

2143 2142.8 vvs 2133 2132.7 vvs

1681 1681.3 vs 1698 1698.4 vs

1434-1421 1434-1421 br 1434-1421 1434-1421 br

1379 1379.1 w 1365 1364.7 s

1345 1345.0 s 1343 1344.4 s

1169 1169.6 s 1221 1221.0 s

1015 1015.0 w 1081 1081.0 w

956 955.9 m 1021 1021.3 w

889 889.0 w 997 997.0 m

663.3 (m), 661 (m), 636.5 (m), 641.2 (s), 600.9 (s), 607.5 (w), 528.8 (m), 512.3 (s), 497 (m), 488 (w), cm-1,

Frequencies from 750 to 400 cm-1are assigned based on photolysis studies and quantum mechanical

calculations (see text for details). s = strong, vs = very strong, m = medium, b = broad, sh = shoulder, w =

weak, vw = very weak, vvw = very very weak

O O

O

O

C

O

HO

H

CO

H3CH3C

H3C

H3CO

H3C

H2C

OH

H2O CO CO2

H2C CH

H3C CH2H2C CH

H3C

CH2

H3C

C

H3C

CH2

H3CC

H3C CH3

OH

Scheme 4.12: High vacuum thermolysis products of Htbaoac at 650 °C

Page 152: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 141 -

Table 4.7: List of reported IR frequencies of tert. butanol in argon matrices compared with observed frequencies from this work.

reported

frequencies [cm-1]

observed frequencies

[cm-1]

intensity† assignment

3626.5 3626.4 s 3622.3 3622.3 m

(OH) str.

2988.0 2988.0 b 2973.1 2972.8 s 2942.7 2942.7 m

(CH3) str.s.

2907.6 2907.6 s (CH3) str.s. 2888.5 2888 vvw 2875.5 2875 vvw

(CH3) str.s.

1490 Not Observed vw 1476.6 1476 sh (CH3) def.as. 1469.3 1469 sh (CH3) def.as. 1464.4 1464 w (CH3) def.as. 1450 1450 m (CH3) def.as.

1391.6 1391 sh 1372.5 1371.8 s

1367.1 1367 sh

(CH3) def.s.

1328.2 1328 b (OH) def. 1241.5 1241 sh 1213.5 1213.5 vs

(CCC) str.as.

1183 Not observed w (assoc) str.s. 1145.1 Not observed sh 1139.5 1139.5 vs

(CO) str.s.

1027.1 1027 vw 1019 1019 s

(assoc) str.s.

1013.4 1013.4 m 921.2 Not observed s

(CH3) def.as.

914.6 914 b (CH3) def.s. 747.8 747 vw 745.9 745 w

(CCC) str.s.

461 Not observed vw 456 456 w

(CCO) def.as.

418 418 vw (CCC) def.s.

s = strong, vs = very strong, m = medium, b = broad, sh = shoulder, w = weak,

vw = very weak, vvw = very very weak

Page 153: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 142 -

Table 4.8:List of reported IR frequencies of 2-methylpropene in argon

matrices compared with observed frequencies from this work.

literature reported

frequency [cm-1]

observed frequency

[cm-1] assignment

3085 3019

3085 3019 CH str.

2942 2983 2996

2942 2983 2995

CH3 asym str.

2884 2893

2884 2892 CH3 asym def.

2871 2860

2871 2860 CH3 sym str

1655 1655 C=C str 1442 1461

1442 1461 CH3 asym def.

1377 1383

1377 Not observed CH3 sym def.

1416 887

1278

1415.8 887

1278.4 CH2 wag.

1053 1058 1141

Not observed 1058

Not observed CH3 rock

970 802

970.3 801.8 CC str.

O(RO)2Ti

O

R

CR1R2

CH2

H

(RO)2Ti=O + ROH + CH2=CR1R2

Scheme 4.13

Page 154: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 143 -

4.3.15 Thermolysis of [Ti(OPri)2(tbaoac)2]

Thermolysis experiments with the complex [Ti(OPri)2(tbaoac)2] were performed between

100 to 600 °C. Thermolysis products could be seen already at oven temperatures as low

as 200 °C. Bands assigned to the starting material disappeared completely at an oven

temperature of 500 °C, indicating the completion of thermolysis.

Fig. 4.10 shows a typical IR spectrum of matrix-isolated species from the thermolysis of

[Ti(OPri)2(tbaoac)2] at 500 °C. Like in the case of the ligand Htbaoac, intense IR bands at

2143, 2133, 1681, and 1698 cm-1 clearly show that a mixture of s-Z- and s-E-acetylketene

has formed. However, the similarity between the fragmentation of [Ti(OPri)2(tbaoac)2]

and Htbaoac goes beyond the ketene formation. All species that have been identified in

the case of the free ligand were found in matrices of the thermolysis experiments of the

Table 4.9: List of reported IR frequencies of But-1-ene in argon matrices compared

with observed frequencies from this work.

literature reported

frequency [cm-1]

Gauche Cis

observed frequency

[cm-1]

Gauche Cis

assignment

2886 2889 2886 br 2889 br Str sym CH2

1439 1444 1439 sh 1444 br CH2 bend

1414 1421 1415 1421 CH2 bend

1316 1323 1316.5 1323 br CH2 wag

1262 1258 1262.58 1256.78 CH2 twist

1175 1180 1175 sh 1180 sh CH2 rock

1076 1128 Not observed Not observed CC str

998 976 998 976 CH=CH wag

796 835 796 835 CH2 rock

631 554 631 554 =CH twist

Page 155: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 144 -

Ti precursor too; i.e. acetone, 1-propen-2-ol, cis and gauche 1-butene, 2-methylbutene,

tert-butanol, CO, CO2, and H2O (see assignments in Fig. 4.10).

A new IR absorption at 3639 cm-1, right in the area for O-H stretches, hinted to a

formation of an alcohol and we speculated that iso-propanol was formed. As reported in

section 4.3, in order to obtain reliable IR data of iso-propanol , we measured IR spectra of

iso-propanol in solid argon in a separate set of experiments (refer experimental section).

Based on these data, we could identify iso-propanol among the thermolysis products by

the set of its four most intense IR bands at 3639, 2988, 1253, and 949 cm-1. In addition to

iso-propanol, we found a second new product, namely propene. The most intense IR band

of this alkene appears at 908 cm-1. A list of detected IR absorptions for propene is given

2000 1600 1200 800 400

0.0

0.1

0.2

0.3

A

sZa

wavenumbers [cm-1]

4000 3600 3200 2800 2400

0.0

0.1

0.2

0.3

CO

2

2mp

+ tb

a

P2O

+ IP

+ P

tba2m

p

ack

tba

+ sE

a +

P

IPIP

2mp

P

ack

sEa

IP

sZa

A

sZa

sEa

gBu

sZa

+ sE

a

H

H H

H

H IP

Fig. 4.10: Thermolysis spectrum of [Ti(OPri)2(tbaoac)2] at an oven temperature

of 500 °C. A: acetone, H: water, P: propene, 2mp: 2-

methylpropene: IP: isopropanol, P2O:1-propene-2-ol, gBu: gauche

1-butene, cBu, cis 1-butene, sEa: s-E-acetylketene, sZa: s-Z-

acetylketene, ack: acetylketene, tba: tert-butanol.

Page 156: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 145 -

in the experimental section, which is in good agreement with published values.[22] IR

bands at 757, 1598, 1694, 1820, and 2088 cm-1 are relatively weak but show typical half-

band widths of matrix-isolated small molecules. Up to date, we could not assign these IR

bands to any species.

Table 4.10: List of reported IR frequencies of 1-propen-2-ol in argon matrices compared with observed frequencies from this work.

literature reported

frequencies

[cm-1]

observed

frequencies

[cm-1]

assignment

3622 3622 OH str

2992 2992 CH2 asy. Str

2978 2978 CH2 sym str

2950 2948 CH2 asym str

2921 2922 CH3 asym str

2835 2837 CH3 sym str

1673 1673 C=C str

1466 1465 CH3 def

1439 1439 CH3 def

1379 1379 CC str.

1331 1331 CO str.

1181 1181 OCC asym str.

1050 Not observed CH3 rock

1002 1002 CH3 rock/CO str

963 963 CH2 def

848 848 CH2 def

821 Not observed OCC sym str.

697 Not observed C=CH2 torsion

494 494 CCOC framework

478 478 OH wag

395 395 C-OH torsion

Page 157: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 146 -

The singularly important task is to rationalize the formation of the detected species.

Matrix-isolation FTIR spectroscopy allows the identification of intermediates, but not the

reaction mechanism, which must therefore remain speculative. The fragmentation of the

ligand Htbaoac seems to follow the known fragmentation pattern of β-ketoesters,

resulting in acetylketene and tert-butanol (Scheme 4.11).[28-30] A second fragmentation

path is illustrated in scheme 4.14. It is feasible that the ester Htbaoac directly eliminates

2-methylbutene to give acetyl acidic acid. We could not identify the latter species, which

presumably undergoes a fast decarboxylation reaction to the detected CO2, acetone, and

1-propene-2-ol. Additionally, the intermediate acetyl acidic acid might eliminate water to

give acetylketene.

Wentrup et al. found that the ease of fragmentation of β-ketoesters correlates with the

availability of the enol form b (scheme 4.11).[29] The Ti precursor [Ti(OPri)2(tbaoac)2]

already contains the β-ketoester in enol form. In analogy to the fragmentation depicted in

schemes 4.11 and 4.14, a possible fragmentation of the Ti precursor is illustrated in

scheme 4.15.

There are several possible explanations for the formation of iso-propanol and tert-butanol

in the thermolysis of [Ti(OPri)2(tbaoac)2]. Hydrolysis of a [Ti]-O-R complex with gas-

phase water or surface OH groups would result in an elimination of ROH. Additionally,

O

O O

H

O

O OH

- CO2

O OH

- H2O

O

CO

Scheme 4.14: Proposed fragmentation of Htbaoac, resulting in acetylketene and

2-methylpropene.

Page 158: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 147 -

the proposed titanium complex d (Scheme 4.15, route B), which contains a carboxyl

group, could intramolecularly undergo hydrolysis to eliminate one equivalent of ROH.

The flash vacuum pyrolysis of titanium complexes of the type [Ti(OR)4] has been

investigated using glass and quartz tubes at 550 and 700 °C.[4] Volatile reaction products

were collected in a liquid-N2 trap and subsequently analysed by NMR, GC, and GC-MS

techniques.

The isopropyl complex [Ti(OPri)4] was found to decompose at 550 °C into propene, iso-

propanol, small amounts of acetone and diisopropyl ether. At 700 °C the fragmentation

seems to be cleaner resulting only in two products propene and iso-propanol. For the tert-

butyl compound [Ti(OBut)4] at both pyrolysis temperatures 2-methylbutene and tert-

butanol were the only volatile products. Obviously, all titanium complexes of our

investigation that exhibit at least two alkoxy groups could fragment according to scheme

4.13. Of course, this includes the proposed intermediate titanium compounds of scheme

4.15. There is neither hints nor spectroscopic evidences for the trapping of TiO2 in the

matrices. We speculate, that Al2O3 oven used in the experiment acts as substrate, where

O

O O

[Ti][Ti] O

O

CO

[Ti] = Ti(OPri)2(tbaoac)

path A

O

O O

[Ti]path B

H

O

O O

[Ti]H

d

- CO2

[Ti] O

Scheme 4.15: Proposed fragmentation pathways of Ti(OPri)2(tbaoac)2]

Page 159: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 148 -

TiO2 could be deposited. We did not find any indications that free radicals like R or RO

play a role in the process. The predominant formation of alkenes and alcohols was

interpreted in form of equation (4.13).

4.3.16 The FTIR spectra of ligand 2,2,6,6,-tetramethyl-3,5-heptane dione (Hthd)

As mentioned in section 4.3.12, ß-diketones may exist in five tautomeric forms. Nyquist

has reported that carbon oxygen absorption band of the chelate form of ß-diketones

occurs in the region 1606-1620 cm-1 accompanied by much weaker absorption between

1712 and 1720 cm-1 assigned to the carbon oxygen absorption of the ß-diketonate

structure.[34] The FT-IR spectrum of the matrix isolated Hthd ligand is shown in figure

4.11.

The spectrum is void of absorption bands in the region associated with keto form. Instead

a strong absorption (multiplet) is found centered at 1621 cm-1. It is seen clearly that the

ligand Hthd exists mainly in enolic form as evident by the strong bands at 1621 and 1601

cm-1. A less intense band at 1727 cm-1 represents the ß-diketone form of the ligand. The

band at 2988 cm-1 originated by the CH stretching vibration of the enol form indicates the

presence of this structure.

O OO O

H

Scheme 4.16

Page 160: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 149 -

4.3.17 Thermolysis of ligand Hthd

The ligand Hthd has been previously studied by other groups for thermal decomposition

using matrix isolation techniques. But the focus has been to investigate the chemistry of

pivaloyl ketenes generated during such a reaction. [35] In addition, Wentrup et al. reported

the flash vacuum pyrolysis of the a-tert-butyl-ß-ketoesters resulting in the formation of

pivaloyl ketenes.[29] But a clear mechanism involving the a-oxoketenes is not reported so

far. Based on known results so far we understand that the ß-diketonates with a-tert-butyl

groups exist predominantly in enol forms and decompose only sluggishly to afford tert-

butyl a-oxoketenes. High vacuum thermolyses of Hthd were carried out in the

temperature range of ambient to 1000 °C. It was found that Hthd is a stable molecule

compared to ß-ketoester ligand Htbaoac. There were no thermolysis products till the oven

temperature of 400 °C and spectra indicate the presence of intact molecule up to 400 °C.

New IR bands could be seen at 500 °C along with the bands of starting material Hthd.

4000 3500 3000 2500 2000 1500 1000 500

0.0

0.5

1.0

1.5

2.0

2.5

3.0

A

Wavenumber [cm-1] Fig. 4.11: FTIR spectra of matrix-isolated ligand 2,2,6,6,-tetramethyl-3,5-heptane

dione (Hthd)

Page 161: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 150 -

Even at 1000 °C the bands belonging to starting material appear in mid level intensity

and assignments of the bands are hindered to some extent. Due to limitations posed by

the overheating of the oven element and necessity to maintain the matrix at low

temperatures, further experiments with higher temperatures could not be performed.

Figure 4.12 shows a representative IR spectrum of matrix-isolated gas-phase species from

a thermolysis of Hthd at 1000 °C. The most intense IR bands appear in the range of 2100

to 2200 cm-1 indicating the presence of ketene intermediates. From the enlargement of the

ketene CO stretching region in Fig. 4.12 it can be seen that the most intense bands are

found at 2142.7 and a shoulder with lower intensity at 2134.6 cm-1. In accordance with

the literature, these two main IR frequencies are always accompanied by two minor IR

bands at 1667 and 1681 cm-1. Based on experimental and theoretical data,[29] R. L.

Toung, et al. could assign the two IR bands at 2142.7 and 2134.6 cm-1 to s-Z-

pivaloylketene and s-E-pivaloylketene respectively. It was observed that s-Z-

pivaloylketene is more favored conformer and hence bands belonging to s-Z-

pivaloylketene were observed in higher intensities compared to those of s-E-

pivaloylketene.[29]

O O

O

H

C

O

O

H

CO

CH2

C

H3C CH3

H3C

HC CH2

(4.17)s-E-pivaloylketene s-Z-pivaloylketene

Scheme 4.17: Thermolysis products of Hthd ligand at 1000 °C.

Page 162: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 151 -

There are no published data over whole mid IR range for pivaloylketene. The bands of

pivaloylketene in the ketene region reported by the thermolysis of the standard ß-

ketoester ligand and those generated by the thermolysis of Hthd ligand found to match

very well. DFT calculations were carried out using Gaussian 98 [37] on s-E-pivaloyketne

and s-Z-pivaloylketne employing B3LYP/6-31 G(d) level calculations and simulated

spectra are compared with observed spectrum. The comparison of observed and

simulated spectra is given in table 4.11. Besides pivaloylketene, 2-methylpropene was

observed in high intensity at a temperature of 500 °C. The temperature range from 500

°C to 900 °C is dominated by the bands of these two species. At 1000 °C, bands due to

propene could also be detected.

2200 2000 1800 1600 1400 1200 1000 800 600 400

0.0

0.1

0.2

0.3

0.4

0.5

0.64000 3800 3600 3400 3200 3000 2800 2600 2400 2200

0.0

0.1

0.2

0.3

0.4

0.5

* ** *

Wavenumbers [cm-1]

p pk

X X X

2mp

p

CO

2

p

H

H

2mp

2mp

ppk

2mp

p

H

H

pk

2mp

+pk

+ p

2mp

2mp

+ p

pk

pk +

CO

A

Fig. 4.12: Thermolysis spectrum of Hthd at an oven temperature of 1000 °C.†

(* indicates the bands due to starting material)

H: water, P: propene, 2mp: 2-methylpropene: pk-pivaloylketene, X-

unidentified bands

Page 163: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 152 -

4.3.18 Thermolysis of [Ti(OPri)2(thd)2]

A series of high-vacuum thermolyses of [Ti(OPri)2(thd)2] was carried out. Between

ambient temperature and oven temperatures of 400 °C the IR spectra of the isolated

complex remains relatively unchanged. Around 500 °C of oven temperature, new IR

bands appeared, indicating the beginning of the thermal decomposition. Thermolyses

experiments up to 1000 °C were carried out. Even at this temperature the bands due to

starting material still exist but their intensities are lowered. As mentioned above, due to

limitations posed by operating at higher temperatures, further high temperature

experiments were not carried out. But data collected up to 1000 °C convincingly provides

the evidence of pivaloyl ketene as intermediate formed during thermal decomposition.

Right from 500 °C up to 1000 °C the bands due to pivaloylketene could be detected in

varying intensities.

2150 2145 2140 2135 2130

0.0

0.1

0.2

0.3

0.4

2134

.6

2149

2138

.7

A

Wavenumbers [cm-1]21

42.7

Fig. 4.13: Ketene region of the spectrum, the most intense band at 2142.7 cm-1 is

assigned to s-Z-pivaloylketene. Bands due to CO2 can be seen at 2138.7

and 1249 cm-1. The weak band at 2134.6 is assigned to s-E-

pivaloylketene.

Page 164: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 153 -

Table 4.11: B3LYP/6-31 G(d) level calculations of frequencies pivaloylketene conformers

and comparison with observed frequencies

IR frequencies s-Z-pivaloylketene

(cm-1)

IR intensities (km/mol)

IR frequencies s-E-

pivaloylketene (cm-1)

IR intensities (km/mol)

Observed frequencies for pivaloylketene

(cm-1)

Relative intensity

556.8 16.1 576.8 29.2 432.6 w

588.1 27.6 768.6 16.3 612 w

756.1 10.2 1041.0 30.6 629.1 w

896.3 11.9 1094.8 101.7 637.8 w

1002.9 103.8 1157.3 91.2 658.4 w

1074.5 64.2 1279.0 15.3 661.3 s

1391.7 255.9 1377.7 75.4 663.2 s

1434.6 17.1 1507.2 11.0 901 w

1507.1 10.7 1517.6 23.3 916 w

1517.6 33.4 1718.7 417.4 932.5 w

1710.9 218.2 2204.4 816.7 997.8 s

2223.1 1075.3 3028.7 23.2 1058.5 m

3024.4 24 3030.6 13.4 1171.9 w

3026.9 18.9 3039.2 21.8 1207 w

3037.9 24.2 576.8 29.2 1230.2 w

3089.8 33.5 768.6 16.3 1355.1 w

3097.7 49.6 1041.0 30.6 1362.7 w

3104.2 33.2 1094.8 101.7 1373.8 m

1380.3 m

1439.3 w

1667 w

1681 w

2134.6 m

2142.7 vvs

3036.5 m

3064.7 m

Page 165: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 154 -

2000 1800 1600 1400 1200 1000 800 600 400

0.00

0.04

0.08

0.12

4000 3600 3200 2800 2400 2000

0.00

0.04

0.08

0.12

Wavenumbers [cm-1]

pk

IP +

p2m

p

pk

IP

p

2mp

?

X

X

X

X

2mp

A

IP +

pk

IP

cBuIP

AIPA

CO2

p

gBu

2mp

p

IP

p A

IP

IPIP

pk &CO

pk

pk

HH

H

HH

**

A

Fig. 4.14: Thermolysis spectrum of [Ti(OPri)2(thd)2] at an oven temperature of

1000 °C. (* indicate the bands due to starting material) A: acetone, H: water,

P: propene, 2mp: 2-methylpropene: IP: isopropanol, P2O: 1-propene-2-ol,

gBu: gauche 1-butene, cBu, cis 1-butene, pk-pivaloylketene, X-unidentified

bands.

The most intense bands observed were those of iso-propanol; the most intense band of

iso-propanol was detected at 1253.4 cm-1 and other bands due to iso-propanol listed in

table 4.2. Two different C4 alkenes could be identified among the thermolysis products,

2-methylpropene and 1-butene. 2-methylpropene has intense band at 887 cm-1. Also other

bands assigned to 2-methylpropene show high intensity throughout the thermolyses

series. Bands due to 2-methylpropene are listed in table 4.8. The ketene region of the

spectrum shows the bands due to pivaloylketene appearing at 2142.7 and 2134.6 cm-1.

These bands are accompanied by bands at 2138.7 and 2149 cm-1 which are due to CO

trapped in matrices. The intense bands of the spectrum are at 2345 and 2339 cm-1 which

are assigned for CO2 trapped in matrix. Propene was observed at 100 °C and it has

Page 166: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 155 -

intense band at 908.7 cm-1 and other bands due to propene are listed in table 4.3. The

intense band assigned to acetone is detected at 1721 cm-1. Other bands due to acetone are

listed in table 4.4. Small amounts of cis-butene and trans-butene were also detected

though all the bands due to these species were not detected; the most intense ones were

clearly visible (refer table 4.9).

The photolysis experiments were carried out along the temperature series to identify the

bands belonging to pivaloylketene. But under different photolysis conditions the new

bands started to appear in the ketene region which increased the complexity of the spectra

and bands due to conformers of pivaloylketene could not be separated as done in case of

acetylketenes. In addition the bands due to other products started to disappear as well

with increasing number of new bands throughout the spectrum. In order to rationalize the

formation of detected species during the thermolysis of the [Ti(OPri)2(thd)2], the

understanding of the structural aspects of the titanium complexes is necessary. Mixed

alkoxide complexes of titanium with ß-diketonate chelating ligands are widely studied. It

has been established that the alkoxy ligands induce a strong trans-effect on the chelating

ligands thus ß-diketonates form a shorter bite angles to the titanium center. One of the

two titanium-oxygen bonds is shorter than the other in the chelate ring. This elongation of

the bond occurs trans to alkoxy ligand attached to the metal center. We speculate that,

this bond is susceptible for cleavage during thermal decomposition. As a result, the

chelate ring opens up and eliminates pivaloylketene and isobutene in the process. The

cleavage of second titanium–oxygen bond leaves behind highly reactive Ti(III) center.

This [R1]2[R2] Ti(III) complex has highly reactive Ti(III) center and we reason out that it

gets easily hydrolyzed by gas phase water or surface OH groups leading to the formation

of iso-propanol.

As experienced in the earlier case there are neither hints nor spectroscopic evidences for

the trapping of TiO2 in the matrices. We speculate, that Al2O3 oven used in the

experiment acts as substrate, where TiO2 could be deposited. Also we could not find any

indications that free radicals like R or RO play a role in the process. The predominant

formation of alkenes and alcohols was interpreted in form of equation (4.13). During our

studies at 100 °C we could observe the propene and formation of propene by thermal

Page 167: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 156 -

decomposition of 2-methylpropene at high temperatures has already been reported

through a radical mechanism proposed in scheme 4.18.[36]

The pivaloylketene intermediate offers a new insight in to the precursor chemistry of

Hthd ligand. The observations during the thermoylses indicate the formation of

pivaloylketene as intermediate over a temperature range of 500 to 1000 °C. The

annealing of matrices to 28 K lead to increase in intensity of CO and CO2 bands at the

cost of bands due to pivaloylketene. This strongly indicates the nature of intermediate

pivaloylketene which is short-lived species in the gas phase. Also this study sheds some

light on the role of Hthd ligand in titanium precursor chemistry. The most plausible

pathway for the decomposition of [Ti(OPri)2(thd)2] is shown in scheme 4.17. The Hthd

ligand not only provides high stability by chelating to the metal center but also on

thermal decomposition, it helps in reducing the metal center and rendering it more

reactive.

O O

O O

H H

-H

O

CO

H

CH3

CH3

C HH

H

-H

CH2

H3C CH3

Scheme 4.16: Decomposition mechanism of Hthd ligand through the formation

of 2-methylpropene and pivaloylketene intermediates.

Page 168: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 157 -

[R1]2[R2]Ti O C

C

CH

C

O

C

[R1]2[R2]Ti O C

C

CH

CO

C

[R1]2[R2]Ti(III)O C

C

CH

CO

O

C

C CC

O

H

R1 = OPri , R2 = thd[R1]2[R2]Ti

O

O

++

-H

Scheme 4.17: Decomposition mechanism of the complex [Ti(OPri)2(thd)2]

through the formation of pivaloylketene intermediates.

Page 169: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 158 -

Band notations:

Following notations were used in the above figures:

† A: acetone, H: water, P: propene, 2mp: 2-methylpropene: IP: isopropanol, P2O: 1-

propene-2-ol, gBu: gauche 1-butene, cBu, cis 1-butene, sEa: s-E-acetylketene, sZa: s-Z-

acetylketene, ack: acetylketene, tba: tert-butanol.

Band intensities: s (strong), vs (very strong), br (broad), w (weak), vw (very weak), m

(medium).

H3C

H3C

CH2 H3C + H2C C CH3

H2C C CH3

H3C

H3C

CH2 +

H2CHC CH3

+

H2C C CH2

CH3

Scheme 4.18: High temperature decomposition of 2-methylpropene to form

propene. No radical intermediates were detected as reported in the

literature.[36]

Page 170: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 159 -

4.4 Summary

The use of matrix isolation coupled with FTIR as a tool for studying the fragmentation of

the MOCVD precursor is demonstrated. To investigate the molecular mechanism

involved in the decomposition of precursor a matrix isolation apparatus was designed and

fabricated. The smaller dimensions of the thermolysis oven used in the present study and

the need for using low pressures in matrix isolation set up increases the Knudsen number

to be at 1532.28 which clearly indicate the heterogeneous surface reactions taking place

inside the oven surface. There is unavoidable coupling between homogeneous gas phase

reactions and heterogeneous surface reaction in such a system.

The thermolysis studies on selected titanium precursors were carried out using MI-IR

techniques. Titanium tetraisopropoxide (TTIP), [Ti(OPri)2(thd)2] are the standard

precursors used for the deposition of TiO2 thin films and newly developed precursor

[Ti(OPri)2(tbaoac)2] were the three systems investigated for thermal decomposition using

MI-IR techniques. The main aim of employing the MI-IR technique to above mentioned

systems is to provide qualitative information about the gas phase during thermal

decomposition of titanium precursors under isolated conditions.

TTIP has been studied for decomposition using several methods and it was found that

iso-propanol is one of the thermolysis products. In the absence of any data on matrix

isolated iso-propanol in the mid IR range, we had used glass balloon method of dilution

to get FTIR spectrum of matrix isolated iso-propanol. Based on this spectrum the iso-

propanol formed during the thermal decomposition of TTIP and other alkoxide

complexes, we could identify IR peaks belonging to iso-propanol.

Thermolyses experiments of TTIP were carried out between ambient temperature and

600 °C. A small amount of iso-propanol was found at temperatures as low as 100 °C. The

most dynamic changes were observed during temperature of 300-400 °C and hence this

range was studied in steps of 20 °C. The themolysis was complete at 390 °C indicated by

the disappearance of bands due to starting material. Three different temperature regimes

where different products could be seen were identified. Up to 300 °C bands due to iso-

propanol were predominant and the bands due to propene were in low intensity. Above

Page 171: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 160 -

320 °C of oven temperature bands due to propene intensified with increasing temperature

till 400 °C. In addition to iso-propanol and propene, bands due to acetone appear from a

temperature of 320 °C and persist along with propene bands till 390 °C. Water was

observed as one of the reaction products indicated by the increase in the intensity of

bands compared to background water of the matrix. There was no hint for possible

trapping of TiO2 in the matrices. It was speculated that TiO2 was deposited on the inner

wall of the matrix oven and so it could not have been trapped in the matrices. Based on

these observations a probable mechanism for the decomposition under different

temperature regimes was proposed.

Matrix isolation studies on the thermolyses of newly developed precursor

[Ti(OPri)2(tbaoac)2] were carried out from ambient to 750 °C. In order to identify the

bands due to different products, a series of thermolyses experiments with ligand Htbaoac

was carried out. It was found that the ligand Htbaoac undergoes thermal decomposition

forming acetylketene intermediates (having the most intense bands in the region 2120-

2150 cm-1) and tert. Butanol. It was found that the acetylketene has two conformers s-E-

acetylketene and s-Z-acetylketene having almost equal energies. So, bands due to these

conformers were detected in almost equal intensities. Other thermolyses products like cis

and trans butane, 2-methyl propene and acetone, 1-propen-2-ol were also observed which

were not reported in earlier studies. Having known the decomposition products of parent

alkoxide TTIP and the ligand Htbaoac it was relatively easy for analysis spectra

generated by the thermolyses of complex [Ti(OPri)2(tbaoac)2]. Based on these

observations the plausible mechanism for the decomposition of the complex

[Ti(OPri)2(tbaoac)2] was proposed. Though there were few bands which could not be

assigned to any species, we failed to assign them to single isolated species. In addition,

we have not observed any bands due to free radicals. Thermoylses of the well known

titanium precursor complex [Ti(OPri)2(thd)2] were carried out. It was thought that a

comparative study on the decomposition of well known precursor and the newly

developed precursor would help us to improve the understanding of the designing issues

in terms of inclusion of ester moieties to well known ß-keto systems.

As a supplementary experiment, the ligand Hthd was thermolysed in a series of

experiments from ambient to 1000 °C. The ligand Hthd started to decompose only at 500

Page 172: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 161 -

°C. The thermolysis is not complete till 1000 °C where we could detect bands due to

starting material along with bands due to the products. The most intense bands due to

products appear in the range 2120-2150 cm-1 and it was found that pivaloyl ketene has

been formed as an intermediate. The most intense bands at 2142.7 cm-1 and 2134.6 cm-1

were assigned to two different conformers of pivaloylketenes namely s-Z-pivaloylketene

and s-E-pivaloylketene. In addition to ketenes, the thermolysis experiments resulted in

the formation of 2-methylpropene and propene were detected.

The complex [Ti(OPri)2(thd)2] was studied for thermal decomposition from ambient to

1000 °C. Up to an oven temperature of 400 °C, the spectrum of the [Ti(OPri)2(thd)2]

remains relatively unchanged. Bands due to products could only be seen at temperatures

above 500 °C. The most intense bands are assigned to pivaloylketene, iso-propanol, 2-

methyl propene, propene, and acetone. Several bands due to starting material were

detected even at temperature of 1000 °C. Based on the above observations the tentative

mechanism involving the cleavage of metal to oxygen bond of the thd ligand and

subsequent formation of pivaloylketene intermediates were proposed.

Page 173: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 162 -

4.5 References

[1] M. L. Hitchman, K. F. Jensen, Chemical vapor deposition-Principles and

applications, Academic Press, 1992

[2] B.-O. Cho, J. Wang, J. P. Changa, J. Appl. Phys. 2002, 92, 4238.

[3] N. Dietz, H. Born, M. Strassburg, V. Woods, Mat. Res. Soc. Symp. Proc. 2004,

798, Y10.45.1.

[4] M. Nandi, D. Rhubright, A. Sen, Inorg. Chem. 1990, 29, 3065.

[5] S.-I. Cho, C.-H. Chung, S. H. Moon, J. Electrochem. Soc. 2001, 148, C599.

[6] S. Cho, -I., C.-H. Chung, S. H. Moon, Thin solid films 2002, 409, 98.

[7] C. P. Fictorie, J. F. Evans, W. L. Gladfelter, J. Vac. Sci. Technol. A 1994, 12,

1108.

[8] J. P. A. M. Driessen, J. Schoonman, K. F. Jensen, J. Electrochem. Soc. 2001,

148, G178.

[9] I. R. Dunkin, Matrix-isolation techniques, Oxford University Press, Oxford New

York Tokyo, 1998.

[10] Fang Yan, PhD. Thesis, Drexel University, 2003.

[11] S.-I. Cho, C.-H. Chung, S. H. Moon, Thin solid films 2002, 409, 98.

[12] P. Comba, H. Jakob, B. Nuber, B. K. Keppler, Inorg. Chem. 1994, 33, 3396.

[13] A. C. Jones, J. Mater. Chem. 2002, 12, 2576.

[14] A. E. Turgambaeva, V. V. Krisyuk, S. V. Sysoev, I. K. Igumenov, Chem. Vap.

Deposition 2001, 7, 121.

[15] Stephen Cradock , A. J. Hinchcliffe, Matrix isolation : a technique for the study of

reactive inorganic species Cambridge Univ. Press, 1975.

[16] C. G. Barraclough, D. C. Bradley, J. Lewis, I. M. Thomas, J. Chem. Soc. 1961,

2601.

[17] O. Poncelet, J. C. Robert, J. Guilment, Mater. Res. Soc. Symp. Proc. 1992, 271,

249.

[18] V. A. Zeitler, C. A. Brown, J. Phys. Chem. 1957, 61, 1174.

[19] J. Livage, M. Henry, C. Sanchez, Prog. Solid State Chem. 1989, 18, 259.

Page 174: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 163 -

[20] D. C. Bradley, R. C. Mehrotra, P. D. Gaur, Metal alkoxides 1978, Academic

press: New York, NY, 1978. p 63.

[21] P. D. Moran, G. A. Bowmaker, R. P. Cooney, Inorg. Chem. 1998, 37, 2741.

[22] A. J. Barnes, J. D. R. Howels, Faraday Transactions 1973, 69, 532.

[23] A. Engdahl, Chem. Phys. 1993, 178, 305.

[24] K. L. Siefering, G. L. Griffin, J. Electrochem. Soc. 1990, 137, 814.

[25] K. L. Siefering, G. L. Griffin, J. Electrochem. Soc. 1990, 137, 1206.

[26] Q. Zhang, G. L. Griffin, Thin Solid Films 1995, 263, 65.

[27] M. M. Schiavoni, H. E. Di Loreto, A. Hermann, H.-G. Mack, S. E. Ulic, C. O.

Della Vedova, J. Raman Spectrosc. 2001, 32, 319.

[28] B. Freiermuth, C. Wentrup, J. Org. Chem. 1991, 56, 2286.

[29] R. L. Toung, C. Wentrup, Tetrahedron 1992, 48, 7641.

[30] C. O. Kappe, W. M.W., C. Wentrup, J.Org.Chem. 1995, 60, 1686.

[31] J. K.-À. Tommola, Spectrochimica Acta 1978, 34A, 1077.

[32] H. Dubost, L. A. Marguin, Chem. Phys. Lett. 1972, 17, 269.

[33] X. K. Zhang, J. M. Parnis, E. G. Lewars, R. E. March, Can. J. Chem. 1997, 75,

276.

[34] R. A. Nyquist, The Interpretation of Vapor Phase Infrared Spectra, Group

Frequency Data, Vol. 1, Sadtler Research Laboratories, Philadelphia,

Pennsylvania, 1984.

[35] V. A. Nikolaev, Y. Frenk`h, I. K. Korobitsyna, J. Org. Chem. USSR 1978, 14,

1338, 1147, 1433.

[36] M. Szwarc, The Journal of chemical physics 1949, 17, 292.

[37] Gaussian 98, Revision A.11.1, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E.

Scuseria, M. A. Robb, J. R. Cheeseman, V. G. Zakrzewski, J. A. Montgomery,

Jr., R. E. Stratmann, J. C. Burant, S. Dapprich, J. M. Millam, A. D. Daniels, K.

N. Kudin, M. C. Strain, O. Farkas, J. Tomasi, V. Barone, M. Cossi, R. Cammi, B.

Mennucci, C. Pomelli, C. Adamo, S. Clifford, J. Ochterski, G. A. Petersson, P. Y.

Ayala, Q. Cui, K. Morokuma, P. Salvador, J. J. Dannenberg, D. K. Malick, A. D.

Rabuck, K. Raghavachari, J. B. Foresman, J. Cioslowski, J. V. Ortiz, A. G.

Baboul, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R.

Page 175: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 164 -

Gomperts, R. L. Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A.

Nanayakkara, M. Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W.

Wong, J. L. Andres, C. Gonzalez, M. Head-Gordon, E. S. Replogle, and J. A.

Pople, Gaussian, Inc., Pittsburgh PA, 2001.

Page 176: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 165 -

Chapter 5

Thermal properties and hydrolysis behavior of newly

developed titanium precursors Abstract Rationally developed precursors have been screened for their thermal properties in detail.

The newly developed precursors were analyzed by simultaneous TG-DTA analyses. The

simultaneous TG-DTA was performed under ambient of nitrogen and under normal

pressure. The onset of volatilization of the precursors were found to be between 80-150

°C which are lower compared to commercially available titanium precursor

[Ti(OPri)2(thd)2]. Clean volatilization was observed mostly in single step with small

amount of residue being left behind.

The sublimation rates of three precursors [Ti(OPri)2(tbaoac)2] (A), [Ti(OEt)2(tbaoac)2]

(B), [Ti(OPri)2(meaoac)2] (C), which were used for TiO2 depositions using home built

CVD reactor were carried out using isothermal studies. The precursor

[Ti(OPri)2(tbaoac)2] (A) was studied for shelf life using TG analyses and it was found

that residue left behind after TG increased with ageing.

Hydrolytic stability of the selected ß-ketoester complexes was investigated using NMR as

an analytical tool. A comparison to standard ß-diketonate complexes with similar ligand

system was done. It was found that ß-keto ester complexes are susceptible for hydrolysis

at a faster rate compared their ß-diketonate analogues.

Page 177: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 166 -

5.1 Introduction

Precursors play a crucial role for the development of materials by a CVD process. From

the point of view of precursor chemistry, the interaction at the molecular level which

influences the precursor purity, volatility and long term stability as well as reactivity and

decomposition kinetics needs to be investigated. The problem associated with the CVD

of multicomponent oxides is the combination of individual precursors as they have large

difference in their thermal and chemical reactivity. It is necessary to design evenly

matched individual precursor with regard to their thermal properties. The basic thermal

properties essential for a molecular precursor are discussed briefly in following sections.

5.1.1 Volatility

The volatility of a compound is a complex function of intermolecular forces (van der

Waals interactions, pi-stacking or hydrogen bonds) which are affected by molecular

weight, geometry and, for solids, lattice structure.[1] The monomeric precursor molecules

are preferred because they tend to be more volatile compared to oligomeric molecules.[2]

In order to control the oligomerization and to increase volatility, tuning of the steric bulk

around metal center is widely practiced. In addition the manipulation of Lewis acid–base

reactions and thus formation of adducts and the use of fluorinated ligands help to certain

extent to tailoring volatility.[3] Highly fluorinated ligands are known to increase volatility.

It is reported that the substitution of H, CH3, or CR3 groups with fluorine ligands causes

intrermolecular repulsion due to the resulting negative “charge envelope”- and thus

contribute to increase in volatility. [4-5] It is well known fact that the polar groups or the

polarizable groups and high molecular masses tend to reduce volatility. To achieve the

volatility, the strength of the polar interactions has to be minimized. This concept is well

utilized in using donor functionalized ligands. Preferably the donor groups like ether, or

amino groups are widely considered as donor moieties for inducing volatility in resulting

complexes.

Volatility of a substance is best characterized by its vapor pressure. If vapor pressures of

precursor compounds are known, it would be easy to compare their volatilities and find

Page 178: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 167 -

the right evaporation temperature for each precursor so that the precursor flux would be

maintained in the reactor at the right level.[6] However, vapor pressures of many possible

precursor compounds are unknown due to the difficulty of measuring low vapor pressures

of solid substances. Vapor pressure of a liquid is measured easily by observing the

pressure and the temperature at which the liquid boils. Measurement of vapor pressure of

a solid is much more complicated and is usually done with special equipment.[7]

In CVD, the volatility of precursors plays an important role in determining the properties

of a deposited film.[8] Gaseous precursors are preferable to liquids and solids as they can

be readily metered and transported to the reactor.[9] In practice, however, precursors are

often solids or liquids. Liquid and solid precursors should have high vapor pressures and

should not decompose at the vaporizing temperature. Solid precursors are not desired in

CVD because the sublimation rate of a solid may not be constant due to particle sintering

effects and therefore a homogeneous precursor flux may be difficult to maintain.

In conventional CVD of group III-V compound semiconductors, the precursors are

vaporized from a stainless steel container (bubbler) held at moderate temperatures (~ 60-

80 °C) and precursor vapors are carried by the carrier gas flow to the deposition zone.

But in case of oxide materials precursors often possess very low vapor pressure (< 1 torr

at room temperature). This naturally demands higher evaporation temperatures to

maintain the precursor flux to the reaction zone. This means the transport lines leading to

the reactor need to be maintained at higher temperature to prevent the condensation of the

precursor. So according to the need, reactor design has to be modified which needs

additional effort to optimize each precursor for a CVD process.[3]

In addition, incase of the multicomponent system, the individual precursors used should

have matching thermal properties i.e. the temperature of volatilization of individual

precursors and their decomposition temperatures should be as near as possible. The

difference in volatilization and decomposition temperatures of individual precursors

results in difficulty in handling them. Premature reactions taking place on the reactor

walls or lines pose a difficult problem to handle. Thus the CVD of multicomponent

oxides becomes even more difficult with these added process variables.

Page 179: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 168 -

5.1.2 Long term stability

In a CVD process, generally the precursor container is held at an evaporation temperature

as long as the deposition is taking place. This can be for few minutes to several hours.

The ideal precursor should withstand this treatment and still should retain its physical and

chemical characteristics. Added to this, the precursors used in CVD are not freshly

prepared for every run. They are stored under the conditions specified by the

manufacturer before putting them to use. Precursors should be stable to certain period of

time and should perform efficiently during this period.

However, it is difficult to find all these properties in a single precursor. The precursor

may undergo chemical and physical change after prolonged heating. Thereafter it may

not retain same efficiency as before. Any change in the physical and chemical property of

the precursor affects the film quality and reproducibility. Storage of precursor is an

important issue which needs to taken care. There are several possibilities wherein the

precursor degradation can occur. One important possibility is the reaction with air,

moisture and light. Precursors having highly unsaturated metal centers react easily with

air and moisture leading to hydrolysis of the precursor. There are several methods to

prevent and stabilize the metal center such as using appropriate chelating agents,

polyether adducts etc. But one can not expect them to be stable for infinite period in air or

moisture. There is a certain time period within which they withstand the attack by oxygen

and moisture but thereafter they hydrolyze. This leads to inconsistency in depositions.

Also air and moisture and light sensitive precursors need to be studied for stability by the

supplier prior to putting them to use.

As the sublimation rate of precursor is strongly dependent on temperature, variation of

vaporizer temperature, under conditions of constant carrier gas flow rate provides a

means to examine film growth by CVD as a function of precursor partial vapor pressure

under conditions of kinetically limited growth. The poor stability of the precursor often

leads to inconsistent run-to-run reproducibility.[10]

In a CVD process where precursors are solid materials, the sublimation rate of the

precursor in the flowing carrier gas ambient actually determines the rate of precursor

delivery into the reactor. The sublimation rate of a solid precursor is easier to determine

Page 180: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 169 -

and more useful for developing a CVD process than a knowledge of the equilibrium

vapor pressure at a given temperature.[10]

Most of the compounds used as precursors for oxide materials are solids and their

stability is usually estimated by repeated sublimations. The traditional delivery

techniques used for solids and/or liquids with low vapor pressures are direct sublimation

or inert or reactive gas-bubbler methods. These require well-controlled, high-temperature

delivery lines in order to avoid condensation of the precursors. The complexity of the

delivery systems increases with the number of lines required for a multicomponent

material. The mass transport of a solid is a function of the surface area of the powder and

therefore the effective transport rate drops as the solid is consumed.

Moreover, some precursors for oxide materials are unstable at high temperature over long

periods of time. The quest for more stable and more volatile oxide precursors needed for

dielectric, multicomponent ferroelectrics and optical thin films is an example of the

challenges in precursor chemistry in order to ensure reproducibility of the deposits, scale-

up and thus industrial applications. During the course of this work, the focus has been to

synthesize new and improved precursors for the CVD of titanium dioxide thin films.

Therefore the objective of this work has been to study the effect of the variation of the

terminal groups of the mixed alkoxy-ß-ketoester complexes of titanium on their volatility,

thermal stability and decomposition.

5.1.3 Hydrolytic stability

Compounds synthesized during this work are not indefinitely air/moisture stable. The

reaction of a solid, air-sensitive compound with moisture or oxygen upon exposure to air

depends, besides on chemical factors, on several other factors such as the particle size,

relative humidity, temperature, and turnover of the air, just to cite a few of them. These

reactions are normally carried out in a less defined manner, and many of the important

parameters are not exactly known. Therefore, experiments were performed to estimate

quantitatively whether an inclusion of ester moiety in the ligand sphere has significant

influence on the hydrolytic stability of the precursors.

When a new precursor is tested for CVD applications, it would be important to know the

right evaporation temperature so that the flow of vapors and consumption could be tuned

Page 181: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 170 -

to an appropriate level already from the very beginning of the experiments. The above

mentioned issues associated with thermal properties of CVD precursors have to be

analyzed in order to have a qualitative data on the volatility, stability, reactivity of a

precursor before employing them as CVD precursors.

In general, TG (thermogravimetric) measurements have proven to be highly valuable

when evaluating the properties of possible precursor compounds.[11] Dynamic

measurements are used as a routine to evaluate the thermal stability and volatility of the

compounds. Isothermal measurements give information about evaporation rates at certain

temperatures. With thermal analyses performed under reduced pressure, the low pressure

CVD process conditions may be simulated.

5.1.4 Thermal analyses

A group of techniques in which a physical property of a substance and/or its reaction

products is measured as a function of temperature whilst the substance is subjected to a

controlled temperature program.[12-14] There are over a dozen thermal analysis methods

which differ in properties measured and the temperature programs. For the purpose of

evaluating the thermal properties relevant to CVD we have used thermogravimetry and

differential thermal analysis (DTA).

In a thermogravimetric analysis, the mass of a sample in a controlled atmosphere is

recorded continuously as a function of temperature or time as the temperature of the

sample is increased (usually linearly with time). A plot of mass or mass percent as a

function of time is called a thermogram.

Differential thermal analysis is a technique in which the difference in temperature

between a substance and a reference material is measured as a function of temperature

while the substance and reference material are subjected to a controlled temperature

program. Usually, the temperature program involves heating the sample and reference

material in such a way that the temperature of the sample Ts is increased linearly with

time. The difference in temperature ? T between the sample temperature and the reference

temperature Tr (? T = Tr - Ts) is then monitored and plotted against sample temperature to

give a differential thermogram.[15]

Page 182: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 171 -

A single thermal method does not always give sufficient information to allow conclusive

analysis. For example, a downward peak in a DTA experiment means an endothermic

change is occurring at a particular temperature range. One can not infer whether this is a

chemical reaction or a physical change such as melting, or whether any gases are

evolved. A TG experiment on the same sample may show a mass loss over this

temperature range, thereby ruling out melting, but still not identifying any volatiles. So

the combination of several analytical methods gives better profile of the changes taking

place. If any two of the techniques are performed on a single sample at the same time,

then they are known as simultaneous techniques.

Successful synthesis of new class of precursors, (refer chapter 2) has led us to test the

physical and chemical properties of this precursors relevant to CVD using simultaneous

TG-DTA analyses. In addition to thermal properties, the hydrolytic stability of the

selected precursors was carried out using NMR as an analytical tool.

5.2 Experimental section

The titanium complexes, titanium bis(isopropoxide) bis(methylacetoacetate)

[Ti(OPri)2(meaoac)2] (1), titanium bis(ethoxide) bis(methylacetoacetate)

[Ti(OEt)2(meaoac)2] (2), titanium bis(isopropoxide) bis(tert-Butylacetoacetate)

[Ti(OPri)2(tbaoac)2] (3), titanium bis(ethoxide) bis(tert-Butylacetoacetate)

[Ti(OEt)2(tbaoac)2] (4) and titanium bis(isopropoxide) bis(N,N-diethylacetoacetamide)

[Ti(OPri)2(deacam)2] (5) and bis-[(di-ethylmalonato) tetra(isopropoxy)-µ-ethoxy-

titanium(IV)] [Ti2(µ-OEt)2(OPri)4(deml)2] (6) were synthesized by according to the

procedure reported in the literature.[16] (The synthesis and chemical characterization are

described in chapter 2).

Page 183: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 172 -

The complexes were recrystallized in hexane to obtain high purity single crystals of

these complexes. The melting points of the precursor complexes were determined in glass

capillaries and are given in table 5.1. The precursors were purified by repeated

sublimation and the temperature of sublimation at a reduced pressure of 3.0 x 10-2 mbar is

listed in table 5.1 for all the newly synthesized titanium compounds.

The thermal characteristics of the compounds relevant to their suitability as precursors for

CVD were studied by simultaneous thermogravimetric analysis and differential thermal

analysis. Thermal measurements were made using a Seiko 6300S11 system. About 10-15

mg of the finely powdered sample were weighed inside a glove box in to aluminum

crucibles. Analyses were made under pre-purified nitrogen flowing at a rate of 300

ml/min. All samples were heated at a typical heating rate of 5 °C/min.

Thermogravimetry was also used to determine the sublimation rates of three different

precursors used for CVD applications. This was done by recording the mass of the

sample as a function of time after equilibrating the sample temperature at a chosen value.

Hydrolysis studies were carried out using Bruker Advance DPX 250 NMR spectrometer.

1 mmol of the compounds were weighed into the NMR tube and molar amounts of water

were added using micro syringe. NMR was taken immediately after the addition of water

and growing hydrolysis product, iso-propanol was monitored over a period of six hours.

Table 5.1: Melting points and sublimation temperatures (at 3.0 x 10-2 mbar) for

various titanium complexes.

Precursor complex Melting point (°C)

Sublimation temperature

(°C) [Ti(OPri)2(meaoac)2] (1) 59 85

[Ti(OEt)2(meaoac)2] (2) 57 80

[Ti(OPri)2(tbaoac)2] (3) 58 85

[Ti(OEt)2(tbaoac)2] (4) 55 80

[Ti(OPri)2(deacam)2] (5) 88 115

[Ti2(µ-OEt)2(OPri)4(deml)2] (6) 89 110

Page 184: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 173 -

5.3 Results and Discussion

There are several metalorganic precursors used for the deposition of titanium containing

oxide thin films. The most widely used precursors are the halides such as TiCl4, the

alkoxides such as [Ti(OPri)4] and the mixed alkoxide-ß-diketonate complex

[Ti(OPri)2(thd)2]. Of all these precursors [Ti(OPri)2(thd)2] is the most widely used

precursor for thin film deposition of titanium dioxide. This is because the compound

possesses the volatility associated with alkoxides and the stability associated with the ß-

diketonates.

In addition, the thermal properties are more closely compatible with other precursors

such as [Sr(thd)2] and [Ba(thd)2] which are used widely for thin film depositions of

BaSrTiO3.

The bench mark precursor, [Ti(OPri)2(thd)2] has been investigated for thermal behavior in

earlier studies.[17] The DTA curve shows three endothermic peaks. First one at 56 °C,

second one at 176 °C and the third one at 224 °C (Fig. 5.1). The only exothermic peak

was centered at 391 °C. The endothermic peak at 176 °C was assigned to the melting and

the one at 224 °C was assigned for the vaporization of the precursor. Also the exothermic

75 150 225 300 375 450

0

20

40

60

80

100

5.3 %

Temperature [°C]

Wei

ght [

%]

-2

0

2

4

6

DT

A [µV

]

Fig. 5.1: Simultaneous TG-DTA curves for bench mark titanium precursor

[Ti(OPri)2(thd)2].

Page 185: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 174 -

peak at 391 °C was assigned to decomposition of the precursor. The bench mark

precursor shows notable thermal stability compared to parent alkoxide [Ti(OPri)4].

Increased thermal stability of [Ti(OPri)2(thd)2] precursor was successfully utilized for the

deposition of multicomponent oxide thin films using [Sr(thd)2] and [Ba(thd)2] precursors

which showed compatibility in thermal stabilities. Though [Sr(thd)2] and [Ba(thd)2] are

thermally stable, prolonged heating of these precursors during vaporization led to

oligomerization thus reducing the volatility of the precursor.[18]

The simultaneous TG-DTA curves of complexes 1-6 are shown in Fig. 5.2 to 5.7

respectively. The temperature onset of volatilization for the compound,

[Ti(OPri)2(meaoac)2] (1) (Fig. 5.2) begins at 150 °C and there are two steps observed in

the TG curve. The step observed at 150 °C could be attributed to the volatilization and at

220 °C to the onset of decomposition which is evident by the appearance of an

exothermic peak at the same temperature. Beyond 225 °C there is no change in weight

loss observed and the amount of residue left behind is quite small (6.3 %). A sharp

100 200 300 400

0

20

40

60

80

100

6.28 %

Temperature [°C]

Wei

ght [

%]

-6

-4

-2

0

2

4

6

8

10

12

DT

A [µV

]

Fig. 5.2: Simultaneous TG-DTA curves of [Ti(OPri)2(meaoac)2]

Page 186: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 175 -

endotherm in the DTA curve at 45 °C corresponds to the melting point of the compound

which was verified by the melting point determination using the capillary mode.

Although the baseline of the DTA curve for the compound [Ti(OEt)2(meaoac)2] (2) (Fig.

5.3) is poor, there is an endotherm at 57 °C which may be attributed to the melting point

of the compound 2, which was confirmed by melting point measurements by capillary

mode. The precursor seems to decompose in the temperature range 200 to 265 °C. A rest

mass of about 16% is left behind at temperatures above 250 °C.

The compound [Ti(OPri)2(tbaoac)2] (3) was the most widely studied precursor complex

during this work. As depicted in Fig. 5. 4 (a), the compound volatilizes monotonically

without any steps during volatilization. As can be seen from the sharp endotherm in

DTA, the compound melts at 58 °C and volatilizes thereafter. The onset of volatilization

begins at 150 °C and is complete at around 230 °C with almost no compound left behind

indicating that the compound vaporizes completely without decomposition. The residue

left behind is less than 2 % above 250 °C. In order to test the behavior of the complex in

air, one set of TG-DTA analysis was carried out on the precursor complex

100 200 300 4000

20

40

60

80

100

16.0 %

Temperature [°C]

Wei

ght [

%]

-10

0

10

20

30

40

50

60

DT

A [µV

]

Fig. 5.3: Simultaneous TG-DTA curves of [Ti(OEt)2(meaoac)2]

Page 187: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 176 -

[Ti(OPri)2(tbaoac)2] (3), as shown in Fig. 5.4 (b).Though the TG curve shows no

significant deviation from the curve taken with the nitrogen flow, the differences were

observed in DTA curve.

100 200 300 400 500

0

20

40

60

80

100

Temperature [°C]

Wei

ght [

%]

-4

-2

0

2

4

6

8

10

(a)

1.6 %

DT

A [µV

]

100 200 300 400 500

0

20

40

60

80

100

Temperature [°C]

Wei

ght [

%]

-4

-2

0

2

4(b)

6 %

DT

A [µV

]

Fig. 5.4. Simultaneous TG-DTA curves for [Ti(OPri)2(tbaoac)2] (a) under flowing argon

(b) in the absence of argon gas flow

Page 188: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 177 -

In order to evaluate the stability of the precursor when exposed to air, TG-DTA analyses

were carried out under ambient atmosphere. It is seen that the compound shows no

significant change in the TG curve when compared to Fig. 5.4 (a). There were only some

differences observed in the DTA curve and the residue left behind was slightly higher (~

6 %) compared to that carried out under inert atmosphere (1.6 %) This shows that the

precursor is having a better stability compared to the parent alkoxide. One important

point to be noted in such studies is that the sensitivity of the compound towards air or

moisture could not be quantified based only on these experiments. Other analytical

techniques are required wherein controlled exposure to air and moisture would help to

certain extent for quantitative analyses.

The simultaneous TG-DTA curves for compound [Ti(OEt)2(tbaoac)2] (4) are shown in

Fig. 5.5. Apparently mass loss is monotonic without any steps. The onset of volatilization

begins in considerable amount at 170 °C and nearly complete at 225 °C. There are two

endothermic peaks in corresponding DTA curve. First one beginning at 56 °C is sharp

and indicative of melting of the precursor. The second endotherm starts at around 170 °C

and ends at around 225 °C. This peak corresponds to bulk sublimation of the precursor.

The exothermic peak starting at 229 °C could be attributed to the decomposition of the

precursor. Above 280 °C, no mass loss was recorded. The mass loss is not total with

100 200 300 400 500

0

20

40

60

80

100

3.7 %

Temperature [°C]

Wei

ght [

%]

-6

-4

-2

0

2

4

DT

A [µV

]

Fig. 5.5: Simultaneous TG-DTA curves for [Ti(OEt)2(tbaoac)2]

Page 189: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 178 -

residue stabilizing beyond 280 °C at about 3.7 % of the original mass. It should be noted

that in most of the newly developed precursors there is significant weight loss and there

is not enough material to show any characteristic DTA peaks because of the small

amount of residue left behind at high temperatures.

The compound [Ti(OPri)2(deacam2] (5) melts at 64 °C as indicated by the endothermic

peak in the DTA curve. It volatilizes thereafter and in a single step till 400 °C. The

maximum volatilization takes place between 170 and 280 °C where most of the mass loss

was recorded in the TG curve. The decomposition of the compound is marked by a broad

exothermic peak in the DTA curve which begins at around 300 °C. The formation of

TiO2 corresponds to 16.5 % of the initial mass. A residue of around 16.5 % of the initial

mass is left behind which could be TiO2 formed by the decomposition of the precursor

assuming it to be cleanly decomposed.

The TG-DTA data for the complex [Ti2(µ-OEt)2(OPri)4(deml)2] is shown in the Fig. 5.7

[Ti2(µ-OEt)2(OPri)4(deml)2] has a dimeric structure with bridging ethoxy moieties. Also

the metal center is surrounded by three different environments influenced by three

different types of ligands, namely ethoxide, iso-propoxide and diisopropyl malonate. This

100 200 300 400 5000

20

40

60

80

100

16.5 %

Tempearature [°C]

Wei

ght [

%]

-2

0

2

4

6

8

10

12

DT

A [µV

]

Fig. 5.5: Simultaneous TG-DTA curves for [Ti(OPri)2(deacam2].

Page 190: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 179 -

may lead to a dynamic thermal behavior. The endothermic peaks could be probably

attributed to the loss of the ligands as the metal center is surrounded by three types of

ligands. The complex melts at 89 °C marked by an endothermic peak in the DTA curve.

The mass loss occurs even before melting. Most of the mass loss occurs within

temperature range of 150-275 °C. Two endotherms were observed centered around 200

and 275 °C. TG curve shows inflexion points between 225 °C and 275 °C. At

temperatures beyond 275 °C no considerable weight loss was observed with a rest mass

of 15 % is left behind.

During the course of this work, the design changes were included in to the well known ß-

keto systems in the form of ester moieties. In order to evaluate the performance of these

precursors, a comparison of their thermal properties was made. For comparison the two

of the reported precursors were synthesized and thermal properties were analyzed.

Fig. 5.8 shows the comparison of the parent alkoxide [Ti(OPri)4], the bench mark

precursor [Ti(OPri)2(thd)2], and recently reported titanium precursor [Ti(2meip)2] with

newly developed precursor [Ti(OPri)2(tbaoac)2]. The chelating ligands such as Hthd and

Hmeip stabilize the metal center to a great extent. The effect of chelation to these ligands

50 100 150 200 250 300 350 4000

20

40

60

80

100

14.8 %

Temperature [°C]

Wei

ght [

%]

-4

-2

0

2

4

6

8

10

12

14

DT

A [µV

]

Fig. 5.7: Simultaneous TG-DTA curves for [Ti2(µ-OEt)2(OPri)4(deml)2] (6)

Page 191: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 180 -

is that the volatilization is also shifted to higher temperatures. The enhanced chemical

stability has to be compromised for decreased volatility.

It can be seen from the Fig. 5.8 the onset of volatilization of the newly developed

precursor is shifted to higher temperature compared to the parent alkoxide [Ti(OPri)4] and

the temperature onset of volatilization is lower compared to bench mark precursor. Thus

by the inclusion of small design changes in the well established ligand structures it is

possible to tune the thermal properties to a certain extent. And the newly developed

precursors show superior thermal properties like clean volatilization, lower

decomposition temperature, sufficient window between volatilization and decomposition

and negligible residue.

100 200 300 400 500

0

20

40

60

80

100

Wei

ght [

%]

Temperature [°C]

[Ti(OPri)2(thd)

2]

[Ti(Meip)2]

[Ti(OPri)2(tbaoac)

2]

[Ti(OPri)4]

Fig. 5.8: Comparison of TG curves of [Ti(OPri)2(tbaoac)2] (3) with the TG curves of

standard and recently developed titanium precursors

Page 192: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 181 -

5.3.1 Sublimation studies

Three different precursors [Ti(OPri)2(meaoac)2] (1), [Ti(OPri)2(tbaoac)2] (3) and

[Ti(OEt)2(tbaoac)2] (4) were used for deposition of TiO2 thin films by CVD. All these

precursors are solids, and in a CVD process where precursor is a solid material, the

sublimation rate of the precursor in the flowing carrier gas ambient actually determines

the rate of the precursor delivery into the reactor. Therefore detailed sublimation studies

were necessary for the above mentioned precursors to determine the weight loss of each

complex as a function of time at different sublimation temperatures.

For practical use of as a CVD precursor, the sublimation (vaporization) rate at the chosen

temperature must be steady over periods of time involved in typical CVD growth runs,

i.e., 2-3 hours. As depicted in Fig. 5.9, the mass loss is constant over long periods of time

at three different temperatures. This ensures constant mass transport during a CVD

growth process. The sublimation rate increases as a function of temperature.

0 50 100 150 200 25050

60

70

80

90

100

150 °C

125 °C

Wei

ght [

%]

Time [min]

100 °C

Fig. 5.9: Mass loss as a function of time a three different temperatures for

[Ti(OPri)2(meaoac)2]

Page 193: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 182 -

Table 5.2: Sublimation rates of [Ti(OPri)2(meaoac)2] as a function of time at different

temperatures (at atmospheric pressure) as obtained from thermogravimetric

analysis carried out at atmospheric pressure.

Table 5.2 gives the various sublimation rates in mg/min for [Ti(OPri)2(meaoac)2] (1), at

three different temperatures. With increase of sublimation temperature from 150 °C to

175 °C the sublimation rate increased almost four times over from 3.69 x 10-3 mg/min to

18.61 x 10-3 mg/min.

Fig. 5.10: Mass loss as a function of time a three different temperatures for

[Ti(OPri)2(tbaoac)2]

0 50 100 150 200

50

60

70

80

90

100

110

175 °C

125 °C

Wei

ght [

%]

Time [Min]

150 °C

Temperature

(°C)

Sublimation rate

(mg/min)

100 1.35 x 10-3

125 3.70 x 10-3

150 18.60 x 10-3

Page 194: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 183 -

Table 5.3: Sublimation rates of [Ti(OPri)2(tbaoac)2] as a function of time at different

temperatures as obtained from thermogravimetric analysis carried out at

atmospheric pressure.

The mass loss as a function of time for the precursor [Ti(OPri)2(tbaoac)2] (3)

[Ti(OEt)2(tbaoac)2] (4) and is shown in Fig. 5.10 and 5.11 and corresponding

sublimation rates are given in tables 5.3 and 5.4 respectively. It can be seen from these

figures that mass loss is constant over long periods of time and over the temperature

range use. From the Fig. 5.10 the sublimation rate at 175 °C is not constant but is slightly

deviated. The precursor sublimes at a much higher rate at 175 °C.

0 50 100 150 200 25075

80

85

90

95

100

120 °C

100 °C

85 °C

Wei

ght [

%]

Time [min] Fig. 5.11: Mass loss as a function of time a three different temperatures for

[Ti(OEt)2(tbaoac)2]

Temperature

(°C)

Sublimation rate

mg/min

125 3.70 x 10-3

150 10.75 x 10-3

175 15.40 x 10-3

Page 195: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 184 -

Table 5.4: Sublimation rates of [Ti(OEt)2(tbaoac)2] as a function of time at different

temperatures (at atmospheric pressure) as obtained from thermogravimetric

analysis carried out at atmospheric pressure.

5.3.2 Shelf life

The long term stability (shelf life) of the precursor [Ti(OPri)2(tbaoac)2] was studied using

thermogravimetry. One batch of precursor was stored for the purpose in the glove box to

prevent it from degradation. This study was essential because this precursor was used

extensively for deposition of TiO2 and SrTiO3 thin films. The storage either as a solid or

in solution using some organic solvent is essential in order to employ this precursor for

various depositions. In order to address the stability over long term some quantification is

needed. This prompted us to study the precursor over a period of two years. Fig. 5.12

shows the TG curves for the precursor over a period of 2 years. From the TG curves one

can observe that volatilization temperature in case of aged samples is shifted to higher

temperatures by about 20 °C. The decomposition temperature range of all the three

samples remains between 220 at 260 °C. In addition, the residues left behind by the

precursor increased with ageing of the precursor.

Temperature

(°C)

Sublimation rate

mg/min

85 2.20 x 10-3

100 5.90 x 10-3

120 9.25 x 10-3

Page 196: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 185 -

Fig. 5.12: Stability of [Ti(OPri)2(tbaoac)2] in storage; determined by the TG analysis.

It can be rationalized by the fact that the precursor crystallites agglomerate over the

period of time, correspondingly the surface area and hence volatility is affected. In order

to ascertain the chemical stability of the precursor, NMR studies were done on the aged

precursor. The aged samples showed no change when compared to the freshly prepared

sample. Also the NMR studies on the precursor left over in the evaporator after CVD

showed no change when compared to freshly prepared sample. This is advantageous for a

CVD process as the precursor is stable for repeated heating cycles under inert

atmosphere.

5.3.3 Hydrolysis studies

Newly developed precursor systems have to be tested for stability in air and moisture.

This is because the large scale depositions are usually carried out on the shop floor by

non-chemists and sometimes the precursors are handled in situations wherein air and

moisture can not be excluded. But the data on the stability of the precursor in air and

50 100 150 200 250 300 350 400 450

0

20

40

60

80

100

8.3 %

Residues

3.9 %1.6 %

Wei

ght [

%]

Temperature [°C]

After 2 years After 1 year Freshly prepared

Page 197: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 186 -

moisture certainly provides a guideline for handling the precursor in an appropriate

manner.

Fig. 5.13: Hydrolysis studies on the precursors [Ti(OPri)2(tbaoac)2] and

[Ti(OPri)2(meaoac)2] in comparison with well known ß-diketonate

complexes.

During the course of this work, new titanium precursors were developed by the inclusion

of design changes in the ligand sphere of well known ß-diketonate systems in the form of

ester moieties. These changes in the ligand sphere have resulted in the formation of

monomeric structures and found to possess superior thermal properties as discussed

earlier in this chapter. Two of the newly developed precursors namely

[Ti(OPri)2(meaoac)2] (1) and [Ti(OPri)2(tbaoac)2] (3) were studied for hydrolysis

behavior and compared with the similar ß-diketonate complexes. It was thought that the

hydrolysis studies would give us a hint about dependency of hydrolytic stability on the

bulk of the side chain as well on the ester moiety embedded in the side chain. Fig. 5.13

shows the results, when the deliberate hydrolysis was carried out in millimolar amounts

using NMR as an analytical tool.

0 50 100 150 200 250 300 350 400

0

10

20

30

40

50

60

70

80

90

100

Hyd

roly

sis

prod

uct [

%]

Time [Min]

[Ti(OPri)2(acac)

2]

[Ti(OPri)2(meaoac)

2]

[Ti(OPri)2(thd)

2]

[Ti(OPri)2(tbaoac)

2]

Page 198: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 187 -

Two ß-diketonate complexes [Ti(OPri)2(acac)2] and [Ti(OPri)2(thd)2] were selected

because of their wide acceptance as titanium precursors and proximity to the ligand

system used in this study. It can be seen from the Fig. 5.13 that the hydrolysis is fastest in

case of [Ti(OPri)2(meaoac)2] and slowest in case of [Ti(OPri)2(thd)2]. The hydrolysis rates

of [Ti(OPri)2(tbaoac)2] and [Ti(OPri)2(acac)2] are comparable. It can be explained in a

simple way that inclusion of ester moiety in the side chain of ß-keto system, stabilizes the

metal center in monomeric form, but is susceptible for faster hydrolysis than its alkyl

counterparts. This is evident by comparison of hydrolysis curves of [Ti(OPri)2(meaoac)2]

and [Ti(OPri)2(acac)2]. In case of ß-ketoester complex the hydrolysis is complete in two

hours while that of ß-diketonate complex was hydrolyzed only 80 % in five hours.

Interestingly, the bulk on the side chain of the ligand seems to play significant role in

stabilizing the complex against hydrolysis. This is evident from the fact that

[Ti(OPri)2(thd)2] complex is more stable to hydrolysis even after 5 hours indicating only

about 20 % of hydrolysis products where as the [Ti(OPri)2(acac)2] is hydrolyzed almost

90 % during the same period.

Page 199: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 188 -

5.4 Summary

During the course of present study several precursors were developed and tested for CVD

applications. The most important criteria to use them as precursors for CVD application

are thermal properties. Simultaneous TG-DTA studies were carried out on all of the

newly developed precursors. During the course of this study the precursors titanium

bis(isopropoxide) bis(tert-Butylacetoacetate) [Ti(OPri)2(tbaoac)2] (A), titanium

bis(ethoxide) bis(tert-Butylacetoacetate) [Ti(OEt)2(tbaoac)2] (B), Titanium

bis(isopropoxide) bis(methylacetoacetate) [Ti(OPri)2(meaoac)2] (C), were used for the

deposition of TiO2 thin films using home built horizontal cold wall reactor. The

simultaneous TG-DTA was performed under ambient of nitrogen and under normal

pressure. Most of the newly developed precursors showed onset of volatilization between

80-150 °C.

The sublimation rates of these precursors were determined at different temperatures

based on isothermal studies. It was found that the precursor [Ti(OPri)2(tbaoac)2] (A) has

highest sublimation rate of 15.4 x 10-3 mg/min at 175 °C. Sublimation rate of

[Ti(OEt)2(tbaoac)2] (B) at 120 °C was found to be 9.25 x 10-3 mg/min. The sublimation

rate of [Ti(OPri)2(meaoac)2] (C) was found to be of the order of 18.60 x 10-3 mg/min at

150 °C. The compound [Ti(OPri)2(tbaoac)2] (A) was extensively studied for shelf life

after storing for three years under argon ambient. It was found that ageing of the

precursor increased the residue after TG-DTA analysis by about 7% of initial mass. But

the volatilization trend was observed to be preserved with slight deviation from the

freshly prepared samples.

Hydrolysis studies carried on the newly developed precursors using NMR as analytical

tool. The hydrolytic stabilities were compared with standard [Ti(OR)2(ß-diketonate)2]

types of precursors. It was found that inclusion of ester moieties led to decreased

hydrolytic stability of the precursors. Use of homoleptic ß-diketonates as ligands

provided better hydrolytic stability in addition to higher thermal stability. Also the bulk

on the side chain of the ester moiety found to have an effect on the hydrolytic stability of

the precursors. The more bulky ß-ketoesterate complex [Ti(OPri)2(tbaoac)2] has better

hydrolytic stability compared to less bulkier [Ti(OPri)2(meaoac)2].

Page 200: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 189 -

5.5 References

[1] L. G. Hubert-Pfalzgraf, H. Guillon, Appl. Organometal. Chem. 1998, 12, 221.

[2] M. Becht, T. Gerfin, K. H. Dahmen, Chem. Mater. 1993, 5, 137.

[3] A. C. Jones, J. Mater. Chem. 2002, 12, 2576.

[4] A. P. Purdy, A. D. Berry, R. T. Holm, M. Fatemi, D. K. Gaskill, Inorg. Chem.

1989, 28, 2799.

[5] R. Gardiner, D. W. Brown, P. S. Kirlin, A. Rheingold, Chem. Mater. 1991, 3, 45.

[6] A. Niskanen, T. Hatanpää, M. Ritala, M. Leskelä, J. Therm. Anal. Cal. 2001, 64,

955.

[7] G. V. Sidorenko, D. N. Suglobov, Soviet Radiochemistry 1983, 24, 646.

[8] A. N. Gleizes, Chem. Vap. Deposition. 2000, 6, 155.

[9] R. A. Gardiner, P. C. van Buskirk, P. S. Kirlin, Mater. Res. Soc. Symp. Proc.

1994, 335, 221.

[10] A. Devi, Ph. D. Thesis, Indian Institute of Science 1997.

[11] T. Ozawa, Thermochim. Acta 1991, 174, 185.

[12] J. O. Hill, For Better Thermal Analysis and Calorimetry III, ICTA, 1991.

[13] R. C. Mackenzie, Thermochim. Acta 1979, 28, 1.

[14] W. W. Wendlandt, Thermal Analysis, 3rd ed., Wiley, New York, 1985.

[15] D. A. Skoog, F. J. Holler, T. A. Nieman, Principles of Instrumental Analysis,

Saunders College Publishing, 1997.

[16] R. J. Errington, J. Ridland, W. Clegg, R. A. Coxall, J. M. Sherwood, Polyhedron

1998, 17, 659.

[17] J.-H. Lee, S.-W. Rhee, J. Electrochem. Soc. 2001, 148 (6), C409.8

[18] S. R. Drake, M. B. Hursthouse, K. M. Abdul Malik, A. S. Miller, J. Chem. Soc.,

Chem. Comm. 1993, 478.

Page 201: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 190 -

Appendix

General characterization techniques

NMR analysis

Proton- 1H and carbon- 13C NMR spectra were recorded for all synthesized compounds.

The spectra were referenced to residual protic impurities of the internal solvent and

corrected to tetramethylsilane. The integration of peaks and peak intensity analyses were

done using Mestrec® software version 2.30. Bruker Advance DPX 200 and Bruker

Advance DPX 250 spectrometers were used.

CHN analysis

The elemental analyses were performed in the spectrometry and chromatography section

at the Ruhr University Bochum using Elemental, CHNSO Vario EL, Hanau.

Mass spectrometry

Electron Ionization (EI) mass spectra were recorded using ionization energies between 24

eV to 70 eV using CHS-Mass spectrometer “Varian MAT” (Bremen). Output spectra was

given as specific masses (m/z) based on abundant isotopes, 1 1H, 12

6C 14 7N 16 8O, 48

22Ti.

Thermal analysis

Thermal measurements were made using a Seiko 6300S11 system. About 10-15 mg of

the finely powdered sample were weighed inside a glove box in to aluminum crucibles.

Analyses were performed under pre-purified nitrogen flowing at a rate of 300 ml/min. All

samples were heated at a typical heating rate of 5 °C/min.

X-ray diffraction analysis

X-ray diffraction anaylsis were carried out employing a D8-Advance Bruker axs

dffractometer using CuKa radiation (? = 1.5418 Å). XRD instrument consists of CuKa as

X-ray source with nickel filter, and a Goebel mirror with a parallel plate collimator and

Page 202: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 191 -

OED detector. High angle XRD measurements were carried out with ?- 2 ? geometry in

the range 20 – 80° using position sensitive detector.

Matrix isolation

The matrix isolation apparatus consists of a vacuum line (Pfeiffer TMH 261; Pfeiffer

DUO 5) and an ARS 8200 cryogenic closed cycle system (ARS cryogenics Inc.). The

starting compound is kept at constant temperature in a small stainless steel vaporizer

connected to high vacuum line through Swagelok® fittings. A small (~25 mm) window

(optically polished cesium iodide suitable for infrared work) is suspended at the tip of the

cryostat within the vacuum shroud and can be cooled to temperatures as low as 9 K.

Vacuum windows (CsI 40 mm diameter, 3 mm thick) on the chamber permit

spectroscopic measurements of samples prepared within. Additional ports permit the

admission of the inert gaseous matrix material (usually argon for IR work) and vacuum

ultraviolet light for sample photolysis. Typically, argon (purity, 6.0) is used as the carrier

gas and passed over the compound using a mass flow controller (flow =1.25 sccm) and

the gaseous mixture is passed through an Al2O3 tube (inner diameter of 1mm; heated by

tungsten wire coiled around the last 15 mm). The hot end of the pyrolysis oven is

stationed 25 mm away from the cooled CsI window to ensure that a maximum amount of

volatile fragments emerging from the oven can be trapped in the matrix. The IR spectra

of the matrices, cooled down to 10 K, is recorded on a Bruker EQUINOX 55 with a KBr

beam splitter in the range of 400 to 4000 cm-1 with a resolution of 1.0 cm-1.

Scanning electron microscopy

Scanning electron microscopy (SEM) analyses were carried out using LEO-1530 Gemini

SEM instrument equipped with an energy dispersive X-ray analysis unit, Oxford ISIS

EDX system. Prior to analyses, metallic Au was evaporated on the surface of the

specimen to form a conducting layer to avoid electrostatic charging. This facility was

provided by faculty of Geology, Ruhr University, Bochum. Surface morphology of the

films was studied with AFM (SIS pico station situated at Forschungszentrum Jülich).

Page 203: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 192 -

X-ray photoelectron spectroscopy

XPS analyses were carried out with a modified Fisons X-ray photoelectron spectrometer

equipped with an Al Ka X-ray source and a CLAM3 electron energy analyzer. The pass

energy was set to 50 eV. The typical operating pressure was less than 10-8 mbar. All

binding energies were referenced to the substrate signal. Survey X-ray photoelectron

spectra and high resolution spectra were recorded for desired elements.

Electrical measurements

LCR (Inductance/Capacitance/Resistance) meter

The HP 4284 LCR meter is a general purpose LCR meter used for evaluating LCR

components, materials, and semiconductor devices over a wide range of frequencies

(20Hz to 1MHz) and test signal levels (5 mV to 2Vrms, 50µA to 20mArms). The LCR

meter is mostly controlled automatically by a computer with an HP-IB interface (standard

interface for HP equipment for automatic test system). The probe station consists of a

microscope, a wafer chuck and a manipulator. The wafer chuck is connected to the

ground and equipped with vacuum for wafer holding. The manipulator is a device that

enables the user to move the tip in any designated direction. Three knobs move tip: up

and down (1), left and right (2), back and forth (3).

Electrical properties of the metal-insulator-semiconductor (MIS) structures, Capacitance-

Voltage (C-V) characteristics are obtained using HP4284 LCR meter by sweeping the

voltage from inversion to accumulation and back. These analyses were carried out using

facilities situated at Forschungszentrum Jülich.

X-ray fluorescence analyses

X-ray fluorescence (XRF, RIGAKU ZSX-100e) was used for the determination of molar

amount of the individual element in the deposited films. For a particular energy

(wavelength) of fluorescent light emitted by an element, the number of photons per unit

time (generally referred to as peak intensity or count rate) is related to the amount of that

analyte in the sample. The counting rates for all detectable elements within a sample are

usually calculated by counting, for a set time, the number of photons that are detected for

the various analytes "characteristic" X-ray energy lines. By determining the energy of the

Page 204: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 193 -

X-ray peaks in spectrum of a sample, and by calculating the count rate of the various

elemental peaks, it is possible to qualitatively establish the elemental composition of the

sample and to quantitatively measure the concentration of these elements. These analyses

were carried out using the facilities situated at Forschungszentrum Jülich.

Page 205: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 194 -

List of publications

1. Synthesis of nano-scale TiO2 particles by a nonhydrolytic approach,

H. Parala, A. Devi, R. Bhakta, R. A. Fischer, J. Mater. Chem., 2002, 1625-1627.

2. Mononuclear Mixed-ß-Ketoester-alkoxide Compound of Titanium as a

Promising Precursor for Low Temperature MOCVD of TiO2 Thin Films

R. Bhakta, F. Hipler, A. Devi, S. Regnery, P. Ehrhart and R. Waser, Chem. Vap.

Deposition, 2003, 9, 295.

3. MOCVD of TiO2 thin films using a new class of metalorganic precursors

R. Bhakta, U. Patil and A. Devi, Electrochem. Soc. Proc. 2003, 08, 1477.

4. MOCVD of TiO2 thin films and studies on the nature of molecular

mechanisms involved in the decomposition of [Ti(OPri)2(tbaoac)2]

R. Bhakta, R. Thomas, F. Hipler, H. F. Bettinger, J. Müller, P. Ehrhart and A.

Devi, J. Mater. Chem. 2004, 14, 3231.

5. High k dielectric materials by metalorganic chemical vapor deposition;

growth and characterization

R. Thomas, S. Regnery, P. Eharhart, R. Waser, U. Patil, R. Bhakta, and A. Devi,

Ferroelectrics, 2005, in press.

6. Engineered precursors for MOCVD of titanium containing oxide thin films;

precursor chemistry and thin film growth

R. Bhakta, R. Thomas, A. Baunemann, M. Winter, P. Ehrhart, R. Waser and

A. Devi, submitted to Chemisty of Materials, 2004.

7. Gas phase decomposition mechanism involved in the thermal decomposition

of [Ti(OPri)2(thd)2] using matrix isolation technique

R. Bhakta, H.F. Bettinger, J. Müller, and A. Devi, under preparation.

Page 206: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 195 -

Presentations at conferences

MOCVD of TiO2 thin films using a new class of metalorganic precursors

R. Bhakta, U. Patil, S. Regnery, F. Hipler, R. A. Fischer, P. Ehrhart, R. Waser, and

A. Devi. Electrochemical Society, 203rd meeting and EUROCVD XIV Paris, France

April 27- May 2, 2003 (Poster)

MOCVD of TiO2 thin films and studies on the nature of molecular mechanisms

involved in the decomposition of [Ti(OPri)2(tbaoac)2]

From Molecules to Materials: Materials Discussion 7, Queens Mary College, London,

13-15 September, 2004. (Oral presentation)

Page 207: Precursor Chemistry, Thin Film Deposition, Mechanistic Studies

- 196 -

Personal details

Name Raghunandan Krishna Bhakta

Date of birth 20. 06. 1974

Place of birth Chitradurga, India

Nationality Indian

Marital status Married

Academic details

1980-1987 Primary school, MHPS, Nilekani, Sirsi, India

1987-1990 High school, Ave Maria, Sirsi, India

1990-1992 Pre university, MMAS College Sirsi, India

1992-1995 B.Sc., MMAS College, Sirsi, India

1995-1997 M.Sc., Mangalore University, Mangalore, India

Professional experience

1997-1998 Rallis India Ltd. Bangalore, India, as research fellow

1998-2001 Centre for Electronics Design and Technology, Indian Institute of

Science, Bangalore, India, as project assistant

April 2001 Registered for Ph.D at Ruhr University Bochum, Germany