copyright by wei guo 2021

162
Copyright By Wei Guo 2021

Upload: others

Post on 16-Apr-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Copyright By Wei Guo 2021

Copyright

By

Wei Guo

2021

Page 2: Copyright By Wei Guo 2021

The Dissertation Committee for Wei Guo Certifies that this is the approved version

of the following Dissertation:

Transition Metal Oxide Thin Films Integration on SrTiO3

Committee:

Alexander A. Demkov, Supervisor

John G. Ekerdt

Alejandro L. De Lozanne

Maxim Tsoi

Xiaoqin Li

Page 3: Copyright By Wei Guo 2021

Transition Metal Oxide Thin Films Integration on SrTiO3

by

Wei Guo

Dissertation

Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

The University of Texas at Austin

May 2021

Page 4: Copyright By Wei Guo 2021

Dedication

To all the pains and gains in the past years.

Page 5: Copyright By Wei Guo 2021

v

Acknowledgements

It has been a long, long journey and it finally comes to an end. Thanks to all the

pains I have been through in the past years. They grinded me. They completed me.

I would like to thank my parents, for unconditionally supporting me in everything.

For my overseas study, we sacrificed a lot of time that we were supposed to be together.

They raised me up with everything they could give. And I am proud to be their son.

I gratefully thank my advisor, Prof. Alex Demkov, for his guidance and help for

the past years. I admire his intelligent and knowledgeable mind. I appreciate his hard

working and workout spirit.

Many thanks to our lab manager, Agham Posadas, for all his training and help he

has given me. He is always reliable in the lab and knows everything! I also thank all my

group members, Patrick, Kristy, Elliot, Tobi, Marc, Lingyuan, Hosung, Ali, Donghan,

Wente, Fatima, Therese, Ilya, Jackie, and group members from Prof. John Ekerdt group,

Ed, Bryce, Shen, Di, and Annie. Thank you for all your help and for being good lab mates!

Also I thank my friends in Prof. Keji Lai’s group, Lu, Di, Zhaodong, Xiaoyu. Especially

Lu as a good listener and gym partner, and a horse rider. My research life would have been

much worse if I did not know all these fun, nice people!

I thank Prof. Vladimir Strokov for his help during my visit to Switzerland. Thanks

to my undergraduate advisor, Prof. Min Xiao and Prof. Chunfeng Zhang, who has led me

into the world of physics research and provided me with opportunities.

I want to give special thanks to my friend Yuntian Song who gave me faith and

support during my darkest time. I hope he continues well to fulfill his ambition and I

believe he will.

Page 6: Copyright By Wei Guo 2021

vi

I also want to thank my high school physics teacher, Pei Liu, who inspired me and

opened the gate of the physics world to me. I would have never chosen physics as my major

and be a physics Ph.D. if he did not teach me.

And at last, I thank all the time I was alone and not alone. All the sleepless nights I

have been through. All the days filled with frustration and self-doubt. All the experiences

and all the thoughts. I learned more than physics and research in the past seven years. It is

a terminal point of my campus life. But it is just a node in life. The next chapter of my

future is opening. I will keep moving.

Page 7: Copyright By Wei Guo 2021

vii

Abstract

Transition Metal Oxide Thin Films Integration on SrTiO3

Wei Guo

The University of Texas at Austin, 2021

Supervisor: Alexander A. Demkov

Transition metals (TMs) have an immense range of intriguing physical properties

and phenomena. Transition metals and transition metal oxides (TMOs) play a very

important role in modern day scientific research and industry. TMs and TMOs exhibit an

even broader range of structural, electrical, magnetic, and optical properties when

fabricated as thin films on other materials compared to their the bulk forms. SrTiO3 (STO)

is a widely used cubic perovskite oxide material known for its excellent electronic

properties. When TMs and TMOs are integrated on STO, effects based on their interaction

with STO have lead researchers to explore the physics behind such effects and their

possible industrial device applications. In this thesis, we will mainly focus on the

integration of TMs and TMOs with STO and with epitaxial STO grown on Si.

The interactions of the transition metal Pt and rare earth metal Eu when deposited

on STO by molecular beam epitaxy (MBE) were investigated. For Pt growth on STO, I

investigated the properties of ultrathin Pt as a function of coverage on TiO2-terminated

SrTiO3 substrate at different temperatures. I used in situ x-ray photoelectron spectroscopy

(XPS), ex situ scanning electron microscopy (SEM) and atomic force microscopy (AFM)

to observe the evolution of the electronic structure and surface morphology of Pt. I

compared the electronic structure of Pt and the different growth patterns at low and high

Page 8: Copyright By Wei Guo 2021

viii

temperatures. I also performed high temperature annealing of low temperature-grown

samples and found a “bubble-up” behavior of the continuous film.

We also performed ultra-high vacuum deposition of Eu metal on STO (001) and

achieved EuO epitaxy on STO via oxygen scavenging. I explored the oxygen scavenging

behavior of Eu using STO films on Si by varying the STO thickness and Eu deposition

temperature. In situ XPS was used to investigate the electronic structure of the nominal

Eu/STO/Si stack. Our XPS results on the Eu/EuO stack revealed an unusual downward

band bending at the interface. This is supported by density functional theory calculations

by Gao. This work has been published in J. Appl. Phys. 121, 105302 (2017) and J. Appl.

Phys. 124, 235301 (2018). Theoretical calculations performed in our group predicted the

existence of a 2-dimensional electron gas (2DEG) at the EuO/STO interface and

demonstrated that the 2DEG location can be controlled if an additional layer of BaTiO3 is

included. To explore this effect on the 2DEG, I performed soft x-ray angle-resolved

photoemission spectroscopy (SX-ARPES) with our collaborators at the Swiss Light

Source. The results are currently being summarized for publication.

For possible applications in Si photonics, I performed a detailed study of dry

oxidation of a Si substrate below a thin epitaxial SrTiO3. Annealing time and temperature

are the key factors to optimize the SiO2 thickness. I developed a theoretical model based

on a modification of the Deal-Grove-Massoud formalism that predicted the thickness of

SiO2 formed underneath STO as a function of time and temperature. The model fits the

experimental data well. This work has been published in J. Appl. Phys. 127, 055302 (2020).

In addition, I performed preliminary studies on free-standing STO membranes. I developed

a fabrication process and performed Raman measurements.

We also proposed a quantum well structure design of BaSnO3/SrTiO3/Al2O3 to

make use of the conduction offset as large as ~3.5eV between BaSnO3 (BSO) and Al2O3.

The whole deposition process is done by MBE and characterized by reflection high energy

electron diffraction (RHEED) and XPS to confirm the BSO film quality. We are still

working on improving the quantum well quality to be able to make multiple quantum well

structures.

Page 9: Copyright By Wei Guo 2021

ix

Table of Contents

List of Tables .................................................................................................................... xii

List of Figures .................................................................................................................. xiii

Chapter 1: Introduction ........................................................................................................1

1.1 Transition metal and oxide thin films ...................................................................1

1.2 SrTiO3 on Si ..........................................................................................................3

1.2.1 Si ............................................................................................................3

1.2.2 SrTiO3 ....................................................................................................4

1.2.3 SrTiO3 growth on Si ..............................................................................5

1.2.4 SrTiO3/Si and SrTiO3 free-standing membranes ...................................8

1.3 Chapter overview ................................................................................................12

1.4 Reference ............................................................................................................14

Chapter 2: Experiment techniques .....................................................................................19

2.1 Molecular beam epitaxy (MBE) .........................................................................20

2.1.1 Operating principles and instrument illustration .................................20

2.2.2 Preparation for MBE growth ...............................................................23

2.2.3 Maintaining vacuum and cryopump ....................................................24

2.2.4 Effusion cells .......................................................................................25

2.2.5 Electron beam evaporator ....................................................................25

2.2.6 Quartz crystal monitor (QCM) ............................................................26

2.2 Reflection high-energy electron diffraction (RHEED) .......................................26

2.3 X-ray photoelectron spectroscopy (XPS) ...........................................................29

2.3.1 Operating principles .............................................................................30

Page 10: Copyright By Wei Guo 2021

x

2.3.2 Surface sensitivity ................................................................................32

2.3.4 XPS peak analysis ................................................................................32

2.4 References ...........................................................................................................33

Chapter 3: Temperature dependence of the morphology and electronic structure of

ultrathin platinum on TiO2-teminated SrTiO3 (001) ....................................................34

3.1 Introduction .........................................................................................................34

3.2 Experiment ..........................................................................................................37

3.3 Results and Discussion .......................................................................................38

3.4 Conclusions .........................................................................................................47

3.5 Acknowledgment ................................................................................................48

3.6 References ...........................................................................................................48

Chapter 4: EuO epitaxy on SrTiO3 by oxygen scavenging ................................................50

4.1 Introduction .........................................................................................................51

4.2 Experiments on Oxygen Scavenging of STO by Eu...........................................56

4.3 Results and Discussion .......................................................................................58

4.3.1 High temperature deposition with thick STO film ..............................58

4.3.2 High temperature deposition with thin STO film ................................61

4.3.3 Low temperature deposition ................................................................66

4.3.4 Theoretical verification by DFT ..........................................................69

4.4 Further Research on Field-Effect of 2DEG in EuO/STO ...................................74

4.5 Conclusions .........................................................................................................82

4.6 Acknowledgement ..............................................................................................84

4.7 References ...........................................................................................................84

Page 11: Copyright By Wei Guo 2021

xi

Chapter 5: Thermal oxidation of Si buried under thin SrTiO3 and free-standing

SrTiO3 membranes .......................................................................................................87

5.1 Introduction to Dry Oxidation of Si ....................................................................87

5.2 Si Oxidation Experiments ...................................................................................90

5.3 Theoretical Model of Si oxidation ......................................................................97

5.4 Fabrication of Free-standing STO Membranes ................................................106

5.5 Measurements of STO membranes ...................................................................108

5.6 Conclusions .......................................................................................................115

5.7 Acknowledgements ...........................................................................................116

5.8 References .........................................................................................................116

Chapter 6: Advanced design of BaSnO3/SrTiO3/Al2O3 quantum wells ..........................120

6.1 Introduction .......................................................................................................120

6.2 Growth and QW design ....................................................................................122

6.3 characterization .................................................................................................126

6.4 Outlook .............................................................................................................128

6.5 References .........................................................................................................129

Chapter 7: Summary and future work ..............................................................................130

7.1 Summary ...........................................................................................................130

7.2 Future work .......................................................................................................132

Bibliography ....................................................................................................................133

Page 12: Copyright By Wei Guo 2021

xii

List of Tables

Table 1. Summary of the growth conditions and the corresponding lattice parameters

from different growth methods reported in the literature (with published

years). Blanks indicate the corresponding paper did not include detailed

data. (Pressure unit is Torr if not marked differently). .................................11

Table 2. Oxidation parameters of the Massoud model [21] for temperature less than

1000°C. .......................................................................................................103

Table 3. Phonon branch assignments for second-order Raman peaks measured from

STO bulk substrate. .....................................................................................111

Page 13: Copyright By Wei Guo 2021

xiii

List of Figures

Figure 1.1: Transition metals (yellow) in the periodic table frame with lanthanides

and actinides shown separately (light yellow) ................................................2

Figure 1.2: Atomic crystal structure of ABO3 perovskite oxide. This is a unit cell

schematic with green A-site atoms and blue B-site atoms. All small red

spheres are oxygen atoms. ..............................................................................2

Figure 1.3. (a)Top view of the atomic arrangement of STO on the Si (100) surface,

with the STO unit cell rotated 45º with respect to the Si unit cell (b) Side

view of the STO/Si interface. The dimers of Si are bonded to the half

ML of Sr and to oxygen, allowing STO to be epitaxially grown. ..................6

Figure 1.4. Schematic of a waveguide phase modulator. ...............................................12

Figure 2.1: Schematic of the vacuum system at the Advanced Atomic Design Lab. All

instruments are linked to the main transfer line. ...........................................19

Figure 2.2: Front view of MBE with components labeled.................................................21

Figure 2.3: Back view of MBE with components labeled .................................................22

Figure 2.4: Schematic of the principle of RHEED. The angle of incidence of the beam

is usually less than 5º from the surface. ........................................................27

Figure 2.5: RHEED images of (a) commercial STO substrate bought from MTI

Crystal (b) 5 uc of STO grown on a Si substrate. .........................................28

Figure 2.6: Picture of the VG Scienta R3000 XPS system. (Front view) ..........................30

Figure 2.7: Conventional X-ray source for an XPS instrument. (Figure adopted from

[11])...............................................................................................................31

Page 14: Copyright By Wei Guo 2021

xiv

Figure 3.1: Pt 4f core level spectrum for deposition at 200 ºC on STO as a function of

Pt thickness. At 0.25 ML coverage (the lowest) the Pt 4f7/2 peak has a

binding energy of 72.4 eV. It gradually shifts by 1.1 eV to the bulk

metallic value (71.3 eV) when Pt reaches 2 ML. ..........................................39

Figure 3.2: Pt 4f7/2 binding energy as a function of Pt thickness for different

deposition temperatures (200–800 ºC). All values shift to the bulk Pt

metal value (~71.3 eV) at different temperatures when Pt is sufficiently

thick. Lower deposition temperature will lead to larger total binding

energy shift while also taking more layers to reach the bulk metal value. ...41

Figure 3.3: Sr 3d, Ti 2p, O 1s binding energy shift as a function of Pt thickness. There

is a ~0.3 eV shift for all three core levels while Pt is accumulating,

which is consistent with an upward band bending........................................41

Figure 3.4: Pt/STO valence band spectrum grown at 800 ºC as a function of thickness.

It shows the Pt contribution to the signal gradually increase in the range

of 0–3 eV. At 2 ML, it shows nonzero intensity at 0 eV which means the

Fermi edge starts to form. .............................................................................42

Figure 3.5: (a) STO surface after water boiling treatment. It shows a terrace structure

with flat steps (b) AFM measurement of 2ML Pt/STO grown at 800 ºC

and corresponding line profile measurement. The Pt nanoclusters (with

~150nm diameter and 0.5 nm height) arrange themselves in rows aligned

along the substrate terrace edges. ..................................................................43

Page 15: Copyright By Wei Guo 2021

xv

Figure 3.6: (a) SEM image of Pt/STO grown at 200 ºC for 10 ML. At the nanometer

scale we can only see a featureless surface with a clear edge when we

check it at micrometer scale (as shown in the inset). The left area in the

inset has no Pt grown on because of shadowing by the sample holder

during deposition. (b) SEM image of Pt/STO grown at 800 ºC for 1 ML.

It shows uniform nanoclusters with average lateral size of ~80 nm. ............45

Figure 3.7: (a) SEM image of Pt/STO with 10 ML of Pt grown at 200 ºC and

annealed at 800 ºC for an hour in air. The flat surface breaks out and

splits into small nanoclusters. Pt clusters have radii in the 3-5 nm range

with a separation of ~25-30 nm. (b) Histogram of the cluster sizes

obtained from the SEM image. The first two columns are gray because

of the overcounting of the algorithm at very small radius. Red line is the

normal distribution fit of the rest of the columns centered at 3.75 nm. ........47

Figure 4.1: Schematic of EuO/STO epitaxial arrangement: (a) side view (b) top view. ...55

Figure 4.2: RHEED pattern for EuO grown on 10 nm STO/Si at 300 °C. ........................59

Figure 4.3: Aberration-corrected TEM images of the EuO/STO interface with the

atomic model overlaid. Pink circles are Eu, red circles are oxygen, green

circles are Sr, and blue circles are Ti. (a) and (b) HAADF and BF pair of

images for EuO [110]/STO [100] projections. (c) and (d) HAADF and

BF pair of images for EuO [100]/STO [110] projections. ............................60

Figure 4.4: VB spectrum of Eu grown on 10 nm STO/Si. Only Eu2+ and oxygen

features are visible. .......................................................................................61

Figure 4.5: RHEED pattern evolution during Eu deposition on a thin (2 nm) STO

layer...............................................................................................................62

Page 16: Copyright By Wei Guo 2021

xvi

Figure 4.6: (a) VB spectra of Eu on 2-nm STO/Si. Two types of Eu peaks are visible.

(b) Si 2p signal of Eu deposited on 2-nm STO/Si. Silicon metal signal is

almost gone but SiOx and EuSiy appear. .......................................................63

Figure 4.7: VB spectra of step-by-step Eu growth on 3-nm STO. Growth times are

shown with different colors that correspond with the spectra. .....................65

Figure 4.8: (a) Valence band spectrum of Eu on 2-nm STO/Si at 20 °C. (b) Eu 4d

spectrum with clearly resolved multiplets characteristic of Eu metal. .........67

Figure 4.9: Valence band evolution of Eu deposition on 2-nm STO/Si at 20 °C. .............68

Figure 4.10: (a) Valence-band XPS spectrum of 10 min Eu deposition on 2-nm

STO/Si pseudo-substrate at 20 °C. This is consistent with the spectrum

from bulk EuO. (b) Valence-band XPS spectrum of 60 min Eu

deposition at 20 °C. A Eu/EuO/STO/Si structure is formed. (c) The

attenuated and broadened 4f DOS of the whole heterostructure

(“calculated pseudo-XPS spectrum”). The Eu metal 4f DOS is marked as

dark blue, the 4f DOS of each EuO layer is marked as magenta, the total

EuO 4f DOS is marked as violet, and the total 4f DOS is marked as red.

The inset gives the 4f DOS of each EuO layer on a larger scale. (e) A

three-dimensional representation of the “calculated-pseudo-XPS

spectrum.” Here the spectrum of each layer is shown separately along

the y direction. The peak position of 4f EuO DOS of each layer is

connected with a violet line. (Figure is also published in Ref. [35]) ............71

Figure 4.11: Valence band scan of a EuO/STO/BTO/Si sample after transportation in

ambient. .........................................................................................................76

Figure 4.12: Eu resonant map spectrum at the Eu 3d core level. .......................................77

Figure 4.13: Ti resonant map with strong Eu2+ 4f signal. ..................................................78

Page 17: Copyright By Wei Guo 2021

xvii

Figure 4.14: Zoom-in Ti resonant map with possible 2DEG signals (circled in red). .......79

Figure 4.15: Oxygen 1s core level comparison for 1) Ge/EuO/STO/BTO/Si sample

before it was taken out of vacuum (red); 2) The same sample in step 1)

left in air for over 2 days (green); 3) The same sample in step 2) after

being heated in vacuum at 350 ºC for 1 hour (blue). Oxygen from EuO

and GeOx are marked above the peaks. ........................................................80

Figure 4.16: Valence band comparison for 1) Ge/EuO/STO/BTO/Si sample before it

was taken out of vacuum (red); 2) The same sample in step 1) after air

exposure for over 2 days (green); 3) The same sample in step 2) after

heating in vacuum at 350 ºC for 1 hour (blue). Eu signal positions are

marked above the peaks. ...............................................................................82

Figure 5.1 Example ellipsometry measurement with the corresponding model fit.

These are the measurement results at 45°. This gives a STO layer with

17.54 nm and SiO2 layer with 7.54 nm thickness. The red one is the

curve and the blue one is the curve. Measured data are shown as open

circles and the fit curves as solid lines. .........................................................92

Figure 5.2: (a) RHEED image for a 10 nm STO/SiO2/Si after the growth and before

anneal. STO pattern is shown along the [110] direction. (b) RHEED

image for a 10 nm STO/SiO2/Si after 800°C anneal for 2 hours. STO

pattern along the [110] direction is still sharp and clear. ..............................94

Figure 5.3: XPS spectrum for the Si 2p region. The major peak around 104.5 eV is

SiO2 and the minor peak around 103 eV is SiOx, Si metal (99.5 eV) is

also still visible. ............................................................................................95

Page 18: Copyright By Wei Guo 2021

xviii

Figure 5.4: Comparison of the XRD out-of-plane full scans before and after

annealing. STO peaks decrease a little and Si is a little higher. All peaks

do not show any obvious deformation other than intensity variation. ..........96

Figure 5.5: The schematic of oxygen propagation in the SrTiO3/SiO2/Si structure.

The inset shows the atomic structure schematic. ..........................................98

Figure 5.6: Diffusivity of oxygen in SrTiO3 (D1) and SiO2 (D2) in the 300-1200°C

temperature range. Inset: Diffusivity of oxygen in SrTiO3 and SiO2 in the

300-700°C temperature range. The STO low temperature (<700°C) data

is obtained by inverse relationship projection from existing diffusivity at

higher temperature (>700°C) ......................................................................100

Figure 5.7: The SiO2 thickness as a function of oxidation time at different

temperatures from equations (12) and (13). The lines represent the

model and the shapes represent the experimental data. Diamonds are

samples annealed under 800°C, circle is under 750°C and square is

under 700°C. All thickness data have an error bar of ~1 nm. .....................104

Figure 5.8: Schematic of the procedure to fabricate STO membranes ............................107

Figure 5.9: (a) (b) are SEM images from different spots on STO/Si etched sample.

The biggest membrane outlined by yellow is ~15 × 15 μm2 ......................108

Figure 5.10: Raman spectrum of a bulk STO substrate. Characteristic peak positions

are labeled and marked with red arrows. ....................................................110

Figure 5.11: Raman spectrum of 20 nm STO thin film on Si. (Thin film grown by

MBE)...........................................................................................................112

Figure 5.12: Raman signal of STO membranes made from 20 nm STO/Si ....................114

Figure 6.1: Band alignment of a single quantum well. The conduction band offset

between BSO and Al2O3 is 3.5 eV. .............................................................124

Page 19: Copyright By Wei Guo 2021

xix

Figure 6.2: Schematic of the heterostructures to be grown. ............................................124

Figure 6.3: Simulation results for possible energy states in a single quantum well. .......126

Figure 6.4: RHEED evolution from (a) thick Al2O3 on STO buffer layer on Si (b) 3 uc

STO on previous Al2O3 surface (c) 10 uc BSO on previous STO surface..128

Page 20: Copyright By Wei Guo 2021

1

Chapter 1: Introduction

1.1 TRANSITION METAL AND OXIDE THIN FILMS

Transition metals play a very important role in modern day scientific research and

in industry. Transition metals (TMs) are commonly taken as the set of elements located in

Group 3 to 12 of the periodic table, as shown in Figure 1.1. Elements with a partially filled

d sub-shell are identified as transition metals. In addition, lanthanide and actinide metals,

with a fully filled d shell but only partially filled f sub-shell, are also sometimes considered

as transition metals. Electrons of the partially filled sub-shell provide not only multiple

stable valence states, but are responsible for the tremendous variety of interesting physical

properties and phenomena in TMs, including magnetism [1], superconductivity [2]-[4],

catalytic properties [5], ferroelectricity [6], multiferroicity [7], [8], and many more [9].

Among all the many kinds of TM compounds, transition metal oxides (TMOs) are

the main focus of this thesis. Multiple oxidation states of TMs enrich the oxide family with

different electronic configurations. The oxidation process transfers the outer s electrons to

bond with oxygen leaving partially occupied d shell as the highest occupied electronic

states. These outer shell electrons determine the physical properties of TMOs [9], [10].

TMOs crystallize in the perovskite, rock salt, rutile, corundum, spinel and other oxide

crystal structures. Perovskite oxides, described by the chemical formula ABO3 (Figure 1.2),

have been widely studied over many decades due to multiple phenomena they exhibit

including, but not limited to, ferroelectricity [6], ferromagnetism [11] high-κ dielectric

constant [12], metal-insulator transitions [13]. We will focus TMs and TMOs deposited on

SrTiO3 in this thesis and study the physical phenomena in these material systems.

Page 21: Copyright By Wei Guo 2021

2

Figure 1.1: Transition metals (yellow) in the periodic table frame with lanthanides and

actinides shown separately (light yellow)

Figure 1.2: Atomic crystal structure of ABO3 perovskite

oxide. This is a unit cell schematic with green A-site

atoms and blue B-site atoms. All small red spheres are

oxygen atoms.

TMOs exhibit a broad range of structural, electrical,

magnetic or optical properties when fabricated as thin films

instead of bulk crystals [10], [14], [15]. The reduction of the

dimensionality from 3D to 2D provides an extra degree of

freedom for tuning the physical properties and gives even more possibilities when a 2D

film is integrated on different substrates. Symmetry breaking and interfaces of thin film

heterostructures result in phenomena like charge transfer and are the origin of novel

physical phenomena not observed in bulk materials [16], [17]. For example, a

Page 22: Copyright By Wei Guo 2021

3

heterojunction of two insulating thin films LaAlO3/SrTiO3 shows the unexpected

interfacial superconductivity [18]-[20]. Interfacial magnetism was discovered in the same

materials system even though both oxides are non-magnetic in bulk [19]-[22]. A very

strong polarization enhancement can be obtained when a non-ferroelectric material is

fabricated as a part of a superlattice with a traditional ferroelectric material [23]. The

immense potential of the integration of TMO thin films with semiconductors leads to many

more exciting possibilities both in fundamental physics and in novel applications.

1.2 SRTIO3 ON SI

1.2.1 Si

Si is arguably the most important semiconductor material today. It is widely used

as a platform for the modern electronics industry based on the complementary metal–

oxide–semiconductor (CMOS) technology [24] and the metal–oxide–semiconductor field-

effect transistor (MOSFET), which is probably THE most important and widely

manufactured device in the information age. Si is also an excellent integration platform for

photonic applications due the ability to fabricate Si waveguides with very low losses and

high optical mode confinement [25]-[27]. The seemingly unlimited supply of Si on earth

is another reason that makes Si so important. The fundamental reason of the irreplaceability

of Si is in its structural and electronic properties.

Si has an indirect band gap of 1.12 eV [28] and can be doped to be a conductor with

dopants like As or P for n-type doping, and B for p-type doping [29]. It has a diamond

crystal structure with covalent bonds, which results in a very high melting point of 1414

°C. The oxidation of Si is another key factor for its widespread industrial application.

The native oxide SiO2 has a tightly bonded, nearly defect-free interface with Si. Great

Page 23: Copyright By Wei Guo 2021

4

interface quality obtained through modern day growth methods affords great transistor

performance in state-of-art industrial production [29]. The wide bandgap of SiO2 (~9 eV

[30]) limits the gate leakage current to an extremely low level when used as a gate

dielectric. The great thermal and mechanical stability [31], high dielectric strength [32],

and convenience of growth, all made SiO2, until very recently, a passivation layer of choice

in the fabrication and manufacturing process of electronic devices.

Due to rapid developments in the semiconductor industry and the miniaturization of

modern semiconductor devices that came along with it, Moore’s Law is no longer followed as

devices approached the fundamental physical limits of traditional materials. For decades,

Si/SiO2 used to be the only option in the semiconductor industry. However, it faced challenges

when the dimension requirements for gate dielectrics scaled down to just several Å in thickness

in recent years [33]. High leakage current at this thickness caused increasing power usage and

device failures, which made people seek new approaches at this length scale [33]. The use of

high-κ gate dielectrics is one of the implemented solutions. The new gate stack can still have a

relatively large dielectric thickness (d) while maintaining a higher capacitance as the

dependency is 𝐶 ∝ 𝜅

𝑑. Intel Corporation launched their 45 nm technology node in 2007 using

hafnium dioxide (HfO2) as gate dielectric [34], [35]. As of today, people are still seeking

alternative methods both material-wise (MoS2 [36], graphene [37]) and architecture-wise, with

various types of FETs (FinFET [38], GAAFET [39]) being studied and developed.

1.2.2 SrTiO3

SrTiO3 (STO) is a widely used cubic perovskite oxide material known for its

excellent electronic properties. STO exhibits water photolysis [40], superconductivity [41],

room temperature ferroelectricity (when strained) [42], and many other fascinating

Page 24: Copyright By Wei Guo 2021

5

properties. It has a lattice constant of 3.905 Å which is lattice-match to many other complex

oxides, which makes it favorable to serve as a substrate or in a superlattice. The interface

of STO with other complex oxides also exhibits a large array of fascinating phenomena

like ferromagnetism and superconductivity in LaAlO3/SrTiO3 interfacial two-dimensional

electron gas (2DEG) [18]-[22], large positive linear magnetoresistance at the EuO/SrTiO3

interface [43], magnetic ordering in LaMnO3/SrTiO3 superlattices [44], etc. These STO-

based effects lead to the further exploration of the physics behind such effects and spurred

the search for possible industrial device implementations. As Si is still the main

“workhorse” in the modern semiconductor industry, integration of STO onto Si would open

new possibilities to utilize these electronic properties and create more efficient devices with

new functionality. Fortunately, such a process of integration does exist [45]. In this thesis,

I will mainly focus on the integration of TMs and TMOs with STO and with STO on Si.

1.2.3 SrTiO3 growth on Si

The discovery by Mckee et al. in 1998 of a way to epitaxially nucleate STO on Si

directly by molecular beam epitaxy (MBE) has opened the possibility of integrating other

perovskite materials on Si [45]. A typical process involves first removing the surface SiO2

and then depositing 0.5 monolayer (ML) of Sr metal in an MBE chamber [46]-[50]. This

Sr sub-monolayer prevents Si oxidation during the initial nucleation of STO [51]. A thin

(1-4 nm), barely crystalline STO layer is then deposited at low temperature (<300°C) with

the correct stoichiometric ratio of Sr and Ti in the presence of modest oxygen pressure (10-

8 ~ 10-7 Torr) [50]. This layer is then fully crystallized in vacuum at 500-550°C

(crystallization temperature could be higher if STO is slightly off stoichiometry), which

results in SiO2-free epitaxial STO on Si [52]. We show a top view of the atomic

Page 25: Copyright By Wei Guo 2021

6

arrangement of STO on the Si (100) surface and a side view of the interface in Figure 1.3.

From the top view, the STO unit cell is rotated by 45º with respect to the conventional Si

unit cell due to the lattice parameter of Si being ~√2 times that of STO. Other perovskites,

like BaTiO3 or LaAlO3, which are epitaxially grown on the initial STO layer are expected

to be crystallographically aligned with the STO template. From the side view, the half ML

of Sr is located in the gaps of the dangling bonds of the clean Si surface and serves as a

surface that allows STO to epitaxially crystallize on top and form the perovskite structure.

Figure 1.3. (a)Top view of the atomic arrangement of STO on the Si (100) surface, with

the STO unit cell rotated 45º with respect to the Si unit cell (b) Side view of

the STO/Si interface. The dimers of Si are bonded to the half ML of Sr and

to oxygen, allowing STO to be epitaxially grown.

Page 26: Copyright By Wei Guo 2021

7

Once the initial template of 4-5 ML of STO is crystallized, additional STO can be

grown with two possible options. If the additional STO is again grown at low temperature

(<300°C) with modest oxygen supply (mid 10-7 Torr) followed by annealing in vacuum to

fully crystallize it, one would get a clean SiO2-free interface between STO and Si. If instead

one grows additional STO, treating the template as if it were a single crystal STO substrate,

additional SiO2 will be formed since oxygen has enough diffusivity in STO at the normal

STO growth temperature of 500-550°C to reach Si. The resulting SiO2 thickness in this

case will increase as the total deposition time is prolonged. For STO layers less than 20 nm

thick, this typically produces about 1-2 nm of SiO2 interlayer between STO and Si [52].

Depending on whether or not the presence of this SiO2 interlayer is critical for an

application, the appropriate process for growing additional STO should be used. Note, that

despite the presence of this amorphous SiO2 interlayer, the resulting STO layer is still

single crystal and in epitaxial registry to Si. The growth of the STO template layer is very

crucial to those epitaxially-grown materials on top. The template crystal quality and strain

highly affect the quality and strain of following epitaxial film, which can introduce some

unexpected features or unwanted defects. Interfacial SiO2, due to high temperature growth,

is also very useful as a film strain engineering tool for STO and other materials deposited

on STO [52]-[54]. Zhang et al. have reported a thermal engineering process for tuning

strain during STO/Si growth [54]. By using the large mismatch of the thermal expansion

coefficients of STO and Si, they were able to produce a continuously tunable STO strain

by varying the growth temperature. Their main idea is to utilize the misfit dislocations

formed during growth and the amorphous SiO2 interlayer between STO and Si to control

residual stress. In-plane tensile stress is built up during the cooling process from the growth

temperature to room temperature. This thermal engineering method provides a practical

Page 27: Copyright By Wei Guo 2021

8

way to manipulate the STO template and the epitaxial film grown on top since the strain

level of the template layer directly affects the film quality, relaxation process, and lattice

parameters of the epitaxial film.

1.2.4 SrTiO3/Si and SrTiO3 free-standing membranes

With the STO/Si deposition technique and so many STO-based complex material

systems and fascinating phenomena, STO can serve as an excellent bridge material to

connect Si and other TMs or TMOs, enabling the integration of these materials on Si to

open more possibilities of novel devices. We will list several well-studied materials

systems based on STO/Si below.

High mobility interfacial 2DEG in LaAlO3/SrTiO3 was first reported by Ohtomo

and Hwang in 2004 [55]. This discovery attracted huge interest and new interesting

properties of this interface have also been studied, as mentioned before [18]-[22]. STO and

LAO in bulk are both insulating materials and their band gaps are 3.3 eV and 5.6 eV,

respectively [56], [57]. While a lot of papers discuss the LAO/STO system as the next

generation electronic device [58]-[60], Ortmann et al. published a series of papers on the

implementation of the LAO/STO quantum well (QW) structure for novel electro-optic

(EO) devices [61]-[64]. They overcame many fundamental limitations for the integration

of thick STO/LAO QW heterostructures on Si, such as the significant structural distortions

at STO/LAO interfaces, which previously restricted thick heterostructure growth [65].

They managed to make 10 periods of 7 uc LAO/4 uc 1% La:STO QW on Si with STO

buffer layer via molecular beam epitaxy (MBE) and confirmed great crystal quality by

reflection high-energy electron diffraction (RHEED), X-ray diffraction (XRD), and

scanning transmission electron microscopy (STEM) measurements [64]. They also

Page 28: Copyright By Wei Guo 2021

9

presented calculations based on the LAO/STO QW on Si for novel electro-optical devices

operating at near-infrared optical wavelength [63]. This work has opened a new avenue for

the development of a wide range of novel optical, electro-optical devices, sensors, light

sources and photonics.

Another significant materials system based on STO/Si is ferroelectric perovskites on

STO. The Pockels effect is an electric field-induced change in the refractive index of a crystal.

It was first demonstrated by Friedrich Pockels in 1893 [66]. BaTiO3 (BTO) is a highly

promising ferroelectric material for the fabrication of electro-optic (EO) modulators because it

exhibits one of the largest Pockels coefficients among the EO materials [67]-[71]. It also has

the added benefit of being readily integrated on a Si materials platform with an STO buffer

layer. These two characteristics make epitaxial BTO ideal for use in next generation silicon

photonics application with fast, low-power optical switches, or even for new forms of

computing including neuromorphic and quantum computing [68], [71]. At room temperature,

bulk BTO has a tetragonal crystal structure with space group P4mm and lattice constants a =

3.994 Å and c = 4.0335 Å. It has a remnant ferroelectric polarization of 26 μC/cm2 and a Curie

temperature of 130ºC, above which it becomes a cubic paraelectric material with a lattice

constant of 4.006 Å and space group Pm3m [70]. Being a ferroelectric, BTO has a unique

crystallographic direction in which the ferroelectric polarization points. For EO modulators,

because the polarization direction controls the coupling between light and an external electric

field, it is important to understand how different growth methods and subsequent processing

affect the direction of the ferroelectric polarization. Certain electro-optic devices may require

polarization to be in the plane of the film (in-plane switching liquid crystal devices), while

other applications may require it to be normal to the plane of the film (Mach-Zehnder

modulator).

Page 29: Copyright By Wei Guo 2021

10

Orientation of BTO is the key factor of the EO response of the film. At room

temperature, there is a ~4% lattice mismatch between BTO and Si (calculated as (aSi-

aBTO)/aBTO). We call a BTO film c-axis-oriented if its c-axis is pointing normal to the substrate

surface plane. On the other hand, we call a BTO film a-axis-oriented if its c-axis lies in the

plane of the film. There is also a large difference between the coefficients of thermal expansion

of BTO (9.0×10-6 K-1) and Si (2.6×10-6 K-1), which leads to an additional in-plane biaxial

tensile stress in the BTO layer during cooling [70]. This additional stress can cause orientation

changes in the BTO polarization. In this section, we summarize several of the most relevant

publications about epitaxial BTO on Si and compare the BTO structural parameters they

obtained with each other.

In Table 1, I have summarized the growth conditions and the corresponding BTO

lattice parameters obtained from the different growth methods [72]. The techniques used to

grow BTO have several things in common: thicker BTO on STO/Si tends to produce more a-

axis orientation; post-deposition annealing of at least 600 ºC in air or oxygen can transform a

c-axis BTO film to an a-axis one; higher oxygen pressure enhances a-axis orientation. In

addition, the quality and the thickness of the STO template layer, and the presence and amount

of an SiO2 interlayer, can significantly affect the manner and degree of BTO film relaxation.

Method Tempera

ture (ºC)

Oxygen

pressure (Torr)

STO thick-

ness (nm)

BTO thick-

ness (nm)

a (nm) c (nm) Orient

ation

MBE (2006) 550-570 4-5×10-8 30 (BST) 10 3.9996±0

.0005

4.025±

0.003

c

MBE (2007) 600 2 30 4.01 4.05 c

MBE (2011) 620 1×10-6 5 40 3.978 4.057 c

MBE (2013) 700 5×10-6 3.9/6.2 8 3.993 3.038 c

MBE (2013) 700 5×10-6 3.9/6.2 40 3.993 4.015 a/c

MBE (2013) 600 1×10-5mbar 4 8 4.038 c

MBE (2014) 440-525 1-5×10-7 4 7 3.996 4.027 c

MBE (2013) 500 1×10-7mbar 2 8 4.032 c

Page 30: Copyright By Wei Guo 2021

11

MBE (2013) 600 1×10-5mbar 4 130 3.997 4.032 a

MBE (2014) 8 80 3.998 4.03 a

MBE (2015) 4-6 90 4.03 c

4-6 90 3.99(850

ºC RTA)

a

MBE (2017) 630 1.6×10-6 5 100 4.00 4.06 c

40 100 4.00 4.02 a

MBE (2019) 650 5×10-7 15 40 3.996 4.046 c

ALD (2014) 1.6 7-20 3.93 4.02 c

ALD (2019) 2 66 4.007(65

0ºC

RTA)

a

ALD (2020) 3.6 10 3.98 a

Rf sputtering

(2012)

700 1.5-3×10-4 20 120 4.024 c

Rf sputtering

(2013)

500-600 10-5 bar 4 100 3.966 4.062 a

Rf sputtering

(2020)

650 2 Pa 10 60 4.003 4.042 a

PLD (2018) 450 0.02 mbar LNO/CeO2/

YSZ

110 4.04 4.12 c

PLD (2018) 700 0.1 mbar LNO/CeO2/

YSZ

110 3.99 4.02 a

PLD (2020) 650 0.01 mbar 10 100 4.09 c

Sol-gel (2020) 2 20 3.980 4.036 a

Table 1. Summary of the growth conditions and the corresponding lattice parameters

from different growth methods reported in the literature (with published

years). Blanks indicate the corresponding paper did not include detailed

data. (Pressure unit is Torr if not marked differently).

A typical waveguide phase modulator based on BTO/Si system is shown in Figure

1.4. The waveguide is located between the metal contacts, and the electric field is

horizontal, or parallel to the surface. The confinement is achieved by placing a Si

waveguide atop the perovskite (indicated with an orange bar in Figure1.4). As a matter of

fact, the majority of the mode is in Si, and only about 20% is in BTO. The electro-optically

induced index change results in a phase shift that depends on the applied voltage, length of

Page 31: Copyright By Wei Guo 2021

12

the modulator, wavelength of light, spacing between the electrodes G, and the overlap

integral between the applied electric field and the optical mode. Details of calculations can

be found elsewhere [72]. I published a review paper [72] that reviewed the growth of

epitaxial BTO on Si by a variety of deposition methods including MBE, pulsed laser

deposition (PLD), and RF sputtering. I summarize the resulting BTO film structure and

quality based on the reported characterization results and discussed EO measurements of

basic devices made from this materials platform, where such data is available.

Figure 1.4. Schematic of a waveguide phase modulator.

1.3 CHAPTER OVERVIEW

In this thesis, my research mainly focuses on TM and TMO integration on STO and

STO/Si. Chapter 2 contains an introduction to the experimental equipment used in my

research at the Advanced Atomic Design Lab in the University of Texas at Austin. Our

major deposition method is MBE equipped with RHEED, and an in situ characterization

tool, which is x-ray photoelectron spectroscopy (XPS).

Page 32: Copyright By Wei Guo 2021

13

Chapter 3 explores the temperature dependence of the morphology and electronic

structure of ultrathin coverage of platinum (Pt) on STO. High and low growth temperatures

applied to Pt results in total different surface morphology and electronic structure

evolution. Some interesting annealing results are also shown.

Chapter 4 describes a series of transition metal depositions on STO [74]. We find

that Eu deposited under UHV condition results an unstable monoxide to form. Previous

work on other metals has been done by former group members [43]. We explore the oxygen

scavenging behavior and manage to epitaxially grow EuO on STO. We vary the STO

thickness and growth temperature to study this interface. Further research is to insert a

BTO film and use angle-resolved photoemission spectroscopy (ARPES) at the Paul Scherer

Institute (PSI), Switzerland, to explore the effect of the ferroelectric BTO layer on the

interfacial 2DEG.

Chapter 5 focuses on the STO/Si pseudo-substrate. We first study the dry oxidation

of buried Si as a function of annealing time and determine the maximum temperature for

which a relatively thin STO layer (10 nm) remains intact. We developed a theoretical model

based on a modification of the Deal–Grove–Massoud formalism that predicts the thickness

of SiO2 formed underneath STO as a function of time and temperature and report a robust

recipe for dry oxidation of Si buried under an epitaxial layer of STO. The following

research is to fabricate a thick SiO2 interlayer on STO/Si template and use selective etch to

remove SiO2 and produce free-standing STO membranes. Free-standing 2D STO may

show different properties than either thin film or bulk STO. In a preliminarily study we

used Raman spectroscopy to characterize STO membranes.

Chapter 6 is our prototype design of BaSnO3/SrTiO3/Al2O3 quantum wells. This

design is based on the large conduction band offset (~3.5 eV) between BSO and Al2O3.

Page 33: Copyright By Wei Guo 2021

14

The deposition process is done by MBE and characterized by RHEED and XPS to confirm

the film quality.

1.4 REFERENCE

1. B. N. Figgis and J.Lewis, The Magnetochemistry of Complex Compounds in Modern

Co-ordination Chemistry, Ed. J. Lewis and C. Wilkins, Interscience, New York (1960).

2. T. Sakudo, Ferroelectrics, 95, 37 (1989).

3. K. Takada, H. Sakurai, E. Takayama-Muromachi, F. Izumi, R. A. Dilanian, and T.

Sasaki, Nature, 422, 53 (2003).

4. Y. Kamihara, H. Hiramatsu, M. Hirano, R. Kawamura, H. Yanagi, T. Kamiya, and H.

Hosono, J. Am. Chem. Soc. 128, 10012 (2006).

5. K. A. Kovacs, P. Grof, L. Burai, M. Riedel. J. Phys. Chem. A. 108, 11026 (2004).

6. H. N. Lee, H. M. Christen, M. F. Chisholm, C. M. Rouleau, and D. H. Lowndes. Nature,

433, 395 (2005).

7. S. W. Cheong and M. Mostovoy. Nat. Mater. 6, 13 (2007).

8. M. Mostovoy, Nat. Mater. 7, 269 (2008).

9. S. Maekawa, T. Tohyama, S. E. Barnes, S. Ishihara, W. Koshibae, and G. Khaliullin,

Physics of Transition Metal Oxides, (Springer, 2004).

10. A. A. Demkov and A. B. Posadas, Integration of Functional Oxides with

Semiconductors (Springer New York, 2014).

11. N. C. Bristowe, J. Varignon, D. Fontaine, E. Bousquet and Ph. Ghosez, Nat. Comm. 6,

6677 (2015)

12. J. A. Kittl, K. Opsomer, M. Popovici, N. Menou, B. Kaczer, X. P. Wang, C. Adelmann,

M. A. Pawlak, K. Tomida, A. Rothschild, B. Govoreanu, R. Degraeve, M. Schaekers,

M. Zahid, A. Delabie, J. Meersschaut, W. Polspoel, S. Clima, G. Pourtois, W. Knaepen,

C. Detavernier, V. V. Afanas’ev, T. Blomberg, D. Pierreux, J. Swerts, P. Fischer, J. W.

Maes, D. Manger, W. Vandervorst, T. Conard, A. Franquet, P. Favia, H. Bender, B.

Brijs, S. Van Elshocht, M. Jurczak, J. Van Houdt, D. J. Wouters, Microelectron. Engin.

86, 1789 (2009).

13. M. Imada, A. Fujimori, Y. Tokura, Rev. Mod. Phys. 70, 1039 (1998).

14. H. Y. Hwang, Y. Iwasa, M. Kawasaki, B. Keimer, N. Nagaosa, and Y. Tokura, Nat.

Mater. 11, 103 (2012).

15. J. Chakhalian, J. W. Freeland, A. J. Millis, C. Panagopoulos, and J. M. Rondinelli, Rev.

Mod. Phys. 86, 1189 (2014).

Page 34: Copyright By Wei Guo 2021

15

16. N. Sai, B. Meyer, and D. Vanderbilt, Phys. Rev. Lett. 84, 5636 (2000).

17. H. Yamada, M. Kawasaki, Y. Ogawa, and Y. Tokura, Appl. Phys. Lett. 81, 4793 (2002).

18. N. Reyren, S. Thiel, A. D. Caviglia, L. F. Kourkoutis, G. Hammerl, C. Richter, C. W.

Schneider, T. Kopp, A.-S. Rüetschi, D. Jaccard, M. Gabay, D. A. Muller, J.-M.

Triscone, and J. Mannhart, Science. 317, 5842 (2007).

19. L. Li, C. Richter, J. Mannhart, and R. C. Ashoori, Nat. Phys. 7, 762 (2011).

20. J. A. Bert, B. Kalisky, C. Bell, M. Kim, Y. Hikita, H. Y. Hwang, and K. A. Moler, Nat.

Phys. 7. 767(2011).

21. A. Brinkman, M. Huijben, M. van Zalk, J. Huijben, U. Zeitler, J. C. Maan, W. G. van

der Wiel, G. Rijnders, D. H. A. Blank, and H. Hilgenkamp, Nat. Mater. 6, 493 (2007).

22. B. Kalisky, J. A. Bert, B. B. Klopfer, C. Bell, H. K. Sato, M. Hosoda, Y. Hikita, H. Y.

Hwang, and K. A. Moler. Nat. Commun. 3, 922 (2012).

23. H. N. Lee, H. M. Christen, M. F. Chisholm, C. M. Rouleau, and D. H. Lowndes, Nature,

433, 395 (2005).

24. D. J. Frank, R. H. Dennard, E. Nowak, P. M. Solomon, Y. Taur and Hon-Sum Philip

Wong, Proceedings of the IEEE. 89, 259 (2001).

25. S. Abel, J. Fompeyrine, “Electro-Optically Active Oxides on Silicon for Photonics” in

Thin Films on Silicon (V. Narayanan, M. Frank and A. A. Demkov. eds.) 455 (2016).

26. M. Hochberg and T. Baehr-Jones, Nat. Photon. 4, 492(2010).

27. G. T. Reed, G. Mashanovich, F. Y. Gardes and D. J. Thomson, Nat. Photon. 4, 518

(2010).

28. B. G. Streetman, S. Banerjee, Solid State electronic Devices, 524 (Hall, New Jersey,

2000).

29. J. D. Plummer, M. D. Deal, P. B. Griffin, Silicon VLSI Technology, (Prentice Hall,

Upper Saddle River, 2013).

30. E. Vella, F. Messina, M. Cannas, R. Boscaino, Phys. Rev. B. 83, 174201 (2011).

31. B. El-Kareh, Fundamentals of Semiconductor Processing Technologies, (Springer US,

New York, 1995).

32. E. Harari, J. Appl. Phys. 49, 2478-2489 (1978).

33. H. Iwai, Microelectron. Eng. 86, 1520 (2009).

34. K. Mistry, C. Allen, C. Auth, B. Beattie, D. Bergstrom, M. Bost, M. Brazier, M.

Buehler, A. Cappellani, R. Chau, C.-H. Choi, G. Ding, K. Fischer, T. Ghani, R. Grover,

W. Han, D. Hanken, M. Hattendorf, J. He, J. Hicks, R. Heussner, D. Ingerly, P. Jain,

R. James, L. Jong, S. Joshi, C. Kenyon, K. Kuhn, K. Lee, H. Liu, J. Maiz, B. McIntyre,

Page 35: Copyright By Wei Guo 2021

16

P. Moon, J. Neirynck, S. Pae, C. Parker, D. Parsons, C. Prasad, L. Pipes, M. Prince, P.

Ranade, T. Reynolds, J. Sandford, L. Shifren, J. Sebastian, J. Seiple, D. Simon, S.

Sivakumar, P. Smith, C. Thomas, T. Troeger, P. Vandervoorn, S. Williams, K.

Zawadzki, Int. El. Devices Meet. 247 (2007).

35. M. T. Bohr, R. S. Chau, T. Ghani, K. Mistry, IEEE Spectrum. 44, 29 (2007).

36. B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti and A. Kis. Nature Nanotech

6, 147(2011).

37. L. Liao, Y. Lin, M. Bao, R. Cheng, J. Bai, Y. Liu, Y. Qu, K. L. Wang, Y. Huang and

X. Duan. Nature 467, 305 (2010).

38. D. Hisamoto, W. Lee, J. Kedzierski, H. Takeuchi, K. Asano, C. Kuo, E. Anderson, T.

King, J. Bokor and C. Hu. IEEE Transactions on Electron Devices. 47 2320 (2000).

39. N. Singh, A. Agarwal, L. K. Bera, T. Y. Liow, R. Yang, S. C. Rustagi, C. H. Tung,

R. Kumar, G. Q. Lo, N. Balasubramanian, and D. L. Kwong. IEEE Electron Device

Letters, 27, 383 (2006)

40. R. Konta, T. Ishii, H. Kato, and A. Kudo, J. Phys. Chem. B. 108, 8992 (2004).

41. K. Ueno, S. Nakamura, H. Shimotani, A. Ohtomo, N. Kimura, T. Nojima, H. Aoki, Y.

Iwasa, and M. Kawasaki, Nat. Mater. 7, 855 (2008).

42. J. H. Haeni, P. Irvin, W. Chang, R. Uecker, P. Reiche, Y. L. Li, S. Choudhury, W. Tian,

M. E. Hawley, B. Craigo, A. K. Tagantsev, X. Q. Pan, S. K. Streiffer, L. Q. Chen, S.

W. Kirchoefer, J. Levy and D. G. Schlom. Nature 430, 758 (2004).

43. K. J. Kormondy, L. Gao, X. Li, S. Lu, A. B. Posadas, S. Shen, M. Tsoi, M. R.

McCartney, D. J. Smith, J. Zhou, L. L. Lev, M.-A. Husanu, V. N. Strocov, and A. A.

Demkov, Sci. Rep. 8, 7721 (2018).

44. W. S. Choi, D. W. Jeong, S. S. A. Seo, Y. S. Lee, T. H. Kim, S. Y. Jang, H. N. Lee,

and K. Myung-Whun, Phys. Rev. B. 83, 195113 (2011).

45. R. A. McKee, F. J. Walker, and M. F. Chisholm, Phys. Rev. Lett. 81, 3014(1998).

46. J. Ramdani, R. Droopad, Z. Yu, J. A. Curless, C. D. Overgaard, J. Finder,K.

Eisenbeiser, J. A. Hallmark, W. J. Ooms, V. Kaushik, P. Alluri, and S. Pietambaram,

Appl. Surf. Sci. 159, 127 (2000).

47. R. A. McKee, F. J. Walker, and M. F. Chisholm, Science 293, 468 (2001).

48. X. Hu H. Li, Y. Liang, Y. Wei, Z. Yu, D. Marshall, J. Edwards, R.Droopad, X. Zhang,

A. A. Demkov, K. Moore, and J. Kulik, Appl. Phys.Lett. 82, 203 (2003).

49. X. Zhang, A. A. Demkov, H. Li, X. Hu, Y. Wei, and J. Kulik, Phys. Rev.B. 68, 125323

(2003).

Page 36: Copyright By Wei Guo 2021

17

50. Y. Wei, X. Hu, Y. Liang, D. Jordan, B. Craigo, R. Droopad, Z. Yu, A. A. Demkov, J.

Edwards, Jr., and W. Ooms, J. Vac. Sci. Technol. B 20, 1402(2002).

51. K. D. Fredrickson, H. Seo and A. A. Demkov, J. Appl. Phys. 120, 065301 (2016).

52. M. Choi, A. Posadas, R. Dargis, C.-K. Shih, A.A. Demkov, D.H. Triyoso, N.D.

Theodore, C. Dubourdieu, J. Bruley, and J. Jordan-Sweet, J. Appl. Phys. 111, 064112

(2012).

53. K. J. Kormondy, S. Abel, F. Fallegger, Y. Popoff, P. Ponath, A. B. Posadas, M.

Sousa, D. Caimi, H. Siegwart, E. Uccelli, L. Czornomaz, C.Marchiori, J. Fompeyrine,

A. A. Demkov. Microelectron. Eng. 147, 215 (2015).

54. L. Zhang, Y. Yuan, J. Lapano, M. Brahlek, S. Lei, B. Kabius, V. Gopalan, and R.

Engel-Herbert, ACS Nano, 12, 1306(2018)

55. A. Ohtomo and H. Y. Hwang, “A high-mobility electron gas at the LaAlO3/SrTiO3

heterointerface,” Nature, 427, 423 (2004).

56. J. Robertson, J. Vac. Sci. Technol. B. 18. 1785(2000).

57. S.G. Lim, S. Kriventsov, T. N. Jackson, J. H. Haeni, D. G. Schlom, A. M. Balbashov,

R. Uecker, P. Reiche, J. L. Freeouf, and G. Lucovsky, J. Appl. Phys. 91, 4500 (2002).

58. R. Ramesh and D. G. Schlom, MRS Bull. 33, 1006(2008).

59. D. F. Bogorin, P. Irvin, C. Cen, and J. Levy, “LaAlO3/SrTiO3-based device concepts,”

in Multifunctional Oxide Heterostructures, Oxford University Press, 364(2010).

60. L. Bjaalie, B. Himmetoglu, L. Weston, A. Janotti, and C. G. Van de Walle. New J.

Phys. 16, 025005 (2014).

61. J. E. Ortmann, N. Nookala, Q. He, L. Gao, C. Lin, A. B. Posadas, A. Y. Borisevich, M.

A. Belkin, and A. A. Demkov, ACS NANO 12, 7682 (2018).

62. J. E. Ortmann, A. B. Posadas and A. A. Demkov, J. Appl. Phys. 124, 015301 (2018).

63. J. E. Ortmann, M. A. Duncan and A. A. Demkov, Optical Materials Express 9, 2982

(2019).

64. J. E. Ortmann, S. Kwon, A. B. Posadas, M. J. Kim and A. A. Demkov, J. Appl. Phys.

125, 155302 (2019).

65. N. Nakagawa, H. Y. Hwang, and D. A. Muller. Nat. Mater. 5, 204 (2006).

66. F. Pockels, Lehrbuch Der Kristalloptik (Leipzig und Berlin, 1906).

67. B. W. Wessels, Annu. Rev. Mater. Res. 37, 659 (2007).

68. S. Abel, T. Stöferle, C. Marchiori, C. Rossel, M.D. Rossell, R. Erni, D. Caimi, M.

Sousa, A. Chelnokov, B.J. Offrein, J. Fompeyrine, Nat. Commun. 4, 1671(2013).

69. D. P. Kumah, J. H. Ngai and L. Kornblum, Adv. Funct. Mater. 30, 1901597(2019)

Page 37: Copyright By Wei Guo 2021

18

70. L. Mazet, S. M. Yang, S. V. Kalinin, S. Schamm-Chardon and C. Dubourdieu, Sci.

Technol. Adv. Mater. 16, 036005 (2015).

71. S.Abel, F. Eltes, J. E. Ortmann, A. Messner, P. Castera, T. Wagner, D. Urbonas, A.

Rosa, A. M. Gutierrez, D. Tulli, P. Ma, B. Baeuerle, A. Josten, W. Heni, D. Caimi, L.

Czornomaz, A. A. Demkov, J. Leuthold, P. Sanchis. J. Fompeyrine. Nature Material,

18, 42 (2019).

72. W. Guo, A. B. Posadas, A. A. Demkov, J. Vac. Sci. Technol. A 39, 030804 (2021)

73. R. C. Alferness, IEEE Transactions on Microwave Theory and Techniques, 82,

1121(1982).

74. A. B. Posadas, K. J. Kormondy, W. Guo, P. Ponath, J. Geler-Kremer, T. Hadamek,

and A. A. Demkov, J. Appl. Phys. 121, 105302 (2017)

Page 38: Copyright By Wei Guo 2021

19

Chapter 2: Experiment techniques

I will now introduce the principles and instrument details of the equipment I have

used in my research. The Advanced Atomic Design Lab has an MBE reactor equipped with

RHEED and an in situ XPS. A schematic of our system is shown in Figure 2.1. I will

explain these in detail in this chapter.

Figure 2.1: Schematic of the vacuum system at the Advanced Atomic Design Lab. All

instruments are linked to the main transfer line.

Page 39: Copyright By Wei Guo 2021

20

2.1 MOLECULAR BEAM EPITAXY (MBE)

Molecular beam epitaxy (MBE) is a physical vapor deposition process that deposits

high quality thin films epitaxially under ultra-high vacuum (UHV). This is a widely used

method both in the semiconductor industry and in research labs. Günther et al. invented

the principal idea of MBE in 1958 [1]. It attracted much attention and enjoyed fast

development since its invention due to the high demand for high quality III-V

semiconductor devices [2]. Cho and Arthur et al. lead the work at Bell Laboratories and

developed MBE in the late 1960s [3].

2.1.1 Operating principles and instrument illustration

The basic principle of MBE is to evaporate high-purity materials under UHV

environment and generate molecular beams going straight to the substrate wafers. Under

appropriate conditions (temperature, flux, crystalline orientation, etc.), gaseous species

arriving on the wafer will condense and react with each other to form an epitaxial film. The

major advantage of MBE over other deposition tools is the precise control of the deposition

process with relatively low flux rate, benefiting from the UHV environment. MBE can

achieve sub-monolayer deposition control, which allows one to have more control over

film properties and allow for tuning the band gap, lattice constant or crystalline orientation.

Our lab is equipped with a custom-built DCA M600 Oxide-MBE. The major

components are listed below and I label them on a picture of the chamber in Figure 2.1.

Page 40: Copyright By Wei Guo 2021

21

Figure 2.2: Front view of MBE with components labeled.

Page 41: Copyright By Wei Guo 2021

22

Figure 2.3: Back view of MBE with components labeled

Page 42: Copyright By Wei Guo 2021

23

(1) Cryopump, pressure gauges and residual gas composition analyzer

(2) Cryopanel and water cooling system

(3) Manipulator with thermal control and sample transfer

(4) Effusion cells and e-beam evaporator.

(5) Quartz crystal monitor (QCM) for flux measurement.

(6) Oxygen source with plasma generator.

(7) RHEED system for monitoring the surface crystallinity.

2.2.2 Preparation for MBE growth

Samples need pre-treatment to be suitable for a vacuum process. The preparation

usually consists of: 1) ultrasonic degrease in acetone, isopropanol and deionized (18

MΩ/cm) water for 5-10 minutes each to remove surface grease and dust as much as

possible, 2) ozone treatment for 15-30 min to remove surface residual carbon.

These steps are very basic procedures for sample preparation. Different substrates

require different additional treatments. For example, the TiO2-terminated STO surface [4]

requires boiling water soaking of the sample and up to 4 hours oxygen anneal to form

atomically flat TiO2-terminated surface; A clean Ge surface requires ex situ wet etching

using HCl as the etchant and H2O2 as the oxidant [5].

After the sample is well-cleaned and treated, it is loaded into the load lock of the

vacuum system and pumped down to the same vacuum level as the system. It is then

transferred to other chambers for the subsequent experiments.

Page 43: Copyright By Wei Guo 2021

24

2.2.3 Maintaining vacuum and cryopump

The vacuum level in the MBE main chamber is within the ultra-high vacuum

(UHV) regime, meaning pressures ~10-9 Torr or lower. Our MBE has a base pressure of

3×10-10 Torr when all cells are cold and the chamber is well-baked out.

Maintaining vacuum is crucial for the operation of an MBE as it determines the

level of atom flux control one can achieve. The mean free path of a gaseous species is the

distance it travels on average without collisions with other atoms or molecules of the gas.

Only a sufficiently high vacuum level will guarantee that the molecular flux will not be

interrupted or scattered while it travels from the effusion cells to the wafer. This usually

requires the mean free path to be higher than 0.5-1 m for MBE. Under normal atmospheric

pressure (760 Torr), the typical mean free path is ~100 nm. In a UHV environment with

10-10 Torr, one obtains a ~100 km mean free path. At first glance, it seems wasteful to

maintain UHV instead of a lower level of vacuum, but the growth process has a lot of

heating, gas injection and other steps, all of which will decrease the vacuum level of the

chamber. A typical oxide MBE operation vacuum level with hot effusion cells and oxygen

flow provided is in the high vacuum regime (8×10-4 ~8×10-8 Torr). This vacuum level

yields mean free paths in the tens of cm to 1 km range, so we can still be sure the mean

free path longer than the distance from the cells to the substrate, as long as we maintain the

vacuum level below ~5×10-5 Torr.

Our MBE system uses a cryogenic pump to maintain UHV. A cryogenic pump

works by having a large area cold surface connected to the vacuum chamber to condense

and adsorb gaseous species. This cold surface is cooled by compressed helium refrigeration

resulting in temperatures of 10-15 K temperature enabling one to condense most gases,

effectively removing them from the chamber to maintain UHV. The walls of the MBE

Page 44: Copyright By Wei Guo 2021

25

chamber are shielded with a cryo-panel that circulates chilled ethanol. The cooling of the

stainless-steel chamber walls helps keep the pressures low during operation when sources

are hot by preventing adsorbed atoms and molecules from re-evaporating.

2.2.4 Effusion cells

Effusion cells (also called Knudsen cells) are miniature furnaces with heating coils

and inner crucibles. High purity metals are loaded in the crucible and evaporated by heating

the crucible to a sufficiently high temperature, which depends on the material (up to ~1800

ºC in the case of Ti). A water-cooling system and radiation shield maintains the exterior

temperature to avoid heating the rest of the reactor during operation. .

Our MBE chamber is equipped with 6 effusion cells, which are of 3 types. For Ti

and La, which have very high melting temperatures, we use high-temperature effusion cells

that allow one to use operating temperatures of 1400 ºC - 1800 ºC. For Ba, Sr and Eu, low-

temperature effusion cells with pyrolytic boron nitride (PBN) crucibles are enough for

operating temperatures below 1000 ºC. Metallic Al requires a special cold-lip cell to prevent

liquid Al from creeping out of the crucible.

All the cells have compressed air-controlled shutters that serve as “caps”. When

closed, shutters physically block line of sight from the effusion cell to the substrate,

effectively stopping any deposition from that cell when that element is not needed during

the growth.

2.2.5 Electron beam evaporator

For those metals with extremely high melting temperatures, which exceed the

temperature range of effusion cells, the electron beam evaporator or so-called E-gun is

used.

Page 45: Copyright By Wei Guo 2021

26

A beam of electrons from a tungsten filament is focused and accelerated (~7.75 kV)

towards the target crucibles by electrostatic and magnetic fields. The electron beam will

heat the metal in the crucible, melting and evaporating it. The MBE in our lab is equipped

with a 4-pocket E-gun. A viewport located above the pockets (see Fig. 2.3) allows one to

monitor them and to adjust the electron beam position and sweep on the metal in the pocket

for optimum coverage.

2.2.6 Quartz crystal monitor (QCM)

QCM is the tool we use to measure the atomic or molecular beam flux. It is located

right below the manipulator and substrate position, to yield flux readings as close as

possible to the flux arriving at the substrate. The QCM determines the flux by measuring

the accumulated film thickness deposited on a quartz crystal [6]. The operating principle

of the sensor is that deposited film on the sensor crystal will cause a mass change and cause

a certain shift in the resonance frequency of the quartz oscillator [7]. We can then determine

the film thickness that accumulates in a certain time interval, from which we can calculate

the flux.

Flux control is a very important factor in MBE deposition, since the stoichiometry

of the film and hence the film quality is largely determined by it.

2.2 REFLECTION HIGH-ENERGY ELECTRON DIFFRACTION (RHEED)

Reflection high-energy electron diffraction (RHEED) is a surface diffraction

technique that is commonly used in deposition tools, especially MBE, to monitor the

surface crystallinity during growth. The principle of RHEED is to use a beam of highly

energetic electrons that strike the sample at a very small angle relative to its surface. The

Page 46: Copyright By Wei Guo 2021

27

surface crystal structure of the sample diffracts the incident electrons, causing an

interference pattern to form, which can be seen on the phosphor screen.

A schematic of principle of RHEED is shown in Figure 2.4.

Figure 2.4: Schematic of the principle of RHEED. The angle of incidence of the beam is

usually less than 5º from the surface.

In Figure 2.4, the incident beam is diffracted by the crystal structure of the sample.

Due to the small incident angle and the high energy of the incident electrons, the

penetration depth of the beam is only several atomic layers and only atoms near the top

surface will contribute to the diffraction. An infinite 3D lattice in real space corresponds to

a 3D array of points in reciprocal space. Considering just the ideal surface as a two-

dimensional lattice, this translates to reciprocal space as a 2D array of reciprocal lattice

rods. When these reciprocal rods intersect the Ewald sphere corresponding to the electron

energy, it produces a diffraction pattern that is projected onto a phosphor screen. A charge-

coupled device (CCD) camera is then used to capture the image on the screen and send it to

Page 47: Copyright By Wei Guo 2021

28

the software. Our manipulator can rotate the sample so we can change the orientation of the

surface and get diffraction patterns from different crystallographic directions like [100], [111]

or [110]. A more detailed momentum and energy analysis and discussion of the working

principles can be found elsewhere [8], [9].

RHEED images taken from real samples are shown in Figure 2.5. Here I present the

RHEED patterns for a commercial STO single crystal substrate and for an MBE-grown STO

thin film on Si.

Figure 2.5: RHEED images of (a) commercial STO substrate bought from MTI Crystal

(b) 5 uc of STO grown on a Si substrate.

In Figure 2.5 (a), the pattern consists of sharp dots sitting along a circular arc, which

is the ideal diffraction pattern from an atomically flat surface with large coherence length.

The pattern of dots is determined by the intersection of the reciprocal rods and the Ewald

sphere. Atomically flat surfaces typically have narrow reciprocal lattice rods such that the

diffraction spots are smaller and shaper. In Figure 2.5 (b), the pattern has elongated into

streaks and is not as sharp as in (a). This shows a surface that is not as flat as that of a

commercially bought substrate but still relatively flat. Due to disorder and smaller

Page 48: Copyright By Wei Guo 2021

29

coherence length due to uncorrelated surface steps, the reciprocal rods are wider and have

finite length, resulting in this slightly less sharp streaky pattern.

RHEED is the most important monitoring tool in MBE deposition. We can adjust

the growth conditions (flux and growth temperature) in real-time to ensure high film quality

using RHEED. For example, the element flux does drift especially for long time

depositions and the crystal quality of the top surface can get worse with time. RHEED is

very sensitive to surface quality changes and we can make the corresponding adjustment

to the flux to maintain the right stoichiometry.

2.3 X-RAY PHOTOELECTRON SPECTROSCOPY (XPS)

X-ray photoelectron spectroscopy (XPS) is a powerful characterization tool for

identifying existing elements in samples and their chemical state, composition, electronic

structure and density of the electronic states. The principle of XPS is based on the

photoelectric effect explained by Einstein in the early 1900s for which he received a Nobel

Prize [10]. Our lab has an in situ VG Scienta R3000 XPS system connected to the common

transfer line along with the MBE. We can measure the samples after growth without

breaking the vacuum and avoid possible sample reaction with air or other contaminants. A

front view picture of our XPS system is shown in Figure 2.6 with the main components

labeled.

Page 49: Copyright By Wei Guo 2021

30

Figure 2.6: Picture of the VG Scienta R3000 XPS system. (Front view)

2.3.1 Operating principles

The photoelectric effect happens when electromagnetic radiation is absorbed by a

material. The electrons in the atoms of the material absorb the energy of the radiation and

are ejected from the material if the photon energy is high enough to overcome the binding

energy due to the atomic nucleus. The binding energy of the emitted electron is determined

by the following equation: Ebinding + ϕ = Ephoton - Ekinetic. Here ϕ is the work function, which

is the minimum energy needed to remove an electron from the material, and Ekinetic is the

measured kinetic energy of the emitted electron.

Page 50: Copyright By Wei Guo 2021

31

In laboratory XPS systems, Ephoton is usually a fixed value corresponding to the

energy emitted by the x-ray source of the system. By measuring the kinetic energy of the

emitted electrons (and correcting for ϕ), we can calculate the binding energy of emitted

electrons. The electron energy analyzer and detector allows one to get a count of how many

electron are emitted at a given energy, producing an energy spectrum. Figure 2.7 shows

the schematic of a conventional XPS instrument.

Figure 2.7: Conventional X-ray source for an XPS instrument. (Figure adopted from [11])

X-rays are generated from the x-ray source, which is then usually monochromated

and focused onto the sample. Electrostatic focusing lenses then help to collect the emitted

electrons from a range of angles then send them through a hemispherical analyzer to filter

electrons by kinetic energy and then count them.

Page 51: Copyright By Wei Guo 2021

32

2.3.2 Surface sensitivity

XPS usually requires UHV conditions, especially for surfaces sensitive to

oxidation. Recalling that the vacuum level determines the mean free path, UHV conditions

will guarantee that: 1) the emitted electrons are not scattered by gas molecules along the

path from sample to detector; 2) attenuation and distortion of the spectra by surface

contamination is avoided [11]. High vacuum conditions (<10-5 Torr) usually satisfies

condition 1, but the sample surface will also adsorb the residual gas molecules at these

pressures in minutes and the contamination cannot be readily ignored under such vacuum

conditions. The surface contamination rate under high vacuum can be as fast as 1 s/ML,

meaning it only takes 1 second to accumulate 1 ML of residual molecules, assuming unity

sticking coefficient. Under UHV conditions, the contamination rate is much slower – about

104 s/ML in UHV [11]. This is a few times longer than a typical set of XPS scans thus

ensuring the sample surface remains contamination-free during the scans.

2.3.4 XPS peak analysis

Here we explain a typical peak analysis of an XPS experiment. We will see many

results of the analysis described here in the following chapters.

Our XPS system uses an Al Kα X-ray source with a photon energy of 1486.6 eV.

When we scan through the whole binding energy range from 0 eV to 1486.6 eV, the emitted

electrons will come from all the core levels from all the elements in the material (except

for H and He), as well as from the valence band. Their intensities (normalized with

photoionization cross-section) reflect the composition of the material. Characteristic core

level peaks are usually the strongest peaks in a pure element spectrum. We analyze these

characteristic core levels to determine the electronic structure and oxidation state of that

element in the sample.

Page 52: Copyright By Wei Guo 2021

33

With the exception of 1s spectra, core level spectra are doublets corresponding to

different total angular momentum quantum numbers. This arises from spin-orbit splitting.

For example, a 3d core level has n = 3, l = 2 and s = 1

2 . The total angular momentum j =

l ± s will then be 3/2 or 5/2. Since the relative peak intensities of doublets are determined

by the ratio of their degeneracy 2j+1 [11], the theoretical intensity ratio of the doublets is

2:3 for a 3d core level.

2.4 REFERENCES

1. K. G. Günther, Z. Naturforschg. A. 13 (12), 1081-1089 (1958).

2. A. Y. Cho, F. K. Reinhart, Appl. Phys. Lett. 21, 355 (1972).

3. J. R. Arthur Jr., J. Appl. Phys. 39, 4032 (1968).

4. R. C. Hatch, M. Choi, A. B. Posadas, and A. A. Demkov, J. Vac. Sci. Technol. B. 33,

061204 (2015).

5. P. Ponath, A. B. Posadas, R. C. Hatch and Alexander A. Demkov, J. Vac. Sci. Technol.

B. 31, 031201 (2013).

6. J.E. Mahan, Physical Vapor Deposition of Thin Films, Wiley-VCH, 2000.

7. M. Rodahl, F. Höök, A. Krozer, P. Brzezinski, B. Kasemo, Rev. Sci. Instrum. 66, 3924

(1995).

8. A. Ichimiya, P. I. Cohen, Reflection High Energy Electron Diffraction Cambridge

University Press (2004).

9. Y. Horio, Y. Hashimoto, A. Ichimiya, Appl. Surf. Sci. 100, 292 (1996).

10. A. Einstein, Ann. Phys. 322, 132 (1905).

11. S. Hofmann, Auger- and X-Ray Photoelectron Spectroscopy in Materials Science: A

User-Oriented Guide, Springer Berlin Heidelberg, Berlin, Heidelberg, (2013).

Page 53: Copyright By Wei Guo 2021

34

Chapter 3: Temperature dependence of the morphology and electronic

structure of ultrathin platinum on TiO2-teminated SrTiO3 (001)

In this chapter, we investigate properties of ultrathin Pt as a function of coverage

(up to 10 monolayers) on TiO2-terminated SrTiO3 (001) substrate at different temperatures

(200–800 ºC). In situ x-ray photoelectron spectroscopy, scanning electron microscopy, and

atomic force microscopy are used to observe the electronic structure and surface

morphology evolution of Pt. The authors find that although Pt will not wet SrTiO3 in the

thermodynamic sense, it forms a continuous film when deposited at 200 ºC due to the low

surface mobility. At 800 ºC, even at very low coverage, Pt forms nanoclusters showing

bulk-like metallic features in the photoemission spectra. We compare the observed

electronic structure evolution of Pt and the different growth patterns at low and high

temperatures with available theoretical calculations.

The contents of this chapter were published in: W. Guo, A. B. Posadas, and A. A.

Demkov, J. Vac. Sci. and Technol. B. 35, 061203 (2017); and A. B. Posadas, K. J.

Kormondy, W. Guo, P. Ponath, J. Geler-Kremer, T. Hadamek, and A. A. Demkov, J. Appl.

Phys. 121, 105302 (2017).

3.1 INTRODUCTION

Platinum (Pt) is a noble transition metal that has excellent chemical and thermal

stability as well as high electrical conductivity (9.4×106 S/m). These make Pt a popular

contact material, especially on oxides, because of its inertness toward oxidation [1]–[4]. In

addition, it is also a very efficient co-catalyst on titania and various other oxide catalysts

for enhancing the reduction half-reaction in solar water splitting to generate hydrogen [5],

[6]. On the other hand, SrTiO3 (STO) is a widely used substrate material in oxide epitaxy.

Page 54: Copyright By Wei Guo 2021

35

It has the cubic perovskite structure with a lattice constant of 0.3905 nm [7], [8]. In

addition, STO has a rather large dielectric constant (~300) and can readily be made

conductive by doping [8]. When these two materials are placed in contact, a Schottky

barrier (SB) with values ranging from 0.6 to 1.3 eV is formed at the interface, with upward

band bending in the STO [1], [3], [9]. Pt has a very high work function of ~5.7 eV [10],

while the electron affinity of STO is ~3.9 eV [8]. With a lattice parameter of 0.3923 nm

for Pt and 0.3905 nm for STO and both materials having a cubic structure, Pt and STO

appear to be well lattice-matched and suitable for epitaxial growth. However, it turns out

that Pt (001) does not normally grow epitaxially on STO (001) and instead prefers to grow

in the (111) orientation even on the lattice-matched STO (001) TiO2-terminated surface

[11].

There are a number of metals such as Al, Ti, Eu and Nb that can scavenge oxygen

from STO to form oxygen vacancies even at moderate temperature (200–400 ºC) and under

relatively high oxygen pressure (up to 10-2 Torr) [4]. However, Pt is one of the small set of

metals that has good stability to oxidation when deposited on STO. Through a combination

of the low oxide enthalpy of formation and high work function of Pt, electrons prefer to

transfer from STO into the metal, and hence STO is not reduced. This makes Pt on STO an

ideal system to study the formation of the initial Schottky barrier without the complications

of surface reactions.

Although there are many papers discussing the Pt/STO heterostructure, particularly

where Pt is used as an electrode [1], they mostly focus on micrometer-thick Pt films. There

are not as many studies about the first few monolayers (MLs) of Pt and its development

into a fully metallic layer under different growth temperatures. Chung and Weissbard

showed that the SB formation is a gradual process during the Pt accumulation in the first

Page 55: Copyright By Wei Guo 2021

36

monolayer, accompanied by charge transfer from STO to Pt [12]. They also reported that

the Pt core level binding energy (BE) shifts to lower values as the coverage of Pt increases

because of gradual charge transfer from STO to Pt [13]. The same type of a SB forms at

the Pt/BaTiO3/(Ba,Sr)TiO3 interface and a similar shift of the Pt core level has also been

reported for this system [3], [14]. Copel et al. measured the band bending in STO when Pt

is deposited on it and described the metallicity of Pt as it accumulates on STO. Theoretical

calculations using density functional theory (DFT) have been performed to study the SB

formation at the Pt/STO interface. [11], [15], [16]. Seo and Demkov explored the growth

of Pt on STO within the first ML using DFT and reported the electronic density of states

for different Pt coverage [11]. They suggested that Pt nanoclusters will start to form from

submonolayer coverage, which can be verified by measuring the core level binding energy

of Pt. By performing thin Pt layer by layer growth at different temperatures, we can verify

this theoretical result and obtain more understanding about the Schottky barrier formation,

electronic structure evolution, and surface morphology of Pt.

The surface morphology of Pt on STO has also been recently reported [17]-[19].

Polli et al. grew 20 nm Pt by molecular beam epitaxy (MBE) on differently terminated

STO substrates and showed by atomic force microscopy (AFM) and conventional

transmission electron microscopy [18] that the TiO2 termination resulted in a more uniform

and ordered orientation and size distribution of Pt clusters. Christensen et al. showed

scanning electron microscopy (SEM) images of Pt nanoparticles [19]. They point out that

with a Pt average coverage of 1 ML, the clusters aggregate into nanoparticles without

coalescing.

In this chapter, we present a systematic study of depositing 0.25, 0.5, 1, 2, 4, and

10 ML of Pt metal on a TiO2-terminated (001) STO surface in ultrahigh vacuum at

Page 56: Copyright By Wei Guo 2021

37

temperatures ranging from 200 to 800 ºC, and investigate the evolution of the electronic

structure by in situ x-ray photoelectron spectroscopy (XPS). We also use SEM and AFM

to investigate the morphology of the surface.

3.2 EXPERIMENT

The deposition is performed in a customized DCA Instruments M600 MBE

chamber with base pressure below 6 × 10-10 Torr. Nb:STO (001) (0.7 wt. % doped)

substrates of 5 × 5 mm2 from MTI Crystal are treated using the water boiling process to

yield TiO2-terminated surfaces [20]. The substrates are degreased ultrasonically in acetone,

isopropanol, and deionized (18 MΩ/cm) water for 5 min each, then soaked in boiling

deionized water for half an hour to dissolve excess surface SrO. The substrate is then

annealed in flowing oxygen for 4 h in a tube furnace at 950 ºC to form a well-ordered TiO2-

terminated surface. This is then followed by a 725 ºC anneal in the vacuum chamber for 1

h to remove volatile surface contaminants from air exposure. This process results in a single

TiO2-terminated surface with extremely low surface defect density [20]. Pt is evaporated

by an electron beam source operated at an emission current of ~110mA with an electron

energy of 7.75 keV. A quartz crystal monitor is used to measure the deposition rate of Pt.

The rate is kept around 0.1 nm/ min (~0.5 ML of Pt per minute) by adjusting the emission

current of the electron beam source. Reflection-high-energy electron diffraction with 21

keV electrons is used to monitor the surface quality and the orientation of Pt.

Pt deposition is commenced after the substrate temperature reaches the desired

value. An amount of Pt equivalent to a cumulative coverage of 0.25, 0.5, 1, 2, 4, and 10

ML is deposited on the substrate. After each target coverage is deposited, the sample is

cooled down and transferred in situ to the XPS chamber to measure the core level and

Page 57: Copyright By Wei Guo 2021

38

valence band spectra. The sample is then transferred back to the growth chamber for the

next deposition, and so on. Our XPS system uses monochromated Al Kα radiation as the

x-ray source and a VG Scienta R3000 hemispherical electron energy analyzer. The core

level spectra of Pt 4f, Sr 3d, Ti 2p, and O 1s were scanned for each coverage. A Hitachi

S5500 SEM is used to image the surface of samples grown at 200 and 800 ºC to determine

the effect of temperature on surface morphology. An Asylum MFP-3D atomic force

microscope (AFM) is used to image and quantify the surface morphology of the films.

3.3 RESULTS AND DISCUSSION

In Figure 3.1, we show the Pt 4f core level binding energy as a function of coverage

for deposition at 200 ºC. The Pt 4f7/2 peak is at ~72.4 eV at a Pt coverage of 0.25 ML and

gradually shifts to ~71.3 eV after a coverage of 2 ML is reached. This indicates a transition

of the Pt from a “partially oxidized” state to a fully metallic state. This result is consistent

with recent DFT calculations [11] that pointed out that, initially, Pt atoms on the TiO2-

terminated STO surface are located at the “hollow” site surrounded by four O atoms and

forming Pt–O bonds. This results in a partially oxidized state for the initial Pt causing the

core level to be shifted to higher BE. As more Pt is added, the Pt–Pt bonds gradually start

to become dominant. Finally, the fully metallic Pt layer forms and the BE corresponds to

that in bulk Pt metal. This transition has also been observed by Chung and coworkers [12],

[13].

Page 58: Copyright By Wei Guo 2021

39

Figure 3.1: Pt 4f core level spectrum for deposition at 200 ºC on STO as a function of Pt

thickness. At 0.25 ML coverage (the lowest) the Pt 4f7/2 peak has a binding

energy of 72.4 eV. It gradually shifts by 1.1 eV to the bulk metallic value

(71.3 eV) when Pt reaches 2 ML.

When examining the BE change as a function of deposition temperature, we find

the behavior shown in Figure 3.2. When depositing Pt at 200 ºC, the total 4f7/2 BE shift

with thickness, going from 0.25 to 2 ML of Pt, is about 1.0 eV. But when we increase the

growth temperature, the total shift is reduced, becoming less than 0.5 eV when depositing

at 800 ºC. This trend can be understood a follows. At lower temperature, Pt is less likely

to move away from the energetically preferred oxygen hollow site [11], [21]. At higher

temperature, however, as the surface mobility increases, Pt will form particles or clusters

more and more easily with increasing coverage as it does not wet STO [16]. At low

temperature, initially, Pt atoms are bonded to oxygen, with the BE of this partially oxidized

Page 59: Copyright By Wei Guo 2021

40

metal shifted significantly from that of the metal. This results in a large energy shift when

plotted as a function of coverage. As the temperature goes up, Pt atoms gain enough energy

to move across the surface and form metal clusters. Thus, even at low coverage the BE is

closer to the bulk metallic value. Thus, for the same Pt coverage, higher growth temperature

results in a lower BE value. The BE of Pt 4f7/2 at 2 ML coverage grown at 800 ºC almost

reaches the bulk metallic value, while it takes several more layers for the lower deposition

temperatures. At this coverage, the deposited Pt is not yet the same as the bulk Pt BE

because the Pt forms as nanoclusters (or nanoparticles). This effect is commonly explained

as the incomplete screening of metal clusters [22], with smaller metal clusters having larger

BE shifts. We have also measured the core level spectra of Sr 3d, Ti 2p, and O 1s. Their

binding energies as a function of Pt coverage are shown in Figure 3.3. All of them show a

uniform systematic shift of about 0.3 eV to lower binding energy while Pt is accumulating.

This is consistent with the shift being caused by upward band bending in STO. We find

that a Schottky barrier of ~0.65 eV is built up gradually as Pt is accumulating.

Page 60: Copyright By Wei Guo 2021

41

Figure 3.2: Pt 4f7/2 binding energy as a function of Pt thickness for different deposition

temperatures (200–800 ºC). All values shift to the bulk Pt metal value

(~71.3 eV) at different temperatures when Pt is sufficiently thick. Lower

deposition temperature will lead to larger total binding energy shift while

also taking more layers to reach the bulk metal value.

Figure 3.3: Sr 3d, Ti 2p, O 1s binding energy shift as a function of Pt thickness. There is

a ~0.3 eV shift for all three core levels while Pt is accumulating, which is

consistent with an upward band bending.

A high-count scan of the valence band spectrum as a function of coverage grown

at 800 ºC is shown in Figure 3.4. It can be seen that the Pt signal emerges gradually in the

region around 0–3 eV as the film thickness increases to 2 ML, indicating the presence of

individual Pt states at low coverage. At 2 ML, we see significant nonzero intensity at the

Fermi level and the onset of a metal-like valence band shape. This means that at 2 ML

Page 61: Copyright By Wei Guo 2021

42

coverage, Pt clusters are big enough to have a Fermi edge. We note that the growth

temperature also affects the size of Pt clusters, with higher temperatures yielding bigger

clusters due to the increased surface mobility. We use AFM to analyze the surface structure

of 2 ML Pt grown at 800 ºC on STO. In Figure 3.5, we can see the structure of the Pt

clusters. Our treatment to form TiO2-terminated STO surface produces a terrace structure

[23] as shown in Figure 3.5(a) and the Pt clusters in the same terrace align in a row

following the STO surface morphology. The average diameter of a cluster is 150 nm and

the average height is 0.5 nm. The number of Pt atoms contained in one such cluster is on

the order of ~106. This is roughly the critical size of Pt clusters needed to develop a Fermi

edge and have a bulk metal core level binding energy.

Figure 3.4: Pt/STO valence band spectrum grown at 800 ºC as a function of thickness. It

shows the Pt contribution to the signal gradually increase in the range of 0–3

eV. At 2 ML, it shows nonzero intensity at 0 eV which means the Fermi

edge starts to form.

Page 62: Copyright By Wei Guo 2021

43

Figure 3.5: (a) STO surface after water boiling treatment. It shows a terrace structure with

flat steps (b) AFM measurement of 2ML Pt/STO grown at 800 ºC and

corresponding line profile measurement. The Pt nanoclusters (with ~150nm

diameter and 0.5 nm height) arrange themselves in rows aligned along the

substrate terrace edges.

Page 63: Copyright By Wei Guo 2021

44

To gain further insight, we acquired SEM images for Pt films grown at 200 and 800

ºC, which are shown in Figure 3.6. For 200 ºC deposition, it is hard to infer a clear surface

structure for the first several monolayers of Pt with the resolution of our equipment. Even

for a 10 ML Pt/STO sample, there is still no clear cluster or particle structure seen at the

nanometer-length scale [as shown in Figure 3.6(a)]. Instead, the image shows a flat,

featureless coverage with a visible edge at the area blocked by sample holder when we

check it at micrometer scale [as shown in Figure 3.6(a) inset]. Although Pt does not wet

the STO surface [16], at low growth temperature, the mobility of Pt atoms is quite low, and

most of Pt atoms deposited on the substrate will stay where they arrive and will not move

around. As more Pt is deposited on the surface, a film with a flat surface forms. However,

at 800 ºC, things are quite different. Pt atoms have enough thermal energy to freely move

across the surface and form nanosize clusters with diameter ~80 nm at an average Pt

coverage of 1 ML [Figure 3.6(b)]. This agrees with Christensen’s high-resolution SEM

image of 1 ML Pt on STO which shows a similar nanocluster structure [19]. When we heat

up a sample grown at 200 ºC to high temperature (800 ºC for an hour), the surface

morphology drastically changes. The SEM image of such a sample is displayed in Figure

3.7(a) and shows that the initially flat film breaks and Pt clusters of various sizes form. We

use Hough transform to analyze the distribution of the cluster size and obtain the histogram

shown in Figure 3.7(b). Note that the algorithm is not accurate for very small clusters

(below 3 nm) due to the resolution of the images. However, we are still able to determine

that the vast majority of Pt clusters have radii in the 3–5 nm range with a separation of

~25–30 nm.

Page 64: Copyright By Wei Guo 2021

45

Figure 3.6: (a) SEM image of Pt/STO grown at 200 ºC for 10 ML. At the nanometer scale

we can only see a featureless surface with a clear edge when we check it at

micrometer scale (as shown in the inset). The left area in the inset has no Pt

grown on because of shadowing by the sample holder during deposition. (b)

SEM image of Pt/STO grown at 800 ºC for 1 ML. It shows uniform

nanoclusters with average lateral size of ~80 nm.

Page 65: Copyright By Wei Guo 2021

46

Page 66: Copyright By Wei Guo 2021

47

Figure 3.7: (a) SEM image of Pt/STO with 10 ML of Pt grown at 200 ºC and annealed at

800 ºC for an hour in air. The flat surface breaks out and splits into small

nanoclusters. Pt clusters have radii in the 3-5 nm range with a separation of

~25-30 nm. (b) Histogram of the cluster sizes obtained from the SEM

image. The first two columns are gray because of the overcounting of the

algorithm at very small radius. Red line is the normal distribution fit of the

rest of the columns centered at 3.75 nm.

3.4 CONCLUSIONS

We investigated the deposition of Pt on atomically smooth, TiO2-terminated (001)-

oriented STO substrates for coverages up to 10 ML in the temperature range from 200 to

800 ºC. In situ XPS, AFM and SEM were used to explore the core level binding energy,

valence band spectrum, and surface morphology as functions of coverage and deposition

temperature. We found that owing to the relatively low mobility of Pt at low temperature,

initially, Pt atoms tend to stay at the low energy oxygen hollow site of the STO surface and

thus exhibit a partially oxidized state in core level XPS in agreement with DFT calculations

[11]. As coverage increases the Pt valence band develops along with the formation of a flat

Pt film, culminating in the emergence of the Fermi edge after 2 ML. However, when the

growth temperature is increased, Pt has enough mobility to move across the surface and

form clusters. Higher growth temperature results in larger Pt clusters that demonstrate

metallic XPS features at lower coverage than if the deposition were done at lower

temperature. SEM clearly shows nanoclusters forming at high growth temperature and a

continuous film at low temperature. The critical Pt cluster size required to exhibit a clear

Fermi edge and have bulk metallic core level is found to be ~106 atoms per cluster by AFM.

Page 67: Copyright By Wei Guo 2021

48

3.5 ACKNOWLEDGMENT

The authors thank Hosung Seo for insightful discussions. This work was supported

by the Air Force Office of Scientific Research under Grant Nos. FA9550-12-10494 and

FA9550-14-1-0090.

3.6 REFERENCES

1. Q. Fu and T. Wagner, Surf. Sci. Rep. 62, 431 (2007).

2. M. Mrovec, J.-M. Albina, B. Meyer, and C. Els€asser, Phys. Rev. B 79, 245121

(2009).

3. R. Schafranek, S. Payan, M. Maglione, and A. Klein, Phys. Rev. B 77, 195310

(2008).

4. A. B. Posadas, K. J. Kormondy, W. Guo, P. Ponath, J. Geler-Kremer, T. Hadamek,

and A. A. Demkov, J. Appl. Phys. 121, 105302 (2017).

5. M. S. Chen and D. W. Goodman, Science 306, 252 (2004).

6. M. Kitano, K. Tsujimaru, and M. Anpo, Top. Catal. 49, 4 (2008).

7. B. Reihl, J. G. Bednorz, K. A. M€uller, Y. Jugnet, G. Landgren, and J. F. Morar,

Phys. Rev. B 30, 803 (1984).

8. A. A. Demkov and A. B. Posadas, Integration of Functional Oxides with

Semiconductors (Springer, New York, 2014).

9. M. Copel, P. R. Duncombe, D. A. Neumayer, T. M. Shaw, and R. M. Tromp, Appl.

Phys. Lett. 70, 3227 (1997).

10. J. Robertson, J. Vac. Sci. Technol. B 18, 1785 (2000).

11. H. Seo and A. A. Demkov, Phys. Rev. B 92, 245301 (2015).

12. Y. W. Chung and W. B. Weissbard, Phys. Rev. B 20, 3456 (1979).

13. M. K. Bahl, S. C. Tsai, and Y. W. Chung, Phys. Rev. B 21, 1344 (1980).

14. F. Rao, M. Kim, A. J. Freeman, S. Tang, and M. Anthony, Phys. Rev. B 55, 13953

(1997).

15. A. Asthagiri and D. S. Sholl, J. Chem. Phys. 116, 9914 (2002).

16. H. Seo, A. B. Posadas, and A. A. Demkov, J. Vac. Sci. Technol., B 30, 04E108

(2012).

17. J. Son, J. Cagnon, and S. Stemmer, J. Appl. Phys. 106, 043525 (2009).

Page 68: Copyright By Wei Guo 2021

49

18. A. D. Polli, T. Wagner, T. Gemming, and M. R€uhle, Surf. Sci. 448, 279 (2000).

19. S. T. Christensen, B. Lee, Z. Feng, M. C. Hersam, and M. J. Bedzyk, Appl. Surf.

Sci. 256, 423 (2009).

20. R. C. Hatch, M. Choi, A. B. Posadas, and A. A. Demkov, J. Vac. Sci. Technol., B

33, 061204 (2015).

21. H. Brune, Surf. Sci. Rep. 31, 125 (1998).

22. G. K. Wertheim and S. B. DiCenzo, Phys. Rev. B 37, 844 (1988).

23. M. Choi, “Monolithic integration of functional perovskite structures on Si” Ph.D.

thesis (The University of Texas at Austin, Austin, TX, 2014).

Page 69: Copyright By Wei Guo 2021

50

Chapter 4: EuO epitaxy on SrTiO3 by oxygen scavenging

The EuO/SrTiO3 heterojunction is a promising combination of a ferromagnetic

material and a two-dimensional electron system. We explore the deposition of Eu metal on

STO/Si pseudo-substrates, with varying STO thickness, under ultrahigh vacuum

conditions. By varying the thickness of the STO layer (2-10 nm) and the deposition

temperature (20-300 °C), we investigate the process by which oxygen is scavenged from

STO by Eu. In situ x-ray photoelectron spectroscopy is used to investigate the electronic

structure of the nominal Eu/STO/Si stack. We find that as a result of Eu deposition,

epitaxial EuO is formed on thick STO (6-10 nm), leaving behind a highly oxygen-deficient

SrTiO3-δ layer of ∼4 nm in thickness. However, if the thickness of the STO layer is

comparable to or less than the scavenging depth, the crystal structure of STO is disrupted

and a solid state reaction between Eu, Si, and STO occurs when the deposition is done at a

high temperature (300 °C). On the other hand, at a low temperature (20 °C), only a 1-2 nm-

thick EuO interlayer is grown, on top of which the Eu metal appears to be stable. By

analyzing the Eu/EuO interface, we also find that electrons transfer from Eu metal into

EuO and induce an unexpected downward band bending at the interface. We use density

functional theory (DFT) to calculate the atomic and electronic structure of the interface and

find that the theoretical and experimental results agree with each other. For further study

on the manipulation of the 2DEG at the EuO/STO interface, we insert a BTO layer under

STO and explore the field-effect arising from the BTO polarization on the 2DEG using soft

x-ray angle-resolved photoemission spectroscopy (SX-ARPES). This study elucidates the

EuO growth process under different conditions and provides better understanding and

control of this system.

Page 70: Copyright By Wei Guo 2021

51

The major content of this chapter was published in: W. Guo, A. B. Posadas, S. Lu,

D. J. Smith and A. A. Demkov, J. Appl. Phys. 124, 235301 (2018). And L. Gao, W. Guo, A.

B. Posadas and A. A. Demkov, Phys. Rev. Materials 3, 094403(2019). Collaborators at

Arizona State University are responsible for the TEM results. ARPES measurements were

greatly supported by Prof. Strocov at the Paul Scherrer Institute (PSI).

4.1 INTRODUCTION

The ferromagnetic semiconductor EuO is a promising material for spintronic

devices [1]-[3] due to its very large magnetic moment (7 μB per Eu atom) and large spin-

splitting of the conduction band leading to almost 100% spin polarization of carriers [4],

[5]. The semiconducting behavior combined with the high magnetic moment makes it an

ideal spin-filter material below its Curie temperature of 69 K [6]. EuO has been epitaxially

integrated onto graphene [7], silicon [1], [8], GaN [1], and several oxide substrates [9] to

provide ferromagnetism in these systems. In addition, EuO also exhibits large

magnetoresistance, strong magneto-optical effect, and unusual transport properties [1], [8]-

[11]. These features make EuO a fascinating oxide for fundamental research, in addition to

the possibility of enabling new kinds of device applications involving spin transport.

The integration of EuO on heterostructures exhibiting a two-dimensional electron

gas (2DEG) is a potentially interesting approach to explore the interaction of strong

magnetism with the sheet charge, particularly for the case of the 2DEG at STO interfaces

[9], [12], [13]. By controlling the deposition of an oxide with a large negative enthalpy of

formation like EuO on STO [14], one can stabilize a highly confined conductive layer of

oxygen-deficient STO at the interface [15]. This heterostructure then offers a way to

combine strong ferromagnetism and a 2DEG in one system. Kormondy et al. have recently

Page 71: Copyright By Wei Guo 2021

52

reported the observation of a large linear positive magnetoresistance in the 2DEG at the

EuO/SrTiO3 interface [9]. Lömker et al. also report a magnetically tunable two-

dimensional electron system in the same structure [12]. Prinz et al. created quantum wells

with ultrathin EuO layers and widened the bulk EuO bandgap from 1.19 eV to ∼1.4 eV in

ultrathin films by the quantum confinement effect [16].

Growing thin films of high quality, stoichiometric EuO is difficult because of the

high stability of the competing Eu2O3 phase [6], [9]. Thin-film growth of EuO has been

reported in several studies [6]-[9], [12], [17], [18] and it is crucial to control the oxygen

partial pressure in order to form stoichiometric EuO and prevent over-oxidation [18].

Alternatively, stoichiometric EuO can be grown on certain substrates under Eu adsorption-

controlled conditions [7], [8], [19]. EuO integration on Si has been reported by several

groups [6], [8], [18], [20], and the EuO/Si interface has been predicted to be

thermodynamically stable if the two materials are in contact [21]. Lettieri et al. reported an

epitaxial EuO films grown on Si by molecular beam epitaxy (MBE) using an SrO buffer

[18]. Sr deoxidation is used to remove native SiO2 [22], [23] and 5 monolayers (ML) of

SrO are grown to provide a lattice-matched surface for EuO growth (0.3% mismatch).

Caspers et al. grew polycrystalline EuO directly on HF-treated, H-passivated Si substrates

by precisely controlling the Eu and oxygen flux [6]. These authors also grew epitaxial EuO

[001] on Si [001] by passivating the Si surface with a 13-Å silicon sub-oxide (SiOx) buffer

layer prior to EuO deposition [20]. However, this process also led to interfacial silicide

formation. Direct EuO growth on Si was reported by Averyanov et al. who used sub-

monolayer reconstructions of Eu on Si to create a Zintl-like template to avoid Si oxidation

[8]. There are also reports of epitaxial EuO growth on STO with buffer layers such as BaO

or SrO to prevent over-oxidation of EuO due to oxygen out-diffusion from STO [24], [25].

Page 72: Copyright By Wei Guo 2021

53

Common to all reported growth methods, EuO growth is typically performed at

temperatures between 300 and 450 °C. Furthermore, controlling the oxygen pressure is

always crucial to form high quality, stoichiometric EuO. Posadas et al. found that EuO can

be grown directly on an STO substrate by Eu deposition in ultra-high vacuum (UHV)

without providing any oxygen [14]. It has been shown that when metals with large enthalpy

of oxide formation (such as Al and Eu) are deposited on STO substrates, oxygen can be

scavenged from the top STO layers and form oxides on top even without additional oxygen.

The oxygen vacancies created in STO create a conductive layer on the STO side of the

interface. This conductive layer in STO due to oxygen vacancies has also been reported in

other studies [9], [14], [26].

Eu, with very large oxide formation enthalpy and a low work function of 2.5 eV, is

expected to scavenge significant amounts of oxygen from STO to form EuO, leaving

behind a highly conductive oxygen-deficient SrTiO3-δ layer [14]. Kormondy et al. used this

process to grow epitaxial EuO on STO single-crystal substrates at 300 °C in UHV. The

spin-polarized 2DEG formed at the EuO/STO interface displayed positive linear

magnetoresistance when the EuO was below its Curie temperature. From electron energy

loss spectroscopy (EELS) analysis in cross-sectional scanning transmission electron

microscopy, the oxygen-deficient SrTiO3-δ layer was observed to be ∼4 nm deep [9]. This

thickness appears to be the “oxygen scavenging depth” of Eu when deposited on STO at a

temperature of 300 °C. It is expected that this depth will vary strongly with temperature

since oxygen scavenging involves atomic diffusion. Therefore, it is important to investigate

the effect of temperature, and also STO layer thickness, on the scavenging of oxygen in

STO by Eu. In particular, it would be interesting to establish what happens when the STO

Page 73: Copyright By Wei Guo 2021

54

thickness becomes comparable to or smaller than the oxygen scavenging depth. In other

words, how much oxygen can Eu scavenge from STO before STO starts to be degraded?

EuO has a cubic rock-salt structure with a lattice constant of 5.14 Å [27]. The lattice

matching of EuO to a cubic perovskite structure requires a 45° in-plane rotation, essentially

lining up with the rock salt-like layers of the perovskite. The nominal matching is therefore

between [100] direction of the perovskite and [110] direction of rock salt. Schematics of

the matching interface structure are displayed in Figure 4.1. In the case of EuO matching

with STO (a = 3.905 Å), the lattice mismatch is ∼7%.

Page 74: Copyright By Wei Guo 2021

55

Figure 4.1: Schematic of EuO/STO epitaxial arrangement: (a) side view (b) top view.

In this chapter, we report on the fabrication of EuO/SrTiO3/Si heterostructures by

depositing metallic Eu under UHV conditions on STO/Si pseudo-substrates with varying

STO layer thickness. Different STO thicknesses allow us to place a limit on the amount of

available oxygen for reaction with Eu. We investigate the oxygen scavenging mechanism

Page 75: Copyright By Wei Guo 2021

56

by varying both the STO thickness (from 2 to 10 nm) and the temperature of Eu deposition

(from 20 °C to 300 °C). Crystallinity of the growing film is monitored by in situ reflection

high-energy electron diffraction (RHEED). We analyze the composition of the resulting

structure by in situ x-ray photoelectron spectroscopy (XPS). DFT is also used to calculate

the electronic structure of the Eu/EuO interface, which we then compare to experimental

XPS results. Further study on the field-effect of the 2DEG in EuO/STO has also been

started but more experiments are needed to draw any conclusions from this study.

4.2 EXPERIMENTS ON OXYGEN SCAVENGING OF STO BY EU

The deposition is performed in a customized DCA 600 MBE system with a base

pressure of 6 × 10−10 Torr. The Si substrates (p-type with a resistivity of 0.1-0.2 Ω cm from

University Wafer) are diced into 20 × 20 mm2 pieces and degreased ultrasonically in

acetone, isopropanol, and deionized water (18MΩ/cm) for 5 min each and then exposed to

UV/ozone to remove carbon from the surface. The Si substrates are then annealed in the

vacuum chamber at 700 °C for 10 min followed by the growth of a 2-nm-thick STO

epitaxial film at 250 °C which is crystallized by annealing in vacuum at 500 °C [21], [28],

[29]. The STO films are grown such that there is no interfacial SiO2 present by separating

the oxygen gas introduction with the STO crystallization anneal [28]–[31]. Additional STO

layers can be grown without SiO2 formation by depositing disordered STO at 250 °C using

co-deposition under moderate oxygen pressure (∼mid 10-7 Torr), and then annealing in

UHV at 500 °C to further crystallize the STO film. It is important to prevent SiO2 formation

to ensure that the only oxygen source for scavenging by Eu is from the STO layer.

We prepared STO/Si pseudo-substrates with STO layer thicknesses of 2, 3, 4, 6,

and 10 nm. For each STO layer thickness, different coverages of Eu are deposited from an

Page 76: Copyright By Wei Guo 2021

57

effusion cell with the cell temperature fixed at 530 °C. This cell temperature has been

determined to yield an Eu metal evaporation rate of ∼2.5 Å/min. An accurate measurement

of the arrival rate of Eu is done by measuring the thickness (using x-ray diffraction finite

size oscillations) of epitaxial Eu2O3 on GaN substrates following growth for a fixed amount

of time [32]. As the sticking coefficient of Eu is unity under Eu2O3 growth conditions, the

arrival rate of Eu on the substrate can be accurately determined.

The Eu metal is deposited under two substrate temperatures designated here as high

(300 °C) and low (20 °C). RHEED with 21 keV electrons is used to monitor the surface

crystallinity changes going from the bare STO surface to EuO during growth. In situ x-ray

photoelectron spectroscopy (XPS) with monochromated Al Kα radiation and a VG Scienta

R3000 hemispherical electron energy analyzer are used to check the core levels of relevant

elements as well as the valence band (VB).

The Eu metal is deposited in a step-by-step fashion and checked with XPS after

each step. Each Eu deposition step lasts for a few minutes, with a deposition pause after

each step in order to transfer the sample into the in situ XPS chamber for analysis, and then

moved back to the MBE growth chamber to deposit the next step. This process is repeated

until the total Eu deposition is finished. This procedure allows us to see the evolution of

the chemical and electronic structure of the sample as Eu is deposited. The core levels Eu

3d, Eu 4d, Si 2p, Ti 2p, and O 1s are measured as a function of Eu coverage to obtain

information about the scavenging process.

Page 77: Copyright By Wei Guo 2021

58

4.3 RESULTS AND DISCUSSION

4.3.1 High temperature deposition with thick STO film

We previously showed that EuO can be epitaxially grown on STO single crystal

substrates by oxygen scavenging from direct Eu metal deposition at 300 °C with a ∼4 nm

scavenging depth [9]. In the present work with EuO/SrTiO3/Si heterostructures, we first

perform Eu deposition on 10 nm STO on Si at 300 °C. This STO thickness appears to be

sufficient to provide enough oxygen to Eu metal in order to form stoichiometric EuO with

several nm thickness, similar to our earlier results on STO single crystal substrates [9]. The

Eu metal was deposited for ∼30 min, which is expected to form roughly 75 Å of Eu metal,

assuming unity sticking coefficient. Eu metal has a body-centered-cubic structure with 4.58

Å lattice constant, and EuO has a rock-salt structure with a lattice constant of 5.14 Å [27].

Assuming the entire 75 Å Eu metal layer is oxidized to EuO by oxygen scavenging, this

would result in a EuO thickness of ∼54 Å. RHEED patterns after Eu deposition along the

[110] and [100] STO azimuths are shown in Figure 4.2. The pattern observed is associated

with the rock-salt surface indicating the formation of epitaxial EuO. During Eu deposition,

the RHEED pattern for the STO layer first becomes blurry and fades away after 5 min

(∼12.5 Å of Eu metal) as newly arrived Eu covers STO in a disordered manner. As Eu

deposition is continued, the RHEED pattern for EuO gradually appears, becoming sharper

and brighter as deposition proceeds. This is a clear indication that oxygen atoms have been

scavenged from the STO layer through the growing EuO layer up to the surface to form

crystalline EuO.

Page 78: Copyright By Wei Guo 2021

59

Figure 4.2: RHEED pattern for EuO grown on 10 nm STO/Si at 300 °C.

To better understand the EuO/STO interface, we performed detailed electron

microscopy observation for several EuO films grown on STO single-crystal substrates by

oxygen scavenging. Representative images are shown in Figure 4.3. The samples were

prepared in cross-section geometry using standard mechanical polishing and argon ion-

milling, and images were recorded with an ARM200F scanning transmission electron

microscope operated at 200 keV. The beam convergence angle was set at 20 mrad, and the

collection angles were 0-22 mrad for large-angle bright-field (BF) imaging and 90-150

mrad for high-angle annular-dark-field (HAADF) imaging.

Page 79: Copyright By Wei Guo 2021

60

Figure 4.3: Aberration-corrected TEM images of the EuO/STO interface with the atomic

model overlaid. Pink circles are Eu, red circles are oxygen, green circles are

Sr, and blue circles are Ti. (a) and (b) HAADF and BF pair of images for

EuO [110]/STO [100] projections. (c) and (d) HAADF and BF pair of

images for EuO [100]/STO [110] projections.

Figures 4.3(a) and 4.3(b) are a pair of HAADF and BF images of the EuO/STO

interface for EuO [110]/STO [100] crystal projections. The abrupt interface and the high

quality EuO epitaxial growth are clearly apparent. Figures 4.3(c) and 4.3(d) show another

HAADF/BF pair of images for EuO [100]/STO [110] projections, again confirming the

excellent crystallinity of the epitaxial EuO layer.

In situ XPS scans confirm a pure Eu2+ signal (no Eu0 or Eu3+) consistent with

stoichiometric EuO from both the Eu 3d and 4d core levels (not shown), and also from the

valence band (VB) as shown in Figure 4.4. The EuO VB shows a strong peak at ∼2 eV

from the Eu 4f states and a smaller feature at higher binding energy associated with O 2p

Page 80: Copyright By Wei Guo 2021

61

in EuO. It is obvious from the XPS data that no Eu3+ oxidation state is present, which would

normally be found in the region from 6 to 8 eV. This result means that EuO is not over-

oxidized and that Eu oxidation from scavenging is self-limited to the +2 oxidation state at

this deposition temperature. These results confirm that crystalline EuO can be grown on

thick (10 nm or more) STO/Si pseudo-substrates, with the same behavior as on a bulk STO

substrate.

Figure 4.4: VB spectrum of Eu grown on 10 nm STO/Si. Only Eu2+ and oxygen features

are visible.

4.3.2 High temperature deposition with thin STO film

Next, Eu metal is deposited on thinner STO layers at the same temperature (300

°C). For STO layers with a thickness of 6 nm, the same oxygen-scavenging process as that

observed on bulk STO occurs, and EuO is epitaxially formed for several nm. However,

when the STO layer becomes thinner than the scavenging depth (∼4 nm) at this

Page 81: Copyright By Wei Guo 2021

62

temperature, the RHEED pattern evolution is different. A pattern similar to Figure 4.2

emerges briefly but vanishes later during growth. The sequence of RHEED patterns

observed is shown in Figure 4.5. The same sequence of RHEED patterns is also observed

for growth on thinner STO layers (2-3 nm).

Figure 4.5: RHEED pattern evolution during Eu deposition on a thin (2 nm) STO layer.

The disappearance of the EuO RHEED pattern indicates either the loss of EuO

crystallinity or the accumulation of some other non-crystalline phase on the surface. By

checking the XPS of Eu deposited on <4 nm STO film, we find the presence of another

feature at ∼1 eV binding energy in the VB spectrum [Figure 4.6(a)], which is associated

with EuSiy. This is also confirmed in Eu 3d and 4d core level spectra (not shown), which

also have an extra EuSiy feature that causes a ∼0.4 eV shift of the leading edge to lower

binding energy [19].

Page 82: Copyright By Wei Guo 2021

63

Figure 4.6: (a) VB spectra of Eu on 2-nm STO/Si. Two types of Eu peaks are visible. (b)

Si 2p signal of Eu deposited on 2-nm STO/Si. Silicon metal signal is almost

gone but SiOx and EuSiy appear.

Page 83: Copyright By Wei Guo 2021

64

Surprisingly, the normally buried Si 2p signal also now appears in the XPS scans

and shows both silicide (Si− at 98.2 eV) and silicate (Si+ at 100.8 eV) features. Only a small

amount of elemental Si remains (Si0), indicating a relatively thick Si reaction layer [Figure

4.6(b)]. We overlay the unreacted EuO VB spectra and Si metal signal for comparison in

Figure 4.6 with dashed lines, so that the shifts are more obvious. The appearance of Eu

silicide indicates that Eu reacts with the Si substrate at some point during growth, most

likely when the EuO RHEED pattern disappears.

To further explore how this reaction takes place, we performed a step-by-step Eu

deposition on thin STO (3 nm) to see the evolution of the electronic structure. The evolution

of the VB spectrum is shown in Figure 4.7.

Page 84: Copyright By Wei Guo 2021

65

Figure 4.7: VB spectra of step-by-step Eu growth on 3-nm STO. Growth times are shown

with different colors that correspond with the spectra.

For the first 10 min (∼25 Å Eu metal), the spectra show only EuO in the VB (Figure

4.7, red line) and elemental Si in the Si 2p scans (not shown here). The RHEED pattern

also corresponds to that of crystalline EuO during this time. When more Eu is deposited,

the RHEED pattern becomes blurry and the VB spectrum starts to show a gradually

increasing EuSiy signal, indicated by the leading edge shifting to lower binding energy.

The presence of possible Eu metal (Eu 5d feature) cannot be excluded from the VB

spectrum alone because it is too small to distinguish from the broad EuSiy signal. The Eu

4d signal shows that there is likely no Eu metal present because the characteristic multiplet

splitting in the metallic Eu 4d spectrum is not observed [32]. At the same time, the Si 2p

spectrum also starts to show both silicide and silicate features. From these results, it is

inferred that Eu oxygen scavenging happens during the first 15-20 min (∼50 Å Eu metal).

After this point, there is not enough oxygen in the thin STO for it to maintain its structure.

We know at 300 °C, Eu has the ability to scavenge oxygen from STO up to 4 nm

deep. When the STO film is thinner than 4 nm, excess Eu results in complete oxygen

depletion of STO. The STO is unable to maintain its structure and is essentially

decomposed by losing too much oxygen. At this temperature, Si atoms are able to diffuse

outward and react with EuO (as well as SrO + Ti). The existence of EuSiy and silicates is

clearly seen from the Eu and Si core level and VB spectra. The likely reaction that could

occur is EuO + SrTiO3 + Si→TiSi2 + SrSiOx + EuSiy. The disappearance of the EuO

RHEED pattern is likely because the reaction forms a mixture of several compounds that

physically break up the EuO layer so that it is no longer crystalline. This is an intriguing

result because it appears to contradict the prediction that EuO is thermodynamically stable

in contact with silicon [19]. Other groups have grown EuO on buffered Si at temperatures

Page 85: Copyright By Wei Guo 2021

66

higher than 300 °C [6], [8], [18], which implies that the decomposition of STO is acting as

some kind of catalyst for the reaction of EuO and Si.

4.3.3 Low temperature deposition

The degree of oxygen scavenging from STO by an overlayer must depend on the

mobility of oxygen atoms/ions through the oxygen-deficient STO and through the growing

metal oxide layer. Because oxygen transport is required, this phenomenon must be strongly

dependent on the substrate temperature. We also perform Eu deposition on thin STO (<4

nm) at room temperature (20 °C) to try and suppress the scavenging process and confirm

the temperature dependence. Initially, a RHEED pattern associated with crystalline EuO

still shows up with clear and sharp features, which means that the usual oxygen scavenging

process occurs first, even at a low temperature. This is expected because oxygen atoms do

not initially have to travel far. The crystalline EuO pattern, however, quickly becomes very

diffuse and dim and stays like that for the rest of the deposition. XPS analysis of this sample

showed no EuSiy and SiOx signals in the Si 2p spectrum, indicating that the thin STO layer

remained intact and prevented Si from diffusing outward. The VB spectrum also showed a

clear Eu metal signal [Figure 4.8(a)]. The feature around 0 eV [Figure 4.8(a), black peak],

which is separated from the main Eu 4f peak at 2 eV, comes from Eu 5d electrons at the

Fermi level, which only appears in Eu metal [33], [34]. The Eu 4f peak can be decomposed

into two components, coming from Eu metal and Eu2+. This observation is also supported

by recent density functional theory (DFT) calculations of the Eu/EuO interface by Gao et

al. [35]. The Eu 4d core level spectrum further confirms the existence of Eu metal because

of the clearly separated multiplet peaks around 127-132 eV corresponding to different total

angular momentum J quantum numbers [32] [Figure 4.8(b)].

Page 86: Copyright By Wei Guo 2021

67

Figure 4.8: (a) Valence band spectrum of Eu on 2-nm STO/Si at 20 °C. (b) Eu 4d

spectrum with clearly resolved multiplets characteristic of Eu metal.

Page 87: Copyright By Wei Guo 2021

68

Figure 4.9: Valence band evolution of Eu deposition on 2-nm STO/Si at 20 °C.

Step-by-step Eu deposition on 2 nm STO/Si at 20 °C indicates that Eu metal shows

up after ∼20 min (∼50 Å Eu metal) of Eu deposition (Figure 4.9). Only crystalline EuO is

formed in the first 10 min as determined by XPS VB spectra and RHEED with an Eu 4f

peak located at a binding energy of ∼2.2 eV. After ∼20 min, the Eu 5d metal feature starts

to appear and increases in intensity [Figure 4.9, orange line]. This means that a 1-2 nm-

thick EuO layer forms, beyond which point Eu cannot scavenge more oxygen from the

STO layer and only Eu metal accumulates on the surface. The Eu metal 4f peak centered

at a binding energy of ∼1.9 eV continuously increases after 20 min while the Eu2+ 4f peak

at 2.5 eV remains and decreases, eventually causing the leading edge of the combined 4f

peak to shift to lower binding energy. The O 2p feature in the VB is gradually buried by

Page 88: Copyright By Wei Guo 2021

69

the Eu metal signal which has broad peaks in the VB spectrum in the region 3-10 eV. It is

also worth mentioning that Eu metal is very reactive to any residual oxygen present. Eu

metal films will show a strong Eu3+ component due to the residual oxygen even when

stored in ultra-high vacuum (1 × 10−9 Torr) after a day or two.

We also looked at the Si 2p and O 1s spectra in the step-by-step Eu growth at 20

°C (not shown). The Si 2p signal only showed elemental Si peaks with a continuously

decreasing intensity with virtually no chemical shift. This confirmed no reaction with the

Si at this temperature. The O 1s signal maintained its intensity in the first 10min (∼25 Å

Eu metal) and started to decrease after 10min. This indicated that EuO stopped forming

after 10 min and Eu metal started to bury the EuO after this time. Combining all these

results, we conclude that at a low temperature, oxygen scavenging is limited by the very

low mobility of oxygen atoms/ions. Only a small amount of oxygen is scavenged from

STO right at the interface, allowing the STO layer to maintain its structure during excess

Eu deposition. Without any oxygen supply, more Eu can only stack in metallic form. We

also performed the same experiment at 150 °C, and the results appear to be the same as

that at 20 °C. Further experiments confirm that the oxygen scavenging process is

suppressed below 150 °C growth temperature.

4.3.4 Theoretical verification by DFT

Gao et al. [35] used DFT to study the atomic and electronic structure of the Eu/EuO

interface, as we have seen from low temperature Eu deposition on STO. Calculations

predict that electrons transfer from Eu metal into EuO and induce an unexpected downward

band bending at the interface. Accounting for spectral broadening and attenuation of the

signal from subsurface layers, the calculated layer-resolved total density of states agrees

Page 89: Copyright By Wei Guo 2021

70

well with experimental XPS in the valence-band region. The total 4f spectrum contains

contributions from both Eu and EuO, with the latter component significantly broadened as

a result of band bending. This bending and charge transfer originate from Eu Fermi-level

pinning at the EuO charge neutrality level, which has been suggested to be located above

the conduction-band bottom.

Calculation details are not shown here and people who interested can refer to

original paper published by Gao et al. [35]. We present below the related theoretical

verification and corresponding XPS spectra and fitting results.

Figures 4.10(a) and 4.10(b) show the experimental XPS data for Eu step-by-step

deposition on STO/Si pseudo-substrate at 20 °C. In Figure 4.10(a), we show the valence-

band (VB) XPS spectrum for a sample after 10 min of Eu deposition on STO. Only rocksalt

EuO forms at this time as a result of oxygen scavenged from STO by Eu atoms. The EuO

is crystalline as indicated by RHEED, as we mentioned in previous sections. The Eu2+ 4f

peak is observed at a binding energy of 2.2 eV, and the smaller and broader peak centered

at 5 eV is from the O 2p band. From 6 to 15 eV the spectrum is flat without any peaks,

indicating that there is no Eu3+ present. This is important, as the formation of the sesqui-

oxide Eu2O3, which is not ferromagnetic, could reduce the spin-filtering efficiency of a

EuO barrier [28].

Page 90: Copyright By Wei Guo 2021

71

Figure 4.10: (a) Valence-band XPS spectrum of 10 min Eu deposition on 2-nm STO/Si

pseudo-substrate at 20 °C. This is consistent with the spectrum from bulk

EuO. (b) Valence-band XPS spectrum of 60 min Eu deposition at 20 °C. A

Eu/EuO/STO/Si structure is formed. (c) The attenuated and broadened 4f

DOS of the whole heterostructure (“calculated pseudo-XPS spectrum”). The

Eu metal 4f DOS is marked as dark blue, the 4f DOS of each EuO layer is

marked as magenta, the total EuO 4f DOS is marked as violet, and the total

4f DOS is marked as red. The inset gives the 4f DOS of each EuO layer on a

larger scale. (e) A three-dimensional representation of the “calculated-

pseudo-XPS spectrum.” Here the spectrum of each layer is shown separately

along the y direction. The peak position of 4f EuO DOS of each layer is

connected with a violet line. (Figure is also published in Ref. [35])

In Figure 4.10(b), the VB spectrum of a sample after 60 min of Eu deposition is

shown. We can see the presence of the metallic Eu spectrum. Due to the low deposition

Page 91: Copyright By Wei Guo 2021

72

temperature resulting in low oxygen mobility through EuO, Eu is unable to continuously

scavenge oxygen from STO. This limits the thickness of EuO to about 2 nm with Eu metal

accumulating on top of the EuO. Compared with Figure 4.10(a), which shows the spectrum

of pure EuO, several other features can be seen due to the presence of Eu metal. There is a

clear, nonzero intensity at the Fermi energy appearing as a small step, corresponding to the

partially occupied Eu 5d band, which is a clear signature of the metal. This feature emerges

at 30 min deposition time and increases in relative intensity as more Eu is deposited. The

Eu 4f feature becomes asymmetric and is now required to be fit with two components

instead of one, as shown in more detail in Figure 4.10(b). The 4f component at 2.2 eV is

the Eu2+ signal from EuO, while the other 4f component at 1.9 eV is interpreted as an Eu

metal 4f signal. As more Eu metal is deposited, the 4f peak clearly shifts toward lower

binding energy, because the Eu component at 1.9 eV is increasing in intensity while the

EuO component at 2.2 eV is being progressively buried. This is a clear indication that at

some point oxygen is no longer able to diffuse through existing EuO at this temperature

and Eu metal begins to accumulate atop the EuO layer. There are also additional broad and

weak features in the 4 - 14 eV binding energy range that appear in the metallic phase that

quickly bury the EuO oxygen 2p peak. These are likely related to Eu 6s electrons in Eu

metal as seen in synchrotron measurements [36]. It should be noted that because oxygen is

supplied via diffusion from STO, there is likely a slight reduction in oxygen content in the

film as the EuO grows thicker. It is therefore likely that a slightly oxygen-deficient EuO

layer is present right at the interface up to the composition range of the EuO phase, which

is reported to be not more than 1 mol% [37], [38]. However, since oxygen solubility in Eu

is extremely small [39], any oxygen not reacting to form EuO is likely to segregate to the

(oxygen-deficient) EuO side of the interface, resulting in Eu metal starting to accumulate

Page 92: Copyright By Wei Guo 2021

73

once the scavenging depth is reached. This self-limiting behavior of EuO thickness is also

observed in the recent detailed study of the EuO oxygen scavenging process by Lomker

and Muller [40].

To facilitate the comparison with the experiment, we broaden the calculated 4f

DOS of each layer using Gaussian convolution. Based on the width of experimental XPS

spectra, we choose a FWHM parameter σ = 0.90 eV. Considering the inelastic scattering

in an XPS measurement, we attenuate the 4f DOS for each layer by a factor of 𝑒−𝑑

𝐿 where

d is the depth of the layer and L is the inelastic mean free path. The inelastic mean free

path of a 4f electron (kinetic energy of ∼1480 keV) is 23 and 31 Å for EuO and Eu,

respectively, as calculated using the TPP-2M formula [41]. We then add up the attenuated

and broadened DOS of each layer as the “calculated pseudo-XPS spectrum” and show it in

Figure 4.10(c). The 4f DOS of each EuO layer (magenta lines) clearly shows the band

bending. The peak position of the interfacial EuO 4f DOS is 2.7 eV and that of the bulk-

like EuO 4f DOS is at 2.3 eV. This leads to the overall broadening of the EuO 4f spectra.

For the Eu metal 4f state, the peak position is 2.1 or 0.4 eV higher compared to that of the

total EuO 4f DOS. This is close to the experimental measurement (0.3 eV), and the 0.1 eV

discrepancy can be ascribed to the EuO band-gap difference between the calculation and

experiment. As can be seen from the experimental XPS spectrum, the Eu metal component

is much more dominant, as EuO is buried and the corresponding signal has been largely

attenuated due to the longer travel distance of the photoelectrons. We observe that for the

structure of two Eu metal layers over 8.5 layers of EuO, the EuO-to-Eu signal intensity

ratio from the pseudo-XPS spectrum is ∼0.7.

On the other hand, it also influences the spacing between the total 4f peak and Eu

metal 4f peak. In experiment, the peak position of the total 4f spectrum is very close to that

Page 93: Copyright By Wei Guo 2021

74

of the Eu metal 4f spectrum, while in the calculation, the former is separated by 0.2 eV

from the latter. We show the spectrum of each EuO layer separately in Figure 4.10(d).

Following the violet line indicating the peak position of the EuO 4f contribution, the band

bending across the EuO layers can be clearly seen. Overall, accounting for the contributions

by both Eu metal and EuO, the total 4f peak spectral width increases to 1.1 eV. This is

comparable to the 0.8 eV width observed in experiment.

4.4 FURTHER RESEARCH ON FIELD-EFFECT OF 2DEG IN EUO/STO

Kormondy et al. have reported the spin-polarized two-dimensional t2g electron gas

at the EuO/STO interface [9]. Soft x-ray angle-resolved photoemission spectroscopy (SX-

ARPES) revealed the t2g nature of the carriers. They showed that this 2DEG displays a very

large positive linear magnetoresistance below the Curie temperature of EuO of 70K.

However, it would be desirable to shift this sheet charge into EuO, where the spin split

conduction band would ensure 100 % spin-polarization along with the increase of the Curie

temperature, possibly above liquid nitrogen temperature [42].

Ponath et al. demonstrated a ferroelectric field effect and carrier density modulation

in an underlying Ge, by switching the ferroelectric polarization of an epitaxial c-axis-

oriented BTO film grown on Ge (001) by MBE [43]. They verified the effect of the

polarization switching on the conductivity by microwave impedance microscopy (MIM).

Lee et al. performed a theoretical calculation of the LaAlO3/EuO interface and

demonstrated that electrostatic doping by an electric field in the polar oxide leads to a fully

spin-polarized 2DEG at the interface [42]. These results hint at the possibility that by

integrating an EuO/STO junction on top of ferroelectric BTO, we may gain control over

the interface charge distribution, thus achieving an unexplored way of modulating the

Page 94: Copyright By Wei Guo 2021

75

2DEG, and possibly shifting the Curie temperature to the range of practical applications.

This idea was theoretically demonstrated by Li et al. recently [44].

For further research of the 2DEG, we performed a preliminary experimental study

on the field-effect in the 2DEG at the EuO/STO interface. We built an EuO/STO/BTO/Si

heterostructure by MBE and used the SX-ARPES instrument at the Paul Scherrer Institute

(PSI) [45]. The BTO is 40 nm-thick and the STO is 5-10 nm-thick. EuO was deposited

with 1-3 nm thickness with a 1-2 nm Ti capping layer on the top to protect EuO from over-

oxidation during transportation.

ARPES experiments are performed at the SX-ARPES end station [45] of the

ADRESS beamline at the Swiss Light Source (PSI, Switzerland) [46]. The principle of SX-

ARPES is the same as XPS, using the photoelectric effect and getting the core level energy

information by analyzing photoelectron energy distributions. Additionally, SX-ARPES

introduces a precise scanning angle control of the sample enabling one to achieve

momentum space resolution with increased probing depth. By combining photoelectron

kinetic energy and momentum analysis, SX-ARPES can get a detailed energy band map of

the sample. For our measurement, circularly polarized x-rays were incident on the sample

at a grazing angle of ~20°. The sample was cooled down to 10-20 K to reduce the thermal

effects and decrease the coherent k-resolved spectral component at high photoexcitation

energies [47]. The combined (beamline and analyzer) energy resolution was ~100 meV.

X-ray absorption spectroscopy (XAS) was performed first to determine the absorption

peaks for key core levels like Ti 2p and Eu 3d. Core levels were scanned after XAS to

check the basic core level electronic structure information. After we determined the core

level peaks positions, resonant maps were scanned for Eu and Ti to find any potential

2DEG. The resonant map uses a range of photon energies (a core level energy range) with

Page 95: Copyright By Wei Guo 2021

76

continuous energy values to excite the sample. This measurement will have direct

photoemission around the Fermi edge just like usual XPS. If the incident photon has the

resonant energy with a core level, it will also excite one electron to the outer shell, which

will undergo Auger decay. This Auger electron will enhance the total intensity around the

Fermi edge. This allows one to detect very weak energy band structures. We show our

results to date in the following figures. ARPES instrument details can be found in Refs. 45,

46.

The VB scan is shown in Figure 4.11. EuO is very easy to over-oxidize to Eu3+

during transportation even with capping. We see only ~15% of Eu2+ is left compared with

Eu3+. The unreacted VB scan is shown Figure 4.4 where only Eu2+ 4f dominates the

spectrum. These Eu3+ signals will heavily cover the Eu2+ we expect to see.

Figure 4.11: Valence band scan of a EuO/STO/BTO/Si sample after transportation in

ambient.

Page 96: Copyright By Wei Guo 2021

77

Figure 4.12 shows the Eu resonant spectrum within the energy window of the Eu

3d core level. The strongest signal corresponds to Eu3+ at 1133 and 1137 eV. Eu2+ resonated

at photon energy=1131 eV which is weak compared with Eu3+. There is no resonant signal

at the Fermi edge (~0 eV) when we zoom in, meaning there is no hole gas at Fermi level

in EuO. We then scanned the Ti resonant map to see the Ti L-edge resonant signal.

Figure 4.12: Eu resonant map spectrum at the Eu 3d core level.

Figure 4.13 shows the Ti resonant map around the Fermi level. It clearly shows a

strong Eu2+ 4f signal. This signal is too strong and resonant along with the whole energy

window of Ti 2p. Due to the high Eu2+ 4f intensity, the details at the Fermi edge are

overwhelmed. In Fig. 4.14, we show a zoom-in figure of Figure 4.13 and adjust the contrast

to highlight the possible 2DEG signals, which is circled in red. There are two very weak

signals resonant at 465.5 and 460.2 eV, which correspond to Ti 2p peak positions, but the

Page 97: Copyright By Wei Guo 2021

78

intensity is just too low that we need to adjust to a very high contrast to see them. This

means that either the signal comes from a depth out of the scope of soft x-rays (4-5 nm),

or there isn’t too much 2DEG in the sample. There are three possible reasons for the poorly

revealed 2DEG signal: 1) Most of the EuO is over-oxidized and the heavy Eu3+ signal

covers the interface signals; 2) The use of a soft x-ray light source limits the measurement

depth to 4-5 nm from the surface, which is where the EuO/STO interface is in the current

structure; 3) We use a Ti capping layer and Ti signals from the capping layer heavily cover

the 2DEG signal, as the t2g carriers in 2DEG also comes from Ti. We tried to perform a

quick k-space scan, but it did not show any fine band structure at the Fermi edge.

Figure 4.13: Ti resonant map with strong Eu2+ 4f signal.

Page 98: Copyright By Wei Guo 2021

79

Figure 4.14: Zoom-in Ti resonant map with possible 2DEG signals (circled in red).

The main challenge we face is to protect the EuO and avoid over-oxidation during

transportation and keep the EuO/STO interface as close as possible to the surface to have

stronger signals from the interface. It requires a capping layer as thin as possible during the

measurement, but still capable of protecting EuO from oxidation. We have tried various

capping layers and found that a GeOx/Ge capping layer works well to protect EuO.

This new capping process uses a thin, amorphous Ge capping layer (~2 nm)

deposited on EuO at room temperature and the sample is then taken out of the vacuum

system. Oxygen in air will quickly oxidize the top half of the Ge and form GeOx of ~1 nm.

Low oxygen transparency of GeOx at room temperature will protect the deeper Ge,

maintaining its metallic state and preventing the over-oxidation of the EuO layer. The

Page 99: Copyright By Wei Guo 2021

80

advantage of GeOx is that it can be thermally desorbed by heating the sample to ~300-400

ºC in vacuum. This results in an interface very close to the top surface that should have

stronger interface signals. This method solves all three possible reasons we mentioned for

the previous weak signals. We show below XPS spectra for each step to demonstrate the

feasibility of this method.

Figure 4.15: Oxygen 1s core level comparison for 1) Ge/EuO/STO/BTO/Si sample before

it was taken out of vacuum (red); 2) The same sample in step 1) left in air

for over 2 days (green); 3) The same sample in step 2) after being heated in

vacuum at 350 ºC for 1 hour (blue). Oxygen from EuO and GeOx are

marked above the peaks.

Figure 4.15 shows the oxygen 1s core level for the same sample at different points

in the process. Step 1) is to cap EuO with amorphous Ge metal at room temperature.

Page 100: Copyright By Wei Guo 2021

81

Oxygen peak intensity is low because the Ge covers EuO. Step 2) is to take the sample out

of vacuum and let it rest for over 2 days in ambient air to oxidize the surface Ge and form

GeOx. The surface layer is a mixture of GeO and GeO2 and has two oxidation states. Thus,

the peak intensity at 529.5 and 531.2 eV both increased. Step 3) is to load the sample back

into the vacuum system and heat the sample at 350 ºC for 1 hour. These are practical

heating conditions to ensure that EuO does not interact with STO and that most of the GeOx

can be removed. The blue curve in Fog. 4.15 shows a very small fraction of GeOx and the

oxygen signal from EuO increases by 2-3 times. This confirms that about half of the Ge

capping has been removed by oxidation and vacuum heating, with the EuO being closer to

the surface. This is very important for SX-ARPES signal enhancement on our samples.

Page 101: Copyright By Wei Guo 2021

82

Figure 4.16: Valence band comparison for 1) Ge/EuO/STO/BTO/Si sample before it was

taken out of vacuum (red); 2) The same sample in step 1) after air exposure

for over 2 days (green); 3) The same sample in step 2) after heating in

vacuum at 350 ºC for 1 hour (blue). Eu signal positions are marked above

the peaks.

To confirm that our EuO is not over-oxidized, we show the VB scan for the same

sample in each steps. There is only Eu2+ signal in step 1) and the oxygen 2p peak is covered

by VB signal from the Ge capping. After step 2), the Eu2+ signal decreased significantly

and the Eu3+ signal emerged. Considering that oxygen must diffuse from the outside

through the top Ge layer, the EuO next to Ge capping layer will oxidize first. The interface

between EuO and STO remains in the 2+ state. After thermal treatment, theEu2+ signal

increases since the Ge capping is now thinner. This shows that while the thin Ge capping

layer is not 100% protective for all the Eu2+ state, it still prevents over-oxidation of about

half of the 2.5-nm-thick EuO layer. The EuO/STO interface region itself should be well-

protected by the thin Ge capping with this procedure.

We tested this procedure by international shipping of samples with Ge capping.

Under normal transportation conditions, samples are packed in gel pack box with N2 gas

flowing. At least 70% of Eu2+ signal remains after 5 days of international shipping, which

means our protection is feasible for EuO samples, especially thicker ones. This is a great

breakthrough since the protection of EuO has been an issue for EuO characterization and

applications in the past.

4.5 CONCLUSIONS

The oxygen scavenging process for Eu on STO substrates has been explored by

depositing pure Eu metal at different temperatures on different thicknesses of STO films

integrated on Si (001). At 300 °C, Eu takes oxygen from STO films thicker than 4 nm and

Page 102: Copyright By Wei Guo 2021

83

forms EuO, which is similar to what happens on bulk STO substrates [16]. However, when

the STO films are thin (in the range from 2 to 4 nm), Eu takes too much oxygen and

destabilizes the STO layer, allowing for a solid-state reaction with Si that results in the

formation of EuSiy and silicates.

On the other hand, below 150 °C, Eu can only take a very limited amount of oxygen

and forms up to a ∼20 Å thick layer of EuO. Beyond this point, Eu no longer takes oxygen

from STO and Eu metal accumulates on the surface, as evidenced by the core level signal

changes in photoemission. Once the metal and oxide are brought into contact, the Fermi

level of Eu lines up with the charge neutrality level of EuO arising from the spin-up

evanescent state. Consequently, charge transfer of the 5d electrons from the Eu metal ∼15

Å inside EuO induces a downward band bending near the interface. This downward band

bending is confirmed experimentally by analyzing the XPS spectra of the Eu 4f peaks and

their broadening at the Eu/EuO interface.

Exploring the field-effect of the EuO/STO interface provides a new possibility that

we may gain control over the interface charge distribution thus achieving a new way of

modulating the 2DEG, and possibly shifting the Curie temperature to the range of practical

applications. The ARPES measurement for such a system encountered technical problems

with extremely weak signals and we are trying to work around these issues.

The EuO/SrTiO3 is a promising platform for combining a ferromagnetic material

and a 2-dimensional electron system. The oxygen scavenging process of growing EuO on

STO has the advantage of not needing to control oxygen to form stoichiometric EuO under

certain conditions. Our study elucidates the behavior of this growth process under different

conditions and thus provides for better understanding and control of this system.

Page 103: Copyright By Wei Guo 2021

84

4.6 ACKNOWLEDGEMENT

This research was partially supported by the National Science Foundation (NSF)

through the Center for Dynamics and Control of Materials: an NSF MRSEC under

Cooperative Agreement No. DMR-1720595 and by the Air Force Office of Scientific

Research (AFOSR) under Grant No. FA9550-18-1-0053. All calculations were performed

at the Texas Advanced Computing Center (TACC) by Lingyuan Gao. I greatly thank Prof.

Strocov for his support when I visited PSI.

4.7 REFERENCES

1. P. G. Steeneken, L. H. Tjeng, I. Elfimov, G. A. Sawatzky, G. Ghiringhelli, N. B.

Brookes, and D.-J. Huang, Phys. Rev. Lett. 88, 47201 (2002).

2. A. Schmehl et al., Nat. Mater. 6, 882 (2007).

3. A. A. Melville, T. Mairoser, A. Schmehl, D. E. Shai, E. J. Monkman, J. W. Harter, T.

Heeg, B. Hollander, J. Schubert, K. M. Shen, J. Mannhart, and D. G. Schlom, Appl.

Phys. Lett. 100, 222101 (2012).

4. B. T. Matthias, R. M. Bozorth, and J. H. Van Vleck, Phys. Rev. Lett. 7, 160 (1961).

5. J. Schoenes and P. Wachter, Phys. Rev. B. 9, 3097 (1974).

6. C. Caspers, M. Müller, A. X. Gray, A. M. Kaiser, A. Gloskovskii, C. S. Fadley, W.

Drube, and C. M. Schneider, Phys. Rev. B. 84, 205217 (2011).

7. A. G. Swartz, P. M. Odenthal, Y. Hao, R. S. Ruoff, and R. K. Kawakami, ACS Nano

6, 10063 (2013).

8. D. V. Averyanov, Y. G. Sadofyev, A. M. Tokmachev, A. E. Primenko, I. A. Likhachev,

and V. G. Storchak, ACS Appl. Mater. Interfaces 7, 6146–6152 (2015).

9. K. J. Kormondy, L. Gao, X. Li, S. Lu, A. B. Posadas, S. Shen, M. Tsoi, M. R.

McCartney, D. J. Smith, J. Zhou, L. L. Lev, M.-A. Husanu, V. N. Strocov, and A. A.

Demkov, Sci. Rep. 8, 7721 (2018).

10. M. Matsubara, A. Schroer, A. Schmehl, A. Melville, C. Becher, M. Trujillo-Martinez,

D. G. Schlom, J. Mannhart, J. Kroha, and M. Fiebig, Nat. Commun. 6, 6724 (2015).

11. M. Matsubara, A. Schmehl, J. Mannhart, D. G. Schlom, and M. Fiebig, Phys. Rev. B

81, 214447 (2010).

12. P. Lömker, T. C. Rödel, T. Gerber, F. Fortuna, E. Frantzeskakis, P. LeFèvre, F. Bertran,

M. Müller, and A. F. Santander-Syro, Phys. Rev. Mater. 1, 062001(R) (2017).

Page 104: Copyright By Wei Guo 2021

85

13. L. Gao and A. A. Demkov, Phys. Rev. B 97, 125305 (2018).

14. A. B. Posadas, K. J. Kormondy, W. Guo, P. Ponath, J. Geler-Kremer, T. Hadamek, and

A. A. Demkov, J. Appl. Phys. 121, 105302 (2017).

15. K. J. Kormondy, A. B. Posadas, T. Q. Ngo, S. Lu, N. Goble, J. Jordan-Sweet, X. P. A.

Gao, D. J. Smith, M. R. McCartney, J. G. Ekerdt, and A. A. Demkov, J. Appl. Phys.

117, 95303 (2015).

16. G. M. Prinz, T. Gerber, A. Lorke, and M. Müller, Appl. Phys. Lett. 109, 202401 (2016).

17. T. Mairoser, J. A. Mundy, A. Melville, D. Hodash, P. Cueva, R. Held, A. Glavic, J.

Schubert, D. A. Muller, D. G. Schlom, and A. Schmehl, Nat. Commun. 6, 7716 (2015).

18. J. Lettieri, V. Vaithyanathan, S. K. Eah, J. Stephens, V. Sih, D. D. Awschalom, J. Levy,

and D. G. Schlom, Appl. Phys. Lett. 83, 975 (2003).

19. R. W. Ulbricht, A. Schmehl, T. Heeg, J. Schubert, and D. G. Schlom, Appl. Phys. Lett.

93, 102105 (2008).

20. D. V. Averyanov, Y. G. Sadofyev, A. M. Tokmachev, A. E. Primenko, I. A. Likhachev,

and V. G. Storchak, J. Appl. Phys. 113, 17C505 (2013).

21. K. J. Hubbard and D. G. Schlom, J. Mater. Res. 11, 2757 (1996).

22. Y. Wei, X. Hu, Y. Liang, D. C. Jordan, B. Craigo, R. Droopad, Z. Yu, A. A. Demkov,

J. John, L. Edwards, and W. J. Ooms, J. Vac. Sci. Technol. B 20, 1402 (2002).

23. J. Lettieri, J. H. Haeni, and D. G. Schlom, J. Vac. Sci. Technol. A 20, 1332 (2002).

24. H. Miyazaki, T. Ito, H. Im, K. Terashima, S. Yagi, M. Kato, K. Soda, and S. Kimura,

Jpn. J. Appl. Phys. 48, 055504 (2009).

25. H. Miyazaki, H. Momiyama, T. Hajiri, T. Ito, K. Imura, M. Matsunami, and S. Kimura,

J. Phys. Conf. Ser. 391, 012047 (2012).

26. B. I. Edmondson, S. Liu, S. Lu, H. W. Wu, A. B. Posadas, D. J. Smith, X. P. A. Gao,

A. A. Demkov, and J. G. Ekerdt, J. Appl. Phys. 124, 185303 (2018).

27. EuO Crystal Structure: Datasheet from “PAULING FILE Multinaries Edition—2012”

in Springer Materials,

see: http://materials.springer.com/ isp/crystallographic/docs/sd_0553841.

28. R. A. McKee, F. J. Walker, and M. F. Chisholm, Phys. Rev. Lett. 81, 3014 (1998).

29. M. Choi, A. Posadas, R. Dargis, C. Shih, A. A. Demkov, D. H. Triyoso, N. D.

Theodore, C. Dubourdieu, J. Bruley, and J. Jordan-Sweet, J. Appl. Phys. 111, 064112

(2012).

30. H. Li, X. Hu, Y. Wei, Z. Yu, X. Zhang, R. Droopad, A. A. Demkov, J. Edwards, Jr.,

K. Moore, and W. Ooms, J. Appl. Phys. 93, 4521 (2003).

Page 105: Copyright By Wei Guo 2021

86

31. K. D. Fredrickson, H. Seo, and A. A. Demkov, J. Appl. Phys. 120, 065301(2016).

32. T. Hadamek, D. Shin, A. B. Posadas, A. A. Demkov, S. Kwon, Q. Wang, and M. Kim,

Appl. Phys. Lett. 111, 142901 (2017).

33. S. P. KowaIczyk, N. Edelstein, F. R. McFeely, L. Ley, and D. A. Shirley, Chem. Phys.

Lett. 29, 491 (1974).

34. J. K. Lang, Y. Baer, and P. A. Cox, J. Phys. F. 11, 121 (1981).

35. L. Gao, W. Guo, A. B. Posadas, and A. A. Demkov. Phys. Rev. Materials 3, 094403

(2019).

36. X.-X. Wang, H.-N. Li, W.-H. Zhang, and F.-Q. Xu, J. Phys.: Condens. Matter 19,

096001 (2007).

37. M. W. Shafer, J. B. Torrance, and T. Penney, J. Phys. Chem. Solids 33, 2251 (1972).

38. A. Mauger and C. Godart, Phys. Rep. 141, 51 (1986).

39. P. R. Subramanian, Eu-O (europium-oxygen), in Binary Alloy Phase Diagrams, 2nd

ed., edited by T. B. Massalski (ASM International, Materials Park, OH, 1990), Vol. 2,

pp. 1665–1666.

40. P. Lomker and M. Muller, Phys. Rev. Mater. 3, 061401(R)(2019).

41. S. Tanuma, C. J. Powell, and D. R. Penn, Surf. Interface Anal. 36, 1 (2004).

42. J. K. Lee, N. Sai, and A.A. Demkov, Phys. Rev. B 82, 235305 (2010).

43. P. Ponath, K. Fredrickson, A. B. Posadas, Y. Ren, X. Wu, R. K. Vasudevan, M. B.

Okatan, S. Jesse, T. Aoki, M. R. McCartney, D. J. Smith, S. V. Kalinin, K. Lai, and A.

A. Demkov, Nature Comm. 6, 6067 (2015).

44. W. Li, L. Gao and A. A. Demkov, Phys. Rev. B 102, 035308 (2020).

45. V. N. Strocov, X. Wang, M. Shi, M. Kobayashi, J. Krempasky, C. Hess, T. Schmitt

and L. Patthey, J. Synchrotron Radiat. 21, 32(2014).

46. V. N. Strocov, T. Schmitt, U. Flechsig, T. Schmidt, A. Imhof, Q. Chen, J. Raabe, R.

Betemps, D. Zimoch, J. Krempasky, X. Wang, M. Grioni, A. Piazzalunga and L.

Patthey, J. Synchrotron Radiat. 17, 631(2010).

47. J. Braun, J. Minár, S. Mankovsky, V. N. Strocov, N. B. Brookes, L. Plucinski, C. M.

Schneider, C. S. Fadley, and H. Ebert, Phys. Rev. B 88, 205409(2013)

Page 106: Copyright By Wei Guo 2021

87

Chapter 5: Thermal oxidation of Si buried under thin SrTiO3 and

free-standing SrTiO3 membranes

In this chapter, we first present the dry oxidation of Si (001) beneath a thin epitaxial

SrTiO3 (STO) layer using furnace annealing in flowing oxygen. A 10-nm layer of STO is

epitaxially grown on Si with no SiO2 interlayer. For such a structure, an annealing

temperature of 800ºC was found to be the limiting temperature to prevent silicate formation

and disruption of the interface structure. The effect of annealing time on the thickness of

the SiO2 layer was investigated. In situ x-ray photoelectron spectroscopy (XPS) and

reflection-high-energy electron diffraction (RHEED) were used to ensure that the quality

of STO is unchanged after the annealing process. The experimental annealing data is

compared with a theoretical oxygen diffusion model based on one due to Deal, Grove and

Massoud. The model fits the experimental data well, indicating that oxygen diffusion

through the STO layer is not the limiting factor. One can therefore readily control the

thickness of the SiO2 interlayer by simply controlling the annealing time in flowing

oxygen. Using this ability to oxidize the underlying Si, we then show some preliminary

results on the fabrication of free-standing STO membranes, which requires thick SiO2

interlayers in order to properly detach the STO from the Si on which it is grown.

A portion of this chapter is published in: W. Guo, A. B. Posadas, A. A. Demkov, J.

Appl. Phys. 127, 055302 (2020). Our collaborators in Prof. Li’s group at UT Austin are

responsible for the Raman spectroscopy measurements.

5.1 INTRODUCTION TO DRY OXIDATION OF SI

STO is a widely used substrate for metal oxide thin film growth [1], [2]. It has a

rather large dielectric constant (~300) [2] making it attractive for dielectric applications.

The lattice constant of STO (3.9 Å) also makes it a suitable substrate material for the

Page 107: Copyright By Wei Guo 2021

88

epitaxtial growth of many oxides with perovskite, rocksalt, and spinel crystal structures

because of the small lattice mismatch to many such materials, such as LaAlO3 [2]-[4],

BaTiO3 [5]-[7], EuO [8], [9], γ-Al2O3 [10], etc. The discovery by Mckee et al. in 1998 that

one can epitaxially nucleate STO on Si directly without forming SiO2 has opened epitaxial

oxide thin films to the potential for technological development [11]. This process has been

further developed and studied both experimentally and theoretically by several groups [11]-

[14]. Even though STO has excellent dielectric properties making it attractive for use as a

gate oxide in field effect transistors [15], the conduction band offset at the STO/Si

interface is essentially zero, making such an application moot [16], [17]. However, STO

on Si can serve as a bridge material for integrating other oxides epitaxially on Si by serving

as a pseudo-substrate [6], [7], [9], [14]. It is therefore important to understand the behavior

of this system under various temperature and oxygen pressure environments, particularly

with respect to the oxidation of the underlying Si.

Silicon oxidation is a well-studied process that has been long-discussed following

the establishment by Deal and Grove of the oxygen diffusion model for the Si surface in

1965 [18]. This oxygen diffusion and reaction model of Si oxidation agrees well with the

experimental data at different temperatures for relatively thick SiO2 layers (>50 nm). The

Deal-Grove model predicts the resulting oxide thickness for a given temperature as a

function of time. However, significant deviations were found for thinner oxide films.

Massoud improved the Deal-Grove model for the thin SiO2 layer regime by adding an

exponential decay term to the original linear-parabolic model [19]-[21]. Owing to the good

dielectric properties of SiO2 and a well-behaved Si-SiO2 interface, controlled oxidation of

Si has played a fundamental role in the fabrication of Si devices such as metal-oxide-

semiconductor field-effect transistors (MOSFET) and single-electron devices [22]-[26].

Page 108: Copyright By Wei Guo 2021

89

With epitaxial STO covering the Si surface, the oxidation behavior of Si may not

be the same and it is important to know how the processing conditions, such as temperature

and oxygen pressure, affect the oxidation rate of silicon with an STO overlayer present. A

typical process of forming STO on Si involves first removing the surface SiO2 and then

depositing 0.5 monolayer of Sr in a molecular beam epitaxy (MBE) chamber. This

submonolayer Sr prevents Si oxidation during the initial nucleation of STO [12], [13]. It is

then followed by 4-5 unit cells (uc) of STO deposition at low temperature (<300°C) which

is fully crystallized at ~500°C. Once the initial template is crystallized, additional STO can

be grown, treating the template as an STO substrate. Continuing the growth at oxygen

pressures where STO is fully oxidized (~5×10-7 Torr) and temperatures where STO is

crystalline as deposited (~500°C) typically results in some oxygen diffusing through STO

and partially oxidizing the underlying Si. There have been several studies of Si oxidation

and interlayer reaction at the STO/Si interface [14], [15], [27]. It has been shown that STO

is not fully thermodynamically stable in direct contact with Si at very high temperature

under ultrahigh vacuum conditions [28]. Choi et al. grew STO/Si by MBE and used post-

deposition annealing in oxygen (10-7 to 10-5 Torr) to control the strain relaxation of STO

[14]. Cross-section transmission electron microscopy (TEM) was used to monitor the

interlayer structure and thickness under different oxygen partial pressures and annealing

times. Goncharova et al. discussed the thermal stability of the STO/Si interface [15]. They

considered the possible reactions that could happen between the layers at different

annealing temperatures such as STO + Si → SiO + SrO + O2 + TiSi2. A thin interlayer

composed of SrSiOx, SrO, and TiSix was found at annealing temperatures as low as 550°C.

At 850°C or even higher temperatures, the STO film decomposes completely, leaving

behind only TiSix islands [29]. Yong et al. have discussed the thermal stability and possible

Page 109: Copyright By Wei Guo 2021

90

interface reactions of the STO/SiO2/Si interface [27]. For example, Si + SrTiO3 → SrSiO3

+ TiSi2, and Si + TiO2 → SiO2 + TiSi2. Optical microscopy and scanning electron

microscopy (SEM) were used to observe the surface morphology changes of the STO film

after annealing at ~800°C. Eisenbeiser et al. used TEM to show the interfacial layer

between STO and Si after the growth [24].

Here, we study the dry oxidation behavior of the buried Si as a function of annealing

time, and determine the maximum temperature for which a relatively thin STO layer (10

nm) remains intact. The STO is found to be stable during the oxidation anneal at 800°C for

up to 10 hours. We start with the MBE growth of STO on Si (001) and then perform a

flowing oxygen anneal in a tube furnace to oxidize the Si underneath. The STO thickness

is fixed at 10 nm (25 unit cells (uc)) for this study. Too thin an STO layer (<10 uc) results

in STO and Si reacting. The practical annealing temperature is found to be 800°C. Below

this temperature, dry oxidation is very slow and impractical; above this temperature, the

STO in contact with Si is not thermally stable. We developed a theoretical model based on

a modification of the Deal-Grove-Massoud formalism that predicts the thickness of SiO2

formed underneath STO as a function of time and temperature, and report a robust recipe

for dry oxidation of Si buried under an epitaxial layer of STO.

5.2 SI OXIDATION EXPERIMENTS

The STO/Si growth is performed in a customized DCA 600 MBE system with a

base pressure of 6×10-10 Torr. P-type doped Si substrates of 20×20 mm2 size are cut from

a prime Si wafer and degreased ultrasonically in acetone, isopropanol, and deionized (18

MΩ/cm) water for 5 minutes each and then exposed to UV/ozone to remove carbon from

the surface. The Si substrates are then annealed in ultrahigh vacuum at 700°C for 10 min

Page 110: Copyright By Wei Guo 2021

91

followed by Sr-assisted de-oxidation before the growth of a 2-nm-thick STO epitaxial film

[11]-[14]. Prior to the growth of STO, 1/2 ML of Sr is formed on the Si (001) surface. This

Zintl layer has 2×1 symmetry and serves as a template for further deposition [11]. The

initially amorphous STO film is formed by co-deposition of Sr and Ti at low temperature

(200°C) under low oxygen pressure (8×10-8-5×10-7 Torr) and is annealed in vacuum at or

above 550°C to crystallize. This procedure results in no interfacial SiO2 as shown by x-ray

photoemission and transmission electron microscopy [12]. There are two possible ways of

growing additional STO. The first one is to deposit additional amorphous STO near room

temperature in oxygen and anneal it in vacuum to the STO crystallization temperature

(550°C). This will not result in formation of the SiO2 interlayer. Another way is to perform

a co-deposition of Sr and Ti under modest oxygen pressure (~mid 10-7 Torr) at 550°C [14]

(as if growing on an STO substrate). Because of the high oxygen diffusivity in STO,

oxygen can diffuse through and oxidize Si underneath without disrupting the crystal

structure of the already crystallized STO. This results in a very thin SiO2 layer (~2 nm)

between STO and Si. In this study, we use the first method to prepare a SiO2-free interface

prior to annealing. Reflection-high-energy electron diffraction (RHEED) with 21 keV

electrons is used to record the STO crystallinity during growth. In situ x-ray photoelectron

spectroscopy (XPS) using monochromated Al Kα radiation and a VG Scienta R3000

hemispherical electron energy analyzer is used to check stoichiometry of STO. After

annealing, the STO/SiO2/Si samples are then measured using J.A. Woollam M-2000DI

spectroscopic ellipsometer to determine the thickness of SiO2. The instrument uses a

combination deuterium / quartz tungsten halogen lamp as the light source and covers a

wavelength range from 190-1650 nm. The data is collected at three different angles of

incidence (45°, 50°, and 55°). The fit is performed using the built-in software,

Page 111: Copyright By Wei Guo 2021

92

CompleteEase, with three layers (Si substrate, thermal oxide SiO2, and bulk STO) over the

entire wavelength range of the instrument. For the bulk STO optical constants, we use the

parametrized optical constants determined by the Zollner group at New Mexico State

University [30]. All three angles of incidence are fit simultaneously using N, C, S fit

weighting. We show an example of a fit of the ellipsometry data in Figure 5.1. The

measured data are shown as circles and the fit curves as solid lines. The red one is the

curve and the blue one is the curve.

Figure 5.1 Example ellipsometry measurement with the corresponding model fit. These

are the measurement results at 45°. This gives a STO layer with 17.54 nm

and SiO2 layer with 7.54 nm thickness. The red one is the curve and the

blue one is the curve. Measured data are shown as open circles and the fit

curves as solid lines.

Page 112: Copyright By Wei Guo 2021

93

After a 10 nm-thick STO layer is grown in the MBE system (RHEED is shown in

Figure 5.2(a)), the sample is taken out of UHV and ultrasonically cleaned by IPA and DI-

water and inserted into a GSL-1700X-S tube furnace from MTI Corporation. Dry oxygen

flows through the furnace tube during the entire annealing cycle with approximately 1

L/min flow rate at atmospheric pressure. We use dry oxidation to avoid a possible reaction

between STO and water [31]. STO/Si samples are placed in the center of the furnace tube

with the STO side facing up. The tube is heated gradually from room temperature to the

final temperature at 10°C/min. The sample is kept at this temperature for the designated

period of time after which the sample is then cooled to room temperature over a period of

two hours. We verified that the STO film still has good crystallinity after the anneal by

RHEED, which is shown in Figure 5.2(b).

Page 113: Copyright By Wei Guo 2021

94

Figure 5.2: (a) RHEED image for a 10 nm STO/SiO2/Si after the growth and before

anneal. STO pattern is shown along the [110] direction. (b) RHEED image

for a 10 nm STO/SiO2/Si after 800°C anneal for 2 hours. STO pattern along

the [110] direction is still sharp and clear.

To check the interface composition, we grew a relatively thin STO/Si sample with

5 nm of STO and checked it with XPS. The Si 2p spectrum after the 800°C anneal is shown

in Figure 5.3. It shows a major SiO2 peak and a minor SiOx peak, with no TiSi2 signal

visible.

Page 114: Copyright By Wei Guo 2021

95

Figure 5.3: XPS spectrum for the Si 2p region. The major peak around 104.5 eV is SiO2

and the minor peak around 103 eV is SiOx, Si metal (99.5 eV) is also still

visible.

It is known that it takes 15 hours to grow a 10 nm-thick SiO2 layer on bare Si [32]

at 700°C. The SiO2 layer will grow even more slowly at lower temperature (<700°C)

because of the reduced oxygen diffusion rate. Most Si oxidation data in the literature are

from the practically useful 800°C-1300°C temperature range. However, there are also

papers reporting degradation of the STO/Si structure at high temperature (~1000°C) when

STO reacts with SiO2 and Si [15], [29]. Therefore, the practical temperature range to have

an acceptable oxidation rate without destroying STO is ~700-900°C. The annealing

Page 115: Copyright By Wei Guo 2021

96

temperature we use is 800°C, which on the one hand is not high enough to degrade the

STO layer but on the other hand, allows the Si oxidation time to remain practical. Under

this annealing temperature, we compared the out-of-plane XRD full scan spectra before

and after annealing for 4 hours (Figure 5.4). We find that the STO film only shows a little

degradation, with the FWHM of rocking curve of the STO (002) peak becoming slightly

broadened from 0.33° to 0.54°.

Figure 5.4: Comparison of the XRD out-of-plane full scans before and after annealing.

STO peaks decrease a little and Si is a little higher. All peaks do not show

any obvious deformation other than intensity variation.

Page 116: Copyright By Wei Guo 2021

97

5.3 THEORETICAL MODEL OF SI OXIDATION

The standard model of Si oxidation has been proposed almost half a century ago by

Deal and Grove [18]. It is based on describing three steps that result in oxidation: (i) oxygen

(or oxidant species in general) transport from the gas phase to the oxide surface where it is

adsorbed, (ii) oxygen transport through the oxide layer towards Si, and (iii) the interface

oxygen reaction with Si and formation of a new layer of SiO2. However, the model has a

well-known difficulty in predicting the initial stage of oxidation for thin films. Massoud et

al. made modifications to the Deal-Grove model to overcome this issue [19]-[21]. In the

case where an STO overlayer is present, we have to modify the model further as the initial

oxide in our case is not SiO2 but STO, and at later stages oxygen has to diffuse through

both materials in order to reach Si.

Following the Deal-Grove-Massoud logic, we set the oxygen concentration in

different regions and connect them by the oxygen diffusion flux. In Figure 5.5, we show a

schematic of the oxygen propagation through the structure. Here C* is the oxygen

concentration of the gas, C0 is the oxygen concentration at the surface of STO, C1 is the

oxygen concentration at the STO/SiO2 interface, and C2 is the oxygen concentration at the

SiO2/Si interface. X1 and x2 are the thickness of STO and SiO2, respectively. F1 is the

incoming oxygen flux, F2 is the flux inside STO, F3 is the flux inside SiO2, and F4 is the

oxidation rate at the Si interface with SiO2. The steady-state condition is discussed in the

original paper [18].

Page 117: Copyright By Wei Guo 2021

98

Figure 5.5: The schematic of oxygen propagation in the SrTiO3/SiO2/Si structure. The

inset shows the atomic structure schematic.

In the original Si oxidation model [18], one has:

𝐹1 = ℎ(𝐶∗ − 𝐶0) (1)

where h is the transport coefficient (from gas into STO). The diffusion flux

𝐹2 = −𝐷1𝑑𝐶

𝑑𝑥 (2)

Page 118: Copyright By Wei Guo 2021

99

is given by Fick’s Law, and because the gradient is linear, we have:

𝐹2 = 𝐷1(𝐶0−𝐶1)

𝑥1, 𝐹3 = 𝐷2

(𝐶1−𝐶2)

𝑥2, 𝐹4 = 𝑘𝐶2, (3)

where D1 is the diffusivity of oxygen in STO, D2 is the diffusivity of oxygen in

SiO2, and k is the reaction rate of Si oxidation.

Since all fluxes are uniform, we set 𝐹1 = 𝐹2 = 𝐹3 = 𝐹4. Solving this system of

equations, we obtain:

𝐶0

𝐶∗ =1+

𝑘𝑥1𝐷1

+𝑘𝑥2𝐷2

1+𝑘

ℎ+

𝑘𝑥1𝐷1

+𝑘𝑥2𝐷2

, 𝐶1

𝐶∗ =1+

𝑘𝑥2𝐷2

1+𝑘

ℎ+

𝑘𝑥1𝐷1

+𝑘𝑥2𝐷2

, 𝐶2

𝐶∗ =1

1+𝑘

ℎ+

𝑘𝑥1𝐷1

+𝑘𝑥2𝐷2

, (4)

The diffusivities of oxygen in STO and silica are compared in Figure 5.6. We

combine high temperature (>700°C) diffusivity data of STO [33] from literature and

extrapolate to lower temperatures (300-700°C) to find estimated diffusivities at those

temperatures. One can see that in the temperature range from 300°C to 1200°C, oxygen

diffusivity in STO, D1, is always at least three orders of magnitude larger than that in silica,

D2. Thus one can neglect the contribution coming from the term 𝑘𝑥1

𝐷1 in the following

discussion. Therefore:

𝑑𝑥2

𝑑𝑡=

𝐹

𝑁1=

𝑘𝐶∗

𝑁1⁄

1+𝑘

ℎ+

𝑘𝑥2𝐷2

=𝐵

𝐴+2𝑥2 , (5)

N1 here is the oxygen needed to oxidize a unit volume of Si, A and B are defined

as:

Page 119: Copyright By Wei Guo 2021

100

𝐴 = 2𝐷2 (1

𝑘+

1

ℎ) , 𝐵 = 2

𝐶∗𝐷2

𝑁1. (6)

Figure 5.6: Diffusivity of oxygen in SrTiO3 (D1) and SiO2 (D2) in the 300-1200°C

temperature range. Inset: Diffusivity of oxygen in SrTiO3 and SiO2 in the

300-700°C temperature range. The STO low temperature (<700°C) data is

obtained by inverse relationship projection from existing diffusivity at

higher temperature (>700°C)

After the modifications of the Massoud model [19], [20], we obtain:

𝑑𝑥0

𝑑𝑡=

𝐵

𝐴+2𝑥0+ 𝐶1𝑒−𝑥0/𝐿1 + 𝐶2𝑒−𝑥0/𝐿2 (7)

Page 120: Copyright By Wei Guo 2021

101

where A, B, Ci, and Li, can be fitted from the experimental values and are listed in

Massoud’s paper [19], [20].

The steady-state picture can be justified as follows. Consider a sudden change in

oxygen concentration in the silica layer, the total time it takes for the concentration to come

back to a stable value is ∆t:

∆𝑡 =total oxygen needed to come back to original concentration

oxygen flux (8)

Assuming that the total amount of oxygen needed is the same as the amount of

oxygen that must flow out of the layer, the numerator is simply 1

2(𝐶1 − 𝐶2)𝑥2, and the

oxygen flux is 𝐹 = 𝐷2(𝐶1−𝐶2)

𝑥2. Therefore,

∆𝑡 =1

2(𝐶1−𝐶2)𝑥2

𝐷2(𝐶1−𝐶2)

𝑥2

=𝑥2

2

2𝐷2. (9)

Here x2 is on the order of nm and D is on the order of ~108 nm2/hr, so ∆t is of the

order of 10-8 hour. Thus we can assume that the flow of oxygen through the layer is

established very quickly and can be assumed to be steady for all practical experimental

conditions.

This suggests that we can treat the problem as oxidation of bare Si since STO is

essentially transparent to oxygen diffusion compared to SiO2. Using the Massoud model

[20], [21], we have

𝑑𝑥0

𝑑𝑡=

𝐵

𝐴+2𝑥0+ 𝐶1𝑒−𝑥0/𝐿1 + 𝐶2𝑒−𝑥0/𝐿2 . (10)

Comparing with the experimental data, the formula can be re-written as:

Page 121: Copyright By Wei Guo 2021

102

𝑑𝑥0

𝑑𝑡=

𝐵+𝐾1𝑒−𝑡/𝜏1+𝐾2𝑒−𝑡/𝜏2

𝐴+2𝑥0 (11)

Solving this differential equation, we obtain:

𝑥0 =𝐴

2+ √(

𝐴

2)

2

+ 𝐵𝑡 + 𝑀1 [1 − exp (−𝑡

𝜏1)] + 𝑀2[1 − exp (−

𝑡

𝜏2) + 𝑀0

𝑀0 = 𝑥𝑖2 + 𝐴𝑥𝑖 , 𝑀1 = 𝐾1𝜏1, 𝑀2 = 𝐾2𝜏2, (12)

where xi is the initial thickness of SiO2, with the parameters given by:

𝐵 = 𝐶𝐵exp (−𝐸𝐵

𝑘𝑇),

𝐵

𝐴= 𝐶𝐵/𝐴exp (−

𝐸𝐵/𝐴

𝑘𝑇)

𝐾1 = 𝐾10 exp (−

𝐸𝐾1

𝑘𝑇), 𝐾2 = 𝐾2

0 exp (−𝐸𝐾2

𝑘𝑇)

𝜏1 = 𝜏10exp (

𝐸𝜏1

𝑘𝑇), 𝜏2 = 𝜏2

0exp (𝐸𝜏2

𝑘𝑇) (13)

The constants are taken from Massoud [21] and are listed in Table 2 for

convenience.

Page 122: Copyright By Wei Guo 2021

103

Crystal Orientation (100) (111) (110)

CB [nm2/min] 1.70×1011 1.34×109 3.73×108

EB [eV] 2.22 1.71 1.63

CB/A [nm/min] 7.35×106 1.32×107 4.73×108

EB/A [eV] 1.76 1.74 2.10

𝐾10 [nm2/min] 2.49×1011 2.70×1019 4.07×108

EK1 [eV] 2.18 1.74 1.54

𝐾20 [nm2/min] 3.72×1011 1.33×109 1.20×108

EK2 [eV] 2.28 1.76 1.56

𝜏10 [min] 4.14×10-6 1.72×10-6 5.38×10-9

Eτ1 [eV] 1.38 1.45 2.02

𝜏20 [min] 2.71×10-7 1.56×10-7 1.63×10-8

Eτ2 [eV] 1.88 1.90 2.12

Table 2. Oxidation parameters of the Massoud model [21] for temperature less than

1000°C.

Based on these parameters in (12) and (13) we can estimate the silica thickness (x0)

for a given set of annealing time and temperature. We plot the silica thickness as a function

of oxidation time at different temperatures in Figure 5.7. The red diamond shapes in Figure

5.7 are the experimental values from the tube furnace oxidation at 800°C measured by

ellipsometry. We also performed experiments at 750°C and 700°C to verify the prediction

of the model for other temperatures. We want to emphasize that the model described above

is derived based on the Deal-Grove-Massoud model using established parameters

appropriate for Si/SiO2 with no adjustable parameters. We use these established parameters

to compare the model to the measured interfacial SiO2 thicknesses and find reasonable

Page 123: Copyright By Wei Guo 2021

104

agreement between the model predictions and the data, especially at 800°C, with somewhat

worse fits for the lower temperatures measured.

Figure 5.7: The SiO2 thickness as a function of oxidation time at different temperatures

from equations (12) and (13). The lines represent the model and the shapes

represent the experimental data. Diamonds are samples annealed under

800°C, circle is under 750°C and square is under 700°C. All thickness data

have an error bar of ~1 nm.

The uncertainty in the SiO2 thickness comes from the particulars of the fitting of

the ellipsometry data and the fluctuations present in the annealing process, particularly the

oxygen flow rate. The fluctuations of the oxygen flow rate were found to cause drifts in

the surface temperature of the sample resulting in oxidation thickness variations. We

performed three anneals with the same condition and measured the SiO2 thickness of each.

Page 124: Copyright By Wei Guo 2021

105

The samples have the same structure before annealing and we controlled the oxygen flux

at the same level. From this, we determine the random error in the SiO2 thickness under

800°C from the oxidation process to be 0.8 nm. Additionally, there is uncertainty due to

the particulars of the ellipsometry fitting of up to 1 nm due to different wavelength range

of fitting and different fit weighting. By choosing proper parameters, the error from the

data fitting can be reduced to 0.1~0.2 nm. From this, we set the error bar in our plot to be

1 nm. Lower temperatures (under 800 °C) and longer anneal times (> 1 hour) would be

expected to result in slightly larger errors as a result of the accumulation of uncertainty

from the temperature variation due to the fluctuation in oxygen flow rate. This could

explain why the model works well for 800°C and not so well for the lower temperatures.

At 800°C, the oxidation rate is not prohibitively slow, while the temperature would

not cause the STO and Si to react and form SrSiO3 or TiSi2 as has been previously reported

[15], [29], [30]. The data agrees reasonably well with the model, indicating that one can

estimate the dry oxidation time needed to obtain a desired thickness of SiO2 while

maintaining a thin single crystal STO film on top. We have explored temperatures higher

than 800ºC in the range of 900-1200ºC. Unfortunately, at these higher temperatures the

STO layer breaks up and reacts with the Si substrate. However, dry oxidation at 800°C is

a robust method to produce interfacial SiO2 layers in the thickness range of 10-50 nm. We

have also explored other STO thickness (5, 20, 40 nm) and have found that the STO film

thickness does not matter if the STO thickness remains under 100 nm. This is confirmed

by the model, which shows that STO thickness is not important over a wide range of values.

Page 125: Copyright By Wei Guo 2021

106

5.4 FABRICATION OF FREE-STANDING STO MEMBRANES

Based on the ability to control the SiO2 interlayer of STO/Si samples, we are able

to make free-standing STO membranes by etching away the SiO2 layer, which works if the

SiO2 layer is thick enough.

Two-dimensional oxide materials exhibit physical properties that are emergent or

different from the bulk. This opens new possibilities for applications. STO is a widely-used

substrate in thin film growth, making it a very good bridge material for epitaxial growth of

various oxides like LaAlO3, BaTiO3, etc. Here, we are interested in any potential new

physics phenomena that may arise in these quasi-2-dimentional STO membranes.

Recently, Hwang’s group [34], [35] successfully fabricated free-standing STO

membranes via liftoff using a water-soluble template. They used PLD to grow Sr3Al2O6

(SAO) templates, and showed that this material can dissolve in water within seconds. A

group from Nanjing University managed to use the same method to produce monolayer-

thick free-standing STO membranes with mm lateral size [36].

The SAO water-soluble method is very attractive and have attempted to grow it in

our MBE system. While we were able to grow SAO compound on STO substrate and

confirmed that it is water-soluble, we have not been able to obtain a sufficiently good

crystalline SAO layer for epitaxial STO to grow on it. We instead developed another liftoff

process as follows. We first grow ~20 nm thick STO on Si substrate and then use oxygen

annealing to create a 2 - 10 nm-thick SiO2 layer in between STO and Si. We use a diamond

scriber to make scratches on the samples and use 4M NaOH solution to etch away the SiO2

and Si around the scratch. We perform the etching on a heating plate with a magnetic

spinner. Warming up solution and using a spinner to agitate it will speed up the etching

process and minimize potential damage of the solution to STO. This process is indeed able

Page 126: Copyright By Wei Guo 2021

107

to yield free-standing STO membranes along the scratch edge. We show the schematic of

the procedure in Figure 5.8.

Figure 5.8: Schematic of the procedure to fabricate STO membranes

As Figure 5.8 shows, STO membranes are hanging along the edges of the scribed

areas. We optimize STO and SiO2 thickness to achieve better etch results. Heating

temperature and spinning speed are also optimized. We managed to get STO membranes

with sizes of 5-8 m wide and 10-20 m in length. We show the SEM picture of the

membranes in Figure 5.9. We can see that the membranes are hanging or curled back along

the STO “cliff” from the scratched area. We use yellow lines to mark a membrane with 15

× 15 μm2 in Figure 5.9(b). This size level is the largest we have made using our fabrication

procedure. The typical size we obtain for 10-40 nm-thick STO membranes is ~5×10 µm2.

Page 127: Copyright By Wei Guo 2021

108

Figure 5.9: (a) (b) are SEM images from different spots on STO/Si etched sample. The

biggest membrane outlined by yellow is ~15 × 15 μm2

5.5 MEASUREMENTS OF STO MEMBRANES

STO membranes are expected to have somewhat different phonon vibrational

modes when compared with STO bulk and thin films since they have different boundary

Page 128: Copyright By Wei Guo 2021

109

conditions. We use Raman spectroscopy to measure these phonon modes and try to

understand the new vibrational behavior of the membranes.

Raman spectroscopy uses the inelastic scattering of photons when a material is

irradiated by an intense a beam of light. Photons will interact with molecular vibrations,

phonons or other excitations in the material and shift them to higher or lower energies. The

shifted energies will provide information on vibration modes in the material. We use a

Micro-Raman instrument to perform Raman spectroscopy on our STO membranes and

compare the results with bulk and thin film STO Raman spectra. Prof. Li’s group in UT-

Austin performed the measurements shown below. Here, I present only some preliminary

results. There is a lot more research that can done in the future in this area.

First, we performed Raman on a bulk STO substrate to calibrate the system since

this spectrum is well-known and studied [37]-[42]. Bulk STO crystal is centrosymmetric

at room temperature. The degrees of freedom consist of one Flu triply degenerate acoustic

mode, three Flu and one F2u triply degenerate optical modes [38]. Flu and F2u are not Raman

active, and no first-order Raman peaks should exist in defect-free STO single crystal [38].

However, impurities and defects in bulk crystal will still bring first-order Raman-active

modes in the spectrum [37], [41]. We show the data for bulk STO in Figure 5.10. Every

characteristic peak is labeled with position values and marked by red arrows. Here, the 175

cm-1 peak corresponds to the first order TO2 mode [37], [41] with a clear Fano asymmetry.

The rest of the peaks are assigned to possible second order modes as shown in Table 3 [38],

[39]. The bulk STO Raman spectrum measured in our system is consistent with data

published in the literature.

Page 129: Copyright By Wei Guo 2021

110

Figure 5.10: Raman spectrum of a bulk STO substrate. Characteristic peak positions are

labeled and marked with red arrows.

Energy shift (cm-1) Assignment

250 2TA

2TO1

TO1+TA

315 TO2+TA

TO2+TO1

TO4-TO2

2LA

Page 130: Copyright By Wei Guo 2021

111

362 2TO2

TO4-TA

TO4-TO1

LO1+TA

620 TO4+TA

TO4+TO1

LO2+TO3

675 2TO3

TO1+LO3

TO4+LA

TA+LO3

712 TO4+TO2

Table 3. Phonon branch assignments for second-order Raman peaks measured from STO

bulk substrate.

We then performed a Raman measurement on a 20 nm as-grown STO thin film on

Si substrate. STO thin films were studied in the early 2000s and compared with the bulk

STO spectrum [37], [40]. Forbidden zero-center optical phonons are clearly observed in

STO thin films. This was attributed to the interaction of the polar mode with polarization

fluctuations in micropolar regions [37]. We produced and measured an MBE-grown STO

thin film sample on Si as a comparison. The results are shown in Figure 5.11 with all peaks

labeled by their energy positions. The sharp, saturated signal at 521 cm-1 is the only Si-

related peak.

Page 131: Copyright By Wei Guo 2021

112

Figure 5.11: Raman spectrum of 20 nm STO thin film on Si. (Thin film grown by MBE)

First order peaks appear at 163, 265, 482, 551, 792 cm-1 for TO2, TO3, LO2, TO4,

LO4, respectively [37], [43]. There is a temperature dependence for first order Raman peaks

and lower temperatures will give sharper line shapes [37], [43]. Our measurement is

performed at room temperature and our first order peaks are not as well-expressed as they

are at low temperature. The main characteristic peak is at 302 cm-1 and can be assigned to

either 2LA or TA+TO2. This is the only second order peak in a thin film that still has strong

intensity. Other second order peaks that appear in the bulk STO Raman spectrum have

decreased to small peaks. As shown in Table 3 above, the 367 cm-1 peak corresponds to

2TO2 and other assignments at the same position. The 617 cm-1 peak corresponds to

TO4+TA and other assignments at the same position. The 670 cm-1 peak corresponds to

Page 132: Copyright By Wei Guo 2021

113

2TO3 and other assignments at the same position. The 712 cm-1 peak corresponds to

TO4+TO2. The 282 cm-1 peak corresponds to LO3-LO1 [39], which is not present in the

bulk STO Raman spectrum. The closest assignment for the 198 cm-1 peak is 2TA [39],

which represents a shift of 20 cm-1. The closest assignment for the 433 cm-1 peak is

TO3+TA or LO2+LA, but this peak is normally present only in measurements at 4 K on

bulk STO [39]. Peaks at 763, 866 cm-1 don’t have proper first or second order assignments.

The main reasons for different behaviors between bulk and thin film STO are the

stress, lattice mismatches or vacancies [37], [40]. TO phonon peaks in the thin film

measured spectrum indicate a lowering of the crystal symmetry in the films, inversion

breaking, and/or translational symmetries [37]. With different boundary conditions and

defect distribution, we are naturally curious what a quasi-2-dimensions free-standing

membrane with will have in its Raman spectrum. Using the process described in the last

section, we managed to make sufficiently large STO free-standing membranes for Raman

measurement. By focusing the laser on the free-standing STO film, we are able to get a

spectrum with a weak STO signal, as shown in Figure 5.12. The low intensity is due to the

small thickness of membranes (20 nm). Better optimization of the equipment can also

improve the total intensity and lower the noise level in the future.

Page 133: Copyright By Wei Guo 2021

114

Figure 5.12: Raman signal of STO membranes made from 20 nm STO/Si

Here, the 521 cm-1 peak is the Si peak. The major peak at 302 cm-1 in supported

thin film STO almost disappears (only a very tiny peak remains). The STO membrane

Raman spectrum is dominated by a 280 cm-1 peak corresponding to LO3-LO1 [39]

(identical peak with STO thin film Raman). First order peaks show up at 265, 481, 791 cm-

1 for TO3, LO2, LO4, respectively [37], [43]. The TO4 mode disappears in the membrane

spectrum compared with the supported thin film STO Raman spectrum. For second order

peaks as shown in Table 3 [37], [38], the 240 cm-1 peak corresponds to 2TA and other

assignments at the same position; the 367 cm-1 peak corresponds to 2TO2 and other

assignments at the same position; the 604 cm-1 peak corresponds to TO3+LO1 [38]; the 617

cm-1 peak corresponds to TO4+TA and other assignments at the same position.

Page 134: Copyright By Wei Guo 2021

115

Interestingly, the 240 cm-1 peak disappears in supported thin film spectrum but is present

in the membrane spectrum. There is a new, unidentified peak at 571 cm-1 in the membrane

spectrum with relatively high intensity and have not found any assignments for it. The 866

cm-1 peak appears in both STO thin film and membrane spectra but there isn’t a proper

assignment for this peak, either. These need further experiments to reveal the nature of

these modes.

The STO membrane measurement has weak signals with a high noise level. The

range of energy shifts observed is also limited by the optical setup to 150-900 cm-1. Further

optimization of the measurement process is needed to obtain a better and wider spectrum.

5.6 CONCLUSIONS

In this chapter, we first presented dry oxidation of STO/Si heterostructures and

demonstrated that the underlying Si can be safely oxidized at a relatively high temperature

(800ºC) with the STO crystallinity not significantly degraded. We deposited 10-nm of

epitaxial STO on Si and performed flowing oxygen anneals at 800ºC. The SiO2 thickness

is measured by ellipsometry and compared with our Deal-Grove-like oxidation model and

found good agreement between the data and the model. We can use this model to predict

the temperature and time needed to obtain the desired SiO2 thickness for Si that is covered

by a thin layer of STO. This additional knob for controlling the layer structure can enable

one to integrate more complicated oxide structures on this STO/Si pseudo-substrate,

especially for applications requiring complete decoupling between the STO and Si.

The study of Si oxidation under STO provides a precise controllable method of

SiO2 formation in STO/Si and we managed to make free-standing STO membranes based

on that capability. We used NaOH etch to make free-standing STO membranes with typical

Page 135: Copyright By Wei Guo 2021

116

size of ~5×10 µm2. Raman spectroscopy was performed on STO thin film and membranes.

Preliminary results show weak signals from STO membranes. In the future we will work

on the optimization of the Raman setup to reduce noise and amplify the STO signal. We

can also use different thicknesses of STO to study the phonon modes of a confined quasi-

2D STO membrane. With the high-k features of STO, we can also try to design new

transistor devices using our STO membranes. For example, we have come up with an

architecture replacing graphene or hBN membranes by STO with a higher dielectric

constant to build tunnel field-effect transistors [43], [45].

5.7 ACKNOWLEDGEMENTS

This research was partially supported by the National Science Foundation through

the Center for Dynamics and Control of Materials: an NSF MRSEC under Cooperative

Agreement No. DMR-1720595 and by the Air Force Office of Scientific Research under

Grant FA9550-18-1-0053. We thank Prof. Li’s group for performing the Raman

measurements.

5.8 REFERENCES

1. B. Reihl, J. G. Bednorz, K. A. Műller, Y. Jugnet, G. Landgren, and J. F. Morar, Phys.

Rev. B 30, 803 (1984).

2. A. A. Demkov and A. B. Posadas, Integration of Functional Oxides with

Semiconductors (Springer, New York, 2014).

3. J. E. Ortmann, N. Nookala, Q. He, L. Gao, C. Lin, A. B. Posadas, A. Y. Borisevich, M.

A. Belkin, and A. A. Demkov, ACS NANO 12, 7682 (2018).

4. J. E. Ortmann, A. B. Posadas and A. A. Demkov, J. Appl. Phys. 124, 015301 (2018).

5. P. Ponath, A. B. Posadas, M. Schmidt, A.-M. Kelleher, M. White, D. O’Connell, P.

Hurley, R. Duffy and A. A. Demkov, J. Vac. Sci. and Technol. B 36 (2018).

6. K. D. Fredrickson, V. V. Vogler-Neuling, K. J. Kormondy, D. Caimi, F. Eltes, M.

Sousa, J. Fompeyrine, S. Abel, and A. A. Demkov, Phys. Rev. B 98, 075136 (2018).

Page 136: Copyright By Wei Guo 2021

117

7. K. J. Kormondy, Y. Popoff, M. Sousa, F. Eltes, D. Caimi, M. D. Rossell, M. Fiebig, P.

Hoffmann, C. Marchiori, M. Reinke, M. Trassin, A. A. Demkov, J. Fompeyrine, and

S. Abel, Nanotechnology 28, 075706 (2017).

8. K. J. Kormondy, L. Gao, X. Li, S. Lu, A. B. Posadas, S. Shen, M. Tsoi, M. R.

McCartney, D. J. Smith, J. Zhou, L. L. Lev, M. Husanu, V. N. Strocov, and A. A.

Demkov, Scientific Reports 8, 7721 (2018).

9. W. Guo, A. B. Posadas, Sirong Lu, David J. Smith, and Alexander A. Demkov, J. Appl.

Phys. 124, 235301 (2018).

10. K. Kormondy, A. Posadas, T. Q. Ngo, S. Lu, J. L. Jordan-Sweet, X. Guo, D. J. Smith,

M. R. McCartney, J. G. Ekerdt, and A. A. Demkov, J. Appl. Phys. 117, 095303 (2015).

11. R. A. McKee, F. J. Walker, and M. F. Chisholm, Phys. Rev. Lett. 81, 3014 (1998).

12. H. Li, X. Hu, Y. Wei, Z. Yu, X. Zhang, R. Droopad, A. A. Demkov, J. Edwards, Jr.,

K. Moore, and W. Ooms, J. Appl. Phys. 93, 4521 (2003).

13. Y. Wei, X. Hu, Y. Liang, D. C. Jordan, B. Craigo, R. Droopad, Z. Yu, A. Demkov, J.

John, L. Edwards, and W. J. Ooms, J. Vac. Sci. Technol. B 20, 1402 (2002).

14. M. Choi, A. B. Posadas, R. Dargis, C. K. Shih and A. A. Demkov, D. H. Triyoso, N.

David Theodore, C. Dubourdieu J. Bruley and J. Jordan-Sweet, J. Appl. Phys. 111,

064112, (2012).

15. L. V. Goncharova, D. G. Starodub, E. Garfunkel, T. Gustafsson, V. Vaithyanathan, J.

Lettieri, and D. G. Schlom, J. of Appl. Phy. 100, 014912 (2006);

16. X. Zhang, A. A. Demkov, H. Li, X. Hu, Y. Wei, and J. Kulik, Phys. Rev. B 68, 125323

(2003).

17. F. Amy, A. S. Wan, A. Kahn, F. J. Walker, and R. A. McKee, J. Appl. Phys. 96, 1635

(2004).

18. B. E. Deal, and A. S. Grove, J. Appl. Phys. 36, 3770 (1965).

19. H. Z. Massoud, J. D. Plummer, and E. A. Irene, J. Electrochem. Soc. 132, 1745 (1985).

20. H. Z. Massoud, J. D. Plummer, and E. A. Irene, J. Electrochem. Soc. 132, 2693 (1985).

21. H. Z. Massoud, and J. D. Plummer, J. Appl. Phys. 62, 3416 (1987).

22. A.F. Tasch, T.C. Holloway, K.F. Lee and J.F. Gibbons. Electron. Lett. 15, 435(1979).

23. W. Hiinsch, Th. Vogelsand, R. Kircher and M. Orlowski, Solid-state Electron 32,

839(1989).

24. K. Eisenbeiser, J. M. Finder, Z. Yu, J. Ramdani, J. A. Curless, J. A. Hallmark, R.

Droopad, W. J. Ooms, L. Salem, S. Bradshaw, and C. D. Overgaard. Appl. Phys. Lett.

76, 1324 (2000).

Page 137: Copyright By Wei Guo 2021

118

25. S. Guha and V. Narayanan, Phys. Rev. Lett. 98, 196101 (2007).

26. J. W. Park, D. F. Bogorin, C. Cen, D. A. Felker, Y. Zhang, C. T. Nelson, C. W. Bark,

C. M. Folkman, X. Q. Pan, M. S. Rzchowski, J. Levy, and C. B. Eom, Nat. Commun.

1, 94 (2010).

27. G. J. Yong, Rajeswari M. Kolagani, S. Adhikari, W. Vanderlinde, Y. Liang, K.

Muramatsu, and S. Friedrich, J. Appl. Phys. 108, 033502 (2010);.

28. K. J. Hubbard and D. G. Schlom, J. Mater. Res. 11, 2757(1996).

29. A. Posadas, R. Dargis, M. Choi, A. Slepko, J.J. Kim, D.J. Smith, and A.A. Demkov, J.

Vac. Sci. Technol. B. 23, 1071 (2011).

30. S. Zollner, A.A. Demkov, R. Liu, P.L. Fejes, R.B. Gregory, P. Alluri, J. Curless, Z.

Yu,J. Ramdani, R. Droopad, T.E. Tiwald, J.N. Hilfiker, J.A. Woollam, J. Vac. Sci.

Technol. B 18, 2242-2254 (2000).

31. H.S. Kato, S. Shiraki, M. Nantoh, M. Kawai, Surf. Sci. 544, 722(2003)

32. http://www.lelandstanfordjunior.com/thermaloxide.html

33. P. Pasierb, S. Komornicki, M. Rekas, J. Phys. Chem. Solid. 60, 1835(1999)

34. D. Lu, D. J. Baek, S. S. Hong, L. F. Kourkoutis, Y. Hikita and H. Y. Hwang, Nature

Materials, 15, 1255(2016)

35. S. S. Hong, J. H. Yu, D. Lu, A. F. Marshall, Y. Hikita, Y. Cui, H. Y. Hwang, Sci. Adv.

3, 5173(2017)

36. D. Ji, S. Cai, T.R. Paudel, et al. Nature 570, 87–90 (2019).

37. A. A. Sirenko, I. A. Akimov, J. R. Fox, A. M. Clark, Hong-Cheng Li, Weidong Si, and

X. X. Xi, Phys. Rev. Lett. 82, 4500(1999)

38. W. G. Nilsen and J. G. Skinner, J. Chem. Phys. 48, 2240 (1968)

39. C. H. Perry, Jeanne H. Fertel and T. F. McNelly, J. Chem. Phys. 47, 1619 (1967)

40. D. A. Tenne, A. K. Farrar, C. M. Brooks, T. Heeg, J. Schubert, H. W. Jang, C. W. Bark,

C. M. Folkman, C. B. Eom, and D. G. Schlom, Appl. Phys. Lett. 97, 142901 (2010)

41. L. Hilt Tisinger, R. Liu, J. Kulik, X. Zhang, J. Ramdani, and A. A. Demkov. J. Vac.

Sci. Technol. B, 21, 53 (2003)

42. R. Ouillon, J-P. Pinan-Lucarre, P.Ranson, P. Pruzan, S. K. Mishra, R. Ranjan and D.

Pandey, J. Phys.: Condens. Matter 14, 2079(2002)

43. Y.L. Du, G. Chen, M.S. Zhang, Solid State Commun. 130, 577(2004)

44. Choi, K., Raza, S.R.A., Lee, H.S., Jeon, P.J., Pezeshki, A., Min, S.W., Kim, J.S., Yoon,

W., Ju, S.Y., Lee, K., Im, S. Nanoscale 7, 5617 (2015)

Page 138: Copyright By Wei Guo 2021

119

45. S. Kang, N. Prasad, H.C.P. Movva, A. Rai, K. Kim, X. Mou, T. Taniguchi, K.

Watanabe, L.F. Register, E. Tutuc, and S.K. Banerjee, Nano Lett. 16, 4975 (2016)

Page 139: Copyright By Wei Guo 2021

120

Chapter 6: Advanced design of BaSnO3/SrTiO3/Al2O3 quantum wells

BaSnO3 and SrSnO3 have triggered huge interest in the area of transparent

conductive oxides as they have superior electron conductivity and wide optical band gaps.

Here, we propose an advanced quantum well structure design of BaSnO3/STO/Al2O3

utilizing the large conduction band offset (~3.5 eV) between BaSnO3 and Al2O3. The

deposition process is done by MBE and the heterostructures are characterized by RHEED

and XPS to confirm the film quality and composition. The work in this chapter is a

prototype implementation of this quantum well design. Further electrical and optical

measurements are needed to verify the actual physical properties of this system.

6.1 INTRODUCTION

Lightly La-doped BaSnO3 (BSO) and SrSnO3 (SSO) have very high room-

temperature conductivity and wide band gap, which makes them attractive for novel

transparent conductors and high-power electronic devices [1]-[10]. Since the early 20120s,

many people have studied the materials both theoretically and experimentally, particularly

the stannate growth process. Kim et al. [1] have reported unprecedentedly high mobility at

room temperature for a PLD-grown 4% La-doped BSO sample. Mobility in single crystal

samples reached as high as 320 cm2/V•s while retaining their optical transparency;

meanwhile, the mobility was still only up to 70 cm2/V•s in epitaxial films at that time.

Later, Liu et al. [2] used first-principles calculations to explore the origin of the superior

mobility in alkaline-earth stannates. Their small electron effective masses result from the

large size of Sn ions and the mainly Sn s-orbital derived conduction band edge. Bharat

Jalan’s group performed many studies of the stannate system. They used hybrid MBE with

a metal-organic Sn source to grow stoichiometric BSO [3], which they characterized with

RHEED, scanning transmission electron microscopy (STEM), electron energy-loss

Page 140: Copyright By Wei Guo 2021

121

spectroscopy (EELS), and energy dispersive x-ray spectroscopy (EDX) to confirm the film

quality, epitaxy, crystal structure, and composition. Conduction band offsets were also

determined [4] for BSO/STO and BSO/LAO interfaces to be 3.1 eV and 3.0 eV,

respectively. They also reported adsorption controlled BSO growth by hybrid MBE in 2017

[5]. Films grown within this limited growth window are shown to yield La-doped BSO

films with mobilities of 105 cm2/V•s. Low mobility and charge compensation are induced

in Ba- and Sn-deficient films, with a stronger Sn-deficiency dependence. They also

analyzed the defect-driven localization crossover in La-doped SSO films [6]. They pointed

out that substrate-induced dislocations in the film strongly influence the electron phase

coherence length, which causes two-dimensional to three-dimensional weak localization

crossover. They used epitaxial strain to engineer the SSO film phase and mobility [7] and

achieved over 300% mobility enhancement at room temperature compared with unstrained

low-temperature orthorhombic polymorph.

Susanne Stemmer’s group also reported several MBE-grown BSO films with high

mobility. They used PrScO3 as a lattice-matched substrate and obtained 150 cm2/V•s

mobility at room temperature [8]. Structural images and band gap data of their stannate

films were reported in a later publication [9]. Paik et al. also published an adsorption

controlled BSO growth process by standard MBE [10]. BSO films grown on DyScO3

substrate showed a 183 cm2/V•s mobility at room temperature and 400 cm2/V•s at 10 K.

Stannates have demonstrated excellent mobility and optical transparence. The wide

bandgap is also suitable for building up large band offsets with certain materials. We can

design quantum wells (QWs) by building heterostructures of materials with large band

offsets. By inserting a thin layer of one material between two layers of another material

with even larger band gap, we can make a simple quantum well. The energy in the wells

Page 141: Copyright By Wei Guo 2021

122

becomes discretized in order to satisfy the Schrodinger equations in this confined potential.

The fabrication of novel TMO QW devices opens up new possibilities for advanced

functionalities [11]-[15]. Jackson et al. showed interface-induced magnetism in

GdTiO3/SrTiO3 and SmTiO3/SrTiO3 quantum wells [11]. Need et al. showed that, in

GdTiO3/SrTiO3 quantum wells, the net ferromagnetism inherent to the Mott insulator

GdTiO3 matrix propagates into the nominally nonmagnetic STO quantum wells [12]. Choi

et al. fabricated a PbZrO3/PbTiO3 quantum well [13] with superior dielectric constant (800)

at a stacking period of 1 uc/1 uc (PZO1/PTO1). There are relatively fewer reports of large

band offset quantum wells that allow mid-infrared intersubband absorption. Zhao et al. use

ZnO/ZnMgO quantum wells to achieve mid-infrared intersubband absorption [14].

Ortmann et al. managed to integrate LAO/STO superlattices on Si and showed the ability

to control intersubband absorption energy by changing the width of the STO well layers

[15].

In this chapter, we develop an MBE approach to the growth of BSO and SSO in

our DCA M600 Oxide-MBE system. We also propose a novel BSO/STO/Al2O3 quantum well

structure that can utilize the large band offset between BSO and Al2O3 and the high mobility

of BSO.

6.2 GROWTH AND QW DESIGN

We first grow epitaxial BSO in our MBE system on STO substrates. BSO has a

decent lattice match with STO and grows well on it, and we can also use the STO/Si

pseudo-substrate to integrate BSO on Si. Details of STO on Si deposition are described in

Chapters 4 and 5. Ba is evaporated from effusion cells and Sn comes from an e-beam

evaporator in form of SnO2. Using SnO2 in the e-gun yields a 1-2×10-6 Torr base oxygen

Page 142: Copyright By Wei Guo 2021

123

pressure during deposition, coming from the SnO2 decomposition. Additional oxygen is

needed to maintain a total oxygen pressure at 2-3×10-5 Torr to ensure Sn is fully oxidized.

The real Ba flux is also observed to decrease slightly when the oxygen pressure exceeds

10-5 Torr likely due to source oxidation. This requires one to increase the Ba flux 1.2-1.3

times above the normal flux value to maintain the same atom arrival rate. The deposition

temperature is naintained between 750-800 ˚C. Higher temperatures will result in Sn re-

evaporation and failure to form the correct phase. At lower temperatures, BSO will not

crystallize well.

The tricky part of using an e-gun for stannate deposition is the unstable e-gun flux.

The SnO2 flux decays too quickly, usually 20-40% over a 5 min period when the rate is ~1

monolayer/min. To guarantee the crystal quality, the flux needs to be sufficiently stable to

maintain the stoichiometric condition. Since we are using SnO2 as Sn source, the total

oxygen pressure is directly related to the Sn flux, and this has been verified to have a linear

dependence by flux measurement using a quartz crystal monitor. We use the real-time

measured chamber pressure as a parameter to represent the Sn flux, and actively maintain

the pressure by minor adjustments to the e-gun emission current to compensate the

decaying flux during the entire deposition process.

We propose a BSO/Al2O3 quantum well structure based on the large band offset

between Al2O3 (7.6 eV) and BSO (3.1 eV). Al2O3 can be epitaxially grown on BSO with

good crystallinity (as -Al2O3), but the reverse is not true due to surface energy issues. To

get around this, we introduce an STO layer on top of Al2O3 to allow BSO to wet and

crystallize as a flat layer. The band alignment of the stack is shown in Figure 6.1 and a

schematic of the proposed structure is shown in Figure 6.2.

Page 143: Copyright By Wei Guo 2021

124

Figure 6.1: Band alignment of a single quantum well. The conduction band offset

between BSO and Al2O3 is 3.5 eV.

Figure 6.2: Schematic of the heterostructures to be grown.

Page 144: Copyright By Wei Guo 2021

125

The valence band offsets (VBOs) marked in Figure 6.1 were extracted from XPS

measurements of various stacks consisting of BSO, Al2O3 and STO layers. First, we

measure a core level and the valence band maximum (VBM) binding energies in STO and

BSO bulk-like films. Second, we measure the same core level binding energies in a

BSO/STO stack. Here we use Ti 2p in STO and Ba 3d in BSO. The VBO is calculated as

follows:

VBOSTO/BSO = (ETi 2p- EVBM)STO - (EBa 3d- EVBM)BSO - (ETi 2p- EBa 3d)STO/BSO

With the same method, we get VBOSTO/Al2O3 = 2.3 eV and VBOBSO/Al2O3 = 1 eV. We

also know the experimental band gaps of each material, which allows one to obtain the

conduction band offsets (CBO), as marked by green lines in Figure 6.1. We can see that

the CBO for BSO and Al2O3 is very large at 3.5 eV.

Based on the CBO and the designed well width in this structure, we performed a

simulation using a Poisson-Schrodinger solver. The well consists of 3 uc of STO and 10

uc of BSO with infinitely thick Al2O3 barriers. The simulation was performed by Suyeong

Jang in our group, and the results are shown in Figure 6.3.

Page 145: Copyright By Wei Guo 2021

126

Figure 6.3: Simulation results for possible energy states in a single quantum well.

Based on the energy levels in the well, we see that one can get transitions whose

energies range from visible light to infrared light. This is a very promising quantum well

structure with a wide range of energy absorption possibilities.

6.3 CHARACTERIZATION

We monitored the RHEED pattern from the surface of the sample during the

deposition procedure to check the crystallinity and surface quality of each quantum well

layer as it is being formed. We also used XPS to check the stoichiometry of the materials

and to measure core level energies in determining the band offsets.

In Figure 6.4, we show the RHEED pattern evolution as each layer of the

heterostructure is deposited. This structure is a 10 uc BSO/3 uc STO/thick Al2O3 stack. The

pattern brightness and sharpness decreases from Al2O3 to BSO which means crystal quality

gets worse as more layers are put on top. There is still work needed to optimize the growth

Page 146: Copyright By Wei Guo 2021

127

process in order to maintain a good crystal quality if we want to repeatedly build a

superlattice of this stack.

(a)

(b)

(c)

Page 147: Copyright By Wei Guo 2021

128

Figure 6.4: RHEED evolution from (a) thick Al2O3 on STO buffer layer on Si (b) 3 uc

STO on previous Al2O3 surface (c) 10 uc BSO on previous STO surface

We also show the VB spectrum of a BSO/STO /Al2O3 stack in Figure 6.5. This

sample has 5 uc BSO on 3 uc STO on thick Al2O3, and shows a combination of STO VB

and BSO VB features.

Figure 6.5: VB spectrum of BSO/STO/Al2O3

6.4 OUTLOOK

We are still working on improving the quantum well crystal quality in order to be

able to repeatedly stack these layers for intersubband absorption measurements. The key is

to maintain the BSO crystal quality on the Al2O3 surface with an STO buffer. Transport

measurements of La-doped BSO films grown in our MBE also need to be performed to

optimize the process, especially the doping. This quantum well structure has great potential

for advanced quantum well devices once the process can be optimized.

Page 148: Copyright By Wei Guo 2021

129

6.5 REFERENCES

1. H. J. Kim, U. Kim, H. M. Kim, T. H. Kim, H. S. Mun, B. G. Jeon, K. T. Hong, W.

J. Lee, C. Ju, K. H. Kim and K. Char. Appl. Phys. Express 5, 061102(2012)

2. H. R. Liu, J. H. Yang, H. J. Xiang, X. G. Gong and S. H. Wei, Appl. Phys. Lett.

102, 112109 (2013)

3. A. Prakash, J. Dewey, H. Yun, J. S. Jeong, K. A. Mkhoyan, and B. Jalan, J. Vac.

Sci. Technol. A 33, 060608 (2015)

4. S.A. Chambers, T. C. Kaspar, A. Prakash, G. Haugstad and B. Jalan, Appl. Phys.

Lett. 108, 152104 (2016)

5. A. Prakash, P. Xu, X. Wu, G. Haugstad, X. Wang and B. Jalan, J. Mater. Chem. C

5, 5730(2017)

6. T. Wang, L. R. Thoutam, A. Prakash, W. Nunn, G. Haugstad and B. Jalan, Phys.

Rev. Mater. 1, 061601(2017)

7. T. Wang, A. Prakash, Y. Dong, T. Truttmann, A. Bucsek, R. James, D. D. Fong, J.

Kim, P. J. Ryan, H.Zhou, T. Birol and B. Jalan, ACS Appl. Mater. Interfaces 10,

43802(2018)

8. S. Raghavan, T. Schumann, H. Kim, J. Y. Zhang, T. A. Cain and S. Stemmer, APL

Materials 4, 016106 (2016)

9. T. Schumann, S. Raghavan, K. Ahadi, H. Kim and S. Stemmer. J. Vac. Sci. Technol.

A 34, 050601 (2016)

10. H. Paik, Z. Chen, E. Lochocki, A. Seidner H., A. Verma, N. Tanen, J. Park, M.

Uchida, S. Shang, B. Zhou, M. Brützam, R. Uecker, Z. Liu, D. Jena, K. M. Shen,

D. A. Muller and D. G. Schlom, APL Mater. 5, 116107 (2017)

11. C. A. Jackson, S. Stemmer, Phys. Rev. B: Condens. Matter Mater. Phys. 88,

180403(2013)

12. R. F. Need, B. J. Isaac, B. J. Kirby, J. A. Borchers, S. Stemmer, and Stephen D.

Wilson, Phys. Rev. Lett. 117, 037205(2016)

13. T. Choi, J. Lee, Thin Solid Films 475, 283(2005)

14. K. Zhao, G. Chen, B. Li and A. Shen, Appl. Phys. Lett. 104, 212104 (2014)

15. J. E. Ortmann, N. Nookala, Q. He, L. Gao, C. Lin, A. B. Posadas, A. Y.

Borisevich, M. A. Belkin, and A. A. Demkov, ACS NANO 12, 7682 (2018).

Page 149: Copyright By Wei Guo 2021

130

Chapter 7: Summary and future work

7.1 SUMMARY

In this thesis, I have presented major results of my studies of transition metal and

transition metal oxide integration with SrTiO3 and SrTiO3/Si templates.

Pt deposition on STO was studied using different thickness and under multiple

temperatures. We use in situ XPS, ex situ AFM and SEM to explore the core level binding

energy, valence band spectrum, and surface morphology as functions of coverage and

deposition temperature. At low temperature, Pt atoms tend to stay at the low energy oxygen

hollow site owing to the relatively low mobility of Pt. In essence Pt can cover STO even

though surface energies make it thermodynamically unfavorable. High growth temperature

will give Pt enough mobility to move across the surface and form clusters. Pt clusters will

exhibit metallic XPS features even at very low coverage. SEM imaging clearly shows

nanoclusters forming at high growth temperature and a continuous film at low temperature

for the same coverage.

Eu deposition on STO in UHV results in EuO epitaxial growth on STO. We studied

the scavenging process at different temperatures on different thicknesses of STO films

integrated on Si. At 300°C, Eu takes oxygen from STO films that are thicker than 4 nm and

forms EuO. For thinner STO, Eu takes too much oxygen and destabilizes the STO layer,

allowing for a solid-state reaction with Si that results in the formation of EuSiy and silicates.

Low temperature growth of Eu only forms up to a ∼20 Å thick layer of EuO. Beyond this

point, Eu no longer takes oxygen from STO and Eu metal accumulates on the surface. The

EuO/Eu metal interface was also studied theoretically and experimentally. A downward

band bending is confirmed near the interface. The study of the ferroelectric field-effect of

the EuO/STO interface is very preliminary. It is currently limited by technical difficulties

Page 150: Copyright By Wei Guo 2021

131

of EuO over-oxidation when transporting samples. In the process, we found a workable

solution using a GeOx capping layer. More research needs to be done to optimize this

process.

We also performed studies on STO/Si. We first considered dry oxidation of STO/Si

heterostructures and demonstrate that the underlying Si can be safely oxidized at a

relatively high temperature (800ºC) with the STO crystallinity not significantly degraded.

We built a Deal-Grove-like oxidation model and found good agreement between the data

and the model. This gives us the ability to predict the temperature and time needed to obtain

the desired SiO2 thickness for Si that is covered by a thin layer of STO. This additional

knob for controlling the layer structure can enable one to integrate more complicated oxide

structures on this STO/Si pseudo-substrate, especially for applications requiring complete

decoupling between the STO and Si. With thick SiO2 interlayer available, we were able to

do selective etching of the underlying SiO2 to make free-standing STO membranes. Using

NaOH etching, we obtained free-standing STO membranes with a typical size of ~5×10

µm2. Raman spectroscopy was used to see if there were any differences between STO thin

films and membranes. Preliminary results were obtained but with weak STO membranes

signals.

We proposed a novel BSO/STO/Al2O3 quantum well structure with a large

conduction band offset (~3.5 eV) and the possibility of near- to mid-infrared absorption. We

showed the energy band diagram and the band offsets for such a structure. RHEED and

XPS were used to characterize the surface quality and the stoichiometry of the

heterostructure. We are working on improving the growth process to maintain good surface

crystallinity and make repeated stacks for further energy absorption measurements a

possibility.

Page 151: Copyright By Wei Guo 2021

132

7.2 FUTURE WORK

TMs and TMOs have a wide range of interesting properties and phenomena to

explore. When they are integrated with STO and Si, additional industrial possibilities are

opened up.

For our EuO/STO interface 2DEG research, we are improving the capping layer to

achieve a better ARPES measurement. This study provides a new possibility to control the

interface charge distribution thus achieving an unexplored way of modulating the 2DEG,

and possibly shift the Curie temperature of EuO to the range of practical applications.

STO free-standing membranes are promising for studying physical properties of

reduced dimensionality and also for potential future transistors. We are working on the

optimization of the Raman setup to reduce noise and amplify the STO membrane signal.

We can use different thicknesses of STO membranes to study the phonon modes of such a

quasi-2D structure. We can also explore the design of new transistor devices using our STO

membranes with their high-k values.

The quantum well of BSO/STO/Al2O3 needs to improvements in the crystal quality

in order to make repeated wells for further energy absorption measurements. Extensive

transport measurements of different doping levels of La-doped BSO samples are also

needed.

Page 152: Copyright By Wei Guo 2021

133

Bibliography

• A. A. Demkov and A. B. Posadas, Integration of Functional Oxides with

Semiconductors (Springer New York, 2014).

• A. A. Melville, T. Mairoser, A. Schmehl, D. E. Shai, E. J. Monkman, J. W. Harter, T.

Heeg, B. Hollander, J. Schubert, K. M. Shen, J. Mannhart, and D. G. Schlom, Appl.

Phys. Lett. 100, 222101 (2012).

• A. A. Sirenko, I. A. Akimov, J. R. Fox, A. M. Clark, Hong-Cheng Li, Weidong Si, and

X. X. Xi, Phys. Rev. Lett. 82, 4500(1999)

• A. Asthagiri and D. S. Sholl, J. Chem. Phys. 116, 9914 (2002).

• A. B. Posadas, K. J. Kormondy, W. Guo, P. Ponath, J. Geler-Kremer, T. Hadamek,

and A. A. Demkov, J. Appl. Phys. 121, 105302 (2017)

• A. Brinkman, M. Huijben, M. van Zalk, J. Huijben, U. Zeitler, J. C. Maan, W. G. van

der Wiel, G. Rijnders, D. H. A. Blank, and H. Hilgenkamp, Nat. Mater. 6, 493 (2007).

• A. D. Polli, T. Wagner, T. Gemming, and M. R€uhle, Surf. Sci. 448, 279 (2000).

• A. Einstein, Ann. Phys. 322, 132 (1905).

• A. G. Swartz, P. M. Odenthal, Y. Hao, R. S. Ruoff, and R. K. Kawakami, ACS Nano

6, 10063 (2013).

• A. Ichimiya, P. I. Cohen, Reflection High Energy Electron Diffraction Cambridge

University Press (2004).

• A. Mauger and C. Godart, Phys. Rep. 141, 51 (1986).

• A. Ohtomo and H. Y. Hwang, “A high-mobility electron gas at the LaAlO3/SrTiO3

heterointerface,” Nature, 427, 423 (2004).

• A. Posadas, R. Dargis, M. Choi, A. Slepko, J.J. Kim, D.J. Smith, and A.A. Demkov, J.

Vac. Sci. Technol. B. 23, 1071 (2011).

• A. Prakash, J. Dewey, H. Yun, J. S. Jeong, K. A. Mkhoyan, and B. Jalan, J. Vac. Sci.

Technol. A 33, 060608 (2015)

• A. Prakash, P. Xu, X. Wu, G. Haugstad, X. Wang and B. Jalan, J. Mater. Chem. C 5,

5730(2017)

• A. Schmehl et al., Nat. Mater. 6, 882 (2007).

• A. Y. Cho, F. K. Reinhart, Appl. Phys. Lett. 21, 355 (1972).

• A.F. Tasch, T.C. Holloway, K.F. Lee and J.F. Gibbons. Electron. Lett. 15, 435(1979).

• B. E. Deal, and A. S. Grove, J. Appl. Phys. 36, 3770 (1965).

Page 153: Copyright By Wei Guo 2021

134

• B. El-Kareh, Fundamentals of Semiconductor Processing Technologies, (Springer US,

New York, 1995).

• B. G. Streetman, S. Banerjee, Solid State electronic Devices, 524 (Hall, New Jersey,

2000).

• B. I. Edmondson, S. Liu, S. Lu, H. W. Wu, A. B. Posadas, D. J. Smith, X. P. A. Gao,

A. A. Demkov, and J. G. Ekerdt, J. Appl. Phys. 124, 185303 (2018).

• B. Kalisky, J. A. Bert, B. B. Klopfer, C. Bell, H. K. Sato, M. Hosoda, Y. Hikita, H. Y.

Hwang, and K. A. Moler. Nat. Commun. 3, 922 (2012).

• B. N. Figgis and J.Lewis, The Magnetochemistry of Complex Compounds in Modern

Co-ordination Chemistry, Ed. J. Lewis and C. Wilkins, Interscience, New York (1960).

• B. Radisavljevic, A. Radenovic, J. Brivio, V. Giacometti and A. Kis. Nature Nanotech

6, 147(2011).

• B. Reihl, J. G. Bednorz, K. A. M€uller, Y. Jugnet, G. Landgren, and J. F. Morar, Phys.

Rev. B 30, 803 (1984).

• B. T. Matthias, R. M. Bozorth, and J. H. Van Vleck, Phys. Rev. Lett. 7, 160 (1961).

• B. W. Wessels, Annu. Rev. Mater. Res. 37, 659 (2007).

• C. A. Jackson, S. Stemmer, Phys. Rev. B: Condens. Matter Mater. Phys. 88,

180403(2013)

• C. Caspers, M. Müller, A. X. Gray, A. M. Kaiser, A. Gloskovskii, C. S. Fadley, W.

Drube, and C. M. Schneider, Phys. Rev. B. 84, 205217 (2011).

• C. H. Perry, Jeanne H. Fertel and T. F. McNelly, J. Chem. Phys. 47, 1619 (1967)

• Choi, K., Raza, S.R.A., Lee, H.S., Jeon, P.J., Pezeshki, A., Min, S.W., Kim, J.S., Yoon,

W., Ju, S.Y., Lee, K., Im, S. Nanoscale 7, 5617 (2015)

• D. A. Tenne, A. K. Farrar, C. M. Brooks, T. Heeg, J. Schubert, H. W. Jang, C. W. Bark,

C. M. Folkman, C. B. Eom, and D. G. Schlom, Appl. Phys. Lett. 97, 142901 (2010)

• D. F. Bogorin, P. Irvin, C. Cen, and J. Levy, “LaAlO3/SrTiO3-based device concepts,”

in Multifunctional Oxide Heterostructures, Oxford University Press, 364(2010).

• D. Hisamoto, W. Lee, J. Kedzierski, H. Takeuchi, K. Asano, C. Kuo, E. Anderson, T.

King, J. Bokor and C. Hu. IEEE Transactions on Electron Devices. 47 2320 (2000).

• D. J. Frank, R. H. Dennard, E. Nowak, P. M. Solomon, Y. Taur and Hon-Sum Philip

Wong, Proceedings of the IEEE. 89, 259 (2001).

• D. Ji, S. Cai, T.R. Paudel, et al. Nature 570, 87–90 (2019).

Page 154: Copyright By Wei Guo 2021

135

• D. Lu, D. J. Baek, S. S. Hong, L. F. Kourkoutis, Y. Hikita and H. Y. Hwang, Nature

Materials, 15, 1255(2016)

• D. P. Kumah, J. H. Ngai and L. Kornblum, Adv. Funct. Mater. 30, 1901597(2019)

• D. V. Averyanov, Y. G. Sadofyev, A. M. Tokmachev, A. E. Primenko, I. A. Likhachev,

and V. G. Storchak, ACS Appl. Mater. Interfaces 7, 6146–6152 (2015).

• D. V. Averyanov, Y. G. Sadofyev, A. M. Tokmachev, A. E. Primenko, I. A. Likhachev,

and V. G. Storchak, J. Appl. Phys. 113, 17C505 (2013).

• E. Harari, J. Appl. Phys. 49, 2478-2489 (1978).

• E. Vella, F. Messina, M. Cannas, R. Boscaino, Phys. Rev. B. 83, 174201 (2011).

• EuO Crystal Structure: Datasheet from “PAULING FILE Multinaries Edition—2012”

in Springer Materials,

see: http://materials.springer.com/ isp/crystallographic/docs/sd_0553841.

• F. Amy, A. S. Wan, A. Kahn, F. J. Walker, and R. A. McKee, J. Appl. Phys. 96, 1635

(2004).

• F. Pockels, Lehrbuch Der Kristalloptik (Leipzig und Berlin, 1906).

• F. Rao, M. Kim, A. J. Freeman, S. Tang, and M. Anthony, Phys. Rev. B 55, 13953

(1997).

• G. J. Yong, Rajeswari M. Kolagani, S. Adhikari, W. Vanderlinde, Y. Liang, K.

Muramatsu, and S. Friedrich, J. Appl. Phys. 108, 033502 (2010);.

• G. K. Wertheim and S. B. DiCenzo, Phys. Rev. B 37, 844 (1988).

• G. M. Prinz, T. Gerber, A. Lorke, and M. Müller, Appl. Phys. Lett. 109, 202401 (2016).

• G. T. Reed, G. Mashanovich, F. Y. Gardes and D. J. Thomson, Nat. Photon. 4, 518

(2010).

• H. Brune, Surf. Sci. Rep. 31, 125 (1998).

• H. Iwai, Microelectron. Eng. 86, 1520 (2009).

• H. J. Kim, U. Kim, H. M. Kim, T. H. Kim, H. S. Mun, B. G. Jeon, K. T. Hong, W. J.

Lee, C. Ju, K. H. Kim and K. Char. Appl. Phys. Express 5, 061102(2012)

• H. Li, X. Hu, Y. Wei, Z. Yu, X. Zhang, R. Droopad, A. A. Demkov, J. Edwards, Jr.,

K. Moore, and W. Ooms, J. Appl. Phys. 93, 4521 (2003).

• H. Miyazaki, H. Momiyama, T. Hajiri, T. Ito, K. Imura, M. Matsunami, and S. Kimura,

J. Phys. Conf. Ser. 391, 012047 (2012).

• H. Miyazaki, T. Ito, H. Im, K. Terashima, S. Yagi, M. Kato, K. Soda, and S. Kimura,

Jpn. J. Appl. Phys. 48, 055504 (2009).

Page 155: Copyright By Wei Guo 2021

136

• H. N. Lee, H. M. Christen, M. F. Chisholm, C. M. Rouleau, and D. H. Lowndes. Nature,

433, 395 (2005).

• H. Paik, Z. Chen, E. Lochocki, A. Seidner H., A. Verma, N. Tanen, J. Park, M. Uchida,

S. Shang, B. Zhou, M. Brützam, R. Uecker, Z. Liu, D. Jena, K. M. Shen, D. A. Muller

and D. G. Schlom, APL Mater. 5, 116107 (2017)

• H. R. Liu, J. H. Yang, H. J. Xiang, X. G. Gong and S. H. Wei, Appl. Phys. Lett. 102,

112109 (2013)

• H. Seo and A. A. Demkov, Phys. Rev. B 92, 245301 (2015).

• H. Seo, A. B. Posadas, and A. A. Demkov, J. Vac. Sci. Technol., B 30, 04E108 (2012).

• H. Y. Hwang, Y. Iwasa, M. Kawasaki, B. Keimer, N. Nagaosa, and Y. Tokura, Nat.

Mater. 11, 103 (2012).

• H. Yamada, M. Kawasaki, Y. Ogawa, and Y. Tokura, Appl. Phys. Lett. 81, 4793 (2002).

• H. Z. Massoud, and J. D. Plummer, J. Appl. Phys. 62, 3416 (1987).

• H. Z. Massoud, J. D. Plummer, and E. A. Irene, J. Electrochem. Soc. 132, 1745 (1985).

• H.S. Kato, S. Shiraki, M. Nantoh, M. Kawai, Surf. Sci. 544, 722(2003)

• http://www.lelandstanfordjunior.com/thermaloxide.html

• J. A. Bert, B. Kalisky, C. Bell, M. Kim, Y. Hikita, H. Y. Hwang, and K. A. Moler, Nat.

Phys. 7. 767(2011).

• J. A. Kittl, K. Opsomer, M. Popovici, N. Menou, B. Kaczer, X. P. Wang, C. Adelmann,

M. A. Pawlak, K. Tomida, A. Rothschild, B. Govoreanu, R. Degraeve, M. Schaekers,

M. Zahid, A. Delabie, J. Meersschaut, W. Polspoel, S. Clima, G. Pourtois, W. Knaepen,

C. Detavernier, V. V. Afanas’ev, T. Blomberg, D. Pierreux, J. Swerts, P. Fischer, J. W.

Maes, D. Manger, W. Vandervorst, T. Conard, A. Franquet, P. Favia, H. Bender, B.

Brijs, S. Van Elshocht, M. Jurczak, J. Van Houdt, D. J. Wouters, Microelectron. Engin.

86, 1789 (2009).

• J. Braun, J. Minár, S. Mankovsky, V. N. Strocov, N. B. Brookes, L. Plucinski, C. M.

Schneider, C. S. Fadley, and H. Ebert, Phys. Rev. B 88, 205409(2013)

• J. Chakhalian, J. W. Freeland, A. J. Millis, C. Panagopoulos, and J. M. Rondinelli, Rev.

Mod. Phys. 86, 1189 (2014).

• J. D. Plummer, M. D. Deal, P. B. Griffin, Silicon VLSI Technology, (Prentice Hall,

Upper Saddle River, 2013).

• J. E. Ortmann, A. B. Posadas and A. A. Demkov, J. Appl. Phys. 124, 015301 (2018).

• J. E. Ortmann, M. A. Duncan and A. A. Demkov, Optical Materials Express 9, 2982

(2019).

Page 156: Copyright By Wei Guo 2021

137

• J. E. Ortmann, N. Nookala, Q. He, L. Gao, C. Lin, A. B. Posadas, A. Y. Borisevich, M.

A. Belkin, and A. A. Demkov, ACS NANO 12, 7682 (2018).

• J. E. Ortmann, S. Kwon, A. B. Posadas, M. J. Kim and A. A. Demkov, J. Appl. Phys.

125, 155302 (2019).

• J. H. Haeni, P. Irvin, W. Chang, R. Uecker, P. Reiche, Y. L. Li, S. Choudhury, W. Tian,

M. E. Hawley, B. Craigo, A. K. Tagantsev, X. Q. Pan, S. K. Streiffer, L. Q. Chen, S.

W. Kirchoefer, J. Levy and D. G. Schlom. Nature 430, 758 (2004).

• J. K. Lang, Y. Baer, and P. A. Cox, J. Phys. F. 11, 121 (1981).

• J. K. Lee, N. Sai, and A.A. Demkov, Phys. Rev. B 82, 235305 (2010).

• J. Lettieri, J. H. Haeni, and D. G. Schlom, J. Vac. Sci. Technol. A 20, 1332 (2002).

• J. Lettieri, V. Vaithyanathan, S. K. Eah, J. Stephens, V. Sih, D. D. Awschalom, J. Levy,

and D. G. Schlom, Appl. Phys. Lett. 83, 975 (2003).

• J. R. Arthur Jr., J. Appl. Phys. 39, 4032 (1968).

• J. Ramdani, R. Droopad, Z. Yu, J. A. Curless, C. D. Overgaard, J. Finder,K.

Eisenbeiser, J. A. Hallmark, W. J. Ooms, V. Kaushik, P. Alluri, and S. Pietambaram,

Appl. Surf. Sci. 159, 127 (2000).

• J. Robertson, J. Vac. Sci. Technol. B 18, 1785 (2000).

• J. Schoenes and P. Wachter, Phys. Rev. B. 9, 3097 (1974).

• J. Son, J. Cagnon, and S. Stemmer, J. Appl. Phys. 106, 043525 (2009).

• J. W. Park, D. F. Bogorin, C. Cen, D. A. Felker, Y. Zhang, C. T. Nelson, C. W. Bark,

C. M. Folkman, X. Q. Pan, M. S. Rzchowski, J. Levy, and C. B. Eom, Nat. Commun.

1, 94 (2010).

• J.E. Mahan, Physical Vapor Deposition of Thin Films, Wiley-VCH, 2000.

• K. A. Kovacs, P. Grof, L. Burai, M. Riedel. J. Phys. Chem. A. 108, 11026 (2004).

• K. D. Fredrickson, H. Seo and A. A. Demkov, J. Appl. Phys. 120, 065301 (2016).

• K. D. Fredrickson, V. V. Vogler-Neuling, K. J. Kormondy, D. Caimi, F. Eltes, M.

Sousa, J. Fompeyrine, S. Abel, and A. A. Demkov, Phys. Rev. B 98, 075136 (2018).

• K. Eisenbeiser, J. M. Finder, Z. Yu, J. Ramdani, J. A. Curless, J. A. Hallmark, R.

Droopad, W. J. Ooms, L. Salem, S. Bradshaw, and C. D. Overgaard. Appl. Phys. Lett.

76, 1324 (2000).

• K. G. Günther, Z. Naturforschg. A. 13 (12), 1081-1089 (1958).

• K. J. Hubbard and D. G. Schlom, J. Mater. Res. 11, 2757 (1996).

Page 157: Copyright By Wei Guo 2021

138

• K. J. Hubbard and D. G. Schlom, J. Mater. Res. 11, 2757(1996).

• K. J. Kormondy, A. B. Posadas, T. Q. Ngo, S. Lu, N. Goble, J. Jordan-Sweet, X. P. A.

Gao, D. J. Smith, M. R. McCartney, J. G. Ekerdt, and A. A. Demkov, J. Appl. Phys.

117, 95303 (2015).

• K. J. Kormondy, L. Gao, X. Li, S. Lu, A. B. Posadas, S. Shen, M. Tsoi, M. R.

McCartney, D. J. Smith, J. Zhou, L. L. Lev, M.-A. Husanu, V. N. Strocov, and A. A.

Demkov, Sci. Rep. 8, 7721 (2018).

• K. J. Kormondy, S. Abel, F. Fallegger, Y. Popoff, P. Ponath, A. B. Posadas, M.

Sousa, D. Caimi, H. Siegwart, E. Uccelli, L. Czornomaz, C.Marchiori, J. Fompeyrine,

A. A. Demkov. Microelectron. Eng. 147, 215 (2015).

• K. J. Kormondy, Y. Popoff, M. Sousa, F. Eltes, D. Caimi, M. D. Rossell, M. Fiebig, P.

Hoffmann, C. Marchiori, M. Reinke, M. Trassin, A. A. Demkov, J. Fompeyrine, and

S. Abel, Nanotechnology 28, 075706 (2017).

• K. Kormondy, A. Posadas, T. Q. Ngo, S. Lu, J. L. Jordan-Sweet, X. Guo, D. J. Smith,

M. R. McCartney, J. G. Ekerdt, and A. A. Demkov, J. Appl. Phys. 117, 095303 (2015).

• K. Mistry, C. Allen, C. Auth, B. Beattie, D. Bergstrom, M. Bost, M. Brazier, M.

Buehler, A. Cappellani, R. Chau, C.-H. Choi, G. Ding, K. Fischer, T. Ghani, R. Grover,

W. Han, D. Hanken, M. Hattendorf, J. He, J. Hicks, R. Heussner, D. Ingerly, P. Jain,

R. James, L. Jong, S. Joshi, C. Kenyon, K. Kuhn, K. Lee, H. Liu, J. Maiz, B. McIntyre,

P. Moon, J. Neirynck, S. Pae, C. Parker, D. Parsons, C. Prasad, L. Pipes, M. Prince, P.

Ranade, T. Reynolds, J. Sandford, L. Shifren, J. Sebastian, J. Seiple, D. Simon, S.

Sivakumar, P. Smith, C. Thomas, T. Troeger, P. Vandervoorn, S. Williams, K.

Zawadzki, Int. El. Devices Meet. 247 (2007).

• K. Takada, H. Sakurai, E. Takayama-Muromachi, F. Izumi, R. A. Dilanian, and T.

Sasaki, Nature, 422, 53 (2003).

• K. Ueno, S. Nakamura, H. Shimotani, A. Ohtomo, N. Kimura, T. Nojima, H. Aoki, Y.

Iwasa, and M. Kawasaki, Nat. Mater. 7, 855 (2008).

• K. Zhao, G. Chen, B. Li and A. Shen, Appl. Phys. Lett. 104, 212104 (2014)

• L. Bjaalie, B. Himmetoglu, L. Weston, A. Janotti, and C. G. Van de Walle. New J.

Phys. 16, 025005 (2014).

• L. Gao and A. A. Demkov, Phys. Rev. B 97, 125305 (2018).

• L. Gao, W. Guo, A. B. Posadas, and A. A. Demkov. Phys. Rev. Materials 3, 094403

(2019).

• L. Hilt Tisinger, R. Liu, J. Kulik, X. Zhang, J. Ramdani, and A. A. Demkov. J. Vac.

Sci. Technol. B, 21, 53 (2003)

Page 158: Copyright By Wei Guo 2021

139

• L. Li, C. Richter, J. Mannhart, and R. C. Ashoori, Nat. Phys. 7, 762 (2011).

• L. Liao, Y. Lin, M. Bao, R. Cheng, J. Bai, Y. Liu, Y. Qu, K. L. Wang, Y. Huang and

X. Duan. Nature 467, 305 (2010).

• L. Mazet, S. M. Yang, S. V. Kalinin, S. Schamm-Chardon and C. Dubourdieu, Sci.

Technol. Adv. Mater. 16, 036005 (2015).

• L. V. Goncharova, D. G. Starodub, E. Garfunkel, T. Gustafsson, V. Vaithyanathan, J.

Lettieri, and D. G. Schlom, J. of Appl. Phy. 100, 014912 (2006);

• L. Zhang, Y. Yuan, J. Lapano, M. Brahlek, S. Lei, B. Kabius, V. Gopalan, and R.

Engel-Herbert, ACS Nano, 12, 1306(2018)

• M. Choi, “Monolithic integration of functional perovskite structures on Si” Ph.D. thesis

(The University of Texas at Austin, Austin, TX, 2014).

• M. Choi, A. B. Posadas, R. Dargis, C. K. Shih and A. A. Demkov, D. H. Triyoso, N.

David Theodore, C. Dubourdieu J. Bruley and J. Jordan-Sweet, J. Appl. Phys. 111,

064112, (2012).

• M. Copel, P. R. Duncombe, D. A. Neumayer, T. M. Shaw, and R. M. Tromp, Appl.

Phys. Lett. 70, 3227 (1997).

• M. Hochberg and T. Baehr-Jones, Nat. Photon. 4, 492(2010).

• M. Imada, A. Fujimori, Y. Tokura, Rev. Mod. Phys. 70, 1039 (1998).

• M. K. Bahl, S. C. Tsai, and Y. W. Chung, Phys. Rev. B 21, 1344 (1980).

• M. Kitano, K. Tsujimaru, and M. Anpo, Top. Catal. 49, 4 (2008).

• M. Matsubara, A. Schmehl, J. Mannhart, D. G. Schlom, and M. Fiebig, Phys. Rev. B

81, 214447 (2010).

• M. Matsubara, A. Schroer, A. Schmehl, A. Melville, C. Becher, M. Trujillo-Martinez,

D. G. Schlom, J. Mannhart, J. Kroha, and M. Fiebig, Nat. Commun. 6, 6724 (2015).

• M. Mostovoy, Nat. Mater. 7, 269 (2008).

• M. Mrovec, J.-M. Albina, B. Meyer, and C. Els€asser, Phys. Rev. B 79, 245121 (2009).

• M. Rodahl, F. Höök, A. Krozer, P. Brzezinski, B. Kasemo, Rev. Sci. Instrum. 66, 3924

(1995).

• M. S. Chen and D. W. Goodman, Science 306, 252 (2004).

• M. T. Bohr, R. S. Chau, T. Ghani, K. Mistry, IEEE Spectrum. 44, 29 (2007).

• M. W. Shafer, J. B. Torrance, and T. Penney, J. Phys. Chem. Solids 33, 2251 (1972).

Page 159: Copyright By Wei Guo 2021

140

• N. C. Bristowe, J. Varignon, D. Fontaine, E. Bousquet and Ph. Ghosez, Nat. Comm. 6,

6677 (2015)

• N. Nakagawa, H. Y. Hwang, and D. A. Muller. Nat. Mater. 5, 204 (2006).

• N. Reyren, S. Thiel, A. D. Caviglia, L. F. Kourkoutis, G. Hammerl, C. Richter, C. W.

Schneider, T. Kopp, A.-S. Rüetschi, D. Jaccard, M. Gabay, D. A. Muller, J.-M.

Triscone, and J. Mannhart, Science. 317, 5842 (2007).

• N. Sai, B. Meyer, and D. Vanderbilt, Phys. Rev. Lett. 84, 5636 (2000).

• N. Singh, A. Agarwal, L. K. Bera, T. Y. Liow, R. Yang, S. C. Rustagi, C. H. Tung, R.

Kumar, G. Q. Lo, N. Balasubramanian, and D. L. Kwong. IEEE Electron Device

Letters, 27, 383 (2006)

• P. G. Steeneken, L. H. Tjeng, I. Elfimov, G. A. Sawatzky, G. Ghiringhelli, N. B.

Brookes, and D.-J. Huang, Phys. Rev. Lett. 88, 47201 (2002).

• P. Lomker and M. Muller, Phys. Rev. Mater. 3, 061401(2019).

• P. Lömker, T. C. Rödel, T. Gerber, F. Fortuna, E. Frantzeskakis, P. LeFèvre, F. Bertran,

M. Müller, and A. F. Santander-Syro, Phys. Rev. Mater. 1, 062001(R) (2017).

• P. Pasierb, S. Komornicki, M. Rekas, J. Phys. Chem. Solid. 60, 1835(1999)

• P. Ponath, A. B. Posadas, M. Schmidt, A.-M. Kelleher, M. White, D. O’Connell, P.

Hurley, R. Duffy and A. A. Demkov, J. Vac. Sci. and Technol. B 36 (2018).

• P. Ponath, A. B. Posadas, R. C. Hatch and Alexander A. Demkov, J. Vac. Sci. Technol.

B. 31, 031201 (2013).

• P. Ponath, K. Fredrickson, A. B. Posadas, Y. Ren, X. Wu, R. K. Vasudevan, M. B.

Okatan, S. Jesse, T. Aoki, M. R. McCartney, D. J. Smith, S. V. Kalinin, K. Lai, and A.

A. Demkov, Nature Comm. 6, 6067 (2015).

• P. R. Subramanian, Eu-O (europium-oxygen), in Binary Alloy Phase Diagrams, 2nd

ed., edited by T. B. Massalski (ASM International, Materials Park, OH, 1990), Vol. 2,

pp. 1665–1666.

• Q. Fu and T. Wagner, Surf. Sci. Rep. 62, 431 (2007).

• R. A. McKee, F. J. Walker, and M. F. Chisholm, Phys. Rev. Lett. 81, 3014(1998).

• R. A. McKee, F. J. Walker, and M. F. Chisholm, Science 293, 468 (2001).

• R. C. Alferness, IEEE Transactions on Microwave Theory and Techniques, 82,

1121(1982).

• R. C. Hatch, M. Choi, A. B. Posadas, and A. A. Demkov, J. Vac. Sci. Technol. B. 33,

061204 (2015).

Page 160: Copyright By Wei Guo 2021

141

• R. Konta, T. Ishii, H. Kato, and A. Kudo, J. Phys. Chem. B. 108, 8992 (2004).

• R. Ouillon, J-P. Pinan-Lucarre, P.Ranson, P. Pruzan, S. K. Mishra, R. Ranjan and D.

Pandey, J. Phys.: Condens. Matter 14, 2079(2002)

• R. Ramesh and D. G. Schlom, MRS Bull. 33, 1006(2008).

• R. Schafranek, S. Payan, M. Maglione, and A. Klein, Phys. Rev. B 77, 195310 (2008).

• R. W. Ulbricht, A. Schmehl, T. Heeg, J. Schubert, and D. G. Schlom, Appl. Phys. Lett.

93, 102105 (2008).

• R. F. Need, B. J. Isaac, B. J. Kirby, J. A. Borchers, S. Stemmer, and Stephen D.

Wilson, Phys. Rev. Lett. 117, 037205(2016)

• S. Abel, J. Fompeyrine, “Electro-Optically Active Oxides on Silicon for Photonics” in

Thin Films on Silicon (V. Narayanan, M. Frank and A. A. Demkov. eds.) 455 (2016).

• S. Abel, T. Stöferle, C. Marchiori, C. Rossel, M.D. Rossell, R. Erni, D. Caimi, M.

Sousa, A. Chelnokov, B.J. Offrein, J. Fompeyrine, Nat. Commun. 4, 1671(2013).

• S. Guha and V. Narayanan, Phys. Rev. Lett. 98, 196101 (2007).

• S. Hofmann, Auger- and X-Ray Photoelectron Spectroscopy in Materials Science: A

User-Oriented Guide, Springer Berlin Heidelberg, Berlin, Heidelberg, (2013).

• S. Kang, N. Prasad, H.C.P. Movva, A. Rai, K. Kim, X. Mou, T. Taniguchi, K.

Watanabe, L.F. Register, E. Tutuc, and S.K. Banerjee, Nano Lett. 16, 4975 (2016)

• S. Maekawa, T. Tohyama, S. E. Barnes, S. Ishihara, W. Koshibae, and G. Khaliullin,

Physics of Transition Metal Oxides, (Springer, 2004).

• S. P. KowaIczyk, N. Edelstein, F. R. McFeely, L. Ley, and D. A. Shirley, Chem. Phys.

Lett. 29, 491 (1974).

• S. Raghavan, T. Schumann, H. Kim, J. Y. Zhang, T. A. Cain and S. Stemmer, APL

Materials 4, 016106 (2016)

• S. S. Hong, J. H. Yu, D. Lu, A. F. Marshall, Y. Hikita, Y. Cui, H. Y. Hwang, Sci. Adv.

3, 5173(2017)

• S. T. Christensen, B. Lee, Z. Feng, M. C. Hersam, and M. J. Bedzyk, Appl. Surf. Sci.

256, 423 (2009).

• S. Tanuma, C. J. Powell, and D. R. Penn, Surf. Interface Anal. 36, 1 (2004).

• S. W. Cheong and M. Mostovoy. Nat. Mater. 6, 13 (2007).

• S. Zollner, A.A. Demkov, R. Liu, P.L. Fejes, R.B. Gregory, P. Alluri, J. Curless, Z.

Yu,J. Ramdani, R. Droopad, T.E. Tiwald, J.N. Hilfiker, J.A. Woollam, J. Vac. Sci.

Technol. B 18, 2242-2254 (2000).

Page 161: Copyright By Wei Guo 2021

142

• S. A. Chambers, T. C. Kaspar, A. Prakash, G. Haugstad and B. Jalan, Appl. Phys. Lett.

108, 152104 (2016)

• S. Abel, F. Eltes, J. E. Ortmann, A. Messner, P. Castera, T. Wagner, D. Urbonas, A.

Rosa, A. M. Gutierrez, D. Tulli, P. Ma, B. Baeuerle, A. Josten, W. Heni, D. Caimi, L.

Czornomaz, A. A. Demkov, J. Leuthold, P. Sanchis. J. Fompeyrine. Nature Material,

18, 42 (2019).

• S. G. Lim, S. Kriventsov, T. N. Jackson, J. H. Haeni, D. G. Schlom, A. M. Balbashov,

R. Uecker, P. Reiche, J. L. Freeouf, and G. Lucovsky, J. Appl. Phys. 91, 4500 (2002).

• T. Choi, J. Lee, Thin Solid Films 475, 283(2005)

• T. Hadamek, D. Shin, A. B. Posadas, A. A. Demkov, S. Kwon, Q. Wang, and M. Kim,

Appl. Phys. Lett. 111, 142901 (2017).

• T. Mairoser, J. A. Mundy, A. Melville, D. Hodash, P. Cueva, R. Held, A. Glavic, J.

Schubert, D. A. Muller, D. G. Schlom, and A. Schmehl, Nat. Commun. 6, 7716 (2015).

• T. Sakudo, Ferroelectrics, 95, 37 (1989).

• T. Schumann, S. Raghavan, K. Ahadi, H. Kim and S. Stemmer. J. Vac. Sci. Technol. A

34, 050601 (2016)

• T. Wang, A. Prakash, Y. Dong, T. Truttmann, A. Bucsek, R. James, D. D. Fong, J.

Kim, P. J. Ryan, H.Zhou, T. Birol and B. Jalan, ACS Appl. Mater. Interfaces 10,

43802(2018)

• T. Wang, L. R. Thoutam, A. Prakash, W. Nunn, G. Haugstad and B. Jalan, Phys. Rev.

Mater. 1, 061601(2017)

• V. N. Strocov, T. Schmitt, U. Flechsig, T. Schmidt, A. Imhof, Q. Chen, J. Raabe, R.

Betemps, D. Zimoch, J. Krempasky, X. Wang, M. Grioni, A. Piazzalunga and L.

Patthey, J. Synchrotron Radiat. 17, 631(2010).

• V. N. Strocov, X. Wang, M. Shi, M. Kobayashi, J. Krempasky, C. Hess, T. Schmitt

and L. Patthey, J. Synchrotron Radiat. 21, 32(2014).

• W. G. Nilsen and J. G. Skinner, J. Chem. Phys. 48, 2240 (1968)

• W. Guo, A. B. Posadas, A. A. Demkov, J. Vac. Sci. Technol. A 39, 030804 (2021)

• W. Guo, A. B. Posadas, Sirong Lu, David J. Smith, and Alexander A. Demkov, J. Appl.

Phys. 124, 235301 (2018).

• W. Hiinsch, Th. Vogelsand, R. Kircher and M. Orlowski, Solid-state Electron 32,

839(1989).

• W. Li, L. Gao and A. A. Demkov, Phys. Rev. B 102, 035308 (2020).

Page 162: Copyright By Wei Guo 2021

143

• W. S. Choi, D. W. Jeong, S. S. A. Seo, Y. S. Lee, T. H. Kim, S. Y. Jang, H. N. Lee,

and K. Myung-Whun, Phys. Rev. B. 83, 195113 (2011).

• X. Hu H. Li, Y. Liang, Y. Wei, Z. Yu, D. Marshall, J. Edwards, R.Droopad, X. Zhang,

A. A. Demkov, K. Moore, and J. Kulik, Appl. Phys.Lett. 82, 203 (2003).

• X. Zhang, A. A. Demkov, H. Li, X. Hu, Y. Wei, and J. Kulik, Phys. Rev.B. 68, 125323

(2003).

• X. X. Wang, H. N. Li, W.-H. Zhang, and F.-Q. Xu, J. Phys.: Condens. Matter 19,

096001 (2007).

• Y. Horio, Y. Hashimoto, A. Ichimiya, Appl. Surf. Sci. 100, 292 (1996).

• Y. Kamihara, H. Hiramatsu, M. Hirano, R. Kawamura, H. Yanagi, T. Kamiya, and H.

Hosono, J. Am. Chem. Soc. 128, 10012 (2006).

• Y. W. Chung and W. B. Weissbard, Phys. Rev. B 20, 3456 (1979).

• Y. Wei, X. Hu, Y. Liang, D. C. Jordan, B. Craigo, R. Droopad, Z. Yu, A. A. Demkov,

J. John, L. Edwards, and W. J. Ooms, J. Vac. Sci. Technol. B 20, 1402 (2002).

• Y.L. Du, G. Chen, M.S. Zhang, Solid State Commun. 130, 577(2004)