TESIS DOCTORAL Study of ferroelectric PbTiO...

202
UNIVERSIDAD CARLOS III DE MADRID Instituto de Química y Materiales Álvaro Alonso Barba TESIS DOCTORAL Study of ferroelectric PbTiO 3 nanostructures deposited onto substrates and prepared by a novel microemulsion mediated synthesis Autor: María Torres Sancho Directoras: Dra. Mª Lourdes Calzada Coco Prof. Lorena Pardo Mata Consejo Superior de Investigaciones Científicas Instituto de Ciencia de Materiales de Madrid Leganés, noviembre de 2009

Transcript of TESIS DOCTORAL Study of ferroelectric PbTiO...

UNIVERSIDAD CARLOS III DE MADRID

Instituto de Química y Materiales Álvaro Alonso Barba

TESIS DOCTORAL

Study of ferroelectric PbTiO3

nanostructures deposited onto substrates

and prepared by a novel microemulsion

mediated synthesis

Autor:

María Torres Sancho

Directoras:

Dra. Mª Lourdes Calzada Coco

Prof. Lorena Pardo Mata

Consejo Superior de Investigaciones Científicas

Instituto de Ciencia de Materiales de Madrid

Leganés, noviembre de 2009

TESIS DOCTORAL

Study of ferroelectric PbTiO3 nanostructures deposited onto

substrates and prepared by a novel microemulsion mediated

synthesis

Autor: María Torres Sancho

Directoras: Dra. Mª Lourdes Calzada Coco

Prof. Lorena Pardo Mata

Firma del Tribunal Calificador:

Firma Presidente:

Vocal:

Vocal:

Vocal:

Secretario:

Calificación:

Leganés, de noviembre de 2009

A Dani

Esta Tesis Doctoral se ha realizado en el Departamento de Materiales para la Tecnología de la

Información del Instituto de Ciencia de Materiales de Madrid (ICMM-CSIC), bajo la dirección de la

Dra. Mª Lourdes Calzada Coco y la Prof. Lorena Pardo Mata, gracias a la concesión de una beca del

Programa de Formación de Personal Investigador (FPI) del Ministerio de Ciencia y Tecnología,

asociada al proyecto de investigación “Procesado por sol-gel de materiales nanométricos y

nanocaracterización piezoeléctrica para ferroeléctricos integrados” (MAT2004-02014). Además,

parte de este trabajo ha sido parte de las COST action 539-ELENA y de la EC network of excellence

on Multifunctional and Integrated Piezoelectric Devices NoE 515757.

Muchas son las personas que, directa o indirectamente, me han ayudado a desarrollar este trabajo

y es justo comenzar la redacción de esta tesis agradeciéndoselo.

En primer lugar, me gustaría agradecer a la Dra. Mª Lourdes Calzada Coco y la Prof. Lorena Pardo

Mata su dedicación, esfuerza y ayuda durante todo el tiempo que ha durado esta tesis. Por su

búsqueda constante de la excelencia, por su confianza en mí desde el principio y por enseñarme

tantas cosas, científicas y no científicas.

Al Dr. Jesús Ricote, por su infinita paciencia con mis dudas sobre microscopía de fuerza y por su

gran capacidad didáctica, de la que tanto he aprendido.

A los Dres. Ricardo Jiménez, Miguel Algueró, Harvey Amorín y Pablo Ramos, por interesarse

siempre por la evolución de la tesis y por sus palabras de ánimo.

Al Dr. Iñigo Bretos, por su ayuda en mis primeros y titubeantes días y su apoyo en estos últimos.

A los Dres. Miguel Ángel Martínez y Alejandro Várez, por su ayuda con los trámites necesarios para

presentar esta tesis al comienzo y al final del proceso.

I would like to thank Dr. Zhaorong Huan y Dr. Sue Impey, for accepting me in their group at

Cranfield University from November 18th to December 2nd of 2006, within the frame of the COST

action 539-ELENA. Many thanks for their help and contribution to this work, particularly to

Christine Kimpton for part of the SEM images.

Al Dr. Fuentes Cobas, por aceptarme a mí y a mis muestras en aquel viaje a Stanford. Gracias a ésa y

a sucesivas expediciones científicas, se pudieron analizar las nanoestructuras preparadas en este

trabajo mediante radiación sincrotrón. Al Dr. Fuentes Montero, por permitirme utilizar esa joya de

programa que es ANAELU: Gracias a ambos por las noches de discusión científica sobre

interpretación de resultados. Special acknowledge is due to Dr. Apurva Mehta and the Stanford

Synchrotron Radiation Lightsource for the diffraction measurements and fruitful discussion.

I would like to thank Dr. Marin Alexe, for accepting me in his group at the Max Planck Institute of

Microstructure Physics at Halle, from May 5th to August 11th 2008, within the framework of a short-

term stay in a foreign laboratory granted by the FPI program. Many thanks for so many interesting

talks about PFM, ferroelectricity and scientific life. I would like to acknowledge Dr. Brian Rodriguez

for his help, teaching and valuable discussion about PFM. I would like to thank both for the warm

care they bestowed on me during my stay. I would also like to thank the entire group for making my

stay a pleasure.

A la Dra. María Alonso del Departamento de Nanoestructuras y Superficies del Instituto de Ciencia

de Materiales de Madrid, por realizar las medidas de espectroscopía de electrones Auger y de

difracción de electrones de baja energía, por su ayuda a la hora de interpretar los resultados, así

como por sus valiosas sugerencias sobre la versión final de la discusión de dichos resultados en este

manuscrito.

Al Dr. Pedro Tartaj del Departamento Biomateriales y Materiales Bioinspirados del Instituto de

Ciencia de Materiales de Madrid,por realizar las medidas de Dynamic Light Scattering y por su

ayuda a la hora de interpretar los resultados.

A la Dra. Lidia Martínez, por realizar las medidas de ángulo de contacto y por su ayuda a la hora de

interpretar los resultados.

Al equipo de Nanotec Electrónica, por su ayuda. Especialmente, a Luís Colchero, por su paciencia,

su disponibilidad y por las muchas cosas que me ha enseñado sobre microscopía de fuerzas.

Gracias también a los Dres. Alfredo Álvarez y Pilar Suárez, con quienes empecé esta etapa de mi

vida. Gracias por darme la primera oportunidad y avivar el gusanillo de la Ciencia.

Gracias a Elvira, Lidia y Renaud, por enseñarme a afrontar los problemas con optimismo y a

relativizarlos. A Mercedes, por preocuparse siempre por mí y mi tesis. A Carlos (Carlitos), Ana, César

y Ramón, por su simpatía y por hacer de los momentos de asueto del día un momento divertido.

A Álvaro, por su paciencia con mis múltiples formateos del disco duro del ordenador y por las risas,

tan necesarias en el periodo de escritura. A los que se fueron (Alfonso, Abel y Raquel) y los que

llegaron (David y Hidham).

De todo corazón, a mis amigas María, Aurora e Irene, por perdonarme el haberme perdido

momentos muy importantes (como Pablete) y acompañarme en todas las grandes encrucijadas que

han ido llegando. Porque hemos ido creciendo juntas y siempre nos hemos apoyado.

Tengo que agradecer a mi familia por apoyar mis decisiones personales y profesionales, por estar a

mi lado pase lo que pase. Gracias por inculcarme el tesón, el perfeccionismo en su justa medida y la

eficiencia.

Y no puedo acabar sin agradecer a Dani su apoyo durante todo este periodo. Por hacerme

compañía con su presencia en las noches de escritura, por tenerme siempre la cena calentita con

una sonrisa. Por esperarme después de cada viaje, por entender que la investigación es mi vida y lo

que me hace feliz y no intentar cambiarlo. Por entender que comprometerse conmigo era también

hacerlo con esta tesis. Porque sin ti, Dani, esta tesis hubiera sido mucho más dura y ni la mitad de

divertida. Gracias.

Resumen:

Los materiales ferroeléctricos presentan una serie de propiedades que les hace

apropiados para un gran número de aplicaciones. Estos materiales presentan dos

estados de polarización de igual energía entre los que se puede transitar mediante la

aplicación de un campo eléctrico externo, convirtiéndolos en materiales muy

atractivos para su utilización en memorias RAM. Hoy en día, la tendencia de los

dispositivos de memoria es aumentar la densidad de almacenamiento, que se

encuentra actualmente en los Tb/inch2, manteniendo o disminuyendo su coste. Para

ello, es imprescindible poder fabricar unidades de almacenamiento cada vez más

pequeñas, manteniendo sus propiedades y a un coste lo más bajo posible.

En la presente tesis doctoral se ha desarrollado un método novedoso de procesado

basado en la tecnología “bottom-up” para la obtención de nanoestructuras

ferroeléctricas de PbTiO3 sobre sustratos para su uso en memorias no volátiles, NV-

FeRAMs. Este procedimiento implica el depósito sobre substratos de las disoluciones

micelares resultantes de la mezcla de soles y microemulsiones mediante la técnica de

Depósito Químico de Disoluciones (CSD).

Para comparar los resultados que se obtendrán en el transcurso de esta tesis, se han

preparado nanoestructuras explotando el fenómeno de la inestabilidad estructural de

láminas ultradelgadas. Según este fenómeno, cuando el espesor de éstas es menor de

un valor crítico, la película se rompe, dando lugar a estructuras aisladas. Se

prepararon diferentes muestras, analizando el fenómeno antes y después de este

espesor crítico, obteniéndose nanoestructuras aisladas con un límite inferior de

tamaño lateral ~50 nm. Al analizar los depósitos, no se aprecia orden sobre el

substrato. Mediante difracción de rayos-X de radiación sincrotrón en ángulo rasante,

se ha determinado la orientación de las nanoestructuras, siendo ésta de fibra y

presentando un cono de distribución direcciones de ±15°. Mediante microscopía de

fuerzas en modo piezorespuesta (PFM), se ha comprobado el carácter ferro-

piezoeléctrico a escala local de estas partículas.

En esta tesis, se ha desarrollado una novedosa tecnología de procesado para la

obtención de nanoestructuras ferroeléctricas de PbTiO3, basada en el depósito de

soluciones micelares resultantes de la mezcla de soles y microemulsiones, y en la

funcionalización de la superficie de los substratos. Se hipotetiza que las micelas

formarán una red organizada una vez depositadas sobre el sustrato, de forma que el

depósito de la disolución micelar dará lugar a una disposición ordenada sobre la

superficie del sustrato de las nanoestructuras y que, además, tendrán un tamaño y

forma controladas. Las micelas proporcionan un entorno aislado a las partículas de sol

que se encuentran en su interior, pudiendo producirse las reacciones químicas de

síntesis de los componentes. Esta característica de las micelas, sumada a su capacidad

de auto ordenación, hace de ellas las “building units” o elementos primarios para las

nanoestructuras ferroeléctricas de PbTiO3. Utilizando este método, se prepararon

nanoestructuras sobre sustratos policristalinos de Pt-(100)Si, compatibles con la

tecnología actual del silicio, y sobre sustratos monocristalinos de SrTiO3.

Sobre los sustratos policristalinos de Pt-(100)Si, se obtuvieron nanoestructuras con un

tamaño promedio de ~70 nm y con una morfología semejante a partir de disoluciones

con diferentes concentración. Estas nanoestructuras son el resultado de la

coalescencia entre un número finito de nanoestructuras primarias. Al analizar las

distribuciones de tamaños de las partículas obtenidas, se deduce que éstas crecen de

forma independiente. Esto contrasta con las nanoestructuras obtenidas mediante

inestabilidad microestructural, que siguen un mecanismo de nucleación y difusión

entre las nanoestructuras vecinas. Esta diferencia confirma la hipótesis de que las

micelas actúan de “building units” de estas nanoestructuras. Sin embargo, las

nanoestructuras preparadas mediante este procedimiento sobre los sustratos

policristalinos no se disponen ordenadamente sobre éte, debido fundamentalmente a

los defectos de la superficie del substrato. Para mejorar la calidad de esta superficie

sobre la que se hace el depósito, ésta se funcionalizó, de forma que se modificó con el

depósito previo de una película de microemulsión. Al depositar la solución micelar

sobre esta superficie funcionalizada, se obtuvieron, después de la cristalización,

nanoestructuras agregadas, como en el caso anterior, y nanoestructuras primarias,

que presentan un tamaño promedio de ~21 nm y con una disposición hexagonal sobre

el sustrato, orden de corto alcance. Las nanoestructuras presentan la estructura

cristalina de la perovskita de PbTiO3 con estructura de fibra y dos de cuyos ejes

presentan un cono de direcciones probables de ±20°, como se determinó mediante

difracción de rayos-X de radiación sincrotrón en ángulo rasante. Las medidas de PFM

confirmaron el carácter ferro-piezoeléctrico de las nanoestructuras, midiéndose en la

nanoestructura más pequeña hasta el momento, ~37 nm y altura de ~14 nm. Según

conocimiento de la autora, este tamaño de nanoestructura aislada está por debajo de

los publicados en la literatura para los que se han obtenido respuesta ferroeléctrica.

Con el objetivo de utilizar substratos con superficies más próximas a la ideal que los

policristalinos utilizados anteriormente, se utilizaron substratos monocristalinos de

SrTiO3. De esta manera, tras probar la validez del método del depósito de soluciones

micelares para la obtención de nanoestructuras primarias de tamaño y forma

controladas y con una disposición ordenada sobre el sustrato, se pretendió mejorar la

disposición ordenada de estas nanoestructuras sobre el substrato. Sin embargo, el

mojado del substrato por la disolución micelar fue muy deficiente, lo que conllevó un

recubrimiento no uniforme del substrato monocristalino. Éste comportamiento se

utilizó para determinar el tipo de crecimiento y de orden de las nanoestructuras de

PbTiO3 sobre los sustratos de SrTiO3. Se determinó que el crecimiento era de tipo

Frank-van der Merwe. Los experimentos de difracción de rayos-X de radiación

sincrotrón en ángulo rasante confirmaron la estructura perovskita del PbTiO3, así

como el crecimiento epitaxial sobre el sustrato. Mediante PFM se midió la respuesta

ferro-piezoeléctrica de las nanostructuras.

Por último, para subsanar los defectos del mojado del sustrato de SrTiO3 por la

solución micelar, se funcionalizó la superficie mediante un tratamiento químico y

térmico, de forma que al depositar la solución micelar sobre el sustrato de SrTiO3 tras

modificar su superficie, se observó la uniformidad del recubrimiento en todo el

substrato y un orden a largo alcance de las nanoestructuras de PbTiO3.

Abstract:

Ferroelectric materials present some properties that make them suitable for a large

number of applications. This materials present to states of polarization of equal

energy, switchable by an external electric field. This makes of them a very attractive

material for their use as random access memories (RAM). Nowadays, the trend in

memory devices is to increase the storage density, which is actually in the Tb/inch2,

maintaining or decreasing the fabrication cost. To achieve that, it is mandatory to be

able to fabricate smaller storage units that keep their properties and at the lowest

cost.

In this thesis, a novel processing method based in the “bottom-up” technology is

developed for the fabrication of ferroelectric PbTiO3 nanostructures onto substrates

for their use, for example, as non volatile ferroelectric RAM, NV-FeRAMs. This

procedure implies the deposition onto substrates of micellar solutions, resulting of

the mixture of sols and microemulsions by the Chemical Solution Deposition

technique (CSD).

In order to compare the results obtained in this thesis, nanostructures had been

prepared using the phenomenon of the microstructural instability of ultrathin films.

According to this phenomenon, when the thickness of an ultrathin film is below a

critical one, it breaks, yielding isolated nanostructures. Different samples were

prepared, studying the phenomenon before and after the critical thickness and

obtaining isolated nanostructures of ~50 nm of lateral size. Self-assembly of the

nanostructures onto the substrate was not observed. By synchrotron X-ray diffraction

in grazing incidence, it was possible to determine the texture of the nanostructures: it

is a fiber texture with an orientation distribution cone of ±15°. By Piezoresponse Force

Microscopy (PFM), the ferro-piezoelectric character of the nanostructures at a local

scale was proved.

In this thesis, it is proposed a novel processing technology for the fabrication of

ferroelectric PbTiO3 nanostructures, based in the deposition of micellar solutions

resulting from the mixture of sols and microemulsions and in the functionalization of

the substrates. Micelles are hypothesized to form a self-assembly network, once

deposited onto the substrate, so that the deposit of the micellar solution will rise to a

self-assembly onto the surface of the substrate with controlled size and shape. In

addition, micelles provide an isolated environment for the sol particles in their inside,

and the chemical reactions of synthesis of the components might occur. These

properties of the micelles, makes of them the “building units” for the ferroelectric

PbTiO3 nanostructures. By using this method, nanostructures were prepared onto

polycrystalline Pt-(100)Si substrates, compatible with the actual Si technology, and

onto single crystal SrTiO3 substrates.

Nanostructures were obtained onto the polycrystalline Pt-(100)Si substrates with an

average size of ~70 nm and with a similar morphology from solutions of different

concentration. These nanostructures are the result of the coalescence of a finite

number of primary nanostructures. From the analysis of the size of the obtained

nanostructures it is deduced that they grow independently, which is contrast with the

growing mechanism of the nanostructures obtained from the phenomenon of the

ultrathin films instability, which mechanism is the nucleation and diffusion between

neighbor nanostructures. This difference confirms the hypothesis of micelles acting as

“building units” of the nanostructures. However, nanostructures prepared by this

procedure onto polycrystalline substrates do not self-assemble onto the substrate,

mainly due to the defects of the surface of the substrate. In order to improve the

quality of this surface, it was functionalized by previously depositing a layer of

microemulsion. When the micellar solution was deposited onto this functionalized

surface, merged nanostructures were obtained after the crystallization process, as

before. Primary nanostructures were also obtained with an average size of ~21 nm

and a hexagonal short-range arrangement onto the substrate. The nanostructures

have the crystalline structure of the PbTiO3 perovskite with a fiber texture and two

axes that present an orientation distribution cone of ±20°, as determined by

synchrotron X-ray diffraction in grazing incidence. PFM measurements confirmed the

ferro-piezoelectric character of the nanostructures, measuring it in nanostructures of

~37 nm of lateral size and ~14 nm of height. To the best of the knowledge of this

author, the size of this isolated nanostructure is below those reported in the literature

where ferroelectric response had been measured.

Single crystal SrTiO3 substrates were used in order to utilize substrates with a surface

closer to the ideal one than the polycrystalline ones used previously. Thus, once the

validity of the microemulsion deposition procedure for the fabrication of primary

nanostructures of controlled size and shape and with a self-assembly onto the surface

of the substrate was proved, an improvement of the self-assembly onto the surface of

the substrates was set as a target. However, the wetting of the substrate by the

micellar solution was deficient, which yield a non-uniform coating of the single crystal

substrate. This behavior was exploited to determine the type of growth and

arrangement of the PbTiO3 nanostructures on the SrTiO3 substrates, establishing that

it is a Frank-van der Merwe growing type. The experiments of synchrotron X-ray

diffraction in grazing incidence configuration confirmed the PbTiO3 perovskite

structure as well as the epitaxial growth onto the substrate. By PFM, the ferro-

piezoelectric response of the nanostructures was measured.

Finally, in order to overcome the deficient wetting of the substrate by the micellar

solution, the surface was functionalized by a chemical and thermal treatment, so that

a uniform coating and a large-range arrangement of the PbTiO3 nanostructures are

observed in the whole substrate, when the micellar solution is deposited onto the

SrTiO3 substrate after modifying the surface.

Table of contents

CHAPTER 1. Introduction

1.1. From ferroelectric bulk ceramics to nanostructures. 1

1.2. State of the art and material requirements in FeRAMs. 3

1.3. State of art of the fabrication of ferroelectric nanostructures onto substrates. 4

1.3.1. The top-down approach. 5

1.3.2. Bottom-up techniques. 6

1.3.3. Hybrid methods. 8

1.4. Ferroelectric compositions of interest for FERAMs. 9

1.4.1. PbTiO3 perovskite structure. 10

1.5. Motivation and purpose of this work. 12

Bibliography 14

CHAPTER 2: Experimental Procedure

2.1. Precursor solutions. 17

2.1.1. Synthesis of the sol. 17

2.1.2.Preparation of the microemulsion. 18

2.1.3. Preparation of the micellar solution. 18

2.2. Selection of substrates.. 19

2.2.1. Pt-coated Si (100) substrates. 19

2.2.2. Microemulsion/Pt- coated Si (100) substrates. 19

2.2.3. (100)SrTiO3 substrates. 19

2.2.4. (100) SrTiO3 substrates with controlled surfaces. 20

2.3. Deposition, drying and crystallization of the PbTiO3 nanostructures. 22

2.4. Microscopy and quantitative microstructure characterization. 23

2.4.1. Optical microscopy. 23

2.4.2. Scanning Electron Microscopy. 23

2.4.3.Transmission Electron Microscopy. 24

2.4.4. Scanning Probe Microscopy. 25

2.4.4.1. Fast Fourier Transform and self-convolution images. 29

2.4.5. Image analyses 30

2.5. Structural characterization. 33

2.5.1. Synchrotron X-ray diffraction. 33

2.5.1.1. Grazing incidence. 35

2.5.2. Auger Electron Spectroscopy. 40

2.5.3. Low Energy Electron Diffraction. 42

Table of contents

2.6. Ferro-piezoelectric characterization: Piezo Response Force Microscopy. 43

2.6.1. Image acquisition. 46

2.6.2. Hysteresis loops. 47

Bibliography 49

CHAPTER 3: Ferroelectric nanostructures by the phenomenon of the microstructural

instability of polycrystalline ultrathin films.

3.1. The microstructural instability of polycrystalline ultrathin films. 53

3.2. Nanoscale PbTiO3 structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by using

the phenomenon of the microstructural instability.

55

3.2.1. Microscopy and quantitative microstructure analysis. 55

3.2.2. Structural characterization. 61

3.2.3. Functional characterization. 66

Remarks. 69

Bibliography 70

CHAPTER 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto

Pt/TiO2/SiO2/(100)Si substrates.

4.1. The microemulsion mediated synthesis 73

4.2. Nanoscale PbTiO3 ferroelectric structures onto Pt/TiO2/SiO2/(100)Si substrates

prepared by microemulsion mediated synthesis.

81

4.2.1. Microscopy and quantitative microstructure analysis. 81

4.2.2. Structural characterization 87

4.2.3. Functional characterization 91

4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion

layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated

synthesis.

98

4.3.1. Microscopy and quantitative microstructure analysis. 102

4.3.2. Structural characterization 112

4.3.3. Functional characterization 117

Remarks 121

Bibliography 123

Table of contents

CHAPTER 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion

mediated synthesis

5.1. Towards ideal surfaces 127

5.2. Nanoscale PbTiO3 structures onto commercial as-served SrTiO3 substrates prepared

by microemulsion mediated synthesis.

128

5.2.1. Microscopy analysis. 129

5.2.2. Structural characterization 138

5.2.3. Functional characterization 142

5.3. Nanoscale PbTiO3 structures onto (100)SrTiO3 substrates with controlled surfaces

prepared by microemulsion mediated synthesis.

143

5.3.1 Preparing ideal surfaces: chemical and thermal treatment of the STO surfaces. 145

5.3.2. Microscopy and quantitative microstructure analysis. 156

Remarks 162

Bibliography 164

CHAPTER 6: Conclusions

6.1. Conclusions 167

6.2. Conclusiones 169

List of figures

CHAPTER 1. Introduction

Figure 1.1. Towards the miniaturization of ferroelectric materials and their integration into microelectronic devices.

1

Figure 1.2. Top portion of 512 Mb DDR2 SDRAM stacked capacitors. HSG means hemispherical grain polysilicon.

3

Figure 1.3. Nanostructures and their hysteresis loop prepared by Focused Ion Beam (a) and (b); Nanoimprint Lithography (c) and (d); and Electron Beam Direct Writing (e) and (f).

5

Figure 1.4. Nanostructures prepared by PLD exploiting the Volmer-Webber growing mode (a), using latex microspheres (c) and alumina templates (e) and their corresponding hysteresis loops (b), (d) and (f), respectively. (g) shows nanostructures fabricated exploiting the microstructural instability of ultrathin films and (h) using di-block-copolymers.

7

Figure 1.5. Nanostructures obtained by a hybrid method that combines EBL and CSD (a) and the PFM image that confirms their ferro-piezoelectric character at a local scale (b).

9

Figure 1.6. Interrelationship of piezoelectric and subgroups on the basis of symmetry.

10

Figure 1.7. Perovskite structure of the PbTiO3 above ~490⁰C (a), below ~490⁰C (b) and representation of the polarization states, as the Ti4+ cation can occupy the schematized two stable positions along the c-axis. (c).

11

Figure 1.8. Mechanical model schematized to explain the existence of dielectric hysteresis in any ferroelectric crystal.

12

CHAPTER 2: Experimental Procedure

Figure 2.1. Synthesis of the sol.

17

Figure 2.2. Preparation of the microemulsion.

18

Figure 2.3. Preparation of the micellar solution.

18

Figure 2.4. Thermal treatments in UHV applied to the STO substrates.

20

Figure 2.5. RTP thermal treatments applied to the STO substrates.

21

Figure 2.6. Spin-coating process.

22

Figure 2.7. Thermal recipe used for the crystallization of the nanostructures.

23

Figure 2.8. Experimental set-up of a SPM system.

26

Figure 2.9. Force-distance curve of the interaction between the probe tip and the sample surface.

26

Figure 2.10. Artifacts of the AFM topography image in contact mode in the case of tip radius smaller (a) and larger (b) than the measured nanostructures.

27

List of figures

Figure 2.11. Resonant frequency shift when the probe is affected by attractive or repulsive forces.

28

Figure 2.12. Nanosensors probes for the acquisition of AFM topography images.

29

Figure 2.13. Original sample (a), image after applying the FFT filter (b) and self-correlated image (c).

30

Figure 2.14. Study the porosity reduction and microstructure recovery from the degraded sintering stage by subsequent Hot isostatic pressing (HIP) (a) and the study the limit of the increase in grain size without microstructure degradation by recrystallization of hot pressed samples in Aurivillius ceramics (b).

30

Figure 2.15. Example of the image analysis process previous to the measurement of the nanostructures onto a substrate. (a) is the original image, (b) the binary image and (c) shows the identification by MIP4 of each nanostructure by coloring them.

31

Figure 2.16. Ewald sphere 2-dimensional representation. Green points are the reciprocal

lattice ones.

35

Image 2.17. Schematic illustration of the grazing incidence geometry used for X-ray diffraction.

37

Figure 2.18. Schematical representation of the interaction of an incident X-ray beam on a crystal and the resulting diffraction peaks in the 2-dimensional plate for different planes of the reciprocal lattice.

38

Figure 2.19. Diagram flow of the iterative procedure for the lattice parameters (a) and texture of the nanostructures (b).

39

Figure 2.20. 2-dimensional X-ray transmission pattern of LaB6 (a) and the profile along the yellow line of previous image (b). LaB6 is the standard used for calibration.

40

Figure 2.21. Schematic illustration of the Auger process, indicating the energy levels involved.

41

Figure 2.22. Conventional LEED system.

43

Figure 2.23. PFM configuration for the obtaining of PFM images and hysteresis loops.

44

Figure 2.24. Nanosensors probes for the acquisition of PFM images.

44

Figure 2.25. Scheme of the out-of-plane, in-plane and mixed response of a ferroelectric material when applying an electric field perpendicular to the surface of the nanostructure.

45

Figure 2.26. Piezoresponse, amplitude and phase signals for different cases of electromechanical and electrostatic interactions.

46

Figure 2.27. Shape of the electric field supplied to the substate for the in-field (a) and out-of -field (b) local piezoelectric hysteresis loops. AC field is represented in dark red and DC field in green.

47

CHAPTER 3: Ferroelectric nanostructures by the phenomenon of the microstructural instability of polycrystalline ultrathin films.

Figure 3.1. Spheroidization of a uniform 2-dimensional ultrathin film of initial thickness t and grain size D (left), as the thermal energy provided to the system increases , and its 3-

53

List of figures

dimensional analysis (right) Figure 3.2. Nanoparticles of Pb(Ti, Zr)O3 (PZT) onto SrTiO3 volume distribution.

54

Figure 3.3. AFM topography images of a PbTiO3 continuous ultrathin film obtained at different magnifications: 5x5 µm image (a) and 2x2 µm image (b).

56

Figure 3.4 AFM topography images and representative profiles of the PbTiO3 nanostructures onto the Pt/TiO2/SiO2/(100)Si substrates prepared from sols with different concentrations.

57

Figure 3.5. Proposed growth evolution of the particles deposited as the concentration of the solutions decreases.

58

Figure 3.6. Equivalent diameter distributions of samples prepared from the 4·10-2

M (a) and 3·10

-2 M (b) sol-gel solutions and their corresponding log-normal distributions (c) and (d).

60

Figure 3.7. Experimental 2-D Synchrotron diffraction pattern of a continuous ultrathin film.

62

Figure 3.8 2θ diffraction pattern obtained from the integration of the 2-D experimental

pattern and the simulated diffraction patterns of the PbTiO3, platinum, TiO2 and silicon of the

substrate.

63

Figure 3.9. Simulated 2-D diffraction pattern of polycrystalline platinum with (111) fiber texture (a) and the experimental 2-D diffraction pattern measured from the ultrathin film (b).

64

Figure 3.10. Simulated 2-D diffraction pattern of (100) Si single crystal (a) and the measured 2-D diffraction pattern from the sample (b). The diffractions maxima that forms the triangle are marked with a cross.

64

Figure 3.11. Simulated 2-D diffraction pattern of PbTiO3 nanostructures with (100) fiber

texture and an orientation distribution cone of ±15° (a) and the measured 2-D diffraction

pattern from the sample (b).

65

Figure 3.12. Topography (a) and (d), phase (b) and amplitude (c) images of the in-plane piezoresponse and phase (e) and amplitude (f) of the out-of-plane piezoresponse for the continuous ultrathin PbTiO3 film.

67

Figure 3.13. Topography (a), phase (b) and amplitude (c) images of the out-of-plane piezoresponse for the PbTiO3 nanostructures.

67

Figure 3.14. Out-of-field consecutive local hysteresis loops of a nanoparticle of 50nm.

68

Figure 3.15. Scheme of the pinned layer and the imprint (a) and its effect on a hysteresis loop (b).

68

CHAPTER 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

Figure 4.1. Representation of Brij-30 with the head and tail groups indicated.

73

Figure 4.2. Some of the possible colloidal aggregates of surfactant molecules in an emulsion.

74

Figure 4.3. Schema of the inverse micelles inside the microemulsion prepared in this work.

75

List of figures

Figure 4.4. Schematic formation process of the micellar solution. 76

Figure 4.5. Hypothetical self-assembly of the nanostructures.

78

Figure 4.6. Tyndall effect in the microemulsion (a), sol (b) and micellar solution (c).

79

Figure 4.7. DLS measurements for the microemulsion (a), sol (b) and micellar solution (c).

80

Figure 4.8. Topography by optical microscopy of the limit of the coating area (a) and SEM micrograph of the center of the non-homogenous coating (b) (sample prepared from a 5·10

-

3 M micellar solution).

82

Fig. 4. 9. AFM topography images of the PbTiO3 nanostructures onto Pt/TiO2/SiO2/(100)Si substrates prepared from the micellar solutions with a sol concentration of 10

-2 M (a-b) and

5·10-3

M (c) at different locations of the coated substrate.

83

Figure 4.10. High magnification AFM image of isolated nanostructures fabricated from the 5·10

-3 M micellar solution.

84

Figure 4.11. Proposed drying evolution with time of the micellar layer (a, b) and oxide nanostructures formed after the thermal treatment of crystallization (c).

85

Figure 4.12 Equivalent diameter distributions of nanostructures prepared from the 10-2

M (a) and 5·10

-3 M (b) micellar solution.

86

Figure 4.13. Experimental 2-D synchrotron x-ray diffraction pattern of a sample prepared from a 5·10

-3 M micellar solution.

88

Figure 4.14. 2θ diffraction pattern calculated from the integration of the 2-D experimental pattern of Fig. 4.13 (black solid line) and simulated diffraction patterns of the PbTiO3 perovskite nanostructures (red line), Pt bottom electrode (blue line), TiO2 anatase (green line) and Al holder (orange line).

89

Figure 4.15. Topography (a) and (d), phase (b) and (e) and amplitude (c) and (f) images of the

in-plane and out-of-plane piezoresponse, respectively, for a sample prepared from the micellar solution with a 10

-2 M concentration.

92

Figure 4.16. Topography (a), phase (b) and amplitude (c) out-of-plane piezoresponse images for the isolated nanostructures of a sample prepared from the micellar solution with a 5·10

-3

M concentration.

93

Figure 4.17. Topography (a), phase (b) and amplitude (c) images of the domain structure of a big nanostructure (~200 nm of lateral size).

94

Figure 4.18. Out-of-field local hysteresis loops of isolated nanostructures of ~95 nm of lateral size(phase (a) and amplitude (b) loops) and ~83 nm of lateral size(c) (four consecutive piezoresponse loops) isolated nanostructures fabricated from the 5·10

-3 M micellar solution.

95

Figure 4.19. Proposed switching mechanism for the isolated nanostructures.

96

Figure 4.20. Schematic representation of the bending of the energy bands for a conductor|p-type semiconductor contact.

97

Figure 4.21. Drops of (a) sol, (b) microemulsion and (c) micellar solution onto a Pt/TiO2/SiO2/(100)Si substrate.

99

Figure 4.22. Drops of (a) the micellar solution and (b) the sol onto the 100

List of figures

microemulsion layer/Pt/TiO2/SiO2/(100)Si substrate.

Figure 4.23. Proposed drying evolution with time of the modified procedure where the micellar solution containing the building units is deposited onto a microemulsion layer/Pt/TiO2/SiO2/(100)Si substrate ((a) and (b)) and resulting oxide nanostructures formed after the thermal treatment of crystallization (c).

101

Figure 4.24. Topography by optic microscopy of a sample prepared using the modification of the microemulsion mediated synthesis method.

102

Figure 4. 25. SEM image of a sample prepared by this procedure with x5000 (a) and x100000 (b) magnification.

103

Figure 4.26. Topography AFM images of a sample prepared by the modified microemulsion mediated synthesis. (a) image shows a 3x3 µm area and (b) a 1x1 µm one.

104

Figure 4.27. AFM topography images of two different locations in the substrate and their corresponding self-correlations, (a) and (b).

105

Figure 4.28. The five fundamental 2-dimensional Bravais lattices.

105

Figure 4.29. Model of location of sol particles in reverse micelles (black ball represent the sol nanoparticle.

106

Figure 4.30. Equivalent diameter distributions of the nanostructures measured on the SEM image of Fig. 4.25 (a) and the AFM image of Fig. 4.26 (a).

107

Figure 4.31. Bright-field TEM images of cross section of an isolated nanostructure (a), three nanostructures formed by the coalescence of primary ones (a-c) and simulated primary nanostructures disposition (d) that yield the nanostructure in picture above it.

109

Figure 4.32. Bright field TEM image of the cross section of an isolated primary nanostructure.

110

Figure 4.33. High magnification bright-field HRTEM images of the inside of the nanostructures (a) (101) planes, (b) (101) planes in adjoining parts of the nanostructure with a relative tilt of 17.5 ° and (c) edge dislocation, marked with an arrow.

111

Figure 4.34. Experimental 2-D synchrotron x-ray diffraction pattern of a sample prepared from the 5·10

-3 M micellar solution.

113

Figure 4. 35. 2θ diffraction pattern calculated from the integratiion of the 2-D experimental pattern of Fig. 4.34 (black solid line) and simulated diffraction patterns of the PbTiO3 perovskite nanostructures (red line) and Pt bottom electrode (blue line).

114

Figure 4.36. Experimental 2-D diffraction pattern with reflections corresponding to Pt (a) and PbTiO3 perovskite (b) and the simulated 2-D diffraction pattern of polycrystalline platinum with (111) fiber texture (c) and PbTiO3 nanostructures with (100) fiber texture.

115

Figure 4.37. Topography (a), phase (b) and amplitude (c) out-of-plane piezoresponse images for the isolated nanostructures of a sample prepared by the modified microemulsion synthesis method. Image (d) corresponds to a high magnification image ; its phase profile (e) are marked in blue in the phase image.

118

Figure 4.38. Piezoresponse hysteresis loop obtained in the nanostructure, which phase profile is shown in Fig. 4.37 (e). Its lateral size is ~37 nm and its height is ~15 nm as measured from the images of Fig. 4.37.

119

List of figures

CHAPTER 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis.

Figure 5.1. AFM topography images of an as-served SrTiO3 substrate (a-b) and profile (d) along the blue line of image (c).

128

Figure 5.2. Optical image of the surface of a sample prepared onto a commercial and as-received SrTiO3 substrate.

129

Figure 5.3. SEM images of the morphology of the sample prepared from the 5·10-3

M micellar solution onto an as-received STO substrate in dependence of the thickness of the coating along the substrate surface.

130

Figure 5.4. SEM image of the morphology of a coating of a block-copolymer film.

131

Figure 5.5. Schematic theoretical cross-section views of the three modes of thin film growth.

133

Figure 5.6. AFM topography images of three different zones of the sample prepared from the 5·10

-3 M micellar solution.

135

igure 5.7. Low magnification AFM topography image of zone III-b of Fig. 5.3 (a) and two representative zoom areas where square truncated pyramids (b) and rectangular ones (c) can be found and their profiles along the blue lines marked in the AFM images.

136

Figure 5.8. Samples prepared from the 10-2

M micellar solution at 650°C and re-crystallized at 850°C (a-b) and 1050°C (c-e). Image (b) shows a higher magnification image of (a). Images (d) and (e) are higher magnification images of (c) and are presented here to show details of the merged nanostructures and the modification of the substrate with temperature.

137

Figure 5.9. Experimental 2-D synchrotron x-ray diffraction pattern of a sample prepared from the 5·10

-3 M micellar solution onto as-received STO substrate.

139

Figure 5.10. 2θ diffraction pattern calculated from the integration of the 2-D experimental pattern of Fig. 5.9 (black solid line) and simulated diffraction patterns of the PbTiO3 perovskite nanostructures (red line) and SrTiO3 substrate(blue one).

140

Figure 5.11. Simulated 2-D diffraction pattern of single crystal SrTiO3 with (001) fiber texture (a) and PbTiO3 nanostructures with (001) texture (b) . The experimental 2-D diffraction patterns with reflections corresponding to the SrTiO3 and the PbTiO3 perovskite phases are shown in (c) and (d), respectively.

141

Figure 5.12. Out-of-field local hysteresis loops of isolated nanostructures of ~40 nm of lateral size (phase (a) and amplitude (b) loops) and truncated squared based pyramid of ~400 nm of lateral size (phase (c) and amplitude (d) loops).

142

Figure 5.13. Surface energy for TiO2 and SrO termination as a function of the TiO2 chemical potential (the bulk reference is set to rutile and the zero chemical potential corresponds to the TiO2 rich condition.

144

Figure 5.14. LEED patterns of the surface of substrate 1 -blue border: (a), (c) and (f)-, substrate 2 (b), substrate 3 (d), substrate 4 (e) and substrate 5 (g). Pattern with green border is from the sample prepared using the pH 4.5 BHF etching solution and those with red border are from the substrates soaked in the pH 5.5 BHF solution.

150

List of figures

Figure 5.15. AES spectra of the surface of substrate 1 (black line), substrate 4(blue line) and substrate 5 (red line).

151

Figure 5.16. AFM topography image (a) and profile (b) along the blue line in the topography image of substrate 2.

154

Figure 5.17. AFM topography image of substrate 5 (a) and the profile of an etch pit (b).

155

Figure 5.18 AFM topography images of an as-served SrTiO3 substrate (a) and substrate 5 (b). Note that the measured area is 1x1 µm

2 in both images.

155

Figure 5.19. Distribution of the height of the steps of the SrTiO3 terraces, measured on the Fig. 5.17 (a).

156

Figure 5.20. Optical micrograph of the sample prepared onto a chemically etched and thermally treated substrate: i.e. substrate 5 from previous section.

157

Figure 5.21. AFM topography images of the crystalline PbTiO3 nanostructures onto the substrate 5 at different locations along one of the diagonals of the 10x10 mm

2 substrate after

chemical etching and crystallization of the substrate.

158

Figure 5.22. AFM topography image (a), the corresponding image when the substrate is subtracted (b) ,the profile (c) along the blue line in (b) and the FFT of image (a) for the crystalline PbTiO3 nanostructures prepared onto substrate 5.8.

159

Figure 5.23. Equivalent diameter distributions of the nanostructures prepared onto a chemically and thermally treated substrate 5.

160

CHAPTER 1: INTRODUCTION

1.1. From ferroelectric bulk ceramics to nanostructures.

Ferroelectrics are a type of multifunctional materials that have a crystallographic axis along

which a spontaneous polarization (Ps) exists in the polar phase, consequence of a non-

centrosymmetric arrangement of ions in its unit cell that produces an electric dipole moment

in the material. This spontaneous polarization can be switched by applying an external electric

field.

Since the Second World War, when the first applications for BaTiO3 ferroelectric perovskite

were found [1], ferroelectrics have been used in a wide range of applications. Figure 1.1

illustrates the scaledown in the order of magnitude of the ferroelectrics historically.

Figure 1.1. Towards the miniaturization of ferroelectric materials and their integration into

microelectronic devices [2].

After the Second World War, the technical exploitation of ferroelectric bulk materials began,

based on interesting properties such as a large dielectric permittivity, which enables the

storage of high charge for an applied potential, high piezoelectric and pyroelectric coefficients,

for their use as actuators, sensors and transducers, and non-linear optical effect, that can give

rise to the exchange of energy between a number of optical beams at different frequencies.

At the beginning, ferroelectric materials were mainly fabricated as single crystals or bulk

ceramics. Although the first studies on ferroelectricity were carried out on single crystals [3],

the deal of research on these materials from the point of view of their applications seemed to

decrease due to the difficulties that the preparation of high-quality single crystals involves and

also to the thigh cost related to the crystal growth techniques.

BULK

CERAMICS

THIN

PLATES

THICK

FILM

THIN

FILM

ULTRA-THIN

FILM

NANO-SIZED

SYSTEMS

1 mm 100 m 10 m 1 m 100 nm 1 nm

NANO-TECHNOLOGY

1950 1970 1990 >2000

2 1.1. From ferroelectric bulk ceramics to nanostructuress

By far, the largest number of applications using ferroelectric materials has been carried out on

bulk ceramic form. As example, ferro-piezoelectric bulk ceramics show a wide range of

applications including, for instance, monolithic multilayer capacitors (MLC), piezo motors

(buzzers, loud speakers, actuators), piezo-generators (accelerometers, power supplies,

sonsors), pressure sensors (sonar, medical ultrasounds) and resonant devices (ultrasonic

cleaners, surface acoustic wave filters) [4, 5].

The use of ferroelectric materials as integrated devices with the semiconductor circuit

technology implies the preparation of the materials onto substrates. Thus, the preparation of

high charge capacitors by using thick film (2-20 µm) technology allows the miniaturization of

ceramics that were used in electro-optic and some piezoelectric devices [5].

However, the ‘70s witnessed the evolution of the electronic industry towards the

miniaturization of the electronic components, which led to the development of new thin films

(0.2-2 µm) deposition techniques for ferroelectric oxides. Obviously, thin films are better

integrated into the as-low-as-possible scale of microelectronic devices rather than bulk

materials. Ferroelectric thin film devices perform the same electronic functions with only a

fraction of the volume of devices based on bulk ceramics or single crystals elements.

Furthermore, films are processed at temperatures of several hundred of degree Celsius lower

than those used for sintering bulk ceramics, which can be a deciding factor in their

applicability. Finally, thin films are convenient integrated with the semiconductor integrated

circuit technology, showing additional benefits such as lower operating voltages, higher

writing, reading and access speeds and micro-level designing, and, therefore, areas of

applications have been identified for these materials (non-volatile memories, micro-actuators,

etc) [5, 6].

At present, we are in an era of reduction to the nanoscale of the ferroelectric materials,

needed, for example, to increase the storage density while maintaining the size and shape of

the device, and their applications. After the films, ultrathin films, whose reduction in size to

the nanoscale takes place in one dimension, are under study [7-9] and there is increasing work

on ferroelectric systems reduced in size in two dimensions (nanorods [10, 11] or nanowires

[12, 13]) and all three dimensions (grains in nanostructured ceramics or isolated

nanostructures).

Chapter 1. Introduction 3

Nanostructures are mainly used for ultra high density RAM memories. The state of art of

fabrication of these memories and the ferroelectric nanostructures onto substrates is

summarized in next sections.

1.2. State of the art and material requirements in FeRAMs.

Random access memories (RAMs) memories are a kind of temporarily memories that store

dynamic data for devices such as computers. The word random refers to the fact that any

piece of data can be returned in a constant time, regardless of its physical location and

whether or not it is related to the previous piece of data, unlike hard disks [14].

The drive for smaller and more powerful devices is demanding an improvement in

performance at low cost. For the past decades, scaling was relied to increase capacity at the

same or reduced cost [15]. However, shrinkage in lateral feature size while maintaining the

stability of the single storage units becomes increasingly problematic [16].

Non-volatile random access memory (NVRAM) is a general name used to describe any type of

memory that does not lose information after switching off. This is in contrast to the most

common known RAMs: dynamic random access memories (DRAMs) and static random access

memories (SRAMs) which both require continual power in order to maintain their data. It is

used, for example, to store the BIOS (Basic Input-Output System ), which is a boot firmware,

designed to be the first code run by a PC when powered on. NVRAMs have fast reading,

writing and accessing time, unlike their main competitors: the flash memories.

Figure 1.2. Top portion of 512 Mb DDR2 SDRAM stacked capacitors. HSG means hemispherical grain

polysilicon. The capacitors present a rough surface. This increases the area with respect to previous

technology, increasing the storage charge for the same thickness and material [17].

4 1.2. State of the art and material requirements in FeRAMs

One group of materials that is used as storage mean for RAMs are ferroelectric materials. The

main feature of a ferroelectric material is that the sign of this spontaneous polarization can be

reversed (switched) by applying a suitable electric field. These two stable states, +P and –P,

can be used to encode the 1 and 0 Boolean algebra that forms the basis of memory and logic

circuitry in all modern computers [14]. The main advantages of using ferroelectric materials for

RAMs are the large dielectric constant (ε = 100 to 1000). In the NVRAM, the ferroelectric

polarization contains the stored information, whereas in a ferroelectric DRAM, the

ferroelectric material is merely a high-dielectric capacitor and can have a null polarization

vector.

Nanostructures usable as storage units and capable of being integrated into high-density

device architecture have to fulfill a number of properties: they must present two different and

switchable states, be periodically ordered in large areas and their size and shape must be

uniform.

Current commercially available non volatile ferroelectric random access memories (NV-

FeRAM) are fabricated with 130 CMOS process technology. CMOS stands for Complementary

Metal Oxide Semiconductor and it refers to both a particular style of digital circuitry design,

and the family of processes used to implement that circuitry on integrated circuits. The chip

realizes storage of 128 Mbytes and reading and writing speeds of 1.6 Gbytes/second and the

cell size is 0.254 µm2 [18] which is significantly smaller than the previous highest density

memories that were 0.719 µm2 that enabled a 64 Mbytes store capacity [19]. However,

shrinkage in lateral size maintaining the properties of the material becomes increasingly

problematic

1.3. State of art of the fabrication of ferroelectric

nanostructures onto substrates.

There are three different approaches to the fabrication of ordered nanostructures onto

substrates: the top-down, the bottom-up and the hybrid approaches. The top-down methods

are those based on carving thin or ultrathin films, the bottom-up ones are based on building

the nanostructures from the bottom, using atoms and molecules and promoting their self-

assembly; and last, the hybrid techniques are formed by the combination of a top-down and a

bottom-up technique. This section offers a brief summary of the main techniques of each

category as well as the most promising results obtained for each one and their advantages and

disadvantages [20, 21].

Chapter 1. Introduction 5

1.3.1. The top-down approach.

Top-down methods are based in lithographic techniques such as Focused Ion Beam (FIB),

Nanoimprint Lithography (NIL) or Electron Beam Direct Writing (EBDW).

In the fabrication of nanostructures by FIB, an ion beam is focused on a particular location and

the edges of the future nanostructures are carved until it stands isolated from the rest of the

thin or ultrathin film. Ganpule et al. [22] were able to fabricate nanostructures of

Pb0.1(Nb0.04Zr0.28Ti0.68)O3 down to 70 nm of lateral size using FIB. They found that the

ferroelectric response in these small nanostructures was the same as the larger nanostructures

(above 1 µm). A scanning electron microscopy (SEM) image of these nanostructures prepared

by FIB with different lateral sizes and the ferro-piezoelectric hysteresis loop of one of 70 nm of

lateral size is presented in Fig. 1.3 (a) and (b).

Figure 1.3. Nanostructures and their hysteresis loop prepared by Focused Ion Beam (a) and (b);

Nanoimprint Lithography (c) and (d); and Electron Beam Direct Writing (e) and (f).

Harnagea et al. [23] prepared Pb(Zr, Ti)O3 structures below 300 nm of lateral size using

nanoimprint lithography combined with metal organic deposition (MOD) and sol-gel chemistry

a) b)

c) d)

e) f)

6 1.3. State of the art of the fabrication of ferroelectric nanostructures onto substrates

(Fig. 1.3 (c)). In nanoimprint lithography, a precursor film is deposited onto a substrate. Then,

using a mold, a thickness contrast is created in the film. The resulting structures are then

crystallized. These structures presented ferro-piezoelectric response (Fig. 1.3 (d)). Scaling

down the 300 nm limit was not possible because of the sticking of the precursor solution to

the mold and lost of cell shape during crystallization.

Electron beam direct writing is based in irradiating selected areas of a precursor film. The non-

irradiated areas are removed and the resulting nanostructures are crystallized. Nanostructures

down to 100 nm of lateral size were prepared following this procedure by Alexe et al [24, 25]

(Fig. 1.3(e)). They show ferroelectric response as can be observed in Fig. 1.3(f) for a PZT

nanostructure of 100 nm of lateral size.

These techniques offer a high control of the structures lateral size and periodic arrangement.

However, as they are based in carving films, quality of the surface of the nanostructures is

compromised, making difficult to obtain nanostructures with adequate ferroelectric properties

below 100 nm of lateral size. In addition, they need expensive equipments, representing a

higher cost than the next group of techniques.

1.3.2. Bottom-up techniques.

Pulsed laser deposition (PLD) has given raised to a number of different methods for the

preparation of ferroelectric nanostructures onto substrates. PLD is a versatile thin film

technology [26] and the methods are based either in the growing characteristics of these films

or in depositing using a mask.

A large number of films grow following the Volmer-Webber mode, which is defined by the

growing of 3-dimensional islands on the substrate [27, 28]. Therefore, controlling the

deposition conditions, isolated nanostructures can be fabricated. Alexe et al. reported the

fabrication of nanostructures of Bi4Ti3O12 of 150 nm of lateral size using this method [29]

(Fig. 1.4. (a) and (b)). Ferroelectric response has been reported for a nanoscale capacitor array

consisting of a 180nm thick Bi4Ti3O12 film.

The most promising results for nanostructures obtained by PLD using masks were achieved

when using latex nanospheres [30] (Fig. 1.4. (c) and (d)) and anodized alumina masks [31]

(Fig. 1.4. (e) and (f)). In the first case, well-ordered arrays of pyramid-shaped ferroelectric

BaTiO3 nanostructures of ~350 nm of lateral size are obtained at the empty places left

between the nanospheres. In the case of PLD using anodized alumina masks, Lee et al. [31]

Chapter 1. Introduction 7

prepared well-ordered areas of ferroelectric Pb(Zr0.20Ti0.80)O3 nanostructures of 64 nm of

lateral size.

Figure 1.4. Nanostructures prepared by PLD exploiting the Volmer-Webber growing mode (a), using

latex microspheres (c) and alumina templates (e) and their corresponding hysteresis loops (b), (d) and

(f), respectively. (g) shows nanostructures fabricated exploiting the microstructural instability of

ultrathin films and (h) using di-block-copolymers [29-33].

Another bottom-up method for the preparation of ferroelectric nanostructures onto

substrates exploits the microstructural instability of ultrathin films (Fig. 1.4. (g)). Further

explanations and discussion on the state of art about this method will be done in the

a) b)

c) d)

e) f)

g) h)

8 1.3. State of the art of the fabrication of ferroelectric nanostructures onto substrates

introduction section of Chapter 3, in order to illustrate the advantages of the method

suggested in this work.

In the last decade, a new approach, based in the use of di-block copolymers, has provided

promising results (Fig. 1.4. (h)). These results are further discussed in Section 4.1.

Images of oxide nanostructures using the phenomenon of the microstructural instability of

ultrathin films and di-block copolymers are shown in Fig. 1.4 (g) and (h), respectively. PZT

nanostructures prepared by the phenomenon of the microstructural instability of ultrathin

films present a lateral size of ~50 nm [33], while the one prepared by using di-block

copolymers are a SrTiO3 perovskite and have a lateral size of ~25 nm [32].

Kronholz et al. [34] fabricated PbTiO3 nanograins of ~30 nm of lateral size onto predefined TiO2

nanostructures created on Pt/TiO2/SiO2/(100)Si substrates by using a self-organized template

with the aid of PS-b-PVC di-block copolymer micelles. This results into an incipient order of the

PbTiO3 nanostructures. No proof of the ferroelectric behaviour of the isolated structures is

reported.

All these bottom-up methods provided smaller lateral size nanostructures than the top-down

approaches summarized previously, at a lower cost. However, except when using templates

for PLD, it is difficult to obtain a periodic 2-dimensional arrangement of the nanostructures

onto the substrates and to achieve a uniform size for all the nanostructures onto the substrate.

3.3. Hybrid methods.

Hybrid methods are based in fabricating an array of seeds (usually TiO2) by a top-down

procedure. These seeds act as preferential nucleation points onto which the nanostructures

will be grown by a bottom-up procedure.

In lithography-modulated self-assembly [35], a thick TiO2 film is deposited and then carved by

electron beam lithography (EBL). After removing the mask, deposition of Pb(Zr0.4Ti0.6)O3 was

carried out by an in situ sputtering process. This way, ferroelectric structures of ~250 nm of

lateral size were fabricated.

Clemens et al. [36] also prepared the TiO2 seeds by EBL. After removing of the mask, the

platinized substrates with the seed patterns were thermally annealed to crystallize them.

Then, PbTiO3 precursor solution was deposited by spin-coating. The resulting nanostructures

were pyrolized and, subsequently, crystallized. The average lateral size of the resulting PbTiO3

Chapter 1. Introduction 9

nanostructures was ~50 nm and they showed ferro-piezoresponse at a local scale (Fig. 1.5 (a)

and (b), respectively).

Figure 1.5. Nanostructures obtained by a hybrid method that combines EBL and CSD (a) and the PFM

image that confirms their ferro-piezoelectric character at a local scale (b) [36].

1.4. Ferroelectric compositions of interest for FERAMs.

A ferroelectric can be defined as a crystallographically non-centrosymmetric dielectric, with a

phase transition from a polar (ferroelectric phase) to a non-polar state (paraelectric phase) at a

temperature at which its dielectric constant shows a maximum value. Ferroelectrics lose their

ferroelectric properties above this temperature, known as the Curie temperature.

Figure 1.6 shows the interrelationship between piezoelectrics and subgroups on the basis of

internal crystal symmetry. Four types of ferroelectrics crystal structures have been also

introduced as subcategories of the general group of ferroelectric materials (based on its unit

cell structure) [5]. The most relevant group is the perovskite (ABO3) crystal structure category

because PbTiO3 in its perovskite phase is the structure used in this thesis.

Piezoelectric crystals, such as the ferroelectric crystals are, polarized upon the application of a

mechanical stress [37]. This is known as direct piezoelectric effect. However, as far as the

ferro-piezoelectric characterization of the ferroelectric nanostructures is concerned, only the

inverse piezoelectric effect is of relevance. In this last case, there is a mechanical movement

generated by the application of an electrical field.

The piezoelectric deformation constants couple the electric displacement with the mechanical

stress and strain with the electric field strength, respectively. They are defined as dij where the

a) b)

10 1.4. Ferroelectric compositions of interest for FeRAMs

first subscript refers to the direction of the electric field and the second one to the direction of

the mechanical stress or deformation. Thus, d33, d31 and d15 are defined, respectively, as the

longitudinal coupling factor (the displacement occurs in the same direction than the field), the

transverse coupling factor (the displacement is perpendicular to the field) and the shear

coupling factor (the shear is perpendicular to the field) [38].

Figure 1.6. Interrelationship of piezoelectric and subgroups on the basis of symmetry [5].

Pyroelectricity is the alteration of the spontaneous polarization with temperature and is a

symmetry property of crystals. Ferroelectric materials are, thus, piezoelectric and pyroelectric

materials which spontaneous polarization can be reverse by applying an electric field (coercive

field).

1.4.1. PbTiO3 perovskite structure.

The ferroelectric material used in this work is lead titanate (PbTiO3). It presents a perovskite

structure, as shown in Fig. 1.7.

At high temperatures (above ~490⁰C) it features cubic centrosymmetric phase, where the Pb2+

cations occupy the corner of the unit cell, the Ti4+ one, the central position and the O2- ions the

faces of the cell, forming an octahedron.

At lower temperatures (below ~490⁰C), PbTiO3 presents a tetragonal phase, where the c

(vertical) lattice parameter is elongated. The unit cell parameters of this phase are

a = b = 3.899(9) Å and c = 4.150(0) Å and its tetragonal distorsion c/a = 1.066(8), ~6.7%. In this

phase, there are two possible stable positions for the Ti4+ cation, both distortions along the c

(vertical) axis from the central position: above it or underneath it. As the center of the charges

32 symmetry point groups

21Noncentrosymmetric

11Centrosymmetric

20Piezoelectric

10Pyroelectric

SubgroupFerroelectric spontaneously

polarizedPolarization reversible

TungstenBronze

Oxygen octahedralABO3

Pyrochlore Layer structure

Chapter 1. Introduction 11

does not coincide with the center of the unit cell, there is a permanent dipole moment.

Therefore, these two positions can be related to the up and down polarization vector.

Figure 1.7. Perovskite structure of the PbTiO3 above ~490⁰C (a), below ~490⁰C (b) and representation of

the polarization states, as the Ti4+

cation can occupy the schematized two stable positions along the c-

axis. (c).

In the non-symmetric phase, two stable configurations are possible for the distribution of

charges in the primitive cell. Due to the small deformation, these configurations are separated

by a relatively small energy barrier. Therefore, a small electric field is enough to make one of

these configurations completely stable, while the other is unstable (process explained in Fig.

1.8 (a)-(d)). Thus, the dependence of the spontaneous polarization of a ferroelectric crystal

with the electric field has the shape shown in Fig 1.8 (e). This is general for any ferroelectric

material [39].

Lead titanate perovskite (PbTiO3) was chosen as material of study in this thesis due to the fact

that it does not exhibit phase transition in a wide range of temperature, its large tetragonality,

its high spontaneous polarization at room temperature (Ps > 50 μC/cm2) and the wide

temperature stability of the ferroelectric phase and its high Curie temperature (490 ⁰C). Both

polarization and Curie temperature are the highest among perovskite type structure

ferroelectric material [40]. As we are decreasing the size of the structures to the nanoscale, it

is expected the reduction of the properties of the material together with difficulties for its

characterization and, thus, the better the bulk properties for the material, the better we can

expect to study the ferroelectric nature of the nanostructures.

12 1.4. Ferroelectric compositions of interest for FeRAMs

Figure 1.8. Mechanical model schematized to explain the existence of dielectric hysteresis in any

ferroelectric crystal. (a) and (c) are the two stable states in the absence of an external electric field. The

potential well under electric fields of opposite directions is shown in (b) and (d). (e) represents the

hysteresis loop of a ferroelectric crystal under an external electric field, being the positions

corresponding to the (a)-(d) energy states marked with a red dot. The discontinuous lines mark the

transition from (a) and (b) states to (c) and (d) states (adapted from [39]).

1.5. Motivation and purpose of this work.

Since ferroelectricity is a collective phenomenon, certain minimum number of unit cells is

required. Therefore, the transition from the ferroelectric phase to a non-ferroelectric phase

must occur if the volume of the nanostructures is decreased beyond a limit [41]. A second

problem is that, as small volume nanostructures, the number of unit cells in the surface or the

interface between the nanostructure and the substrate below with respect to the bulk

nanostructure is large. Thus, there will be a larger influence of the surface effects on the

general behavior of the nanostructure.

The aim of this thesis is, on one hand, the development of a technique capable of produce

ferroelectric nanostructures periodically ordered of controlled size and shape. On the other

hand, there is a second scope: the preparation of real systems in order to study basic problems

of nanoferroelectric materials, related to the scale of the size and the surface effects. For that,

ferroelectric nanostructures of PbTiO3 in its perovskite phase will be prepared using a bottom-

U

z

U

z

U

z

U

z

a)

c)

b)

d)

P

E

a

c

b

d

e)

Chapter 1. Introduction 13

up method. This procedure will ensure that the nanostructures are grown damage-free, unlike

the ones prepared from any of the top-down procedures, as it was explained previously and,

thus, that the size and surface effects observed are those caused by the scaling of the size of

the nanostructure rather than by the preparation technique.

14 Bibliography

Bibliography

[1] R.B. Gray. Transducer and method of making the same. 1949.

[2] I. Bretos, "Low-toxic chemical solution deposition methods for the preparation of

multifunctional (Pb1-xCax)TiO3 thin films", Departamento de Química Inorgánica, 2006, Madrid:

Universidad Autónoma de Madrid. Ph.D. Thesis.

[3] J. Valasek, "Piezoelectric and allied phenomena in Rochelle salt", Physical Review, 15,

1920, p:537

[4] N. Setter and R. Waser, "Electroceramic materials", Acta Materialia, 48, 2000, p:151

[5] G.H. Haertling, "Ferroelectric ceramics: history and technology", Journal of the

American Ceramic Society, 82 (4), 1999, p:797

[6] N. Setter, "Electroceramics: looking ahead", Journal of the European Ceramic Society,

21, 2001, p:1279

[7] C.H. Ahn, K.M. Rabe and J.M. Triscone, "Ferroelectricity at the nanoscale: Local

polarization in oxide thin films and heterostructures", Science, 303 (5657), 2004, p:488

[8] D.D. Fong, G.B. Stephenson, S.K. Streiffer, J.A. Eastman, O. Auciello, P.H. Fuoss and C.

Thompson, "Ferroelectricity in ultrathin perovskite films", Science, 304 (5677), 2004, p:1650

[9] J. Junquera and P. Ghosez, "Critical thickness for ferroelectricity in perovskite ultrathin

films", Nature, 422 (6931), 2003, p:506

[10] A.N. Morozovska, E.A. Eliseev and M.D. Glinchuk, "Ferroelectricity enhancement in

confined nanorods: Direct variational method", Physical Review B, 73 (21), 2006,

[11] Naumov, II, L. Bellaiche and H.X. Fu, "Unusual phase transitions in ferroelectric

nanodisks and nanorods", Nature, 432 (7018), 2004, p:737

[12] J.E. Spanier, A.M. Kolpak, J.J. Urban, I. Grinberg, O.Y. Lian, W.S. Yun, A.M. Rappe and H.

Park, "Ferroelectric phase transition in individual single-crystalline BaTiO3 nanowires", Nano

Letters, 6 (4), 2006, p:735

[13] W.S. Yun, J.J. Urban, Q. Gu and H. Park, "Ferroelectric properties of individual barium

titanate nanowires investigated by scanned probe Microscopy", Nano Letters, 2 (5), 2002,

p:447

[14] J.F. Scott, Ferroelectric memories. Advanced microelectronics, ed. K. Itoh and T.

Sakurami. 2000, Berlin: Springer.

[15] C. Sealy, "Winning the memory race", Materials today, 11 (6), 2008, p:16

[16] G.I. Meijer, "Who Wins the Nonvolatile Memory Race?", Science, 319, 2008, p:1625

[17] L.A. Zheng. Method of forming inside rough and outside smooth HSG electrodes and

capacitor structure. 2007; US7459746.

Bibliography 15

[18] Toshiba Develops World's Highest-Bandwidth, Highest Density Non-volatile RAM. 2009.

http://www.toshiba.co.jp/about/press/2009_02/pr0902.htm

[19] Toshiba Develops World's Fastest, Highest Density FeRAM. 2006.

http://www.toshiba.co.jp/about/press/2006_02/pr0701.htm

[20] M. Alexe, C. Harnagea and D. Hesse, "Non-conventional micro- and nanopatterning

techniques for electroceramics", Journal of Electroceramics, 12, 2004, p:69

[21] M. Alexe and D. Hesse, "Self-assembled nanoscale ferroelectrics", Journal of Materials

Science, 41, 2006, p:1

[22] C. Ganpule, A. Stanishevsky, Q. Su, S. Aggarwal, J. Melngailis, E. Williams and R.

Ramesh, "Scaling of ferroelectric properties in thin films", Applied Physics Letters, 75, 1999,

p:409

[23] C. Harnagea, M. Alexe, J. Schilling, J. Choi, R.B. Wehrspohn, D. Hesse and U. Gösele,

"Mesoscopic ferroelectric cell arrays prepared by imprint lithography", Applied Physics Letters,

83 (9), 2003, p:1827

[24] M. Alexe, C. Harnagea, D. Hesse and U. Gösele, "Patterning and switching of nanosize

ferroelectric memory cells", Applied Physics Letters, 75, 1999, p:1793

[25] M. Alexe, C. Harnagea, D. Hesse and U. Gösele, "Polarization imprint and size effects in

mesoscopic ferroelectric structures", Applied Physics Letters, 79, 2001, p:242

[26] G.K. Hubler, Comparison and vacuum deposition techniques, in Pulsed Laser Deposition

of Thin Films, G.K.H. D. B. Chrisey, Editor. 1994, Wiley: New York.

[27] A. Milchev, Electrocrystallization. Fundamentals of nucleation and growth. 1st ed.

2002: Springer.

[28] J. Sun, P. Jina, Z.G. Wanga, H.Z. Zhangb, W. Z.Y. and L.Z. Hu, "Changing planar thin film

growth into self-assembled island formation by adjusting experimental conditions", Thin Solid

Films, 476, 2005, p:68

[29] M. Alexe, J.F. Scott, C. Curran, N.D. Zakharov, D. Hesse and A. Pignolet, "Self-patterning

nano-electrodes on ferroelectric thin films for gigabit memory applications", Applied Physics

Letters, 73 (11), 1998, p:1592

[30] W. Ma, M. Alexe and U. Gösele, "Formation of Ferroelectric Perovskite Nanostructure

Patterns Using Latex Sphere Monolayers as Masks: An Ambient Gas Pressure Effect during

Pulsed Laser Deposition ", Small, 1 (8-9), 2005, p:837

[31] W. Lee, H. Han, A. Lotnyk, M.A. Schubert, A. Senz, M. Alexe, D. Hesse, S. Baik and U.

Gösele, "Individually addressable epitaxial ferroelectric nanocapacitor arrays with near Tb

inch2 density", Nature nanotechnology, 3 (7), 2008, p:402

16 Bibliography

[32] D. Grosso, C. Boissiere, B. Smarsly, T. Brezesinski, N. Pinna, P.A. Albouy, H. Amenitsch,

M. Antonietti and C. Sanchez, "Periodically ordered nanoscale islands and mesoporous films

composed of nanocrystalline multimetallic oxides", Nature Materials, 3 (11), 2004, p:787

[33] I. Szafraniak, C. Harnagea, R. Schloz, S. Bhattacharyya, D. Hesse and M. Alexe,

"Ferroelectric epitaxial nanocrystals obtained by a self-patterning method", Applied Physics

Letters, 83 (11), 2003, p:2211

[34] S. Kronholz, S. Rathgeber, S. Karthauser, H. Kohlstedt, S. Clemens and T. Schneller,

"Self-assembly of diblock-copolymer micelles for template-based preparation of PbTiO3

nanograins", Advanced Functional Materials, 16 (18), 2006, p:2346

[35] S. Bühlmann, P. Muralt and S. Von Allmen, "Lithography-modulated self-assembly of

small ferroelectric Pb(Zr,Ti)O3 single crystals", Applied Physics Letters, 84, 2004, p:2614

[36] S. Clemens, S. Rohrig, A. Rudiger, T. Schneller and R. Waser, "Embedded ferroelectric

nanostructure arrays", Nanotechnology, 20 (7), 2009, p:5

[37] B. Jaffe, W.R. Cook and H. Haffe. 1971, New York: Academic Press.

[38] B. Jaffe, W.R. Cook and H. Haffe, Piezoelectric Ceramics. 1971, New York: Academic

Press.

[39] B.A. Strukov and A.P. Levanyuk, Principios de ferroelectricidad. 1988, Madrid: Ediciones

de la Universidad Autónoma de Madrid.

[40] B.A. Tuttle, D.A. Payne and J.L. Mukherjee, "Spontaneous Polarization for ferroelectric

materials", MRS bulletin, 19 (7), 1994, p:20

[41] A. Roelofs, T. Schneller, K. Szot and R. Waser, "Towards the limit of ferroelectric

nanosized grains", Nanotechnology, 14, 2003, p:250

CHAPTER 2: EXPERIMENTAL PROCEDURE

In this chapter, the details of the preparation process of the nanoparticles as well as the

explanation of the methods for the characterization of their properties will be presented.

2.1. Precursor solutions.

Nanoparticles will be prepared by deposition of a micellar solution onto the substrates by

three different procedures that will be explained in this Experimental Procedure Chapter. For

the preparation of micellar solutions, a PbTiO3 sol and a microemulsion are mixed. The

procedures to obtain those are here described.

2.1.1. Synthesis of the sol.

The precursor sol was synthesized using the route generally known in the literature as the

“diol-route” [1]. Therefore, lead (II) acetate trihydrate (Pb(OCOCH3)2·3H20, Aldrich, 99%) and

1,3-propanediol (HO(CH2)3OH, Aldrich 98%) were refluxed in air at ~ 155°C for 1 h. Then,

titanium (IV) di-isopropoxide bis-acetylacetonate (Ti(OC3H7)2(CH3COCHCOCH3)2, Aldrich,

75 wt% solution in 2-propanol) was added to the mixture and reflux was maintained in air at

~110°C for 8 h. After this step, byproducts were partially distilled off the solution. The volume

of the distilled liquid was the 80% volume of the total 2-propanol (CH3CHOHCH3) that the

synthesized sol contains [2]. An air stable and precipitate-free Pb(II)-Ti(IV) sol was obtained.

Stock sols are obtained with an average concentration of ~1.44 M (equivalent moles of PbTiO3

per liter of sol) and ~1.46 g/ml of density. Fig. 2.1. schematizes the route of synthesis.

Figure 2.1. Synthesis of the sol.

Pb(OCOCH3)2·3H2O + HO(CH2)3OH

Ti(OC3H7)2(CH3COCHCOCH3)2

Pb(II)-Ti(IV) sol

Reflux in air (155°C, 1h)

Ti (IV):1,3 propanediol1:5

18 2.1. Precursor solutions

2.1.2. Preparation of the microemulsion.

Microemulsions were prepared by mixing 0.420 g of Brij-30 (CH3(CH2)12(OCH2CH2)4OH, Aldrich

99%), 1.950 g of cyclohexane (C6H12, Aldrich, 99+%) and 0.034 g of deionized water. The

water:surfactant ratio is 1.0:1.6. The water-clear appearance of the mixture upon vigorous

stirring indicated the formation of the microemulsion. Fig. 2.2. schematizes the preparation

method.

Figure 2.2. Preparation of the microemulsion.

The microemulsions were kept for 24h and ultrasonically stirred for 10 minutes before the

deposition step.

2.1.3. Preparation of the micellar solution

The micellar solutions were prepared by mixing the sol with the microemulsion and adding 1,3-

propanediol (HO(CH2)3OH, Aldrich, 98%) to obtain the desired concentration. The preparation

method is schematized in Fig. 2.3.

Figure 2.3. Preparation of the micellar solution.

Brij-300.420 g

C6H12

1.950 g

Deionized H2O0.034 g

Microemulsion

Brij-30:H2O1.0:1.6

Microemulsion Pb(II)-Ti(IV) sol

1,3-propanediol

Micellar solution

Chapter 2: Experimental procedure 19

The micellar solutions were kept for 24 h and ultrasonically stirred for 10 minutes before the

deposition step.

Micellar solutions were also prepared by adding Ethyl-Hexanol and deionized water. Water

caused the precipitation of the sol, and was, thus avoided in the preparation of micellar

solution. Ethyl-Hexanol, instead was immiscible with cyclohexane. Thus, 1,3-propanediol was

chosen as solvent.

2.2. Selection of substrates.

Two different types of substrates were chosen for the fabrication of the nanostructures. A

polycrystalline substrate -Pt-coated Si(100) substrates-, compatible with the current Si

technology, and a single-crystal one –SrTiO3-, chosen because its crystal structure similarity

which, in addition, is supposed to lead to epitaxial growth.

2.2.1. Pt-coated Si(100) substrates

Selected Pt-coated Si(100) substrates are formed by Pt/TiO2/SiO2/(100)Si. The SiO2 layer

appears spontaneously when Si is in contact with air. Pt and TiO2 layers are deposited onto the

Si wafer (Crystal GmbH) by radiofrequency magnetron sputtering (Alcatel SCM 450) at the

“Laboratorio de sensores” of the “Centro de Tecnologías Físicas Leonardo Torres Quevedo”

(CSIC) in Madrid, with resulting thickness of ~100 nm and ~50 nm for the Pt and the TiO2,

respectively.

The substrates were cleaned before the deposition by ultrasonically soaking them in

trichloroethylene (Cl2CClCH, Panreac, stabilized with ethanol, 99%) for 5 min, in acetone

(CH3COCH3, Panreac, 99.5%) for 4 min and in 2-propanol (CH3CHOHCH3, Panreac, 99.5%) for

3 min, consecutively.

2.2.2. Microemulsion/Pt-coated Si(100) substrates

Functionalized substrates were prepared by depositing a layer of microemulsion by spin-

coating (50 s, 2000 rpm) onto the cleaned Pt/TiO2/SiO2/(100)Si substrates described in

previous subsection 2.2.1.

2.2.3. (100)SrTiO3 substrates

Commercial SrTiO3 (STO) and Nb-doped SrTiO3 substrates from Crystal GmbH are used in this

work as-served, without additional cleaning.

20 2.2. Selection of substrates

2.2.4. (100)SrTiO3 substrates with controlled surfaces

Chemical treatments of the substrates consisted on two different procedures: in the first one,

substrates were soaked in an etching solution (Buffered Hydrofluoric Acid, BHF, pH 4.5, 20 s);

in the second one, they were ultrasonically soaked in deionized water for 10 min and then

soaked in an etching solution (BHF, pH 5.5, 30 s).

BHF with pH 4.5 was prepared by adding 9.69 g of ammonium fluoride (NH4F, Aldrich, 98+%) to

10 ml of deionized water. Hydrofluoric acid (HF, Panreac, 99%) was added up to pH 4.3

(measured with a Thermo scientific Orion pHmeter). Deionized water is added up to complete

25 ml in order to obtain a 10 M ammonium fluoride solution.

BHF with pH 5.5 was prepared by adding 9.69 g ammonium fluoride to 10 ml of deionized

water. Hydrofluoric acid was added up to pH 5.5. Deionized water is added up to complete 25

ml.

Figure 2.4. Thermal treatments in UHV applied to the STO substrates.

Three different thermal treatments (annealings) were carried out with no oxygen flow,

contrary to the processes described previously in the literature [3, 4]. Substrates were

annealed in Ultra High Vacuum (UHV) using a radiative furnace, with the heater facing sample

backside and feedback controlled by a PID system. Since the furnace thermocouple was

located close to the substrate, but not in direct contact with it, the temperature of the STO

surface was calibrated by means of radiation thermometry at the center of the test Si(100)

~0.1 °C

815 ± 5 °C, 3600s

~0.03 °C

~0.2 °C

842 ± 7 °C, 7200s

~0.09 °C

a)

b)

Chapter 2: Experimental procedure 21

samples. The pressure of the UHV chamber (initially below 2x10-10 Torr) kept below 1x10-8 Torr

during these thermal treatments. Annealings in UHV were carried out with two different

treatments described in Fig. 2.4: in the first one, a slow heating rate of ~0.1 ⁰C/s up to

815±5 ⁰C was used. The temperature was maintained for 3600 s. The average cooling rate

applied is very low (0.03 ⁰C/s) down to room temperature. In the second case, an average

heating rate of ~0.2 ⁰C/s is applied until 842±7 ⁰C. The temperature is maintained for 7200 s.

The cooling rate applied in this second thermal treatment is ~0.09 ⁰C/s down to room

temperature.

In the case of the thermal treatments in air, the substrate was annealed with a heating rate of

~30 ⁰C/s using rapid thermal processing (RTP), keeping the annealing temperature (1050⁰C) for

3600 s and with a very low average cooling rate of ~0.07 ⁰C/s down to room temperature,

using soaking times of 1800 s at the temperatures indicated in Fig.2.5.

Figure 2.5. RTP thermal treatments applied to the STO substrates.

Table 2.1 summarizes the treatments performed for controlling the substrate surface.

Table 2.1. Chemical and thermal treatments carried out in order to control the surface of the STO

substrates.

Soaking in deionized water Chemical etching Thermal treatment

- pH 4.3, 20 s UHV, 842±7 ⁰C for 7200 s

Ultrasonically soaking for 10

min pH 5.5, 30 s UHV, 815±5 ⁰C for 3600 s

Ultrasonically soaking for 10

min pH 5.5, 30 s In air, 1050ºC for 3600 s

Further information about these substrates is provided in section 5.3.

1050°C, 3600s

~30°C/s

950°C, 1800s

850°C, 1800s

750°C, 1800s

650°C, 1800s

~0.1°C/s

22 2.3. Deposition, drying and crystallization of the PbTiO3 nanostructures

2.3. Deposition, drying and crystallization of the PbTiO3

nanostructures.

The deposition of the micellar solutions has been carried out by spin-coating. This process can

be divided into four stages [5]: deposition, spin-up, spin-off and evaporation, as shown in Fig.

2.6. During the deposition stage, a liquid excess is dispensed onto the substrate surface. This

liquid flows radially outward in the spin-up stage, driven by centrifugal force. In the spin-off

stage, the liquid excess flows to the perimeter of the substrate surface and leaves it as droplets.

Finally, evaporation of solvent occurs during the last stage.

Figure 2.6. Spin-coating process.

Glass syringes equipped with non-sterile plastic nozzles and coupled with filters (Millipore) of

0.2 µm pore size were used in the deposition of the sols for the nanostructures fabricated by

using the phenomenon of the microstructural instability of polycrystalline ultrathin films. In

the case of nanostructures prepared by using micellar solution, no filters had been used, as the

micelles size is larger (see section 4.2 for further details) than the pore size of the filters.

Spin-coating was carried out in a spinner equipment (TP 6000 gyrset system of SET-Micro-

Controle group model), working at 2000 rpm for 45 s.

When fabricating nanostructures by using the phenomenon of the microstructural instability

of polycrystalline ultrathin films, the wet precursor films thus obtained were dried and

pyrolysed in order to remove the entrapped solvent and the majority of the organic species

present within the gel network. The wet layer was subjected to a thermal treatment on a hot

plate (Selecta) stabilized at 350 ⁰C for 60 s.

In the case of the fabrication of nanostructures by the microemulsion mediated synthesis, in

order to promote their evaporation induced self-assembly, the micellar layer was dried at

ω

ω

dω/dt≠0

ω ω

Deposition Spin-up Spin-off Evaporation

Chapter 2: Experimental procedure 23

~30 ⁰C for 7 days in an oven and the relative humidity was kept constant at ~30 % (controlled

by an hygrometer HD2101.2 by Delta Ohm).

The coatings so obtained were crystallized into the desired oxide phase (i.e. perovskite) by

Rapid Thermal Processing (RTP, carried out at a Jetstar 100T JIPELEC equipment). The samples

were annealed in air at 650 ⁰C for 50 s with a heating and cooling rate of 30 ⁰C/s. Fig. 2.7.

shows the thermal program used for the crystallization of the PbTiO3 nanostructures.

Figure 2.7. Thermal recipe used for the crystallization of the nanostructures.

2.4. Microscopy and quantitative microstructure characterization

A description and fundamental definitions concerning the characterization techniques of the

microstructures used in this work are given in the following sections.

2.4.1. Optical microscopy

Optical microscopy images were obtained with an optical microscope, at x22 magnification.

Calibration of the images was done with a micrometer (Leitz).

2.4.2. Scanning Electron Microscopy

The Scanning Electronic Microscopy (SEM) [6] is based on the use of a beam of highly energetic

electrons to examine objects at a very fine scale. In SEM, the image is formed by an electron

beam which is focused on the specimen surface. This beam is first produced at the top of the

microscope by a thermoionic electron gun. Electrons are accelerated towards the specimen

using a positive electric potential of 2-40 kV. The electron beam is confined and focused using

metal apertures and magnetic lenses into a thin, monochromatic beam, directed down to the

sample. The interaction between the primary electrons and the sample gives rise to various

physical phenomena (backscattering, cathodoluminiscence, secondary electrons, x-ray, Auger

emission, etc). In the case of SEM, the detector collects secondary electrons and converts

them into a signal that is sent to a viewing screen, producing thus an image. Secondary

600 ⁰C

650 ⁰C, 50s

1s

1s

1s1s

1s

1s

550 ⁰C

500 ⁰C

100 ⁰C

20s

24 2.4. Microscopy and quantitative microstructure characterization

electrons are excited electrons of the specimen produced by the inelastic collision with the

primary electron beam which results ejected from the atoms and, after undergoing additional

scattering events while travelling through the sample, emerged from its surface with a low

energy value (< 50 eV). Due to this low energy, only secondary electrons that are near the

surface can escape and be examined. Hence, the information obtained by SEM is related to the

topography of the sample.

SEM images of this work have been obtained in three different electronic microscopes: JSM

6335F NT (FEG-SEM) microscope at the “Centro de Microscopía Electrónica Luís Bru” of the

“Universidad Complutense de Madrid” at Madrid, a Philips ESEM XL 30 FEG at the

Interdisziplinäres Zentrum für Materialwissenschaften at Halle (Germany) and a FEI XL30-SFEG

at Cranfield University at Cranfield (United Kingdom).

2.4.3. Transmission electron microscopy (TEM)

A transmission electron microscope essentially works as a SEM one, with the main difference

that the image formed by TEM (Transmission Electron Microscopy) is obtained from the

electrons transmitted through the specimen [7]. Since electronic absorption by the sample is

much more efficient than electronic transmission through it, TEM microscopes usually work

with a high voltage electron beam of 80-400 keV. Furthermore, materials for TEM analysis

must be specially thinned to get specimens which allow electrons to be transmitted through

them (100-200 Å thickness).

A TEM microscope basically consists in a column, set at ultrahigh vacuum, where the

illumination system is displayed along. The electron gun produces a high-energy stream of

monochromatic electrons which are focused into a thin, coherent beam by the use of

condenser lenses. The first lens determines the spot size (general size range of the final spot

that strikes the sample), whereas the second one controls its intensity (from a wide dispersed

spot to a pinpoint beam). The beam is restricted by the condenser aperture until it strikes the

specimen and parts of the electrons are transmitted through. The transmitted electrons are

focused by the objective lens into an intermediate image, which is enlarged by the

intermediate and projector lenses until it impinges on a phosphor image screen and light is

generated, allowing the user to see the image.

Two types of information (images) can be obtained; the direct image projected by the entire

specimen (microstructure) and the electron diffraction pattern resulted when the electron

beam crosses an orderly crystallographic pattern (crystallographic structure). With the

Chapter 2: Experimental procedure 25

objective aperture the image contrast is enhanced by blocking out high-angle diffracted

electrons, whereas the selected area aperture allows collecting the diffraction patters

obtained. From the electron diffraction patterns, the interplanar distances dhkl and indexation

of planes {hkl} were calculated, according to Bragg’s Law and assuming low diffraction angles θ,

by the following expression

𝑑ℎ𝑘𝑙 = 𝜆𝐿/𝑅 (2.1)

where L is the camera length of the microscope (specimen-screen distance), λ is the

wavelength of the electron beam and R is the distance of the diffracted spot respect to the

incident beam position (in the electron pattern).

TEM images can be bright or dark field ones. Bright field images are obtained by opening the

lens of the microscope, which eliminates the diffracted beams. In contrast, each diffracted

beam form an image called dark field image in which the maximum intensity is observed for

the portions of the sample that diffracts in a certain angle.

TEM measurements have been carried out at the “Universidad Carlos III de Madrid”, in Madrid

(Spain), using a JEOL-200 FX II microscope operating at 200 kV. Cross section specimens of the

films were prepared from two pieces of the sample which were stacked with the

nanostructures layers facing and glued. Lateral thickness was decreased until 20 µm, using a

tripod polisher. Large electron-transparent areas were obtained by Ar+ ion milling (acceleration

voltage of 5 kV, beam intensity of 5 mA and incidence angle between 8-10°).

2.4.4. Scanning Probe Microscopy

Scanning Probe Microscopy (SPM) [8, 9] is a general name for a group of techniques in which a

sharp probe scans the surface of a sample, measuring some property of it. The tip is mounted

onto a flexible cantilever of known geometrical and material properties, so that it is possible to

separate all the contributions of the interaction with a high sensitivity. The interactions of the

tip with the sample are sensed by the resultant deflection of the cantilever onto which the tip

is mounted. This deflection is measured by reflecting a laser beam off the cantilever. The

reflected laser beam strikes a position-sensitive photo-detector consisting of four-segment

photo-detector. The differences of the intensity of the signals between the segments of photo-

detector indicate the position of the laser spot on the detector and thus the angular

deflections of the cantilever.

26 2.4. Microscopy and quantitative microstructure characterization

A typical SPM system consists of a tip mounted onto a cantilever, a xyz piezo-scanner attached

either to the sample or to the cantilever, a laser and a deflection sensor or detector as

schematized in Fig. 2.8.

Figure 2.8. Experimental set-up of a SPM system.

As explained before, the sample investigation is possible by studying the forces acting between

the sharp probe and the surface of the sample. Depending on the distance between them, one

or another force dominates the system, that is to say, they are position-dependent. Fig. 2.9

represents the force-distance curve for the interaction between the probe and the sample.

Figure 2.9. Force-distance curve of the interaction between the probe tip and the sample surface.

Depending on the point of the curve where measurements take place, images might be

acquired in contact, intermittent-contact or non-contact modes.

detector

sample

XYZ piezo-scanner

cantilever

Laser beam

tip

Forc

e(F

)

Distance (d)

Repulsive forces

Attractive forces

Chapter 2: Experimental procedure 27

In contact mode, the tip is constantly adjusted to maintain a constant force over the sample as

it scans it, and measuring the deflection of the cantilever provides the information about the

interaction between the probe and the surface. The probe mostly senses the repulsive short-

range interatomic forces.

During scanning, "profile broadening" artifact due to the tip-sample convolution can appear. In

Fig. 2.10, the elementary tip-sample convolution phenomenon is examined for the AFM

operating in the contact mode, where the two main phenomena are schematized. The 2D

height profile appears these ways because of the geometry of the tip: it has a finite size and

shape so it is not able to fit into all of the space not filled by the particle being imaged. Some

negative spaces are even inaccessible and will contribute to the appearance of the height

profile (Fig. 2.10 (b)).

Figure 2.10. Artifacts of the AFM topography image in contact mode in the case of tip radius smaller (a)

and larger (b) than the nanostructures, measured.

In the first case (Fig. 2.10 (a)), the tip radius is much smaller than the object. This object is

imaged broader than the real one, while the height is the same. Supposing a perfect tip that

ends in a cone, broading of the objects follow the next equation:

𝑅𝑐 = 𝑅 cos𝜃 + 1 + sin𝜃 tan𝜃 (2.2)

In the second case (Fig. 2.10 (b)), the tip radius is much larger than the spherical object. In this

case, the tip move across the object surface and can be approximated by a sphere of radius R’

moving along the sphere of radius R, i.e. the tip describes an arc of radius R’+R.

θ

2R

2R0

a)

R’2R

R’2R

b)

28 2.4. Microscopy and quantitative microstructure characterization

A second mode of operating is the so called, intermittent-contact, dynamic mode or tapping

mode. In this mode [10-13], the cantilever is forced by an actuator placed at its base to

oscillate up and down at its resonant frequency. Part of the oscillation extends into the

repulsive regime, so the tip intermittently touches or “taps” the surface. This is the reason why

this mode is also referred as tapping mode. Constant amplitude is maintained by a feedback

mechanism while scanning the sample. In this manner an image of the surface topography is

generated [14]. Fig. 2.11 shows how oscillation resonant frequency shifts with the repulsive or

attractive forces between the sample and the probe.

Figure 2.11. Resonant frequency shift when the probe is affected by attractive or repulsive forces.

Tapping or intermittent-contact mode tends to be more applicable to general imaging in air,

particularly for soft samples, as the lateral resolution is higher than in contact mode while the

forces applied to the sample are lower and less damaging.

In non-contact mode, the cantilever-tip system vibrates at its resonant frequency above the

sample and the changes in the topography are detected by changes in the amplitude or the

phase of the vibration. It is basically the same as tapping mode, but the tip never touches the

sample. This can lead to artifacts such as imaging water drops on the surface as if they were

part of the topography of the samples.

ωR ω'Rω‘’R

Δω

ΔA

AR

A’R

A

ω

Attractive forces Repulsive forces

Chapter 2: Experimental procedure 29

Atomic Force Microscopy (AFM) is a SPM technique based on the detection of the topography

of the surface by measuring the atomic force between the tip probe and said surface. In this

work, AFM topography images are acquired with a commercial SPM microscope (Nanotec

Electronica) controlled by WSxM software [15]. The probe used in this work (Nanosensors) is

presented in Fig. 2.12.

Figure 2.12. Nanosensors probes for the acquisition of AFM topography images.

2.4.4.1. Fast Fourier Transform and self-convolution images.

Self-correlation is defined as [16]:

𝐺 𝑘1 ,𝑘2 = 𝑓 𝑥,𝑦 · 𝑓 𝑥 + 𝑘1,𝑦 + 𝑘2 (2.3)

Where f(x,y) is the image matrix. This equation takes the image and the same image is shifted

a distance k1 and k2 in the X and Y axis with respect to the center of the image. The resulting

image, G(k1, k2), is a measure of how different the two images are. The more similar the image

and the shifted image are, the higher the value of the self correlation. In self-correlation, the

highest value is obtained at the center of the image (where k1 and k2 are zero). Any periodicity

in the original image will be shown as a periodic pattern in the self correlation.

The Fast Fourier Transform (FFT) is an efficient algorithm to compute the Discrete Fourier

Transform (DFT) and its inverse, reducing the number of computations needed for N points

from 2·N2 to 2·N·log2N. DFT is a generalized case of the Continuous Fourier Transform for a

discrete function and it is extremely useful because it reveals periodicity in input data as well

as the relative strengths of any periodic components by transposing from time-domain to

frequency-domain [16]. The DFT can be transformed efficiently using a FFT.

In the system used in this work, SPM signals are discrete and obtained by using an Analogical

to Digital Converter (ADC) chip placed at the Digital Signal Processor (DSP). The DSP do not

perform integral calculus and thus, the signal is obtained by smashing numbers using the

“Butterfly” algorithm [16]. This algorithm is a standard procedure that implements the

SSS-NCH• SuperSharpSilicon - tip• Non-Contact High frequency• C = 42 N/m, fo = 330 kHz

PPP-NCH• Pointprobe® Plus - tip• Non-Contact High frequency• C = 42 N/m, fo = 330 kHz

PPP-NCHPt• Pointprobe® Plus - tip• Non-Contact High frequency• C = 42 N/m, fo = 330 kHz• PtIr5 coated probe

PPP-CONT• Pointprobe® Plus - tip• Contact Mode

• C = 0.2 N/m, fo = 13 kH

30 2.4. Microscopy and quantitative microstructure characterization

numeric conversion of a time-domain signal into a frequency spectrum, being both discrete,

performing the FFT.

Figure 2.13 illustrates the self-correlation and the FFT filter of a high order sample.

Figure 2.13. Original sample (a), image after applying the FFT filter (b) and self-correlated image (c).

2.4.5. Image analyses.

Size distribution of the nanostructures fabricated in this work will be obtained by

measurements on microscopy images and analyzed later by a graphic method using probability

plots. This allows the precise determination of average values and standard deviations of the

distributions. And, besides, microstructural phenomena that escape the less detailed

microstructural analysis are easily studied, like the process of normal grain growth [17]. For

example it has been used for the study of relevant phenomena concerning microstucture-

properties relationships in piezoceramics such as the ones presented in Fig. 2.14.

Figure 2.14. Study the porosity reduction and microstructure recovery from the degraded sintering

stage by subsequent Hot isostatic pressing (HIP) [18] (a) and the study the limit of the increase in grain

size without microstructure degradation by recrystallization of hot pressed samples in Aurivillius

ceramics [19] (b).

a) b) c)

(a)

Chapter 2: Experimental procedure 31

After obtaining the AFM or SEM topography images, they are processed by an image

manipulation software (Gimp, GPU license) and a graphic analysis software (MIP4, Digital

Image System) based on the one designed to study microstructure in ceramics [20]. First,

image analysis of the original image (Fig. 2.15 (a)) is carried out to enhance the contrast and

select the particles to be measured using Gimp. Then, segmentation of the image or

conversion into a binary image is carried out by selecting a gray threshold level and pixels of

the image that correspond to the isolated nanostructures are painted in white and those of the

substrates or non isolated nanostructures in black (Fig. 2.15 (b)).Finally, the object

identification, that is to say, the automatic identification of each isolated area that

corresponds to a single nanostructure is made (Fig. 2.15 (c)). Subsequently, the software

carries out the measurement of the selected parameters for each isolated area. This procedure

is illustrated in Fig. 2.15.

Figure 2.15. Example of the image analysis process previous to the measurement of the nanostructures

onto a substrate. (a) is the original image, (b) the binary image and (c) shows the identification by MIP4

of each nanostructure by coloring them.

The size of an object may be characterized using a number of linear parameters (interception

length, perimeter, Feret diameter, etc.) or the area inside the object, all which distribution can

be precisely obtained by computer-aided image measurements. The characteristic parameter

selection for the best description and study of an object depends on its shape. In this work,

nanostructures are characterized by the equivalent diameter to their circular shape, as they

have more or less the same dimensions in all directions. The equivalent diameter is calculated

as 𝐷𝑒𝑞 = 4𝐴𝑟𝑒𝑎𝜋 . At least 200 nanostructures are measured in each image in order to

obtain reliable statistical distributions.

Distributions in this work are either Gaussians or Lognormal. Gaussians distributions describe

data that clusters around a mean or average value. The probability density function for

Gaussin or normal distribution is given by

a) b) c)

32 2.4. Microscopy and quantitative microstructure characterization

𝑓 𝑥, < 𝑠 >,𝜍 =1

σ 2πe−

1

2

x−<𝑠>

σ

2

(2.4)

where <s> is the mean, which, in the case of a Gaussian is equal to the average value and σ is

the standard deviation.

Lognormal distributions are those for which the distribution of the logarithm of the data is a

Gaussian. The probability density function for a lognormal distribution can be expressed as:

𝑓 𝑥, < 𝑠 >,𝜍 =1

x·σ 2πe−

1

2

lnx −<𝑠>

σ

2

(2.5)

where <s> is the mean, which, in the case of a Gaussian is equal to the average value and σ is

the standard deviation of lnx.

From these parameters, it is possible to obtain the average size, <s>, and the standard

deviation, σ, of a lognormal distribution from the fitting of the probabilistic line. If a line is

obtained when representing the size distribution with respect to the accumulated frequency,

then the distribution is Gaussian. When the line is obtained only if representing the logarithm

of the size distribution, then, the distribution is lognormal (see Fig. 2.14 (b)).

The different parameters can be determined by the following expressions:

𝑦 =1

𝜍𝑥 −

<𝑠>

𝜍 (2.6)

where y is the accumulative frequency and x, the studied variable, i.e. the size of the

nanostructures. The expression is the same if instead of representing the size of the

nanostructure, its logarithm is represented, as is the case of lognormal distributions.

The mixture of two distribution functions can be also detected, by changes in the slope of the

probabilistic curve. Bimodal distributions result in curves that are asymptotic to their parent

distributions. Information (average size and standard deviation) about parent distributions can

be obtained here by the fitting lines to which the distributions are asymptotic [18, 20-22] (see

Fig. 2.14 (a)).

Chapter 2: Experimental procedure 33

2.5. Structural characterization

2.5.1. Synchrotron X-Ray diffraction.

The crystalline structure of the nanostructures has been studied by synchrotron X-ray

diffraction in grazing incidence geometry.

Synchrotron radiation is usually defined as the electromagnetic radiation emitted by a

relativistic charged particle (generally electrons) moving on a circular orbit, e.g. in a

synchrotron. In a synchrotron, the electric (accelerating) and magnetic (deflecting) fields are

synchronized to ensure that the particles follow a prescribed geometric and energetic path

[23].

The emission of electromagnetical radiation from an accelerated charge is a classical

phenomenon [24]. In a synchrotron, centripetal acceleration generates an extremely intense

beam of synchrotron radiation that is emitted tangent to the charges (electrons) path. Under

the operation conditions of a synchrotron, the emitted radiation typically goes from ultraviolet

(1016 Hz) to X-ray (1018 Hz).

One of the sources of synchrotron radiation is wigglers. A wiggler consists on a series of

magnets designed to periodically laterally deflect ('wiggle') a beam of charged particles inside a

storage ring of a synchrotron and, thus, generate synchrotron radiation.

When an X-ray beam impinges on a crystal (i.e. periodically ordered of atoms) which

interatomic distances are about the same as the wave length (on the order of a few

angstroms), part of the waves can be scattered by the atoms present in the crystal. If the

scattered waves conserve the energy of the incident beam (elastic scattering) and are also able

to interfere with each other (coherence), the diffraction phenomenon may occur. The

diffraction arises when scattered beams are combined in phase with each other, producing an

unmoving distribution of constructive interference and, hence, a visible interference x-ray

pattern. All former conditions can be summarized in the Bragg’s Law:

2 · 𝑑ℎ𝑘𝑙 · sin𝜃 = 𝑛 · 𝜆 (2.7)

Where dhkl is the distance between scattering centers (atomic {hkl} planes), θ is the diffraction

angle, n is an integer known as the order of the diffracted beam and λ is the wavelength of the

incident beam.

34 2.5. Structural characterization

Each crystal presents a particular and characteristic diffraction pattern, since the different

interatomic distances between the {hkl} planes are specific from its primitive unit cell. In this

way, the unit cell vectors which define the structure of the crystal can be determined from the

X-ray diffraction patterns. On the other hand, the intensity of the diffracted beams is related to

the nature and arrangement of the atoms in the crystalline network. Intensity variations are

indicative of preferred orientation of one or more particular crystallographic planes in the

crystal.

The reciprocal lattice of a Bravais lattice of position vector 𝑅 is mathematically defined as the

lattice with position vectors 𝑘 that fulfills the next equation:

𝑒𝑖𝑘 𝑅 = 1 (2.8)

Each point (hkl) in the reciprocal lattice corresponds to a set of lattice planes (hkl) in the real

space lattice. The direction of the reciprocal lattice vector corresponds to the normal to the

real space planes, and the magnitude of the reciprocal lattice vector is equal to the reciprocal

of the interplanar spacing of the real space planes.

The reciprocal lattice plays a fundamental role in most analytic studies of periodic structures,

particularly in the theory of diffraction.

Ewald sphere was introduced in connection with the concept of the reciprocal lattice and is the

geometrical representation of the Bragg’s Law [25-27]. The aim of the Ewald sphere is to

determine which lattice planes (represented by the grid points on the reciprocal lattice) will

result in a diffracted signal for a given wavelength, λ, of incident radiation.

The incident plane wave falling on the crystal has a wave vector 𝑘𝑖 whose length is 2π / λ. The

diffracted plane wave has a wave vector 𝑘𝑓 . If no energy is gained or lost in the diffraction

process (it is elastic) then 𝑘𝑓 has the same length as 𝑘𝑖 . The difference between the wave-

vectors of diffracted and incident wave is defined as scattering vector ∆𝑘 = 𝑘𝑓 − 𝑘 𝑖. Since 𝑘𝑖

and 𝑘𝑓 have the same length, the scattering vector must lie on the surface of a sphere of

radius 2π / λ. This sphere is called the Ewald sphere. A 2-dimensional representations is shown

in Fig. 2.13.

The reciprocal lattice points are the values of momentum transfer where the Bragg diffraction

condition is satisfied and for diffraction to occur the scattering vector must be equal to a

reciprocal lattice vector. Geometrically this means that if the origin of reciprocal space is

Chapter 2: Experimental procedure 35

placed at the tip of 𝑘𝑖 , then diffraction will occur only for reciprocal lattice points that lie on

the surface of the Ewald sphere.

Figure 2.16. Ewald sphere 2-dimensional representation. Green points are the reciprocal lattice ones.

2.5.1.1. Grazing incidence

The conventional Bragg-Brentano geometry (θ-2θ) may, however, not be the best

configuration for the analysis of nanostructures, since reflections coming from the substrate

could overlap the diffraction pattern of the nanostructure. The asymmetric Bragg geometry,

also known as grazing incidence geometry, is used in the analysis of polycrystalline materials

onto substrates, as is the case of the nanostructures of this work. In this configuration, the

incidence angle θ is fixed at a certain value (α). Low incidence angles reduce the penetration of

the incident X-rays within the material and, therefore, reflections coming from the substrate

are minimized.

To determine the appropriate incidence angle to use in the grazing incidence X-ray diffraction,

the critical angle, αc, i.e. the angle below which, the total reflection of the X-rays takes place,

and the penetration depth of the X-rays for different α values were calculated for the PbTiO3

nanostrucures. The critical angle was calculated from the Snell’s law in total reflection

conditions [28] and from the refraction index for X-rays of a material [29] according to the

following expression:

sin𝛼𝑐 = 2.6 · 𝑒−6𝜌𝜆2 (2.9)

where ρ is the theoretical density of the material and (7.97 g/cm3 in the case of PbTiO3), λ is

the wavelength of the X-ray beam (0.97513 Å for this synchrotron radiation). The total

reflection for the PbTiO3 will occur at 0.41 °.

·· · · · · ·

·· · · · · ·

·· · · · · ··· · · · · ··· · · · · ·

·· · · · · ·

θθ

θ

ki

kf

36 2.5. Structural characterization

Therefore, incident angles below αc cannot be used in this experiment, since the total

reflection is produced.

The penetration depth of the X-ray beams through the film is related to the absorption of the

sample as:

𝐼 = 𝐼0𝑒−𝜇𝑥

sin 𝛼 (2.10)

where µ is the X-ray absorption coefficient of the material, x is the thickness of the material

penetrated by the X-rays and α is the incident angle of the X-rays. The equivalent penetration,

Λ, is defined as the thickness that is penetrated by X-rays of Io intensity in order to decrease

their intensity by 1/e:

Λ =sin 𝛼

𝜇 (2.11)

The µ value calculated for the PbTiO3 is 0.022 µm-1. Therefore, the penetration calculated for a

various incidence angles are:

α (°) 0.450 0.500 0.750 1.000 2.000

Λ (µm) 0.357 0.397 0.595 0.793 1.586

According to these values, angles above the critical angle that produce total reflection, will

yield penetration of substrates by the X-ray beam. The equivalent penetration is calculated

also for the Pt (which is the last layer of the Pt/TiO2/SiO2/(100)Si substrates) and the SrTiO3,

being µ, 0.455 µm-1 and 0.010 µm-1, respectively.

α (°) 0.450 0.500 0.750 1.000 2.000

Λ(Pt) (µm) 0.017 0.019 0.029 0.038 0.077

Λ(SrTiO3) (µm) 0.785 0.872 1.309 1.745 3.490

Therefore, diffraction experiments should take into account the substrate contributions to the

2-dimensional diffraction patterns.

In this work synchrotron X-ray diffraction experiments were carried out in grazing incidence at

the beamline station 11-3 of the Stanford Synchrotron Radiation Lightsource (SSRL) in Stanford

(USA). A monochromatic beam of 12.7 keV was generated by a 26 pole wiggler, vertically and

horizontally focused by a Rh coated flat mirror and a single crystal (311) bent monochromator,

respectively. The diffracted radiation is detected by an imaging plate detector of 345 mm of

Chapter 2: Experimental procedure 37

diameter with a pixel size of 150 µm x 150 µm (MAR345 image plate, Marresearch GmbH). Fig.

2.17 schematizes the configuration of the experiment.

Image 2.17. Schematic illustration of the grazing incidence geometry used for X-ray diffraction.

Imaging plate detectors are a kind of 2-dimensional detectors in which a layer of BaF(Br,I):Eu2+

that contains color centers is disposed onto a robust film-like base “plate”. The plate is then

exposed to the X-rays. The image is later scanned by an online scanner. Scanning the image

consists of exciting the colors centers and then detecting the induced radiation. Stimulating

color centers does not require much energy, generally red laser suffices.

The great advantages of imaging plates are their large size, low cost and their high dynamic

range. The latter quality has made it the detector of choice for 2-dimensional diffraction. Their

major drawback is the dead time associated with the time-consuming scanning [30].

A 2-dimensional plate and the detecting mechanism are schematized in Fig. 2.18. Red, green

and white points of the reciprocal lattice are those that fulfill the Law condition of diffraction.

The corresponding vectors are projected to the plate and will result into the final pattern. Note

that only the points which projection intersects the plate will yield excitation of the plate pixels

and, therefore, observed diffraction peaks.

38 2.5. Structural characterization

Figure 2.18. Schematical representation of the interaction of an incident X-ray beam on a crystal and

the resulting diffraction peaks in the 2-dimensional plate for different planes of the reciprocal lattice

(l=0, l=1 and l=2).

One of the factors that determines the number of peaks observed is the distance between the

sample and the detector. In this work, the distance has been kept to the minimum in order to

detect the larger number of peaks possible to determine the orientation of the nanostructures.

Thus distances are as follow for the samples to be analyzed in the next chapters 3,4 and 5:

Type of analyzed particles Section in this work Sample-detector distance

Prepared by using the phenomenon of the instability of ultrathin fims onto Pt/TiO2/SiO2/(100)Si wafer substrates

3.2.2 150 mm

Prepared by the micromulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si wafer substrates

4.2.2 125 mm

Prepared by the micromulsion mediated synthesis onto microemulsion-layer/ Pt/TiO2/SiO2/(100)Si wafer substrates

4.3.2 305 mm

Prepared by the microemulsion mediated synthesis onto as-received (100)SrTiO3 substrates

5.2.2 360 mm

Chapter 2: Experimental procedure 39

Distances are varied depending on the relative intensities observed of the PbTiO3

nanostructures and the different substrates. It was explained before, that for angles above the

critical angles that produces total reflection, the penetration distance is larger than the height

of the nanostructures. The contribution of the substrates will depend not only on the

incidence angle but also on the orientation of the substrates crystallites or single-crystal planes

with respect to the incidence angle.

2-dimensional diffractions patterns are simulated by using Anaelu software [31-33]. When

crystals present a circular symmetry about some sample axis, they are said to have a fiber

texture. It is possible to calculate the inverse pole figure known the fiber texture and the unit

cell of crystals, and, from said inverse pole figure, the simulated diffraction pattern can be

calculated [34].

In this work, simulated diffraction patterns are calculated from the inverse pole figure and by

rotating the unit cell, using Anaelu software. The diagram flow of the iterative procedure for

the lattice parameters and texture of the nanostructures is schematized in Fig. 2.19.

Figure 2.19. Diagram flow of the iterative procedure for the lattice parameters (a) and texture of the

nanostructures (b).

b)

Simulated θ-2θ diffraction pattern

(PowderCell)

Comparasion with the integrated

pattern

Modification of lattice parameters

Experimental lattice parameters

a)Simulated 2-D

diffracion pattern (experimentals

parameters)(ANAELU)

Does ithave fiber

texture

Simulated diffraction pattern

from the IPF(ANAELU)

Comparasion with the integrated

pattern

Modification of texture

Experimental texture

determination

Simulated diffraction pattern

from the rotation of

the unit cell(ANAELU)

Comparasion with the integrated

pattern

Modification of texture

Experimental texture

determination

YES NO

Calculated latticeparameters fromthe experimental

patterns

Calculated latticeparameters fromthe experimental

patterns

Calculated latticeparameters fromthe experimental

patterns

40 2.5. Structural characterization

Fig. 2.20 shows a 2-dimensional pattern obtained when using the 2-dimensional diffraction

pattern.

Figure 2.20. 2-dimensional X-ray transmission pattern of LaB6 (a) and the profile along the yellow line of

previous image (b). LaB6 is the standard used for calibration.

Each radius is a conventional θ-2θ diffraction pattern. For any point in the pattern, the

intensity depends on orientation of the crystallographic planes nanostructures and is a

function of the distance to the center (the position of the peak) and the angle ψ. The

integration of the diffraction pattern is carried out by the equation:

𝐼 𝜃 = 𝐼 𝑟,𝜓 · 𝑑𝜓𝜓𝑓

𝜓𝑜 (2.12)

The diffraction pattern calculated this way contains information about all the diffraction peaks

in the pattern.

2.5.2. Auger electron spectroscopy

Auger electron spectroscopy (AES) is based on the effect of the same name, which starts by

removing an electron from an inner atomic shell (pink arrow in Fig. 2.21). Several processes

can produce this core hole, but bombardment with a high energy (2-5 KeV) electron beam is

the most common one. The core hole is then filled by a second atomic electron from a higher

shell (green arrow in Fig. 2.21). As the energy difference must be simultaneously released, a

third electron, the Auger electron, can escape from the solid (red arrow in Fig. 2.21), carrying

the excess energy in a radiationless transition. This process of an excited ion decaying into a

doubly charged ion by ejection of an electron is called the “Auger process”, and the

corresponding transition will be labeled with the name of three electronic levels involved (see

a) b)

Chapter 2: Experimental procedure 41

e.g. the L2M2,3M4,5 Auger transition of Titanium depicted in Fig. 2.21). Note that the kinetic

energy of the Auger electron only depends on the energies of these three orbitals (although

considering hole relaxation effects), and not on the excitation energy. Thus, it only depends

on the nature of the excited atom, and can be used as a “fingerprint” of the corresponding

element.

AES is a surface sensitive technique that takes advantage of the short mean free path of

electrons in matter: as the range of energies of the Auger electrons generally detected is

among 5-2000 eV, the mean free path of these electrons is on the order of some angstroms (4-

60 Å) and so, the information obtained comes from the first atomic layers. This technique

provides information about the composition and stoichiometry of the surface, being perfectly

suited for surface elemental analysis: since the Auger transitions are characteristic of each

element and their positions are tabulated, it is relatively easy to identify the elements present

at the surface. Note, nevertheless, that these energy levels are also sensitive to the chemical

environment of the excited atoms, and thus, “chemical shifts” can affect the Auger emission

lines, modifying the energy and line shape of the measured peaks.

Figure 2.21. Schematic illustration of the Auger process, indicating the energy levels involved. The

example corresponds to the L2M2,3M4,5 Auger transition of Titanium.

The intensity of a given Auger transition depends (among other factors) on the atomic

concentration of the respective element in each of the probed surface layers, and therefore, it

can be used to get quantitative information on the surface stoichiometry. There are significant

difficulties, however, in the proper absolute quantification of AES. A usual pragmatic approach

is to compare the ratios of the Auger signals (IA/IB , IC/IB,…) from different elements (A,B,C,…)

characteristic of a sample surface of known stoichiometry, with the respective ratios of other

Fermi level

Auger Process

3d M4,5

3d M2,3

3d M1

2p L3

2p L2

2p L1

1s

……

42 2.5. Structural characterization

sample surfaces whose stoichiometry we want to know (measured under similar experimental

conditions). This is the approach used here (see Chapter 5) for the analysis of SrTiO3 surfaces

prepared by different procedures.

The principal disadvantage of AES is that the incident electron beam charges up a non-

conducting sample. This effect, which also affects LEED measurements, may become relevant

in the case of STO surfaces.

The most usual form of measuring and representing AES spectra is that followed here (see e.g.

Fig. 5.15), which involves a differentiation with respect to energy, in order to suppress the

large secondary background on which the Auger signal is mounted, thus emphasizing the

detection of the Auger peaks.

Note that AES and LEED techniques require an ultra-high vacuum (UHV) environment, since

they involve the detection of relatively low energy electrons. In the present case, a commercial

rear view LEED optics (see Fig. 2.22), also adapted for AES mesurements, was available in the

same chamber used for substrate treatments in UHV, and it was employed for the surface

characterization of the STO samples. Auger spectra in the 20 to 550 eV range were taken in the

first derivative mode (dN/dE), using a 2 KeV incident electron beam. Special care was placed to

use the same measurement parameters (emission current, sample to detector distance, peak-

to-peak voltage, amplitude or sensitivity, pass energy, d-well time, etc …) for all the surfaces

under study.

2.5.3. Low energy electron diffraction

Low Energy Electron Diffraction (LEED) is a technique well suited for the determination of the

surface structure of crystalline materials. Surfaces are bombarded with a collimated beam of

low energy electrons (30-500eV). The wave lengths that correspond to those energies are of

the order of the interplanar distances, and their inelastic dispersion cross sections are maxima,

being the electron mean free path in the range of 10-20 Å (i.e., less than 6 atomic layers).

Therefore, the detected diffracted electrons come from the first atomic layers, making LEED a

classic technique for the crystallographic analysis of surfaces.

Figure 2.22 schematizes a conventional LEED system, similar to that used here, where the

sample is placed at the centre of the optic axis. The backscattered electrons are detected by a

fluorescent semispherical screen, generating the diffraction pattern. The LEED detector usually

contains three or four hemispherical concentric grids (used for screening out the inelastically

scattered electrons) and a phosphor screen or other position-sensitive detector.

Chapter 2: Experimental procedure 43

Figure 2.22. Conventional LEED system.

If the incident beam is normal to the sample surface, the pattern observed on the fluorescent

screen is a direct picture of the reciprocal lattice of the surface. The size of the Ewald's sphere

and hence, the number of diffraction spots on the screen is controlled by the incident electron

energy. From the knowledge of the reciprocal lattice, information on the real space lattice can

be derived, at least qualitatively (in terms of the surface periodicity and the point group).

2.6. Ferro-piezo electric characterization: Piezoresponse Force

Microscopy.

Piezoresponse Force Microscopy (PFM) is based on the detection of local vibrations of a

sample induced by an electric field applied between the conductive tip of the scanning force

microscope and the bottom electrode of the sample [35, 36]. The local oscillations of the

sample surface are transmitted to the tip and detected using a usual lock-in technique. The

signal at the lock-in output is denoted piezoresponse signal (PRS) [35]. PFM is the only

available technique suitable for the functional characterization of the nanostructures onto

substrates.

Voltage applied to the tip can be expressed as

𝑉𝑡𝑖𝑝 = 𝑉𝑑𝑐 + 𝑉𝑎𝑐 sin𝜔𝑡 (2.8)

The piezoelectric response of the surface is detected as a first harmonic component of bias-

induced tip deflection:

𝑑 = 𝑑0 + 𝐴 · cos 𝜔𝑡 + 𝜙 (2.9)

Electron gunSample

Fluorescent screen

Electron energyselecting grids

Incident beam

Elastically diffractedelectrons

44 2.6. Ferro-piezoelectric characterization

PFM is carried out in contact mode, under a constrained force. Fig. 2.23 shows the

configuration used in this work. PFM measurements were carried out using the SFM

microscope previously described in section 2.4.5 implemented with a SR7265 DSP lock-in

amplifier (Signal Recovery) and a Agilent 33120 function waveform generator. The AC signal is

applied through the bottom electrode.

Figure 2.23. PFM configuration for the obtaining of PFM images and hysteresis loops.

Conductive commercial Pt/Ir coated tips (Nanosensors) on cantilevers were used to apply an

AC voltage of 1-3V at 50 kHz and 245 kHz. These cantilevers are summarized in Fig. 2.24.

Figure 2.24. Nanosensors probes for the acquisition of PFM images. PPP-CONT tips present a higher

resolution due to the fact that the tip is clearly visible from the top, making more accurate the

positioning of the laser spot onto it.

In PFM, cantilevers not only are the probes to measure the reverse piezoelectric effect but also

act as temporally top electrode for the measurement. When the polarization is parallel to the

applied electric field, the deflection of the cantilever takes place in the out-of-plane direction,

and gives rise to a vertical movement of the laser spot in the photosensitive diode. The in-

plane deflection of the cantilever is due to a polarization perpendicular to the applied field. In

detector

samplesubstrate

Laser beam

V(t)

Lock-in amplifier

Function generator

signal output

feedback

reference

A B

C D

PPP-NCHPt• Pointprobe® Plus - tip• Non-Contact High frequency

• C = 42 N/m, fo = 330 kHz• PtIr5 coated probe

ATEC-EFM• Pointprobe® Plus - tip• Contact Mode

• C = 0.7 N/m, fo = 503 kH

Chapter 2: Experimental procedure 45

this case, a horizontal movement of the laser spot is detected by the photosensitive diode.

When the polarization vector presents both parallel and perpendicular contributions, both the

in-plane and the out-of-plane signals are detected.

When the polarization vector pointing downwards (i.e., c- domains), the application of a

positive tip bias results in the expansion of the sample and bias-induced surface oscillations

are in phase with tip voltage f=0. For polarization pointing up-wards (i.e., c+ domains) f =180°.

In the case of the in-plane response, that is to say, when the response is orthogonal to the

applied field, the application of both positive and negative bias results into a shear movement

of the sample [37]. Fig. 2.25 schematizes these effects as well as the most common case of

polarization that presents both the in-plane and out-of-plane contributions.

Figure 2.25. Scheme of the out-of-plane, in-plane and mixed response of a ferroelectric material when

applying an electric field perpendicular to the surface of the nanostructure [38].

46 2.6. Ferro-piezoelectric characterization

PFM signal can be enhanced by measuring near the resonant frequency of the cantilever.

When in contact, the resonance is slightly enhanced near the free resonance of the cantilever

and highly enhanced near the contact resonance of the cantilever [39, 40].

2.6.1. Image acquisition

Images are acquired by scanning the surface of the sample in contact mode while inducing the

vibration of the sample applying an AC electric field between the top electrode (the conductive

tip) and the bottom electrode (the substrate of the sample) [35].

When the tip is in contact, both electrostatic and electromechanical interactions contribute.

The total response contains both local and non-local contributions. Fig. 2.26 shows the

piezoresponse, amplitude and phase signals when there is electromechanical interaction only,

when there is, in addition, a weak electrostatic contribution and when the electrostatic

interaction is dominant.

Figure 2.26. Piezoresponse, amplitude and phase signals for different cases of electromechanical and

electrostatic interactions [41].

Electromechanicalinteraction only

Weak electrostaticcontribution

Dominant electrostaticcontribution

PR

SA

mp

litu

de

Ph

ase

Chapter 2: Experimental procedure 47

2.6.2. Hysteresis loops.

PFM microscopes can be used in order to obtain local piezoelectric hysteresis loops. There are

two different approaches for obtaining PFM hysteresis loops. In the first one, an increasing

bias is applied in a staircase way, while measuring the PRS or the decoupled amplitude and

phase signals in a stationary state with a small amplitude AC signal. This is called in-field

hysteresis loop (Fig. 2.27 (a)).

In the second type of hysteresis loop, the bias is applied in pulses of increasing or decreasing

value, while the AC electric field is kept constantly on. This kind of hysteresis loops is called

out-of-field or remnant hysteresis loop because the PRS or the decoupled amplitude and phase

signals are measured in the remnant steady state (Fig. 2.27 (b)).

Figure 2.27. Shape of the electric field supplied to the substate for the in-field (a) and out-of -field (b)

local piezoelectric hysteresis loops. AC field is represented in dark red and DC field in green.

a)

b)

48 2.6. Ferro-piezoelectric characterization

Both kinds of hysteresis loops provide complementary information. While the in-field

hysteresis loops shows clearly the electrostatic contribution (as a straight line that passes

through the origin and has a slope proportional to said contribution), the out-of-field ones

provides remnant information.

Measurements are made using the tip of the cantilever as top-electrode. This implies that the

applied field is not uniform with field lines resembling an umbrella. Thus, the measured values

of d33 must be considered as effective values [35].

Bibliography 49

Bibliography

[1] N.J. Phillips, M.L. Calzada and S.J. Milne, "Sol gel-derived lead titanate films", Journal of

non crystaline solids, 147&148, 1992, p:285

[2] R. Sirera, "Síntesis por sol-gel de soluciones de titanato de plomo modificado para la

preparación de láminas delgadas ferroeléctricas", 1997, Madrid: Universidad Autónoma de

Madrid. Ph.D. Thesis.

[3] M. Kawasaki, K. Takahashi, T. Maeda, R. Tsuchiya, M. Shinohara, O. Ishiyama, T.

Yonezawa, M. Yoshimoto and H. Koinuma, "Atomic control of the SrTiO3 crystal-surface",

Science, 266 (5190), 1994, p:1540

[4] T. Ohnishi, K. Shibuya, M. Lippmaa, D. Kobayashi, H. Kumigashira, M. Oshima and H.

Koinuma, "Preparation of thermally stable TiO2-terminated SrTiO3 (100) substrate surfaces",

Applied Physics Letters, 85 (2), 2004, p:272

[5] D.E. Bornside, C.W. Macosko and L.E. Scriven, "On the modeling of spin coating",

Journal of Imaging Technology, 13 (4), 1987, p:122

[6] J.I. Goldstein, D.E. Newbury, P. Echlin, D.C. Joy, C. Friori and E. Lifshin, Scannig Electron

Microscopy and X-ray microanalysis. 1981, New York: Plenum Press.

[7] D.B. Williams and C.B. Carter, Transmision Electron Microscopy. 1996, New York:

Plenum Press.

[8] G. Binnig, C.F. Quate and C. Gerber, "Atomic Force Microscope", Physical Review

Letters, 56, 1986, p:930

[9] S.V. Kalinin, R. Shao and D.A. Bonnell, "Local phenomena in oxides by advanced

scanning probe microscopy ", Journal of the American Ceramic Society, 88 (5), 2005, p:1077

[10] R. Garcia and R. Perez, "Dynamic atomic force microscopy methods", Surface Science

Reports, 47 (6-8), 2002, p:197

[11] P.K. Hansma, J.P. Cleveland, M. Radmacher, D.A. Walters, P.E. Hillner, M. Bezanilla, M.

Fritz, D. Vie, H.G. Hansma, C.B. Prater, J. Massie, L. Fukunaga, J. Gurley and V. Elings, "Tapping

mode atomic-force microscopy in liquids", Applied Physics Letters, 64 (13), 1994, p:1738

[12] C.A.J. Putman, K.O. Vanderwerf, B.G. Degrooth, N.F. Vanhulst and J. Greve, "Tapping

mode atomic-force microscopy in liquid", Applied Physics Letters, 64 (18), 1994, p:2454

[13] Q. Zhong, D. Inniss, K. Kjoller and V.B. Elings, "Fractured polymeric silica fiber surface

studied by tapping mode atomic-force microscopy", Surface Science, 290 (1-2), 1993, p:L688

[14] R.D. Jaggi, A. Franco-Obregon, P. Studerus and K. Ensslin, "Detailed analysis of forces

influencing lateral resolution for Q-control and tapping mode", Applied Physics Letters, 79 (1),

2001, p:135

50 Bibliography

[15] I. Horcas, R. Fernandez, J.M. Gomez-Rodriguez, J. Colchero, J. Gomez-Herrero and A.M.

Baro, "WSXM: A software for scanning probe microscopy and a tool for nanotechnology",

Review of Scientific Instruments, 78 (1), 2007, p:013705

[16] R.E. Blahut, Fast Algorithms for Digital Signal Processing. 1985, New York: Addison-

Wesley.

[17] S.K. Kurtz and F.M.A. Carpay, "Microstructure and normal grain growth in metals and

ceramics. Part II: Experiments", Journal of Applied Physics, 51 (11), 1980, p:5745

[18] J. Ricote and L. Pardo, "Improvement of calcium modified lead titanate piezoceramics

by hot isostatic pressing", Journal of the European Ceramic Society, 20 (11), 2000, p:1677

[19] A. Moure, A. Castro and L. Pardo, "Improvement by recrystallisation of Aurivillius-type

structure piezoceramics from mechanically activated precursors", Acta Materialia, 52, 2004,

p:945

[20] J. Ricote and L. Pardo, "Microstructure-properties relationships in samarium modified

lead titanate piezoceramics--I. Quantitative study of the microstructure", Acta Materialia, 44

(3), 1996, p:1155

[21] J. Ricote, C. Alemany, L. Pardo and C.E. Millar, "Microstructure-properties relationships

in samarium modified lead titanate piezoceramics--II. Dielectric, piezoelectric and mechanical

properties", Acta Materialia, 44 (3), 1996, p:1169

[22] J. Ricote, C. Alemany and L. Pardo, "Microstructural effects on dielectric and

piezoelectric behavior of calcium-modified lead titanate ceramics", Journal of Materials

Research, 10 (12), 1995, p:3194

[23] L. Fuentes, Synchrotron Radiation Diffraction and Scattering in Ferroelectrics, in

Multifunctional polycrystalline ferroelectric materials, L. Pardo and J. Ricote, Editors. in press,

Springer-Verlag

[24] H. Wiedemann, Synchrotron Radiation. 2003, Berlin: Springer-Verlag.

[25] J.D. Bernal, "The structure of graphite", Proceedings of the Royal Society of London

Series a-Containing Papers of a Mathematical and Physical Character, 106 (740), 1924, p:749

[26] P.P. Ewald, "Remark on the work by M Laue - The trinumeric-symmetric x-ray pictures

on regular crystals", Physikalische Zeitschrift, 14, 1913, p:1038

[27] P.P. Ewald, "The theory of the interference of X-rays in crystals", Physikalische

Zeitschrift, 14, 1913, p:465

[28] F.A. Jenkins and H.e. Whtie, Fundamentals of Optics. 1976, Tokyo: Mc Graw-Hill.

[29] B.k. Vainshtein, Modern Crystallography I. 1981, Berlin: Springer-Verlag.

[30] J.K. Cockcroft and A.N. Fitch, Experimental Setups, in Powder Diffraction: Theory and

Practice D. R.E. and S. Billinge, Editors. 2008, The Royal Society of Chemistry: Cambridge.

Bibliography 51

[31] L. Fuentes-Montero, "Software “Anaelu” para Análisis de Patrones Bidimensionales de

Difracción de Rayos X", Fisica de Materiales, 2008, Chihuahua: Centro de Investigacion en

Materiales Avanzados, S. C. Ph.D. Thesis.

[32] L. Fuentes-Montero and L. Fuentes-Cobas. "Modelling of texture effect on 2d

diffraction patterns". in SSRL/LCLS User's conference. 2009. SSRL.

[33] L. Fuentes-Montero, M.E. Montero-Cabrera, L. Calzada, M.P. De la Rosa, O. Raymond,

R. Font, M. Garcia, A. Mehta, M. Torres and L. Fuentes. "Synchrotron Techniques Applied to

Ferroelectrics: Some Representative Cases". in Symposium on Ferroelectricity and

Piezoelectricity held at the 15th International Materials Research Congress (IMRC). 2006.

Cancun, MEXICO.

[34] L. Fuentes, "Anomalous scattering and null-domain ghost corrections for fibre

textures", Textures and Microstructures, 10, 1989, p:347

[35] C. Harnagea, A. Pignolet, M. Alexe and D. Hesse, "Piezoresponse scanning force

microscopy: what quantitative information can we really get out of piezoresponse

measurements on ferroelectric thin films", Integrated Ferroelectrics, 44, 2002, p:113

[36] A. Gruverman, O. Aucello and H. Tokumoto, "Imaging and control of domain structures

in ferroelectric thin films via scanning force microscopy ", Annual review of materials science,

28, 1998, p:101

[37] G. Catalan, B. Noheda, J. McAneney, L.J. Sinnamon and J.M. Gregg, "Strain gradients in

epitaxial ferroelectrics", Physical Review B, 72 (2), 2005, p:020102

[38] A. Rüdinger, S. T., R. A., T. S., S. T. and W. R., "Nanosize ferroelectric oxide - tracking

down the superparaelectric limit", Applied physics A-Materials science & processing, 80 (6),

2005, p:1247

[39] S. Jesse, B. Mirman and S.V. Kalinin, "Resonance enhancement in piezoresponse force

microscopy: Mapping electromechanical activity, contact stiffness, and Q factor", Applied

Physics Letters, 89 (2), 2006,

[40] T. Stoica, R. Calarco, R. Meijers and H. Luth, "Nanoscale imaging of surface

piezoresponse on GaN epitaxial layers", Applied Surface Science, 253 (9), 2007, p:4300

[41] S.V. Kalinin and D.A. Bonnell, Electric scanning probe imaging and modification of

ferroelectric surfaces, in Nanoscale Characterisation of Ferroelectric Materials. Scanning probe

microscopy approach, M. Alexe and A. Gruverman, Editors. 2004, Springer: Berlin. p. 282.

CHAPER 3: FERROELECTRIC NANOSTRUCTURES BY THE

PHENOMENON OF THE MICROSTRUCTURAL

INSTABILITY OF POLYCRYSTALLINE ULTRATHIN FILMS

3.1. The microstructural instability of polycrystalline

ultrathin films

Ultrathin ferroelectric oxide films are those with a thickness below 50 nm [1-3]. They present a

microstructural instability, a phenomenon that makes possible to obtain isolated

nanostructures onto substrates. This effect was first modeled by Lange and co-workers for

epitaxial ultrathin films of PbTiO3 onto SrTiO3 single crystal substrates [4-6].

When the thickness of an ultrathin film is below a critical value, the microstructural instability

causes the film to break into isolated grains in order to lower the free energy of the system, as

explained through the diagram in Fig. 3.1.

CONTINUOUS FILM

Figure 3.1 Spheroidization of a uniform 2-dimensional ultrathin film of initial thickness t and grain size D

(left), as the thermal energy provided to the system increases , and its 3-dimensional analysis (right) [6].

a)

c)

e)

g)

b)

d)

f)

h)

54 3.1. The microstructural instability of polycrystalline ultrathin films

The model film (Fig 3.1 left) is composed of uniform grains of initial grain size D and thickness

t. These grains were allowed to spheroidize at constant volume and the grain centers of mass

are assumed to remain fixed by the substrate. Initially, the configurational change can be

described by the angle between adjacent grains, Ψ, which is allowed to decrease from an

initial value of π degrees. Calculations show that films with an initial D/t ratio less than 8/π will

retain a boundary between the grains even when Ψ decreases to 0 degrees. Films with a D/t

ratio larger than 8/π will, however, reach a point where the grain boundary disappears.

By thermodynamic calculations, Miller et al. [6] demonstrate that the breakup of the ultrathin

film lowers the free energy of the system when the grain-size to film-thickness exceeds a

critical value.

Figure 3.2. Nanoparticles of Pb(Ti, Zr)O3 (PZT) onto SrTiO3 volume distribution [7].

Later on, Dawber et al. [7] and Harnagea et al. [8] prepared ferroelectric nanoislands onto

conductive polycrystalline Si-based substrates. They found that the nanostructures show a

different morphology depending on the sol-gel precursor solution concentration, that is to say,

on the initial thickness and the crystallization temperature [7]. Further studies found that PZT

nanostructures prepared following this procedure do not show single crystal quality [9]. A

scheme of their results is showed in Fig. 3.2.

Chaper 3: Nanoparticles fabrication by the phenomenom of the microstructural instability of polycrystalline ultrathin films

55

Chemical Solution Deposition is a low cost method that enables the preparation of complex

mixed oxide materials supported onto substrates with a well-adjusted stoichiometry. In this

process, the liquid precursors are deposited onto the substrate, obtaining an amorphous

precursor layer that, after thermal treatment leads to a polycrystalline film. This is a powerful

technique for the fabrication of thin and ultrathin films that has been successfully used since

the early 80’s for polycrystalline ferroelectric films.

Using sol-gel, the precursor solution is spun-coated on a substrate to obtain a thin

metalorganic gel layer that subsequently is converted into an amorphous oxide film by a

thermal annealing process. During this thermal treatment of crystallization, depending on film

parameters and annealing conditions, the system transforms either into a continuous ultrathin

film or it patterns itself into nanostructures of a large variety of shapes and sizes. PbTiO3

ultrathin films develop holes during this crystallization process. Upon further annealing, these

holes grow to a stable size, or even cause the film to break up into single crystal islands under

the conditions described previously [9].

This chapter is presented with the aim of a comparative study of the characteristics of the

nanostructures resulting from this procedure and those, original to this thesis, that will be

presented in the following chapters. The crystal structure, microstructure and properties of

the nanostructures prepared from ultrathin films below that critical thickness have been

studied by AFM, SEM, Synchrotron Radiation Diffraction and PFM.

3.2. Nanoscale PbTiO3 structures onto Pt/TiO2/SiO2/(100)Si

substrates prepared by using the phenomenon of the

microstructural instability

The lead titanate precursor solution used in this work has been synthesized by the process

described previously in Chapter 2 (Experimental procedure, Section 2.1.1). In this study,

different sol concentrations are used: 10-1 M, 4·10-2 M, 3·10-2 M, 2·10-2 M, 10-2 M and 5·10-3 M

for the preparation of the nanostructures onto Pt/TiO2/SiO2/(100)Si substrates.

3.2.1. Microscopy and quantitative microstructure analysis

The ultrathin film breaks for a sol dilution below 5·10-3 M when using 1,3-propanodiol as

solvent, as referred in the literature [10, 11]. Sample prepared from the sol dilution 10-1 M is

56 3.2. Nanoscale PbTiO3 structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by using the

phenomenon of the microstructural instability

shown in Fig. 3.3 to illustrate the appearance of a PbTiO3 continuous ultrathin film as seen by

AFM.

Figure 3.3. AFM topography images of a PbTiO3 continuous ultrathin film obtained at different

magnifications: 5x5 µm image (a) and 2x2 µm image (b).

Fig. 3.4 shows the AFM topography images of different samples prepared from sols with

different concentrations, in order to establish the growth mechanism of the particles and for

the sake of comparison with the methods that will follow. For the highest concentration shown

in the figure 4·10-2 M - rounded particles are grown not covering the substrate uniformly. Fig

3.4 (a) shows the topography of this sample at the outer part of the coating, where the

isolated nanostructures can be found. As the concentration decreases (Fig. 3.4 (c)), the

continuous coated area ceases to exist. The particles shape is less rounded, as shown in the

profile plot, with a flat top facet (Fig. 3.4(d)), indicating that the instability phenomenon takes

place throughout the substrate. The resulting structures for a sol with a concentration of

2·10-2 M are needle-shape type that preferentially nucleates around the substrate defects, in

the radial direction. For sol concentrations below 10-2 M, the resulting structures are flat grains

of irregular shape, and lateral dimensions bigger than 200 nm, as shown in the AFM profile

(Fig. 3.4(h)). As in the case of the needles, they preferentially nucleate around the defects.

When the ultrathin film breaks, the resulting particles try to grow with the minimum surface

energy configuration [4-6]. In the case of the PbTiO3 nanostructures onto SrTiO3 single crystal

substrates, this configuration is a pyramid [12], while in the case of PbTiO3 nanostructures

onto Pt/TiO2/SiO2/(100)Si substrates, the minimum surface energy configuration is a rounded

shape [7]. The different shape, which depends on the substrate, is due to the different degree

of mismatch between the lattices of the substrates and that of the nanostructures as well as to

the different relative surface tension values.

1.0µm

21.52 nm

0nm

400nm

8.64 nm

-7.80 nm

Chaper 3: Nanoparticles fabrication by the phenomenom of the microstructural instability of polycrystalline ultrathin films

57

Concentration Topography Profile

4·10-2 M

3·10-2 M

2·10-2 M

10-2 M

5·10-3 M

Figure 3.4 AFM topography images and representative profiles of the PbTiO3 nanostructures onto the

Pt/TiO2/SiO2/(100)Si substrates prepared from sols with different concentrations.

20.92 nm

0.00 nm

400350300250200150100500

8

7

6

5

4

3

2

1

0

X[nm]

Z[n

m]

600nm

38.54 nm

0.00 nm

400350300250200150100500

20

15

10

5

0

X[nm]

Z[n

m]

600nm

27.94 nm

0.00 nm

400350300250200150100500

8

7

6

5

4

3

2

1

0

X[nm]

Z[n

m]

longitudinal direction

transversal direction

600nm

82.48 nm

0.00 nm

400350300250200150100500

35

30

25

20

15

10

5

0

X[nm]

Z[n

m]

600nm

64.90 nm

0.00 nm

400350300250200150100500

15

10

5

0

X[nm]

Z[n

m]

a) b)

c) d)

e) f)

g) h)

i) j)

58 3.2. Nanoscale PbTiO3 structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by using the

phenomenon of the microstructural instability

When decreasing the concentration of the precursor sol to 3·10-2 M, it is observed that some

areas contain isolated particles, which are rounded, and other areas have a group of non

isolated structures that have a flat top facet (Fig. 3.4). If the amount of material is low enough,

the continuity of the ultrathin gel layer will be lost before any thermal treatment is carried out.

One of the parameters that is critical for the evolution of the discontinuity of the films is how

apart are these areas: as they grow, the mass that was initially occupying these areas must be

redistributed, increasing the thickness of the film and, maybe, filling in some other areas. So,

for some areas, the film thickness will be increased dramatically, meaning that the particles

will not be able to complete the isolation and subsequent spheroidization and leading to these

non isolated grains with a flat top facet.

For the sample prepared from the 2·10-2 M diluted sol, there is not enough material to form

the initial amorphous continuous ultrathin film, so precursor drops nucleus of crystallization

will appear on heating and form needles, which is a common crystalline growth habit of PbTiO3

crystals.

Concentration Before the crystallization After the crystallization

4·10-2 M

non homogeneous

3·10-2 M

2·10-2 M

<10-2 M

Figure 3.5. Proposed growth evolution of the particles deposited as the concentration of the solutions

decreases.

When the concentration of the diluted sol is 10-2 M or smaller, all the precursor material will

set around the defects (Figure 3.4). When the samples are crystallized, large and flat grains will

be formed around them by coalescence, giving rise to larger particles, but in a smaller number.

The morphologies of the particles in all four growth stages are summarized in Fig. 3.5.

Substrate

D

t

Ψ

Substrate

D

t

Ψ Ψ

Substrate

θ

Substrate Substrate

Substrate

?

Substrate

Chaper 3: Nanoparticles fabrication by the phenomenom of the microstructural instability of polycrystalline ultrathin films

59

Previous works [7, 8] have been devoted to the growth of nanostructures as a function of the

initial film thickness and the crystallization temperature. The appearance of flat-faceted

structures has been observed [7], as well as the preferential nucleation of lead titanate

particles at the substrate imperfections, such as grain boundaries of polycrystalline Pt-coated

substrates, when the material amount is small enough [13]. However, to the knowledge of this

author, the formation of needles prior to that effect has never been reported, which may be

due to the different morphology of the platinum coating of the substrates used for these

samples and the ones used in the literature: while the ones used for the above samples have

small grains (smaller than the tips of AFM cantilevers) and roughness of about 25-30 nm, the

ones used at the works referred have flat grains of some hundreds of nanometers.

In Section 2.4.6, it was explained how to calculate the average size of a given distribution of

size of an object, <s>, and the standard deviation of a lognormal distribution, σs, from the

fitting of the line obtained at the probability plot of such distribution. Here the size of the

nanostructures can be well defined at high concentrations by the equivalent diameter to their

circular shape, since they have more or less the same dimensions in all directions (Fig. 3.4(a) y

(c)), whereas such a parameter will not be valid for needle-shape particles.

From the AFM images of the samples deposited from the 4·10-2 M and 3·10-2 M diluted sols,

the distribution of the particle size was measured using MIP4 software by Digital Image

Systems. Fig 3.6 shows the equivalent diameter distributions of the nanoparticles as well as

the corresponding probability plots. Both distributions are log-normal and the fitting of their

probabilistic lines is analitically expressed by:

y = -18.88 + 4.85·x R = 0.99 (3.1)

for the nanostructures deposited from the 4·10-2 M sol and

y = -13.49 + 3.52·x R = 0.99 (3.2)

for those deposited from the 3·10-2 M sol.

Following this procedure, the average equivalent diameter of the nanostructures derived from

the 4·10-2 M and 3·10-2 M sols are 50 nm and 48 nm with a standard deviation of 11 nm and 14

nm, respectively.

60 3.2. Nanoscale PbTiO3 structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by using the

phenomenon of the microstructural instability

20 30 40 50 60 70 80 90 100

0

10

20

30

40

Co

un

t

Deq (nm)

20 40 60 80 100

0

5

10

15

20

25

30

35

40

45

50

Co

un

t

Deq

(nm)

0,01

0,5

2

10

30

50

70

90

98

99,5

3,2 3,4 3,6 3,8 4,0 4,2 4,4 4,6 4,8

3,2 3,4 3,6 3,8 4,0 4,2 4,4 4,6 4,8

0

10

20

30

40

Co

un

ts

Ln(Deq

)

Cu

mu

lative

Fre

qu

en

cy

Figure 3.6 Equivalent diameter distributions of samples prepared from the 4·10

-2 M (a) and 3·10

-2 M (b)

sol-gel solutions and their corresponding log-normal distributions (c) and (d).

The equivalent diameter distributions are log-normal, as it is currently found in the bulk

ceramics when mass transport occurs at high temperature at the surface of the original

powder particles, allowing the growth of big particles by disappearance of the small ones

around them and subsequent disappearance of the interparticle space, in the well-known

growth phenomenon of sintering [14, 15]. In thin films technologies the literature refers to a

similar phenomena of coalescence of nuclei of growth onto the substrate. The lognormal

shape of the size distributions of the nanoparticles obtained here points to the occurrence of

mechanisms in the formation of the nanostructures, where the diffusion of mass at high

0.01

0.5

2

10

30

50

70

90

98

99.5

3.2 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8

3.2 3.4 3.6 3.8 4.0 4.2 4.4 4.6 4.8

0

10

20

30

40

50

Co

un

ts

Ln(Deq)

Cu

mu

lative

Fre

qu

en

cy

c) d)

a) b)

Chaper 3: Nanoparticles fabrication by the phenomenom of the microstructural instability of polycrystalline ultrathin films

61

temperature and coalescence of the structures play an important role during their growth

process.

From the Fig. 3.6 (a) and (b), one can observe how the distribution is becoming more log-

normal and broadening when the concentration of the diluted sol used for these samples

decreases. This behavior has been reported previously [4].

3.2.2. Structural characterization

Fig. 3.7 shows the 2-D diffraction pattern obtained with synchrotron X-ray diffraction in grazing

incidence configuration at the Stanford Synchrotron Radiation Lightsource for the continuous

ultrathin film prepared from the 2·10-2 M sol, using a grazing incidence configuration.

This pattern consists of some sharp rings plus some broad ring portions, located in angle values

close to the said sharp rings (Fig. 3.7 (a)). Furthermore, there are four clear broad spots; three

of them located at the center of the diffraction pattern (indicated by a black point at their

centre) and the fourth on the second quadrant, at low angle positions (Fig. 3.7 (b)).

The integration of the experimental diffraction pattern (as explained in the Experimental

Chapter, 2.5.1.3) is shown in Fig. 3.8. The PbTiO3 perovskite pattern is shown in green and the

integrated pattern is the solid black line. Each peak corresponds to one of the sharp rings or

broad ring portions of the 2-D diffraction pattern of Fig. 3.7. It presents four very intense peaks

as well as some other peaks that are less intense. The simulation of the diffraction pattern for

the crystalline phases present at the sample is shown in colors. The most intense peaks

correspond to the polycrystalline platinum of the substrate (Pt/TiO2/SiO2/(100)Si). The less

intense peaks can be explained by the PbTiO3 perovskite phase. One can observe that the four

most intense peaks have a small contribution of this phase as well. Reflections corresponding

to the PbTiO3 perovskite phase as well as to platinum and titanium oxide are detected in the

diffractogram of Fig. 3.8.

62 3.2. Nanoscale PbTiO3 structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by using the

phenomenon of the microstructural instability

Figu

re 3

.7 E

xper

imen

tal 2

-D S

ynch

rotr

on

dif

frac

tio

n p

atte

rn o

f a

con

tin

uo

us

ult

rath

in f

ilm.

b)

a)

··

·

Chaper 3: Nanoparticles fabrication by the phenomenom of the microstructural instability of polycrystalline ultrathin films

63

Figu

re 3

.8 2

θ d

iffr

acti

on

pat

tern

ob

tain

ed f

rom

th

e in

tegr

atio

n o

f th

e 2

-D e

xper

imen

tal p

atte

rn a

nd

th

e si

mu

late

d d

iffr

acti

on

pat

tern

s o

f th

e P

bTi

O3,

pla

tin

um

,

TiO

2 an

d s

ilico

n o

f th

e su

bst

rate

.

1015

2025

3035

4045

50

3080

422

1540

211

0

Po

wd

er

Ce

ll

2.

2

AN

AT

AS

E 3

.6%

101

103

004

112

200

202

105

211

213

116

220

107

215

301

206

008

303

312

PbTiO3 14.3%

001

100

101

110

111

002

200

102

211

202

003

221

301

Pt 7.1%

111

200

220

311

RUTILE 3.6%

110

101

200

111

220

002

310

212

Si 71.4% (pref.Or)

111

220

222

331

APT18_01_41.X_Y

1015

2025

3035

4045

50

3080

422

1540

211

0

Po

wd

er

Ce

ll

2.

2

AN

AT

AS

E 3

.6%

101

103

004

112

200

202

105

211

213

116

220

107

215

301

206

008

303

312

PbTiO3 14.3%

001

100

101

110

111

002

200

102

211

202

003

221

301Pt 7.1%

111

200

220

311

RUTILE 3.6%

110

101

200

111

220

002

310

212

Si 71.4% (pref.Or)

111

220

222

331

APT18_01_41.X_Y

64 3.2. Nanoscale PbTiO3 structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by using the

phenomenon of the microstructural instability

Figure 3.9-11 shows the more detailed identification of the rings, ring sectors and spots of the

experimental diffractogram shown in Fig. 3.7.

Figure 3.9 Simulated 2-D diffraction pattern of polycrystalline platinum with (111) fiber texture (a) and

the experimental 2-D diffraction pattern measured from the ultrathin film (b).

Figure 3.9 (a) shows the simulated diffraction pattern for a polycrystalline platinum with a

(111) fiber texture with an orientation distribution cone of ±3°, calculated from the inverse

pole figure (see Chapter 2, Experimental Procedure, Section 2.5.1.2, for further explanations).

The intense peak that corresponds to the (331) direction, shown in the simulation by a blue

line over a weak spot.

Figure 3.10 (a) shows the simulated diffraction pattern for (100) Si. It was calculated from the

Inverse Pole Figure. It explains the three interconnected spots at the central part of the

diffractogram, that correspond to the diffraction of the plane families indicated at Fig. 3.10

(a). The lines that interconnect the spots can be understood as diffuse radiation, which is

related to high ordered structures [16] as is the case of the (100)Si wafer. There is another very

intense peak that corresponds to the (331) direction, shown in the simulation by a blue line

over a weak spot.

Figure 3.10 Simulated 2-D diffraction pattern of (100) Si single crystal (a) and the measured 2-D

diffraction pattern from the sample (b). The diffractions maxima that forms the triangle are marked

with a cross (see also Fig. 3.7 (b)).

(311)

(220)

(200)

(111)

a) (311)

(220)

(200)

(111)

b)

a)

(114)

(2-12)(-122)

(331)

(114)

(2-12)(-122)

(331)+

++

b)

Chaper 3: Nanoparticles fabrication by the phenomenom of the microstructural instability of polycrystalline ultrathin films

65

Figu

re 3

.11

Sim

ula

ted

2-D

dif

frac

tio

n p

atte

rn o

f P

bTi

O3

nan

ost

ruct

ure

s w

ith

(1

00

) fi

ber

tex

ture

an

d a

n o

rien

tati

on

dis

trib

uti

on

co

ne

of

±15

° (a

) an

d t

he

mea

sure

d

2-D

dif

frac

tio

n p

atte

rn f

rom

th

e sa

mp

le (

b).

(22

2)

(11

3)

(10

3)

(00

2)

(11

0)

(10

0)

(20

0)

(21

1)

(00

3)

(11

2) (1

11

)

(20

2)

(20

0)

b)

(22

2)

(11

3)

(10

3)

(00

2)

(11

0)

(10

0)

(20

0)

(21

1)

(00

3)

(11

2)

(11

1)

(20

2)

(20

0)

a)b

)

66 3.2. Nanoscale PbTiO3 structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by using the

phenomenon of the microstructural instability

Fig. 3.11 (a) shows the simulated 2-D diffraction pattern of the PbTiO3 perovskite phase,

calculated from the inverse pole figure, considering the lattice parameters to be c = 4.130(1) Å

and a = 3.912(1) Å . The crystalline distortion is c/a = 1.055(7), smaller than the theoretical one

(1.0668) which correspond to a theoretical c = 4.150(0) Å and a = b = 3.890(0) Å. Previous

studies reported the effects of stresses in ferroelectric thin films [17, 18]. They point to three

different reasons for the strains: the different mismatch between substrate and film, the

different thermal expansion of both and the phase transformation of the film with

temperature. The last reason is minimized when cooling down from temperatures above 500

⁰C, which is done in this case. Only lattice and thermal expansion mismatches can result into

strains. There is also another effect that must be considered: the film is subjected to stresses

induced by the substrate [17]. However, this contribution decreases with the thickness of the

film and should be minimized in the case here considered.

All considered, the stresses responsible of the smaller tetragonal distortion found for this thin

film might be mainly due to the different lattice mismatch and thermal expansion of the

PbTiO3 film and the polycrystalline Pt surface.

All the sharp rings not explained yet can be identified by the (100) fiber structures of the

PbTiO3 perovskite phase. Orientation distribution cone is ±15°. Fig. 3.11 indicates to which

family of planes corresponds each ring presented at the diffractogram.

3.2.3. Functional characterization

PFM studies were carried out for the continuous ultrathin film and the isolated nanostructures

prepared from the 10-2 M and 4·10-2 M diluted sol, respectively. Images corresponding to the

amplitude and phase of the piezoresponse of the particles are shown in Fig 3.12 (ultrathin film)

and Fig 3.13 (nanostructures).

The measurements of Fig. 3.12 were performed at the Materials Science Institute of Madrid

under an AC field of amplitude 2V -peak to peak- and a frequency of 50 kHz. The image shows

that almost all grains have a polarization that is not purely in-plane or out-of-plane but rather

have both contributions. There are two areas that present no ferro-piezoresponse: one at the

bottom right part of the image and another one at the centre.

Chaper 3: Nanoparticles fabrication by the phenomenom of the microstructural instability of polycrystalline ultrathin films

67

Figure 3.12. Topography (a) and (d), phase (b) and amplitude (c) images of the in-plane piezoresponse

and phase (e) and amplitude (f) of the out-of-plane piezoresponse for the continuous ultrathin PbTiO3

film.

In the case of the piezoresponse of the nanostructures of Fig. 3.13, the measurements were

carried out at the Max Planck Institute for Microstructure Physics under an AC field of

amplitude 1.5V -peak to peak- and a frequency close to the contact resonance of the cantilever

Figure 3.13. Topography (a), phase (b) and amplitude (c) images of the out-of-plane piezoresponse for

the PbTiO3 nanostructures.

(245kHz), in order to enhance the measurements using the ATEC-EFM cantilevers, as explained

in Chapter 2 (Experimental Procedure, Section 2.6.3).

The image shows different contrast and amplitude at the top facet of the nanostructures than

at the lateral. Most probably, the coating or the cantilever tip is damage and, consequently,

25.46 nm

0nm

PZ

125 pm/V

0.00

PFM Image Amplitude image

PFM Image Phase imageAFM Image

PFM Image Amplitude image

PFM Image Phase imageAFM Image

25.46 nm

0nm

PZ

81pm/V

0.00

25.46 nm

0nm

PZ

0.00

a) b) c)

d) e) f)

PFM Image Amplitude image

PFM Image Phase imageAFM Image

100nm100nm100nm

PZ

18.21 nm

0nm

30 pm/V

0.00

a) b) c)

68 3.2. Nanoscale PbTiO3 structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by using the

phenomenon of the microstructural instability

the response is enhanced when the part of the tip that is not damaged still is in contact with

the nanostructure.

Measurement conditions: Vac = 1.5V

Freq ac = 245kHz F = 25 nN

Tbias = 0.2s Tmeasurement = 0.3s

SEN = 1mV TC = 20 ms

Figure 3.14. Out-of-field consecutive local hysteresis loops of a nanoparticle of 50nm.

Furthermore, the out-of-field consecutive local hysteresis loops of the isolated nanostructures,

shown in Fig. 3.14, verify the ferro-piezoelectric character of these PbTiO3 nanostructures.

Hysteresis loops are asymmetric with respect to the piezoresponse axis. In ferroelectric

nanoparticles, Rodriguez et al. [19] have already found this effect. It is due to a pinned layer at

the bottom part of the nanostructure that clamps the switching of the polarization. The nature

of the pinned layer is unknown, but most probably related with certain crystal structure

defects occurring at the interface between the substrate and the ferroelectric particle (Fig.

3.15). When the field is applied in the same direction as the pinned polarization, its

contribution adds to that switched by the field, whereas when the field is applied in the

opposite polarity this contribution is subtracted from the switched one giving place to the

asymmetry of the loop. The hysteresis loops also present an asymmetry with respect to the

voltage axis which is related to a certain imprint of the nanostructures. This effect is well-

known as it takes place in most thin films [20].

Figure 3.15. Scheme of the pinned layer and the imprint (a) and its effect on a hysteresis loop (b)[21]

-10 -5 0 5 10

-40

-20

0

20

40

60

80

100

Pie

zo

-Re

sp

on

se

(a

.u)

Voltage (V)

1st cycle

2nd cycle

3rd cycle

4th cycle

5th cycle

Chaper 3: Nanoparticles fabrication by the phenomenom of the microstructural instability of polycrystalline ultrathin films

69

Remarks

1. Isolated structures of PbTiO3 onto Pt/TiO2/SiO2/(100)Si substrates prepared by using

the phenomenon of the microstructural instability of polycrystalline ultrathin films

change the shape and size when the sol concentration, and thus the film thickness, is

decreased. Also, nanostructures in the range of the 50 nm are obtained from 4·10-2 M

and 3·10-2 M sols.

2. The equivalent diameter distributions of the nanostructures are log-normal. This

means that the mechanisms of growth involve coalescence and diffusion among

neighbor particles.

3. The amount of structures decreases with the concentration. For sol concentrations

below 3·10-2 M, structure size increases as the sol concentration decreases, up to some

300 nm of lateral dimensions. At 2·10-2 M needle structures are formed. For lower sol

concentration (10-2 M and below), all the structures nucleate and grow around the

defects of the substrate surface.

4. No self-arrangement of the nanostructures is obtained by this procedure onto

Pt/TiO2/SiO2/(100)Si substrates.

5. 2-D synchrotron radiation grazing-incidence diffractograms of a continuous ultrathin

film present reflections corresponding to those of the PbTiO3 perovskite crystal

structure. The orientation of the crystallites is a fiber one with the (100) axis

perpendicular to the surface of the substrate and a direction distribution cone of ±15°.

Cell parameters of the PbTiO3 perovskite deduced from this pattern are c = 4.130(1) Å,

a = 3.912(1) Å.

6. The isolated PbTiO3 nanostructures here grown in the range of the 50 nm show a

ferro-piezoelectric local behavior.

70 Bibliography

Bibliography

[1] G.L. Brennecka, C.M. Parish, B.A. Tuttle, L.N. Brewer and M.A. Rodriguez, "Reversibility

of the perovskite-to-fluorite phase transformation in lead-based thin and ultrathin films",

Advanced materials, 20 (8), 2008, p:1407

[2] G.L. Brennecka, C.M. Parish, B.A. Tuttle, L.N. Brewer and M.A. Rodriguez, "Multilayer

thin and ultrathin film capacitors fabricated by chemical solution deposition", Journal of

Materials Research, 23, 2008, p:176

[3] G.L. Brennecka and B.A. Tuttle, "Fabrication of ultrathin film capacitors by chemical

solution deposition", Journal of Materials Research, 22 (10), 2007, p:2868

[4] A. Seifert, A. Vojta, J.S. Speck and F.F. Lange, "Microstructural instability in single-

crystal thin films", Journal of Material Research, 11 (6), 1996, p:1470

[5] J.H. Kim and F.F. Lange, "Seeded epitaxial growth of PbTiO3 thin films on (001) LaAlO3

using the chemical solution deposition method", Journal of Material Research, 14 (4), 1999,

p:1626

[6] K.T. Miller, F.F. Lange and D.B. Marshall, "The instability of polycrystalline thin films:

Experiment and theory", Journal of Material Research, 5 (1), 1990, p:151

[7] M. Dawber, I. Szafraniak, M. Alexe and J.F. Scott, "Self-patterning of arrays of

ferroelectric capacitors: description by theory of substrate mediated strain interactions",

Journal of Physics: Condensed Matter, 15, 2006, p:L667

[8] C. Harnagea, A. Pignolet, M. Alexe and D. Hesse, "Piezoresponse Scanning Force

Microscopy: What Quantitative Information Can We Really Get Out of Piezoresponse

Measurements on Ferroelectric Thin Films", Integrated Ferroelectrics, 44, 2002, p:113

[9] M. Alexe and D. Hesse, "Self-assembled nanoscale ferroelectrics", Journal of Material

Research, 41, 2006, p:1

[10] J. Ricote, S. Holgado, P. Ramos and M.L. Calzada, "Piezoelectric ultrathin lead titanate

films prepared by deposition of aquo-diol solutions", IEEE Transactions on Ultrasonics

Ferroelectrics and Frequency Control, 53 (12), 2006, p:2299

[11] J. Ricote, S. Holgado, Z. Huang, P. Ramos, R. Fernandez and M.L. Calzada, "Fabrication

of continuous ultrathin ferroelectric films by chemical solution deposition methods", Journal of

Materials Research, 23 (10), 2008, p:2787

[12] I. Szafraniak, C. Harnagea, R. Schloz, S. Bhattacharyya, D. Hesse and M. Alexe,

"Ferroelectric epitaxial nanocrystals obtained by a self-patterning method", Applied Physics

Letters, 83 (11), 2003, p:2211

Bibliography 71

[13] R.W. Schwartz, T. Schneller and R. Waser, "Chemical solution deposition of electronic

oxide films", Comptes Rendus Chimie, 7, 2004, p:433

[14] J.S. Reed, Principles of Ceramic Processing. 2nd ed. 1995, New York: Wiley-Interscience

Publication.

[15] C.S. Pande and S.P. Marsh, "The analytical modeling of normal grain-growth", Jom-

Journal of the Minerals Metals & Materials Society, 44 (9), 1992, p:25-29

[16] P.R. Comès, M. Lambert and A. Guinier, "Désordre Linéaire dans le Cristaux (cas du

Silicium, du Quartz, et des Pérovskites Ferroelectriques).", Acta Crystallographica A26, 1970,

p:244-254

[17] J. Mendiola, M.L. Calzada, P. Ramos, M.J. Martin and F. Agulló-Rueda, "On the effects

of stresses in ferroelectric (Pb, Ca)TiO3 thin films", Thin Solid Films, 315, 1998, p:195-201

[18] G.A.C.M. Spierings, G.J.M. Dormans, W.G.J. Moors, M.J.E. Ulenaers and P.K. Larsen,

"Stresses in Pt/Pb(Zr, Ti)O3/Pt thin-films stacks for integrated ferroelectric capacitors", Journal

of Applied Physics, 78 (3), 1995, p:1926-1933

[19] B.J. Rodriguez, S. Jesse, M. Alexe and S.V. Kalinin, "Spatially Resolved Mapping of

Polarization Switching Behavior in Nanoscale Ferroelectrics", Advanced materials, 20 (15),

2008, p:102

[20] H.N. AlShareef, D. Dimos, W.L. Warren and B.A. Tuttle, "Voltage offsets and imprint

mechanism in SrBi2Ta2O9 thin films", Journal of Applied Physics, 80 (8), 1996, p:4573-4577

[21] C.M.P. G.L. Brennecka, B.A. Tuttle, L.N. Brewer, M.A. Rodriguez, "Multilayer thin and

ultrathin film capacitors fabricated by chemical solution deposition", Journal of Material

Research, 23, 2008, p:176

CHAPTER 4: FERROELECTRIC NANOSTRUCTURES BY

MICROEMULSION MEDIATED SYNTHESIS ONTO

Pt/TiO2/SiO2/(100)Si SUBSTRATES

4.1. The microemulsion mediated synthesis

The term surfactant comes from surface acting agent. A surfactant is an amphiphilic molecule,

that is to say, it has a hydrophilic and a hydrophobic part, which are called “head” and “tail”,

respectively. Because they have these two distinguishable parts, they usually are found at the

interfaces and can be used in a wide range of applications: microemulsions, detergents,

membranes, liquid crystals or liposomes [1].

Surfactants can be divided into anionic, cationic and non-ionic surfactants, depending on the

structure of their head. In this work, a non-ionic molecule will be used: Polyoxyethylene (4)

lauryl ether which chemical formula is CH3-(CH2)10-CH2-(O-CH2-CH2)4-OH and commercial name

is Brij-30. A representation of the molecule is shown in Fig. 4.1.

Figure 4.1. Representation of Brij-30 with the head and tail groups indicated.

If surfactant molecules are added to an emulsion (mixture of two immiscible liquids), they try

to organize in order to minimize the chemical potential, stabilizing the groups of one of the

phases in the other and preventing them from reverting into two different layers.

Emulsions are usually described as water-in-oil and oil-in-water type or direct and inverse

emulsions, respectively, depending on which is the continuous medium. Also, depending on

the size of the non-continuous phase drops, they can be divided in microemulsions,

miniemulsions or emulsions [2, 3].

The shape and size of the colloidal aggregates depend on the type of surfactant and the nature

and relative quantity of the two phases present in the emulsion. Fig. 4.2 shows some of the

possible aggregates.

74 4.1. The microemulsion mediated synthesis

Figure 4.2. Some of the possible colloidal aggregates of surfactant molecules in an emulsion. One of the

phases of the emulsion is presented in light blue and the other one in white. Surfactant head is

represented by a circle and the tail by a line.

One of these possible assemblies are micelles. Micelles are a grouping of surfactant molecules

where either the hydrophobic (in a polar continuous phase) or the hydrophilic (in a non-polar

continuous phase) ends cluster inward to escape the continuous phase, keeping one of the

liquids of the emulsion inside. If the heads form the inside of the aggregate, then it is called an

inverse micelle and in the opposite case, it is called a direct micelle (see Fig. 4.2). Micelles are

formed by the core, a liquid pool inside, and the shell, formed by the surfactant.

In this work, microemulsions will be prepared from a mixture of cyclohexane (as oil), water and

Brij-30. The molar ratio of water/surfactant is kept constant ~1.2, which gives place to inverse

micelles with a spherical shape [4]. Fig. 4.3 schematizes the configuration of the microemulsion.

Surfactant molecule Inverse micelle Direct micelle

Planar lamellar phase Onion like lamelar phase

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

75

Figure 4.3. Schema of the inverse micelles inside the microemulsion prepared in this work. Core is

identified as medium blue and shell as dark blue. A detailed image shows the configuration of the

surfactant molecules.

Microemulsion mediated synthesis includes all the approaches that involve the formation of

micelles from a surfactant in a mixture of water and oil. This group of procedures had been

inspired by Nature. The assembly process in biological organization is controlled under

thermodynamic and hydrophobic effects. The hydrophobic effect is a unique organizing force,

based on repulsion by the solvent instead of attractive forces at the site of organization and is

responsible for assembly of membranes of cells [5, 6].

When a microemulsion is added to a sol, sol particles will introduce themselves inside the core

of the micelle, keeping the water content unaltered. This resulting solution will be denoted

hereinafter as micellar solution. In this work, it is hypothesized that these micelles that contain

water and sol particles will be the building units from which the crystalline nanostructures will

be formed.

Fig. 4.4 schematizes the formation of the micellar solution, considering the solvents, sol and

surfactant that will be used in this work. Note that most of the micelles will contain sol

particles inside, but they will be some with only water in their core.

76 4.1. The microemulsion mediated synthesis

Figure 4.4. Schematic formation process of the micellar solution. The dark blue annulus represents the

shell of the micelle and of the building unit.

Further self-organization of micelles into periodic hexagonal, cubic or lamellar phases places

both the organic and inorganic compounds into 3-D arrangements, as a result of hydrophobic

and thermodynamic effects. That makes of the drying step a very important one. Theoretically,

if there is no interaction between the solvent (1,3-propanediol) and the micelles, the liquid

from the pores in between the micelles would evaporate leaving the network exposed.

However, adsorption and capillary forces tend to produce the collapse of the micelles network,

if a controlled drying is not carried out. This also causes solvent to flow from the interior of the

coating in order to replace that liquid that is evaporating. Slow drying process, consequently,

helps to the self-assembly of the micelles (whether with only water at the core, or both water

and sol particles). This is the premise of the evaporation-induced self-assembly process (EISA)

[7, 8], where the drying process is controlled to produce the desired order of the building units

in the coating. In this thesis, this type of drying will be understood as an intermediate step in

the CSD preparation of the self-assembled ferroelectric nanostructures onto substrates. Thus,

1,3-propanediol solvent

H2O

C6H12

O

OO

OOH

O

O

O

O

OH

O

O

O

O

OH

O

O

O

O

OH

O

O

O

O

OH

O

O

O

O

OH

O

O

O

O

OH

O

O

O

O

OH

H2O

H2O

H2O

H2O

1,3-propanediol

H20 + sol

H2O

C6H12 drops

Sol nanoparticles

Microemulsion Sol

Micellar solution

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

77

drying times will be long at a low temperature and with a controlled humidity in order to be

under conditions as close as possible to those used in the evaporation-induced self-assembly

process.

Microemulsion mediated synthesis has been extensively used in many different fields [9-17],

and just scarcely in the last two decades for the preparation of perovskite structured

nanocrystalline complex oxides [18-24].

Ferroelectric nanopowders have been prepared by microemulsion mediated synthesis, with

good control of particle size, fully dense and of high purity. In addition, they are processed at

reduced temperatures [18, 22, 24-26], compared with conventional methods of processing of

oxide powders. They have been used as precursors for the conformation and sintering of bulk

ferroelectric ceramics [27]. Ferroelectric thin films onto substrates have also been prepared by

the microemulsion mediated synthesis either by a micellar solution [28] or by adding a certain

amount of surfactant to a precursor sol [29, 30]. Films with a high control of the preferred

orientation and improved ferroelectric properties are obtained.

In particular, their use in the preparation of isolated self-assembled ferroelectric

nanostructures onto substrates has been rarely attempted. One important advantage of this

procedure is that controlled size and shape of the micelles can be achieved, and consequently,

resulting isolated structures would be also obtained with a controlled shape and size.

The work of Bhattacharyya et al. [19] was one of the first attempts to prepare BaTiO3 isolated

nanostructures onto silicon substrates. In order to obtain them, crystalline nanopowders were

prepared from a microemulsion using the procedure established by Herrig and Hempelmann

[22]. These nanopowders were dispersed in an alcohol solvent and the resulting dispersion

deposited afterwards onto a conductive substrate by chemical solution deposition, obtaining

nanostructures with an average size of 50-60 nm. Order of the nanostructures onto the

substrate was not observed and ferroelectric response was not reported.

Grosso et al. [21] chose the di-block copolymer approach, for the preparation of SrTiO3,

MgTa2O6 and CoxTi1-xO2-x nanostructures onto single-crystal Si wafer combining it with dip-

coating process. By this procedure, a low-range hexagonal order of the structures onto the

substrate is obtained. The nanostructures are ~25 nm rounded ones. In this work, the

dielectric characteristics of the SrTiO3 nanoislands were not showed.

78 4.1. The microemulsion mediated synthesis

Kronholz et al. [23] fabricated PbTiO3 nanograins of ~30 nm of lateral size onto predefined TiO2

nanostructures created on Pt/TiO2/SiO2/(100)Si substrates by using a self-organized template

with the aid of PS-b-PVC di-block copolymer micelles. This procedure combines top-down and

bottom-up approaches. This mixed mechanism results into an incipient order of the PbTiO3

nanostructures. Proof of the ferroelectric behaviour of the isolated structures is not reported.

For the development of this work, it is hypothesized that pre-organized templates of inverse

micelles combined with sol-gel synthesis will lead to self-assembled isolated nanostructures

onto substrates with controlled shape and size. In this approach, microemulsion mediated

synthesis will be combined with chemical solution deposition for the preparation of

ferroelectric PbTiO3 isolated nanostructures onto Pt/TiO2/SiO2/(100)Si substrates.

Figure 4.5. Hypothetical self-assembly of the nanostructures: a self-assembly of the micelles (whether

building units or micelles with only water inside) will lead to self-assembly nanostructures after drying

and crystallization. The building units and micelles of the deposited micellar solution are displayed as

violet circles, the micelles with only water inside in blue and the resulting nanoparticles, after thermal

treatment, are represented in orange. A defect in the Pt coating is indicated by a non brown depicted

area.

Pt polycrystalline

TiO2

SiO2

Si wafer

H2OH2O

+ Sol

H2O +

Sol

H2O +

Sol H2O +

SolH2O

H2O +

SolH2O

H2OH2O

+ Sol

H2O

+ Sol

H2O

H2OH2O

+ Sol

H2O

+ Sol

H2O

+ Sol H2O

+ Sol

H2O +

Sol

H2O +

Sol

H2O +

Sol

Drying

Pt coating defect

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

79

In addition, micelles may act as nanoreactors, providing an isolated place where the chemical

reactions and synthesis of the compounds might take place [4, 18, 22, 26, 31], helped by the

high pressures inside the core. Thus, the combination of self-assembly ability and the fact that

they work as nanoreactors would turn micelles into the building units for the ferroelectric

nanostructures that will be fabricated in this Chapter.

Thus, the hypothesis is that coating the substrate with the micellar solution will yield self-

assembled nanostructures onto the Pt/TiO2/SiO2/(100)Si substrate. Fig. 4.5 shows a scheme of

the self-assembly process.

Note that not all the micelles will contain sol inside them and, thus, only the previously

denoted as building units will yield oxide nanostructures. This is represented in Fig. 4.5.

As colloids, micellar solutions, microemulsions and sols show the Tyndall effect [32]. This

consists on the light scattering by the colloidal particles or particles in a suspension. When

using a laser beam, a straight line can be seen crossing the colloidal solution (Fig. 4.6). This

qualitative experiment is a proof of the colloidal character of the microemulsion, sol and

micellar solution, prepared in this work.

In order to investigate the size of micelles and sol particles and to unveil the nature of the

particles in the micellar solution, DLS measurements were carried out (Fig. 4.7).

Figure 4.6. Tyndall effect in the microemulsion (a), sol (b) and micellar solution (c)

Sol

b)

Micellar solution

c)

Microemulsion

a)

80 4.1. The microemulsion mediated synthesis

Figure 4.7. DLS measurements for the microemulsion (a), sol (b) and micellar solution (c)

0 200 400 600 800 1000

0,0

0,2

0,4

0,6

0,8

1,0

Norm

alis

ed

inte

nsi

ty

Hydrodynamic radius (nm)

Microemulsiona)

0 10 20 30 40

0,0

0,2

0,4

0,6

0,8

1,0

No

rma

lise

d in

ten

sity

Hydrodynamic radius (nm)

Solb)

0 200 400 600 800 1000

0,0

0,2

0,4

0,6

0,8

1,0

No

rma

lised

inte

nsi

ty

Hydrodynamic radius (nm)

Micellar solutionc)

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

81

From these measurements, it was found that micelles formed by Brij-30 have an average

radius of ~275 nm in the microemulsion, while the sol nanoparticles are ~12 nm. The particles

in the micellar solutions have an average radius of ~290 nm. The fact that no particle of ~12nm

is observed in the DLS measurements carried out in the micellar solution indicates that all sol

particles are incorporated into the core of the micelles, protected by the surfactant, in this

micellar solution, as it is needed to give place to isolated nanostructures. Also, we can observe

a slight increment in the radius of the micelles when the sol is set inside, as we would expect. It

is a well-known fact that a change in the viscosity of the core of a micelle with respect to the

solvent one will change its radius and adding sol nanoparticles to the water pool will change

the viscosity of the core [33].

4.2. Nanoscale PbTiO3 ferroelectric structures onto

Pt/TiO2/SiO2/(100)Si substrates prepared by microemulsion mediated

synthesis.

Micellar solution with building units of ~290 nm of average size will be used as precursor

solutions for the preparation of the lead titanate nanostructures onto a conductive coated Si-

wafer. This method is compatible with the silicon technology, and will make these

nanostructures suitable for their use as NV-FeRAM.

4.2.1. Microscopy and quantitative microstructure analysis.

Coatings of the substrate obtained by the deposition of the micellar solutions are not uniform.

The deficient wetting of the substrate by the micellar layer gives place to uncoated areas of

the substrate and coated ones with a thickness gradient. To the naked eye, it can be

distinguished in the coated substrate, after drying and crystallization, a large brown area with

an irregular shape located at the centre of the substrate with a darker boundary, while the

outer part of the substrate remains uncovered (see Fig. 4.8 (a)). Fig. 4.8 (b) shows a

micrograph of the limit of the brown area, where the darker colour is found. Lead titanate

nanostructures correspond to the brighter spots, while the darker scratches on the substrate

reveal the TiO2 underlying stratum of the substrate.

Fig. 4.9 displays the AFM images of samples prepared from the 10-2 M and 5·10-3 M micellar

solutions, taken at the central brown area and at the outer part, close to the edge of this

brown area shown in Fig. 4.8 (a). Fig. 4.9 (a) shows a continuous film that leads to isolated

structures when the coating is thin enough (outer part, Fig. 4.9 (b)). This film can be compared

82 4.2. Nanoscale PbTiO3 ferroelectric structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

with that shown in Fig. 3.3 (a) of previous Chapter 3, where a continuous ultrathin film was

prepared directly from the sol alone (without adding the microemulsion).

Figure 4.8. Topography by optical microscopy of the limit of the coating area (a) and SEM micrograph

of the center of the non-homogenous coating (b) (sample prepared from a 5·10-3

M micellar solution).

In the latter image, brighter gray corresponds to PbTiO3, black to TiO2 and the medium gray to Pt.

This is consistent with the DLS experimental results of the sol nano particles positioning in the

core of the micelles in the micellar solution and the hypothesis of the micelles acting as

building units, thus giving place to a layer of deposited micelles that after drying, elimination of

a)

50 µm

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

83

Fi

g. 4

. 9

. A

FM t

op

ogr

aph

y im

age

s o

f th

e P

bTi

O3 n

ano

stru

ctu

res

on

to P

t/Ti

O2/S

iO2/

(10

0)S

i su

bst

rate

s p

rep

ared

fro

m t

he

mic

ella

r so

luti

on

s w

ith

a s

ol

con

cen

trat

ion

of

10

-2 M

(a-

b)

and

5·1

0-3

M (

c) a

t d

iffe

ren

t lo

cati

on

s o

f th

e co

ated

su

bst

rate

.

Co

nc

0.0

1M

y 0

.005

M

Dis

tan

ceto

the

cen

ter

of

the

sub

stra

ted

= 0

64

.08

nm

0n

m

60

.00

nm

0n

m

55

.62

nm

0n

m

a)b

)

c)

84 4.2. Nanoscale PbTiO3 ferroelectric structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

organics, and crystallization give place to isolated smaller structures than those obtained when

the deposited sol-gel coating breaks due to the phenomenon of the microstructural instability.

If the microemulsion is not used, the deposited sol would crystallize in flat structures of

irregular shape, and lateral dimensions bigger than 200 nm (as explained in Chapter 3, section

3.2.1, Fig. 3.4 (g) and (i)). That is to say, instead of having free precursor particles that will set

around the defects, building units will be distributed uniformly onto the substrate, carrying the

precursor material in their inside from which the oxide nanostructures will be formed.

Figure 4.10. High magnification AFM image of isolated nanostructures fabricated from the 5·10-3

M

micellar solution. The profile of each nanostructure that corresponds to the light violet line is

represented underneath them.

Fig. 4.9 (b) and (c) show nanostructures of similar shape and size. Only their amount on the

surface differs, decreasing with the concentration of the micellar solution. The number of

micelles in the coating layer depends on the amount of water and surfactant in the

microemulsion and, thus in the micellar solution. However, the number of building units with

precursor particles inside depends on the concentration of the sol. Consequently, decreasing

the concentration of the sol leads to a reduced number of crystalline nanostructures onto the

substrate.

The observed nanostructures seem to be formed by coalescence of smaller primary ones. This

can be deduced from the high magnification topography image of Fig. 4.10, where some of the

nanostructures show an imperfect spherical shape that would correspond to the merged of

primary ones. These primary nanostructures would be those resulting from crystallization from

the building units. The characteristic dimensions of the entrapped sol dropplets would be

determined by the physico-chemistry of the microemulsion and, in turn, will determine the

21

150100500

20

15

10

5

0

-5

X[nm]

Z[n

m]

250200150100500

25

20

15

10

5

0

-5

X[nm]

Z[n

m]

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

85

Figure 4.11. Proposed drying evolution with time of the micellar layer (a, b) and oxide nanostructures

formed after the thermal treatment of crystallization (c).

H2O

Pt polycrystalline

TiO2

SiO2

Si wafer

H2O +

Sol

H2O + Sol 1,3-propanediol

H2O +

Sol

H2OH2O +

Sol

H2O + Sol

H2O +

Sol

Drying

Thermaltreatment

Capillary forces

Pt coating defects

a)

b)

c)

86 4.2. Nanoscale PbTiO3 ferroelectric structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

size of the primary nanostructures. Contrary to this, the size of the nanostructures here

observed will depend on the homogeneity of the coating on the substrate and on the

arrangement of the micelles, which ultimately will determine how close the building units and,

subsequently, the merging primary nanostructures are. We can understand this as the result of

inhomogeneity that the surface of the polycrystalline Pt substrate presents and the thickness

of the coating which is not a monolayer of micelles. The deposited micelles tend to

concentrate in certain points during the long process of drying and shrinkage of the coating,

when the solvent is eliminated, leading, during the thermal treatment, to a coalescence of the

primary nanostructures. These nanostructures are not periodically arranged onto the substrate

(Fig. 4.9). In addition to the substrate surface inhomogeneity, there is a second effect that

must be taken into consideration: the deficient wetting of the substrate by the micellar

solution that is partially responsible of the large shrinkage of the coating during the drying step

[34], leading to the collapse of the micelles at the centre of the substrate and yielding the

brown area described before (see Fig. 4.8 (a)).

Fig. 4.11 describes the proposed drying and crystallization processes. As explained in the

introductory part of this Chapter (section 4.1), during drying, adsorption and capillary forces

oppose to the exposure of the micelle network [34]. This will make the micelles and building

units to become closer (Fig. 4.11 (b)) and will finally make a number of primary nanostructures

to be close enough (Fig. 4.11) to coalescence during the thermal treatment and, thus, yielding

a single nanostructure after that (Fig. 4.11 (c)).

Figure 4.12 Equivalent diameter distributions of nanostructures prepared from the 10-2

M (a) and 5·10-

3 M (b) micellar solution.

0.01

0.5

2

10

30

50

70

90

98

99.5

0 20 40 60 80 100 120 140

0 20 40 60 80 100 120 140

0

10

20

30

40

50

Co

un

ts

Deq

(nm)

Cu

mu

lative

fre

qu

en

cy

0,01

0,5

2

10

30

50

70

90

98

99,5

0 20 40 60 80 100 120 140

0 20 40 60 80 100 120 140

0

10

20

30

40

50

Co

un

ts

Deq

(nm)

Cu

mu

lative

fre

qu

en

cy

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

87

Fig. 4.12 presents the size distributions of the nanostructures prepared from the 10-2 M and

5·10-3 M micellar solutions, calculated from the AFM images (Fig. 4.9 (b) and (c)).

Both distributions are Gaussian and the fitting of their probabilistic lines is analitically

expressed by:

y = -3.48 + 0.05·x R = 0.99 (3.1)

for the nanostructures deposited from the 10-2 M sol and

y = -3.70 + 0.06·x R = 0.99 (3.2)

for those deposited from the 5·10-3 M sol.

Following this procedure, the average equivalent diameter of the nanostructures derived from

the 10-2 M and 5·10-3 M sols are 75 nm and 67 nm with a standard deviation of 21 nm and 18

nm, respectively.

The fact that both distributions are Gaussian means that the formation of the nanostructures

does not take place via a normal grain growth, which would give place to lognormal

distributions as those measured for nanostructures obtained using the phenomenon of the

microstructural instability of ultrathin films (see Chapter 3, Fig. 3.6). Rather, the

nanostructures grow from independent nucleation points, either isolated building units or

groups of these. As it has been explained for other systems, micelles may act as nanoreactors

[35-38] and so, hydrolysis and crystallization may occur in the core, at least partly, even before

the thermal treatment takes place.

Primary oxide nanostructures could be formed inside the micelles prior to the so-called

thermal treatment. During such thermal treatment, merging between the primary

nanostructures occurs, as revealed by the AFM images of Fig. 4.10, via the usual thermally

stimulated mass transport (diffusion) at the nanostructure surfaces.

4.2.2. Structural characterization

Experiments of synchrotron X-ray diffraction using a 2-dimensional detector were carried out.

Fig. 4.13 shows the resulting pattern of the sample prepared from the 5·10-3 M micellar

solution. It consists of three spots forming a triangle, some broad spots (one of them circular,

the others oval type) and some sharp rings or circular sectors of lower intensity.

88 4.2. Nanoscale PbTiO3 ferroelectric structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

Figu

re 4

.13

. Exp

erim

enta

l 2-D

syn

chro

tro

n x

-ray

dif

frac

tio

n p

atte

rn o

f a

sam

ple

pre

par

ed f

rom

a 5

·10

-3 M

mic

ella

r so

luti

on

.

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

89

Figu

re 4

.14

. 2

θ d

iffr

acti

on

pat

tern

cal

cula

ted

fro

m t

he

inte

grat

ion

of

the

2-D

exp

erim

enta

l pat

tern

of

Fig.

4.1

3 (

bla

ck s

olid

lin

e) a

nd

sim

ula

ted

dif

frac

tio

n p

atte

rns

of

the

Pb

TiO

3 p

ero

vski

te n

ano

stru

ctu

res

(red

lin

e), P

t b

ott

om

ele

ctro

de

(blu

e lin

e), T

iO2

anat

ase

(gr

een

lin

e)

and

Al h

old

er (

ora

nge

lin

e).

15

20

25

30

35

40

45

50

354

177

0

PbT

iO3 32.3

%

001

100

101

110

111

002

200

102

201

112

211

202

003

212

221103

301310

311

222

Al 32.3%

111

200

220

311

222Pt 3.2%

111

200

220

311RUTILE 32.3%

110

101

200

111

210

211

220

002

310

221

301

112

311

320

202

212

400

5PHI000-PHI020-STPW002-5mins_01_03.X_Y

Intensity (a.u.)

90 4. 2. Nanoscale PbTiO3 ferroelectric structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by microemulsion mediated synthesis

Fig. 4.14 presents, as a solid black line, the integration of the above described 2-D diffraction

pattern. Each peak corresponds to one of the sharp rings, ring portions or circular or oval peak

of the said pattern. Each simulated diffraction pattern of the possible materials on the sample

is represented in a different color, as explained in the caption at the top left corner.

There is a certain contribution of the aluminium holder to the pattern. This can easily be

observed for the peak at 22°. Note that the holder is set at a different height than the sample

and, therefore, the peak is shift from its theoretical position to a lower angle. Also, the

incidence angle is very low (0.1°) meaning that the path of the beam on the sample is very long

(~ 8.6 cm) compared with the sample (~ 1 cm) and, thus, the peaks can be shifted with respect

to their theoretical position. The distance considered for the above calculations and the ones

that will be explained later in this section had been done considering an average value that fit

the peaks of the Pt present on the substrate.

Fig. 3.9 of previous chapter shows the simulated peaks for a (111) texture of the Pt of the

substrate, with an orientation distribution cone of ±3°, as it was previously observed for the

ultrathin film in previous section 3.2.2. This simulated pattern was calculated from the inverse

pole figure and fits explain the broad oval peaks in the 3-dimensional pattern of Fig. 4.13.

The silicon wafer is responsible of the broad circular intense peak located at the second

quadrant and the triangle at the center of the 2-D pattern in Fig. 4.13. Fig. 3.10 in previous

chapter shows the simulated diffraction pattern for (100) diffraction pattern that explains the

above peaks.

The peak at 20° in Fig. 4.14 can only be explained if there is PbTiO3 perovskite in the sample.

This assures that the PbTiO3 nanostructures studied in previous sections are perovskite, but

their relative intensities, the fact that other peaks have contributions from other materials on

the sample and the long shadow of the beam on the sample and the subsequent shift of the

peaks, makes impossible to determine a possible texture of these nanostructures.

However, the lattice parameters had been fitted to a = b = 3.890(0) Å and c = 4.120(0) Å, and

the tetragonal distorsion 1.059(1), being slightly different from the theoretical one (theoretical

values: a = b = 3.899(9) Å, c = 4.150(0) Å, and tetragonal distortion 1.066(8)). As the

geometrical facts of this experiment that were explained in the paragraphs above have clearly

an influence on these measurements, it is hard to explain the difference in the parameters and

any consideration on stresses and strains would be merely speculative.

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

91

4.2.3. Functional characterization

Samples prepared from the 10-2 M and 5·10-3 M micellar solutions exhibit local ferro-

piezoelectric local behaviour as displayed in the PFM images of Figs. 4.15-17 and the hysteresis

loops shown in Fig. 4.18. The measurements were performed for an AC frequency close to the

free resonance frequency of the cantilever in order to enhance the signal obtained (see section

2.6 for further explanation).

Fig. 4.15 shows the in-plane and out-of plane piezoresponse images, together with the

corresponding AFM topography images for the same area of the sample prepared from the 10-

2 M micellar solution. It displays areas of agglomerated nanostructures that show a clear ferro-

piezoresponse. The polarization vector has both the in-plane and the out-of-plane components,

although the perpendicular to the surface is stronger (scales for the amplitude are the same

for both types of signals). There are also two single isolated nanostructures (marked with a

black point (·)) that exhibit a ferro-piezoelectric behaviour also with both the in-plane and out-

of plane components. A detailed observation of Fig. 4.15 (f) reveals lighter areas that

correspond to the non-agglomerated nanostructures of Fig. 4.15 (d). Although they are weaker

than the one for the agglomerates, it means that there is a mechanical displacement of the

cantilever and, thus, a ferro-piezoelectric response also in isolated nanostructures.

Fig. 4.16 displays the topography and corresponding out-of-plane response of isolated

nanostructures fabricated from the 5·10-3 M micellar solution. Only out-of-plane response was

found. All of the nanostructures distinguishable at the topography image (Fig. 4.16 (a)) exhibit

ferro-piezoelectric response. Pt is a conductive layer and so, the electrostatic contribution is

not zero but a constant value greater than the amplitude of the piezoresponse of the

nanostructures [39], explaining the inverse contrast observed at image (c) of Fig. 4.16.

Fig. 4.17 shows the topography, phase and amplitude image of an isolated nanostructure

fabricated from the 5·10-3 M micellar solution with domains inside. Domains can be found at

the largest nanostructures and their width is 20-30 nm, approximately.

Out-of-field piezoresponse hysteresis loops for an AC voltage at the free resonance of the

cantilever (Fig. 4.18 (a) and (b)) and at its contact resonance (Fig. 4.18 (c)) have been measured

at an isolated nanostructure of ~93 nm and ~83 nm of lateral size, respectively, from the

92 4. 2. Nanoscale PbTiO3 ferroelectric structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by microemulsion mediated synthesis

.

Figu

re 4

.15

. To

po

grap

hy

(a)

and

(d

), p

has

e (b

) an

d (

e) a

nd

am

plit

ud

e (c

) an

d (

f) im

age

s o

f th

e in

-pla

ne

and

ou

t-o

f-p

lan

e p

iezo

resp

on

se,

resp

ect

ivel

y, f

or

a sa

mp

le p

rep

ared

fro

m t

he

mic

ella

r so

luti

on

wit

h a

10

-2 M

co

nce

ntr

atio

n.

PZ

40

.81

nm

0n

m

31

pm

/V

0.0

0

PZ

40

.81

nm

0n

m

31

pm

/V

0.0

0

PFM

Im

age

Am

plit

ude im

age

PFM

Im

age

Phase im

age

AFM

Im

age

Out-

of-

pla

ne

In-p

lane

a)

b)

c)

d)

e)

f)

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

93

Fig

ure

4.1

6.

Top

ogr

aph

y (a

), p

has

e (b

) an

d a

mp

litu

de

(c)

ou

t-o

f-p

lan

e p

iezo

resp

on

se i

mag

es

for

the

iso

late

d n

ano

stru

ctu

res

of

a sa

mp

le p

rep

ared

fro

m t

he

mic

ella

r

solu

tio

n w

ith

a 5

·10

-3 M

co

nce

ntr

atio

n.

19

.46

nm

-6.2

1 n

m

PZ

1.4

5 p

m/V

0.0

0

20

0nm

a)

200nm

c)

20

0nm

b)

b)

Out-

of-

pla

ne

PFM

Im

age

Am

plit

ude im

age

PFM

Im

age

Phase im

age

AFM

Im

age

94 4. 2. Nanoscale PbTiO3 ferroelectric structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by microemulsion mediated synthesis

Figu

re 4

.17

. To

po

grap

hy

(a),

ph

ase

(b)

and

am

plit

ud

e (c

) im

ages

of

the

do

mai

n s

tru

ctu

re o

f a

big

nan

ost

ruct

ure

(~2

00

nm

of

late

ral s

ize)

.

19

.46

nm

-6.2

1 n

m

PZ

1.4

5 p

m/V

0.0

0

a)

c)

b)

Out-

of-

pla

ne

PFM

Im

age

Am

plit

ude im

age

PFM

Im

age

Phase im

age

AFM

Im

age

25

02

00

15

01

00

50

0

5 4 3 2 1 0

X[n

m]

Z[µm]e)

d)

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

95

sample prepared from the 5·10-3 M micellar solution. As expected, piezoresponse is enhanced

when measuring with an AC field close to the contact resonance of the cantilever [40].

Hysteresis loops are symmetric with respect to the voltage axis as well as to the piezoresponse

axis. This is in contrast to the results obtained previously in Chapter 3, showed in Fig. 3.15,

where an isolated nanostructure of 50 nm of lateral size presented an asymmetric hysteresis

loop with respect to both axes.

Measurement conditions:

Vac = 1.0 V, Freq = 65 kHz, F = 100nN, Tbias = 0.1s, Tmeasurement = 4.5s, SEN = 5mV,

TC = 10 ms

Measurement conditions:

Vac=1.5V

Freq= 245 kHz

F= 25nN

Tbias= 0.2s

Tmeasurement= 0.3s

SEN = 5mV

TC= 20 ms

Figure 4.18. Out-of-field local hysteresis loops of isolated nanostructures of ~95 nm of lateral size(phase

(a) and amplitude (b) loops) and ~83 nm of lateral size(c) (four consecutive piezoresponse loops)

isolated nanostructures fabricated from the 5·10-3

M micellar solution.

-10 -5 0 5 10-200

-150

-100

-50

0

50

100

150

200

Phase (

º)

Voltage (V)

-10 -5 0 5 10

-1,5

-1,0

-0,5

0,0

0,5

1,0

1,5

d3

3

eff(p

m/V

)

Voltage (V)

a) b)

-8 -6 -4 -2 0 2 4 6 8

-1000

-800

-600

-400

-200

0

200

400

600

800

1000

1200

Pie

zo

resp

on

se

(a

.u.)

Voltage (V)

c)

96 4. 2. Nanoscale PbTiO3 ferroelectric structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by microemulsion mediated synthesis

In Fig. 4.18 (c), the first hysteresis loop measured is different from the subsequent ones. It is

symmetric with respect to both the voltage and piezoresponse axis while the next ones are not,

becoming more asymmetric with respect to the piezoresponse axis. It can be attributed to an

imprint [41] that becomes more relevant as the nanostructure suffers subsequent loops.

On the other hand, the area over the voltage axis becomes smaller, while the one underneath

remains less altered. The nanostructures have an initial polarization that can be attributed to

the condition of the ferroelectric/metallic interface of the substrate. As the voltage field

increases, the polarization reverses, reaching the saturation voltage. Subsequently, the voltage

field decreases, until it becomes negative. In this step, a certain clamped volume is formed,

with the polarization set in the same direction as it was initially. When the field is in the

opposite direction and big enough to make the polarization switch, this clamped polarized

layer will stay unaltered, decreasing the piezoresponse in only one direction.

Figure 4.19 Proposed switching mechanism for the isolated nanostructures.

Fig. 4.19 shows a schematic drawing of the proposed mechanism of switching at the

nanostructures, being (a) the first hysteresis loop and (b) the second one. Hysteresis loop (a)

has initial polarization of unknown nature but that should be compensated by charges on the

surface. When an external electric field is applied, two effects take place. On one hand,

Schottky barrier formed at the metal-ferroelectric junction will decrease. Typically,

ferroelectric titanates present a region near the interface ferroelectric/metal more oxygen

deficient than the inside, creating a gradient of oxygen vacancies and causing the interface to

be slighter p-type than the inside. This makes the energy band to bend as shown in Fig. 4.20

and enables a charge injection towards the ferroelectric nanostructure [42].

-12 -10 -8 -6 -4 -2 0 2 4 6 8 10 12

-1400

-1200

-1000

-800

-600

-400

-200

0

200

400

600

800

1000

1200

1400

Pie

zo

resp

on

se

(a

.u.)

Voltage (V)

+ + + + + +

Substrate

P+ + + + + +

Substrate

+ + + + + +

Substrate

P

P

a)

+ + + + + +

Substrate

+ + + + + +

Substrate

P

P

1

2

3

4

51

2

3

4

5

-12 -10 -8 -6 -4 -2 0 2 4 6 8 10 12

-1400

-1200

-1000

-800

-600

-400

-200

0

200

400

600

800

1000

1200

1400

Pie

zo

resp

on

se

(a

.u.)

Voltage (V)

b)

+ + + + + +

Substrate

+ + + + + +

Substrate+ + + + + +

Substrate

+ + + + + +

Substrate

PP

P P

6

78

96

7

8

9

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

97

On the other hand, real ferroelectric materials present a certain number of charged defects,

such as impurities or vacancies that can be moved under an electric field [43]. When these

defects reach areas of chemical instability (e.g. dislocations, domain walls or grain boundaries),

they stay trapped in there.

When the field is applied at the beginning of the first loop (stage 1), it is not high enough to

reverse the initial polarization of the domain in the nanostructure. When the polarization

reverses (stage 2), the kinetics of polarization switching in these systems is governed by a

combination of nucleation and domain wall motion [39, 44]. Thus, the first time an electric

field with orientation opposed to that of the initial polarization is applied, domains will switch,

but when the state is reversed to the initial one (stage 4 in Fig. 4.19 (a)), surface charges will

stabilize the domain walls and create a pinned volume that will remain unaltered throughout

the subsequent hysteresis loops (Fig. 4.19 (b), stages 6-9). Switching destabilized the

equilibrium state resulting in the migration of the charges to the inside of the nanostructures.

These charges will act as screening charges and are localized inside the ferroelectric

nanostructures, near the pinned layer boundary [45]. The result is that the second and

subsequent hysteresis loops (coloured hysteresis loops in Fig. 18 (c)) are shifted with respect

to the piezoresponse axis [41].

Figure 4.20. Schematic representation of the bending of the energy bands for a conductor|p-

type semiconductor contact [43].

Phase transition due to the pressure of the tip over the sample was also considered as a

possible explanation but, for the phase transition to occur and be reversible, the applied

Metal Semiconductor

Conduction band

Valence band

Conduction band

Valence band

Fermi level

98 4. 2. Nanoscale PbTiO3 ferroelectric structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by microemulsion mediated synthesis

pressure should be within the range of 7.5-10 GPa [46], and the calculated one for the

measurements here presented is ~100 MPa, two orders of magnitude smaller.

The increase of the piezoresponse near the coercive voltage has been explained as a 90:

switching of an a domain into a c domain [47]. Nanostructures from this sample only show out-

of-plane polarization, unlike the ones prepared from the 10-2 M micellar solution, which also

show in-plane polarization. In the case that those nanostructures have the crystalline axis c

perpendicular to the surface of the substrate, that would mean that the in-plane response is

due to a domains.

All considered, it means that nanostructures prepared by this procedure show ferro-

piezoelectric properties. To the best of the knowledge of this author, this is the first time that

ferroelectricity has been proved in nanostructures prepared by a microemulsion mediated

synthesis.

4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion

layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified

microemulsion mediated synthesis.

In previous section, a preparation route involving microemulsion mediated synthesis and

chemical solution deposition was described. This procedure was proved as a valid one for the

preparation of ferroelectric nanostructures of controlled shape and size. However, due to the

deficient wetting of the substrate by the micellar solution, shrinkage of the coating during the

drying step and the subsequent collapse of the building units network in the deposited layer, it

was not possible to prepare self-assembled nanostructures.

Fig. 4.21 shows drops of sol, microemulsion and micellar solution onto a Pt/TiO2/SiO2/(100)Si

substrates. Micellar solution and sol have a similar contact angle (~42.6° and ~44.4°,

respectively), while the microemulsion wets much better the substrate (contact angle of ~0°).

While having a similar contact angle, micellar solution and sol coatings behave in a different

way as presented in Chapter 3 and previous section. This supports the hypothesis established

at the introductory section of this chapter in which we supposed the micelles would isolate the

nanoparticles of sol.

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

99

Figure 4.21. Drops of (a) sol, (b) microemulsion and (c) micellar solution onto a Pt/TiO2/SiO2/(100)Si

substrate.

In order to improve the coating of the substrate by the micellar layer, prevent a large

shrinkage of the wet deposited layer during drying and obtain self-assembled isolated

crystalline nanostructures, a modification of the preparation procedure is proposed in which

the surface of the substrate is first modified by the deposition of a microemulsion layer.

Substrate here in after will be considered to consist of microemulsion

layer/Pt/TiO2/SiO2/(100)Si. Measurements of the contact angle were carried out to check the

wetting of the micellar solution and sol onto this new substrate and the results are shown in

Fig. 4.22.

Sola)

~44.4°

Microemulsionb)

Micellar solutionc)

~42.6°

100 4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

Figure 4.22. Drops of (a) the micellar solution and (b) the sol onto the

microemulsion layer/Pt/TiO2/SiO2/(100)Si substrate.

One can see that both the sol and the micellar solution contact angles decreased as compared

with those of Fig. 4.21. The contact angle values are ~17.6° for the micellar solution drop and

~17.5° for the sol one. Surfactants have been used extensively in many fields in order to

modify the surface tension of substrates as well as to functionalize them [39], preventing a

large shrinkage of the coating.

As a result, the procedure will be modified. Fig. 4.23 schematizes the proposed drying

mechanism of the modified microemulsion mediated synthesis procedure. Capillary forces will

be opposed by the friction forces between the micelles and the building units or simple

micelles of the micellar solution. In addition, by improving the wetting of the new substrate by

the micellar solution, the shrinkage of the coating during the drying step will be minimized and

so, it is less probable that the building units will collapse and give place to merged

nanostructures.

Micellar solutiona)

~17.6°

Solb)

~17.5°

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

101

Figure 4.23. Proposed drying evolution with time of the modified procedure where the micellar solution

containing the building units is deposited onto a microemulsion layer/Pt/TiO2/SiO2/(100)Si substrate ((a)

and (b)) and resulting oxide nanostructures formed after the thermal treatment of crystallization (c).

Pt polycrystalline

TiO2

SiO2

Si wafer

C6H12

Micellar layerH2OH2OH2O

H2OH2OH2O

H2O +

Sol

H2O +

Sol

H2O +

Sol

1,3-propanediol

Micellar solution

deposition +Drying

Thermal

treatment

Capillary forces

Friction forces

a)

b)

c)

sub

stra

tesu

bst

rate

Mic

ella

r so

luti

on

102 4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

Therefore, the hypothesis is that, by this modified processing method, a self-assembly network

of building units will be formed on top of this microemulsion layer/Pt/TiO2/SiO2/(100)Si

substrate. These building units will be able to maintaining their periodic order onto the

substrate during the drying and the thermal treatment of crystallization, giving rise to final

arrays of oxide nanostructures supported onto conductive coated semiconductor substrates.

The characteristics and the properties of the crystalline nanostructures resulting from this

modified microemulsion mediated synthesis procedure are shown in next sections.

4.3.1. Microscopy and quantitative microstructure analysis

Samples prepared by this procedure show uniform coatings of the substrate by the micellar

solutions and, after the thermal treatment, the surface is undistinguishable from that of a non-

coated substrate, as observed by optical microscopy (Fig. 4.24).

Figure 4.24. Topography by optic microscopy of a sample prepared using the modification of the

microemulsion mediated synthesis method.

Fig. 4.25 shows the SEM images of the surface of the sample. Low magnification image displays

bright spots (Fig, 4. 25 (a)), while the high magnification one (Fig. 4. 25 (b)) presents spots that

seem to be arranged into lines. Three bigger nanostructures, marked with a white cross, are

also observed in Fig. 4. 25 (b), which size can be compared to the bright spots of Fig. 4.25 (a).

50 µm

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

103

Figure 4. 25. SEM image of a sample prepared by this procedure with x5000 (a) and x100000 (b)

magnification. Three lines in light blue mark the direction followed by the nanostructures.

AFM topography images (Fig. 4.26) show small, primary nanostructures and bigger ones. These

images are representative of a 1.5x1.5 cm2 sample area, where similar images can be found at

any zone of the sample, indicating a high uniformity of the coating of the substrate by the

nanostructures. These nanostructures are smaller than the ones of the non-modified

a)

b)

x x

x

104 4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

procedure, described in previous section (Fig. 4.5). The deposited microemulsion layer prior to

the micellar solution deposition functionalizes the substrate, i.e. it establishes a smoother and

non-defective surface onto which the building units are deposited more homogeneously,

distributing uniformly throughout the entire substrate surface, due to the improvement of the

wetting.

Figure 4.26 Topography AFM images of a sample prepared by the modified microemulsion mediated

synthesis. (a) image shows a 3x3 µm area and (b) a 1x1 µm one.

Thus, a high uniform and homogenous coating is obtained and it does not suffer the large

shrinkage described in Fig. 4.11, but the drying evolution shown in Fig. 4.23. As a result, the

nanostructures do not merge into the nanostructures showed at the topography images of

previous section (Fig. 4.9-10).

In addition to enhance the homogeneity of the coating and production of primary

nanostructures, the fact that two consecutive layers of micelles -one from the microemulsion

layer and the next one a building unit from the micellar solution- are deposited gives place to

the self-assembly of the resulting oxide nanostructures. Micelles can self-assemble into

network as in the case of detergents [7], due to the similar amphiphilic character that exhibits

both and, after drying and crystallization, this network will yield self-assembled nanostructures.

A very long drying period, as the one that is taken for these samples, promotes a steady state

in which self-assembly is the most stable state [7].

17.23 nm

0nm

40.81 nm

0nm

a) b)

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

105

Figure 4.27. AFM topography images of two different locations in the substrate and their corresponding

self-correlations, (a) and (b).

Fig. 4.27 shows self correlation images of portions of two different AFM topography images, in

which bigger nanostructures are not observed. It demonstrates the periodic order self-

assembly of the primary nanostructures up to the third neighbours. In the case of 2-D

structures, there are five possible Bravais lattices: square, hexagonal, rectangular, rectangular

centred, and oblique [40], as the ones presented in Fig. 4.28.

Figure 4.28. The five fundamental 2-dimensional Bravais lattices[40].

16nm

24

nma) b) c) c)

106 4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

Self-correlations of pictures of Fig. 4.27 (a) and (b) display a distorted hexagonal network

(angles of the unit cell are not exactly 60°). Therefore, a long-range periodic order of the

crystalline nanostructures seems to be obtained by using this modified procedure for the

preparation of nanostructures. Grosso et al. also found this type of array of nanostructures

when using di-block copolymer synthesis [21].

However, the self-assembly of the nanoparticles is not totally perfect. There are three reasons

for that, all of them complementary. First of all, DLS measurements provide the size of the

colloids but this size does not determine the position of the sol nanoparticle inside the micelle.

Fig 4.29 explains the six possible positions of a sol particle (black) with respect to a micelle [4].

It is known, from the DLS measurements that the sol particles are in the core of the micelle, so

the three possible theoretical situations are (c), (d) or (e). Anyway, when all the surfactant and

the solvent are evaporated, crystallization takes places, and the resulting primary

nanoparticles are obtained on the substrate, if they are not all in the case of Fig. 4.29 (c), then,

they will not be positioned in a perfect self-assembly.

Figure 4.29. Model of location of sol particles in reverse micelles (black ball represent the sol

nanoparticle) [37].

On the other hand, the substrate surface is not perfectly flat and so, the micelles will not form

a perfect ordered layer. Finally, another reason, shown in Fig. 4.5, is that not all the micelles

are building units, and thus, these micelles with only water inside break the continuity of order

of the resulting nanostructures. Consequently, nanostructures derived from the micelles

cannot self-arrange in a perfect network unless these factors are controlled.

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

107

The size distributions of the nanostructures shown in the SEM (Fig. 4.25 (a)) and AFM

topography images (Fig. 4.26) are represented in Fig. 4.30. The size distribution of the

nanostructures presented in Fig. 4.30 (a) is bimodal and the one corresponding to the primary

nanoparticles at the AFM topography image is Gaussian (Fig. 4.30 (b)).

The fitting of their probabilistic lines is analitically expressed by:

y = -2.29 + 0.03·x R = 0.99 (4.3)

y = -2.24 + 0.02·x R = 0.99 (4.4)

for the nanostructures shown at the SEM image Fig. 4.20 (a)

y = -5.76 + 0.25·x R = 0.99 (4.5)

for those displayed at the AFM topography image Fig. 4.23 (a).

Figure 4.30. Equivalent diameter distributions of the nanostructures measured on the SEM image of Fig.

4.25 (a) and the AFM image of Fig. 4.26 (a). Straight lines in the cumulative frequency graph represent

the fitting of the probabilistic lines. There are two in the case of the bimodal distribution (red and blue

lines of (a)).

Following this procedure, the average equivalent diameter calculated for the nanostructures of

the SEM image are 65 nm and 92 nm with a standard deviation of 28 nm and 41 nm,

0.01

0.5

2

10

30

50

70

90

98

99.5

0 100 200

0 100 200

0

25

50

75

Co

un

ts

Deq (nm)

Cu

mu

lative

Fre

qu

en

cy

a)

0,01

0,5

2

10

30

50

70

90

98

99,5

12 14 16 18 20 22 24 26 28 30 32

12 14 16 18 20 22 24 26 28 30 32

0

10

20

30

40

Co

un

ts

Deq(nm)

Cu

mu

lative

Fre

qu

en

cy

b)

108 4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

respectively. As for the AFM images, the equivalent diameter calculated for the primary

nanostructures is 21 nm with a standard deviation of 4 nm.

Both the bimodal and the single distributions are Gaussian, supporting the hypothesis

established in the previous section: nanostructures obtained by this method are formed by the

merging of primary ones, instead of the most common process in ceramics and thin films in

which coalescence and growth from different nucleus of growth takes place and gives rise to

lognormal distribution of particle size.

The nanostructures prepared by the microemulsion mediated synthesis in previous section 4.2

from a micellar solution of sol concentration 5·10-3 M present an average equivalent diameter

of 67 nm (section 4.2.1). This value is comparable with one of the average sizes calculated for

the bimodal distribution, confirming the hypothesis of nanostructures prepared by the first

procedure as resulting from coalescence of primary ones.

By the modified procedure presented in this section, primary nanostructures are, mainly,

prevented from merging into larger ones and those are the ones corresponding to the single

distribution of average size 21 nm. However, there are still a number of primary nanoparticles

that merge into bigger ones, and those are the ones that have a bimodal distribution and a size

comparable with the ones in section 4.2. In that section, the calculated distribution is a

Gaussian and it is not bi-modal, probably because there is not enough statistics to see it.

TEM measurements were carried out in order to investigate the structure of the

nanostructures as well as the substrate/nanostructure interface.

Fig. 4.31 shows bright-field TEM images of cross sections of three nanostructures: two of them,

of ~162 and ~175 nm of lateral size and a smaller one of ~25 nm. Their profiles present

rounded irregular top facets of ~25 nm of width. This support the hypothesis established in

section 4.2.1 of large nanostructures formed by merging of primary ones. In Fig. 4.31 (d), the

primary nanostructures that might yield the observed ~175 nm nanostructure (Fig. 4.31 (c)),

are highlighted in blue. The size of these hypothetical primary nanostructures is smaller than

25 nm of lateral size. AFM topography images provided an average lateral size of ~21 nm,

which is in agreement with the hypothetical size of the primary nanostructures. From these

images it is clear that primary nanostructures set one aside the next one and not on top of it.

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

109

Figure 4.31 Bright-field TEM images of cross section of an isolated nanostructure (a), three

nanostructures formed by the coalescence of primary ones (a-c) and simulated primary nanostructures

disposition (d) that yield the nanostructure in picture above it.

Fig. 4.32 shows the cross-section image of an isolated nanostructure of ~9 nm of lateral size

and ~7 nm of height. (100) planes are easily observed, parallel to the surface of the substrate.

110 4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

The edge of the nanostructure is rounded, almost hemispherical. Isolated small nanostructures

(primary ones) are spherical as confirmed by these TEM images. A sphere is the minimum

energy configuration as so, it is logical that the isolated nanostructures and the corrugation of

the merged big nanostructures are spheres. One can argue that, since the micelles are

spherical and spherical nanopowders have been fabricated from them, it is consistent to find

semi-spherical nanostructures on top the substrate. However, after the drying step, a thermal

treatment is performed. The energy supplied to the system is enough to change the

morphology of the nanostructures if necessary for minimum energy reasons.

Figure 4.32. Bright field TEM image of the cross section of an isolated primary nanostructure

Fig. 4.33 (a-b) shows different higher magnification images of the nanostructures, where the

(101) planes are easily distinguishable. In Fig. 4.33 (a), they all have the same orientation with

respect to the substrate surface. Fig. 4.33 (b) shows two different regions where (100) planes

can be observed with a relative rotation of 17.5° (the edge of each region is marked by the last

planes, in blue). The merging region or tilt boundary presents a number of edge dislocations of

spacing D = 1.63 nm and Burger vector |b| = 4.9 Å, calculated using the relation |b| = D·sin θ.

This kind of boundary is usually called “pure tilt boundary” and it was suggested [41] that low-

angle boundaries between adjoining crystallites or crystal grains consist of arrays of

dislocations, as in this case.

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

111

Fig. 4.33 (c) shows the (001) planes of a PbTiO3 nanostructure. In this image, the edge

dislocations inside of the nanostructures are highlighted. This result, combined with the one

obtained in Fig. 4.32 indicates that the orientation of the nanostructure is (001). The sample is

crystallized in the cubic phase (650 :C) and during the cooling step and subsequent phase

transition is when the unit cells lengthen in one direction and set the orientation as (100) or

(001). This is the minimum energy type of growth [42].

Figure 4.33 High magnification bright-field HRTEM images of the inside of the nanostructures (a) (101)

planes, (b) (101) planes in adjoining parts of the nanostructure with a relative tilt of 17.5 ° and (c) edge

dislocation, marked with an arrow.

112 4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

Defects as vacancies or impurities might clamp their movement, clamping the ferroelectric

domains and yielding piezoresponse hysteresis loops with the shape of the one presented in

Fig. 4.18 (c).

In all the images presented in the above TEM study, it is clear that the substrate

/nanostructure interface is clean and that some nanostructures contain dislocations. Residuals

coming from the preparation procedure used (i.e. a layer with residuals of the precursor

microemulsion, amorphous phases, …) are not observed.

The (101) planes, shown in Fig. 4.33 (a-b), correspond to the same nanostructure. The distance

between them is ~2.9 Å of average. Fig. 4.33 (c) presents the (001) planes with an interplanar

distance of 4.0 Å of average. The planes (100) of Fig. 4.32 are ~3.7 Å of average. This indicates

a tetragonal unit cell and thus a perovskite phase with lattice parameters smaller than the

theoretical ones. However, these measurements are local ones and it is difficult to extrapolate

these results without further statistics.

To conclude, these measurements confirm the perovskite phase and the small size of the

primary nanostructures. They also confirm the formation of larger nanostructures by the

merging of primary ones.

4.3.2. Structural characterization

Fig. 4.34 displays the 2-dimensional synchrotron X-ray diffraction pattern of the sample

prepared from the 5·10-3 M micellar solution.

The pattern consists of four broad peaks and four sharp rings. The maximum intensity is found

at the broad peaks and at certain positions of the rings sectors.

Fig. 4.35 shows the integration of the experimental 2-D diffraction pattern, in solid black and

the simulated lead titanate and platinum diffraction peaks in red and blue, respectively. Each

peak corresponds to one of the sharp rings or broad ring portions of the 2-D diffraction pattern.

It presents four peaks. The most intense peaks correspond to the polycrystalline platinum of

the substrate, having also contributions of PbTiO3 perovskite phase.

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

113

Fi

gure

4.3

4. E

xper

imen

tal 2

-D s

ynch

rotr

on

x-r

ay d

iffr

acti

on

pat

tern

of

a sa

mp

le p

rep

ared

fro

m t

he

5·1

0-3

M m

icel

lar

solu

tio

n.

114 4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

Figu

re 4

. 35

dif

frac

tio

n p

atte

rn c

alcu

late

d f

rom

th

e in

tegr

atiio

n o

f th

e 2

-D e

xper

imen

tal p

atte

rn o

f Fi

g. 4

.34

(b

lack

so

lid li

ne)

an

d s

imu

late

d d

iffr

acti

on

pat

tern

s o

f th

e P

bTi

O3

per

ovs

kite

nan

ost

ruct

ure

s (r

ed

lin

e) a

nd

Pt

bo

tto

m e

lect

rod

e (b

lue

line)

.

Intensity(a.u.)

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

115

Figu

re

4.3

6.

Exp

erim

enta

l 2

-D

dif

frac

tio

n

pat

tern

w

ith

re

fle

ctio

ns

corr

esp

on

din

g to

P

t (a

) an

d

Pb

TiO

3 p

ero

vski

te

(b)

and

th

e si

mu

late

d

2-D

d

iffr

acti

on

p

atte

rn

of

po

lycr

ysta

llin

e p

lati

nu

m w

ith

(1

11

) fi

ber

tex

ture

(c)

an

d P

bTi

O3

nan

ost

ruct

ure

s w

ith

(1

00

) fi

ber

tex

ture

.

(111)

(111)

(200)

(200)

(111)

(111)

(200)

(200)

(010)

(111)

(110)

(100)

(110)

(100)

a)

b)

c)

d)

(010)

(111)

(110)

(100)

(110)

(100)

116 4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

Fig. 4.36 shows a more detailed identification of the broad spots and ring sectors of the 2-D

diffraction pattern. Simulation of the diffraction peaks for the polycrystalline platinum layer

explains the broad diffraction peaks as resulting of a (111) fiber texture with an orientation

distribution cone of ±3°, as was previously observed for the ultrathin film in section 3.2.2. The

simulated pattern was calculated from the inverse pole figure. Fig. 4.36 (c) shows this

simulation.

Fig. 4.36 (d) shows the simulated diffraction pattern for the PbTiO3 perovskite phase,

considering the lattice parameters to be a = b = 3.890(0) Å and c = 4.056(7) Å. It is the result of

adding up the simulated diffraction patterns for all the crystals with axis rotated (-7:, -7:, ±20:)

and (20:, -7:, -7:±20:) as seen in the TEM image of Fig. 4.32 (b), where (101) planes presented

a tilt angle of 17.5°. The crystalline distortion is c/a = 1.042(8), much smaller than the

theoretical one (1.066(8) being, a = 3.899(9) Å and c = 4.150(0) Å). The results for the ultrathin

film in previous section 3.2.2 imply lattice parameters equal to a = b = 3.912(1) Å and

c = 4.130(1) Å.

A previous study [43] indicates that, after the drying step in a sol-gel deposition procedure, a

thin film is subjected to tensile stresses, which explains the lattice distortion of the ultrathin

film. However, it is reported in the same work as well as others [44, 45] that this stress is

minimized with film thickness. Primary nanostructures fabricated by this procedure are ~10

nm of height and the uncoated substrate surface is much larger than the area occupied by the

oxide nanostructures. In addition, drying time is very long, so this source of stress should be

minimized and cannot explain by itself the smaller distortion of the nanostructures.

Previous works in nanopowders fabricated by a microemulsion mediated synthesis proved the

high crystallinity of the ones prepared by this procedure [18, 22], which means that variation in

the crystallinity throughout the sample is not contemplated as a reason for a low tetragonal

distortion. The size of the nanostructures here presented is one of the smallest for individual

structures onto substrates reported in the literature. Due to the so small size of the

nanostructures, the crystal lattice is subjected to stress, yielding the distortion in the c

direction that explains such a small tetragonal distortion.

The 2-D diffraction pattern of an ultrathin film (section 3.2.2), also shows a crystalline

distortion smaller than the theoretical one that was attributed to stresses due to the mismatch

between the substrate and the PbTiO3 perovskite phase and the different thermal expansion

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

117

of the platinum layer and the PbTiO3 ultrathin film. Both reasons apply here and most probably

are6 the main cause for the stress found.

The primary nanostructures observed by bright-field HRTEM presented an interplanar distance

of ~2.85 Å for the (101) planes of one merged nanostructure and ~3.71 Å for the (100) planes

of a primary nanostructure. These distances imply different values of the lattice parameters of

these individual nanostructures. However, the values obtained by X-ray diffraction represent

an average value of all the nanostructures present at the sample.

It was not possible to simulate the diffraction pattern from the inverse pole figure. The sample

does not have a well defined fiber texture and it is not possible to apply the statistical

describer the Anaelu software uses [46]. Not being a continuous sample, it might be that there

is not a polycrystalline structure with a valid statistic. Thus, we have to speak about crystals or

small groups of crystals. This reinforces the hypothesis of these particles as primary ones.

4.3.3. Functional characterization

PFM measurements were performed on the isolated nanostructures prepared by this modified

microemulsion mediated synthesis, using a 5·10-3 M. Images corresponding to the out-of-plane

polarization are shown in Fig. 4.37.

The experiments were carried out at the Max Plank Institute for Microstructure Physics under

an AC field of amplitude 1.5V -peak to peak- and at a frequency close to the contact resonance

of the cantilever (245kHz) in order to enhance the measurements using the ATEC-EFM

cantilevers, as explained in Chapter 2 (Experimental Procedure, Section 2.6).

The image shows nanostructures with different phase and amplitude. 180: phase difference

can be observed at Fig. 4.37 (d) and through the amplitude profile (Fig. 4. 37 (e)).

The out-of–field hysteresis loop of an isolated nanostructure, shown in Fig, 4.38, verifies the

ferro-piezoelectric character of these PbTiO3 nanoparticles. It is asymmetric with respect to

the piezoresponse and the voltage axis. As discussed in section 3.2.3, it means that the

nanostructure has a certain imprint [47] and a certain offset due to the presence of a pinned

layer [48]. In a previous Section, Fig. 4.33 (b) showed two regions of the same nanostructure

with a relative tilt of the (101) planes and edge dislocations that enable such tilt. This is a

probable area where charges can clamp the motions of the dislocations creating a pinned area

that would be too small to switch.

118 4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

Figu

re 4

.37

. To

po

grap

hy

(a),

ph

ase

(b)

and

am

plit

ud

e (c

) o

ut-

of-

pla

ne

pie

zore

spo

nse

im

ages

fo

r th

e is

ola

ted

nan

ost

ruct

ure

s o

f a

sam

ple

pre

par

ed b

y th

e

mo

dif

ied

mic

roem

uls

ion

syn

thes

is m

eth

od

. Im

age

(d)

corr

esp

on

ds

to a

hig

h m

agn

ific

atio

n im

age

; it

s p

has

e p

rofi

le (

e) a

re m

arke

d in

blu

e in

th

e p

has

e im

age.

20

0nm

20

0n

m2

00

nm

PZ

27

.80

nm

0n

m

11

pm

/V

0.0

0

PFM

Im

age

Am

plit

ude im

age

PFM

Im

age

Phase im

age

AFM

Im

age

Out-

of-

pla

ne

a)

b)

c)

150

100

50

0

12

10 8 6 4 2 0

X[n

m]

Z[µm]

+90º

-90º0º

d)

e)

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

119

This hysteresis loop shows both kinds of asymmetries from the very first loop. As in the case of

the merged nanostructures from previous section, there is a certain initial polarization, which

means that there are charges on the surface of the substrate prior to the application of any DC

field. The fact that regardless the sample and the procedure followed to obtain the isolated

nanostructures the same initial polarization is obtained, might mean that microemulsion

mediated synthesis, combined with a CSD method as the one here used, yield to

nanostructures with a defined initial polarization that gives rise to the surface charges and not

the other way around. Previous studies on ferroelectric perovskite materials grown epitaxial

onto single crystal substrates suggested that the elastic accommodation of the in-plane strain

could lead to an enhancement of the spontaneous polarization of simple ferroelectric

perovskites along the out-of-plane direction [49].

Measurement conditions: Vac = 1.5V

Freq ac = 235kHz F = 30 nN

Tbias = 0.2s Tmeasurement = 0.3s

SEN = 2mV TC = 0.1 s

Figure 4.38. Piezoresponse hysteresis loop obtained in the nanostructure, which phase profile is shown

in Fig. 4.37 (e). Its lateral size is ~37 nm and its height is ~15 nm as measured from the images of

Fig. 4.37.

The effect of the pinned layer is more important in this case than it was for the nanostructures

prepared by the instability of ultrathin films in previous Chapter 3 (section 3.2) and the

nanostructures prepared by the non modified procedure of previous section 4.2. It was

explained before that this effect occurs at the interface between the substrate and the

nanostructure and that the smaller the nanostructure is, the more relevant this effect

becomes. Consequently, the hysteresis loop is shifted so much that, most of it, is in the

positive side of the piezoresponse axis.

-10 -8 -6 -4 -2 0 2 4 6 8 10

-10

0

10

20

30

40

50

Pie

zo

resp

on

se

(a

.u.)

Voltage (V)

120 4.3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si substrates prepared by the modified microemulsion mediated synthesis

In this case, an increase of the piezoresponse near the coercive voltage is not observed, as it

was for the merged nanostructures. This might indicate that the non-pinned layer is a single

domain that switches 180°. Ferroelectric PbTiO3 nanodots were found to present a stable

polarization due to surface charges, instead of the supposed vortex state [50]. This state could

only be reverted by 180° switching, as it is happening in the nanostructures prepared by this

modified method.

The smallest size where ferroelectricity has been demostrated by PFM has been in

nanostructures of 41 nm of lateral size [51, 52], while the nanostructure which hysteresis loop

is here represented is ~37 nm, becoming one of the smallest nanostructures where ferro-

piezoresponse is proved by PFM to date. Note that the resolution of SPM in contact mode is

compromised when the dimension of the measured nanostructures is equal or smaller than

the one of the tip [53]. Actually, the real dimensions of the nanostructure are even smaller, as

shown previously by TEM, where primary nanostructures with ~8.99 are observed.

Chapter 4: Ferroelectric nanostructures by microemulsion mediated synthesis onto Pt/TiO2/SiO2/(100)Si substrates.

121

Remarks

1. At the micellar solution, micelles act as building units, isolating the sol nanoparticles

inside the core of the micelle and yielding, after drying and thermal treatment of the

coatings, the nanostructures onto the Pt/TiO2/SiO2/(100)Si substrates.

2. By using a combination of microemulsion mediated synthesis and CSD, nanostructures

with controlled size and shape are obtained. Size does not depend on the sol

concentration: the average size of the crystalline nanostructures obtained by this

procedure is ~75 nm for the ones prepared from the 10-2 M micellar solution and

~67 nm for those prepared from the 5·10-3 M one. Only the number of nanostructures

on the substrate does. Those nanostructures are not ordered.

3. Nanostructures prepared by this procedure are formed by merging of primary ones.

Size distributions are Gaussian, i.e. the formation of the nanostructures does not take

place via normal grain growth, but from a limited number of primary nanostructures.

This is due to the capillary forces during drying that attract micelles to each other, as

well as to the defects of the substrate surface and bad wetting of the coating.

4. PFM images show ferro-piezoresponse of the nanostructures, proving their

ferroelectric character. Mechanism of switching in these nanostructures is complex,

involving 90° switching of a domains and the creation of a pinned layer, stabilized by

the charges on the surface of the platinum bottom electrode of the substrate that

introduce themselves at the domain wall.

5. This first procedure does not provide a periodic array of ordered nanostructures onto

the substrate. In order to improve the wetting and to minimize the shrinkage of the

micellar solution coating during drying, a modification of the previous preparation

method is proposed, in which a microemulsion layer will constitute the surface onto

which the micellar solution will be spun-coated. In this way, isolated primary

nanostructures with size distributions of average size ~21 nm are obtained as observed

on the AFM images, as well as a reduced number of merged bigger ones.

6. TEM results proved that isolated nanostructures down to ~9 nm of lateral size can be

observed. These nanostructures are found all over the sample, showing the uniformity

and quality of the coating. Both the primary nanostructures and larger, merged ones

show Gaussians size distributions.

122 Remarks

7. The primary nanostructures prepared by this late modified procedure shows

hexagonal order. A long-range order is not observed due to the presence of some

merged bigger nanoparticles, the inhomogeneous surface of the platinum substrate ,

the fact that not all the deposited micelles are building units and, probably, to the fact

that not all sol nanoparticles are exactly at the center of the building units.

8. The synchrotron radiation x-ray diffraction pattern of the nanostructures obtained by

the modified method shows PbTiO3 perovskite reflections with cell parameters

a=b=3.890(0) Å and c=4.056(7) Å and crystal axis rotated (-7°, -7°, ±20°) and (20°, -7°, -

7°±20°). The reason for this tetragonal distortion smaller than the theoretical one

might be the stress due to the large surface area to volume ratio of the sample that

implies the small size of the nanostructures, the mismatch between substrate and

nanostructure lattices and the different thermal expansion coefficients of both of

them.

9. Nanostructures obtained by this modified procedure are ferroelectric and seem to be

single domains over a pinned layer. Ferro-piezoresponse is measured in this work in

one of the smallest nanostructures where this kind of measurements could be carried

out by PFM to date, with lateral size of ~37 nm as determined by AFM topography

images.

Bibliography. 123

Bibliography

[1] H. Wennerstrom and B. Lindman, "Micelles - physical-chemistry of surfactant

association", Physics Reports-Review Section of Physics Letters, 52 (1), 1979, p:1

[2] W.D. Harkins, "A general theory of the reaction loci in emulsion polymerization II", The

journal of chemical physics, 13, 1945, p:47

[3] K. Landfester, "The generation of nanoparticles in miniemulsions", Advanced materials,

13 (10), 2001, p:765

[4] M.P. Pileni, "Reverse micelles as microreactors", Journal of Physical Chemistry, 97,

1993, p:6961

[5] C. Tanford, "The hydrophobic effect and the organization of living matter", Science,

200, 1978, p:1012

[6] S. Mann, S.L. Burkett, S.A. Davis, C.E. Fowler, N.H. Mendelson, S.D. Sims, D. Walsh and

N.T. Whilton, "Sol-gel synthesis of organized matter", Chemistry of Materials, 9 (11), 1997,

p:2300

[7] C.J. Brinker, Y.F. Lu, A. Sellinger and H.Y. Fan, "Evaporation-induced self-assembly:

Nanostructures made easy", Advanced materials, 11 (7), 1999, p:579

[8] D. Grosso, F. Cagnol, G. Soler-Illia, E.L. Crepaldi, H. Amenitsch, A. Brunet-Bruneau, A.

Bourgeois and C. Sanchez, "Fundamentals of mesostructuring through evaporation-induced

self-assembly", Advanced Functional Materials, 14 (4), 2004, p:309

[9] G. Schinkel, I. Garrn, B. Frank, U. Gernert, H. Schubert and R. Schomacker, "Fabrication

of alumina ceramics from powders made by sol-gel type hydrolysis in microemulsions",

Materials Chemistry and Physics, 111 (2-3), 2008, p:570

[10] J. Tartaj and P. Tartaj, "Two-Stage Sintering of Nanosize Pure Zirconia", Journal of the

American Ceramic Society, 92 (1), 2009, p:S103

[11] D.E. Zhang, Z.W. Tong, S.Z. Li, X.B. Zhang and A.L. Ying, "Fabrication and

characterization of hollow Fe3O4 nanospheres in a microemulsion", Materials letters, 62 (24),

2008, p:4053

[12] M. Boutonnet, J. Kizling, P. Stenius and G. Maire, "The preparation of monodisperse

colloidal metal particles from microemulsions", Colloids and Surfaces, 5 (3), 1982, p:209

[13] S. Hingorani, V. Pillai, P. Kumar, M.S. Multani and D.O. Shah, "Microemulsion mediated

synthesis of zinc-oxide nanoparticles for varistor studies", Materials Research Bulletin, 28 (12),

1993, p:1303

[14] Y.F. Lu, H.Y. Fan, A. Stump, T.L. Ward, T. Rieker and C.J. Brinker, "Aerosol-assisted self-

assembly of mesostructured spherical nanoparticles", Nature, 398 (6724), 1999, p:223

124 Bibliography

[15] K.P. Das and J.E. Kinsella, "Stability of food emulsions: physicochemical role of protein

and nonprotein emulsifiers", Advances in Food and Nutrition Research, 34, 1990, p:81

[16] Y.L. Khmelnitsky, A.V. Levashov, N.L. Klyachko and K. Martinek, "Engineering

biocatalytic systems in organic media with low water-content", Enzyme and Microbial

Technology, 10 (12), 1988, p:710

[17] T.H. McHugh, "Protein-lipid interactions in edible films and coatings", Nahrung-Food,

44 (3), 2000, p:148

[18] C. Beck, W. Härtl and R. Hempelmann, "Size-controlled synthesis of nanocrystalline

BaTiO3 by a sol-gel type hydrolysis in microemulsion-provided nanoreactors", Journal of

Materials Research, 13 (11), 1998, p:3174

[19] S. Bhattacharyya, S. Chattopadhyay and M. Alexe. "Fabrication of isolated ferroelectric

nanostructures". in Materials Research Society Symposium. 2003. San Francisco.

[20] S. Clemens, S. Rohrig, A. Rudiger, T. Schneller and R. Waser, "Embedded ferroelectric

nanostructure arrays", Nanotechnology, 20 (7), 2009, p:5

[21] D. Grosso, C. Boissiere, B. Smarsly, T. Brezesinski, N. Pinna, P.A. Albouy, H. Amenitsch,

M. Antonietti and C. Sanchez, "Periodically ordered nanoscale islands and mesoporous films

composed of nanocrystalline multimetallic oxides", Nature Materials, 3 (11), 2004, p:787

[22] H. Herrig and R. Hempelmann, "A colloidal approach to nanometre-sized mixed oxide

ceramic powders", Materials letters, 27, 1996, p:287

[23] S. Kronholz, S. Rathgeber, S. Karthauser, H. Kohlstedt, S. Clemens and T. Schneller,

"Self-assembly of diblock-copolymer micelles for template-based preparation of PbTiO3

nanograins", Advanced Functional Materials, 16 (18), 2006, p:2346

[24] L.S. Ee, J. Wang, S.C. Ng and L.M. Gan, "Low temperature synthesis of PZT powders via

microemulsion processing", Materials Research Bulletin, 33 (7), 1998, p:1045

[25] C. Pithan, Y. Shiratori, R. Waser, J. Dornseiffer and F.H. Haegel, "Preparation,

processing, and characterization of nano-crystalline BaTiO3 powders and ceramics derived

from microemulsion-mediated synthesis", Journal of the American Ceramic Society, 89, 2006,

p:2908

[26] H. Herrig and R. Hempelmann, "Microemulsion mediated synthesis of ternary and

quaternary nanoscale mixed oxide ceramic powders", Nanostructured Materials, 9 (1-8), 1997,

p:241

[27] C. Pithan, Y. Shiratori, A. Magrez, S.B. Mi, J. Dornseiffer and R. Waser, "Consolidation,

microstructure and crystallography of dense NaNbO3 ceramics with ultra-fine grain size",

Journal of the Ceramic Society of Japan, 114 (1335), 2006, p:995

Bibliography. 125

[28] Y. Yamashita, H. Yamamoto and Y. Sakabe, "Dielectric properties of BaTiO3 thin films

derived from clear emulsion of well-dispersed nanosized BaTiO3 particles", Japanese Journal of

Applied Physics, 43 (9B), 2004, p:6521

[29] J. Kim, J.K. Kim, S. Heo and H.S. Lee, "Ferroelectric properties of sol-gel prepared La-

and Nd-substituted, and Nb-co-substituted bismuth titanate using polymeric additives", Thin

Solid Films, 503 (1-2), 2006, p:60

[30] S.S. Kim, E.K. Choi, J.K. Kim, J.S. Kim, T.K. Song and J. Kim, "Effects of surfactant on

surface morphology and orientation of Nb-doped Bi4Ti3O12 thin films", Journal of the Korean

Physical Society, 42, 2003, p:S1126

[31] M.P. Pileni, "The role of soft colloidal templates in controlling the size and shape of

inorganic nanocrystals", Nature Materials, 2, 2003, p:145

[32] J. Tyndall, "On the Blue Colour of the Sky, the Polarization of Skylight, and on the

Polarization of Light by Cloudy Matter Generally", Proceedings of the Royal Society of London,

17, 1868, p:223

[33] M. Ramirez, J. Bullon, J. Anderez, I. Mira and J.L. Salager, "Drop size distribution

bimodality and its effect on O/W emulsion viscosity", Journal of Dispersion Science and

Technology, 23 (1-3), 2002, p:309

[34] C.J. Brinker and G.W. Scherer, Sol-gel science. The physics and chemistry of sol-gel

processing. 1990, San Diego: Academic Press.

[35] C. Beck, W. Härtl and R. Hempelmann, "Size-controlled synthesis of nanocrystalline

BaTiO3 by a sol-gel type hydrolysis in microemulsion-provided nanoreactors", Journal of

Materials Research, 13 (11), 1998, p:3174-3180

[36] H. Herrig and R. Hempelmann, "A colloidal approach to nanometre-sized mixed oxide

ceramic powders", Materials letters, 27, 1996, p:287-292

[37] M.P. Pileni, "Reverse micelles as microreactors", Journal of Physical Chemistry, 97,

1993, p:6961-6973

[38] M.P. Pileni, "The role of soft colloidal templates in controlling the size and shape of

inorganic nanocrystals", Nature Materials, 2, 2003, p:145-150

[39] G. Decher, Layered nanoarchitectures via directed assembly of anionic and cationic

molecules. Comprenhesice supramolecular chemistry, ed. J.-P. Sauvage and M.W. Hosseini. Vol.

9. 1996, Oxford: Pergamon Press.

[40] C. Hammond, Introduction to crystallography. Royal Microscopical Society. Microscopy

Handdbooks. Vol. 19. 1992, Oxford: Oxford university press.

[41] C. Kittel, Introduction to solid state physics. 6th ed. 1986, New York: Wiley.

126 Bibliography

[42] J. Ricote, R. Poyato, M. Algueró, L. Pardo and L. Calzada, "Texture development in

modified lead titanate thin films obtained by chemical solution deposition on silicon-based

substrates", Journal of the American Ceramic Society, 86 (9), 2003, p:1571

[43] J. Mendiola, M.L. Calzada, P. Ramos, M.J. Martin and F. Agulló-Rueda, "On the effects

of stresses in ferroelectric (Pb, Ca)TiO3 thin films", Thin Solid Films, 315, 1998, p:195

[44] A. Bartasyte, O. Chaix-Pluchery, J. Kreisel, C. Jimenez, F. Weiss, A. Abrutis, Z. Saltyte

and M. Boudard, "Investigation of thickness-dependent stress in PbTiO3 thin films", Journal of

Applied Physics, 103 (1), 2008, p:014103

[45] T. Ohno, B. Malic, H. Fukazawa, N. Wakiya, H. Suzuki, T. Matsuda and M. Kosec. "Origin

of compressive residual stress in alkoxide derived PbTiO3 thin film on Si wafer". in 25th

Meeting on Ferroelectric Materials and Their Applications (FMA-25). 2008. Kyoto, JAPAN.

[46] L. Fuentes, "Anomalous scattering and null-domain ghost corrections for fibre

textures", Textures and Microstructures, 10, 1989, p:347

[47] H.N. AlShareef, D. Dimos, W.L. Warren and B.A. Tuttle, "Voltage offsets and imprint

mechanism in SrBi2Ta2O9 thin films", Journal of Applied Physics, 80 (8), 1996, p:4573

[48] B.J. Rodriguez, S. Jesse, M. Alexe and S.V. Kalinin, "Spatially Resolved Mapping of

Polarization Switching Behavior in Nanoscale Ferroelectrics", Advanced materials, 20 (15),

2008, p:102

[49] M.-W. Chu, I. Szafraniak, R. Scholz, C. Harnagea, D. Hesse, M. Alexe and U. Gösele,

"Impact of misfit dislocations on the polarization instability of epitaxial nanostructured

ferroelectric perovskites", Nature Materials, 3, 2004, p:87

[50] J. Wang and M. Kamlah, "Domain control in ferroelectric nanodots through surface

charges", Applied Physics Letters, 93 (26), 2008, p:262904

[51] A. Roelofs, I. Schneller, K. Szot and R. Waser, "Piezoresponse force microscopy of lead-

titanate nanograins possibly reaching the limit of ferroelectricity", Applied Physics Letters, 81

(27), 2002, p:5231

[52] A. Roelofs, T. Schneller, K. Szot and R. Waser. "Towards the limit of ferroelectric

nanosized grains". in 3rd International Conference on Trends in Nanotechnology. 2002.

Santiago Composte, Spain: Iop Publishing Ltd.

[53] J.S. Villarrubia, "Algorithms for scanned probe microscope image simulation, surface

reconstruction, and tip estimation", Journal of Research of the National Institute of Standards

and Technology, 102 (4), 1997, p:425

CHAPER 5: NANOSTRUCTURES ONTO SrTiO3 SINGLE-

CRYSTAL SUBSTRATES BY MICROEMULSION MEDIATED

SYNTHESIS

5.1. Towards ideal surfaces

Previously, in Chapter 4, it was shown that the deposition of micellar solutions onto Pt coated

Si based substrates yield PbTiO3 ferroelectric nanostructures and that the self-arrangement of

primary nanostructures can be achieved on polycrystalline substrates. However, such

nanostructures presented a number of defects related to the lattice mismatch between the

polycrystalline Pt top layer and the PbTiO3 grown nanostructures that were observed either

directly (see Fig. 4.28 where dislocations are found inside the bulk nanostructures) or indirectly

from the piezoelectric measurements at local scale (see Fig. 4.17, Fig. 4.35 and the discussion

that follows).

It was proposed as a target of this thesis to learn about effects derived from the scaling down

of the size of the nanostructures. But it is difficult to separate these effects from the extrinsic

ones due to the mismatch between the substrate and the PbTiO3 nanostructures [1]. By using

a single crystal substrate, as will be done in this Chapter, surface, nanostructure crystal

structure and nanostructures defects should be minimized and this can lead to better

ferroelectric performance while keeping the nano-scale size and ordering of the

nanostructures.

The selected single crystal is SrTiO3 (STO) often used to promote epitaxial growth of

ferroelectric perovskites by a number of techniques (sputtering, PLD, etc…) [1-6]. At room

temperature, STO has a cubic cell (space symmetry 𝑃 4 𝑚 3 2 𝑚 , space group 221) with

lattice parameters a = 3.900(5) Å and a perovskite like structure. This structure can be

described as stacks of alternating layers of SrO and TiO2 planes.

In previous Chapter 4, micelles were proved as valid building units to obtain isolated

nanostructures onto Pt coated Si(100) substrates. Thus, the combination of a single crystal

substrate with lattice parameters close to the ones of the PbTiO3 perovskite, and the

microemulsion mediated synthesis previously discussed, is supposed to yield primary

nanostructures with a periodic order and that would grow epitaxially on the substrate surface.

128 5.2. Nanoscale PbTiO3 structures onto commercial as-served SrTiO3 substrates prepared by

microemulsion mediated synthesis

5.2. Nanoscale PbTiO3 structures onto commercial as-served SrTiO3

substrates prepared by microemulsion mediated synthesis.

Quality of the substrate surfaces is important. This surface should be as close as possible to an

ideal one to preserve, over macroscopic surface areas, the expected arrangement of the

micelles in the solution-derived layer and in the resulting crystalline nanostructures after heat

treatment. In this study, one side polished commercial (100)SrTiO3 substrates will be used as

served.

Fig. 5.1 shows the AFM topography images of an as-received STO substrate. Terraces of the

substrate vary in length and width and they present a non-well defined edges.

Figure 5.1 AFM topography images of an as-served SrTiO3 substrate (a-b) and profile (d) along the blue

line of image (c).

The surface of the substrate was analyzed by LEED and AES. It was not possible to measure the

as-received substrates due to the strong charging effects and the, as expected, contaminated

surface. Even if they were used after a surface cleaning treatment, contact with the

5004003002001000

7

6

5

4

3

2

1

0

X[nm]

Z[Å

]a) b)

19.33 Å

0.00 Å

c) d)

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

129

atmosphere would contaminate them and turn them to the previous state. It is a well known

fact that SrO reacts with CO2 and H2O at room temperature to form carbonate and hydroxide

compounds following the reaction:

3SrO + 2CO2 + H2O → Sr(OH)2 + 2SrCO3 (5.1)

and, therefore, contaminating the substrate surface.

Micellar solutions with concentrations of 10-2 M and 5·10-3 M were deposited onto the as-

served STO substrates and thermally treated by RTP at 650°C (as explained in Chapter 2,

section 2.3) to obtained the crystalline PbTiO3 nanostructures that are studied next.

5.2.1. Microscopy analysis.

When using the microemulsion mediated synthesis combined with CSD technology onto a

commercial as-served SrTiO3 substrate, coating of the substrate is not uniform. Fluid steam

rings, as shown in Fig. 5.2, are formed during deposition, drying and thermal treatment of

crystallization, mainly due to a non-uniform wetting of the substrate by the micellar solution

and, consequently, non homogenous evaporation of the solvent during the spin-coating and

drying steps.

Figure 5.2. Optical image of the surface of a sample prepared onto a commercial and as-received SrTiO3

substrate.

A detailed study using SEM was performed along one of the radii of the quasi-circular coating,

revealing the different morphology of the coating depending on the distance to the thickest

area (Fig. 5.3). Thus, the coating formed by the inner rings, which are closer, presents an

50 µm

130 5.2. Nanoscale PbTiO3 structures onto commercial as-served SrTiO3 substrates prepared by

microemulsion mediated synthesis

Figu

re 5

.3.

SEM

im

ages

of

the

mo

rph

olo

gy o

f th

e sa

mp

le p

rep

ared

fro

m t

he

5·1

0-3

M m

icel

lar

solu

tio

n o

nto

an

as-

rece

ived

STO

su

bst

rate

in

dep

end

ence

of

the

thic

knes

s o

f th

e co

atin

g al

on

g th

e su

bst

rate

su

rfac

e.

d=

120

0μm

d=

640

0μm

d=

320

0

μm

III

IIIIV

V

IIIV

V

IIII

-bIII

-a

[100

][0

10

][1

00]

[01

0]

[100

][0

10]

[100

][0

10

][1

00]

[010

]

Ce

nte

r o

f th

eco

ati

ng

Bo

rder

of

the

coat

ing

[10

0]

[010

]

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

131

almost continuous deposit, where some pyramids of rounded facets protrude (area I in Fig.

5.3). At outer rings, a network of lines of similar width can be observed (area II in Fig. 5.3),

following the [100] and [010] directions of the substrate. When the amount of deposited

material decreases at outer rings, the continuous network breaks into truncated pyramids of

square and rectangular basis that may merge (Fig. 5.3, area III) following the same directions

than the initial network of lines. For even outer rings, there is so little material that the

incipient truncated pyramids coexist with round small PbTiO3 nanostructures (area IV, Fig. 5.3).

Despite the non uniform coating, in some areas, a 2-dimensional short range periodicity of the

nanostructures is found. These arrays follow the *100+ direction and a 45⁰ direction, most

probably the [110] one. Directions are established with respect to the facets of the single

crystal. A similar morphology is found, despite of the concentration of the micellar solution

used.

The estimated size of the square based pyramids of area III-b (Fig. 5.3) is ~50 nm while the

rectangular based ones are ~20-50 nm wide and ~100 nm long. The isolated nanostructures

formed in area V are ~30 nm of lateral size.

The fact that areas with different thickness and topography are found when using

microemulsions have been already reported by Carvalho et al. [7]. They prepared block

copolymers films and observed a similar phenomenon to the one here described that depends

on the relative thickness of the coating of the different areas, as seen in Fig. 5.4.

Figure 5.4. SEM image of the morphology of a coating of a block-copolymer film [7].

However, the morphology of the different areas depends on the type of growth of the

nanostructures onto the substrate. When the phase formation phenomena do not depend on

132 5.2. Nanoscale PbTiO3 structures onto commercial as-served SrTiO3 substrates prepared by

microemulsion mediated synthesis

the kinetics of nucleation and growth, the mode of nucleation on foreign substrate may be

formulated on the basis of macroscopic thermodynamic considerations [8], taking into account

the surface and interface energies for lattice mismatch systems [9]. Results shown in previous

Chapter 4 already proved that this is the case when using the microemulsion mediated

synthesis, as the nanostructures prepared by this method present Gaussian size distributions,

which are related to independent nucleation points, either isolated building units or groups of

these (see sections 4.2.1 and 4.3.1).

Growth can be categorized in three different modes, summarized in Fig. 5.5.

a) the Frank-van der Merwe mode , when the adhesion between the substrate and the

growing structure is strong and the relative mismatch is small. It is a layer by layer type

of growth and before growing a new layer, the old one should be complete.

b) the Stranski-Krastanov mode, in the case of strong adhesion between the growing

structure and the substrate and a relatively large mismatch. After the completion of a

certain number of monolayers, 3-dimensional cluster of structures grow onto them.

c) the Volmer-Webber mode, for a weak adhesion between substrate and growing

structure. 3-dimensional cluster are observed right from the beginning. The relative

mismatch between the crystalline lattices determines the strain to which the 3-

dimension nuclei are subjected.

If the SEM images are considered as a scheme of how the isolated nanostructures and

subsequent continuous films grow onto the substrate (supposing the growing steps in this

order: Fig 5.3 (V), (IV), (III-a), (III-b), (II), (I)), it can be concluded that they follow the Frank-van

der Merwe growing mode, as it would be expected for the small lattice mismatch between

SrTiO3 and PbTiO3 and the strong adhesion when the last layer of the substrate or the first one

from the nanostructure is the shared TiO2 one.

AFM topography images of Fig. 5.6 confirm the results obtained from the SEM images and

provide additional information about the arrangement of the nanostructures onto the

substrate surface. Topography images of this figure can be related to those of Fig. 5.3. From

the profile of the topography images, it can be estimated that the nanostructures grow onto

the cotinuous thin film (zone I) with a lateral size of ~ 25-30 nm, the square pyramids of

zone III have a ~ 100 nm base and the rectangular ones a 50x250 nm2 base. Nanostructures of

zone V in Fig. 5.6 seem not to be isolated as it was expected from the SEM images. As it was

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

133

Figu

re 5

.5.

Sch

emat

ic t

heo

reti

cal

cro

ss-s

ecti

on

vie

ws

of

the

thre

e m

od

es

of

thin

film

gro

wth

. Ea

ch m

od

e is

sh

ow

n f

or

con

secu

tive

per

iod

s o

f ti

me

and

su

bse

qu

ent

surf

ace

cove

rage

[8

, 9].

t=t 1

sub

stra

tesu

bst

rate

sub

stra

te

t=t 2

>t 1

sub

stra

tesu

bst

rate

sub

stra

te

t=t 3

>t 2

>t 1

sub

stra

tesu

bst

rate

sub

stra

te

a) F

ran

k-va

n d

er

Me

rwe

mo

de

b) S

tran

ski-

Kra

stan

ov

mo

de

c) V

olm

er-

We

bb

er m

od

e

134 5.2. Nanoscale PbTiO3 structures onto commercial as-served SrTiO3 substrates prepared by

microemulsion mediated synthesis

already explained in section 2.4.5, lateral resolution in tapping mode is compromised when the

measured structures have the same size or are smaller than the probe. Therefore, in this case,

these nanostructures are observed as non-isolated nanostructures of ~37 nm of lateral size,

even if SEM images provided more accurate data.

A lower magnification image of the same of zone III of Fig. 5.6 is displayed in Fig. 5.7. Long

truncated rectangular based pyramids (Fig. 5.7(c)) coexist with squared based ones (Fig. 5.7

(b)). The first ones are in the micron range (~1 µm long, ~150 nm wide), while the last ones can

be classified as nanostructures of lateral size of ~150 nm. Their height is similar, despite the

shape or the orientation as can be confirmed by the color scale. These nanostructures onto the

substrate follow the [100] and [010] directions, setting at the surface minimum energy areas,

that is to say, the terraces edges.

These results indicate an important role of the surface of the substrate on the shape and

positioning of the nanostructures depending on the configuration of the terraces of the SrTiO3

substrate surface.

As seen in Fig. 5.6, nanostructures grown onto these substrates try to be placed in the

minimum energy locations of the substrate surface. Thus, they try to follow the terraces of the

substrate. By annealing the samples at high temperatures (850°C and 1050°C), it was observed

how the nanostructures settle at the terraces, as very small isolated particles at the edge (Fig.

5.8 (b) and (d)) or as pyramids with facets parallels to the terraces and placed at their corners

(Fig. 5.8 (a),(c) and (e)), depending on the amount of material (and subsequent relative

thickness) of the initial coating. The morphology of the surface of the substrate plays a major

role in the arrangement of the nanostructures. However, re-crystallization at high temperature

modifies the morphology of the nanostructures, as it provides additional energy and promotes

merging of the already present nanostructures

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

135

Figu

re 5

.6.

AFM

to

po

grap

hy

imag

es

of

thre

e d

iffe

ren

t zo

nes

of

the

sam

ple

pre

par

ed f

rom

th

e 5

·10

-3 M

mic

ella

r so

luti

on

. Th

e p

rofi

les

dis

pla

yed

bel

ow

eac

h i

mag

e

corr

esp

on

d t

o t

he

blu

e lin

es d

epic

ted

in t

he

AFM

imag

es

abo

ve t

hem

.

III

IIIIV

V

17

.43

nm

0.0

0 n

m

25

02

00

15

01

00

50

0

5 4 3 2 1 0

X[n

m]

Z[nm]

Com

plet

edth

infi

lm

Nan

ostr

uctu

res

27

.49

nm

0.0

0 n

m

Squa

repy

ram

id

Rec

tang

ular

pyr

amid

11

.13

nm

0.0

0 n

m

200nm

Na

no

stru

ctu

res

VIII

I

136 5.2. Nanoscale PbTiO3 structures onto commercial as-served SrTiO3 substrates prepared by

microemulsion mediated synthesis

Figu

re 5

.7.

Low

mag

nif

icat

ion

AFM

to

po

grap

hy

imag

e o

f zo

ne

III-

b o

f Fi

g. 5

.3 (

a) a

nd

tw

o r

epre

sen

tati

ve z

oo

m a

reas

wh

ere

sq

uar

e tr

un

cate

d p

yram

ids

(b)

and

rect

angu

lar

on

es (

c) c

an b

e fo

un

d a

nd

th

eir

pro

file

s al

on

g th

e b

lue

lines

mar

ked

in t

he

AFM

imag

es.

44

.45

nm

0.0

0 n

m

1.6

1.4

1.2

10

.80

.60

.40

.20

25

20

15

10 5 0

X[µ

m]

Z[nm]

17

.62

nm

0.0

0 n

m

25

02

00

15

01

00

50

0

16

14

12

10 8 6 4 2 0

X[n

m]

Z[nm]

71

.19

nm

0.0

0 n

m

tran

sver

sal

lon

gitu

din

al

b)

c)

a)

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

137

Figu

re 5

.8.

Sam

ple

s p

rep

ared

fro

m t

he

10

-2 M

mic

ella

r so

luti

on

at

65

0°C

an

d r

e-c

ryst

alliz

ed a

t 8

50

°C (

a-b

) an

d 1

05

0°C

(c-

e).

Imag

e (b

) sh

ow

s a

hig

her

mag

nif

icat

ion

imag

e o

f (a

). Im

ages

(d

) an

d (

e) a

re h

igh

er m

agn

ific

atio

n im

ages

of

(c)

and

are

pre

sen

ted

her

e to

sh

ow

det

ails

of

the

mer

ged

nan

ost

ruct

ure

s an

d t

he

mo

dif

icat

ion

of

the

sub

stra

te w

ith

tem

per

atu

re.

5.9

8 n

m

-4.9

9 n

m

3.6

3 n

m

-3.0

8 n

m

a)b

)

26

.54

nm

-6.8

5 n

m

5.6

5 n

m

-3.0

1 n

m

200nm

42

.06

nm

-6.5

0 n

m

c)d)

e)

138 5.2. Nanoscale PbTiO3 structures onto commercial as-served SrTiO3 substrates prepared by

microemulsion mediated synthesis

5.2.2. Structural characterization

Fig. 5.9 shows the experimental 2-dimensional synchrotron X-ray diffraction pattern of one the

samples prepared in this section.

The pattern consists of sharp short arcs and some diffuse spots. The most representative spots

are zoomed as inset in Fig. 5.9.

Integration of this experimental 2-D diffraction pattern is displayed in Fig. 5.10 as a black line.

The simulated PbTiO3 and SrTiO3 crystal perovskite phases are also presented in red and blue,

respectively. The most intense peaks correspond to the (100) SrTiO3 single crystal substrate,

having also contributions of the PbTiO3 perovskite phase.

Fig. 5.11 presents the simulated patterns of the SrTiO3 substrate and the PbTiO3

nanostructures in their perovskite structure. Note that not all the simulated peaks of the

PbTiO3 are present at the real pattern due to the low relative intensities of such peaks. A

certain cone of directions of angular width of ±3⁰ had been used for the simulations. The

lattice parameters calculated for the PbTiO3 from the simulations were a = b = 3.896(6)Å and

c = 4.149(4) Å, almost the theoretical ones (a = b = 3.899(9)Å and c = 4.140(0) Å).

The experimental Debye sharp ring sections agree well with the diffraction maxima of SrTiO3.

Measured SrTiO3 ring sections correspond to crystal orientation characterized by (001) poles

approximately normal to the sample surface. The diffuse spots of Fig. 5.9 always appear on the

low angle side of the SrTiO3 sharp peaks. This indicates a crystal phase with larger lattice

parameters than SrTiO3. These maxima fit to PbTiO3 perovskite reflections. The orientation

similarity and close cell parameters of the substrate and the PbTiO3 phase indicate that the

PbTiO3 nanostructures have grown epitaxially onto the SrTiO3 substrate. This epitaxial growth

explain partially why the lattice parameters are closer to the theoretical ones than the ones

obtained onto Pt/TiO2/SiO2/(100)Si substrates.

Besides, from the semiquantitative analysis of the broadening of the peaks assigned to PbTiO3,

it is deduced that these structures have nanometric sizes, as previously observed from the

topographic image of Fig. 5.3 (zone V), and that they are under strained conditions, which also

agrees with results reported for Pb(Zr0.42Ti0.48)O3 perovskite structures onto SrTiO3 substrates

[10, 11].

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

139

Figu

re 5

.9. E

xper

imen

tal 2

-D s

ynch

rotr

on

x-r

ay d

iffr

acti

on

pat

tern

of

a sa

mp

le p

rep

ared

fro

m t

he

5·1

0-3

M m

icel

lar

solu

tio

n o

nto

as-

rece

ived

STO

su

bst

ate.

140 5.2. Nanoscale PbTiO3 structures onto commercial as-served SrTiO3 substrates prepared by

microemulsion mediated synthesis

Figu

re 5

.10

. 2θ

dif

frac

tio

n p

atte

rn c

alcu

late

d f

rom

th

e in

tegr

atio

n o

f th

e 2

-D e

xper

imen

tal p

atte

rn o

f Fi

g. 5

.9 (

bla

ck s

olid

lin

e) a

nd

sim

ula

ted

dif

frac

tio

n p

atte

rns

of

the

Pb

TiO

3 p

ero

vski

te n

ano

stru

ctu

res

(red

lin

e) a

nd

SrT

iO3

sub

stra

te(b

lue

on

e).

Intensity(counts)

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

141

Figu

re 5

.11

. Si

mu

late

d 2

-D d

iffr

acti

on

pat

tern

of

sin

gle

crys

tal S

rTiO

3 w

ith

(0

01

) fi

ber

te

xtu

re (

a) a

nd

Pb

TiO

3 n

ano

stru

ctu

res

wit

h (

00

1)

text

ure

(b

) .

The

exp

erim

enta

l 2-D

dif

frac

tio

n p

atte

rns

wit

h r

efle

ctio

ns

corr

esp

on

din

g to

th

e Sr

TiO

3 an

d t

he

Pb

TiO

3 p

ero

vski

te p

has

es a

re s

ho

wn

in (

c) a

nd

(d

), r

esp

ect

ivel

y.

a)b

)

c)d

)

(11

0)

(11

0)

(11

1)

(20

0)

(21

1)

(21

1)

(22

0)

(22

0)

(11

0)

(11

1) (2

00

)

(21

1)

(21

1)

(22

0)

(22

0)

(11

0)

(11

0)

(11

0)

(11

1)

(20

0)

(21

1)

(21

1)

(22

0)

(22

0)

(11

0)

(11

1) (2

00

)

(21

1)

(21

1)

(22

0)

(22

0)

(11

0)

(20

0)

(00

1)

(10

1)

(11

0)

(20

0)(1

11

)

(11

2)

(21

1)(2

02

)

(22

0)

(10

1) (1

10

)(2

00

)

(11

1)

(11

2)

(21

1)

(20

2)

(22

0)

(20

0)

(20

0)

(00

1)

(10

1)

(11

0)

(20

0)(1

11

)

(11

2)

(21

1)(2

02

)

(22

0)

(10

1) (1

10

)(2

00

)

(11

1)

(11

2)

(21

1)

(20

2)

(22

0)

(20

0)

142 5.2. Nanoscale PbTiO3 structures onto commercial as-served SrTiO3 substrates prepared by

microemulsion mediated synthesis

Therefore, these analyses show that the nanostructures prepared by the microemulsion aided

sol–gel method onto as-received STO (100) substrates have a PbTiO3 perovskite structure with

a (100) preferred orientation.

5.2.3. Functional characterization

In order to prove the ferroelectric character of these nanostructures, measurements were

carried out on them by PFM. For this, new samples were deposited onto as-received

conductive Nb-doped (100)SrTiO3 substrates from the 5·10-3 M micellar solution.

Local piezoelectric and phase hysteresis loops were measured in the regions of the coating

where the self-arranged isolated nanostructures are observed (see zone V, Fig. 5.3). Also, the

piezoelectric activity of the larger particles with pyramid morphology was evaluated.

Fig. 5.12 shows the out-of-field piezoresponse hysteresis loops for an AC voltage at the free

resonance of the cantilever corresponding to the isolated nanostructures -(a) and (b)- and the

truncated pyramids -(c) and (d)-.

Figure 5.12. Out-of-field local hysteresis loops of isolated nanostructures of ~40 nm of lateral size

(phase (a) and amplitude (b) loops) and truncated squared based pyramid of ~400 nm of lateral size

(phase (c) and amplitude (d) loops).

In both cases, the hysteresis loops are symmetric with respect to the voltage axis and show a

high bias. This high bias is related to a certain imprint of the nanostructure [12], a well known

b)

d)c)

a)

d3

3ef

f(a

.u.)

d3

3ef

f(a

.u.)

Voltage (V) Voltage (V)

Voltage (V) Voltage (V)

Phas

e(⁰

)P

ha

se(⁰

)

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

143

effect in thin films [13] that was also observed in this thesis, in the case of nanostructures

prepared by using the microstructural instability phenomenon (Fig. 3.14) and not on those

prepared by the microemulsion mediated synthesis (Fig. 4.18 and Fig. 4.38).

Both pairs of hysteresis loops are symmetric with respect to the voltage axis, in contrast to the

experiments in section 4.2.3 and 4.3.3 where nanostructures prepared onto

Pt/TiO2/SiO2/(100)Si substrates were non symmetric (Fig. 4.38) or become non symmetric after

the first loop (Fig. 4.18 (c)). These effects were attributed to a pinned volume that is unable to

switch but has a certain polarization. The nature of this volume remains unknown, but is most

probably related to the defects –dislocations- observed in the TEM cross-section images of

section 4.2.1, which movement is clamped by free charges from the surface that are injected

towards the nanostructure. Here, this phenomenon is not observed, indicating that either

there is no defect in the bulk nanostructure or the free charges do not penetrate inside the

nanostructure and, consequently, they do not clamp a domain.

However, measurements on samples prepared onto Nb-doped STO substrates present

problems because, after the thermal treatment of crystallization of the nanostructures, the

substrate becomes less conductor. This destroys the capacitor formed by the bottom

substrate, the PbTiO3 nanostructure and the conductive probe, making difficult the

measurement of the piezoelectric hysteresis loops, most probably due to the migration of the

Nb atoms that increases the resistivity of the substrate. Therefore, it is difficult to extract

useful information from the functional characterization here shown and to compare these

results with the ones from previous chapters, onto Pt-coated (100)Si substrates.

5.3. Nanoscale PbTiO3 structures onto (100)SrTiO3 substrates

with controlled surfaces prepared by microemulsion mediated

synthesis.

As it was observed in previous section 5.2.1, the surface of the substrate plays a key role in the

growth of the nanostructures: it influences the type of growth, the adhesion to the substrate,

the epitaxy and the arrangement of the nanostructures. Therefore, a change in the physico-

chemical properties of the substrate surface will result in a different coating of the substrate

by the solution and, consequently, in a different shape and arrangement of the nanostructures.

This was exploited, for instance, in the work by Gilber et al. [14], who reported the possibility

of preparing, by CSD, self-organized heteroepitaxial CeO2 nanodots by controlling the surface

of LaAlO3 single crystal substrate.

144 5.3. Nanoscale PbTiO3 structures onto SrTiO3 substrates with controlled surfaces prepared by

microemulsion mediated synthesis

Concerning the role of the substrate surface in the PbTiO3/SrTiO3(100) system, Fig. 5.6 showed

that nanostructures preferentially grow at the edges of the STO terraces. Also, Fig. 5.8

illustrated the modification of the surface morphology when the substrate was annealed at

high temperatures. Therefore, it seems that the morphology of the substrate surface actually

influences the arrangement of the deposited nanostructures. However, this is not the only

factor to take into account to grow ferroelectric PbTiO3 nanostructures. In addition, in order to

improve the epitaxy between the SrTiO3(100) single crystal surface and the PbTiO3

nanostructures, the substrate should be finished in TiO2, which is the common layer between

both perovskite structures. In this way, epitaxial growth between nanostructures and

substrate would be enhanced, a less defective structure achieved (because of a uniquely

defined interface) and, consequently, adhesion improved.

As it was said before, the SrTiO3 perovskite structure consists of alternating layers of non-polar

SrO and TiO2, so that the terminating layer of a (100) crystal is not uniquely defined. Since the

SrO and TiO2 terminations present very different surface energies, SrO being more stable than

TiO2 for a wider range of chemical potentials (see Fig. 5.13), one would expect to find a

preferential SrO termination on nominally ideal STO surfaces (e.g., in absence of atmospheric

contaminants). In real surfaces, however, the ratio of the SrO and TiO2 terminated areas is

found to generally depend on surface preparation conditions.

Figure 5.13. Surface energy for TiO2 and SrO termination as a function of the TiO2 chemical potential

(the bulk reference is set to rutile and the zero chemical potential corresponds to the TiO2 rich

condition) [15].

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

145

An additional key point is the fact that re-crystallization at high temperatures (850 °C, 1050°C)

changes the morphology of the nanostructures and does not lead to nanostructures of

controlled shape and size, and uniformly distributed onto the substrate. Controlling such

characteristics is a subject of major importance, since the most probable application of these

ferroelectric nanostructures was pointed out to be their use as NV-FeRAM, Therefore, in the

next sections the surfaces of as-received (100) SrTiO3 substrates will be modified by different

treatments, studying the effect of these modifications on nanostructure growth and

positioning.

5.3.1 Preparing ideal surfaces: chemical and thermal treatments of the

STO surfaces.

The ideal STO (100) surface for heteroepitaxy purposes should be clean, smooth, well terraced

and have a unique type of terminating layer. Different treatments have been previously

reported in the literature for improving the surfaces of SrTiO3 substrates and achieve such

features.

Initially, Kawasaki et al. [16] proposed a mechanism of preparation of atomically smooth

surfaces of STO terminated by a TiO2 layer. Their procedure consists on selectively dissolving

the SrO surface planes, using the fact that the nature of both oxides is different: SrO is a basic

oxide while TiO2 is an acidic one. Thus, controlling the pH of the wet etching solution (Buffered

HydroFluoric acid (BHF)) would result into a substrate finished by the desired oxide layer. Their

chemical treatment consisted on soaking the as received polished substrates in the etching

solution (BHF, pH 4.4-4.6, 10 min), followed by a rinse in pure water and ethanol; finally,

substrates were dried in a nitrogen stream. Based on AFM, low energy ion scattering

spectroscopy (ISS) and reflection high energy electron diffraction (RHEED) measurements, the

authors claimed that the surface of these BHF-treated substrates (without any thermal

treatment) exhibited atomically flat terraces with o.4 nm high steps, and that the terminating

layer was TiO2 with a coverage factor of 100%.

Koster et al. [17] proposed to soak the substrates in a more basic BHF solution (pH= 5.5) and

for less time (30 s), preceding this treatment with a ultrasonically soaking of the as-received

substrate in water, in order to selectively promote the hydroxylation of the topmost SrO layer,

to make it more soluble in the acidic solution used for the etching of the substrate surface. It is

a well known fact that SrO reacts with CO2 and H2O from the atmosphere at room temperature

to form stable Sr compounds (like Sr(OH)2 and SrCO3). At the same time, TiO2 is unlikely to

react with water, due to its high stability. Koster et al. proposed to ultrasonically soak the as-

146 5.3. Nanoscale PbTiO3 structures onto SrTiO3 substrates with controlled surfaces prepared by

microemulsion mediated synthesis

received substrate in water in order to enhance the formation of Sr(OH)2 and thus, the etch-

selectivity of SrO relative to TiO2. To remove possible remnants and facilitate the

recrystallization of step ledges, a final thermal treatment (in a tube oven with flowing O2 ) was

carried out at 950 °C for 1 h, raising the temperature at a rate of about 0.5 °C/s, and cooling

down slowly (in 3 h) to room temperature. Based on AFM, LFM , RHEED and X-ray

photoelectron spectroscopy (XPS) measurements, these authors [17] claim that their

preparation method improves surface morphology (by reducing deep etch pits) and the

reproducibility of the results (by reducing the critical dependence on etching time and pH),

leading to atomically flat -and relatively clean- surfaces with straight terrace ledges and a

nearly perfect single (TiO2) termination.

Others authors [18-22] proposed slights variations of the above procedures and revealed that

a more complex scenario actually accompanies the preparation of such ideal STO surfaces.

Some of these works emphasized the convenience of adding thermal treatments to the

chemical etching proposed in ref. [15], in order to get better results, e.g., to promote ledge

recrystallization and achieve a well terraced morphology with straight steps [18, 21, 22]. To

reduce the pitchs created by the etching, a previous annealing stage it has been suggested [22],

while a repetition of the etching and heating stages it has been proposed to elliminate (or at

least reduce) the Sr-species that are observed to segregate to the surface during post-etching

thermal treatments [18]. Several works [18, 20] pointed out the importance of using specific

structural and compositional techniques to characterize the resulting surfaces, since only

conventional AFM measurements in air may not be able to detect the presence (on the

surface) of small segregated Sr-compounds or of islands half a unit cell high. Finally, a detailed

study by grazing incidence X-ray diffraction (GIXD) has recently found [20] that substrates

prepared by the procedure reported in ref. [17] (or small variations) have a rather high

TiO2/SrO surface termination ratio, but not a 100% TiO2 termination: results suggested that

about 75% of the surface is terminated by a TiO2 layer and about 25% by a SrO layer.

Inspired by these two recipes [16, 17], different processing strategies were carried out here

trying to achieve appropriate STO surfaces terminated by TiO2 planes; they consisted in a

chemical treatment of the commercial (polished) 10x10x1 mm3 SrTiO3(100) substrates,

followed by a thermal treatment. The processed substrates were then studied by AFM, AES

and LEED techniques to analyze the topography, composition and crystal structure of the

resulting surfaces.

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

147

As far as the chemical treatment is concerned, two kinds of experimental procedures were

performed. In the first one (etching-I), substrates were soaked for 20 seconds in a BHF etching

solution of pH 4.3; i.e., within the range reported [16] as adequate to avoid extensive surface

damage. In the second one (etching-II), substrates were ultrasonically soaked in deionized

water for 10 minutes before soaking them in a BHF etching solution of pH = 5.56 for 30 s. In

both procedures, substrates were rinsed in pure (Milli-Q) water after etching, and dried in a N2

stream.

Thermal treatments were carried out either in ultra high vacuum (UHV) or in air, in both cases

without oxygen gas flow. Details on the respective experimental conditions can be found in

section. Substrates were annealed up to temperatures high enough to allow recrystallization

processes. In particular they were heated above 800 ⁰C, the reported threshold for regrowth

of the step edges [17, 22]. Extremely low cooling rates were selected in all cases, in order to

promote the steady state [17, 23] and avoid the formation of undesired droplet-like features

on the STO surfaces [23].

Figure 2.4 in Chapter 2 schematically describes the two experimental procedures followed in

UHV. In the first one (thermal-A), a slow heating rate of ~0.1 ⁰C /s was used up to reach 815 ±

5 ⁰C on the substrate surface; this temperature was maintained for 3600 s, and the cooling

rate subsequently applied was even lower (0.03 ⁰C/s). In the second case (thermal-B) values

are slightly higher: a heating rate of ~0.2 ⁰C/s was applied until reaching a surface temperature

of 842 ± 7 ⁰C, which was maintained for 7200 s; cooling down to room temperature was also

performed with a very low rate ( ~0.09 ⁰C/s).

In the case of the thermal treatment in air (thermal-C), the STO substrate was annealed much

faster (heating rate of ~30 ⁰C/s) by using a rapid thermal processor (RTP). The annealing

temperature, kept for 3600 s, was also significantly higher (1050 ⁰C). Nevertheless, for cooling

down to room temperature, a very low rate was selected: ~0.1 ⁰C/s.

Table 5.1 summarizes the different STO (100) substrates processed here, together with their

respective treatments. It can be seen that, besides the samples chemically etched, there is

another one (labelled “substrate 1”) simply cleaned by Ar+ ion bombardment (60 min, 3 µA,

0.6-1 kV) and annealing in UHV (thermal-A conditions). This susbtrate was prepared in order to

have a clean surface reference for the Auger signals and LEED patterns, because some of the

procedures here investigated have not been previously analyzed in the literature (or at least

148 5.3. Nanoscale PbTiO3 structures onto SrTiO3 substrates with controlled surfaces prepared by

microemulsion mediated synthesis

not by AES and LEED), while the preparation of STO surfaces by Ar+ bombardment and

annealing in UHV is well documented [16, 24-26].

Table 5.1. Summary of the chemical etching and thermal treatments carried out here in order to control

the surface of the STO substrates. Note that samples soaked in BHF with pH 5.5 have been previously

soaked in deionized water for 10 min. For the sake of comparison, a substrate annealed in UHV after

Ar+ bombardment (instead of chemically etched) is also included

Substrate Deionized water

BHF Thermal treatment

Substrate 1 (Ar

+ bombarded and

annealed in UHV) No No

Thermal-A UHV, 815 ± 5 ⁰C for 3600 s

slow heating and cooling rates

Substrate 2 No pH 4.3, 20 s Thermal-B

UHV, 842 ± 7 ⁰C for 7200 s slow heating and cooling rates

Substrate 3 Yes pH 5.5, 30 s

No

Substrate 4 Yes pH 5.5, 30 s Thermal-A

UHV, 815 ± 5 ⁰C for 3600 s slow heating and cooling rates

Substrate 5 Yes pH 5.5, 30 s Thermal-C

In air, RTP at 1050ºC for 3600 s Fast heating & slow cooling (rates)

The LEED and AES results obtained for the differently processed STO substrates are

summarized in Figs. 5.14 and 5.15 and Table 5.2. As explained before (Chapter 2), such

measurements have been performed in the same UHV chamber (base pressure < 2x10-10 Torr)

used for thermal treatments in vacuum (thermal-A and thermal-B). This has permitted us to

record the LEED and AES data of substrates 1, 2 and 4 immediately after each thermal

treatment and without breaking vacuum. In contrast, substrates 3 and 5 have been exposed to

atmosphere (surfaces becoming thus contaminated) before being characterized by AES and

LEED. Substrate 5 was probably the most affected by the exposure to atmospheric

contamination, because a detailed AFM study was carried out before the AES/LEED

measurements (taken two weeks after the thermal treatment, keeping the substrate in air

meanwhile).

Before discussing surface composition and order, a short comment on sample conductivity will

be introduced, since AES and LEED measurements also provided information on this property,

as judged by the existence and strength of charging effects. “As received” substrates (mirror

polished samples without further treatments) are known to exhibit important charging effects,

and so was observed here. It is also well known that heat treatments in vacuum create oxygen

vacancies in the STO samples, which, as a consequence, become n-type doped [24, 27-30]. This

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

149

doping causes a darkening of the substrates and substantially increases sample conductivity,

thus facilitating surface analysis by STM or electron spectroscopy and diffraction techniques.

Both effects have been observed here for substrates 1, 2 and 4: after UHV heating, sample

colour changed from light yellow to dark grey, and no charging was detected during AES or

LEED measurements. In contrast, substrate 5 (annealed in air) exhibited some charging effects

(less than “as received” substrates) and did not become dark. The most remarkable result was

found on substrate 3: this “as etched” surface (without further thermal treatments) shows a

significant reduction of charging effects, suggesting surface conductivities not only better than

those of “as received” samples, but even better than that of substrate 5, (i.e., better than

samples annealed in air after etching).

Illustrative LEED patterns of all the substrates are presented in Fig. 5.14. Since they have been

recorded for different energies and sample azimuths, in order to facilitate data analysis, the

energy of the incident electron beam is indicated in each case, identifying also a few selected

diffraction spots. Patterns displayed on the left column correspond to substrate 1, while those

on the central and right columns belong to chemically etched substrates; among the later,

those depicted on the top and central raws correspond to substrates annealed at high

temperatures.

The surface of substrate 1 exhibits excellent 1x1- SrTiO3 (100) LEED patterns (Fig. 5.14 (a), (c),

(f)). This surface thus keeps the 1x1 bulk structure without additional reconstructions, in

agreement with previous reports for similar substrate treatments (Ar+ bombardment and

annealing in UHV) [24-26]. Note the low background and very sharp and brilliant spots of these

patterns, which indicate a well defined 2D nature and global ordering (i.e. good crystal quality)

of the surface region, suggesting also surface cleanness.

150 5.3. Nanoscale PbTiO3 structures onto SrTiO3 substrates with controlled surfaces prepared by

microemulsion mediated synthesis

Figu

re 5

.14

. LE

ED p

atte

rns

of

the

surf

ace

of

sub

stra

te 1

-b

lue

bo

rder

: (a

), (

c) a

nd

(f)

-, s

ub

stra

te 2

(b

), s

ub

stra

te 3

(d

), s

ub

stra

te 4

(e)

an

d s

ub

stra

te 5

(g)

. P

atte

rn w

ith

gree

n b

ord

er is

fro

m t

he

sam

ple

pre

par

ed u

sin

g th

e p

H 4

.5 B

HF

etch

ing

solu

tio

n a

nd

th

ose

wit

h r

ed b

ord

er a

re f

rom

th

e su

bst

rate

s so

aked

in t

he

pH

5.5

BH

F so

luti

on

.

Each

ro

w o

f fi

gure

s sh

ow

LEE

D p

atte

rns

carr

ied

ou

t fo

r ap

pro

xim

atel

y th

e sa

me

inci

den

t en

ergy

.

60

eV(0

,1)

(1,0

)

(0,1

)7

6.2

eV

60

eV

(0,1

)

(1,-

1)

90

.8eV

16

0.7

eV(0

,2)

(2,0

)

(0,2

)

16

2eV

(0,2

)

(2,0

)

91

eV

(0,2

)

18

6eV

(0,1

)

(3,0

)

152

eV

(0,2

)

(a)

(b)

(c)

(d)

(e)

(f)

(g)

(h)

(i)

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

151

Such cleanness is further confirmed by the Auger spectra of substrate 1 (see black line in Fig.

5.15), at least within the detection limit of the instrument (i.e., AES signals from possible

contaminants, like carbon, if exist, are below 0.1 atomic monolayer). On the other hand, peaks

from the Sr(MMN), Ti(LMM) and O(KLL) Auger transitions are clearly visible (energy regions of

50-130eV, 350-430 eV and 470-520 eV, respectively), with line-shapes and energies

characteristic of STO surfaces [25, 29, 31]. The Sr(104 eV), Ti(382 eV) and O(510 eV) peaks

(marked by arrows in Fig. 5.15) have been selected for the surface composition analysis, their

relative Sr-to-Ti and O-to-Ti intensity ratios being shown in Table 5.2. It should be noted that

the O-to-Ti intensity ratios measured for substrate 1 (Table 5.2) are found to be in agreement

with literature results for similar substrate treatments [25]. Hence, substrate 1 can be

confidently used as a reference surface for the LEED and AES analysis of STO samples.

Figure 5.15. AES spectra of the surface of substrate 1 (black line), substrate 4(blue line) and substrate 5 (red line), showing the Sr(MMN), Ti(LMM) and O(KLL) Auger transitions in the derivative mode. Arrows mark the peaks considered in the intensity ratio analysis of Table 5.2.

Table 5.2 summarizes the relative intensities of the Sr and O peaks with respect to the Ti, as

well as the observed LEED reconstruction (if any) for the as-received substrates and chemically

etched and/or thermally treated.

50 100 150 200 250 300 350 400 450 500

20000

25000

30000

35000

Inte

nsity

Energy (eV)

Sr peaksTi peaks

Oxygen

50 100 150 200 250 300 350 400 450 500

25000

30000

35000

40000

Inte

nsity

Energy (eV)

50 100 150 200 250 300 350 400 450 500

25000

30000

35000

40000

50 100 150 200 250 300 350 400 450 500

30000

35000

40000

50 100 150 200 250 300 350 400 450 500

20000

25000

30000

35000

Inte

nsity

Energy (eV)

Sr peaksTi peaks

Oxygen

50 100 150 200 250 300 350 400 450 500

20000

25000

30000

35000

Inte

nsity

Energy (eV)

Sr peaksTi peaks

Oxygen

Sr

Ti

O

Substrate 1

Substrate 4

Substrate 5

152 5.3. Nanoscale PbTiO3 structures onto SrTiO3 substrates with controlled surfaces prepared by

microemulsion mediated synthesis

Table 5.2. Auger intensity ratios obtained for the Sr(104 eV), Ti(382 eV) and O(510 eV) peaks, as measured on different STO (100) substrates: “as received” and prepared by the procedures detailed in Table 5.1 The respective surface structures observed by LEED are also indicated.

SrTiO3 substrate LEED IO/ITi ISr/ITi

As-received - 2.5 ± 0.5 1.4 ± 0.15

Substrate 1 1x1 2.7 ± 0.25 1.7 ± 0.2

Substrate 2 c(2x2) 4.1 ± 0.4 5.4 ± 0.5

Substrate 3 (1x1) 2.8 ± 0.2 2.9 ± 0.2

Substrate 4 c(2x2) 3.3 ± 0.3 4.4 ± 0.3

Substrate 5 (1x1) 2.2 ± 0.3 1.2 ± 0.3

Similar LEED patterns are observed for substrate 2 (Fig. 5.14 (b)) and substrate 4 (Fig. 5.14 (c),

(e)), which not only indicate good crystal order in the top layers, but also reveal the presence

of a superstructure. These two substrates (chemically etched and annealed in UHV) exhibit

patterns with the 1x1 diffraction spots characteristic of the bulk structure, and additional

(fractional) spots forming what we think is a centered c(2x2) superstructure (also known as

√2x√2-R245⁰). Note that the surface symmetry observed in both substrates is the same,

independently of whether “etch-I” or “etch-II” procedures have been employed. Their

respective Auger spectra show surfaces free of carbon contamination (see e.g., blue line in Fig.

5. 15, corresponding to substrate 4). Hence, the c(2x2) superstructure is probably due to a

surface reconstruction produced by the rearrangement of oxygen vacancies in the top layers.

In fact, several superstructures (twinned (2x1), c(6x2), c(4x2), (2x2), (√5x√5-R26.6⁰), … ) have

been observed by LEED, RHEED, GIXD, STM or TEM techniques on clean STO (100) surfaces

after diverse thermal treatments, being interpreted as related to different arrangements and

densities of the oxygen vacancies in the surface region [20, 21, 26-30, 32-35]. Noticeably, this

is the first time (to our knowledge) that a c(2x2) superstructure is reported for clean SrTiO3

(100) surfaces, but we are not aware either of structural studies of STO (100) surfaces

prepared in the same conditions as substrates 2 and 4 ( i.e., by chemical etching with a BHF

solution followed by high-T annealing in UHV with very slow thermal ramps and no oxygen

flow). Perhaps the closest procedures are those employed in refs. [20, 21], for which 1x1

and/or 2x2 surface lattices have been reported (depending on the annealing conditions).

The analysis of the Auger signals from substrates 2 and 4 lead to somehow unexpected results,

in the sense that their O-to-Ti and Sr-to-Ti intensity ratios are remarkably high (See Table 5.2),

and do not support the achievement of TiO2 terminated surfaces; on the contrary, the AES

results indicate Ti-deficient surface regions. In particular, substrate 2 exhibits the lowest Ti

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

153

signals and the highest O-to-Ti and Sr-to-Ti Auger ratios of all the samples analyzed here (Table

5.2), and therefore, it presents the most Ti-deficient surface. Such a result might be related to

the too acid pH (pH 4.5) of the etching solution, but in any case, it undoubtedly shows that the

treatment used for substrate 2 has not improved the surface as one would expect from ref.

[21].

For substrate 3, the Auger data indicate a higher Ti-content of the surface layers in comparison

to substrates 2 and 4 (see Table 5.2). Thus, it seems that “as etched” surfaces (just prepared by

the etch-II protocol, without any thermal treatment) are richer in Ti than those subsequently

annealed in UHV. Nevertheless, the Sr signal of substrate 3 is still too high to correspond to

ideal TiO2 -terminated surfaces: note that, while the O-to-Ti intensity ratio of substrate 3 is

rather similar to that measured on substrate 1, the Sr-to-Ti ratio is clearly higher (the Ti-

content thus lower) than those obtained for substrate1 and “as received” samples.

One should also remark the clear improvement of certain surface features (like cleanliness,

conductivity and crystalline order) observed for substrate 3 respect to “as received” samples.

The Auger signal from carbon in substrate 3 is much lower than that of “as received”

substrates, and --as judged by the peak energy and lineshape-- it does not seem to be due to

carbonate species formed by reaction with Sr-oxide from the STO surface [17] (like happens in

“as received” substrates), but to the adsorption of C-compounds during a short exposure to

atmospheric pressure (before the AES/LEED analysis). Such improvement in cleanliness and

conductivity has permitted us to record LEED patterns on substrate 3 (Fig. 5.14 (h) (i)), while no

pattern at all could be observed on “as received” samples. Moreover, taking into account that

the surface of substrate 3 was not strictly clean when it was examined by LEED, the relatively

good quality of the patterns (showing sharp, and in some cases even brilliant, spots) indicate a

rather good crystal ordering in the surface region, which is noticeable for “as etched”

substrates already at this stage (without thermal treatments). Only 1x1 diffraction spots are

detected for substrate 3, thus suggesting that no surface reconstructions are present. A 1x1

LEED pattern has also been reported for surfaces prepared by chemical etching with BHF

solutions before thermal treatments (or after soft annealings) [21].

The LEED pattern observed for substrate 5 (see Fig. 5.14 (f)) is rather similar to that of

substrate 3, although with a higher background and not so brilliant diffraction spots. Both

features can be simply related to the stronger charging effects and the higher C contamination

present in substrate 5; such C contamination (confirmed by AES) is probably just the result of a

longer period of exposure to atmosphere (between substrate processing and the AES or LEED

154 5.3. Nanoscale PbTiO3 structures onto SrTiO3 substrates with controlled surfaces prepared by

microemulsion mediated synthesis

measurements). Taking into account these circunstamces, the observed diffraction pattern can

be interpreted as indicative of a reasonably good crystalline order in the surface region. Only

1x1 diffraction spots are detected, which suggest the absence of reconstructions in this surface

(prepared by the etching-II procedure and subsequently annealed in air (RTP, thermal-C)).

Note that a 1x1 structure would be in agreement with GIXD results recently reported for

surfaces prepared by similar chemical and thermal treatments [20].

The AES spectrum of substrate 5 is displayed in Fig. 5.15 (blue line). The highest Ti content of

all the surfaces here analyzed seems to correspond to this sample, for which the lowest values

of the Sr-to-Ti and O-to-Ti Auger intensity ratios are achieved (see Table 5.2). Moreover, the

occupancy fraction of the TiO2 termination in substrate 5 may be in the 80% to 100% range,

since the O/Ti and Sr/Ti Auger intensity ratios measured for this substrate are even lower than

those obtained for “as received” (polished) samples, and a 75% to 95% TiO2 termination (or a

5% to 25% SrO termination) has been reported for commercial as-polished substrates [16, 21].

Hence, one of the main conclusions of this combined LEED-AES analysis is that, the preparation

method used for substrate 5 permits to produce rather clean and ordered surfaces mostly

terminated by TiO2 (occupancy fractions in the 80% to 100% range) .

A representative AFM topography image of the surface of the substrate 2 is shown in Fig. 5.16.

Figure 5.16. AFM topography image (a) and profile (b) along the blue line in the topography image of substrate 2.

Square etch pits can be found throughout the whole surface of the substrate. They are ~1.0-

1.5 µm of lateral size and 100-300 nm high. Also, island like residues of ~150-200 nm of lateral

size and ~ 10-20 nm high are observed. Kawasaki et al. [16] reported the presence of etch pits

when using a more acidic BHF solutions (pH<4) than the one used here (pH 4.5) and island like

293.77 nm

0.00 nm

43.532.521.510.50

120

100

80

60

40

20

0

X[µm]

Z[n

m]

a) b)

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

155

residues, of lower size than the ones here observed (0.2-0.4 nm high) when using a more basic

etching solution (pH>5).

Fig. 5.17 shows the AFM topography image of the surface of the substrate with the less Ti

deficient surface (substrate 5). It has well defined terraces and some etch pits. These pits are

no more than 6 nm height and ~100 nm wide. No particle residues are observed in the image.

Figure 5.17. AFM topography image of substrate 5 (a) and the profile of an etch pit (b).

Fig. 5.18 displays a before and after the chemical and thermal treatment pictures of the

surface of substrate 5 at the same scale. Improvement of the substrate surface was achieved

by getting well defined terraces that are less Ti deficient, as deduced from the AES spectrum.

Figure 5.18 AFM topography images of an as-served SrTiO3 substrate (a) and substrate 5 (b). Note that

the measured area is 1x1 µm2 in both images.

In order to prove that the surfaces of the substrate 5 are atomically flat after the chemical

etching and thermal treatment discussed above, the steps between terraces were measured.

2.86 nm

0.00 nm

150100500

6

5

4

3

2

1

0

X[nm]

Z[n

m]

a) b)

a) b)

1.41.210.80.60.40.20

1.2

1

0.8

0.6

0.4

0.2

X[µm]

Z[n

m]

c)

156 5.3. Nanoscale PbTiO3 structures onto SrTiO3 substrates with controlled surfaces prepared by

microemulsion mediated synthesis

Images were taken in the 12 x 12µm2 area of Fig. 5.17 (a) and the steps measured. The results

are shown in Fig. 5.19. 52% of the steps are two unit cells high, 26% are one and 22% are three

unit cells, considering the theoretical unit cell of STO perovskite phase is cubic of parameters

a = b = c = 3.900(5) Å. Therefore, it can be stated that the surface is atomically flat with steps

between the terraces of few multiples of a unit cell.

Figure 5.19. Distribution of the height of the steps of the SrTiO3 terraces, measured on the Fig. 5.17 (a).

As a result, the ultrasonically soaking in deionized water joined to the chemical etching in the

BHF etching solution of pH5.5 for 20 s and RTP annealing at 1050°C in air for 3600 s and very

slow cooling rates, provide STO substrates with controlled surfaces. The PbTiO3 nanostructures

will be prepared on these atomically flat surfaces with minimum defects, to study how

microemulsion aided sol-gel deposition method proposed in this thesis, determines the

ordering, size and properties of the resulting PbTiO3 nanostructures.

5.3.2. Microscopy and quantitative microstructure analysis.

Fig. 5.20 displays an optical micrograph of the resulting coating deposited onto the STO

substrate which surface topography has been shown in Fig. 5.17 (a), i.e. substrate 5. No stream

rings is observed and it indicates a uniform coating of the substrate.

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

157

Figure 5.20. Optical micrograph of the sample prepared onto a chemically etched and thermally treated substrate: i.e. substrate 5 from previous section.

A study of the topography of the sample along one of the diagonals of the substrate surface

was carried out by AFM and the results are presented in Fig. 5.21. The morphology of the

nanostructures deposited onto the substrate varies slightly from one point to another,

showing a large uniformity in shape and arrangement. At the images taken at the zones 1 and

5, larger and merged nanostructures are found. This is due to the well-known border effect of

the spin-coating deposition [36].

Fig. 5.22 (a) represents the same image as in area 3 of Fig. 5.21 and the corresponding image

when the substrate is subtracted (b). It is clear how the bright spots that correspond to the

grown primary nanostructures are aligned along the [010] and [100] directions of the substrate.

Fig. 5.22 (c) shows a profile of the three nanostructures marked by a blue line in Fig. 5.22 (b).

These nanostructures are representative of the whole image, having similar size and shape

(~35 nm and 7 nm of lateral dimension and height, respectively).

The Fast Fourier Transform (FFT) of Fig. 5.22 (a) is represented in Fig. 5.22 (d). Two bright lines,

are observed, indicating that the nanostructures settle, preferentially, along the [100] and

[010] directions of the single crystal STO substrate, thus demonstrating a long-range order of

the nanostructures onto the substrate. These directions are established with respect to the

facets of the single crystal.

500 nm

158 5.3. Nanoscale PbTiO3 structures onto SrTiO3 substrates with controlled surfaces prepared by

microemulsion mediated synthesis

Figu

re 5

.21

. A

FM t

op

ogr

aph

y im

age

s o

f th

e cr

ysta

llin

e P

bTi

O3

nan

ost

ruct

ure

s o

nto

th

e su

bst

rate

5 a

t d

iffe

ren

t lo

cati

on

s al

on

g o

ne

of

the

dia

gon

als

of

the

10

x10

mm

2

sub

stra

te a

fter

ch

em

ical

etc

hin

g an

d c

ryst

alliz

atio

n o

f th

e su

bst

rate

. 2

3

1

45

17

.63

nm

0.0

0 n

m

2

27

.35

nm

0.0

0 n

m

4

18

.83

nm

0.0

0 n

m

3

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

159

Figu

re 5

.22

. A

FM t

op

ogr

aph

y im

age

(a),

th

e co

rres

po

nd

ing

imag

e w

hen

th

e su

bst

rate

is s

ub

trac

ted

(b

) ,t

he

pro

file

(c)

alo

ng

the

blu

e lin

e in

(b

) an

d t

he

FFT

of

imag

e (a

)

for

the

crys

talli

ne

Pb

TiO

3 n

ano

stru

ctu

res

pre

par

ed o

nto

su

bst

rate

5.

35

03

00

25

02

00

15

01

00

50

0

8 7 6 5 4 3 2 1 0

X[n

m]

Z[nm]

0

10

.05

nm

0.0

00

10

.47

nm

1.8

9

a) c)

d)b)

160 5.3. Nanoscale PbTiO3 structures onto SrTiO3 substrates with controlled surfaces prepared by

microemulsion mediated synthesis

The size distribution of the nanostructures onto the substrate was calculated from Fig. 5.22 (a)

and is represented in Fig. 5.23. The distribution is Gaussian and its probabilistic line can be

expressed as:

y = -3.10 + 0.08·x R = 0.99 (5.2)

Figure 5.23 Equivalent diameter distributions of the nanostructures prepared onto a chemically and

thermally treated substrate 5.

The average equivalent diameter of the nanostructures is 36 nm with a standard deviation of

12 nm. This size is of the same order than the one obtained for the primary nanostructures of

section 4.3.3 (Fig. 4.30), that was 21 nm with a standard deviation of 4 nm. Also, their rounded

shape seems to be similar. The height of the primary nanostructures of previous chapter 4 is

higher than the one of these nanostructures, but the average volume can be estimated as

constant and equal to ~3750 nm3, considering an average height of 7nm for the

nanostructures prepared in this section and 14 nm for those of previous chapter and a

spherical cap shape. The nanostructures height values can be easily obtained from the scales

of the images.

The different lateral size and height but equal volume indicate different surface energies of the

substrate surfaces, which are 92 erg/cm2 in the case of the (111) Pt [37] and is higher for the

(001) SrTiO3 substrate, either SrO or TiO2 terminated, as seen in the energy versus chemical

potential graph in Fig. 5.13.

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

161

All considered, self-arrangement of the primary nanostructures onto the SrTiO3 depends

strongly on the surface energy of the substrate. When the substrate surface is forced to

terminate in TiO2 stacks and it is atomically flat, self-arrangement of the PbTiO3 nanostructures

is obtained and the microemulsion is uniformly deposited onto the substrate. This results into

primary nanostructures formed from the building units contained in the micellar solution and

self-arrangement of the crystalline nanostructures onto the substrate along the [100] and

[010] directions of the single crystal.

To the best of the knowledge of this author, a long-range order of ferroelectric PbTiO3

nanostructures onto a substrate has not been reported before by using a bottom-up

preparation technique. This is important for the fabrication of arrays of nanostructures to be

used as ferroelectric memories (Fe-NVRAM) with a low cost bottom-up technique.

162 Remarks

Remarks

1. When using the microemulsion mediated synthesis onto as-received commercial

(100)SrTiO3 single crystals, to promote epitaxial growth and less defective

nanostructures of PbTiO3, different morphologies of structures at different locations of

the coating are found due to the deficient wetting of the substrate by the micellar

solution. This different morphology depends on the relative height of the coating and,

consequently, on the amount of material to be crystallized. Non-isolated

nanostructures of ~37 nm of lateral size were found in certain areas, whereas square

and rectangular pyramids up to 250 nm of lateral size were found in some other areas.

2. According to sinchrotron X-ray experiments, these nanostructures have the perovskite

structure expected for PbTiO3, with lattice parameters a = b = 3.896(6)Å and c = 4.149

(4) Å, very close to the theoretical ones of a = b = 3.899(9)Å and c = 4.140 (0) Å

3. When deposited onto a conductive single-crystal substrate such as an as-received Nb+

doped SrTiO3, it was possible to measure the ferro-piezoelectric behavior. However,

contrarily to what was expected, this behavior does not differ significantly from the

one observed for the previously studied nanostructures onto polycrystalline

substrates.

4. For controlling the surface of the (100)SrTiO3 substrates, different chemical etchings

and thermal treatments were used. All of them lead to a modification of the surfaces.

5. Substrates treated with a pH 4.5 BHF solution and UHV annealing at 815°C with slow

heating and cooling rates show a c(2x2) surface reconstruction and Ti-deficient

surfaces. Better results (crystallinity and less Ti-deficiency) are achieved for substrates

soaked in a more basic BHF solution (pH 5.5). Substrates treated with this chemical

etching (before any annealing) display reasonably good results concerning surface

ordering, cleanness, and conductivity, as seen in the LEED patterns.

6. Substrates ultrasonically soaked in water, in a BHF etching solution of pH 5.5 and RTP

annealed in air at 1050°C with a very slow cooling rate show surfaces with the highest

Ti content of the present study, which may correspond to 80%-100% TiO2 termination.

Besides, atomically flat surfaces with terraces of height of a few cell steps between

them.

7. When TiO2 terminated and atomically flat (001)SrTiO3 single crystal substrate surfaces

are used, the PbTiO3 nanostructures grow uniformly along all the surface (10x10mm2).

Round particles of ~36 nm of lateral size and ~7nm of height are periodically

Chapter 5: Nanostructures onto SrTiO3 single-crystal substrates by microemulsion mediated synthesis

163

distributed along the [100] and [010] directions of the substrate, as demonstrated by

the Fast Fourier Transform of the AFM images.

8. These samples with an array of PbTiO3 nanostructures obtained by a bottom-up

method are, in principle, promising for their use in ferroelectric ultra-high density

memories.

164 Bibliography

Bibliography

[1] I. Szafraniak, C. Harnagea, R. Schloz, S. Bhattacharyya, D. Hesse and M. Alexe,

"Ferroelectric epitaxial nanocrystals obtained by a self-patterning method", Applied Physics

Letters, 83 (11), 2003, p:2211

[2] M. Alexe and D. Hesse, "Self-assembled nanoscale ferroelectrics", Journal of Material

Research, 41, 2006, p:1

[3] M. Dawber, K.M. Rabe and J.F. Scott, "Physics of thin-film ferroelectric oxides",

Reviews of modern physics, 77, 2005, p:1083

[4] M. Dawber, C. Lichtensteiger, M. Cantoni, M. Veithen, P. Ghosez, K. Johnston, K.M.

Rabe and J.M. Triscone, "Unusual behavior of the ferroelectric polarization in PbTiO3/SrTiO3

superlattices", Physical Review Letters, 95 (17), 2005, p:177601

[5] C.B. Eom, R.B. Vandover, J.M. Phillips, D.J. Werder, J.H. Marshall, C.H. Chen, R.J. Cava,

R.M. Fleming and D.K. Fork, "Fabrication and properties of epitaxial ferroelectric

heterostructures with (SrRuO3) isotropic metallic oxide electrodes", Applied Physics Letters, 63

(18), 1993, p:2570

[6] I. Vrejoiu, G. Le Rhun, N.D. Zakharov, D. Hesse, L. Pintilie and M. Alexe, "Threading

dislocations in epitaxial ferroelectric PbZr0.2Ti0.8O3 films and their effect on polarization

backswitching", Philosophical Magazine, 86 (28), 2006, p:4477

[7] B.L. Carvalho and E.L. Thomas, "Morphology of steps in terraced block-copolymer

films", Physical Review Letters, 73 (24), 1994, p:3321

[8] A. Milchev, Electrocrystallization. Fundamentals of nucleation and growth. 1st ed.

2002: Springer. 280.

[9] J. Sun, P. Jina, Z.G. Wanga, H.Z. Zhangb, Z.Y. Wangb and L.Z. Hu, "Changing planar thin

film growth into self-assembled island formation by adjusting experimental conditions", Thin

Solid Films, 476, 2005, p:68

[10] M.-W. Chu, I. Szafraniak, R. Scholz, C. Harnagea, D. Hesse, M. Alexe and U. Gösele,

"Impact of misfit dislocations on the polarization instability of epitaxial nanostructured

ferroelectric perovskites", Nature Materials, 3, 2004, p:87

[11] I. Szafraniak, S. Bhattacharyya, C. Harnagea, R. Scholz and M. Alexe, "Self-assembled

ferroelectric nanostructures", Integrated Ferroelectrics, 68, 2004, p:279

[12] B.J. Rodriguez, S. Jesse, M. Alexe and S.V. Kalinin, "Spatially Resolved Mapping of

Polarization Switching Behavior in Nanoscale Ferroelectrics", Advanced materials, 20 (15),

2008, p:102

Bibliography 165

[13] H.N. AlShareef, D. Dimos, W.L. Warren and B.A. Tuttle, "Voltage offsets and imprint

mechanism in SrBi2Ta2O9 thin films", Journal of Applied Physics, 80 (8), 1996, p:4573

[14] M. Gibert, T. Puig, X. Obradors, A. Benedetti, F. Sandiumenge and R. Huhne, "Self-

organization of heteroepitaxial CeO2 nanodots grown from chemical solutions", Advanced

materials, 19 (22), 2007, p:3937

[15] Y. Liang and A.A. Demkov, Interfacial properties of epitaxial oxide/semiconductor

systems, in Materials fundamentals of gate dielectrics, A.A. Demkov and A. Navrotsky, Editors.

2005, Springer: Berlin.

[16] M. Kawasaki, K. Takahashi, T. Maeda, R. Tsuchiya, M. Shinohara, O. Ishiyama, T.

Yonezawa, M. Yoshimoto and H. Koinuma, "Atomic control of the SrTiO3 crystal-surface",

Science, 266 (5190), 1994, p:1540

[17] G. Koster, B.L. Kropman, G.J.H.M. Rijnders, D.H.A. Blank and H. Rogalla, "Quasi-ideal

strontium titanate crystal surfaces through formation of strontium hydroxide", Applied Physics

Letters, 73 (20), 1998, p:2920

[18] T. Ohnishi, K. Shibuya, M. Lippmaa, D. Kobayashi, H. Kumigashira, M. Oshima and H.

Koinuma, "Preparation of thermally stable TiO2-terminated SrTiO3 (100) substrate surfaces",

Applied Physics Letters, 85 (2), 2004, p:272

[19] H.B. Moon, J.H. Cho and J.S. Ahn. "Nanoscale topographic evolutions of SrTiO3 (001)

surfaces". in 14th Symposium on Dielectric and Advanced Matter Physics/6th World on High-

Dielectric and Ferroclectric Devices/Materials. 2005. Muju, SOUTH KOREA: Korean Physical Soc.

[20] A. Fragneto, G.M. De Luca, R. Di Capua, S.d.U. U., M. Salluzzo, X. Torrelles, T.-L. Lee and

J. Zegenhagen, "Ti- and Sr-rich surfaces of SrTiO3 studied by grazing incidence x-ray diffraction",

Applied Physics Letters, 91, 2007, p:101910

[21] M. Kawasaki, A. Ohtomo, T. Arakane, K. Takahashi, M. Yoshimoto and H. Koinuma,

"Atomic control of the SrTiO3 surface for perfect epitaxy of perovskite oxides", Applied Surface

Science, 107, 1996, p:102

[22] G. Koster, B.L. Kropman, G.J.H.M. Rijnders, D.H.A. Blank and H. Rogalla, "Influence of

the surface treatment on the homoepitaxial growth of SrTiO3", Materials Science and

Engineering, B56, 1998, p:209

[23] K. Szot, W. Speier, U. Breuer, R. Meyer, S. J. and R. Waser, "Formation of micro-crystals

on the (100) surface of SrTiO3 at elevated temperatures", Surface Science, 460, 2000, p:112

[24] N. Bickel, G. Schmidt, H. K. and K. Müller, "Ferroelectric relaxation of the SrTiO3 (100)

surface", Physical Review Letters, 62, 1989, p:2009

[25] Y.-W. Chung and W.B. Weissbard, "Surface spectroscopy studies of the SrTiO3 (100)

surface and the platinum-SrTiO3 (100) interface", Physical Review B, 20, 1979, p:3496

166 Bibliography

[26] B. Cord and R. Courths, "Electronic study of SrTiO3(001) surfaces by photoemission",

Surface Science, 62 (1-3), 1985, p:34

[27] Q. Jiang and J. Zegenhagen, "SrTiO3 (001)-c(6x2): a long range, atomically ordered

surface stable in oxygen and ambient air", Surface Science, 367, 1996, p:L42

[28] M.S. Martín González, M.H. Aguirre, E. Morán, M.A. Alario-Franco, V. Pérez-Dieste, J.

Avila and M.C. Asensio, "In situ reduction of (100) SrTiO3", Solid State Sciences, 2, 2000, p:519

[29] P.J. Moller, K. S.A. and E.F. Lazneva, "elective growth of a MgO(100)-c(2x2)

superstructure on a SrTiO3 (100)-(2x2) substrate", Surface Science, 425, 1999, p:15

[30] M. Naito and H. Sato, "Reflection high-energy electron diffraction study on the SrTiO3

surface structure", Physica C, 389, 1994, p:1

[31] J. Brunen and J. Zegenhagen, "Investigation of the SrTiO3 (110) surface by means of

LEED, scanning tunneling microscopy and Auger spectroscopy", Surface Science, 389, 1997,

p:349

[32] N. Erdman and L.D. Marks, "SrTiO3 (001) surface structures under oxidizing conditions",

Surface Science, 526, 2003, p:107

[33] Q. Jiang and J. Zegenhagen, "c(6x2) and c(4x2) reconstruction of SrTiO3 (001)", Surface

Science, 425, 1999, p:343

[34] T. Matsumoto, H. Tanaka, T. Kawai and S. Kawai, "STM-imaging of a SrTiO3 (100)

surface with atomic scale resolution", Surface Science Letters, 278, 1992, p:L153

[35] H. Tanaka, T. Matsumoto, T. Kawai and S. Kawai, "Surface structure and electronic

property of reduced SrTiO3 (100) surface observed by Scanning Tunneling Microscopy/

Spectroscopy", Japanese Journal of Applied Physics, 32, 1993, p:1405

[36] C.J. Brinker and G.W. Scherer, Sol-gel science. The physics and chemistry of sol-gel

processing. 1990, San Diego: Academic Press.

[37] T.K. Galeev, N.N. Bulgakov, G.A. Savelieva and N.M. Popova, "Surface properties of

platinum and palladium", Reaction Kinetics and Catalysis Letters, 14, 1980, p:61

CHAPTER 6: CONCLUSIONS

6.1. Conclusions.

The main conclusions obtained from the experimental results presented and discussed in this

Ph.D Thesis are summarized as follows:

1. Ferroelectric nanostructures by the phenomenon of the microstructural instability of

polycrystalline ultrathin films

Prior to the presentation of the results of the novel method proposed in this thesis for the

preparation of ferroelectric PbTiO3 nanostructures, a known procedure to obtain

nanostructures onto substrates is analyzed. This method is based in the microstructural

instability of solution derived ultrathin films when their thickness is beyond a critical one. The

critical concentration of the deposited sol below which this phenomenon does not control the

growing process of the nanostructures was determined. When the phenomenon applies, the

mechanisms of growth involve coalescence and diffusion among neighbor particles, as deduced

from the size distributions. The average lateral size of the resulting nanostructures is in the

range of 50 nm derived from the 4·10-2

M sol. No self-arrangement of the nanostructures is

observed by this procedure. By synchrotron radiation grazing-incidence experiments, it was

possible to determine the crystal structure –PbTiO3 perovskite-, the orientation of the

crystallites - a fiber one with the (100) axis perpendicular to the surface of the substrate and a

direction distribution cone of ±15°-, and the cell parameters –c = 4.130(1) Å, a = 3.912(1) Å of

the nanostructures. The nanostructures show ferro-piezoelectric response.

2. Nanoscale PbTiO3 ferroelectric structures onto Pt/TiO2/SiO2/(100)Si substrates prepared by

microemulsion mediated synthesis.

Microemulsion mediated synthesis is proposed as a novel procedure for the preparation of

ferroelectric oxides onto substrates. The hipothesis here considered is that the micelles act as

building units, isolating the nanostructures and yielding, after drying and thermal treatment of

the micellar solution coatings, self-assembled PbTiO3 nanostructures onto the substrates. These

nanostructures present controlled size and shape that do not depend on the concentration of

the solution. The average size of the crystalline nanostructures obtained by this procedure is

~75 nm for the ones prepared from the 10-2

M micellar solution and ~67 nm for those prepared

from the 5·10-3

M one. Only the number of nanostructures on the substrate does change.

Those nanostructures are not ordered which is considered an effect of the substrate defects

and inhomogeneity. Their size distributions are Gaussian, in contrast with the lognormal ones

168 6.1. Conclusions

found for the nanoparticles from the microstructural instability. This means that they grow

independently from each other from coalescence of a small number of primary nanoparticles.

This fact evidences the hipothesis of the micelles acting as building units of the nanoparticles.

The ferro-piezoelectric character is proved by the PFM images and hysteresis loops. Mechanism

of switching in these nanostructures is complex, involving 90° switching of a domains and the

creation of a pinned layer, stabilized by the charges on the surface of the platinum bottom

electrode of the substrate that introduce themselves at the domain wall.

3. Nanoscale PbTiO3 ferroelectric structures onto microemulsion layer/Pt/TiO2/SiO2/(100)Si

substrates prepared by the modified microemulsion mediated synthesis.

A modification of the previous method, consisting in the functionalization of the surface of the

substrate by a microemulsion layer is proposed as a valid procedure to improve the wetting

and minimize the shrinkage of the micellar solution coating. Isolated primary nanostructures

with size distributions of average lateral size ~21 nm are obtained as observed on the AFM

images, as well as a reduced number of merged bigger ones. TEM results proved that isolated

nanostructures down to ~9 nm of lateral size can be observed. This is found all over the sample,

showing the uniformity and quality of the coating. Both the primary nanostructures and

merged bigger ones show Gaussians size distributions. The primary nanostructures prepared by

this late modified procedure shows hexagonal short-range order, that veryfies the self-assemby

capacity of the micelles. The crystal structure was determined by grazing incidence synchrotron

diffraction as a PbTiO3 perovskite with cell parameters a=b=3.890(0) Å and c=4.056(7) Å and

crystal axis rotated (-7°, -7°, ±20°) and (20°, -7°, -7°±20°). Ferro-piezoresponse is measured in

this work in one of the smallest nanostructures where this kind of measurements could be

carried out by PFM to date, which lateral size is ~37 nm and ~14 nm of height as determined by

AFM topography images. This nanostructure as well as the rest of them, seem to be single

domains onto a pinned layer.

4. Nanoscale PbTiO3 structures onto commercial as-served SrTiO3 substrates prepared by

microemulsion mediated synthesis.

Nanostructures prepared onto as-served SrTiO3 single crystals present a non-uniform coating,

which morphology depends on the relative height of the coating and is due to a deficient

wetting of the substrate by the micellar solution. At some locations, non-isolated

nanostructures of ~37 nm of lateral size were found. The crystal structure was analyzed by

diffraction of synchrotron radiation in grazing incidence configuration, obtaining a PbTiO3

perovskite lattice with cell parameters a = b = 3.896(6)Å and c = 4.149 (4) Å. It was possible to

measure ferro-piezoelectric response of the nanostructures, when deposited onto SrTiO3-Nb

Chaper 3: Conclusions 169

doped substrates, presenting a response close to the one observed for the nanostructures

deposited onto the polycrystalline Pt-(100) Si substrates.

5. Nanoscale PbTiO3 structures onto (100)SrTiO3 substrates with controlled surfaces prepared by

microemulsion mediated synthesis

In order to promote epitaxial growth and to improve the quality of the coating of the substrate

by the micellar layer, the surface of the substrate was modified. Different chemical and thermal

treatments were tested, establishing that thermal treatment in vacuum lead to a c(2x2)

reconstruction and Ti deficient surfaces. The best results were obtained for the substrates

ultrasonically soaked in water, in a BHF etching solution of pH 5.5 and RTP annealed in air at

1050°C with a very slow cooling rate, presenting surfaces with the highest Ti content of the

present study and atomically flat surfaces with terraces of height of a few cell steps between

them.

When TiO2 terminated and atomically flat (001)SrTiO3 single crystal substrate surfaces are used,

the PbTiO3 nanostructures grow uniformly along all the surface (10x10mm2). Particles of ~36

nm of lateral size and ~7nm of height are periodically distributed with a long-range order.

6. A novel bottom-up method based in the use of microemulsion mediated synthesis and

functionalizing/controlling the substrate surface has been developed in this Ph. D.Thesis,

proving its effectiveness for obtaining ferroelectric PbTiO3 nanostructures with a long-range

order onto the substrates. These materials are promising for their use in ferroelectric ultra-

high density memories.

6.2. Conclusiones.

1. Nanoestructuras ferroeléctricas preparadas utilizando el fenómeno de la inestabilidad

microstructural de láminas ultradelgadas.

Antes de presentar los resultados del método novedoso que se ha propuesto en esta tesis para

la preparación de nanoestructuras ferroeléctricas de PbTiO3, se analiza un método conocido

para la obtención de nanoestructuras sobre sustratos. Este método está basado en la

inestabilidad microestructural que presentan las láminas ultradelgadas derivadas de

disoluciones cuando su espesor está por debajo de uno crítico. Se determinó la concentración a

partir de la cuál este fenómeno ya no rige el proceso de crecimiento de las nanoestructuras.

Cuando este fenómeno tiene lugar, los mecanismos de crecimiento son los de coalescencia y

difusión entre partículas vecinas, como se deduce de de las distribuciones de tamaño. La

dimensión lateral promedio de las nanoestructuras resultantes del sol con concentración

170 6.2. Conclusiones

3·10-2

M está en el rango de los 50 nm. No se observa que las nanoestructuras se distribuyan

uniformemente por sí solas sobre el sustrato. Mediante los experimentos de difracción de

rayos-X de difracción sincrotrón en ángulo rasante, se pudo determinar la estructura cristalina

–perovskita de PbTiO3-, la orientación de los cristalitos –orientación de fibra con el eje (100)

perpendicular a la superficie del sustrato y un cono de distribución de direcciones de ±15°-, y

los parámetros de la celda unitaria –c = 4.130(1) Å, a = 3.912(1) Å de las nanoestructuras. Las

nanoestructuras presentan respuesta ferro-piezoeléctrica.

2. Estructuras nanométricas de PbTiO3 sobre sustratos de Pt/TiO2/SiO2/(100)Si preparadas

mediante síntesis asistida con microemulsión.

El depósito de soluciones micelares se propone como un procedimiento novedoso para la

preparación de óxidos ferroeléctricos sobre sustratos. La hipótesis considerada es la de que las

micelas actúan como “building units”, aislando las nanoestructuras y dando lugar, después del

secado y el tratamiento térmico del recubrimiento de la solución micelar, a las nanoestructuras

PbTiO3 auto-organizadas sobre sustratos. Estas nanoestructuras presentan forma y tamaño

controlados que no dependen de la concentración de la disolución. El tamaño promedio de las

nanoestructuras cristalinas obtenidas por este procedimiento es de ~75 nm para las preparadas

a partir de la solución micelar de concentración 10-2

M y ~67 nm para las preparadas a partir de

la de 5·10-3

M. Sólo el número de nanoestructuras sobre el sustrato varía con la concentración.

Estas nanoestructuras no están ordenadas, lo que se considera un resultado de los defectos e

inhomogeneidad del sustrato. Sus distribuciones de tamaño de grano son Gausinas, en

contraste con las lognormales de las nanoparticulas obtenidas mediante la inestabilidad

microestructural. Ello se interpreta como el crecimiento independiente, resultado de la

coalescencia de algunas nanopartículas primarias. Este hecho pone de manifiesto la veracidad

de la hipótesis de que las micelas actúan como unidades primarias de las que se deriva el

crecimiento de las nanopartículas. El carácter ferro-piezoeléctrico se comprobó mediante

imágenes y ciclos de histéresis obtenidos con el uso de la técnica de PFM. El mecanismo de

conmutación de estas nanoestructuras es complicado e involucra la conmutación de 90⁰ de

dominios a y la creación de una capa que no conmuta y que se estabiliza por las cargas

presentes en la superficie del electrodo inferior de platino del sustrato que se introducen en las

paredes de dominio.

Chaper 3: Conclusions 171

3. Estructuras nanométricas de PbTiO3 sobre sustratos de

película de microemulsión/Pt/TiO2/SiO2/(100)Si preparadas mediante síntesis asistida con

microemulsión modificada.

Con el objetivo de mejorar el mojado y minimizar la contracción del recubrimiento de la

solución micelar, se propone una modificación del método anterior, que consiste en

funcionalizar la superficie del sustrato mediante una capa micelar. Mediante AFM se

observaron nanoestructuras primarias aisladas con un tamaño promedio de ~21 nm, así como

un reducido número de estructuras mayores que han coalescido. Esto se puede encontrar a lo

largo de toda la muestra, demostrando la uniformidad y calidad del recubrimiento. Tanto las

nanoestructuras primarias como las coalescidas presentan distribuciones de tamaño Gausianas.

Las nanoestructuras primarias preparadas por este procedimiento muestran orden hexagonal

de corto alcance, lo que verifica la hipótesis de la capacidad de auto-ordenación de las miceleas.

La estructura cristalina se determinó mediante difracción de rayos-X de radiación sincrotrón en

ángulo rasante, encontrándose la estructura perovskita del PbTiO3 con parámetros de red de

a=b=3.890(0) Å and c=4.056(7) Å y una rotación de los ejes cristalinos de (-7°, -7°, ±20°) y (20°, -

7°, -7°±20°). En este trabajo se ha medido respuesta ferro-piezoeléctrica en una nanoestructura

muy pequeña (el menor tamaño publicado, según conocimiento de la autora), utilizando PFM.

Su dimensión lateral es de ~37 nm y tiene ~14 nm de altura. Esta nanoestructura, como en los

casos anteriores, parece consistir en un mono domonio sobre una capa que no conmuta

4. Estructuras nanométricas de PbTiO3 sobre sustratos comerciales de SrTiO3 preparadas

mediante síntesis depósito de soluciones micelares.

Las nanoestructuras preparadas sobre sustratos monocristalinos comerciales de SrTiO3

presentan un recubrimiento no uniforme, en el que la morfología depende de la altura relativa

de dicho recubrimiento y es debido a un mojado deficiente de la superficie del sustrato por la

solución micelar. En algún punto, es posible observar nanoestructuras aisladas de ~37 nm de

dimensión lateral. La estructura cristalina de las nanoestructuras se analizó mediante difracción

de rayos-X de radiación sincrotrón en ángulo rasante, obteniéndose que ésta se corresponde

con la estructura perovskita del PbTiO3 con parámetros de celda a = b = 3.896(6)Å and c = 4.149

(4) Å. En sustratos conductores de SrTiO3 dopados con Nb, fue posible medir la respuesta ferro-

piezoeléctrica de las nanoestructuras, presentando una respuesta similar a la observada en las

nanoestructuras depositadas sobre sustratos policristalinos de Pt-(100) Si.

172 6.2. Conclusiones

5. Estructuras nanométricas de PbTiO3 sobre sustratos con superficies controladas preparadas

mediante síntesis asistida con microemulsión.

Para promover el crecimiento epitaxial y mejorar la calidad del recubrimiento del sustrato por

la solución micelar, se modificó la superficie de los sustratos. Se probaron diferentes

combinaciones de tratamientos químicos y térmicos, estableciéndose que los tratamientos en

UHV daban lugar a una reconstrucción c(2x2) de la superficie que, además, era deficiente en Ti.

Los mejores resultados se obtuvieron para sustratos sumergidos en agua desionizada y tratados

mediante agitación en ultrasonido y, posteriormente, en una solución BHF tampón de pH 5.5 y

recocidos usando RTP en aire a 1050°C con una velocidad de enfriamiento muy lenta. Estas

superficies presentaban el mayor contenido en Ti de este estudio y eran atómicamente planas

con terrazas de altura algunas celdas unidad.

Cuando se utilizan los sustratos monocristalinos de (001)SrTiO3 con superficies atómicamente

planas y terminadas en TiO2, las nanoestructuras de PbTiO3 se disponen uniformemente a lo

largo de la superficie del sustrato (10x10mm2). Particulas de tamaño promedio ~36 nm de

dimensión lateral and ~7nm de altura se distribuyen de forma periódica con largo alcance.

6. En esta tesis doctoral, se ha desarrollado un método novedoso, de tipo “bottom-up” basado

en el depósito de soluciones micelares. Así mismo, se ha desarrollado un procedimiento de

funcionalización/control de la superficie de los sustratos. Se ha probado su efectividad para

obtener nanoestructuras ferroeléctricas de PbTiO3 con un orden de largo alcance sobre los

sustratos. Estos materiales son prometedores para ser utilizados en memorias ferroeléctricas

de alta densidad.

Part of this work has been presented in the following papers and

conferences:

Papers published in SCI journals:

M.L. Calzada, M. Torres, L.E. Fuentes-Cobas, A. Mehta, J. Ricote and L. Pardo, "Ferroelectric self-

assembled PbTiO3 perovskite nanostructures onto (100)SrTiO3 substrates from a novel

microemulsion aided sol-gel preparation method", Nanotechnology, 18 (37), 2007, p:375603.

L. Fuentes-Montero, M.E. Montero-Cabrera, L. Calzada, M.P. De la Rosa, O. Raymond, R. Font, M.

Garcia, A. Mehta, M. Torres and L. Fuentes. "Synchrotron Techniques Applied to Ferroelectrics:

Some Representative Cases", Integrated Ferroelectrics, 101, 2008, p:113.

M. Torres, J. Ricote, L. Pardo and M.L. Calzada. "Nanosize ferroelectric PbTiO3 structures onto

substrates. Preparation by a novel bottom-up method and nanoscopic characterisation",

Integrated Ferroelectrics, 99, 2008, p:95.

M.L. Calzada, M. Torres, J. Ricote and L. Pardo, "Ferroelectric PbTiO3 nanostructures onto Si-based

substrates with size and shape control", Journal of Nanoparticle Research, 11 (5), 2009, p:1227.

M. Torres, M. Alonso, M.L. Calzada and L. Pardo, “Influence of the substrate surface on the self-

assembly of ferroelectric PbTiO3 nanostructures obtained by microemulsion assisted Chemical

Solution Deposition”, Ferroelectrics, (in press).

M. Torres, M.L. Calzada, B. Rodriguez, M. Alexe, L. Pardo, “SPM studies of ferroelectric

nanostructures prepared by a microemulsion-assisted method onto substrates”, Processing and

Applications of Ceramics, (in press).

International Conferences

M.L. Calzada, M. Torres, A. García, J. Ricote and L. Pardo, “PbTiO3 nanostructures onto silicon

substrates fabricated by bottom-up technology”, in X International Conference on Electroceramics

(Electroceramics X)2006. Toledo (Spain). Poster.

M.L. Calzada, M. Torres, J. Ricote, L. Pardo, “PbTiO3 nanostructures onto substrates prepared by

microemulsion mediated synthesis”, in “COST 539 Workshop”. 2006. Bruxelles (Belgium). Oral

presentation.

L. Fuentes-Montero, M.E. Montero-Cabrera, L. Calzada, M.P. De la Rosa, O. Raymond, R. Font, M.

Garcia, A. Mehta, M. Torres and L. Fuentes. "Synchrotron Techniques Applied to Ferroelectrics:

Some Representative Cases". in Symposium on Ferroelectricity and Piezoelectricity held at the 15th

International Materials Research Congress (IMRC). 2006. Cancun (Mexico). Oral presentation.

M. Torres, J. Ricote, L. Pardo and M.L. Calzada. "Nanosize ferroelectric PbTiO3 structures onto

substrates. Preparation by a novel bottom-up method and nanoscopic characterisation". in 19th

International Symposium on Integrated Ferroelectrics. 2007. Bordeaux, (France). Oral Presentation.

M. Torres, M.L. Calzada, J. Ricote, L.E. Fuentes-Cobas, A. Mehta, L. Pardo, “Preparation by

microemulsion mediated synthesis of PbTiO3 nanostructures onto substrates”, in “MIND Internal

Workshop on films prepared by Chemical Solution Deposition”. 2007. Madrid (Spain).

M. Torres, M. Alonso, M.L. Calzada and L. Pardo, “Influence of the substrate surface on the self-

assembly of ferroelectric PbTiO3 nanostructures obtained by microemulsion assisted Chemical

Solution Deposition”, in 9th European Conference on Applications of Polar Dielectrics. 2008.Roma;

(Italy). Oral Presentation.

M.Torres, L.Pardo and M.L.Calzada, “Control of the PbTiO3 nanostructures size as a function of the

processing route by bottom-up CSD methods”, in XI International Conference on Electroceramics

(Electroceramics XI). 2008. Manchester (UK). Oral presentation.

M. Torres, M.L. Calzada, B. Rodriguez, M. Alexe, L. Pardo, “SPM studies of ferroelectric

nanostructures prepared by a microemulsion-assisted method onto substrates”, in “COST 539

Workshop “Advanced Functional Characterization of Nanostructured Materials”. 2009. Madrid

(Spain).