Superior Pinning Properties in Nano-Engineered YBa2Cu3O7−δ Giorgio Ercolano Department of Materials Science & Metallurgy Corpus Christi College This dissertation is submitted for the degree of Doctor of Philosophy 2011 Si qui forte mearum ineptiarum lectores eritis manusque uestras non horrebitis admouere nobis... Declaration I declare that, except where otherwise stated, this dissertation is the result of my own work and includes nothing which is the outcome of work done in collaboration. No part of this dissertation had been submitted at Cambridge or any other University for a degree, diploma or other qualification. The length including tables, footnotes, bibliography and appendices do not exceedes the limit of 60000 words. Giorgio Ercolano Cambridge April 2011 Acknowledgements This dissertation describes the research work on controlled nano- engineering of high temperature superconducting materials carried out at Cambridge University Materials Science and Metallurgy De- partment in the Device Materials Group between November 2007 and February 2011. I would like to express my gratitude to my supervisor, Prof. Judith MacManus-Driscoll, for the help and valuable advice provided during this project. I would also like to acknowledge... Dr Stuart Wimbush for the helpfull discussions on nanoinclusions flux pinning and on the vortex path model Dr John Durrell and Dr Marcus Weigand for the trainings and help on microfabrication and critical current measures Dr Sophie Harrington for the helpfull discussions on pulsed laser de- position of doped YBa2Cu3O7−δ thin films and target preparation Dr Marco Bianchetti for the practical help with the photolithography and critical currents characterisation of the thin films and for the usefull discussions Prof. Haiyan Wang, Joon Lee Hwan and Dr Lata Sahonta for the amazing TEM work Dr Xavier Moia and Dr Thomas Fix for the valuable discussions on structural data interpretation Dr Nadia Stelmashenko for the help with the atomic force microscope Mary Vickers for the help with the x-ray diffraction analysis and Dr Lara San-Emeterio-Alvarez, Dr David Muoz-Rojas, Diana Iza, Oon Jew Lee and Dr Sohini Kar for the support The work described in this dissertetion was supported by NESPA, Nano-Engineered Superconductors for Power Application, a frame- work of the Marie Curie Research Training Network, funded within the EUs 6◦ framework program. Superior Pinning Properties in Nano-Engineered YBa2Cu3O7−δ Giorgio Ercolano Large electrical current transport in the absence of energy losses is the key factor in commercial applications of high temperature supercon- ductors. This thesis demonstrates an easy and inexpensive bottom-up technique to produce self assembled nanorods, segmented nanorods as well as nanoparticles in YBa2Cu3O7−δ thin films grown by pulsed laser deposition. The structural and morphological characteristic of the pinning landscapes produced are investigated and correlated to their effects on the superconducting properties of the thin films. In particular two pinning landscapes are investigated: Ba2YNbO6 nanorods are grown in YBa2Cu3O7−δ thin films using a Ba2YNbO6 doped YBa2Cu3O7−δ pulsed laser deposition targets and Ba2(Y/Gd)(Nb/Ta)O6 segmented nanorods together with (Y/Gd)2O3 nanoparticles are grown in (Y/Gd)Ba2Cu3O7−δ thin films using a Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ pulsed laser deposition targets. The Ba2YNbO6 + YBa2Cu3O7−δ is deeply characterised and the ef- fects of the deposition parameters are analysed. Ba2YNbO6 is demon- strated to be an interesting novel pinning addition capable to increase the critical current and to reduce the YBa2Cu3O7−δ critical currents angular dependencies anisotropy. The Ba2YNbO6 + Gd3TaO7 + YBa2Cu3O7−δ is found to produce a new complex pinning landscape extremely effective. At high fields the synergetic combination of the different defects typology is shown to generate an interesting new feature in the critical current angular dependencies. Chapter 1 is an introduction to superconductivity, the fundamentals of the field are briefly presented. In chapter 2 the discussion in focused on pinning in high temperature superconductors. Cuprates and in par- ticular YBa2Cu3O7−δ are presented. The pinning phenomenon and the practical pinning engineering in thin films is also discussed in this chapter. Chapter 3 describes the thin films preparation methods and the characterisation techniques used in the research work. Chapter 4 and 5 are focused on the Ba2YNbO6 doped YBa2Cu3O7−δ thin films. Chapter 4 is an introduction to Ba2YNbO6 doped YBa2Cu3O7−δ, the preliminary results obtained on Ba2YNbO6 doped YBa2Cu3O7−δ thin films are shown in this chapter. The crystalline structure, the mor- phology and the superconducting properties of thin films deposited adopting different deposition parameters are analysed and discussed in chapter 5. In chapter 6 the new complex pinning landscape of Ba2(Y/Gd)(Nb/Ta)O6 and (Y/Gd)2O3 in (Y/Gd)Ba2Cu3O7−δ is pre- sented. Concluding remarks on the research described in the work ends the thesis in a brief final chapter 7. Contents Contents vii List of Figures xi 1 Introduction 1 1.1 The discovery of superconductivity . . . . . . . . . . . . . . . . . 1 1.2 The Meissner effect and the London penetration depth . . . . . . 2 1.3 Critical magnetic field, critical current density. Two types of su- perconductors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.4 Finally a microscopic theory of superconductivity . . . . . . . . . 7 1.5 The high temperature superconductors revolution . . . . . . . . . 7 2 Pinning in YBa2Cu3O7−δ thin films 9 2.1 Cuprates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 2.1.1 Weaklinks and the texturing of cuprates . . . . . . . . . . 10 2.2 YBa2Cu3O7−δ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.2.1 Crystal structure . . . . . . . . . . . . . . . . . . . . . . . 11 2.2.2 Oxygen deficiency . . . . . . . . . . . . . . . . . . . . . . . 12 2.3 Pinning . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 2.3.1 The dissipative phenomenon . . . . . . . . . . . . . . . . . 14 2.3.2 Dissipation-free current flow, the pinning force . . . . . . . 15 2.3.3 Defects as pinning centers . . . . . . . . . . . . . . . . . . 16 2.4 Practical Pinning Engineering . . . . . . . . . . . . . . . . . . . . 17 2.4.1 Non superconducting secondary phase addition . . . . . . 17 2.4.2 Alternative routes . . . . . . . . . . . . . . . . . . . . . . . 19 vii CONTENTS 3 Experimental Techniques 21 3.1 Sample Preparation . . . . . . . . . . . . . . . . . . . . . . . . . . 21 3.1.1 Pulsed Laser Deposition . . . . . . . . . . . . . . . . . . . 22 3.1.2 Target Preparation . . . . . . . . . . . . . . . . . . . . . . 25 3.1.3 Sample Patterning . . . . . . . . . . . . . . . . . . . . . . 27 3.2 Structural Morphological characterisation . . . . . . . . . . . . . . 29 3.2.1 Structural Analysis . . . . . . . . . . . . . . . . . . . . . . 29 3.2.2 Microscopy . . . . . . . . . . . . . . . . . . . . . . . . . . 35 3.3 Superconductivity Properties characterisation . . . . . . . . . . . 36 4 Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results 39 4.1 The Nb introduction in the YBa2Cu3O7−δ . . . . . . . . . . . . . 39 4.1.1 Nb doped ceramic samples . . . . . . . . . . . . . . . . . . 40 4.1.2 Ba2YNbO6 use in YBa2Cu3O7−δ thin film . . . . . . . . . 41 4.2 Ba2YNbO6 perovskite additions to YBa2Cu3O7−δ: the preliminary results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 4.2.1 The Ba2YNbO6 synthesis, the pulsed laser deposition tar- get preparation and thin films deposition . . . . . . . . . . 43 4.2.2 Crystalline structure analysis: x-ray diffraction data . . . . 45 4.2.2.1 Crystalline phases identification . . . . . . . . . . 46 4.2.2.2 Crystalline phases orientation . . . . . . . . . . . 47 4.2.2.3 Crystallographic matching and strain . . . . . . . 48 4.2.3 Nanostructure analysis: Atomic Force Microscopy and Trans- mission Electron Microscopy . . . . . . . . . . . . . . . . . 51 4.2.3.1 Surface topography (Atomic Force Microscopy) . 51 4.2.3.2 Cross-section transmission electron microscopy . 53 4.2.4 The superconducting properties: Tc, Jc(B) and Jc(B, θ) . . 58 4.2.4.1 Transition temperature, Tc . . . . . . . . . . . . . 59 4.2.4.2 The critical current density . . . . . . . . . . . . 60 4.2.5 Concluding remarks to the preliminary results . . . . . . . 63 5 Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters 65 viii CONTENTS 5.1 Ba2YNbO6 perovskite additions to YBa2Cu3O7−δ: the effects of the deposition rate on the nanostructure and the superconducting properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 5.1.1 Ba2YNbO6 doped YBa2Cu3O7−δ target preparation and thin film deposition . . . . . . . . . . . . . . . . . . . . . . 66 5.1.2 Crystalline structure analysis: x-ray diffraction data . . . . 67 5.1.2.1 Crystalline phases identification, phases orientation 67 5.1.2.2 Epitaxy quality: rocking curve . . . . . . . . . . 68 5.1.2.3 Epitaxy quality: reciprocal space maps . . . . . . 69 5.1.3 Nanostructure analysis: Atomic Force Microscopy and Trans- mission Electron Microscopy . . . . . . . . . . . . . . . . . 76 5.1.3.1 Surface topography (Atomic Force Microscopy) . 76 5.1.3.2 Cross-section transmission electron microscopy . 79 5.1.4 The superconducting properties: Tc, Jc(B) and Jc(B, θ) . . 82 5.1.4.1 Transition temperature, Tc . . . . . . . . . . . . . 82 5.1.4.2 The critical current density . . . . . . . . . . . . 84 5.1.5 Concluding remarks to the analysis of the effects of the de- position rate on the nanostructure and the superconducting properties . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 5.2 Ba2YNbO6 perovskite additions to YBa2Cu3O7−δ: deposition pa- rameter optimization . . . . . . . . . . . . . . . . . . . . . . . . . 91 5.2.1 Ba2YNbO6 doped YBa2Cu3O7−δ pulsed laser deposition target preparation and thin films deposition . . . . . . . . 91 5.2.2 Surface Topology: Grain Connectivity . . . . . . . . . . . 92 5.2.3 The superconducting properties: Tc, Jc(B) and Jc(B, θ) . . 98 5.2.3.1 Transition temperature, Tc . . . . . . . . . . . . . 98 5.2.3.2 The critical current density . . . . . . . . . . . . 100 5.2.4 Concluding remarks to the deposition parameters optimiza- tion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 6 Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ 104 6.1 Ba2YNbO6 and Gd3TaO7 doped YBa2Cu3O7−δ target preparation and thin films deposition . . . . . . . . . . . . . . . . . . . . . . . 105 ix CONTENTS 6.2 Crystalline structure analysis: x-ray diffraction data . . . . . . . . 106 6.2.0.1 Crystalline phases identification . . . . . . . . . . 106 6.2.1 Crystalline phases orientation . . . . . . . . . . . . . . . . 108 6.3 Cross-section transmission electron microscopy . . . . . . . . . . . 109 6.4 The superconducting properties: Tc, Jc(B) and Jc(B, θ) . . . . . . 112 6.4.1 The critical current density . . . . . . . . . . . . . . . . . 112 6.4.1.1 The critical current density: Jc(B) . . . . . . . . 113 6.4.1.2 The critical current density angular dependence: Jc(B, θ) . . . . . . . . . . . . . . . . . . . . . . . 115 6.4.1.3 A new pinning feature . . . . . . . . . . . . . . . 117 6.4.2 Concluding remarks to Ba2YNbO6 and Gd3TaO7 simulta- neous doping of YBa2Cu3O7−δ . . . . . . . . . . . . . . . . 117 7 Conclusions and further work 120 Appendix 123 References 146 x List of Figures 1.1 Differences in magnetisation behaviours of a perfect conductor and a superconductor, the Meissner effect. a) Perfect conductor cooled in zero field; b) Perfect conductor cooled in applied magnetic field; c) Superconductor cooled in zero field; d) Superconductor cooled in applied magnetic field. . . . . . . . . . . . . . . . . . . . . . . . 3 1.2 Structure of an isolated Abrikosov vortex [1] . . . . . . . . . . . . 6 1.3 Magnetisation curves for a type I and a type II superconductor [1] 6 2.1 a) Crystal Structure of YBa2Cu3O7−δ; b) CuO2 planes; c) CuO chains. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 2.2 a) Superconductive transition temperature Tc for ten samples of Ba2YCuOx with varying x values; b) Refined crystallographic cell parameters for Ba2YCuOx with varying x values; from [2]. . . . . 13 3.1 Schematic of a typical pulsed laser deposition system [3] . . . . . 23 3.2 The target sintering thermal process . . . . . . . . . . . . . . . . 26 3.3 Sketches of a sample at the different stages of a photolithographic process. a) As deposited sample; b) Sample after the first spin coating of the photoresist; c) Sample after the developing of the electrode mask; d) Sample after the silver layer and gold layer deposition. e) Sample after the lift-off process; f) Sample after the second spin caoting of the photoresist; g) Sample after the developing of the trak mask; h) Sample after the milling/cleaning process. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 3.4 Micrography of the surface of a patterned sample . . . . . . . . . 31 xi LIST OF FIGURES 3.5 Schematic of a diffractometer operated in the Bragg-Brentano Ge- ometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31 3.6 Schematic of a diffractometer geometry operating a φ scan . . . . 32 3.7 Schematic of a diffractometer geometry operating a rocking curve scan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 3.8 Schematic of a diffractometer geometry scanning a reciprocal space map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 3.9 Schematic of the geometry for the determination of Jc(B, θ) . . . 38 4.1 Powder diffraction patter for a) pure Ba2YNbO6, b) pure YBa2Cu3O7−δ, c) 1:1 mole mixture of YBa2Cu3O7−δ:Ba2YNbO6 heated at 950 ◦C for 15 hr. [4] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42 4.2 Powder diffraction patter for Ba2YNbO6powder produced. Adapted from [5]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 4.3 X-ray diffraction data for undoped YBa2Cu3O7−δ and 5%mol Ba2YNbO6 doped film deposited on SrTiO3. Adapted from [5]. . . . . . . . . 46 4.4 X-ray diffraction data from φ scans of (101) SrTiO3, (102) YBa2Cu3O7−δ, (202) Ba2YNbO6from a 5%mol Ba2YNbO6 doped film. Adapted from [5]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48 4.5 Crystallographic matching of YBa2Cu3O7−δ with Ba2YNbO6. a) b-axis view and b) c-axis view [5]. . . . . . . . . . . . . . . . . . . 49 4.6 Atomic Force Microscopy surface image of a pure YBa2Cu3O7−δ and a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films. a) 5 µm × 5 µm surface area of pure YBa2Cu3O7−δ; b) 1 µm × 1 µm surface area of pure YBa2Cu3O7−δ; c) 5 µm × 5 µm surface area of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ; d) 1 µm × 1 µm surface area of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ . . . . . . . . . . 52 xii LIST OF FIGURES 4.7 Transmission electron microscope cross-sectional image of a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film. White arrows in the di- rection of the c-axis mark the position of self assembled Ba2YNbO6 nanorods parallel to the c-axis. Larger white arrows parallel to the YBa2Cu3O7−δ ab-planes mark the position of nanoparticles and of the interface between the substrate and the thin film. Inset shows selected area electron diffraction pattern taken from a region in proximity of a Ba2YNbO6 inclusion showing diffraction spot of Ba2YNbO6 YBa2Cu3O7−δ and SrTiO3. (TEM images from Prof. H Wang research group at Texas A&M University). . . . . . . . . 54 4.8 Transmission electron microscope cross-sectional image of a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film. White arrows in the di- rection of the YBa2Cu3O7−δ c-axis mark the position of self assem- bled Ba2YNbO6 nanorods parallel to the c-axis. Arrows parallel to the YBa2Cu3O7−δ ab-planes mark the position of nanoparticles. The Ba2YNbO6 nanorods presence is indicated by the Moire fringes arising from lattice mismatch between Ba2YNbO6 and YBa2Cu3O7−δ. (TEM images from Prof. H Wang research group at Texas A&M University). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 4.9 Resistance variation with the temperature measured on a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film. The inset shows the region of the superconducting transition magnified. . . . . . . . . 59 4.10 Critical current density variation with the applied magnetic field value measured on a pure YBa2Cu3O7−δ and a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films at 77 K and field applied parallel to the c-axis of the YBa2Cu3O7−δ. The inset shows data on a logaritmic axis and the calculated α values. . . . . . . . . . . . . . 61 4.11 Critical current density variation with the direction of the applied magnetic field measured on a pure YBa2Cu3O7−δ and a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films at 77 K and magnetic flux density = 0.5 T . . . . . . . . . . . . . . . . . . . . . . . . . 62 xiii LIST OF FIGURES 5.1 X-ray diffraction reciprocal space map measured on a pure YBa2Cu3O7−δ thin film. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 5.2 X-ray diffraction reciprocal space map measured on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films. a) Film deposition rate of 1 Hz; b) Film deposition rate of 10 Hz. . . . . . . . . . . . . . . . . . . 72 5.3 Half maximum intesity insoline and peak position comparison for the (108)(018) YBa2Cu3O7−δ peaks from figure 5.1 and 5.2. . . . 75 5.4 Atomic Force Microscopy surface image (5 µm × 5 µm) of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a sub- strate temperature of 780 ◦C, repetition rate of 1 Hz and 10 Hz with a thickness of 700 nm and 400 nm. a) 1 Hz and 700 nm; b) 10 Hz and 700 nm; c) 1 Hz and 400 nm; d) 10 Hz and 400 nm. . . 77 5.5 Transmission electron microscope cross-sectional image of a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited with a laser pulse rates of 1 Hz (a, b) and 10 Hz (c, d). White arrows in the di- rection of the YBa2Cu3O7−δ c-axis mark the position of self assem- bled Ba2YNbO6 nanorods parallel to the c-axis. Arrows parallel to the YBa2Cu3O7−δ ab-planes mark the position of nanoparticles. (TEM images from Prof. H Wang research group at Texas A&M University). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 5.6 Resistance variation with the temperature measured on a pure YBa2Cu3O7−δ thin film deposited at a substrate temperature of 780 ◦C and laser pulses repetition rates of 1 Hz and on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a sub- strate temperature of 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 xiv LIST OF FIGURES 5.7 Critical current density variation with the applied magnetic field value measured on a pure YBa2Cu3O7−δ thin film deposited at a substrate temperature of 780 ◦C and laser pulses repetition rates of 1 Hz and on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz. The field is applied parallel to the c-axis of the YBa2Cu3O7−δ, T = 77 K. The data are on a log-linear plot (a) and a log-log plot (b). . . . . . . . . . . . . . . 85 5.8 Pinning force calculated from data reported in figure 5.7 . . . . . 86 5.9 Critical current density variation with the direction of the applied magnetic field measured on a pure YBa2Cu3O7−δ thin film de- posited at a substrate temperature of 780 ◦C and laser pulses repe- tition rates of 1 Hz and on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz. a) Applied magnetic flux density = 0.5 T, T = 77 K; b) Applied magnetic flux density = 1 T, T = 77 K. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 5.10 Atomic Force Microscopy surface image (5 µm × 5 µm) of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a sub- strate temperature of 740 ◦C; a) repetition rate of 1 Hz; b) repeti- tion rate of 5 Hz; c) repetition rate of 10 Hz; . . . . . . . . . . . . 93 5.11 Atomic Force Microscopy surface image (5 µm × 5 µm) of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a sub- strate temperature of 760 ◦C; a) repetition rate of 1 Hz; b) repeti- tion rate of 5 Hz; c) repetition rate of 10 Hz; . . . . . . . . . . . . 94 5.12 Atomic Force Microscopy surface image (5 µm × 5 µm) of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a sub- strate temperature of 780 ◦C; a) repetition rate of 1 Hz; b) repeti- tion rate of 5 Hz; c) repetition rate of 10 Hz; . . . . . . . . . . . . 95 5.13 Average grain size measured from AFM analysis of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited at a substrate tempera- ture of 740, 760 and 780 ◦C as a function of laser pulse rate . . . 97 xv LIST OF FIGURES 5.14 Visualisation of the transition temperature values measured for 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 740, 760 and 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz . . . . . . . . . . . . . . . . . . 98 5.15 Critical current density variation with the applied magnetic field value measured on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 760 and 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz. The field is applied parallel to the c-axis of the YBa2Cu3O7−δ, T = 77 K. . . . . . . . 100 5.16 Critical current density measured at 0.5 T and 1 T on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a sub- strate temperature of 760 and 780 ◦C as a function of laser pulse rate. The field is applied parallel to the c-axis of the YBa2Cu3O7−δ, T = 77 K. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101 5.17 Critical current density variation with the direction of the applied magnetic field measured on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 760 and 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz. a) Applied magnetic flux density = 0.5 T, T = 77 K; b) Applied magnetic flux density = 1 T, T = 77 K. . . . . . . . . . . . . . . . . . . . . 102 6.1 Bragg-Brentano scan of a YBa2Cu3O7−δ + 2.5%mol Ba2YNbO6 + 2.5%mol Gd3TaO7 thin film deposited on SrTiO3. . . . . . . . . . 107 6.2 X-ray diffraction data from φ scans of (102) YBa2Cu3O7−δ, (202) Ba2YNbO6 and (404) (Y/Gd)2O3 from a 2.5%mol Ba2YNbO6 + 2.5%mol Gd3TaO7 doped YBa2Cu3O7−δ thin film. . . . . . . . . . 109 xvi LIST OF FIGURES 6.3 a) TEM image of a YBa2Cu3O7−δ + 2.5%mol Ba2YNbO6 + 2.5%mol Gd3TaO7 thin film cross section; b) TEM image of plate-like nanopar- ticleof (Y/Gd)2O3 nucleated between two Ba2(Y/Gd)(Nb/Ta)O6 nanorods segments; c) Selected area electron diffraction pattern of the image in (a); d) TEM imafe of plate-like nanoparticle of (Y/Gd)2O3, inset (d) Fourier transform of the particle image. (TEM images from Prof. H Wang research group at Texas A&M Univer- sity). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110 6.4 Critical current density variation with the applied magnetic field value measured on a pure YBa2Cu3O7−δ, a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ, a 2.5%mol Ba2YNbO6 - 2.5%mol Gd3TaO7 doped YBa2Cu3O7−δ and a BaZrO3 doped YBa2Cu3O7−δ thin films [6]. The field is applied parallel to the c-axis of the YBa2Cu3O7−δ, T = 77 K. the data are on a log-linear plot (a) and a log-log plot (b). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 6.5 Critical current density variation with the direction of the ap- plied magnetic field measured on a pure YBa2Cu3O7−δ, a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ, a 2.5%mol Ba2YNbO6 - 2.5%mol Gd3TaO7 doped YBa2Cu3O7−δ and a BaZrO3 doped YBa2Cu3O7−δ thin films [6]. The field is applied parallel to the c-axis of the YBa2Cu3O7−δ. a) Applied magnetic flux density = 1 T, T = 77 K; b) Applied magnetic flux density = 3 T, T = 77 K. . . . . . . 116 6.6 a) High field critical current density variation with the direction of the applied magnetic field measured on a 2.5%mol Ba2YNbO6 - 2.5%mol Gd3TaO7 doped YBa2Cu3O7−δ. Applied magnetic flux density = 3 T to 6 T, T = 77 K; b) sketch of vortices interacting si- multaneously with nanorod segments along the c-axis and intrinsic and extrinsic defects along ab-planes. . . . . . . . . . . . . . . . . 118 xvii Chapter 1 Introduction Superconductivity, a state of matter with a century of history. In the introduction of this dissertation is presented a road map of the main scientific milestones in superconductivity, from its discovery in 1911 to the latest high temperature superconductor discovered. 1.1 The discovery of superconductivity First discovered by the Nobel prize winner H. Kamerling Onnes [7], the supercon- ductive state is characterized by one of the most intriguing properties of matter, the capacity to transport an electric current without the appearance of any re- sistive process. In other words the possibility to transfer and transform energy with reduced losses. H. K. Onnes was awarded the Nobel Prize in Physics in 1913 “for his inves- tigations on the properties of matter at low temperatures which led, inter alia, to the production of liquid helium”. In 1908 he was the first scientist able to liquefy helium [8]; in 1911, during his following investigation on the properties of mat- ter at low temperature, he wrote about The Disappearance of the resistance of mercury. He discovered that the resistance of mercury becomes practically zero when this is cooled below 4.2 K [9]. Onnes realized that zero resistance was a new property that would have characterized a different state of matter, as a matter of the fact in his Nobel lecture he said “the mercury at 4.2 K has entered a new 1 1. Introduction state, which owing to its particular electrical properties, can be called the state of superconductivity.” 1.2 The Meissner effect and the London pene- tration depth Another milestone in the superconductivity history was the discovery of the Meissner effect, in 1933 Walther Meissner and Robert Ochsenfeld, during their researches on the magnetoelectric properties of matter in very low temperatures, observed a sudden change in current distribution and in magnetic induction at the beginning of superconductance. They discovered that tin and lead samples expel the magnetic field when cooled below the transition temperature, in other words when the transition from normal to superconductive state occurs [10]. One of the main implications of the Meissner effect is that superconductivity is something different from perfect conductivity. If two different samples, a su- perconductor and an ideal perfect conductor, are cooled in their zero resistance state and after that an external magnetic field is applied, both the sample would react to the external magnetic field in the same way. The rise of Faraday shielding superficial currents would prevent the penetration of the external magnetic field in the samples. On the other hand if the same two samples are cooled in their zero resistance state while the field is applied, the magnetic field distribution would not change in the perfect conductor, because there would be no induced electromotive force, instead the superconductor will expel the flux also in this second scenario (figure 1.1). A superconducting phase does not allow magnetic flux density to exist whether it is field cooled or not, even a magnetic flux preexisting the transition will be expelled. A few years later, in 1935, a theoretical description of the Meissner effect was found by Fritz and Heinz London [11]. They were able to relate the current distribution to the electromagnetic field. A key factor in the London brothers research was the definition of the magnetic flux penetration depth. Flux expulsion is related to the presence of screening currents. Since only a finite current density 2 1. Introduction Figure 1.1: Differences in magnetisation behaviours of a perfect conductor and a superconductor, the Meissner effect. a) Perfect conductor cooled in zero field; b) Perfect conductor cooled in applied magnetic field; c) Superconductor cooled in zero field; d) Superconductor cooled in applied magnetic field. 3 1. Introduction can be carried in a superconductor, a surface layer in which the screening currents flow has to have a finite thickness. For this reason the magnetic flux density does not decay abruptly at the superconductor surface, but decays gradually penetrating the material. The flux penetration depth, λ, is commonly called the London penetration depth. In simple geometries λ is the characteristic length over which the field density decays exponentially. 1.3 Critical magnetic field, critical current den- sity. Two types of superconductors From an experimental point of view it was clear that the superconductive state disappears when a sufficiently large magnetic field is applied. The field at which the superconductivity is destroyed is called the critical magnetic field, and it is usually denoted as Hc. This field value is strictly related to another fundamental value, the critical current density. If an increasing magnetic field is applied to a superconductor the shielding currents will also increase to produce a zero mag- netic flux density, Hc is the value at which the shielding current density reaches the critical value (Jc). When the applied external field strength is equal to Hc the superconductor is incapable of carrying any transport current, this is due to the fact that the total current is the sum of both shielding and transport currents thus if the shielding current has reached the Jc no others current can be added without exceed the critical current density and therefore destroy the superconductivity. In 1950 Vitaly Lazarevich Ginzburg and Lev Davidovicˇ Landau developed a mathematical phenomenological theory to model superconductivity [12]. The the- ory, describes the macroscopic phenomena in superconductivity relying on general thermodynamics arguments. Despite the lack of explanation of the microscopic mechanisms of superconductivity this theory was fundamental to understand the differences between two main classes of superconductors. Based on the general theory of the second order phase transitions proposed by Landau in 1937, the main variable of the model is the “order parameter” ψ which has a finite value below the transition and zero above it. One of the key results of the model is the definition of two characteristic length: the coherence length and penetration 4 1. Introduction depth. The coherence length ξ, expresses the distance over which the order pa- rameters can differ by a significant amount, in other words represents the distance over which a superconductive phase can become normal. The penetration depth is the characteristic length over which the field density decays, the same char- acteristic length derived in the London theory. The Ginzburg-Landau equations allow the calculation of the energy associated to the formation of a boundary be- tween a normal region and a superconducting region, thus it is possible to predict at which ξ/λ ratio such a boundary is thermodynamically favored. The equations of the theory depend only on the dimensionless material constant, κ, defined as the ratio between λ and ξ. The energy associated to the formation of a boundary between a normal region and a superconductive region is at the base of Alexei A. Abrikosov prediction of a mixed state in which superconductivity and magnetic flux coexist. It was known that for κ values above 1/ √ 2 the surface energy between the superconducting and normal layers would have been negative. Abrikosov decided to solve the Ginzburg-Landau equations for κ values above 1/ √ 2. He found that the magnetic flux and superconductivity could indeed coexist as an array of cylinders of normal conducting material parallel to the field direction surrounded by shielding current vortices in a matrix of superconducting material, the geometric features of this array strongly depend on temperature and field strength [13]. He stated that “in the case κ >> 1 (λ >> ξ) every vortex has a “core” of size ξ, where the order parameter varies rapidly, and the outer region of the size λ where the magnetic field decays to zero” [14] (figure 1.2). Between current vortices a repulsive force arises and the normal core are usually ordered in triangular or square lattice [15,16]. The last outstanding result of Abrikosov’s calculations was that the magnetic flux passing through an area bounded by a superconducting electrical current is quantized and that the quantum of magnetic flux is a constant φ0, it is indepen- dent from the material as long as it is superconductive. This quantized flux is normally referred as a fluxon (Φ0 ≈ 2.067× 10−15 Wb). The presence of the mixed state is confirmed by the magnetization behaviour of type II superconductors that is different from that of type I. A type II super- conductor has two different critical field strength values: the lower critical field, 5 1. Introduction Figure 1.2: Structure of an isolated Abrikosov vortex [1] Hc1, that is the value below which the Meissner effect is present and the entire magnetic flux is expelled, and the upper critical field, Hc2, that is the value above which superconductivity is destroyed. For fields value between Hc1 and Hc2 a type II superconductor is in the mixed state predicted by Abrikosov (figure 1.3). Figure 1.3: Magnetisation curves for a type I and a type II superconductor [1] 6 1. Introduction 1.4 Finally a microscopic theory of supercon- ductivity Since the discovery of superconductivity most of the brilliant minds of the twen- tieth century theoretical physics tried to find a microscopic theory to explain the phenomenon and failed. Almost fifty years passed from the the discovery of the phenomenon to the publication of a microscopic theory of superconductivity. In 1957 Bardeen, Cooper and Schrieffer published their theory on a paper en- titled “Theory of Superconductivity” [17], the theory, known as the BCS theory after the scientist initials, explain “a second-order phase transition at the critical temperature, Tc an electronic specific heat varying as exp(−T0/T ) near T = 0K and other evidence for an energy gap for individual particle-like excitations, the Meissner-Ochsenfeld effect (B = 0), effects associated with infinite conductivity (E = 0), and the dependence of Tc on isotopic mass, Tc √ M = const.” The theory is based on the coupling of electrons in Cooper pairs [18] due to phonon-electron interactions. The interactions between the vibrations of the lattice and the electrons generate an attractive force, when this interaction over- come the Coulomb repulsive interaction the coupling is probable. In the BCS this coupling of the electron leads to the formation of a gap in the electrons energy band, thus any further interactions with an associated energy lower than the gap can not occurs. In a superconductor below Tc the scattering of coupled electron by the lattice (dissipative process) has an associated energy that is lower than the gap, thus it does not happen. The BCS theory is a rigorous quantum mechanical description of the super- conductive phenomenon, a more detailed description of the theory would require a discussion which goes beyond the intent of this chapter. 1.5 The high temperature superconductors rev- olution Since Onnes discovery only metallic materials (mainly type I superconductors) and metal alloys (mainly type II superconductors) were thought to be supercon- 7 1. Introduction ductive. From 1986 onward an acceleration in the scientific achievements leaded to discovery of different families of superconductors. In 1986 Johannes Georg Bednorz and Karl Alexander Mu¨ller discovered high Tc superconductivity in the Ba-La-Cu-O system [19], a year later they were awarded with the Noble prize in Physics “for their important breakthrough in the discovery of superconductivity in ceramic materials”. The discovery of the first cuprate based ceramic superconductor with a Tc of 30 K obtained a con- siderable interest and started an incredible and productive research race. The result of this intellectual spring was the achievement of Tc above 100 K that were considered impossible only a few years before. In 1987 the Y-Ba-Cu-O system was discovered, Tc = 93 K [20], was the first material to be in a superconduc- tive state above the liquid nitrogen boiling temperature. A year later two other cuprates were discovered to be superconductive at even higher temperatures, the Bi-Sr-Ca-Cu-O with a Tc of 108 K [21] and the Tl-Ca/Ba-Cu-O with a Tc of 120 K [22]. These materials are hard to process, but the high-Tc together with the high Hc and Jc made of these materials the ideal candidates for most practical applications. However, despite the large amount of research and published work, a complete theoretical knowledge of this class of superconductors is still missing, in particular the superelectron condensation mechanism is still unknown. In the last few years other two non-cuprate materials were discovered to be superconductive and deserve a mention for the interest that the scientific commu- nity has shown to these newcomers. Magnesium Diboride, Tc of 39 K, discovered in 2001 [23] is interesting for the low-cost of the raw material and for the cable manufacturing process easy and inexpensive when compared to that of cuprates based cables [24–27]. And last, a new family of iron-based superconductors was discovered in 2006 [28], as with the Bednorz-Mu¨ller discovery of cuprates this discovery started a productive race in the search of similar compounds with bet- ter properties [29–32]. However in this case after four years the compound with the highest Tc of the family is still below the liquid nitrogen boiling point: the samarium-doped Sr-Fe-As-F has a Tc of 56 K [33]. 8 Chapter 2 Pinning in YBa2Cu3O7−δ thin films YBa2Cu3O7−δ is undoubtedly one the most promising materials to use in high temperature superconducting cables. Companies and research centers are cur- rently working and studying to achieve better performances, lower production costs and a wider understanding of the physical processes that limit loss less current transport capability. 2.1 Cuprates The two most studied superconductors based on cuprate perovskites are the YBCO (YBa2Cu3O7−δ) and the BSCCO system, the most common supercon- ductors used from the BSCCO system are the Bi-2212 (Bi2Sr2CaCu2O8) and the Bi-2223 (Bi2Sr2Ca2Cu3O10). Cuprates are considered as extreme type II super- conductors and show an intrinsic anisotropy, a layered structure and an extremely small coherence length ξ of the order of a few nm. The small ξ implies a large upper critical field, considering that at Hc2 the normal cores are as densely packed as the coherence length limit allows and that each core carries a quantized flux φ it is possible to estimate the upper critical field value as: Hc2 = φ piξ2 (2.1) 9 2. Pinning in YBa2Cu3O7−δ thin films On the other hand disorder on the atomic scale can influence the supercon- ductive properties. 2.1.1 Weaklinks and the texturing of cuprates Despite the large critical current densities measured for YBa2Cu3O7−δ single crys- tals, or for epitaxial YBa2Cu3O7−δ films on single crystals (Jc > 106MAcm−2 at 4.2 K) [34, 35] one of the main problems that kept YBa2Cu3O7−δ, away from the production of superconducting cables were the low Jc values measured and in general the poor connectivity in polycrystalline samples [36]. Soon it became clear that the critical current densities across grain boundaries are a function of the misorientation angle of the grains’ crystalline lattices. The ratio of the grain boundary critical current density to the critical current density in the grains varies by two orders of magnitude, changing from almost 1 when the grain are well aligned to a drastically reduced value of about 1 50 when the grains tilt angle is above 20 [37]. The BSCCO cuprates show similar behaviours [38], however a self texturing process, that aligns the grains when the BSCCO wires are produced by a simple “powder in tube” method, minimize the problematic related to the weaklinks of the cuprates. The self grain alignment is the reason why large critical current densities were “easily” obtained in the BSCCO conductors [39] while, when sim- ilar method were applied to the YBa2Cu3O7−δ, only rather low critical current densities were achieved [40,41]. Due to this fact, despite the higher Jc, the lower anisotropy and the better in-field performance of the epitaxial YBa2Cu3O7−δ films [35], the first cuprate based superconductors to enter the market were the BSCCO tapes. Today a second generation of coated conductors is entering the scene. These new high temperatures superconducting cables are based on YBa2Cu3O7−δ. The texturing issues were solved adopting two new production techniques: epitax- ial YBa2Cu3O7−δ on highly textured buffer layers deposited by using ion beam assisted deposition (IBAD) [42] and epitaxial YBa2Cu3O7−δ on rolling-assisted biaxially-textured substrates (RABiTS) [43]. In both the process the YBa2Cu3O7−δ is deposited epitaxially on a textured buffer layer. In the IBAD 10 2. Pinning in YBa2Cu3O7−δ thin films technique the texturing is obtained during the buffer layer deposition while in the RABiTS the substrate (usually a nickel alloy) is textured by a thermo-mechanical treatment before the buffer layer deposition [44]. Once the production of long length textured YBa2Cu3O7−δ conductors was achieved the YBa2Cu3O7−δ has become the cuprate in use in new coated super- conductors, thus an optimisation of the material properties has become one of the most important topics. 2.2 YBa2Cu3O7−δ An interesting and complex crystal structure underlie the excellent YBa2Cu3O7−δ properties. For this reason any optimal description of the material has to start with a description of the crystalline structure. 2.2.1 Crystal structure Figure 2.1: a) Crystal Structure of YBa2Cu3O7−δ; b) CuO2 planes; c) CuO chains. 11 2. Pinning in YBa2Cu3O7−δ thin films The YBa2Cu3O7−δ crystal structure is described in figure 2.1a, it is an or- thorhombic, distorted, oxygen deficient perovskite [45]. The lattice parameters of this orthorhombic structure are a = 0.3822 nm, b = 0.3891 nm and c = 1.1677 nm [46,47]. The lattice difference between a and b is small enough to allow epitaxial growth of YBa2Cu3O7−δ on cubic substrates. As a result a large number of twin boundaries are usually formed [48,49], therefore the results of most of the macroscopic measurements are an average of the properties measured over the a and b directions. However since the anisotropy within the ab-plane is small usually an uniaxial anisotropy ab-plane c-axis is assumed. Another usual description of the crystal structure is focused on the Cu-O system. In this description YBa2Cu3O7−δ is pictured as a layered structure made of two distorted planes of Cu-O2 separated by Y +3 ions, and Cu-O copper chains coordinated with Ba+2 ions at the edge of the unit cell. The Cu-O2 copper planes are evidenced in figure 2.1b and the Cu-O copper chains in figure 2.1c. 2.2.2 Oxygen deficiency Oxygen content is a key factor in YBa2Cu3O7−δ as it determines to the structure and properties of the materials [2, 50, 51]. A detailed study of the effects of the oxygen content on the superconductive properties as well as the crystal structure was published in 1990 by Cava R J et al. [2]. The key results of their work was the discovery that on changing the oxygen content form 7 to 6 (δ changes from 0 to 1) the YBa2Cu3O7−δ shows changes in Tc from 92 K to 60 K followed by the disappearance of superconductivity (figure 2.2a) as well as a structural transition from orthorhombic to tetragonal (figure 2.2b). Furthermore Cava R J et al. found that the oxygen content variation involves only the oxygen sites along the Cu-O chains. A direct consequence of this knowledge is that good oxygenation is crucial to achieve high Tc and that any effort to improve the YBa2Cu3O7−δ performance has to take into account the importance of oxygen content. As a matter of fact most of the YBa2Cu3O7−δ production processes include an annealing step in a concentrated oxygen environment. 12 2. Pinning in YBa2Cu3O7−δ thin films Figure 2.2: a) Superconductive transition temperature Tc for ten samples of Ba2YCuOx with varying x values; b) Refined crystallographic cell parameters for Ba2YCuOx with varying x values; from [2]. 13 2. Pinning in YBa2Cu3O7−δ thin films 2.3 Pinning In YBa2Cu3O7−δ the loss-less current transport capabilities are usually also lim- ited by the mobility of the Abrikosov vortices. More deeply the critical current density is determined by complex interactions between the transport currents and the induced magnetic flux lines and the interactions between the flux lines and the nanostructure of the materials. Vortex phenomena is a complex field of solid state matter which is still evolv- ing. The aim of this section is to give a brief introduction to the phenomenon in order to understand the possible interactions, and the improvements that an engineered nanostructure can induce on the critical currents. The following brief introduction is based on an exhaustive description of the phenomenon that was published in 1994 by Blatter G. et al. [52]. 2.3.1 The dissipative phenomenon When a transport current is applied, the flux lines start to move. The movement is due to the action of the Lorentz force (note that c is a constant and η is the friction coefficient): FL = j ∧B/c (2.2) fL = (Φ0/c)j ∧ n (2.3) In the absence of external flux pinning mechanisms the only counter force is the friction force: Fη = −ηv (2.4) At the equilibrium v is the steady-state velocity and FL = Fη thus v can be derived as: v = j ∧B/cη (2.5) The flux motion generate a finite electric field E: 14 2. Pinning in YBa2Cu3O7−δ thin films E = B ∧ v/c (2.6) A finite electric field coexisting with a current density generate a dissipation, E and j are parallel thus the power dissipated is: P = (j ∧B)2/c2η (2.7) In a scenario where the friction force is the only force opposing the Lorentz force it would be impossible to realize a dissipation-free current flow. 2.3.2 Dissipation-free current flow, the pinning force To achieve a dissipation-free current flow, the flux lines have to be pinned in place. In other words v has to be equal 0 also when FL 6= 0. An additional static force, a pinning force Fpin, has to counter the Lorentz force and be active when the flux lines are not moving (v = 0). The flux lines interact with any defects in the lattice. Since the formation of a normal region leads to the loss of the condensate state increasing the energy associated to the system, any region of the material in which the superconducting order parameter is already depressed constitute a region in which the presence of a flux line is energetically favourable. Therefore any region in which the super- conducting order parameter is depressed will contribute to a finite pinning force density Fpin. The critical current density is the current at which the Lorentz force is equal to the pinning force since any further increment of the current flow would lead to an imbalanced increment of the Lorentz force (FL > Fpin) and subsequently a flux motion (v 6= 0) and power dissipation. Therefore assuming FL = Fpin and j⊥B, Jc can be derived as: jc = cFpin/B (2.8) In order to increase Jc it is necessary to increase the pinning force Fpin. 15 2. Pinning in YBa2Cu3O7−δ thin films 2.3.3 Defects as pinning centers A single flux line can interact with different class of defects [53, 54]. Each class is characterised by a characteristic dimension. Defects smaller than ξ in each dimension are called point defects and the saved condensation energy is propor- tional to the defect volume. In anisotropic materials, since ξ is a function of direction, the pinning force provided by this class of defects is also a function of the force direction. Defects with one dimension larger than the others are named linear defects while defects with two dimensions larger than the other are planar defects. In high temperature superconductors naturally occurring defects like dis- location are linear defects and act like pinning lines while twin planes, stacking faults are planar defects and act as pinning planes. In YBa2Cu3O7−δ and cuprates in general the layered crystal structure and the consequent ordered stacking of layers with different superconducting order parameter are an additional array of pinning planes parallels to the ab-planes of the crystal figure 2.1. Linear and planar defects, as well as defects with specific spatial arrangements can lead to directional pinning, increasing the pinning potential with respect to a specific direction of the applied magnetic field. As an example the pinning force provided by a dislocation to a flux line will be higher if the flux line and the dislocation are parallel. In the same way, an ordered array of nanoparticles can act as a linear defect providing the higher pinning potential to flux lines that are able to penetrate the sample in the same direction of the array. In conventional YBa2Cu3O7−δ thin films the Jc is the highest when measured with the applied magnetic field parallel to the ab-planes of the crystal. This is a direct consequence of the directional planar pinning provided by the specific cuprates layers. A last thing to note is that, even if the single interactions between flux lines and defects are at the base of the material response, in most field values and temperature conditions (especially high temperatures) the materials properties are given by the simultaneous interactions of many flux lines with many different defects [55,56]. For this reason a theoretical approach to describe the overall ma- terial behaviour is usually too complex, thus an experimental approach is usually more productive when dealing with pinning engineering. Nevertheless the theo- 16 2. Pinning in YBa2Cu3O7−δ thin films retical achievements greatly enhance the understanding of the physics processes underling the complex behaviours of pinning engineered enhanced superconduc- tors. 2.4 Practical Pinning Engineering Defects engineering of YBa2Cu3O7−δ to increase the Jc values has been and is today one of the hot topics of superconductive materials science. Large electri- cal current transport in absence of energy losses is the key factor in commer- cial application of high temperature superconductors. Since energy losses arise from vortices movements a solution has been researched in nanostructured defects landscape capable to pin the vortices in place. The higher is the efficiency of the pinning landscape the higher is the current value threshold at which the losses appears. Different pinning landscapes and different techniques have been adopted dur- ing the years by researchers, in the following sections an overview of the most interesting works is given. 2.4.1 Non superconducting secondary phase addition The secondary phase addition is an attractive way to introduce defects. It can be adopted both in films grown by physical processes and in films grown by chemical processes. The first example of pinning engineering in YBa2Cu3O7−δ thin films by the introduction of a non superconducting epitaxial second phase is the addition of BaZrO3 [57]. MacManus-Driscoll et al. in 2004 were able to produce high quality YBa2Cu3O7−δ thin films with improved Jc (up to a factor of 5) in the magnetic field by introducing BaZrO3 powder in a Pulsed Laser Deposition (PLD) target of YBa2Cu3O7−δ. They achieved an epitaxial growth of a cubic non-superconducting phase in an epitaxial grown high quality YBa2Cu3O7−δ thin film. An improvement of both the random and directional pinning was obtained. In particular a large increment of Jc was measured with the magnetic field applied parallel to the c- axis of the film. Furthermore, by using TEM (Transmission Electron Microscopy) 17 2. Pinning in YBa2Cu3O7−δ thin films analysis, c-axis oriented columnar defects were found evidencing the correlation between the defects landscape and the pinning potential. This pioneering work is the first of a long series of research articles aiming to produce the ideal pinning landscape adopting a standard industrial ready deposition technique and avoiding technological complications [58–63]. BaZrO3 is not the only secondary phase to produce columnar defects, during the years also BaSnO3 [64–66], RE3TaO7 [67], Ba2YTaO7 [68] where found to pro- duce similar columnar defects. All these phases produce c-axis oriented columns, but the mean width, the spacing and the linearity of the columnar defects pro- duced varies. The variation is dependent on both the phase and the process parameters adopted. A common feature of all this heteroepitaxial columnar de- fects is the perovskite crystal structure with the only exception of a pyrochlore phase (RE3TaO7 [67]). In this work the same technique is used to produce nanostructured YBa2Cu3O7−δ thin film with superior pinning properties adopting a Nb based perovskite as secondary phase as well as in a Nb and Ta simultaneous addition that will result in the complex ideal pinning landscape that will be described later in the thesis. When giving also a small overview of the milestones in the YBa2Cu3O7−δ pinning engineering fields the achievements obtained in chemical solution nanos- tructured YBa2Cu3O7−δ films have to be reported. Gutierrez et al., in 2007, were able to obtain a dense nanodispersion of defects in YBa2Cu3O7−δ thin films by introducing BaZrO3 [69]. In their work an entirely chemical process was adopted and randomly oriented BaZrO3 nanoparticles were reported as the basis of a strong isotropic pinning. Another widely used technique used to produce pinning enhanced YBa2Cu3O7−δ is the multiple target ablation in pulsed laser deposition [70–73]. A secondary phase can be introduced by ablating multiple targets in a so called pseudo-multilayer deposition [70,71]. Haugan et al. were able to improve the pin- ning properties of YBa2Cu3O7−δ by introducing particles of nanometric size of YBa2CuO5 (referred as 211) by growth of alternating layers of ultra thin 221 and YBa2Cu3O7−δ [70]. This multiple bilayer structures were obtained by ablating in sequence a target made of pure YBa2Cu3O7−δ and a target of pure YBa2CuO5 us- 18 2. Pinning in YBa2Cu3O7−δ thin films ing a computer driven target’s carousel. Adopting the same technique Campbell et al. deposited Y2O3 - YBa2Cu3O7−δ pseudo-multilayer films achieving similar results to Haugan [71]. Recently a Nb doped YBa2Cu3O7−δ was also produced adopting the multiple target deposition [74]. Multiple target ablation, when compared to single composite target ablation, has the advantage of having a certain degree of control over some dispersion parameters allowing a change in the number of laser pulses for each layer (the amount of material deposited from each target), and, in theory some control can also be achieved on the deposition parameters adopted for each target. In prac- tice the change of certain deposition parameters like oxygen pressure or substrate temperature during the film deposition would over complicate the deposition pro- cess, therefore in almost all the works adopting this technique the only parameters that are tuned during the secondary phase target ablation are pulse frequency and laser energy. 2.4.2 Alternative routes Other possible routes to increase defects concentration in YBa2Cu3O7−δ thin films have been experimented in recent years and deserve citation. The first is substrate decoration which can be achieved adopting different techniques. The idea is to deposit nanoparticles or nanoislands on the substrate before the YBa2Cu3O7−δ film []. The nanoparticles can be deposited by pulsed laser deposition [75–80], sputtering [81–84] or even with gas-phase prepared nanoparticles [85] and TFA- MOD [86]. These particles will then induce stresses and therefore defects in the film by introducing distortions at the substrate-film interface. The last two techniques differ from the others as they are not based on the addition of a secondary phase. Improved pinning can be achieved by exchanging a certain amount of Y with different rare-earth elements [87]. Enhanced low- field pinning is shown by mixed rare-earth barium cuprate films, while Y based film with large radius rare-earth elements substitution have improved pinning up to at least 7 T. Irradiation with high energy heavy ions has also been adopted in YBa2Cu3O7−δ single crystal pinning research [88–93] and in thin film studies 19 2. Pinning in YBa2Cu3O7−δ thin films [94–99] for its capability to introduce linear defects along the ballistic directions. Despite the interesting and unique features of this technique its high complexity make the irradiation process unpractical. 20 Chapter 3 Experimental Techniques This chapter describes the experimental techniques adopted in the sample prepa- ration and characterisation. The first section will focus on the sample deposition followed by a description of the micro bridge patterning together with the elec- trode deposition. The structural and morphological characterisation techniques are described in the second section. These measurements are mainly performed on unpatterned samples (as deposited samples), and with the only exception of the TEM are non destructive measures. The last section of the chapter describes the transport measures, these are the main characterisation of the superconducting properties. 3.1 Sample Preparation As anticipated in the previous chapter, the nanostructuring technique adopted in this thesis is the introduction of one or more secondary phases in the YBa2Cu3O7−δ. These additional phases are included in the films by introducing the materials directly to the targets used in the pulsed laser deposition. The collection of experimental techniques used to prepare the samples anal- ysed in this work are presented in this section. Below is given a description of the pulsed laser deposition system utilized and a discussion of the technical solutions adopted to obtain the best possible control of the process parameters. A tight control of process parameters is crucial in obtaining a good repeatability. The 21 3. Experimental Techniques sintering procedure applied to the pulsed laser deposition target preparation is described after the deposition technique while the photolithographic process per- formed to obtain the micro bridge patterning and the electrodes deposition are explained at the end of the section. 3.1.1 Pulsed Laser Deposition Pulsed laser deposition (PLD) is a valuable choice for depositing extremely pure films because it reproduces exactly the target composition, multilayer materials can be done rather easily, it is the fastest route to prototyping any thin film coating and, most importantly, very high quality YBa2Cu3O7−δ thin films can be easily produced. As a matter of the fact of its extreme versatility, a considerable amount of research work in YBa2Cu3O7−δ pinning engineering has been carried out adopting pulsed laser deposition as deposition technique [100–104]. Pulsed laser deposition is one of the physical vapor deposition (PVD) adopted to grow high quality thin films. A target of the desired composition is ablated by a high power pulsed laser, the material ablated from the target forms a plasma plume of energized ions that after traveling though a controlled atmosphere are deposited on an heated substrate forming a thin coating film. Once these ions reach the film surface thay are usually referred to as adatoms. In particular adatom is the term used to describe a single atom adsorbed on a surface. To obtain high quality thin films it is necessary to control several parameters during the deposition. These parameters are the substrate temperature, the oxygen pressure in the chamber, the laser energy and the frequency of the laser pulses [105,106]. All the samples in this work were deposited in a YBa2Cu3O7−δ dedicated de- position system to minimize contamination issues. This system is equipped with a Lambda Physik KrF excimer laser (λ = 248 nm, fluence = 2 mJcm−2) and the laser beam is admitted to a ultra high vacuum (UHV) chamber through a quartz window. Periodic cleaning of the window as well as a constant monitoring of the energy admitted in chamber ensure homogeneity of the ablating energies over the time. Others quartz windows are mounted on the sides of the UHV chamber and are generally used for growth and plume monitoring, secondary substrate surface 22 3. Experimental Techniques temperature monitoring and for laser-target-substrate alignment. A focal lens is positioned outside the chamber and it is used to focus the laser on the targer in order to form a 2 mm x 3 mm rectangular spot on the target surface. The laser spot geometry influence the plume geometry: a large spot generates a small plume thus a small region in which the deposition is uniform; a small spot generates a large plume and a large region in which the deposition is uniform but as a consequence the deposition rate is reduced. The laser spot geometry adopted in this work allows a uniform deposition in a region ≈ 1.4 cm x 1.4 cm at ≈ 5 cm away from the target surface and uniform films are easily obtained on 1 cm x 0.5 cm substrates. The substrate surface temperature is monitored using an external pyrometer while a secondary visible light laser, aligned to the primary excimer laser, is used during the aligning task. Figure 3.1 shows a schematic of a typical pulsed laser deposition system. Figure 3.1: Schematic of a typical pulsed laser deposition system [3] The substrates used are (001) aligned single crystals of SrTiO3 (STO), these 23 3. Experimental Techniques are glued to an heater/holder with an high thermal conducting silver paste. The temperature of the heater is controlled using a PID controller that controls the power supply of the heating elements measuring the internal temperature of the heater block with a thermocouple. Since the thermocouple is inserted in the heater the use of the external pyrometer to monitor the substrate surface tem- perature is adopted to avoid substrate temperature variations between different depositions. This variations could rise from unavoidable differences in the ther- mal conductivity between the heater and the substrate. A variable ∆T between the thermocouple the pyrometer measure is usually present. This ∆T is mainly due to the position of the thermocouple. The thermocouple is placed inside the heater and at high temperatures a large energy dissipation due to radiating pro- cesses generates large thermal gradients between the center of the heater, where the thermocouple is positioned, and its surface, where the substrate is glued. Furthermore the impossibility to reproduce the same thermal contact in every deposition adds a variability to the discussed ∆T . This variable ∆T and the fact that the substrate surface temperature is the one that determines the atoms mo- bility, make the substrate surface temperature directly measured by the external pyrometer the only possible choice as reference temperature. The target is mounted on a rotating carousel equipped with a plume shutter. The target is rotated to provide a uniform ablation. while the shutter is used to protect the substrate from the plume during a pre-deposition ablation pro- cess. This pre-ablation process is adopted to remove the target surface layer and to ensure that the entire deposition is realized under similar target conditions. The laser ablation induces modification to the surface morphology of the target inducing a continuous fusion-recrystallization cycle. It is important to induce these morphologic modifications before starting the thin film deposition without allowing the plume to reach the substrate. The vacuum system is a common two stage system equipped with a rotary pump and turbomolecular pump, an additional multi channel mass flow controller is also present. The rotary pump is used to pump down the chamber pressure from ambient pressure to mid vacuum (∼ 103 mbar to ∼ 10−3 mbar) and to provide a backing pressure ≤ 10−3 mbar to the turbomolecular pump. The tur- bomolecular pump is used from mid vacuum to high vacuum (∼ 10−3 mbar to 24 3. Experimental Techniques ≤ 10−6 mbar). Depositions are performed in a pure oxygen atmosphere with an oxygen pressure of 0.3 mbar. These pure oxygen deposition atmospheres are realized by first reaching high vacuum (≤ 10−6 mbar) using the turbomolecular pump and then venting oxygen while controlling the chamber pressure by simul- taneously controlling the oxygen incoming flow with the mass controller and the extraction pump power. In a typical deposition the laser is operated in constant energy mode and the fluence is monitored and kept at ≈ 2 mJcm−2, the pulse frequencies adopted are between 1 and 10 Hz. Usually each deposition is performed using ∼ 4500 pulses to obtain a film thickness of ∼ 500 nm. The distance between the target and the substrate is ∼ 5 cm. An optimal oxygenation of the samples is obtained with an in situ low temperature annealing of 1 hr. This step is realized cooling the sample to ≈ 520 ◦C and raising the oxygen pressure to 500 mbar. All thin films produced in this research work were deposited on SrTiO3 (001) single crystal substrates. This substrate is commonly used in YBa2Cu3O7−δ epi- taxial growth by pulsed laser deposition. It allows high quality films to be de- posited and it was chosen for ease of comparison with previous work. In particular the use of SrTiO3 (001) allows the comparison of the properties of the sample produced with a large amount of existing data without introducing unnecessary variables. Oxygen content, grain orientation and film morphology are determinant pa- rameters of the films superconducting properties. All these features are strongly influenced by the deposition parameters adopted. Almost twenty years of research in the field has provided a good knowledge of the processability windows of pure YBa2Cu3O7−δ. High quality YBa2Cu3O7−δ thin films deposition is technological knowledge already acquired and thus will not be discussed. An exhaustive re- view on pulsed laser deposition of pure YBa2Cu3O7−δ thin films was published by Singh et al. [3]. 3.1.2 Target Preparation All the pulsed laser deposition targets used to produce the samples used in this work were prepared from sintered powders. Although commercial pure 25 3. Experimental Techniques YBa2Cu3O7−δ targets are available it was chosen to produce pure YBa2Cu3O7−δ targets in laboratory in order to avoid differences in the superconducting prop- erties which may arise from different target quality rather than different targets composition. The process adopted to produce the targets is the same for both the pure and the composite ones. Powder are pressed in the form of a cylindrical target and the sintered at 950 ◦C in oxygen flow for 12hr in a dedicated tubular furnace. A detailed scheme of the thermal process is reported in figure 3.2. Figure 3.2: The target sintering thermal process Pure YBa2Cu3O7−δ powder (SCI Engineered Materials 99.999%) is used in both pure and composite targets. The secondary phases used in this work are Ba2YNbO6 and Gd3TaO7. Targets produced by mixing pure YBa2Cu3O7−δ with 5 mol% of Ba2YNbO6 powder were adopted in the deposition of the samples discussed in the next chapter. Targets produced by mixing pure YBa2Cu3O7−δ with 2.5 mol% of Ba2YNbO6 powder and 2.5 mol% of Gd3TaO7 precursor were used in the deposition of the samples discussed in the last chapter. The Gd3TaO7 precursor mixed with the pure YBa2Cu3O7−δ and the Ba2YNbO6 in the target were 99.99% Gd2O3 and Ta2O5. 26 3. Experimental Techniques 3.1.3 Sample Patterning While measuring the critical current density Jc a limited critical current Ic is desired. A low Ic reduce a series of problematic side effects and technological complication during the measure. A large Ic would require large current transfer, this implies powerful currents sources, large conducting cables connecting the sample and large electrodes trough which transfer the currents to the sample. Further more since it is impossible to eliminate the contact resistances a large transfer current passing trough any residual contact resistance will cause the dissipation of large energies. To avoid an increment of the sample temperature, caused by the heat generated in the dissipations at the contact, a powerful cooling system should also be provided. Considering equation 3.1 the easiest way to reduce Ic is the reduction of the superconductor’s cross section. Ic = jcSSupercond. (3.1) A reduction of the cross section Ssupercond. simplifies the measure of the large Jc typical of the YBa2Cu3O7−δ thin films (usually Jc > 106 MAcm−2) and is obtained by introducing current tracks of micrometric width with a lithographic process. A reduction of the contact resistances can be obtained with the deposition of an electrode array through which the currents are transferred to the sample. Also the electrodes are deposited with a lithographic process. The electrode deposition is realized with the lift-off process described below. First a layer of a commercial positive photoresist (AZ 4533) is spin coated on the sample (spin at 6000 rpm for 45 s), then the resist layer is baked at 110 ◦C for 1 minute to allow the resist network reticulation (figure 3.3b). An intense blue light is projected on the coated sample surface using a pro- jection mask aligner for 24 s, the mask projected at this stage is composed of two rows of six square holes, the intense blue light is projected on the sample only through these twelve squares. The exposed regions of the positive photoresist become soluble to a developing solution while the unexposed regions remain insoluble, thus after a developing stage the two rows of six square holes printed on the mask are transferred to 27 3. Experimental Techniques the photoresist coating layer. The developing stage is realized by the submersion of the samples in a 3vol : 1vol H2O:AZ351B developing solution (AZ351B is a commercial developer). The duration of the developing stage is a key factor not always predictable: it has to be long enough to ensure that the resist layer is completely removed in the areas that were exposed to the light but at the same time an over lasting developing stage is to be avoided since it can cause resist removal from the unexposed zones. In order to optimally time the duration, the developing step is divided in several steps of reduced duration followed by the observation of the sample surface with an optical microscope until the resist layer is completely removed from the exposed areas (figure 3.3c). Once the developing stage is successfully completed a silver layer and a gold layer are deposited on the sample surface in an ion milling magnetron sputtering hybrid system. The Ag / Au bilayer deposition is preceded by an ion milling step to remove the passivated layer on the superconducting film ensuring the optimal connectivity between the Ag layer and the YBa2Cu3O7−δ surface (figure 3.3d). The last step of the lift-off process, which gives the name to the technique, is realized by submerging the samples in acetone in an ultrasonic bath. The acetone dissolves the photoresist layer and the resist removal causes the lift-off (delamination) of the Ag / Au bilayer deposited on it while the metals deposited directly on the YBa2Cu3O7−δ surface in the areas that were not covered by the resist forms the electrodes. In this way two rows of six Ag / Au electrodes are deposited on the films (figure 3.3e). The track pattern transfer (the micro bridge formation) is obtained with a process similar to the electrode deposition until the developing stage. In this process the mask used project the light in the areas of the samples from which is necessary to remove the superconducting material. Thus after the developing stage, the resist layer is the exact copy of the desired tracks pattern (figure 3.3g), the effective film etching can then by realized either by a physical or chemical process (dry etching or wet etching). Very small features ( ≤ 10 µm) are usually realized with a dry etching because the chemical etching could blur the features. Since the smallest feature in this work, the tracks width, is ≈ 50 µm both processes are equally effective. The dry etching is performed in an argon ion milling system where energized ions removes material from the sample surfaces. 28 3. Experimental Techniques The wet etching is realized by submerging the sample in a HCl diluted solution (0.01M) with the acid dissolving the unprotected YBa2Cu3O7−δ from the sample (figure 3.3h). A schematic of the sample at the different stages of the process is reported in figure 3.3 while a micrography showing a single track and fractions of two electrodes in shown in figure 3.4 Five tracks (≈ 50 µm in width), each one connected to four Ag / Au elec- trodes, are patterned on the samples using the described techniques. 3.2 Structural Morphological characterisation In this section are listed the scientific investigation techniques adopted to charac- terise the crystal structures and the morphology at the nanoscale of the samples. Since these techniques are commonly adopted by materials scientists working in thin films, as well as different fields, they will not be discussed in detail. In this section only a presentation of the answers that can be achieved with the analysis performed is given. 3.2.1 Structural Analysis Phase and orientation analysis are performed using x-ray diffraction. Performing this non destructive technique it is possible to gather information on the crystal phases present in the thin films and their orientation. The distribution of grain orientations is described by the crystalline texture. Random texture describes films with randomly oriented grains, fibre texture is a term used to describe films in which grains have one crystallographic axis parallel to the substrate normal but are randomly oriented in plane, and epitaxial growth (or epitaxial alignment) describes films made of grains that have a fixed orientation in plane. A scan in Bragg-Brentano geometry will give information on the spacing of the crystalline planes that are oriented perpendicularly to the substrate surface. The analysis are usually performed in the geometrical configuration reported in figure 3.5, and in this configuration only the diffraction peaks associated to the planes with the normal oriented perpendicular to the substrate surface will be registered. 29 3. Experimental Techniques Figure 3.3: Sketches of a sample at the different stages of a photolithographic process. a) As deposited sample; b) Sample after the first spin coating of the photoresist; c) Sample after the developing of the electrode mask; d) Sample after the silver layer and gold layer deposition. e) Sample after the lift-off process; f) Sample after the second spin caoting of the photoresist; g) Sample after the developing of the trak mask; h) Sample after the milling/cleaning process. 30 3. Experimental Techniques Figure 3.4: Micrography of the surface of a patterned sample Figure 3.5: Schematic of a diffractometer operated in the Bragg-Brentano Geom- etry 31 3. Experimental Techniques The possible presence of an epitaxial growth is easily evidenced by the presence of only the peaks indexed (00l) since this will be the experimental proof that the crystalline c-axis is the only crystal axis oriented perpendicular to the substrate surface. The analysis of the diffraction peaks gathered in this geometry allows the determination of the crystalline composition of the samples and a first information on the orientation of the phases. A limitation of this typology of x-ray diffraction analysis is to only gather information of the so called out of planes diffraction. The determination of the out of plane orientations allows to distinguish a random texture from a fibre texture or an epitaxial growth. In order to determine whether a film can be described by a fiber texture or an epitaxial growth it is necessary to analyse the in plane orientation of the crystalline phases. The in plane orientation analysis is performed with the so called φ scan. In figure 3.6 a sketch of the geometrical configuration in which the φ scans are done is reported. Figure 3.6: Schematic of a diffractometer geometry operating a φ scan In a φ scan the sample orientation χ, the tube and detectors position θ are predetermined to focus the machinery on a specific in plane diffraction peak. The scan is then performed varying the sample φ angle and measuring the intensity 32 3. Experimental Techniques of the diffracted x-rays at the different φ values. In this way it is possible to asses the in plane orientation of a given crystalline phase and determine whether that specific crystalline phase is growing following the orientation of the substrate crystalline axis or is growing with a determined offset or is not following any specific orientation. This technique is valid if the phases were rightly determined and if the peaks on which the analysis system is focused are unique. These first two x-ray diffraction techniques described are an optimal solution in the determination of the crystalline phases and their orientation but are limited when determining the epitaxy quality. They give little to no information on the on the spread of the orientation angles that is usually below the resolution of systems adopted when operated in the geometrical configuration described. Further more the presence of small fraction of randomly oriented crystalline phases could easily remain undetermined by these analysis. A randomly oriented fraction of a given oriented phase could be interpreted as noise when performing a φ scan. Figure 3.7: Schematic of a diffractometer geometry operating a rocking curve scan In order to address the epitaxy quality and the strain levels rocking curves and reciprocal space maps are gathered. A rocking curve (ω scan) is an analysis of the angular spread of a determined 33 3. Experimental Techniques planar direction. It is realised by scanning with a fixed 2θ (a fixed pi − 2θ) a small windows of positive and negative offsets ω (figure 5.2). The full width half maximum of the peaks measured gives an indication of the angular spread of the planar direction analysed. Figure 3.8: Schematic of a diffractometer geometry scanning a reciprocal space map A reciprocal space map is the detailed study of the x-ray diffraction in a small window of 2θ and ω angles. A reciprocal space map can be obtained by measuring a set of ω/2θ scans with different ω-offset angles around a specific reciprocal space reflection, determined by ω and 2θ (figure 3.8). The intensity of the diffracted signal is then plotted on a map with coordinates qz and qx that are the reciprocal of the spacing dz and dx thus it is possible to directly calculate the lattice parameters. The position of the diffraction peaks and their shape gives information on the mean lattice spacing and on the strain of the phase analysed. It is also possible to establish if a phase is relaxed or strained with respect to the substrate. As a downside the reciprocal space maps are time consuming mea- sures and therefore the windows around a specific diffraction angles are usually the smallest possible. Even if in theory it would be possible to gather a complete maps this are always avoided for the impracticality of the time required. 34 3. Experimental Techniques 3.2.2 Microscopy The knowledge of the crystalline phases, their orientation and strains level rep- resents only a partial knowledge of a film quality. Furthermore since the pinning potential is related to the nanostructuration of the phases, a study of the mor- phology on the nanometric scale is fundamental. Only by knowing the grain mor- phology and the shape, spacing and orientation of a secondary phase it is then possible to relate the defect landscape to the pinning properties. Furthermore it is possible to evaluate the strength and weaknesses of a defects distribution in order to improve the properties increasing the points of strength and decreasing the weaknesses of a pinning landscape generating ideals defects distribution. There are two main scientific investigation tools capable to fulfill the task: the atomic force microscopy (AFM) and the transmission electron microscopy (TEM). AFM [107] is used to investigate the surface topography of the samples, from the topography is possible to evaluate the homogeneity of a film and the roughness of the surface. In addition to shapes and dimensions of the grain in samples characterised by a smooth surface is also possible to investigate the presence and distribution of secondary phases. The technique is non destructive and it is also fast and reliable, unfortunately the only information gathered concern the samples surface, thus only a partial information of the defects distribution can be obtained. As an example it would be impossible from an AFM only investigation to establish whether a secondary phase particle is a c-axis oriented defects or a plate like ab-planes oriented particles. For these reasons it is necessary to complete the information with TEM. TEM gives information on the distribution and orientation of the phases in the sample. It is a powerful tool but is a time consuming and destructive technique. Nevertheless it is fundamental when picturing the pinning landscapes produced in the films. Essentially with this technique are taken images of the samples cross-section at the nanometric scale. A direct observation of the nanoparticles introduced in the samples is obtained. Furthermore analysing the electron diffrac- tion pattern taken with the same technique it is possible to confirm and complete the phase and orientation analysis produced the the x-ray diffraction. A TEM 35 3. Experimental Techniques allows to take different electron diffraction patterns from different locations of the sample thus allows to gather separated electron diffraction pattern from the different phases present in the sample. TEM is also adopted to perform local strain analysis to complete the information on the average strain obtained with the reciprocal space maps. Cross-section TEM images were taken by JEOL 2010 analytical microscope with a point-to-point resolution of 0.20 nm. High resolution TEM and scanning transmission electron microscopy analysis was conducted using FEI Tecnai F20 with a point-to-point resolution of 0.18 nm. Cross-sectional samples for TEM analysis were prepared by a standard manual grinding and thinning procedure followed by a final ion polishing step (Gatan PIPS 691 precision ion polishing system). 3.3 Superconductivity Properties characterisa- tion The final part of this chapter is dedicated to on overview of the techniques adopted to characterize the superconducting transport properties of the sample produced. The three main properties analysed are the transition temperature Tc, the critical current density as a function of the applied magnetic field value Jc(B) and the critical current density as a function of the direction of the applied magnetic field Jc(B, θ). It is evident that the target is to achieve the highest possible Jc and Tc and to minimize the detrimental effect of an increasing value of the applied magnetic field on Jc. This set of measurements gives complete information on the quality and effectiveness of the defects landscape in the creation of an ideal pinning enhanced thin film YBa2Cu3O7−δ based superconductors. The first measure always performed is the determination of the transition temperature Tc. This measure is performed by a straight forward technique. A 4 point resistance measure is realized applying the smallest transport current possible that minimize the noise while not affecting the transition temperature (usually I ≤ 10−6 A). The resistance is measured while reducing the tempera- ture from TAmb to 77 K. The transition temperature is then associated with the 36 3. Experimental Techniques temperature at which the resistance disappears. The attribution of Tc is not unique, in some work it is attributed to the temperature with the maximum of the ∂ρ ∂T or with the mean point of the transition, in this work the most restrictive condition is chosen. The reason to chose the most restrictive condition for the determination of Tc is that this condition takes into account the homogeneity of a sample and that a sample will not be considered superconductive until a complete superconductive link is generated between the probing electrodes. The critical current density as a function of the applied magnetic field value Jc(B) and the critical current density as a function of the direction of the applied magnetic field Jc(B, θ) are measured with the same technique. The only difference is the parameter that is changing during the measurement. In the Jc(B) measure the direction of the applied magnetic field is constant (usually B‖ c-axis and B ‖ ab-planes) and the applied magnetic field value is increased. In the Jc(B, θ) measurement the applied magnetic field value is constant and the direction is changed. The core of the measurement is the individuation of the critical transport current Ic. To fulfill this task the current is increased in small step and a voltage value is measured for each current value until the voltage reaches a maximum value (V = 10−5V ), from the I−V curve obtained the current value at which the voltage is equal to a defined criterion (Vc = 10 −6V cm−1) is derived and that value is defined as the Ic value. Knowing the Ic and the geometry of the micro bridge the Jc is easily calculated using equation 3.1. The derivation of the Ic value can be achieved by adopting different technique: a linear interpolation around the criterion voltage, a polynomial interpolation over a n-point windows around the criterion or fitting the I − V data with the following equation: I = Ic( V Vc )n (3.2) The different Ic derivation techniques lead to little or no variation of the value derived. Never the less the technique adopted in the work is the fitting with the equation 3.2 since it is more reliable in the case of noisy measures. The Jc(B) and the Jc(B, θ) measurements are basically realised with a set of I − V measures in different magnetic fields intensity and geometry. Both Jc(B) 37 3. Experimental Techniques Figure 3.9: Schematic of the geometry for the determination of Jc(B, θ) and Jc(B, θ) measurements are performed in maximum Lorentz force configura- tion (I ⊥ B) figure 3.9. 38 Chapter 4 Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results In this chapter are reported the results of the study of the Nb doping of YBa2Cu3O7−δ thin films. The effects on the nanostructure of the films as well as the superconducting properties are discussed. The study in this chapter includes the preliminary results obtained on Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited by pulsed laser deposition. 4.1 The Nb introduction in the YBa2Cu3O7−δ The introduction of a secondary phase in the YBa2Cu3O7−δ that has been de- scribed in the previous chapter is now a widely used tool to increase the pinning properties of commercial conductor as well as research labs advanced materials. When introducing new pinning additions to YBa2Cu3O7−δ adopting the pulsed laser deposition technique, it is important to consider that the phase which forms may not be the one which is added to the YBa2Cu3O7−δ in the target. The phase that forms is the most thermodynamically and epitaxially stable. As a matter of the fact if Zr4+ is added to the YBa2Cu3O7−δ the phase that forms is Ba(Zr,Y)O3 [57] while when Sn 4+ is added BaSnO3 forms [64–66]. The addition of Ta5+ was reported to form both RE3TaO7 [67] and Ba2YTaO7 [68]. Consid- ering the optimal results obtained with the Ta5+ in this chapter the effects of 39 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results another large, highly charged ion, Nb5+, are presented. 4.1.1 Nb doped ceramic samples The first introduction of Nb in YBa2Cu3O7−δ was reported soon after the dis- covery of the YBa2Cu3O7−δ. In 1988 Kuwabara M et al. reported an increased Jc value in Nb2O5 doped YBa2Cu3O7−δ ceramic samples [108]. The effects of Nb was unclear but the authors reported that almost all the Nb added was not in- corporated in the YBa2Cu3O7−δ and was instead segregated to form a secondary phase. An elemental distribution map showing that the Cu element did not co- exist with the Nb while both Y and Ba did and the discovery that Nb was no longer in the from of oxide made the authors claim the formation of a secondary compound with Y and Ba. In 1989 Greaves C. and Slater P.R. published a study on the Nb and Ta substitution in the REBa2Cu3O7 (RE = Y, Eu, La) [109]. They suggest the presence of two possible products, the REBa2Cu2MO8 and the Ba2REMO6 + CuO (M = Nb, Ta). They observed that the energy difference between the two possible products is small and that, since in the double perovskite structure the rare earth cation is in the small octahedral site, the stability of the perovskite structure is reduced when large RE3+ ions are used. The authors observed that for RE = Y, Eu the phases formed where REBa2Cu3O7, Ba2REMO6 and CuO while a substitution and the consequent formation of the REBa2Cu2MO8 phase was only observed for RE = La. From their studies it is evident that doping YBa2Cu3O7−δ (RE = Y) with Nb or Ta leads to the formation of the Ba2YNbO6 and Ba2YTaO6. A few years later, in 1993, the diffraction pattern of a sample with composition YBa2Cu2.95Nb0.005O7−δ was reported in a paper from Strukova G. K. et al. [110,111]. They also found the presence of an YBa2Cu3O7−δ orthorhombic phase (a = 0.3827 nm, b = 0.3894 nm, c = 1.1718 nm), a Ba2YNbO6 cubic perovskite phase (a = 0.4218 nm) and a small amount of CuO. They also reported that the intensity of the lines corresponding to the Ba2YNbO6 increased with increasing the Nb content. They attributed a greater stability of the Ba2YNbO6 secondary phase when compared to YBa2Cu3−xNbxO7−δ to the low solubility of the Nb in 40 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results the YBa2Cu3O7−δ. Recently, in 2003, Babu N. H. et al. reported the presence of Y2Ba4CuxNb1−xOy, a phase similar to Ba2YNbO6 with Cu partially substituting Nb, in YBa2Cu3O7−δ bulk doped with Nb [112]. Improved Jc was observed in the doped samples [113]. In 2009 Yeoh W. K. et al. reported enhanced flux pin- ning at high field in YBa2Cu3O7−δ single grain bulk samples doped with NbO2 and related the improved pinning to the formation of a Nb rich phase within the YBa2Cu3O7−δ [114]. 4.1.2 Ba2YNbO6 use in YBa2Cu3O7−δ thin film The stability and compatibility of the Nb-based double perovskite Ba2YNbO6 with the YBa2Cu3O7−δ was studied by Paulose K V et al. in 1992 [4]. They found that annealing a 1:1 mole mixture of YBa2Cu3O7−δ and Ba2YNbO6 for 16 hr at 950 ◦C there was absolutely no reaction (figure 4.1). The stability and compatibility of the Ba2YNbO6 with the YBa2Cu3O7−δ made it a candidate for a novel substrate in the YBa2Cu3O7−δ thin film depo- sition. The difficulties encountered in the production of large Ba2YNbO6 single crystal have prevented the use this material as a substrate, nevertheless high quality high temperature superconductor thin films and electronic devices were fabricated adopting Ba2YNbO6 as a new buffer layer [115,116]. The only early report on the use of Ba2YNbO6 to enhance flux pinning in YBa2Cu3O7−δ thin film is a work published in 1995 from Jia J. H. et al. [117]. The authors report on thin Nb added YBa2Cu3O7−δ deposited on ZrO2 substrates by a dc magnetron sputtering method. The secondary phase formed in the film is reported to be the Ba2YNbO6 perovskite but the effects on the critical current registered were negative. The conclusions of this early research were that the Ba2YNbO6 secondary phase does not play a role in the flux pinning process. More then 10 years later, in 2007, Nb doped ErBa2Cu3O7−δ was successfully produced by pulsed laser deposition [118–122], in this work the authors report that c-axis oriented nanorods of Ba2ErNbO6 were formed in the ErBa2Cu3O7−δ. However the effects of the nanorods on the critical current values were not clearly explained. The presence on these c-axis oriented nanorods seemed to have effects 41 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results Figure 4.1: Powder diffraction patter for a) pure Ba2YNbO6, b) pure YBa2Cu3O7−δ, c) 1:1 mole mixture of YBa2Cu3O7−δ:Ba2YNbO6 heated at 950 ◦C for 15 hr. [4] 42 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results on the pinning performances of the films only when the latter were grown under certain condition and a window of processability was undefined. Recently, successful formation of Ba2YNbO6 nanorods in YBa2Cu3O7−δ thin films grown by pulsed laser deposition has been reported by several authors [5, 74,123,124]. The article [5] is part of this thesis, it is the first attempt to produce Ba2YNbO6 - YBa2Cu3O7−δ thin films by pulsed laser ablation of a mixed target, and contains the first results obtained. These results will be described in the following section of this chapter. In conclusion Ba2YNbO6 is a good pinning phase candidate for the following reasons: • The large Nb+5 does not substitute the Cu in YBa2Cu3O7−δ • Ba2YNbO6 is the most stable Nb compound in a YBa2Cu3O7−δ matrix • Ba2YNbO6 forms in a wide processing window • Ba2YNbO6 nanoparticles forms self-assemble in nanorods in a similar man- ner to BaZrO3 • The large lattice mismatch between Ba2YNbO6 and YBa2Cu3O7−δ should provide additional pinning inducing lattice distortion in YBa2Cu3O7−δ 4.2 Ba2YNbO6 perovskite additions to YBa2Cu3O7−δ: the preliminary results In this section is reported the first results obtained by the addition of the Ba2YNbO6 perovskite pinning phase to the YBa2Cu3O7−δ. The results were discussed in a feasibility study published in early 2010 in Superconductor Science and Technol- ogy [5]. 4.2.1 The Ba2YNbO6 synthesis, the pulsed laser deposi- tion target preparation and thin films deposition The technique adopted to produce the single phase Ba2YNbO6 powder is the same described and successfully used by Paulose K V et al. [4]. Ba2YNbO6 43 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results powder were produced mixing and grinding stoichiometric quantity of 99.99% Y2O3, Ba(NO3)2, Nb2O5 and a minimal amount of CuO2 followed by a solid state reaction at 1450 ◦C for 24 h in flowing O2. In order to achieve a complete reaction and thus produce a single phase Ba2YNbO6 two successive grinding and reaction were needed. The minimal amount of CuO2 (≈ 0.5%wt.) was added to the oxides mixture because it is described to greatly enhance the powder reactivity. The reaction enhancing effect of the CuO2 is explained by the fact that CuO2 presence introduce in the system a low energy reaction intermediate lowering the energy barrier of the overall reaction [4]. Figure 4.2: Powder diffraction patter for Ba2YNbO6powder produced. Adapted from [5]. In the figure 4.2 are shown the results of an x-ray analysis in the Bragg- Brentano geometry (θ − 2θ scan) of the Ba2YNbO6 powder produced by the described solid state reaction. No impurities or secondary phase peaks are found, given the minimal amount of CuO2 that was added is not expected to be easily evidenced. The quantity involved is too small to be separated from the noise levels that are usually present in this typologies of measures. The lattice parameters calculated form the x-ray data is a = 0.844 nm ± 0.001 nm. The lattice parameter derived is in agreement with the data recorded in the Joint Committee on Powder 44 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results Diffraction Standards database (JCPDS) as well as the data presented in more recent studies [125]. The Ba2YNbO6 production process described constitutes a reliable source of pure single phase Ba2YNbO6 to use, together the YBa2Cu3O7−δ commercial powder, in the pulsed laser deposition target preparation. The target sintering procedure adopted for the samples discussed in this section is exactly the general procedure described in the previous chapter, the Ba2YNbO6 doping amount in the target adopted in this study is 5%mol. The control pure YBa2Cu3O7−δ target is sintered adopting the same technique. More sophisticated procedures were also adopted to produce advanced targets, the modification to the processes will be described in the appropriate spaces. The pulsed laser deposition parameters adopted for both pure YBa2Cu3O7−δ control thin films and the Ba2YNbO6 added samples are summarized in table 4.1. Parameter Value Substrate Temperature 770 ◦C Chamber Pressure 0.3 mbar flowing O2 Laser Fluence 2 Jcm−2 Repetition Rate 5 Hz Number of Pulses 4500 Annealing Time 1 hr Annealing Temperature 520 ◦C Annealing Pressure 500 mbar O2 Table 4.1: Pulsed laser deposition parameters All the thin films deposited were measured to be of ≈ 0.5 µm thickness. 4.2.2 Crystalline structure analysis: x-ray diffraction data The first analysis realized on deposited films when using a new target composi- tion are usually x-ray diffraction structural analysis to investigate the crystalline phases produced in the samples, their orientation and possibly the strains and lattice distortions arising form the large interfaces of nanostructured composite materials. 45 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results 4.2.2.1 Crystalline phases identification Figure 4.3: X-ray diffraction data for undoped YBa2Cu3O7−δ and 5%mol Ba2YNbO6 doped film deposited on SrTiO3. Adapted from [5]. The first relevant results reported in figure 4.3 is that the x-ray diffraction data collected in the Bragg-Brentano geometry from the deposited thin films (both doped and pure) shows only the diffraction peaks related to the (00l) planes from the YBa2Cu3O7−δ, the Ba2YNbO6 and the SrTiO3. This is evidence that also the doped thin films have the usual c-axis orientation of pure YBa2Cu3O7−δ thin films deposited on (001) SrTiO3 single crystal. The only difference evidenced by the comparison of the diffraction data gath- ered from the pure YBa2Cu3O7−δ control thin films and the Ba2YNbO6 doped ones is the presence of the peaks related to the (00l) planes of the Ba2YNbO6. It is possible to conclude that the stability of the Ba2YNbO6 - YBa2Cu3O7−δ sys- tem during the pulsed laser deposition does not differ from the proved stability at high temperatures [4], thus the only secondary crystalline phase introduced in the thin films deposited is the Ba2YNbO6. Furthermore a first hint of the Ba2YNbO6 orientation is revealed by the presence of only the (00l) planes diffraction peaks, the non superconducting phase is aligned out-of-plane with the YBa2Cu3O7−δ 46 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results matrix. To conclude the crystalline phases identification a small discussion has to be given about a small peak at 2θ ≈ 34◦. This peak is present in both the pure YBa2Cu3O7−δ and the Ba2YNbO6 doped thin films thus its presence can not be associated with the Ba2YNbO6 doping. The intensity is lower than that from the (004) and (008) Ba2YNbO6 peaks, and the Ba2YNbO6 doping level is only 5%mol. A peak at 2θ ≈ 34◦ in the YBa2Cu3O7−δ system could be associated with the (111) YBa2Cu3O7−δ peaks or with the (004) Y2O3. The first could be due to a minimal fraction of misaligned grains, a small fraction a grown grains can explain the peaks presence and it is commonly found at lower deposition temperature that the one adopted in these films. The second could be formed starting from a minimal Y imbalance in the target and considering that the peaks is also present in the pure YBa2Cu3O7−δ control sample, this minimal imbalance should be in the commercial pure YBa2Cu3O7−δ powder adopted in the target sintering process. Nevertheless, since the intensity is very small and the peak is not related to the Ba2YNbO6 doping, its presence does not influence the results. 4.2.2.2 Crystalline phases orientation The assessment of the in-plane texture is realized by φ scans. The technique is described in the previous chapter and the results obtained analysing a 5%mol Ba2YNbO6 doped film are summarized in figure 4.4. The (202) Ba2YNbO6 peak (χ = 45 ◦, 2θ = 30.03◦) reveal an in-plane align- ment since it matches both the (101) SrTiO3 (χ = 45 ◦, 2θ = 32.46◦) and the (202) YBa2Cu3O7−δ (χ = 57.06◦, 2θ = 27.62◦) peaks. This direct observation of a cube on cube growth of the Ba2YNbO6 with the YBa2Cu3O7−δ combined with the out-of-plane c-axis homogeneous orientation demonstrated in figure 4.3 suggest full heteroepitaxy between the YBa2Cu3O7−δ, the Ba2YNbO6 and the SrTiO3. A last point should be made on the diffraction peaks choice for the φ scans. Some authors perform the analysis on the (103) YBa2Cu3O7−δ (χ = 45◦, 2θ = 32.8◦) peak and not the (102) YBa2Cu3O7−δ used in this work. Unfortunately the first is extremely close to (101) SrTiO3 and in most analysers the two peaks 47 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results Figure 4.4: X-ray diffraction data from φ scans of (101) SrTiO3, (102) YBa2Cu3O7−δ, (202) Ba2YNbO6from a 5%mol Ba2YNbO6 doped film. Adapted from [5]. overlap. Since the intensity of the (101) SrTiO3 peak in much higher that the intensity of the (103) YBa2Cu3O7−δ peak choosing the latter as reference peak for the YBa2Cu3O7−δ in-plane orientation assessment could lead to misinterpre- tation. 4.2.2.3 Crystallographic matching and strain Once the phase orientations are known it is possible to propose a crystallo- graphic matching model and to evaluate the lattice mismatch of Ba2YNbO6 with YBa2Cu3O7−δ. In figure 4.5a is presented the matching of YBa2Cu3O7−δ with Ba2YNbO6 viewed along the b-axis ([010]Y BCO direction). In this view 3 unit cells of the cubic Ba2YNbO6 match 2 unit cells of YBa2Cu3O7−δ. The c-axis mismatch (related to YBa2Cu3O7−δ) can be calculated with the following equation: 3aBY NO − 2cY BCO 2cY BCO = +8.34% (4.1) 48 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results Figure 4.5: Crystallographic matching of YBa2Cu3O7−δ with Ba2YNbO6. a) b- axis view and b) c-axis view [5]. 49 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results In order to have a perfect lattice match along the c-axis the Ba2YNbO6 lattice should be compressed along the c-axis. In the doped film the (00l) peaks from the Ba2YNbO6 are indeed shifted to higher angles compared to the bulk values (figure 4.3). The resulting lattice parameter is of 0.830 nm ± 0.001 nm instead of the 0.844 nm ± 0.001 nm resulting from the powder scan (figure 4.2). This confirms the presence of a compressive strain of ≈ 1.7%, this strain is much lower than the theoretically derived mismatch. In addition the (00l) YBa2Cu3O7−δ diffraction peaks are slightly shifted towards lower 2θ values (in the opposite direction of the Ba2YNbO6 diffraction peaks) thus the YBa2Cu3O7−δ c-axis lattice parameter seems to be extended to higher values. A tensile strain of ≈ 0.4% can be derived from the lattice value of 1.173 nm calculated from the diffraction data presented in figure 4.3. Nevertheless the overall strain given by the Ba2YNbO6 contraction and the YBa2Cu3O7−δ extension is lower than the theoretical value thus an additional strain relief mechanism is most likely to be playing a matching role. The formation of misfit dislocations is one the possible strain relief mechanism and such defects formation has been previously reported in BaZrO3 doped YBa2Cu3O7−δ [126]. A semicoherent interface could also be responsible of the reduced strain levels in the crystalline structures. In figure 4.5b the matching of YBa2Cu3O7−δ with Ba2YNbO6 looking along the c-axis ([001]Y BCO direction) is presented. The matching between the YBa2Cu3O7−δ and the Ba2YNbO6 crystalline lattices is realised with 1 unit cell of Ba2YNbO6 and 2 unit cell of YBa2Cu3O7−δ. The averaged in-plane lattice mismatch (averaged along the a-axis and b-axis) is given by the following equa- tion: aBY NO − 2aY BCO 4aY BCO + aBY NO − 2bY BCO 4bY BCO = +9.42% (4.2) This large lattice mismatch, larger than the others studied non superconduct- ing phase additions, should be beneficial for low field pinning thanks to possible localized strain fields at the Ba2YNbO6-YBa2Cu3O7−δ interface [127], but it likely to produce self-assembled columnar arrays of defects that are shorter and wider than those produced with pinning addition with lower lattice mismatch [64], and similar to those produced with BaZrO3 addition [57]. 50 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results 4.2.3 Nanostructure analysis: Atomic Force Microscopy and Transmission Electron Microscopy X-ray diffraction analysis is a powerful tool to investigate the crystalline struc- tures and their orientation but they reveals little information on the nanostruc- turing of these phases. To investigate the nanostructure nanoscale microscopy has to be performed. In this section the surface topography is discussed by analysing atomic force microscopy images while the morphological distribution of the sec- ondary pinning phase and the distortion in the YBa2Cu3O7−δ lattice are studied with cross-section transmission electron microscopy. 4.2.3.1 Surface topography (Atomic Force Microscopy) In figure 4.6 are reported atomic force micrographs of a pure YBa2Cu3O7−δ and a 5%mol Ba2YNbO6 doped film surface. An analysis of the root mean square roughness performed on the surface to- pography of the 5 µm × 5 µm scans gives information on the quality of the two films in examination. The root main square roughness of the pure YBa2Cu3O7−δ thin film is ≈ 9.1 nm (figure 4.6a), a slightly lower value of ≈ 7.8 nm was calcu- lated for the 5%mol Ba2YNbO6 doped thin film (figure 4.6c). Carefully analysing both the 5 µm × 5 µm images it is evident that a small amount of particulate is present on the surfaces of both thin films, these superficial particles of diameter ≥ 250 nm were deposited on the growing film during the deposition and are a well known and studied in consequence of a pulsed laser deposition process in which the substrate is perpendicular to the plume and the target density is not optimal [105, 128]. Improved techniques could be adopted to reduce and eliminate this phenomenon [128–132]; in applications that require a perfectly clean surface adopting these methods is recommended. In this case the presence of the particulate is not a major issue and as shown later is not evidently reducing the overall superconducting properties. Furthermore since the particulate is present on the surfaces of both doped and undoped thin films it should not be related to the Ba2YNbO6 doping. Nevertheless in the next section higher quality films obtained from higher density target are discussed. The grain growth morphology is affected by the Ba2YNbO6 presence in the 51 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results Figure 4.6: Atomic Force Microscopy surface image of a pure YBa2Cu3O7−δ and a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films. a) 5 µm × 5 µm surface area of pure YBa2Cu3O7−δ; b) 1 µm × 1 µm surface area of pure YBa2Cu3O7−δ; c) 5 µm × 5 µm surface area of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ; d) 1 µm × 1 µm surface area of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ 52 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results doped YBa2Cu3O7−δ thin film. The first difference is the fact that in the Ba2YNbO6 sample the diameter of the growth grains appear to be slightly larger, the mean diameter measured in the undoped YBa2Cu3O7−δ film in ≈ 500 nm while the one measured in the Ba2YNbO6 doped YBa2Cu3O7−δ is ≈ 1 µm. This can be explained by a lower nucleation rate in the Ba2YNbO6 doped film. The second difference is the grain boundary shape, rough in Ba2YNbO6 doped YBa2Cu3O7−δ films and smooth in undoped YBa2Cu3O7−δ films. The main difference between the surface morphology of the doped film and that of the pure YBa2Cu3O7−δ is evidenced by the comparison of the 1µm ×1µm scans, figure 4.6b and 4.6d. The latter shows superficial particles of ≈ 15-20 nm in diameter that are absent in the image of taken from the pure YBa2Cu3O7−δ thin film. These superficial particles, that are to be attributed to the Ba2YNbO6 presence, are uniformly distributed over the sample surface and are not segre- gated at the grain boundary. This is a first experimental evidence that a uni- formly distributed array of defects based on Ba2YNbO6 is possibile given that the Ba2YNbO6 is able to grow inside a YBa2Cu3O7−δ crystal in the form of nanomet- ric inclusion and is not expelled from the growing YBa2Cu3O7−δ grains during the deposition (figure 4.6d). 4.2.3.2 Cross-section transmission electron microscopy The morphology of the Ba2YNbO6 inclusions was studied with the use of cross sectional imaging of the sample by transmission electron microscopy. In figure 4.7 is shown cross-section transmission micrography of a 5%mol Ba2YNbO6 doped thin film. Nanorods aligned with the c-axis of the films are evidently shown. The nanorods imaged are ≈ 10 nm in diameter and the spacing between two adjoining rods is ≈ 40 nm. A similar spacing is observed in the superficial par- ticles shown by the atomic force microscopy scan reported in figure 4.6d. These similar spacing values is a validation of the assumption that the particles observed on the surface by using the atomic force microscopy are related to the nanorods terminating at the film surface. The diameter of the particles observed on the surface with the atomic force microscopy appears to be larger then the diameter of the nanorods observed with the transmission electron microscopy. There are 53 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results Figure 4.7: Transmission electron microscope cross-sectional image of a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film. White arrows in the direction of the c-axis mark the position of self assembled Ba2YNbO6 nanorods parallel to the c-axis. Larger white arrows parallel to the YBa2Cu3O7−δ ab-planes mark the position of nanoparticles and of the interface between the substrate and the thin film. Inset shows selected area electron diffraction pattern taken from a region in proximity of a Ba2YNbO6 inclusion showing diffraction spot of Ba2YNbO6 YBa2Cu3O7−δ and SrTiO3. (TEM images from Prof. H Wang research group at Texas A&M University). 54 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results two plausible explanations that are not mutually exclusive but may be jointly responsible for the observed difference. The first explanation is found in the fact that the superficial particles might actually have a larger diameter because at the surface the ions mobility is higher and there may also be further agglomeration immediately after the deposition is terminated. The second possible explanation is related to differences in the resolution of the two methods of analysis used. In particular, the resolution of the atomic force microscopy depends on the diam- eter of the tip adopted. Tips with a larger diameter can make features appear larger than they are in reality. This aberration is absent in the transmission elec- tron microscopy where the captured images are not subjected to the described deformation. Additional information that can be derived from the nanorods spacing is the matching field. The matching field is the field value at which the spatial density of defects is equal to that of the flux lines passing through the superconductor. At this field value the pinning potential is the highest since in theory each and every flux line should be pinned by a single nanorod. Unfortunately, unlike a perfect Abrikosov lattice, it is not possible to identify a triangular or square arrangement of flux lines in superconductors with defects. In fact, the flux lines tend to change their spatial arrangement to overlap the defects distribution. Nevertheless it is at least possible to have an estimation of the matching field by using the mean spacing of the nanorods noticing that for a random distribution this tends to be higher than that derived for a square distribution and lower that that of a triangular distribution. The equations reported below refer to the matching field calculation derived for a square and a triangular distribution (note that Φ0 is the quantized flux ≈ 2.067× 10−15 Wb and d is the defects average spacing) H∗ = Φ0 d2 (4.3) H∗ = 2 √ 3Φ0 3d2 (4.4) Calculating the matching field for a defects average spacing of ≈ 40 nm applying the formula derived for a square distribution 4.3 gives a matching field H∗ ≈ 1.29 T. Applying the formula derived for triangular distribution 4.4 the 55 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results calculated matching field H∗ is ≈ 1.49 T. The applied magnetic field value in which the pinning potential of the analysed thin films should be higher is between 1.29 T and 1.49 T. In order to validate and complete the x-ray diffraction data reported in the previous section an electron diffraction pattern of the cross section is shown in the inset to figure 4.7. It is possible to observe a set of streaky diffraction dots for the (004) Ba2YNbO6. The presence of streaky dots is related to large Ba2YNbO6 c-axis distortions. These distortions were expected due to the fact that the x- ray analysis reported a compressive strain along the c-axis of the Ba2YNbO6 in the order of ≈ 1.7%. The lattice spacing d (004) calculated from the electron diffraction pattern is ≈ 0.214 nm, indicating a lattice parameter a ≈ 0.856 nm. This lattice parameter value, while being broadly consistent with the x- ray measurements and confirming the secondary phase as the Ba2YNbO6 double perovskite, does not confirm the compressive nature of the strain. In the cross sectional images the nanorod inclusions are not the only nanoin- clusions evidenced. A small fraction of nanoparticles are evidenced by the white arrows parallel to the ab-planes. The presence of both self assembled nanorods and nanoparticles is a feature which was already observed in BaZrO3 added YBa2Cu3O7−δ as well as RETa3O7 added YBa2Cu3O7−δ, the amount of nanopar- ticles is related to the deposition condition [133,134]. In the work of Boris Maiorov et al., published in 2009, an interesting discussion of the beneficial effects result- ing from simultaneous presence of both nanorods and nanoparticles is reported. In their work it is shows how the synergetic combination of such 1D defects (nanorods) and point defects (nanoparticles) can strongly enhance the flux pin- ning by preventing some low energy depinning mechanism to take place. In particular it is discussed how the flux jump from two adjacent nanorods with a double kink propagation mechanism is associated with a higher energetic barrier when a nanoparticle is introduced between the two adjacent nanorods or when the nanorods are not highly linear, continuous and are not all perfectly parallel to the c-axis. Analysing the higher magnification images (figure 4.8a and 4.8b) it is clear how the nanorods formed by the Ba2YNbO6 are wider (≈ 10 nm) than the RETa3O7 (≈ 5 nm) and BaZrO3 (≈ 7-8 nm), with a larger angle spread around the c- 56 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results Figure 4.8: Transmission electron microscope cross-sectional image of a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film. White arrows in the direction of the YBa2Cu3O7−δ c-axis mark the position of self assembled Ba2YNbO6 nanorods parallel to the c-axis. Arrows parallel to the YBa2Cu3O7−δ ab-planes mark the position of nanoparticles. The Ba2YNbO6 nanorods presence is indicated by the Moire fringes arising from lattice mismatch between Ba2YNbO6 and YBa2Cu3O7−δ. (TEM images from Prof. H Wang research group at Texas A&M University). 57 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results axis and shorter ≈ 100 nm as opposed to whole film thickness. Furthermore adding the simultaneous presence of nanoparticles and nanorods the resulting pinning landscape should be optimal in minimizing the low energy depinning mechanism efficiency. On the downside the fact that the nanorods tend to be wider indicates a lower rate of nucleation than RETa3O7, BaZrO3 and BaSnO3. The main consequence is a lower defects density at equal volumetric doping that should result in a lower pinning efficiency at high field values. One last thing to notice in this transmission electron microscopy study of the Ba2YNbO6 doped YBa2Cu3O7−δ sample is the evidence of how the YBa2Cu3O7−δ lattice is deformed by the nanorod inclusions. As perfectly pictured in figure 4.8b the YBa2Cu3O7−δ ab-planes are curved around the nanorods. The lattice mismatch between the YBa2Cu3O7−δ and the Ba2YNbO6 induce buckling of the YBa2Cu3O7−δ crystalline planes around the nanorods. Superconductivity is then expected to be depressed in the region surrounding the Ba2YNbO6 nanorods. The high possibility of additional pinning from strain fields in the vicinity of Ba2YNbO6-YBa2Cu3O7−δ interfacial regions is once more evidenced. 4.2.4 The superconducting properties: Tc, Jc(B) and Jc(B, θ) A preliminary study of the effects of Ba2YNbO6 inclusions in YBa2Cu3O7−δ can not be considered complete without the analysis of the main superconducting properties. Once the crystalline structure and the nanostructure of the sam- ple have been discussed it is necessary, in order to have an overall picture of Ba2YNbO6 performance, to evaluate the effective superconducting properties. Considering a power application such as superconducting cable or a more sophis- ticated application such as superconducting coils for magnets, the most important properties to evaluate are the transition temperature Tc, the variation of the crit- ical current density as a function of the applied magnetic field value Jc(B) and the variation of the critical current density as a function of the direction at which the magnetic field is applied Jc(B, θ). 58 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results Figure 4.9: Resistance variation with the temperature measured on a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film. The inset shows the region of the superconducting transition magnified. 4.2.4.1 Transition temperature, Tc In figure 4.9 is reported the resistance measured on a 50 µm width current track patterned on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film. The transition temperature Tc measured as the temperature at which the resistance disappears is Tc = 89 K. The reduction in the Tc is lower than expected. Reduced Tcs have been reported for most of the doped YBa2Cu3O7−δ films produced [57,68,118,119]. The Tc reduction is usually a function of the volumetric amount of the doping and the lattice distortion induced. It is important to notice that similar Tc reductions were observed in 5%mol BaZrO3 doped YBa2Cu3O7−δ thin films [67], but the volumetric amount of Ba2YNbO6 in 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films is higher than that of BaZrO3 in a 5%mol BaZrO3 doped YBa2Cu3O7−δ thin film. In particular a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film contains a 4.2%vol of Ba2YNbO6 while a BaZrO3 in a 5%mol BaZrO3 doped YBa2Cu3O7−δ thin film contains a 1.8%vol of BaZrO3 thus the Tc reduction registered in the Ba2YNbO6 doped scenario is lower than explected. 59 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results Two main informations can be derived from the observation of this small Tc reduction. First, the YBa2Cu3O7−δ planes buckling around the nanorods (pic- tured in the transmission electron microscopy figure 4.8b) is an effective strain sink capable of limiting the strain field propagation in a region close to the nanorods, for this reason the detrimental effects of the strain on the transition temperature are minimal. Second, as it was expected from earlier report on Nb doped YBa2Cu3O7−δ ceramics, the Nb do not substitute in the YBa2Cu3O7−δ and the Ba2YNbO6 do not react with the superconducting lattice. In other words the Nb remains enclosed in the stable Ba2YNbO6 and the only effects on the YBa2Cu3O7−δ crystalline structure are related to the lattice accommodation of the mismatch. 4.2.4.2 The critical current density Despite the slightly reduced Tc the critical current density Jc is improved over the whole field values window analysed. In figure 4.10 is reported Jc(B) measurements up to 1 T with the field applied parallel to the c-axis (B‖c) at 77 K for a pure YBa2Cu3O7−δ thin film and a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film. It is evident that the Jc is higher for the doped thin film and that at 1 T the Jc is a higher by a factor of ≈ 2. In the Jc(B) measurements there is a region for low values of the applied magnetic field in which the critical current Jc can be approximated with the following equation: Jc(B) ≈ Jc(0)−α∗B (4.5) The lower the α value the lower is the Jc reduction with increasing the applied magnetic field value. In the log-log plot shown in the inset are reported the calcu- tad α values for both the films. The lower α value calculated for the Ba2YNbO6 doped YBa2Cu3O7−δ thin film is an indication of a higher pinning potential. The calculated α value for the Ba2YNbO6 doped YBa2Cu3O7−δ sample is α = 0.37 that is lower than the α = 0.51 calculated for YBa2Cu3O7−δ films of this study. The measurements shown in figure 4.10 are exhaustive in describing the Ba2YNbO6 60 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results effects when the field is applied parallel to the c-axis, but in order to have a more complete picture of the Ba2YNbO6 efficiency as a pinning additive it is necessary to investigate the angular dependence of Jc. The angular dependence of the criti- cal current density at 77 K for an applied field value of 0.5 T is reported in figure 4.11. Figure 4.10: Critical current density variation with the applied magnetic field value measured on a pure YBa2Cu3O7−δ and a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films at 77 K and field applied parallel to the c-axis of the YBa2Cu3O7−δ. The inset shows data on a logaritmic axis and the calculated α values. The main characteristic of the Jc(B, θ) measured for the 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film is a strongly reduced anisotropy compared to the Jc(B, θ) measured for the undoped YBa2Cu3O7−δ thin film. The reduced anisotropy is related to an increase of the pinning potential when field is applied parallel to the Ba2YNbO6 nanorods. Since the nanorods are growing parallel to the c-axis of the YBa2Cu3O7−δ (figure 4.7) the pinning potential, and thus the Jc, is increased when the field is applied parallel to the c-axis of the thin film, 61 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results Figure 4.11: Critical current density variation with the direction of the applied magnetic field measured on a pure YBa2Cu3O7−δ and a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films at 77 K and magnetic flux density = 0.5 T which is orthogonal to the sample surface. Furthermore the anisotropy is reduced also by a reduction of the Jc when the field is applied parallel to the ab-plane of the YBa2Cu3O7−δ. The two changes in the pinning potential described can be related to the presence of Ba2YNbO6. The increment of the critical current density, Jc, when the field is applied parallel to the Ba2YNbO6 nanorods was expected and has been described for all the secondary phases forming c-axis aligned nanorods. In general when the field is applied parallel to the direction of any form of one dimensional defects, such as Ba2YNbO6 nanorods, the interaction between the flux lines and the defects is maximized and for this reason the pinning effectiveness of the defects is optimal. At the same time the interruption of the intrinsic pinning layer of the YBa2Cu3O7−δ (ab-planes) generated by the introduction of the Ba2YNbO6 nanorods together with the induced buckling of the YBa2Cu3O7−δ crystalline planes around the nanorods could be responsible for the observed reduction of the critical current density when the field is applied parallel to the ab-planes of the YBa2Cu3O7−δ. 62 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results An interesting feature that can be observed when comparing the effects of the Ba2YNbO6 introduction in the YBa2Cu3O7−δ to the angular dependence of the critical current density with the effects of the other pinning secondary phases (Ba(Zr,Y)O3, RETa3O7, BaSnO3 and Ba2YTaO7) is that the Ba2YNbO6 nanorods appears to be effective over a larger angular range producing a broader c-axis peak. The formation of shorted nanorods not perfectly oriented parallel to the c-axis of the YBa2Cu3O7−δ (discussed in the transmission electron microscopy analysis) could be the reason behind this increased effectiveness over the angles the magnetic field application. To conclude the Jc(B, θ) analysis a small discussion on the effects of the Ba2YNbO6 nanoparticles observed in transmission electron microscopy of the thin film cross-section (figure 4.8a) has to be done. Nanoparticles are usually related to an increase of the isotropic pinning [69]. A strong additional isotropic pinning increases the critical current density, Jc, independently of the angle at which the magnetic field is applied. Thus the presence of nanoparticles would have no effect on the shape of the Jc(B, θ). On the other hand the reduced pinning potential observed when the magnetic field is applied parallel to the ab- planes of YBa2Cu3O7−δ would suggest that an addition of isotropic pinning is absent. In reality the conclusion that can be made is that if nanoparticles are effectively adding an isotropic pinning to the Ba2YNbO6 doped YBa2Cu3O7−δ thin film this is lower than the reduction of the ab-pinning potential generated by the Ba2YNbO6 introduction. Furthermore it is important to remember that the synergetic combination of Ba2YNbO6 nanorods and nanoparticles can strongly enhance the pinning provided by the Ba2YNbO6 nanorods to the fields applied parallel to the c-axis of the YBa2Cu3O7−δ [134]. 4.2.5 Concluding remarks to the preliminary results In this section were reported the first results obtained by the addition of the Ba2YNbO6 perovskite pinning phase to YBa2Cu3O7−δ. The double perovskite Ba2YNbO6 results as a stable secondary phase that can be used with positive results as a pinning additive. Ba2YNbO6 is capable of forming c-axis oriented nanorods which result in shorter, wider nanorods than those generated by the 63 4. Ba2YNbO6 doped YBa2Cu3O7−δ: preliminary results other known pinning additives; furthermore the Ba2YNbO6 nanorods are oriented parallel to the c-axis of the YBa2Cu3O7−δ but with a larger splay angles range. For these reason the Ba2YNbO6 nanorods appear to be effective over a relatively large angular range of magnetic field application and not only when the field is applied parallel to c-axis of the YBa2Cu3O7−δ. 64 Chapter 5 Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters The study in this chapter is an evaluation of how the production process param- eters affect the formation of the secondary phase and the overall epitaxial quality of the thin films. Furthermore it is discussed how the epitaxial quality of the film and the morphology of the secondary phase influence the superconductive properties. 5.1 Ba2YNbO6 perovskite additions to YBa2Cu3O7−δ: the effects of the deposition rate on the nanos- tructure and the superconducting proper- ties The results obtained from the preliminary study on Ba2YNbO6 doped YBa2Cu3O7−δ thin films showed that the secondary phase Ba2YNbO6 can be added into a mixed YBa2Cu3O7−δ-Ba2YNbO6 pulsed laser deposition target to obtain pinning enhanced superconducting thin films. In order to investigate the real potential of the double perovskite Ba2YNbO6 a study of the influence of the two principal deposition parameters on the nanostructure and the supercon- ducting properties of Ba2YNbO6 doped YBa2Cu3O7−δ thin films was performed. 65 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters A set of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films was deposited by varying the deposition rate. The crystalline structure, nanostructure and su- perconducting properties of the thin films produced were analyzed. In this sec- tion are reported the analysis results that allows a better understanding of the Ba2YNbO6-YBa2Cu3O7−δ system overall and they demonstrate the possibility to tune the nano structure and the pinning properties of the Ba2YNbO6 doped YBa2Cu3O7−δ thin films by adjusting the deposition parameters. 5.1.1 Ba2YNbO6 doped YBa2Cu3O7−δ target preparation and thin film deposition The target used in the pulsed laser deposition of Ba2YNbO6 doped thin films analysed in this section is similar to the one that was described in the previous section. Since the preliminary study reported in the previous section evidenced a small fraction of particulates on the thin film surfaces (see 4.2.3.1), a variation to the production process of the target adopted for the analysis described in this section was implemented. Densification of the pulsed laser deposition target is a first countermeasure to the formation of particulate on the deposited thin film surface. In order to achieve a more dense target, once sintered a first time it was subjected to a subsequent regrinding. The powders obtained by the second grinding were again pressed in the form of a cylindrical target and the sintered at 950 ◦C in oxygen flow for additional 12hr. The laser pulses repetition rate was systematically modified to study its ef- fects on the nanostructure and pinning properties of the deposited thin films. The laser pulse repetition rate variation changes the maximum time allowed to the ions migration on the surface of a growing film before new higher energy ions are deposited from a subsequent laser pulse. A higher rate increases the deposition rate but lowers the “free migration” time. This parameter has a direct effect on the migration time and thus a direct effect on the crystalline and nanostructure of the growing films. In particular repetition rates of 1, and 10 Hz and substrate temperatures of 780 ◦C were investigated. In the table below 5.1 are summa- rized the pulsed laser deposition parameters adopted for the deposition of the Ba2YNbO6 doped YBa2Cu3O7−δ thin films analysed in this section. 66 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Parameter Value Substrate Temperature 780 ◦C Chamber Pressure 0.3 mbar flowing O2 Laser Fluence 2 Jcm−2 Repetition Rate 1, 10 Hz Number of Pulses 4500 Annealing Time 1 hr Annealing Temperature 520 ◦C Annealing Pressure 500 mbar O2 Table 5.1: Pulsed laser deposition parameters All the Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited were measured to be of ≈ 0.4 µm thickness. Some additional thicker samples of ≈ 0.7 µm thickness where also deposited using 8000 laser pulses. 5.1.2 Crystalline structure analysis: x-ray diffraction data Similarly to the crystalline phases characterization realised for the first Ba2YNbO6 doped YBa2Cu3O7−δ system, reported in the previous section, the crystalline structures of all the samples produced for the study of the effects of the depo- sition parameters were analysed with an extensive use of the x-ray diffraction technique. 5.1.2.1 Crystalline phases identification, phases orientation The x-ray diffraction data collected in the Bragg-Brentano geometry to identify the crystalline phases as well as the in-plane texture analysis by examination of φ scans don’t add additional information to the ones gathered in the previous section. The results obtained at the different growth condition closely matches each other. For this reason it is hard to understand if there are differences in the crystalline structures of the Ba2YNbO6 doped YBa2Cu3O7−δ thin films that were deposited in the different conditions described. At a first x-ray diffraction analysis all the samples appears to have similar crystalline features. The only diffraction peaks evidenced by x-ray diffraction analysis in the Bragg-Brentano geometry (θ − 2θ scan) are those related to the 67 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters (00l) planes from the YBa2Cu3O7−δ, the Ba2YNbO6 and the SrTiO3. Further- more also the φ scans closely matches for the different thin films and confirms the in-plane orientation discussed in the previous section. In conclusion the θ− 2θ and the φ scans are not capable at identifying differ- ences in the crystalline structures and varying laser pulses repetition rate do not change the fact that the Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited con be mainly described as epitaxial YBa2Cu3O7−δ films with the presence of a cube on cube oriented Ba2YNbO6 perovskite within the YBa2Cu3O7−δ. 5.1.2.2 Epitaxy quality: rocking curve Even if the standard x-ray diffraction analysis does not detect differences in the crystalline structure, a first classification can be made on the basis of the epitaxial quality. The quality of the YBa2Cu3O7−δ epitaxy with the SrTiO3 can be anal- ysed by the rocking curve detection. The half maximum width of the diffraction peaks resulting from a rocking curve scan is an indication of the angular spread of the c-axis orientation. The lower the spread, the higher is the c-axis alignment. Thin Film SrTiO3 YBa2Cu3O7−δ Ba2YNbO6 (002) (005) (004) YBa2Cu3O7−δ (1 Hz) 0.10 0.18 YBa2Cu3O7−δ (10 Hz) 0.12 0.18 doped YBa2Cu3O7−δ (1 Hz) 0.11 0.15 1.0 doped YBa2Cu3O7−δ (10 Hz) 0.10 0.20 2.1 Table 5.2: Full Width Half Maximum values of the SrTiO3 (002), YBa2Cu3O7−δ (005) and Ba2YNbO6 (004) rocking curves measured on pure YBa2Cu3O7−δ and 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate tem- perature of 780 ◦C and repetition rate of 1 and 10 Hz. In the table 5.2 is reported the full width half maximum (FWHM) val- ues of the rocking curves around the (002) SrTiO3, (005) YBa2Cu3O7−δ and (004) Ba2YNbO6 planes of the different samples. The FWHM values of the YBa2Cu3O7−δ do not reveal an evident pattern and the most relevant informa- tion is that the YBa2Cu3O7−δ c-axis orientation do not appear to be influenced by the deposition rate in absence of the Ba2YNbO6 phase. The Ba2YNbO6 FWHM values instead shows that the sample deposited at higher deposition rate (higher 68 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters laser repetition rate) shows a larger FWHM value than the one obtained for the sample deposited at lower deposition rate. Larger “free migration” time allow the formation of a Ba2YNbO6 crystalline lattice with a higher c-axis alignment. One thing to notice is that since the Ba2YNbO6 is growing inside a YBa2Cu3O7−δ lattice its orientation quality is strictly related to YBa2Cu3O7−δ. More precisely the Ba2YNbO6 lattice does not present a better c-axis orientation alignment than the YBa2Cu3O7−δ lattice in which it is formed, on the other hand it is also true that a poorly oriented Ba2YNbO6 can be formed in a perfectly oriented YBa2Cu3O7−δ matrix. While it is possible to use a technique developed to investigate the c-axis alignment quality of thin films to also analyse that of the Ba2YNbO6 nanopar- ticles and nanorods it is not possible to directly compare the FWHM obtained for the Ba2YNbO6 nanoinclusions with those obtained with the YBa2Cu3O7−δ thin films. In general, since the Ba2YNbO6 is a dopant and its quantity is only a fraction of the YBa2Cu3O7−δ, the intensity x-ray diffraction peaks re- lated to the Ba2YNbO6 are ≈ 2 orders of magnitude lowers than those related to the YBa2Cu3O7−δ. Furthermore the morphological nature of the Ba2YNbO6, nanorods and nanoparticles, is responsible for an additional broadening of the Ba2YNbO6 peaks in the diffraction pattern that is related to the size of the phase and not the strain or orientation. For these reason the FWHM values obtained from the Ba2YNbO6 should be taken as a comparison between the different c-axis degrees of the Ba2YNbO6 phase only. 5.1.2.3 Epitaxy quality: reciprocal space maps The rocking curve were used to investigate the quality of the c-axis alignment for both the Ba2YNbO6 and YBa2Cu3O7−δ separately but while a first information on the Ba2YNbO6 phase was revealed it was impossible to obtain additional information on the YBa2Cu3O7−δ . In order to further investigate the effects of the deposition parameters on the overall epitaxial quality of the YBa2Cu3O7−δ and Ba2YNbO6 nanoparticles reciprocal space maps were gathered for a pure YBa2Cu3O7−δ thin film, and two Ba2YNbO6 doped YBa2Cu3O7−δ thin films. In particular the Ba2YNbO6 doped films were deposited adopting different repetition 69 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters rate of the ablating pulsed laser, 1 and 10 Hz, but maintaining the same substrate temperature of 780 ◦C; the pure YBa2Cu3O7−δ was deposited at 1 Hz and the same substrate temperature. Figure 5.1: X-ray diffraction reciprocal space map measured on a pure YBa2Cu3O7−δ thin film. In figure 5.1 is reported the reciprocal space map gathered from a pure YBa2Cu3O7−δ sample. There are three main diffraction peaks, the (103) of the SrTiO3 and the (109) - (019) of the YBa2Cu3O7−δ are overlapping each other while is clearly visible the (108) - (018) of the YBa2Cu3O7−δ. The YBa2Cu3O7−δ peaks are actually formed by two different peaks related to the inevitable twins formation. It is impossible to distinguish between the (108) and the (018) YBa2Cu3O7−δ peaks because the two peaks are too close in the reciprocal space that they appear as a single large peak, the theoretical peaks position indicated in the reciprocal space maps are well between the peaks boundary and the center of the peak is between this two theoretical position as expected in twin formations. It is possible from the reciprocal space maps to calculate the lattice parameters directly from the coordinate of a peak in the reciprocal space if the (h,k,l) index is known. Remembering that the reciprocal space coordinates (qx and qz) are the reciprocal of the lattice spacing (dx and dz) the lattice parameters a,b and c for 70 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters a (h,0,l) and a (0,k,l) can be calculated applying the following formulas: for a (h,0,l) peak a = h qx (5.1) c = l qz (5.2) for a (0,k,l) peak b = k qx (5.3) c = l qz (5.4) In figure 5.1 it is easier to calculate the YBa2Cu3O7−δ lattice parameters from the (108)-(018) peaks since these are not overlapping with the any other phases peak. Its important to notice that similarly to the lattice mismatch calculated in the last section since it not possible to separate the (108) peak from the (018) the lattice parameter calculated from the qx coordinate of the reciprocal space map refers to an average value between the a and b lattice parameters. The average ab calculated from the a and b bulk values (a = 0.382 nm and b = 0.388 nm), is abbulk = 0.385 nm and it is perfectly in line with the average value calculated from the (108)-(018) YBa2Cu3O7−δ taken from the reciprocal space map in figure 5.1 where a qx = 2.60 nm −1 also gives a ab = 0.385 nm. Similarly the YBa2Cu3O7−δ c lattice parameter bulk value is c = 1.168 nm and from a qz = 6.85 nm −1 a c = 1.168 nm is calculated. The YBa2Cu3O7−δ thin film while being epitaxial with the SrTiO3 substrate it is relaxed and the lattice parameters have the same value reported in literature for bulk YBa2Cu3O7−δ. Focusing on the (103) SrTiO3 diffraction peak at qx = 2.56 nm −1 and qz = 7.68 nm−1, it is evident that additional diffraction peaks of (103) are present. The presence of these additional peaks is related to the use of a x-ray incident beam that is not monochrome and, even if the main quantity of x-rays are the Cu Kα with λ = 0.15405 nm, there are also secondary x-rays generated at different λ that are the cause of the additional diffraction peaks presence. The intensity of the secondary peaks is lower and these are only visible for the high intensity SrTiO3 71 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters diffraction peaks. Another feature related to the (103) SrTiO3 is a diagonal line also only visible for the high intensity diffraction generated by the substrate, also this line is an artifact generated by the absence of a monochromator in particular the line is generated by non copper x-rays emissions. Figure 5.2: X-ray diffraction reciprocal space map measured on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films. a) Film deposition rate of 1 Hz; b) Film deposi- tion rate of 10 Hz. 72 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters In figure 5.2 are reported the reciprocal space map gathered from a Ba2YNbO6 doped YBa2Cu3O7−δ sample deposited at 1 Hz repetition rate (figure 5.2a) and a Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited at 10 Hz repetition rate (figure 5.2b). In addition to the three main diffraction peaks also present in figure 5.1, that are related to the YBa2Cu3O7−δ and the substrate, the (103) Ba2YNbO6 peak is also present. Comparing the image in figure 5.1 with the images in figure 5.2 the (103) Ba2YNbO6 peak is not the only evident difference, a broadening of the YBa2Cu3O7−δ related peaks is a second clear effect of the Ba2YNbO6 doping. The broadening of the YBa2Cu3O7−δ peaks in the Ba2YNbO6 doped samples could be related to the buckling of the crystalline YBa2Cu3O7−δ planes around Ba2YNbO6 nanoinclusion discussed in the last section. The fact that overall the lattice parameter of the YBa2Cu3O7−δ does not seem to change confirms that the broadening is related to a local distortion, like planes buckling. Finally a direct indication of the effects of the deposition parameters on the crystalline structure can be obtained by the comparison of the reciprocal space maps of the two different Ba2YNbO6 doped YBa2Cu3O7−δ thin films (figure 5.2a and 5.2b). First the (103) Ba2YNbO6 peak shape and intensity appear to be different in the two samples. In particular the Ba2YNbO6 peak of the 10 Hz sample (figure 5.2b) has a lower intensity and a long tail is present. These two features could be evidence of the existence of a fraction of randomly oriented Ba2YNbO6 that is absent in the 1 Hz sample (figure 5.2a). From the (103) Ba2YNbO6 peak of the 1 Hz sample it is possible to calculate the lattice param- eters a of the Ba2YNbO6 phase in the planar direction and in the YBa2Cu3O7−δ c-axis direction from the qx and qz values respectively. Applying the equations 5.1 and 5.2 on the qx = 2.32 nm −1 and qz = 7.22 nm−1 taken from figure 5.2a the lattice parameters a = 0.430 nm along the planar direction and c = 0.415 nm are calculated. The Ba2YNbO6 phase lattice parameter bulk value is a = 0.422 nm, and comparing this value with those calculated from the reciprocal space map it is clear that the Ba2YNbO6 is subjected to a compressive strain along the YBa2Cu3O7−δ c-axis direction and an elongation strain along the YBa2Cu3O7−δ ab-planes direction. The deformation calculated for the two different direction are + 1.9 % along the ab-planes direction and - 1.66 % along the c-axis direction. Considering the morphological nature of the Ba2YNbO6 inclusions it is possible 73 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters to explain why the compressive strain is only acting in the c-axis direction even if the lattice mismatch calculated from the bulk values (in-plane lattice mismatch = + 9.42 % and c-axis lattice mismatch = + 8.34 %) indicate a compressive strain in all directions. The Ba2YNbO6 is mainly forming nanorods aligned parallel to the c-axis of the YBa2Cu3O7−δ thus the interface between the Ba2YNbO6 and the YBa2Cu3O7−δ is developing prevalentely along the c-axis direction; for this reason the minimization of the mismatch along the c-axis direction is energeti- cally more relevant. The difficulty to identify unique qx and qz values in the low intensity peak of the Ba2YNbO6 measured from the the 10 Hz deposited thin film makes not possible the repetition of the calculations performed on the 1 Hz deposited sample. Analysing the (108)-(018) YBa2Cu3O7−δ peaks in figure 5.2 additional infor- mation on the effects of the deposition rate on the YBa2Cu3O7−δ lattice is ob- tained. The broadening of the peak related to the Ba2YNbO6 doping is slightly more relevant in the Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited with a laser repetition rate of 10 Hz. This indicates that the YBa2Cu3O7−δ region affected by lattice distortions (buckling of the crystalline planes) is larger when the deposition rate is higher. There are two possible explanation of this evidence, in particular either a larger region of YBa2Cu3O7−δ may be affected by lattice distortions around each nanoinclusion or the Ba2YNbO6 nanoinclusions density may be higher. The phase that is more directly affected by diffusion processes, and thus by the deposition rate, is the Ba2YNbO6. The Nb ions must in fact diffuse on larger region of the surface in order to form the Ba2YNbO6 nanoinclu- sions. For this reason the higher density of Ba2YNbO6 inclusions is more likely the cause of the increased lattice distortion. In figure ?? is shown a direct comparison of the (108)-(018) YBa2Cu3O7−δ peaks in the different films, the indication of the peaks’ position as well as the half maximum intensity isolines allow direct visualization of the effects of the presence of Ba2YNbO6 on the YBa2Cu3O7−δ crystalline structure. The broadening of the YBa2Cu3O7−δ peaks is not the only difference, the lat- tice parameters measured from the reciprocal space map of the Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited at 10 Hz is different from the one measured from the film deposited at 1 Hz. In particular, from the (108)-(018) YBa2Cu3O7−δ peak 74 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.3: Half maximum intesity insoline and peak position comparison for the (108)(018) YBa2Cu3O7−δ peaks from figure 5.1 and 5.2. of the reciprocal space map measured on the film deposited at 1 Hz (qx = 2.58 nm−1 and qz = 6.85 nm−1) the lattice parameters ab = 0.388 nm and c = 1.168 nm are calculated, while from the (108)-(018) YBa2Cu3O7−δ peak of the recip- rocal space map measured on the film deposited at 10 Hz (qx = 2.60 nm −1 and qz = 6.83 nm −1) are calculated the lattice parameters ab = 0.385 nm and c = 1.171 nm. In conclusion the YBa2Cu3O7−δ in the film deposited at 1 Hz appears to better match the SrTiO3 substrate while not being affected on the overall c-axis spacing by the Ba2YNbO6 presence at all. On the other hand the YBa2Cu3O7−δ in the film deposited at 10 Hz maintains the ab bulk value (complete relaxation) but the c lattice parameter is slightly larger. This is an indication that the lattice distortion generated by the Ba2YNbO6 doping is influencing a region of the YBa2Cu3O7−δ lattice that is large enough to be noticed when measuring the overall c lattice parameter value. The information obtained from the analysis of the (108)-(018) YBa2Cu3O7−δ peak measured in the reciprocal space is summarised in table 5.3. 75 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Sample Peak coordinates Lattice parameters qx / qz ab 2θ / ω c Undoped YBa2Cu3O7−δ 2.60 nm−1 / 6.85 nm−1 0.385 nm 68.69 ◦ / 13.58 ◦ 1.168 nm Doped YBa2Cu3O7−δ (1 HZ) 2.58 nm−1 / 6.85 nm−1 0.388 nm 68.60 ◦ / 13.66 ◦ 1.168 nm Doped YBa2Cu3O7−δ (10 Hz) 2.60 nm−1 / 6.83 nm−1 0.385 nm 68.59 ◦ / 13.42 ◦ 1.171 nm Table 5.3: (108)-(018) YBa2Cu3O7−δ peak analysis. 5.1.3 Nanostructure analysis: Atomic Force Microscopy and Transmission Electron Microscopy The analysis of the nanostructure reveals interesting differences between the Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited at 10 Hz and those de- posited at 1 Hz. Similarly to what was done in the previous section the sur- face topography is discussed by analysing atomic force microscopy images while the morphological distribution of the secondary pinning phase and the distortion in the YBa2Cu3O7−δ lattice are studied with cross-section transmission electron micrography. 5.1.3.1 Surface topography (Atomic Force Microscopy) A comparison between the surface topography of samples deposited at different rates is reported in figure 5.4, in particular the surfaces of samples of different thickness (≈ 700 nm and 400 nm) deposited at 1 and 10 Hz are investigated. The first evident effect of increasing the laser repetition rate is a the presence of smaller growth grains. Comparing the surfaces of sample with similar thickness (figure 5.4a and 5.4b or 5.4c and 5.4d) it is clearly shown that the crystalline grains are smaller in the samples grown at 10 Hz repetition rate. As expected, a reduced “free migration” time can then be associated with a reduced accretion of the grains. The first result is that the thin film deposited with a repetition rate of 10 Hz is formed of a higher number of smaller grains when compared to a thin film deposited with a repetition rate of 1 Hz. The direct consequence of 76 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.4: Atomic Force Microscopy surface image (5 µm × 5 µm) of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 780 ◦C, repetition rate of 1 Hz and 10 Hz with a thickness of 700 nm and 400 nm. a) 1 Hz and 700 nm; b) 10 Hz and 700 nm; c) 1 Hz and 400 nm; d) 10 Hz and 400 nm. 77 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters the morphology described is clearly shown by the comparison of the thicker films surfaces, figure 5.4a and 5.4b. It is in fact clear that the connectivity shown by the grains in figure 5.4b is reduced. Furthermore by the same comparison it appears that the surfacial morphology of the thin film deposited at the higher repetition rate is a scaled down version of the one shown by the film deposited at 1 Hz. A similar analysis could be made comparing the surface morphologies of the thinner films (figure 5.4c and 5.4d) but it is difficult to visualise the grains boundaries in the smooth surface of the thinner film deposited at 1 Hz. With a more accurate analysis of the images another feature can be deduced. In particular, as expected, the grain dimension is not affected by the film thick- ness. Focusing on the thin films deposited with a repetition rate of 10 Hz (figure 5.4b and 5.4d) the grains dimension appears to be similar. However the shape of the boundaries turn from the well defined straight lines of the thinner film (figure 5.4d) to the more chaotic boundaries of the thicker sample (figure 5.4b). This is a evidence that the grains connectivity is also reduced when the thickness is increased. A fine dispersion of superficial nanoparticles is shown in figure 5.4c. Simi- lar nanoparticles where also evidenced in the previous section (figure 4.6d) and where attributed to the Ba2YNbO6 presence. It is impossible to picture the same small nanoparticles in atomic force micrography taken on samples with a higher roughness (figure 5.4a and 5.4b) because the increased scale flatten the small topographic features. It is unclear why the superficial nanoparticles are absent (or invisible) in the thinner film deposited with a repetition rate of 10 Hz (figure 5.4d). Since its surface is shown in a similar scale to the 1 Hz film the absence of the nanoparticles is an additional indication of a morphological difference induced by the different deposition rate. The root mean square roughness values calculated from the surface topogra- phies shown in figure 5.4 are reported in table 5.4. The calculated roughness values are higher in the 700 nm thick films then in the 400 nm thick films. How- ever, there is no evidence of a direct influence of the repetition rate on these values. A last note on this surface topography study has to be made on the target improved quality. As stated earlier a densified target was prepared to avoid the 78 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Repetition Rate, Thickness Root mean square roughness 1 Hz, 400 nm 2.12 nm 10 Hz, 400 nm 7.12 nm 1 Hz, 700 nm 34.14 nm 10 Hz, 700 nm 21.41 nm Table 5.4: Root mean square roughness calculated from the surface topography of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 780 ◦C, repetition rate of 1 Hz and 10 Hz with a thickness of 700 nm and 400 nm. presence of particulates on the film surface. Comparing the surface topographies shows in this section with the earlier study reported (section 4.2.3.1) it is evident that the changes in the production process of the pulsed laser deposition targets solved the particulate issue. 5.1.3.2 Cross-section transmission electron microscopy The cross-section images taken using the transmission electron microscopy clearly show the strong influence of the laser pulse repetition rate on the nanostructure of the Ba2YNbO6 doped YBa2Cu3O7−δ thin films. In figure 5.5 are reported the transmission electron micrographs taken on the films grown at 1 Hz and 10 Hz at different magnification. Columnar de- fects are shown with a good contrast only in the film deposited at 1 Hz (fig- ure 5.5a and 5.5b) and are indicated by arrows parallel to the c-axis of the YBa2Cu3O7−δ. In the film deposited at 10 Hz some shadows parallel to the c- axis of the YBa2Cu3O7−δ are present and could be related to the presence of Ba2YNbO6 nanorods. Nevertheless the contrast in figure 5.5c and 5.5d is lower and these columnar defects are not clearly identified. The reduced contrast can be related to a lower orientation degree as well as to larger distortion. Once more the Ba2YNbO6 inclusions nanostructure is shown to be the phase influenced more by the deposition rate. Focusing on the images related to the Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited at laser pulse rate of 1 Hz, it is possible to analyse the morphology of the nanorods parallel to the c-axis of the YBa2Cu3O7−δ. The nanorods diam- eter is ≈ 10 nm and their spacing is ≈ 40 nm. Similarly to the case reported in 79 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.5: Transmission electron microscope cross-sectional image of a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited with a laser pulse rates of 1 Hz (a, b) and 10 Hz (c, d). White arrows in the direction of the YBa2Cu3O7−δ c-axis mark the position of self assembled Ba2YNbO6 nanorods parallel to the c-axis. Arrows parallel to the YBa2Cu3O7−δ ab-planes mark the position of nanoparticles. (TEM images from Prof. H Wang research group at Texas A&M University). 80 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters the last section where a Ba2YNbO6 doped YBa2Cu3O7−δ deposited with a laser pulse rate of 5 Hz showed same structures with similar spacing. Repeating once more the matching field calculation also the films deposited at 1 Hz should have a matching field value between 1.29 T and 1.49 T. Comparing figure 5.5a with figure 4.8a two differences are evident: the nanorods in the film deposited at a laser pulse rate of 1 Hz appears to be continuous and longer than the ≈ 100 nm length of the nanorods showed in the film deposited at 5 Hz; in the film deposited at 1 Hz there is no sign of the nanoparticles that where shown in the 5 Hz case. The film deposited at a laser pulses rate of 10 Hz (figure 5.5c and 5.5d) is the one that differs the most. The nanorods are not clearly visible but a large number of nanoparticles is present. In general from the transmission electron microscopy analysis this film appears to have the less ordered YBa2Cu3O7−δ lattice and an almost completely disordered Ba2YNbO6 nanostructure. It is interesting to see that the nanoparticles and nanorods presence is largely influenced by the laser pulses rate. The nanoparticles are absent in the 1 Hz films, their presence increase increasing the laser pulse rate up to the 10 Hz film where they appear to be the dominant nanoinclusion. On the other hand the nanorods are long, linear and almost perfectly parallel to the YBa2Cu3O7−δ c-axis in the film deposited at a laser pulse of 1 Hz; are shorter (≈ 100 nm) with a large angle spread to the YBa2Cu3O7−δ c-axis in the film deposited at a laser pulse of 5 Hz (see section 4.2.3.2); are distorted and not clearly imaged in the film deposited at a laser pulse of 10 Hz. The Ba2YNbO6 inclusions generate distortion in the YBa2Cu3O7−δ lattice. The influence of the laser pulse rate on the nanostructure of the Ba2YNbO6 inclusions is then transferred to the YBa2Cu3O7−δ lattice distorted by those in- clusions. In transmission electron microscope images of the Ba2YNbO6 doped YBa2Cu3O7−δ films deposited at pulse rates of 1 Hz and 5 Hz the YBa2Cu3O7−δ crystalline planes parallel to the ab direction appear to be distorted only in the re- gions close the nanorods and the planes buckling in this small region appear as an effective strain sink. On the other hand a similar buckling of the planes does not appear to be effective in respect of the high density of the nanoparticles growing in the Ba2YNbO6 doped YBa2Cu3O7−δ deposited at 10 Hz. The YBa2Cu3O7−δ lattice appears to be distorted with continuity and the entire lattice seems to be 81 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters affected by the Ba2YNbO6 presence. The high density of nanoparticles appears to be the most likely cause. In particular, the reduced distance between the Ba2YNbO6 inclusions do not allow full relaxation of the YBa2Cu3O7−δ lattice. 5.1.4 The superconducting properties: Tc, Jc(B) and Jc(B, θ) The crystalline structure and the nanostructure of Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at different rates have been analysed. In particular the nanos- tructure was found to be heavily influenced by the laser pulse rate. In this sec- tion the superconducting transport properties (Tc, Jc(B) and the Jc(B, θ)) are analysed and related to the nanostructural feature evidenced. In addition to the Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at 1 Hz and 10 Hz also films deposited at 5 Hz where characterized. The crystalline structure and nanostruc- ture morphology of Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at 5 Hz was analysed in the preliminary study. Nevertheless, since the transport proper- ties characterisation performed in the preliminary study presented was limited to low field values (Jc(B) up to 1 T and Jc(B, θ) measured at 0.5 T), the transport properties of new films deposited at laser pulses repetition rate of 5 Hz adopting the densified pulsed laser deposition target are characterised in order to obtain a more complete picture of the effects of the deposition rate. 5.1.4.1 Transition temperature, Tc The transition temperature does not appear to be effected by the laser pulses repe- tition rate adopted. The Tc measured from all the Ba2YNbO6 doped YBa2Cu3O7−δ thin films produced is reduced by a small amount (≈ 3 K) when compared to the Tc measured from a pure YBa2Cu3O7−δ thin film. In figure 5.6 are reported the resistance measured on a 50 µm width current track patterned on pure YBa2Cu3O7−δ thin film and 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a laser pulses repetition rate of 1, 5 and 10 Hz. The transition temperature Tc, measured as the temperature at which the resistance disappears, is reported in table 5.5. The Tc reduction is in line with the one measured in the preliminary study (section 4.2.4.1), thus the same considerations can be done and the Ba2YNbO6 82 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.6: Resistance variation with the temperature measured on a pure YBa2Cu3O7−δ thin film deposited at a substrate temperature of 780 ◦C and laser pulses repetition rates of 1 Hz and on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 780 ◦C and laser pulses repe- tition rates of 1, 5 and 10 Hz. 83 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters stability within the YBa2Cu3O7−δ matrix is confirmed. Furthermore it is impor- tant to notice that despite the higher disorder pictured by the transmission elec- tron microscopy the transition temperature measured on the Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited at 10 Hz is almost equal to the one measured on the thin film deposited at a laser pulses repetition rate of 1 Hz. In conclusion the fact that the higher defects density is not related to a further reduction of the transition temperature could be the evidence that connected regions of the YBa2Cu3O7−δ lattice only marginally distorted by the Ba2YNbO6 inclusion may be present also in the thin films deposited at laser pulses repetition rate of 10 Hz. Thin Film Transition Temperature Tc pure YBa2Cu3O7−δ (1 Hz) 91.5 K Ba2YNbO6 doped YBa2Cu3O7−δ (1 Hz) 87.8 K Ba2YNbO6 doped YBa2Cu3O7−δ (5 Hz) 88.8 K Ba2YNbO6 doped YBa2Cu3O7−δ (10 Hz) 88.0 K Table 5.5: Transition Temperature Tc collected from the resistance variation with temperature measurements shown in figure 5.6. 5.1.4.2 The critical current density Once again despite the slightly reduced transition temperatures Tc the first ev- ident effect of the Ba2YNbO6 doping is improved critical current densities Jc values over the whole field range analysed. In figure 5.7 are reported Jc(B) measurements up to 6 T with the field applied parallel to the c-axis (B‖c) at 77 K for a pure YBa2Cu3O7−δ thin film and 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a laser pulses repetition rate of 1, 5 and 10 Hz. The Jc(B) values measured on the Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at laser pulses repetition rate of 1 and 5 Hz are similar. This is an evidence that the pinning potential is similar when the field is applied parallel to the YBa2Cu3O7−δ c-axis. The Jc values measured on the film deposited at 10 Hz are lower then those measured on the other Ba2YNbO6 doped YBa2Cu3O7−δ films in the low field region (B ≤ 2 T) but are the highest when the applied field is above 2 T. A possible explanation can be found in the higher density of Ba2YNbO6 nanoinclu- 84 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.7: Critical current density variation with the applied magnetic field value measured on a pure YBa2Cu3O7−δ thin film deposited at a substrate temperature of 780 ◦C and laser pulses repetition rates of 1 Hz and on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz. The field is applied parallel to the c-axis of the YBa2Cu3O7−δ, T = 77 K. The data are on a log-linear plot (a) and a log-log plot (b). 85 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters sions. The matching field values (1.29 T ≤ H∗ ≤ 1.59 T) calculated from the Ba2YNbO6 nanorods were similar in the 1 Hz and 5 Hz films. Unfortunately, due to the high distortions resulting in low transmission electron image contrast it was not possible to calculate a matching field for the 10 Hz film. However, since the lower spacing of the inclusion was evidenced, a higher matching field value was expected. Figure 5.8: Pinning force calculated from data reported in figure 5.7 To better visualize the effect of different matching field values in figure 5.8 is reported the pinning force calculated as Jc(B) × B from the data reported in figure 5.7. The pinning force maximum for the Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at 1 and 5 Hz are located at a field valued of ≈ 1.4 T. It is worth noticing that this value is perfectly in line with the matching field values predicted from the nanorods spacing measured on the images acquired with the electron transmission microscopy. The fact that the pinning force max- imum for the Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited at 10 Hz is shifted towards higher field values confirm the higher matching field value that was expected from the higher Ba2YNbO6 inclusion density. Table 5.6 reporting the α values calculated from data in figure 5.7 complete 86 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Thin Film α pure YBa2Cu3O7−δ (1 Hz) 0.47 Ba2YNbO6 doped YBa2Cu3O7−δ (1 Hz) 0.32 Ba2YNbO6 doped YBa2Cu3O7−δ (5 Hz) 0.37 Ba2YNbO6 doped YBa2Cu3O7−δ (10 Hz) 0.39 Table 5.6: α calculated from the data reported in figure 5.7. the Jc(B) analysis. The α values are reduced when the Ba2YNbO6 is added to YBa2Cu3O7−δ and the higher value is the one calculated for the Ba2YNbO6 doped YBa2Cu3O7−δ deposited at a laser pulse repetition rate of 10 Hz is α = 0.39 which is lower than α = 0.47 calculated for the pure YBa2Cu3O7−δ sample. The reason why the highest α value is the one calculated the 10 Hz Ba2YNbO6 doped YBa2Cu3O7−δ film is that the pinning landscape generated in this film is effective at high field more then it is at low field and that α is calculated in the very low field regime (B ≤ 1 T). At low field values the pinning potential is not fully developed since the high density of nanoinclusions may reduce the energy associated with flux jumps between adjoining defects reducing the pinning potential. To complete the analysis of the effects that different laser pulses repetition rates have on the nanostructure and properties of Ba2YNbO6 doped YBa2Cu3O7−δ thin films the critical current density variation as a function of the external mag- netic field direction is measured. In figure 5.9 are reported Jc(B, θ) measurements at 77 K and 0.5, 1 T for a pure YBa2Cu3O7−δ thin film and 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a laser pulses repetition rate of 1, 5 and 10 Hz. The first important result is that even if all the Ba2YNbO6 doped YBa2Cu3O7−δ thin films shows an higher Jc value when the field is applied parallel to the YBa2Cu3O7−δ c-axis the only thin film that shows higher Jc values than pure YBa2Cu3O7−δ when the field is applied parallel to the ab-planes is the thin film deposited at a laser pulses repetition rate of 1 Hz. It is evident that the pinning potential along the ab-planes of the Ba2YNbO6 doped YBa2Cu3O7−δ thin film de- posited at 5 and 10 Hz is reduced. The cause of this reduction can not be indicated in the nanorods inclusion because these are the dominant inclusion type in the Ba2YNbO6 doped YBa2Cu3O7−δ deposited at 1 Hz and its ab-direction pinning 87 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.9: Critical current density variation with the direction of the applied magnetic field measured on a pure YBa2Cu3O7−δ thin film deposited at a sub- strate temperature of 780 ◦C and laser pulses repetition rates of 1 Hz and on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate tem- perature of 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz. a) Applied magnetic flux density = 0.5 T, T = 77 K; b) Applied magnetic flux density = 1 T, T = 77 K. 88 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters potential is similar if not better then the one shown by the pure YBa2Cu3O7−δ film. The ab-direction reduction of the Jc could be associated with the lower con- nectivity evidenced in the Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at 5 Hz and 10 Hz. It is certain that the loss of pinning capability when the field is applied parallel to the YBa2Cu3O7−δ ab-planar direction is related to the interruption of the crystalline ab-planes continuity. There are two main sources of planar continuity interruption: the nanorods and the grain boundaries. The nanorods generate a local hole through the planes and a possible loss of conti- nuity in the small region around them that is affected by planar buckling. The grain boundaries are a two dimensional interruption of the planar continuity that is extended over lengths that are order of magnitude larger that the nanorods interruption. This together with the evidence that one of the main effects of higher repetition rates is the reduction of the growth grains dimension (figure 5.4) suggest that the deficient pinning potential measured when the field is ap- plied parallel to the ab-planar direction is caused by the interruption of the ab crystalline planes occurring at the grain boundaries. It is worth remembering that the scenario presented for the Ba2YNbO6 doped YBa2Cu3O7−δ deposited at laser pulses repetition rates of 5 and 10 Hz is com- mon in pinning enhanced YBa2Cu3O7−δ thin films in which the secondary phases forms heteroepitaxial nanorods [57, 65, 74, 118] while the presence of a strong c- axis pinning without a reduction of the ab-planes pinning was achieved only in carefully tuned BaZrO3 doped YBa2Cu3O7−δ [6, 134]. In general while it is com- mon to increase the isotropic pinning without disrupting the ab-planar continuity, it is complex to introduce a strong anisotropic pinning along the c-axis direction without depressing the planar Jc. The Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited at a laser pulses repetition rate of 10 Hz shows a strong c-axis pinning potential revealing the presence of c-axis directional defects that were not clearly shown by the trans- mission electron microscopy. Furthermore the isotropic pinning introduced by the densely packed nanoparticles is not effective at low field (B ≤ 1 T) and at these field values the connectivity issues and low energies flux jumps nullify the pinning effects. Similarly the pinning enhancement that should be provided by the simultaneous presence of nanorods and nanoparticles in the Ba2YNbO6 89 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters doped YBa2Cu3O7−δ deposited at 5 Hz are also probably denied by the reduced connectivity. 5.1.5 Concluding remarks to the analysis of the effects of the deposition rate on the nanostructure and the superconducting properties The analysis performed in this section demonstrated that the laser pulses repeti- tion rate heavily influence the YBa2Cu3O7−δ grains morphology and consequently the overall connectivity and continuity of the ab crystalline planes. Higher repe- tition rates generates smaller growth grains with reduced connectivity and planar continuity. Such reduced connectivity and continuity is found to generate a re- duction of the critical current values Jc when the field is applied parallel to the YBa2Cu3O7−δ ab-planes. The laser pulses repetition rate influences also the nanostructure of the Ba2YNbO6. Long linear nanorods are obtained at low deposition rates (1 Hz), at 5 Hz are obtained shorter nanorods and simultaneously nanoparticles and at 10 Hz the nanoparticles are the dominant Ba2YNbO6 nanostructure. It would be pos- sible to tune the Ba2YNbO6 nanorods-nanoparticles ratio to achieve an optimal pinning landscape by tuning the repetition rate as done with the BaZrO3 [134] but the connectivity issues evidenced have a larger detrimental effect on the pinning properties than the enhancing effect expected from the sinergetic combination of 0D nanoparticles and 1D c-axis oriented nanorods. The best results in mid-low field (0 T to 2 T) values where obtained by Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at 1 Hz. On the other hand the high density pinning landscape generated in the Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited at 10 Hz provides enough pinning force to overcome the related connectivity issue when the field applied is above 2 T; a possible application niche for the 10 Hz thin films could then be found in special- ized equipment for high magnetic field. 90 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters 5.2 Ba2YNbO6 perovskite additions to YBa2Cu3O7−δ: deposition parameter optimization There are several parameters that define the process in a pulsed laser deposition. The substrate - target distance, the laser fluence, the oxygen pressure, the sub- strate temperature and the laser pulse repetition rate all influence the quality of the deposited thin film. The substrate - target distance, the laser fluence and the oxygen pressure are related more to the target ablation and plume diffusion process than to the film growth, furthermore they have been optimized over the years. The last two are the parameters that directly influence the film growth process. A complete study on the effects of the laser puses repetition rate is discussed in the previous section, while in this brief final section of the chapter the study of the properties of the Ba2YNbO6 doped YBa2Cu3O7−δ thin film is concluded with the optimisation of the substrate temperature in the pulsed laser deposition. This final section will not include an in depth study of the nanos- tructure similar to the one reported in the previous sections of the chapter but will focus on the grain connectivity and the superconducting properties of a large number of films. 5.2.1 Ba2YNbO6 doped YBa2Cu3O7−δ pulsed laser depo- sition target preparation and thin films deposition The target used in the pulsed laser deposition is the one that was ablated to deposit the thin films analysed in the previous section. The substrate temperature was modified to study its effects on the growth and the superconducting properties of Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at different laser pulses repetition rate. It is difficult to predict the effects of the substrate temperatures on the film growth as it simultaneously affects multiple mechanism of the process. As an example an higher substrate temperature is certainly related to higher mobility of the ions moving on the growing film surface and at the same time high substrate temperatures in a low oxygen pressure environment may speed up an oxygen content reduction process. Thus in order to identify a processability window, if not the optimum, a set 91 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters of Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at different substrate temperatures and different laser pulse repetition rate were grown. In particular repetition rate of 1, 5 and 10 Hz and substrate temperatures of 740 and 760 ◦C were investigated and compared to the results obtained on thin films deposited at a substrate temperature of 780 ◦C. In the table 5.7 below are summarized the pulsed laser deposition parameters adopted for the deposition of the Ba2YNbO6 doped YBa2Cu3O7−δ thin films analysed in this section. Parameter Value Substrate Temperature 740,760,780 ◦C Chamber Pressure 0.3 mbar flowing O2 Laser Fluence 2 Jcm−2 Repetition Rate 1, 5, 10 Hz Number of Pulses 4500 Annealing Time 1 hr Annealing Temperature 520 ◦C Annealing Pressure 500 mbar O2 Table 5.7: Pulsed laser deposition parameters All the Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited were measured to be of ≈ 0.4 µm thickness. 5.2.2 Surface Topology: Grain Connectivity The analysis discussed in the previous section showed that the grain connectivity has a major influence on the superconducting properties, in particular when the magnetic field is applied parallel to the YBa2Cu3O7−δ ab-planar direction the crit- ical current values appear to be reduced in the Ba2YNbO6 doped YBa2Cu3O7−δ thin films featuring small growth grains. For this reason a study of surface to- pography can not be neglected even in this brief analysis of the superconducting properties. In order to provide clear visual impact the images of the surface topography were separated into three. In figure 5.10 are reported the surface topographies for all the thin films produced with a substrate temperature of 740 ◦C, in figure 5.11 those deposited at 760 ◦C and in figure 5.12 those at 780 ◦C. In all the figures from top to down the laser pulses rate varies from 1 Hz to 10 Hz. 92 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.10: Atomic Force Microscopy surface image (5 µm × 5 µm) of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 740 ◦C; a) repetition rate of 1 Hz; b) repetition rate of 5 Hz; c) repetition rate of 10 Hz; 93 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.11: Atomic Force Microscopy surface image (5 µm × 5 µm) of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 760 ◦C; a) repetition rate of 1 Hz; b) repetition rate of 5 Hz; c) repetition rate of 10 Hz; 94 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.12: Atomic Force Microscopy surface image (5 µm × 5 µm) of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 780 ◦C; a) repetition rate of 1 Hz; b) repetition rate of 5 Hz; c) repetition rate of 10 Hz; 95 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters The first clear feature evidenced by these atomic force microscopy study is the presence of outgrowths on the surfaces of the thin films deposited at a substrate temperature of 740 ◦C (figure 5.10). These outgrowths where recently reported in the case of pure YBa2Cu3O7−δ on RABiTS for substrate temperatures from 740 ◦C and below [135]. Nevertheless several authors have investigated and char- acterized the outgrowths on the surface of pulsed laser ablated YBa2Cu3O7−δ on different substrates [136,137]. The outgrowths are reported to be a-axis YBa2Cu3O7−δ grains, misaligned c- axis grains and non superconducting grains made by Y2BaCuO5, CuYO2, CuO, BaCuO2 and Ba2CuO3. It is possible to distinguish between the outgrowths formed by superconducting YBa2Cu3O7−δ grains and those formed by non su- perconducting phases with the atomic force microscopy. The outgrowths formed by YBa2Cu3O7−δ (a-axis YBa2Cu3O7−δ and misaligned c-axis YBa2Cu3O7−δ) are reported as block-shaped while those formed by non superconducting phases are semi-spherical. From the images reported in figure 5.10 it is evident the out- growths are block-shaped thus are YBa2Cu3O7−δ grains. The superconducting outgrowths are described as predominantly c-axis YBa2Cu3O7−δ grains paral- lel to the substrate surface that nucleate on the substrate at the same time of the desired c-axis YBa2Cu3O7−δ perpendicular to the substrate surface when the substrate temperature is below a certain threshold. In the present case this lower temperature limit is found to be 740 ◦C. Further- more a significant increase in the size of the outgrowths is also reported to occurs during long deposition. This size increment is related to precipitates formation at the edge of the outgrowths as well as to the merging of neighboring c-axis parallel grains. A similar behaviour is found in figure 5.10 where the size of the outgrowths on the surface of the thin film deposited at a laser pulses repetition rate of 1 Hz is much larger than that observed for the thin films deposited at 5 and 10 Hz. Although the quality of these films is reduced by the presence of these outgrowths, it is significant to make a final comment on the dimension and shape of the c-axis perpendicular grains that are visible below the outgrowths. These grains are almost perfectly rectangular, their size is ≈ 300 nm and do not seem to change with the variation of laser pulses repetition rate. From the images in figure 5.11 and 5.12 it is evident that the outgrowths are absent. 96 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.13: Average grain size measured from AFM analysis of 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin film deposited at a substrate temperature of 740, 760 and 780 ◦C as a function of laser pulse rate Figure 5.13 compares the average grain size measured from the AFM in figure 5.10, 5.11 and 5.12, which shows a well defined trend. The size of the grains and thus the connectivity increases when the substrate temperature is increased and when the laser pulses repetition rate is decreased. It is important remembering that a reduction of the pulses laser repetition rate corresponds to an increase in the maximum time allowed to the ions migration on the surface. Referring to a standard model of nucleation and growth, the grain size analysis shows that an increase in ions mobility as well as the time allowed to their migration shifts the balance between nucleation and growth towards the latter. In the previous section it was shown that thin films composed of larger grains have better super- conducting properties than films composed by smaller grains. Therefore, if this trend were to be confirmed, the films with the best superconducting properties should be those deposited with a substrate temperature of 780 ◦C. A last consideration can be made comparing the surface topographies of the thin film deposited at 1 Hz and 760 ◦C (figure 5.11a) with the one deposited at 5 Hz and 780 ◦C (figure 5.12b). It is evident that the two images are similar thus indicating the possibility to achieve similar results with lower mobility (lower substrate temperature) and longer migration times (lower laser pulse repetition 97 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters rate) or higher mobility and smaller migration times. 5.2.3 The superconducting properties: Tc, Jc(B) and Jc(B, θ) To conclude the last section of the chapter are shown the results obtained from the analysis of the superconducting properties. The variation of transition tem- perature Tc and critical current density as a function of the applied magnetic field intensity Jc(B) and as a function of the applied magnetic field direction Jc(B, θ) are reported. 5.2.3.1 Transition temperature, Tc In figure 5.14 is visualized the Tc variation with the substrate temperature. Each curve is referred to a laser pulse repetition rate. The Tc values are also summarized in tabular form in table 5.8. Figure 5.14: Visualisation of the transition temperature values measured for 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate tem- perature of 740, 760 and 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz The transition temperature values measured for the thin films deposited at a 98 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters substrate temperature of 760 ◦C has the same behaviuor discussed in the previous section for the thin films deposited on the substrate kept at 780 ◦C. The Tc reduction is ≈ 3 K as Tc ≈ 88 K are measured for all the samples. A different Tc trend is observed in thin films deposited on substrates kept at 740 ◦C. In these thin films the Tc values appears to be influenced by the laser pulses repetition rate and the Tc measured for the thin films deposited at 10 Hz shows a severe reduction down to ≈ 79 K. The main difference in these films is the presence of the outgrowths discussed earlier in the section. It is possible that the YBa2Cu3O7−δ c-axis grains aligned parallel to the substrate surface introduce high levels of strain and distortion reducing the Tc. The fact that the size of these outgrowths changes with the variation of the laser pulses repetition rate could be the reason why only in when these outgrowth are present the Tc values also changes with the variation of the repetition rate. A possible explanation to the Tc variation with the laser pulses repetition rate can be found considering that the strain and distortion are responsible for the reduced Tc values. These are generated at the interfaces between two phases (or c- axis parallel YBa2Cu3O7−δ grains and c-axis perpendicular YBa2Cu3O7−δ grains) thus larger interfaces induce larger Tc reduction. Taking into account that the an increase of the laser pulses repetition rate decrease the sizes of the outgrowths but at same time increase their density and thus increase the interfacial area (figure 5.10), the relation between Tc values and the laser pulses repetition rates is evident. 1 Hz 5 Hz 10 Hz 740 ◦C 87.8 K 85K 79 K 760 ◦C 87.9 K 89 K 88.2 K 780 ◦C 87.8 K 88.8 K 88.8 K Table 5.8: Transition temperature values for 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 740, 760 and 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz. 99 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters 5.2.3.2 The critical current density The critical current values measured on thin films deposited on substrate kept at temperature values smaller then 780 ◦C are all reduced (figure 5.15). A reduction of the substrate temperatures appears to reduce the critical current values regard- less of the repetition rate. As a matter of the fact that the detrimental effects generated by connectivity issues has been evidenced and discussed in the previous section and that the atomic force microscopy analysis revealed that larger grains (better connectivity) where generated at high substrate temperatures a similar behaviour of the critical current values was expected. Figure 5.15: Critical current density variation with the applied magnetic field value measured on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 760 and 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz. The field is applied parallel to the c-axis of the YBa2Cu3O7−δ, T = 77 K. A self explanatory summary plot is reported in figure 5.16, the critical current density values measured at 0.5 T and 1 T are plotted as a function of the pulse rate and for the samples deposited at a substrate temperature of 760 ◦C and 780 ◦C. In figure 5.17 the critical current density variation with direction of the applied 100 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.16: Critical current density measured at 0.5 T and 1 T on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate tempera- ture of 760 and 780 ◦C as a function of laser pulse rate. The field is applied parallel to the c-axis of the YBa2Cu3O7−δ, T = 77 K. magnetic field is reported. These data do not add new knowledge on the system analysed and the consideration done in the last section could be repeated. Once again the relevant results is that the best performances are obtained by the thin films deposited on substrate kept at 780 ◦C. An interesting observation is obtained by the comparison of the Jc(B) and the Jc(B, θ) of the thin film deposited at 760 ◦C and 1 Hz with those of the thin film deposited at 780 ◦C and 5 Hz. The Jc values obtained in these films are similar and also the topographies were similar. This is an important evidence that, as stated before, it is possible to reduce the deposition time (increase the laser pulses repetition rate) without affecting the superconducting properties in- creasing the ions mobility (increase the substrate temperature). Unfortunately, even if the above is true, the substrate temperature (mobility) has an upper limit. When temperatures above 780 ◦C are used new phenomena are introduced and the superconductive properties are worsened probably by reductions in oxygen content. In particular Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited on substrate kept at temperatures of 800 and 820 ◦C shows Tc values below 80 K. 101 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters Figure 5.17: Critical current density variation with the direction of the applied magnetic field measured on 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at a substrate temperature of 760 and 780 ◦C and laser pulses repetition rates of 1, 5 and 10 Hz. a) Applied magnetic flux density = 0.5 T, T = 77 K; b) Applied magnetic flux density = 1 T, T = 77 K. 102 5. Ba2YNbO6 doped YBa2Cu3O7−δ: the deposition parameters 5.2.4 Concluding remarks to the deposition parameters optimization In this last section the relevance of grain size on the Ba2YNbO6 doped YBa2Cu3O7−δ thin films is once again evidenced. The connection between the temperature at which the substrate is kept during the deposition with the films grains morphol- ogy is analysed. Within the parameter range analysed the optimal superconducting properties for the Ba2YNbO6 doped YBa2Cu3O7−δ system are found in the thin films that are deposited on substrates kept at 780 ◦C with a laser pulse repetition rate of 1 Hz. A temperature of ≈ 780 ◦C seems to be an optimal value since further increments require additional processing to equilibrate the oxygen content. If post annealing processes are not considered a substrate temperature ≈ 780 ◦C is an upper limit on the processability parameter variation range. On the other hand, the laser pulses repetition rate of 1 Hz is the lowest rate analysed and the trend observed indicated that reducing the repetition rate would lead to larger growth grains and more ordered nanorods arrays. However, it is important to notice that even if there is not lower limit of processability for the laser pulse repetition rate, there is a lower efficacy limit. In other words adopting increasingly smaller repetition rate at a certain point will cease to be beneficial and from industry point of view will start to be disadvantageous. In fact, once the process ceases to be kinetically limited and reaches the thermodynamic equilibrium any further reductions in the repetition rate would have no influence. 103 Chapter 6 Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ In the previous chapter an in-depth analysis of the effects of Nb addition to YBa2Cu3O7−δ is presented. It is proved that niobium additions to YBa2Cu3O7−δ produce Ba2YNbO6 nanorods and that in general additions to REBa2Cu3O7−δ produce Ba2RENbO6 nanorods [5,118,138] which are similar to BaZrO3 nanorods, ≈ 10 nm in diameter and ≈ 100 nm in length but with a larger splay than BaZrO3 around the c-axis. Another element, the tantallum, also showed similar properties to the nio- bium. Both niobium and tantallum are highly charged ions and do not substitute in the YBa2Cu3O7−δ lattice but form non superconducting phases. One of the phases that has been reported to form in the YBa2Cu3O7−δ doped with tanta- llum is the Ba2YTaO6 perovskite with a crystalline lattice that is identical to the Ba2YNbO6 [124]. Referring to the Ba2YNbO6 and Ba2YTaO6 perovskites the Nb and Ta are present as Nb+4 and Ta+4 coordinating 6 oxygen in an octahedral lattice site and the ionic radius of both the element is 0.082 nm. As a matter of the fact the Ba2YNbO6 and Ba2YTaO6 have identical crystalline lattices. On the other hand Ba2YTaO6 is not the only phase reported to be gener- ated by the tantallum addition, a defective pyrochlore RE3TaO7 has also been 104 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ reported [67]. Despite the different chemical composition and crystalline struc- ture attributed to the nanorods formed by the addition of tantallum these have been described as very fine (≈ 5 nm in diameter), continuous over the entire film thickness, highly linear and densely distributed thus different from the wider (≈ 10 nm in diameter), shorter and less linear nanorods which are formed by the niobium addition. Since the niobate and tantalate nanorods in YBa2Cu3O7−δ are of rather dif- ferent morphology the YBa2Cu3O7−δ doped with niobium and the YBa2Cu3O7−δ doped with tantallum have different superconducting properties. It is interest- ing to evaluate whether the addition of both niobium and tantallum of the same overall doping level results in an averaging effects or in an increased complexity of the system that could yield to an entirely different and new pinning landscape. In this chapter is reported a study of simultaneous doping of YBa2Cu3O7−δ with both Ba2YNbO6 and Gd3TaO7. The reason tantallum was added in the form of RE3TaO7 is that this was previously studied in our group as a pinning additive to YBa2Cu3O7−δ thin films. 6.1 Ba2YNbO6 and Gd3TaO7 doped YBa2Cu3O7−δ target preparation and thin films deposition The Ba2YNbO6 and Gd3TaO7 doped YBa2Cu3O7−δ target for the pulsed laser deposition was sintered adopting the process described in the previous chapter to produce densified target and minimize the particulate amount deposited on film surfaces (section 5.1.1). As anticipated in chapter 3, unlike the Ba2YNbO6 per- ovskite, the Gd3TaO7 powders were not prepared prior to the target sintering and the desired amount of 99.99% Gd2O3 and 99.99% Ta2O5 powders were directly added to a mixture of YBa2Cu3O7−δ and Ba2YNbO6 powders. The final target stochiometry is 2.5%mol Ba2YNbO6 + 2.5%mol Gd3TaO7 doped YBa2Cu3O7−δ doped YBa2Cu3O7−δ. The pulsed laser deposition parameters chosen for the thin films analysed in this chapter are those that showed optimised growth for the 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ. In particular, the substrate temperature was kept at 780 105 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ ◦C and the laser pulses repetition rate was set at 1 Hz. In table 6.1 is summarized the complete set of deposition parameters studied. Parameter Value Substrate Temperature 780 ◦C Chamber Pressure 0.3 mbar flowing O2 Laser Fluence 2 Jcm−2 Repetition Rate 1 Hz Number of Pulses 4500 Annealing Time 1 hr Annealing Temperature 520 ◦C Annealing Pressure 500 mbar O2 Table 6.1: Pulsed laser deposition parameters The Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ thin films deposited were measured to be of ≈ 0.35 µm thickness. 6.2 Crystalline structure analysis: x-ray diffrac- tion data Similarly to the crystalline structure characterisation performed on the first Ba2YNbO6 doped YBa2Cu3O7−δ thin films produced the first results presented in this chapter are those obtained from the x-ray diffraction analysis. It is funda- mental to investigate the crystalline structure and the orientation of the phases introduced in the thin films. 6.2.0.1 Crystalline phases identification Figure 6.1 shows the x-ray diffraction pattern of a typical thin film. The (00l) peaks of YBa2Cu3O7−δ (or (Y/Gd)Ba2Cu3O7−δ) as well as the (002), (004) and (008) peaks of Ba2(Y/Gd)(Nb/Ta)O6 are labeled. A peak at 2θ = 33.3◦ can be identified as the (004) diffraction peak of (Y/Gd)2O3. Further- more traces of a copper rich cuprate are also present as indicated by the diffraction peaks (008),(0010) and (0012) at 2θ = 25.9◦, 2θ = 33.2◦ and 2θ = 41.4◦ that can be associated to (Y/Gd)Ba2Cu4O8. 106 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ Figure 6.1: Bragg-Brentano scan of a YBa2Cu3O7−δ + 2.5%mol Ba2YNbO6 + 2.5%mol Gd3TaO7 thin film deposited on SrTiO3. It is evident that Gd3TaO7 is absent and that a chemical reaction has gen- erated new phases. It is possible to express a balanced chemical reaction that summarize the phases transformation and that is consistent with the phases ob- served by the x-ray diffraction reported in figure 6.1. 0.95 YBa2Cu3O7−δ + 0.025 Ba2YNbO6 + 0.025 Gd3TaO7 ↓ 0.85 (Y/Gd)Ba2Cu3O7−δ+ 0.05 Ba2(Y/Gd)(Nb/Ta)O6 + 0.0375 (Y/Gd)2O3 + 0.075 (Y/Gd)Ba2Cu4O8 Considering the nature of the ions present in the system is easy to understand the phases evolution observed. Gd can substitute onto the Y site since they have similar ionic radii. For the same reason Nb and Ta can cross-substitute. This allows Ta and Gd introduced as Gd3TaO7 together with the Ba2YNbO6 and Ba ions subtracted to the YBa2Cu3O7−δ to form the complex perovskite Ba2(Y/Gd)(Nb/Ta)O6. Furthermore the Gd3TaO7 → Ba2(Y/Gd)(Nb/Ta)O6 transformation generates a Gd excess and a Ba deficiency that together with the Gd ↔ Y mutual substitution could be at the origin of the (Y/Gd)2O3 and 107 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ (Y/Gd)Ba2Cu4O8 particles formation. The previously reported [67] Gd3TaO7 phase does not form and the Ta partic- ipate to the formation of a perovskite structure similar to the reported Ba2YTaO6 [68, 109]. Ba2YNbO6 and Ba2YTaO6 have identical crystalline structure and the newly formed Ba2(Y/Gd)(Nb/Ta)O6 do not shows evident differences. (Y/Gd)Ba2Cu4O8 formation has been previously observed in coated YBa2Cu3O7−δ superconductors affected by barium deficiency. In particular a bar- ium depletion by reaction with a CeO2 buffer layer has been reported to generate (Y/Gd)Ba2Cu4O8 [139]. In the Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ the (Y/Gd)Ba2Cu4O8 formation can be related to the barium deficiency in the Gd3TaO7 reactant. (Y/Gd)Ba2Cu4O8 has been reported also in films grown by metal-organic deposition, where it is described to be in the form of stacking fault defects [140]. Concluding the phase identification it is clear that the Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ is a Ba2(Y/Gd)(Nb/Ta)O6 + (Y/Gd)2O3 doped (Y/Gd)Ba2Cu3O7−δ thin film with additional traces of (Y/Gd)Ba2Cu4O8. 6.2.1 Crystalline phases orientation The phase orientation analysis is reported in figure 6.2. The SrTiO3 curve is omitted since the (Y/Gd)Ba2Cu3O7−δ can be taken as an orientation reference. The cube on cube growth of the Ba2(Y/Gd)(Nb/Ta)O6 in the (Y/Gd)Ba2Cu3O7−δ is evidenced by the x-ray φ scans: the (202) Ba2(Y/Gd)(Nb/Ta)O6 peak (χ = 45 ◦, 2θ = 30.03◦) matches the (102) (Y/Gd)Ba2Cu3O7−δ (χ = 57.06◦, 2θ = 27.62◦) peaks. Furthermore considering the out-of-plane c-axis homogeneous orientation a full heteroepitaxy between the (Y/Gd)Ba2Cu3O7−δ, the Ba2(Y/Gd)(Nb/Ta)O6 and the SrTiO3 substrate is estabilished. The φ scan performed on the (404) peak (χ = 45◦, 2θ = 48.518◦) is evidence that the (Y/Gd)2O3 is rotated 45 ◦ in-plane, as previously reported [141]. At a more accurate analysis a small fraction of the (Y/Gd)2O3 particles appears to be oriented with a minimum offset from the a and b crystalline directions of the (Y/Gd)Ba2Cu3O7−δ, a few degrees away from a cube on cube orientation. These broad (Y/Gd)2O3 φ peaks have been explained as a near coincidence site 108 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ Figure 6.2: X-ray diffraction data from φ scans of (102) YBa2Cu3O7−δ, (202) Ba2YNbO6 and (404) (Y/Gd)2O3 from a 2.5%mol Ba2YNbO6 + 2.5%mol Gd3TaO7 doped YBa2Cu3O7−δ thin film. lattice matching that can possibly accommodate the large strain and structural differences between (Y/Gd)2O3 and YBa2Cu3O7−δ [142]. However observing the intensity ratio between the peaks related to the (Y/Gd)2O3 rotated 45 ◦ and those related to the (Y/Gd)2O3 in a near coincidence matching it is evident that only traces of the latter can be found and that the (Y/Gd)2O3 can be considered as a rotated 45◦ in plane. 6.3 Cross-section transmission electron microscopy Transmission electron microscopy cross-sectional images are reported in figure 6.3. At least two different nanostructural features can be clearly observed. The first and more obvious feature is a set of fine segmented nanorods of ≈ 7 nm in diameter parallel to the (Y/Gd)Ba2Cu3O7−δ c-axis. These nanorods are smaller than the Ba2YNbO6 nanorods which are ≈ 10 nm in diameter [5, 124], but larger than the nanorods of Gd3TaO7 [67] and Ba2YTaO6 [68] which are ≈ 5 nm in diameter. It is important to notice that while in the case of Gd3TaO7 109 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ Figure 6.3: a) TEM image of a YBa2Cu3O7−δ + 2.5%mol Ba2YNbO6 + 2.5%mol Gd3TaO7 thin film cross section; b) TEM image of plate-like nanoparticleof (Y/Gd)2O3 nucleated between two Ba2(Y/Gd)(Nb/Ta)O6 nanorods segments; c) Selected area electron diffraction pattern of the image in (a); d) TEM imafe of plate-like nanoparticle of (Y/Gd)2O3, inset (d) Fourier transform of the par- ticle image. (TEM images from Prof. H Wang research group at Texas A&M University). 110 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ nanorods different nanostructural features could be related to different strains and lattice mismatch in the case of Ba2YTaO6 this is not possible because Ba2YTaO6 and Ba2YNbO6 have the identical crystalline lattice. However despite having the same crystalline lattice the Ba2YNbO6 forms nanorods larger in diameter than the Ba2YTaO6 hence this has to be related to different kinetics of the niobium and the tantallum. Ba2(Y/Gd)(Nb/Ta)O6 could be described as a mixture of Ba2YNbO6 and Ba2YTaO6 (with an additional Gd ↔ Y substitution), hence an averaged growth kinetics of Ba2YNbO6 and Ba2YTaO6 is obtained , as might be expected. The average nanorod segment length is ≈ 30 nm, this is shorter than both the continuous Ba2YTaO6 that are long the entire film thickness and the short Ba2YNbO6 that are ≈ 80-100 nm. An average nanorods spacing of ≈ 28 nm nm can be calculated from image in figure 6.3a. Calculating the matching field for a triangular and a quadratic rods distribution (see section 4.2.3.2) the applied magnetic field value in which the pinning potential of the analysed thin films should be higher is between 2.64 T and 3 T, these values are larger than those reported for the Ba2YNbO6 inclusion with the same overall doping level. A second nanostructural feature is clearly shown in figure 6.3b: a plate-like nanoparticle of RE2O3 (indicated as (Y/Gd)2O3) parallel to the (Y/Gd)Ba2Cu3O7−δ ab-planes is connecting two adjoining Ba2(Y/Gd)(Nb/Ta)O6 nanorods segments. To confirm the crystalline nature of this plate-like nanopar- ticles as RE2O3 a Fourier transform confirming the structure of a (Y/Gd)2O3 particle is shown in figure 6.3d. These plate-like particles of (Y/Gd)2O3 are in the 25-30 nm width range and their extensions along the (Y/Gd)Ba2Cu3O7−δ c-axis is ≈ 7 nm. In order to confirm the phases identification obtained with the x-ray diffraction analysis in figure 6.3c is reported a selected area electron diffraction pattern of a large cross-sectional region. Diffraction spots that can related to SrTiO3, (Y/Gd)Ba2Cu3O7−δ, Ba2(Y/Gd)(Nb/Ta)O6, (Y/Gd)Ba2Cu4O8 are labeled and are consistent with the x-ray diffraction pattern reported in figure 6.1. To conclude, the nanostructure investigation reported in this section is impor- tant to emphasize that the nanorods segmentation is a new structural feature, and that this is the first time it has been observed. The segmentation of the nanorods yields to an average segment length of ≈ 30 nm. An explanation of the phe- 111 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ nomenon may be in kinetic terms: the tantallum ion is heavier than the niobium thus it has a slower diffusion, this may generate a shortage in the tantallum sup- ply as a reactant for the nanorods’ growth thus preventing these from maintaining continuity as the film growth progresses. A new segment nucleates aligned along the (Y/Gd)Ba2Cu3O7−δ c-axis with a preexisting segment thanks to the presence of a nucleation enhancing strain field generated within the (Y/Gd)Ba2Cu3O7−δ by the nanoinclusion similarly to the BaZrO3 and Ba2YNbO6 [133]. 6.4 The superconducting properties: Tc, Jc(B) and Jc(B, θ) The complex thin film crystalline structure composition made of three different nanoinclusion phases ((Y/Gd)2O3, Ba2(Y/Gd)(Nb/Ta)O6, (Y/Gd)Ba2Cu4O8) and a possible RE (Gd ↔ Y) variation do not influence the transition temperature. No suppression of the superconducting transition temperature was observed, the Tc values measured were ≈ 89 K. Similarly to what observed in Ba2YNbO6 doped YBa2Cu3O7−δ thin films a small reduction of the transition temperature values from the transition of the pure YBa2Cu3O7−δ thin films is expected [5,57,68,118,119], when this is limited to a few Kelvin (≈ 2 K), it is evident that the secondary non superconducting phases are stable and that the only source of Tc reduction are the induced YBa2Cu3O7−δ lattice distortions. 6.4.1 The critical current density In this section are compared the effects on the critical current density of Ba2YNbO6, BaZrO3, and simultaneous Ba2YNbO6 - Gd3TaO7 doping. The critical cur- rent density reported are all measured on thin films of ≈ 300 nm thickness. The BaZrO3 doped YBa2Cu3O7−δ data are taken from one of the best per- forming films in literature deposited adopting a YBa2Cu3O7−δ target covered by yttria-stabilized zirconia on 2% of the surface area [6] while a Ba2YNbO6 doped YBa2Cu3O7−δ thin films of ≈ 350 nm thickness was specifically grown and measured for this comparison. 112 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ 6.4.1.1 The critical current density: Jc(B) Figure 6.4: Critical current density variation with the applied magnetic field value measured on a pure YBa2Cu3O7−δ, a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ, a 2.5%mol Ba2YNbO6 - 2.5%mol Gd3TaO7 doped YBa2Cu3O7−δ and a BaZrO3 doped YBa2Cu3O7−δ thin films [6]. The field is applied parallel to the c-axis of the YBa2Cu3O7−δ, T = 77 K. the data are on a log-linear plot (a) and a log-log plot (b). 113 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ In figure 6.4 is reported Jc(B) measurements up to 4 T with the field ap- plied parallel to the c-axis (B‖c) at 77 K for a pure YBa2Cu3O7−δ, a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ, a 2.5%mol Ba2YNbO6 - 2.5%mol Gd3TaO7 doped YBa2Cu3O7−δ and a BaZrO3 doped YBa2Cu3O7−δ thin films. The thin films with the simultaneous doping of Ba2YNbO6 and Gd3TaO7 show the highest Jc of all the films studied in the entire field range analysed. An outstanding Jc value of 1.8 MAcm−2 at 1 T and values above 1 MAcm−2 up to 2.5 T are the best indications of the excellent potential of this new pinning landscape. Furthermore, from the data shown in figure 6.4b it is clear that the exponential Jc decay range (linear on double logaritmic axis) is extended to fields values above 2 T. Another unusual feature related to this new pinning landscape is the net change in the slope of Jc(B) that occurs as the field intensity approaches the value of matching fields (≈ 2.6 T). This could be an indication of the effective defusing of the low energy depinning mechanisms and of a strong pinning capacity of the single nanorods. The Jc, in fact, start to decrease only when the fluxons are about to saturate the pinning sites. The high pinning capacity of the Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ thin films can also be deduced by the α values reported in table 6.2. Dopant α YBa2Cu3O7−δ 0.47 Ba2YNbO6 doped YBa2Cu3O7−δ 0.35 BaZrO3 doped YBa2Cu3O7−δ 0.25 Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ 0.14 Table 6.2: α values calculated from the Jc(B) curves shown in figure 6.4. The α value ≈ 0.14 calculated for the Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ thin film is lower than the lowest previously reported value, an α ≈ 0.19 reported in a specially grown BaZrO3 doped YBa2Cu3O7−δ where it was shown that an optimisation of the deposition process could produce a mixture of BaZrO3 randomly distributed nanoparticles and splayed columnar defects that could synergetically reduce the depinning mechanism effectiveness [134]. In fact, the new pinning landscape produced by the Ba2YNbO6 + Gd3TaO7 simultaneous doping consists of segmented nanorods and plate-like nanoparticles that form a 114 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ synergetic pinning landscape similar to the one produce in the specially grown BaZrO3 doped YBa2Cu3O7−δ. The two pinning landscapes show two key differ- ences: the Ba2(Y/Gd)(Nb/Ta)O6 nanorods are smaller in diameter (≈ 7 nm) than the BaZrO3 nanorods (≈ 10-15 nm); the Ba2(Y/Gd)(Nb/Ta)O6 nanorods are segmented while the BaZrO3, similarly to the Ba2YNbO6, while not being continuous have an average length of ≈ 100 nm. The segmentation could be the key factor to explain the lower α, segmented rods could in fact gave rise to strongly pinned staircase vortices. 6.4.1.2 The critical current density angular dependence: Jc(B, θ) Investigating the angular dependence of Jc (figure 6.5) it is evident that the Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ thin film shows superior pinning properties over a large angle ranges and not only when the field is applied parallel the YBa2Cu3O7−δ c-axis. The Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ films show at low field val- ues (figure 6.5a) a strong, narrow c-axis pinning peak and no variation of critical current respect the pure YBa2Cu3O7−δ thin film is measured when the field is applied parallel to the YBa2Cu3O7−δ ab-planes. The Jc values measured for this new pinning landscape with the field applied parallel to the YBa2Cu3O7−δ c- axis of 1.8 MAcm−2 is 3 times higher then the one measured for the Ba2YNbO6 doped YBa2Cu3O7−δ and 6 times higher then the one measured for the undoped YBa2Cu3O7−δ. The Ba2YNbO6 doped YBa2Cu3O7−δ thin film shows results that are identical to those reported in the previous chapter: a noticeable Jc increase when the field is applied parallel to the YBa2Cu3O7−δ c-axis together with a small Jc increase with fields applied parallel to the YBa2Cu3O7−δ ab-planes. Increasing the field intensity the strong, narrow c-axis pinning peaks of the Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ films changes to a strong and broad peak. The segmented nanorods provide strong pinning to the vortices over a wide range of angular directions. The Jc values measured on the Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ are higher then those measured on Ba2YNbO6 doped YBa2Cu3O7−δ indicating that the pinning of the finer, segmented 115 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ Figure 6.5: Critical current density variation with the direction of the applied magnetic field measured on a pure YBa2Cu3O7−δ, a 5%mol Ba2YNbO6 doped YBa2Cu3O7−δ, a 2.5%mol Ba2YNbO6 - 2.5%mol Gd3TaO7 doped YBa2Cu3O7−δ and a BaZrO3 doped YBa2Cu3O7−δ thin films [6]. The field is applied parallel to the c-axis of the YBa2Cu3O7−δ. a) Applied magnetic flux density = 1 T, T = 77 K; b) Applied magnetic flux density = 3 T, T = 77 K. 116 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ Ba2(Y/Gd)(Nb/Ta)O6 nanorods together with (Y/Gd)2O3 plate-like nanopar- ticles is more effective compared to the coarser, non segmented Ba2YNbO6 rods. 6.4.1.3 A new pinning feature A last interesting piece of information is the evolution of the angular Jc increasing the field (figure 6.6a). If a low field the most prominent feature is the strong c-axis pinning peak and substantially unmodified ab-planes peaks (figure 6.5a) increasing the field values a new feature start to become visible. This new pinning feature can be initially observed at 3 T and it appears as shoulders on the broad c-axis peak and increasing the field value these shoulders resolves into distinct peaks around ± 60◦. The magnitude of this two peaks increases with increasing the field values and for field values of 5 T and above they become the dominant pinning peaks. This additional pinning preferential direction is a new feature that is directly related to the new pinning landscape generated in the Ba2YNbO6 + Gd3TaO7 doped YBa2Cu3O7−δ thin films. The existence of this additional preferential pin- ning direction can be explained using the vortex path model [143]. In this model the combination of c-axis pinning structures (segmented Ba2(Y/Gd)(Nb/Ta)O6 nanorods) and ab-planes structures (plate-like (Y/Gd)2O3 nanoparticles and YBa2Cu3O7−δ ab-planes) results in staircase vortices that are pinned simulta- neously by the different structures. These vortices are subjected to the strongest pinning at a characteristic angle determined by the distribution of defects and defects length (figure 6.6b). Since in this type of pinning are involved both the available species of additional pinning structure (Ba2(Y/Gd)(Nb/Ta)O6 nanorods and (Y/Gd)2O3 nanoparticles) the hypothetical matching field would have an higher value thus this pinning is more effective at high field values. 6.4.2 Concluding remarks to Ba2YNbO6 and Gd3TaO7 si- multaneous doping of YBa2Cu3O7−δ The study reported in this chapter is on the Ba2YNbO6 and Gd3TaO7 simul- taneous doping of YBa2Cu3O7−δ. The expected mixture of Ba2YNbO6 coarse splayed nanorods and Gd3TaO7 fine dense linear nanorods was not generated. 117 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ Figure 6.6: a) High field critical current density variation with the direction of the applied magnetic field measured on a 2.5%mol Ba2YNbO6 - 2.5%mol Gd3TaO7 doped YBa2Cu3O7−δ. Applied magnetic flux density = 3 T to 6 T, T = 77 K; b) sketch of vortices interacting simultaneously with nanorod segments along the c-axis and intrinsic and extrinsic defects along ab-planes. 118 6. Ba2YNbO6 and Gd3TaO7 simultaneous doping of YBa2Cu3O7−δ Combining Ba2YNbO6 and Gd3TaO7, two well studied phases, a completely new self-assembled pinning landscape was generated instead. This pinning landscape is not a mere mixture of the landscape generated by Ba2YNbO6 and Gd3TaO7 but presents an optimal nanorods’ architecture of fine, straight, segmented rods. Furthermore (Y/Gd)2O3 nanoparticles were fortuitously generated because the nanorods of Ba2(Y/Gd)(Nb/Ta)O6 formed were of different composition (poorer in rare earth, richer in barium) than the reactants Ba2YNbO6 and Gd3TaO7 that were added to the YBa2Cu3O7−δ pulsed laser deposition target. The critical current densities measured at 77 K are greatly enhanced com- pared to previously studied pinning additives. In particular Jc values above 1 MAcm−2 for fields up to 2.5 T were achieved and this is a new benchmark in the YBa2Cu3O7−δ thin films properties. Furthermore new pinning features around 60◦ were observed for the first time proving the possibility of tuning the pinning landscape in order to generate thin films with an arbitrary preferential pinning direction. Applications where the magnetic field is applied obliquely to the su- perconductor could greatly benefit from the use of this new pinning landascape. 119 Chapter 7 Conclusions and further work The study presented in this thesis investigated the use of novel secondary non superconducting phases to generate nanostructured YBa2Cu3O7−δ thin films pro- duced by pulsed laser deposition from a single composite target with enhanced flux pinning and improved performance in applied magnetic fields. First an intensive study on the addition of a niobium based non superconduct- ing phase was completed. At the time the pinning additive known to be effective for the production of nanorods and the creation of a c-axis pinning peak were BaZrO3 (a zirconium based perovskite) [57], RE3TaO7 (tantallum based pyro- chore) [67] and BaSnO3 (a tin based perovskite) [64]. With the exception of the RE3TaO7 all the other phases are barium based perovskites. These perovskites form a different pinning landscape thanks to different ions kinetics and differ- ent lattice mismatches. Thus the YBa2Cu3O7−δ thin films doped with different perovskites have different properties. It was known that in bulk samples adding niobium or tantallum to YBa2Cu3O7−δ forms Ba2YNbO6 or Ba2YTaO6 [109] but at the time of that study RE3TaO7 was the only phases reported in pulsed laser deposited YBa2Cu3O7−δ thin films doped with tantallum while an early study reported the successfull production of Ba2NbErO6 perovskite in ErBa2Cu3O(7−δ) [118,119]. The first step of the research work reported in this thesis is the demonstration of the possibility to introduce Ba2YNbO6 nanorods in YBa2Cu3O7−δ together with providing the experimental evidence of the potential of broad c-axis pinning peaks. The broad c-axis pinning peak was related to the morphology of the novel 120 7. Conclusions and further work Ba2YNbO6 nanorods, that was found to be different from those reported before. The nanorods resulted to be shorter, wider and with a larger c-axis splay than the those generated by the other known pinning additives [5]. Once discovered, the potential of Ba2YNbO6 as pinning additive, a study of the effects of the deposition rate and substrate temperatures on Ba2YNbO6 doped YBa2Cu3O7−δ thin films was performed. The study demonstrated that fast deposition rates induce smaller growth grains and the formation of Ba2YNbO6 nanoparticles togheter with the ordered Ba2YNbO6 nanorods. The thin films produced adopting high deposition rates while proved to be highly effective at high field values show a severe reduction of the intrinsic YBa2Cu3O7−δ ab-planes pinning. This reduction is related to the large amount of discontinuity intro- duced to the YBa2Cu3O7−δ crystall ab-planes. On the other hand Ba2YNbO6 doped YBa2Cu3O7−δ thin films deposited at low deposition rate showed that it is possible to introduce Ba2YNbO6 and increase the pinning potential along the c- axis without reducing the connectivity or the YBa2Cu3O7−δ crystall quality thus without reducing the ab-plane pinning potential. In conclusion Ba2YNbO6 doped YBa2Cu3O7−δ thin films with enanched c-axis togheter with good ab-planes pin- ning potential were succesfully deposited. These Ba2YNbO6 doped YBa2Cu3O7−δ thin films stand out because they show a low anisotropy of the angular depen- dence of the critical current greatly reducing the design difficulties encountered in many coated conductors applications. The last part of the thesis is devoted at a new research that arises from the desire to investigate the effects on the pinning of a synergistic combination of morphologically different nanoinclusions. In particular the combination of the wide, short splayed Ba2YNbO6 nanorods and the fine, long, higly linear Gd3TaO7 nanorods was investigated. This work demonstrated that in a niobium and tan- tallum combined scenario the tantallum partecipate to the formation of a per- ovskyte and do not form Gd3TaO7. A tantallum based perovskite Ba2YTaO6 was also recently reported in tantallum doped YBa2Cu3O7−δ [68], however also the Ba2YTaO6 perovskite is morphologically different from the Ba2YNbO6 being fine, long and highly linear like the previously reported Gd3TaO7 nanorods. A completely new pinning landscape was produced by depositing thin films from a Ba2YNbO6 + Gd3TaO7 + YBa2Cu3O7−δ PLD target. The structural 121 7. Conclusions and further work and morphological characterization of these films showed that the these films are mainly made of (Y/Gd)Ba2Cu3O7−δ with fine, linear, segmented nanorods of Ba2(Y/Gd)(Nb/Ta)O6 and (Y/Gd)2O3 plate-like nanoparticles inclusions. This new defect landscape produced in the thin films is capable of generating state-of- the-art pinning properties. Astonishingly high critical currents are not the only effect: for the first time possibility to achieve self segmentation of nanorods was demostrated, and for the first time additional new pinning features around 60◦ were observed. Owing to the very high critical currents obtained, exceeding the industry standard BaZrO3, it is likely that the results obtained in this work will at- tract the interest of many research groups from different area of the field. Fur- thermore, modifying the ratio of niobium and tantallum and the ratio between Ba2(Y/Gd)(Nb/Ta)O6 and (Y/Gd)2O3 will almost certanly demonstrate the tun- ability of the pinning landscape, and a new era will begin in which it will be possible to produce superconductors specifically designed for pinning in a prede- termined direction. A direct continuation of this work should investigate the effects of the simul- taneous niobium tantallum doping in a simplified system. At first YBa2Cu3O7−δ films doped only with Ba2YNbO6 and Ba2YTaO6 produced with different of Nb:Ta ratios could be grown in order to investigate the tunability of the seg- mentation. In a second stage (Y/Gd)2O3 could be added to the Ba2YNbO6 and Ba2YTaO6 doping to evaluate the importance of the synergistic combination, and find the optimal balance between plate like nanoparticles and segmented nanorods. In a last stage Gd, as well as other rare earths, could be used to partially substitue the Y riproducing the otpimal pinning landscape obtained in this work but with an increased knoledge of the single constituent effects and increased freedom on the composition of the films. In addition a study of the pinning properties at temperatures below 77 K of the system demonstrated in this dissertation could provide additional evidence of the effective possibility of tuning nanoengineered YBa2Cu3O7−δ in order to achieve optimal pinning along arbitrary directions. 122 Appendix Publications related to the research described in this dissertation are presented in this section. The publications order is the following: • Ercolano G et al. “Enhanced flux pinning in YBa2Cu3O7−δ thin films using Nb-based double perovskite additions”. Supercond. Sci. Technol., 23:022003, 2010. • MacManus-Driscoll J L et al. “High current, low cost YBCO conductors- what’s next?”. Supercond. Sci. Technol., 23:034009, 2010. • Ercolano G et al. “State-of-the-art flux pinning in YBa2Cu3O7−delta by the creation of highly linear, segmented nanorods of Ba2(Y/Gd)(Nb/Ta)O6 together with nanoparticles of (Y/Gd)2O3 and (Y/Gd)Ba2Cu4O8”. Sub- mitted to Supercond. Sci. Technol. A general table of the crystallographic parameters of non-superconductive second phase materials used as pinning phases in YBa2Cu3O7−δ films is also reported. 123 IOP PUBLISHING SUPERCONDUCTOR SCIENCE AND TECHNOLOGY Supercond. Sci. Technol. 23 (2010) 022003 (6pp) doi:10.1088/0953-2048/23/2/022003 RAPID COMMUNICATION Enhanced flux pinning in YBa2Cu3O7−δ thin films using Nb-based double perovskite additions G Ercolano1, S A Harrington1, H Wang2, C F Tsai2 and J L MacManus-Driscoll1 1 Department of Materials Science and Metallurgy, University of Cambridge, Pembroke Street, Cambridge CB2 3QZ, UK 2 Department of Electrical and Computer Engineering, Texas A&M University, College Station, TX 77843, USA E-mail: ge228@cam.ac.uk Received 31 August 2009, in final form 11 December 2009 Published 13 January 2010 Online at stacks.iop.org/SUST/23/022003 Abstract The addition of a new niobate double perovskite pinning phase to YBa2Cu3O7−δ thin films grown by pulsed laser deposition is reported. The YBa2NbO6 phase self-assembles into stacks of ∼10 nm second phase particles, aligned with the c-axis of the YBa2Cu3O7−δ. The YBa2Cu3O7−δ/YBa2NbO6 composite thin films have enhanced critical current, by a factor of 2, at 1 T (H ‖ c) over the pure YBa2Cu3O7−δ, whilst maintaining a high transition temperature. Niobium does not substitute in the YBa2Cu3O7−δ matrix. This, together with the high stability of the second phase formed, makes it an ideal pinning additive. 1. Introduction Improved performance of high temperature superconductors (HTS) in magnetic fields is a highly sought after goal to make HTS conductors commercially viable. There are two routes to achieve higher current carrying performance, (a) controlled nano-engineering of superconducting materials to enhance flux pinning [1–8] and (b) growth of thicker highly epitaxial material in a simple and cost-effective manner through applying novel processing routes. With industrial applications in mind, the current focus should now be to produce efficient, reproducible and cost-effective nanopinning defect arrays. To achieve these end goals it is necessary to: (1) select an ion or phase addition which produces a secondary pinning phase which forms in a wide processing window, and (2) to understand and hence control the second phase assemblage. The group IV and group V ion additions, more specifically Zr4+, Hf4, Ta5+ and, as shown in this work, Nb5+, appear to be the best to meet the first goal [9] and we are now at the stage of beginning to understanding the second one. YBa2Cu3O7−δ (YBCO) thin films containing such additions with improved Jc (up to a factor of 5) in magnetic field was first demonstrated by the introduction of BaZrO3 (BZO) [10]. Through careful process optimization, the results with BZO have been further substantially improved over the last 5 years [11, 12]. More recently, with the aim of producing minimally chemically and structural perturbative effects, rare earth tantalates (RE3TaO7, RTO) have been studied. Compared to BZO, these have been found to yield greater tunability of particle assemblage, and no Tc reduction. In a short space of time, this has led to a Jc increase by up to a factor of 10 [13]. Further studies on the RTO addition effects and on the process optimization have been performed and will be published in the future. It is important to note that for the PLD technique, when deciding on new pinning additions to YBCO the phase which forms is not necessarily the one which is added to the YBCO, but is the one which is the most thermodynamically and epitaxially stable. Hence, if Zr4+ is added to YBCO, Ba(Zr, Y)O3 forms [10]. In this paper we present the results of addition of another large, highly charged ion, Nb5+, (see table 1) which does not substitute for the Cu site in YBCO and hence which has great potential to form a benign second 0953-2048/10/022003+06$30.00 © 2010 IOP Publishing Ltd Printed in the UK1 Supercond. Sci. Technol. 23 (2010) 022003 Rapid Communication Table 1. Ionic charge and radius of critical ion additions for second phase formation, and percentage lattice mismatch with respect to YBCO (along ab and c). Critical pinning ion addition indicates that the ion is not a component of the YBCO phase and that addition of the pure element to YBCO will, for reasons of thermodynamic stability and epitaxial stabilisation, lead to the pinning addition phase in column 1. Critical ion pinning addition/ionic radius (A˚) for 6 fold co-ordination % Misfit to YBCO ab % Misfit to YBCO c YBa2NbO6 Nb5+/0.78 9.42 8.34 BaZrO3 Zr4+/0.86 8.11 7.03 BaSnO3 Sn4+/0.83 6.77 5.72 Sm3TaO7 Ta5+/0.78 −2.33 −7.99 Yb3TaO7 Ta5+/0.78 −4.74 −11.04 phase addition. The charge and size of the ion are important for minimizing the level of substitution in the YBCO lattice. Zr4+ and Sn4+ (table 1) can substitute for Y3+ (whose ion radius is 1.159 A˚ for 8 fold co-ordination), whereas Ta5+ and Nb5+ are less likely to substitute for Y3+ because of the more dissimilar ion sizes and ion valences. We study the addition of a double perovskite Nb-based phase (YBa2NbO6) to YBCO. This phase is the most stable Nb-related compound when in equilibrium with YBCO, explaining why when we add Nb2O5 or BaNbO3 to YBCO, the double perovskite YBa2NbO6 still forms [14–16]. We find that nanoparticles of the phase self-assemble in a manner similar to the tantalates [13] and BZO [17]. Addition of Nb2O5 to bulk YBCO has been shown to enhance Jc [18]. However this enhanced material was not well understood until later [14–16] when segregation into YBCO + YBa2NbO6 was observed. The stability and compatibility of the YBa2NbO6 compound with YBCO made it a candidate for novel substrates [19], as well as for new buffer layers for high quality HTS thin films and electronic devices [20, 21]. Double perovskites, of general formula A2BB′O6, have attracted much interest in other research fields because their high flexibility in lattice parameter and magnetic behaviour [22–25]. YBa2NbO6 has the advantage over many of the double perovskite phases, e.g. the well studied FeSr2MoO6 phase, in that there is less propensity for the cations to have a mixed valence [26], which would then cause unwanted local oxygen stoichiometry changes in the YBCO lattice. The use of YBa2NbO6 as a secondary phase to improve pinning in YBCO epitaxial films on ZrO2 single crystalline substrates has been reported with negative results [27]. The physical properties and chemical compatibility make YBa2NbO6 an ideal candidate as a pinning addition to YBCO. However, difficulty in obtaining adequate oxygenation of composite bulk material [28] resulted in depressed transport properties. This is not an obstacle to YBa2NbO6 doped thin film growth as high surface to volume ratio and small grain size allows full oxygenation in a rapid manner. Here we present success in producing pinning engineered YBa2Cu3O7−δ/YBa2NbO6 thin films with high critical currents grown by pulsed laser deposition. 2. Experimental methods PLD targets were made by mixing and grinding pure YBCO powder (SCI Engineered Materials) (99.999%) with the desired amount of YBa2NbO6 powder in an agate mortar, the latter were produced as single phase by the solid state reaction method, namely mixing, grinding of 99.99% Y2O3, Ba(NO3)2, and Nb2O5 powders followed by reaction at 1450 ◦C for 24 h. The mixed powders were pressed in the form of a cylindrical target and sintered at 950 ◦C for 12 h in flowing O2 in a dedicated tubular furnace. Films were grown by PLD on (001) SrTiO3 (5 mm × 10 mm) single crystal substrate (Pi-Kem Ltd). Deposition were performed with a Lambda Physik KrF excimer laser (λ = 248 nm) in 30 Pa flowing O2, and the substrate temperature was kept at 770 ◦C. A repetition rate of 5 Hz and 4500 pulses were used for all films, which were measured to be of 0.5 μm thickness. The transport critical current density was measured using a conventional four-point probe method and a ∼1 μV cm−1 criterion on photolithographically patterned bridges of 250 μm width. The angular dependence of Jc at 77 K and 0.5 T was measured with the applied magnetic field rotated in a plane perpendicular to the current flow direction to an angle θ with the c-axis of the film. The crystallographic structure and orientation of the films were studied using x-ray diffraction in Bragg–Brentano geometry. Lattice parameters were refined using eva software, using the positions of high angle peaks to minimize errors. Cross-sectional transmission electron microscopy (TEM) and selected area electron diffraction patterns (SADP) were also undertaken to determine the crystal structure and sizes of the nanoparticles formed within the YBCO matrix. 3. Results and discussion Figure 1(a) displays the results of x-ray analysis of the YBa2NbO6 powder produced by solid state reaction. These results agree with the data recorded in the JCPDS or in more recent studies [29]. No impurities or secondary phases peaks are found. The lattice parameter calculated from the x-ray data is 0.844 nm ± 0.001 nm. The x-ray diffraction data from the deposited films (figure 1(b)) shows peaks related to (00l) planes from the YBCO and the SrTiO3 for both undoped and doped samples. This means that the films are c-axis oriented, which is the standard orientation for YBCO thin films on (001) SrTiO3. Additionally, the data from the doped sample only shows the peaks related to the (00l) planes of YBa2NbO6, indicating that the additive phase is aligned out-of-plane with the YBCO. A small peak at 34◦, possibly arising from the (111) planes of YBCO is present in both the undoped and doped sample data. No other secondary phases were present. The (00l) orientations of both phases is expected from the crystallographic matching of these double and triple perovskite cells (shown schematically in figure 2(a)). Looking along the [010]YBCO direction, 3 unit cells of the cubic YBa2NbO6 match 2 unit cells of YBCO giving a lattice mismatch strain along the c-axis of YBCO, ((3aYBa2NbO6 −2cYBCO)/2cYBCO), of +8.34%. 2 Supercond. Sci. Technol. 23 (2010) 022003 Rapid Communication Figure 1. X-ray diffraction data for YBa2NbO6 in the form of pure precursor powder and when embedded within YBCO films. (a) X-ray diffraction spectrum of YBa2NbO6 powder reacted at 1450 ◦C. (b) X-ray diffraction data for undoped YBa2Cu3O7−δ and a 5 mol% YBa2NbO6 doped film grown on SrTiO3; (c) φ scan of (101) SrTiO3, (102) YBCO, (202) YBa2NbO6 from a 5 mol% YBa2NbO6 doped film. In the doped film all the (00l) peaks from the YBa2NbO6 are shifted to higher angles compared to the bulk values giving a lattice parameter of 0.830 nm ± 0.001 nm compared to the bulk value of 0.844 nm ± 0.001 nm. Hence, there is a compressive strain of 1.7% along the c-axis, much lower than the theoretical value of 8.34% assuming the YBCO lattice is not distorted (table 1). The YBCO (00l) peaks are shifted to in the opposite direction to the YBa2NbO6 peaks, i.e. to lower angles, and the ‘c’-axis is extended to a value of 11.73 A˚ giving Figure 2. Crystallographic matching of YBCO with YBa2NbO6. (a) b-axis view and (b) c-axis view. a tensile strain of 0.4%. The extended c-axis is not likely to arise from reduced oxygen content since the sample Tc is not reduced, as shown later. The overall strain is lower than the theoretical value and indicates partial strain relief by formation of misfit dislocations. Dislocation formation has previously been reported in BZO doped YBCO [17]. Nevertheless, the fact that the YBCO is locally partially strained in tension gives the possibility of additional pinning from strain fields in the vicinity of the YBCO/YBa2NbO6 interfacial regions. To assess the in-plane texture, phi scans were recorded as shown in figure 1(c). The (202) YBa2NbO6 peaks reveal in-plane alignment and match both the (101) STO and (102) YBCO peaks, confirming that the perovskite particles have grown aligned ‘cube on cube’ with the YBCO. These results combined with the out-of-plane x-ray data suggest full heteroepitaxy between the YBCO, YBa2NbO6 and the substrate. The cube-on-cube in-plane orientation determined from the phi scan is represented in the crystallographic model of figure 2(b). Looking down the [001]YBCO direction, 1 unit cell of YBa2NbO6 matches 2 unit cells of YBCO giving an in-plane lattice mismatch (a and b average) strain of +9.42%. The lattice mismatch of YBa2NbO6 with YBCO is larger than other widely studied pinning phases additions (table 1). This large mismatch should be beneficial for low field (<1 T, 77 K) pinning due the localized strain fields that will 3 Supercond. Sci. Technol. 23 (2010) 022003 Rapid Communication arise [30] but likely produces shorter and wider self-assembled columnar array, contrasting with tantalate columns where the lattice mismatch is very small and long, narrow columns are produced [13]. They are more similar to BaZrO3 nanorods which have similar lattice mismatch with YBCO [17, 31]. The next basic research challenges to be addressed for niobate pinning additions to YBCO are (a) determining the optimum level of niobate addition, (b) tuning the column sizes and distributions. Cross-section transmission electron micrographs of a 5 mol% YBa2NbO6 doped sample (figures 3(a) and (b)) show nanorods aligned along the c-axis of the YBCO. The rods are 10–15 nm in diameter and their spacing is roughly 40 nm, giving a matching field of 1.3 T. Some random particles were also observed as indicated by the white arrows drawn along the ab planes. The inset to figure 3(a) shows a selected area diffraction pattern of a region around the nanoparticles. A set of streaky diffraction dots for (004) YBa2NbO6 is observed. Streaky dots rather than clearly distinguished spots has been reported in similar systems previously [33] and is an indication of a large YBCO c-plane distortion created by the YBa2NbO6 nanorod additions. d(004) estimated from the diffraction pattern is ∼0.214 nm, indicating a ∼ 0.856 nm, broadly consistent with the x-ray measurements. Figure 3(b) shows a higher magnification image revealing fringes arising from the lattice mismatch between the YBa2NbO6 and YBCO lattices. Since a relatively slow growth rate, 5 Hz, was used, the appearance of both self-assembled columns and random particles are anticipated, similar to the case of RETa3O7 additions [32]. The rods are wider (up 15 nm as opposed to 5 nm) and shorter (∼100 nm as opposed to the whole film thickness) than tantalate rods, most likely because of the large mismatch strain which enhances the self-assembly kinetics [13]. At the interface between the film and substrate, a dark layer (about 10–20 nm thick) is observed which clearly shows different contrast than the rest of the film. This is in agreement with previous observations of asymmetric in-plane Jc seen in YBCO + RETa3O7 films arising from interface disorder [34]. It is not clear if the nanorods nucleated within this darker layer or on the substrate surface. An atomic force micrograph of the film surface (figure 3(c)) shows superficial particles ∼15–20 nm in diameter uniformly distributed over the surface. These particles are consistent with the nanorods terminating at the sample surface. The size of the particles observed with the AFM is larger than the rods observed with the TEM. This could be due to the fact that we are observing the surface where there is higher mobility and possible agglomeration of surface particles immediately after the growth is terminated. Despite the 5 mol% doping level and the large lattice mismatch between the phases, only a minimal reduction of the transition temperature is observed (figure 4(a)), with a transition temperature of 89 K. This value indicates that the Nb does not poison the YBCO, but remains ‘locked’ in the second phase. In addition, the 0.4% strain level measured in the YBCO does not lead to any significant Tc reduction. Similar Tc reductions were also observed in ∼5 mol% BZO doped Figure 3. Microstructures of YBCO + 5 mol%YBa2NbO6 films. (a) and (b) TEM cross-sectional images (lower and higher magnifications, respectively). White arrows in the direction of the c-axis mark the positions of self-assembled nanorods along c. White arrows perpendicular to the c-axis indicate the positions of random nanoparticles as well as the interface between the substrate and the film. Inset to (a) shows selected area diffraction pattern of region in vicinity of nanoparticles. Moire fringes arising from lattice mismatch between YBa2NbO6 and YBCO are observed in (b); (c) AFM image showing nanoparticles at the surface. systems [13]. However, in this case, since we are forming a double perovskite instead of a single perovskite as for BZO, the same mol% addition of double perovskite yields twice the volume of inclusions as the single perovskite. Hence, the small Tc reduction is, in fact, less than might be expected. Jc versus magnetic field measurements up to 1 T (H ‖ c) at 4 Supercond. Sci. Technol. 23 (2010) 022003 Rapid Communication Figure 4. Transport properties of 0.5 μm micron thick YBCO + 5 mol%YBa2NbO6 films on STO. (a) Resistance versus temperature for a thin film doped with 5 mol% YBa2NbO6. Inset shows magnified region of the superconducting transition; (b) critical current density versus applied magnetic field at 77 K, H ‖ c. Inset shows data on log–log axis; (c) angular dependence of the critical current density at 77 K, 0.5 T. 77 K (figure 4(b), with log–log plot shown inset) indicates that the doped sample has better performance over the whole field regime analysed, with Jc of the doped sample being by a factor of ∼2 higher at 1 T. The improved angular dependence of Jc in the YBa2NbO6 doped sample compared with pure YBCO, measured at 0.5 T and 77 K, is shown in figure 4(c). The doped sample shows a reduced anisotropy compared with the pure YBCO. This lower anisotropy results from the increased Jc when the field is aligned with the YBa2NbO6 nanorods. The ab peak is depressed somewhat presumably by the interruption of the intrinsic layer pinning by the highly continuous niobate nanorods. Compared with the typical shape of the angular dependence of pinning enhanced YBCO, for example in YBCO + BZO and YBCO + RTO, the YBa2NbO6 nanorods produce a very broad ‘c’-axis peak which make the particles effective over a relatively large angular range. The large breadth of the peak is possibly related to the large mismatch strain which produces relatively short and wide rods (figure 3). A significant fraction of random isotropic pinning which would improve Jc for all angles but have little effect on the shape of the angular dependence of Jc with magnetic field does not seem to be present as the ab peak is lower in the doped sample than in the pure YBCO. In conclusion a new pinning additive, the double perovskite YBa2NbO6, was used to produce self-assembled non-superconducting nanorods within YBCO thin films. Additional flux pinning was observed not only when the magnetic field was aligned along the c-axis but also over a relatively large angular range. Improvements in Jc of a factor ∼2 at 1 T at 77 K were measured. The minimal poisoning and potential for high tunability of YBa2NbO6 puts it in a similar category to RE3TaO7 pinning additions (which has distinct advantages to BZO). The higher lattice mismatch of YBa2NbO6 with YBCO also gives potential to induce extra pinning from strain effects. Further optimization of growth conditions starting from our preliminary results will almost certainly lead to further enhancements of critical current density. Acknowledgments The work was supported by the European Commission as part of NESPA, Nano-Engineered Superconductors for Power Application, a framework of the Marie Curie Research Training Network, funded within the EU’s 6th framework programme. JLM-D acknowledges support from the Marie Curie Excellence Grant ‘NanoFen’, (MEXT-CT-2004- 014156), and HW acknowledge support from the US NSF (DMR-0709831). References [1] Civale L et al 1990 Phys. Rev. Lett. 65 1164 [2] Civale L et al 1991 Phys. Rev. Lett. 67 648 [3] Konczykowski M et al 1991 Phys. Rev. 44 7167 [4] Kim D H et al 1997 IEEE Trans. Appl. Supercond. 7 1997 [5] Puig T et al 2008 Supercond. Sci. Technol. 21 034008 [6] Sparing M et al 2007 Supercond. Sci. Technol. 20 S239–46 [7] Maiorov B et al 2007 Supercond. Sci. Technol. 20 S223–9 [8] Kursumovic A et al 2009 Supercond. Sci. Technol. 22 015009 [9] Driscoll J et al 2006 US Patent Specification 0025310 [10] MacManus-Driscoll J L et al 2004 Nat. Mater. 3 439 [11] Mele P et al 2008 Supercond. Sci. Technol. 21 015019 5 Supercond. Sci. Technol. 23 (2010) 022003 Rapid Communication [12] Mairov B et al 2009 Nat. Mater. 8 398 [13] Harrington S A et al 2009 Supercond. Sci. Technol. 22 022001 [14] Strukova G K et al 1993 Supercond. Sci. Technol. 6 589 [15] Strukova G K et al 1996 Physica C 267 67 [16] Pillai C G S and George A M 1992 J. Mater. Sci. Lett. 11 1639 [17] Goyal A et al 2005 Supercond. Sci. Technol. 18 1533 [18] Kuwabara M and Kusaka N 1988 Japan. J. Appl. Phys. 27 L1504 [19] Paulose K V et al 1992 Physica C 193 273 [20] Sathiraju S et al 2005 IEEE Trans. Appl. Supercond. 15 3009 [21] Grekhov I et al 1997 Physica C 276 18 [22] Brandle C D and Fratello V J 1990 J. Mater. Res. 5 2160 [23] Koshy J et al 1999 Bull. Mater. Sci. 22 243 [24] Aguiar J A et al 1998 Physica C 307 189 [25] Aguiar J A et al 1997 Phys. Rev. 58 2454 [26] Rager J et al 2004 J. Am. Ceram. Soc. 87 1330 [27] Jia J-H et al 1995 Mod. Phys. Lett. 9 439 [28] Vlakhov E et al 1997 J. Mater. Sci. Lett. 16 763 [29] Barnes P W et al 2006 Acta Crystallogr. B 62 384 [30] MacManus-Driscoll J L et al 2004 Appl. Phys. Lett. 84 5329 [31] Li J et al 2006 Microsc. Microanal. 12 566 [32] Harrington S et al 2010 Nanotechnology accepted [33] Yamada K et al 2008 Appl. Phys. Lett. 92 112503 [34] Harrington S et al 2009 Appl. Phys. Lett. 95 022518 6 IOP PUBLISHING SUPERCONDUCTOR SCIENCE AND TECHNOLOGY Supercond. Sci. Technol. 23 (2010) 034009 (5pp) doi:10.1088/0953-2048/23/3/034009 High current, low cost YBCO conductors—what’s next? J L MacManus-Driscoll1, S A Harrington1, J H Durrell1, G Ercolano1, H Wang2, J H Lee2, C F Tsai2, B Maiorov3, A Kursumovic1 and S C Wimbush1 1 Department of Materials Science and Metallurgy, University of Cambridge, Pembroke Street, Cambridge CB2 3QZ, UK 2 Department of Electrical and Computer Engineering, Texas A&M University, College Station, TX 77843, USA 3 MPA-STC, Los Alamos National Laboratory, Los Alamos, NM 87545, USA E-mail: jld35@cam.ac.uk Received 12 November 2009, in final form 1 December 2009 Published 22 February 2010 Online at stacks.iop.org/SUST/23/034009 Abstract The Holy Grail for high temperature superconducting conductors is achieving high current material in a simple and cost-effective way. The current status is encouraging but even after more than twenty years of intense worldwide research, there are still many new avenues to be explored. Innovative functional oxide materials science is central to future progress. This paper discusses three key areas of our research focusing on new directions: highly tailored flux pinning using the new core pinning additives R3TaO7 and RBa2NbO6 for control of nanostructure formation; pinning using magnetic phase additives such as RFeO3 with the potential for a magnetic contribution to the flux pinning; and the use of liquid assisted growth enabling very high growth rates leading to thick films with no critical current degradation. 1. Introduction Over the last decade, progress in the manufacture of high quality, long length, high current coated conductors based on the second generation superconducting material YBa2Cu3O7−δ (YBCO) for a variety of transport and magnetic applications has been impressive [1]. In particular, in just the past few years, great progress has been made with practical flux pinning approaches [2, 3]. There has also been some success in minimizing the complicated series of buffer and seed layers required between metal substrate and YBCO coating [4]. Less attention has been paid to exploring radical new routes to forming conductors by fast, more cost-effective routes. Of course, the weak link problem of YBCO grains [5] means that high angle grain boundaries need to be circumvented and this is no mean feat. The future for coated conductors will involve attaining the required performance at minimal cost. Approaches involving both optimization and further scaling up of the current processing routes as well as exploring new horizons for achieving higher performance by more scalable routes are required. The areas which we believe represent new frontiers and could ultimately lead to a step-change include highly tailored core pinning, practical magnetic pinning, and the use of rapid liquid assisted growth [6]. 2. Experimental details Composite deposition targets were prepared in-house from commercial YBa2Cu3Ox powder (SCI Engineered Materials, 99.99%) together with the appropriate additives of R2O3 and Ta2O5 (for R3TaO7 where R is a rare earth) or Y2O3, BaCO3 and Nb2O5 (for YBa2NbO6) or Y2O3 and FeO (for YFeO3). Targets of varying compositions from 0.5 to 10 mol% were prepared by mixing, pressing and sintering at 985 ◦C in flowing oxygen for 12–24 h. A pure YBCO target was also prepared under the same conditions for control purposes. The substrates used were single crystal SrTiO3(001) or buffered metallic tapes (∼150 nm La2Zr2O7 grown on RABiTS Ni–W by dip coating and ex situ processing [7]). Samples were grown in a deposition atmosphere of 30 Pa of flowing oxygen at either 760–795 ◦C by standard pulsed laser deposition (PLD) or at 815 ◦C by PLD incorporating hybrid liquid phase epitaxy (HLPE). In the HLPE process, a sub-micron thick BaO–CuO liquid layer is first deposited and 0953-2048/10/034009+05$30.00 © 2010 IOP Publishing Ltd Printed in the UK1 Supercond. Sci. Technol. 23 (2010) 034009 J L MacManus-Driscoll et al Figure 1. Cross-sectional TEM of films of YBCO + 5 mol% Gd3TaO7 and YBCO + 5 mol% YBa2NbO6 pinning additions grown under similar conditions, namely 770 ◦C and 790 ◦C respectively at 5 Hz on SrTiO3. Self-assembled nanocolumns of the respective pinning addition are indicated with arrows. growth of the YBCO occurs under this layer [8]. Subsequent to deposition, all samples were annealed in situ in a static oxygen pressure of 50 kPa to achieve optimum oxygenation. Film thicknesses were typically in the range 0.5–2.5 μm. Films were structurally characterized by x-ray diffraction (XRD) and transmission electron microscopy (TEM). Mag- netic properties were measured in a cryogenic vibrating sample magnetometer (VSM) equipped with a 1 T electromagnet. Electrical properties were measured by a standard four- probe technique in an 8 T superconducting magnet following photolithographic patterning and ion beam etching to form bridge structures of typically 50–125 μm width and 1 mm length. For angular measurements of the critical current density, Jc, the field was applied in the maximum Lorentz force configuration and rotated from normal to the film surface (θ = 0◦) to parallel to it (θ = 90◦). 3. Results 3.1. Core pinning Several new approaches to the enhancement of core pinning in YBCO through the incorporation of novel nanostructural inclusions have been tested and developed. The most widely studied and successful method so far involves incorporation of BaZrO3 nanoparticles but there are many other additions which are also effective [9]. The current research focus should now be on tailoring nanopinning additions to achieve desired and controlled arrays, the optimal array (whether it be random, correlated or a combination of both geometries) being dependent on the application field and temperature, and the growth method to be used. In order to nanoengineer the materials, the kinetics and thermodynamics of phase formation of the particular phase(s) formed when additions are made to YBCO films needs to be understood. The mechanism of formation of the nanostructured inclusions depends on the lattice mismatch with the YBCO and the kinetics of self- assembly (mobility of the constituent ions) which depend on the growth temperature relative to the melting point of the phase, but is also readily controllable by PLD by altering the growth rate. Table 1. Pinning ion additions made to YBCO films, stable phases formed, and lattice misfits to YBCO. Pinning ion addition Stable phase in YBCO film Misfit to YBCO ab Misfit to YBCO c Ta5+ Yb3TaO7 −4.8%a −10.9% Ta5+ Gd3TaO7 −2.4%a −8.7% Zr4+ BaZrO3 +8.4%b +7.5% Nb5+ YBa2NbO6 +9.3%b +8.5% a Rotated 45◦ to cube-on-cube. b Cube-on-cube. With the knowledge that Group IV and Group V ions do not readily substitute into the YBCO lattice, but instead form stable heteroepitaxial second phases [10], we have investigated Ta5+ and Nb5+ additions to YBCO films [11–13] in the form of Yb3TaO7 (a = 1.039 nm), Gd3TaO7 (a = 1.065 nm) and YBa2NbO6 (a = 0.8436 nm), as alternatives to the common Zr4+ ion addition in the form of BaZrO3 (a = 0.4181 nm). The lattice mismatches of each of these phases to YBCO (a = 0.3828 nm, b = 0.3887 nm, c = 1.1665 nm), calculated from the bulk lattice parameters, for the most promising core pinning additions we have developed are shown in table 1. Figure 1 compares TEM images of nanocolumns of Gd3TaO7 and YBa2NbO6, showing Gd3TaO7 nanocolumns to be straighter and more continuous than YBa2NbO6 nanocolumns. If we also consider the known form of BaZrO3 nanocolumns [14], we find that lattice misfit and nanocolumn size and perfection are interrelated, the finest and most continuous columns arising from lower in-plane lattice misfit to YBCO [11]. Angular critical current measurements of so far optimized R3TaO7-added samples show the superior flux pinning obtained with these additives compared to pure films. By growing under conditions which produce a composite random nanoparticle/correlated nanocolumn microstructure, Jc values of 1.1 MA cm−2 at 1 T (figure 2(a)) and 0.7 MA cm−2 at 3 T were obtained compared to 0.2 MA cm−2 and 0.04 MA cm−2 at the same fields for pure YBCO (all at 77 K, H ‖ c). The YBa2NbO6-added samples generally show broader c- axis peaks which are consistent with the more discontinuous columns obtained (figure 2(b)). 2 Supercond. Sci. Technol. 23 (2010) 034009 J L MacManus-Driscoll et al Figure 2. Angular Jc data at 77 K for 0.5 μm thick YBCO films with additions of Yb3TaO7 or YBa2NbO6. (a) Data at 1 T for 5 mol% Yb3TaO7 additions compared to pure YBCO showing a large enhancement in Jc, particularly for fields applied in the c direction. (b) Data at 0.5 T comparing 1.5 mol% Yb3TaO7 and 5 mol% YBa2NbO6 additions normalized to Jc ‖ c in order to emphasize the broader c-axis peak obtained for YBa2NbO6. 3.2. Pinning by magnetic phase additions In practical high temperature superconductors, even with the introduction of artificial pinning centres, the self-field Jc is still 5–10 times lower than the theoretical limit. The realizable advantages of magnetic pinning over ‘standard’ non-magnetic pinning originate from the additional magnetic interaction that this provides [15]. Whereas normal materials interact with the superconducting order parameter via the coherence length ξ (on the order of nanometres in the high- Tc materials), magnetic materials have a potential interaction over length scales of the penetration depth λ which is much larger (hundreds of nanometres). This larger interaction volume implies that a lower density of pinning centres will be required to achieve effective pinning and that consequently the disruption to the superconducting phase (and its resultant properties) will be minimized. Beyond the basic studies of magnetic pinning [16–18], the main hurdle to overcome is to introduce magnetic material in such a way as to avoid significant ‘poisoning’ of the superconducting phase. This requires forming very stable second phases that effectively ‘lock away’ the magnetic ion addition while remaining magnetic themselves. Our exploratory studies [19] have indicated YFeO3 to be a suitable addition. YFeO3 belongs to the family of rare earth orthoferrites, RFeO3, which exhibit weak ferromagnetism as a result of a slight canting of their predominantly antiferromagnetically coupled Fe3+ moments [20]. Their structure is that of four distorted perovskite units assembled into an orthorhombic unit cell (a = 0.5282 nm, b = 0.5596 nm, c = 0.7605 nm) [21], thereby suggesting a high likelihood of structural compatibility with YBCO. There is no measurable reduction in the superconducting critical temperature Tc for a 1 mol% YFeO3 addition [22, 23], indicating an effective segregation of the magnetic species. Randomly dispersed magnetic nanoparticles of sub-5 nm size (upper right panel of figure 3) were observed leading to a 2– 3 times increase in Jc at self-field and a 10 times increase at 6 T compared to pure YBCO. Increasing the dopant amount leads to a slight suppression of Tc and consequent reduction in Figure 3. Angular Jc measurements at 77 K, 1 T of a 1 μm thick film of YBCO with 1 mol% YFeO3 additions compared with a control sample of pure YBCO. Upper right panel shows a cross-sectional TEM image of a 1 mol% magnetically doped film, with arrows indicating the positions of several sub-5 nm dopant nanoparticles. Lower right panel is the room temperature (RT) magnetic hysteresis loop (magnetic moment m versus applied field H ) of a sample with 5 mol% YFeO3 additions. the Jc enhancement. The particles have a good in-plane lattice parameter match to YBCO resulting in a strain of just 0.12% in-plane (for a 45◦ in-plane rotation of the YFeO3 unit cell relative to the YBCO) and 2.4% out-of-plane. The coercivity μ0 Hc is found to be rather small at around 6 mT (lower right panel of figure 3), due to the small particle size causing a superparamagnetic response at the temperatures of interest, and the measured saturation field is around 0.2 T. The random particle arrangement results in random pinning producing an upwards shift in Jc for all field angles (figure 3). 3.3. Rapid, innovative growth of coated conductors Rapid liquid assisted growth of YBCO + 5 mol% BaZrO3 coated conductors was achieved on metallic substrates by HLPE. Samples of two different thicknesses were grown in 4 min (1 μm film) and 10 min (2.5 μm film). The XRD rocking curves of the (004) peak of the La2Zr2O7 buffer gave a rolling direction FWHM of 5.4◦ and a perpendicular direction FWHM of 9.1◦. The XRD rocking curves for the 3 Supercond. Sci. Technol. 23 (2010) 034009 J L MacManus-Driscoll et al Figure 4. Characterization of coated conductors of YBCO + 5 mol% BaZrO3 grown by HLPE on ∼100 nm dip-coated La2Zr2O7-buffered RABiTS Ni-W. (a) X-ray rocking curves of YBCO layer and La2Zr2O7 buffer layer, and (b) Jc as a function of field applied parallel to c and (inset) as a function of applied field angle θ at 1 T. (005) peak of the YBCO layer grown on the buffer by HLPE are improved over the buffer by approximately 0.7◦ in each case, to 4.7◦ and 8.3◦, respectively (figure 4(a)). Transport measurements (figure 4(b)) show that high Jc material can be achieved on a rapidly grown, dip-coated buffer and that there is no degradation in the Jc of the superconductor with thickness. Hence, it is relatively easy to grow 5 μm thick material in situ in under 20 min and still achieve self-field Jcs of around 1 MA cm−2. The several advantages of the HLPE method can be summarized as follows: (a) the film texture is improved over the substrate and so the method allows for the use of much cheaper, rapidly grown substrate-buffer materials; (b) the growth rate is several times faster than standard methods; (c) there is no degradation in Jc with thickness. Finally, we note that for a given Ic, a potential advantage in being able to grow thick material having a moderate Jc over thinner material with higher Jc is that preliminary investigations indicate that flux creep rates are lower in thicker material [24]. 4. Conclusions There are several potential new avenues to be explored for achieving low cost, high performance superconducting conductors. Highly tailored core pinning, practical magnetic pinning, and rapid liquid assisted growth are key areas. Phases formed from Group IV and V ion additions are ideal for achieving strong, tunable core pinning. The strain between the particles and the YBCO matrix influences the form of their distribution which directly impacts on the field and temperature dependent pinning behaviour. The main challenge for practical magnetic pinning is to form stable, non-poisoning phases. YFeO3 is one such phase and has been shown to enhance pinning at low (1 mol%) levels. Finally, rapid, easy growth of thick YBCO by HLPE on ∼100 nm dip-coated La2Zr2O7- buffered RABiTS Ni–W gives conductors with Ic > 250 A grown in a matter of minutes. Acknowledgments This work has been supported by the European Commission (Marie Curie Excellence Grant ‘NanoFen’, MEXT-CT-2004- 014156), the UK EPSRC (‘Novel materials nanoengineering for enhanced performance of superconducting coated conduc- tors and functional oxides’), and the EU RTN ‘NESPA’. SAH is supported by the Mays Wild Research Fellowship of Downing College, Cambridge. SCW is supported by The Leverhulme Trust with supplementary funding from The Isaac Newton Trust. References [1] American Superconductor, Superpower, Southwire 2009 USDOE High Temperature Superconductivity Program Peer Review (Alexandria, VA, Aug. 2009) available at http://www. htspeerreview.com [2] Yamada Y et al 2005 Appl. Phys. Lett. 87 132502 [3] Kang S et al 2006 Science 311 1911 [4] Stan L, Feldmann D M, Usov I O, Holesinger T G, Maiorov B, Civale L, DePaula R F, Dowden P C and Jia Q X 2009 IEEE Trans. Appl. Supercond. 19 3459 [5] Dimos D, Chaudhari P and Mannhart J 1990 Phys. Rev. B 41 4038 [6] Ohnishi T, Huh J-U, Hammond R H and Jo W 2004 J. Mater. Res. 19 977 [7] University of Cambridge and Nexans SuperConductors GmbH, Chemiepark Knapsack, 50351 Hu¨rth, Germany 2008 European Patent Application 201373 [8] Kursumovic A, Maiorov B, Durrell J H, Wang H, Zhou H, Stan L, Harrington S, Wimbush S, Holesinger T G and MacManus-Driscoll J L 2009 Supercond. Sci. Technol. 22 015009 [9] Foltyn S R, Civale L, MacManus-Driscoll J L, Jia Q X, Maiorov B, Wang H and Maley M 2007 Nat. Mater. 6 631 [10] Driscoll J L and Foltyn S R 2004 US Patent Specifiation 20060025310 [11] Harrington S A, Durrell J H, Maiorov B, Wang H, Wimbush S C, Kursumovic A, Lee J H and MacManus-Driscoll J L 2009 Supercond. Sci. Technol. 22 022001 [12] Harrington S A, Durrell J H, Wang H, Wimbush S C, Tsai C F and MacManus-Driscoll J L 2009 Nanotechnology at press [13] Ercolano G, Harrington S A, Wang H, Tsai C F and MacManus-Driscoll J L 2009 Supercond. Sci. Technol. at press [14] Mele P, Matsumoto K, Horide T, Ichinose A, Mukaida M, Yoshida Y, Horii S and Kita R 2008 Supercond. Sci. Technol. 21 032002 [15] Blamire M G, Dinner R B, Wimbush S C and MacManus-Driscoll J L 2009 Supercond. Sci. Technol. 22 025017 4 Supercond. Sci. Technol. 23 (2010) 034009 J L MacManus-Driscoll et al [16] Jan D B, Coulter J Y, Hawley M E, Bulaevskii L N, Maley M P, Jia Q X, Maranville B B, Hellman F and Pan X Q 2003 Appl. Phys. Lett. 82 778 [17] Lange M, Van Bael M J, Moshchalkov V V and Bruynseraede Y 2002 Appl. Phys. Lett. 81 322 [18] Yang Z, Lange M, Volodin A, Szymczak R and Moshchalkov V V 2004 Nat. Mater. 3 793 [19] Wimbush S C, Yu R, Bali R, Durrell J H and MacManus-Driscoll J L 2009 Physica C at press doi:10.1016/j.physc.2009.10.117 [20] Treves D 1962 Phys. Rev. 125 1843 [21] Coppens P and Eibschu¨tz M 1965 Acta Crystallogr. 19 524 PDF 73-1345 [22] Wimbush S C, Durrell J H, Bali R, Yu R, Wang H, Harrington S A and MacManus-Driscoll J L 2009 IEEE Trans. Appl. Supercond. 19 3148 [23] Wimbush S C, Durrell J H, Tsai C F, Wang H, Jia Q X, Blamire M G and MacManus-Driscoll J L 2009 Supercond. Sci. Technol. submitted [24] Feldmann M, Holesinger T, Maiorov B and Civale L 2009 Results presented at the USDOE High Temperature Superconductivity Program Peer Review (Alexandria, VA, Aug. 2009) available at http://www.htspeerreview.com 5 State-of-the-art flux pinning in YBa2Cu3O7-δ by the creation of highly linear, segmented nanorods of Ba2(Y/Gd)(Nb/Ta)O6 together with nanoparticles of (Y/Gd)2O3 and (Y/Gd)Ba2Cu4O8 G. Ercolano1, M. Bianchetti1, S. C. Wimbush1, S. A. Harrington1, H. Wang2, J. H. Lee2 and J. L. MacManus-Driscoll1 1Department of Materials Science and Metallurgy, University of Cambridge, Pembroke Street, Cambridge CB2 3QZ, UK 2Department of Electrical and Computer Engineering, Texas A&M University, College Station, TX 77843, USA E-mail: ge228@cam.ac.uk Abstract. Self-assembled, segmented nanorods of c-axis aligned Ba2(Y/Gd)(Nb/Ta)O6 as well as randomly-distributed nanoparticles of (Y/Gd)2O3 and (Y/Gd)Ba2Cu4O8 were grown into YBa2Cu3O7-δ (YBCO) thin films by pulsed laser deposition. The complex pinning landscape proves to be extremely effective, particularly at higher fields where the segmented vortices yield a plateau in critical current density (Jc) with field angle around 60°. In 0.3 µm thick films, the Jc values are higher than 1 MAcm-2 at 2.5 T (H||c-axis). Owing to the combined interactions of the vortices with the different pinning centres, interesting new features are observed at high fields in the angle-dependence of Jc. 1. Introduction Even though conductors based on YBa2Cu3O7-δ (YBCO) have reached an extremely sophisticated stage of technological development with exemplary critical current performance arising from nanoengineering of pinning centres [1-8], since higher current conductors mean lower production costs, achieving even stronger flux pinning by simple means is still a highly sought-after goal. Probably the most successful method for engineering flux pinning in YBCO is by the introduction of an epitaxial non-superconducting secondary phase, e.g. BaZrO3 [9-11]. Ta and Nb additions to YBCO have more recently been studied based on the fact that they are highly charged (5+) ions that should perform in a similar way to Zr and not substitute into the YBCO lattice. Tantalate additions to YBCO produce excellent pinning performance via the formation of very fine, dense nanorods whose composition has been ascribed to either a defective pyrochlore, RE3TaO7 (RE = rare earth, Gd being the most widely studied) [12], or to the double perovskite, Ba2YTaO6 [13]. Niobate additions produce Ba2RENbO6 nanorods [14-17] which are similar to BaZrO3 nanorods — wider, shorter and less linear than the tantalate rods [14] (diameter ~ 10–15 nm, splayed around the c-axis). The niobate rods remain highly effective pinning centres at low fields aligned along the c-axis, although considering the entire field range tantalate rods are superior overall. Since the niobate and tantalate nanorods in YBCO are of rather different morphology and hence give different performance characteristics, it is interesting to consider whether addition of both of these second phases of the same overall level results in an averaging effect or whether further complexity is induced in the system, thereby yielding an entirely different pinning landscape. In fact, the situation is a mix of the former and the latter with the final result that the nanostructure created fortuitously yields an overall superior pinning performance, particularly at higher fields. 2. Experimental methods Films were grown by pulsed laser ablation of a composite target of appropriate composition. Targets were prepared by mixing and grinding pure YBa2Cu3O7-δ powder (SCI Engineered Materials 99.99%) with 2.5 mol% of Ba2YNbO6 powder and 2.5 mol% of Gd3TaO7. Ba2YNbO6 powder was produced by mixing and grinding stoichiometric quantities of 99.99% Y2O3, Ba(NO3)2 and Nb2O5 followed by solid state reaction at Confidential: not for distribution. Submitted to IOP Publishing for peer review 27 January 2011 1450°C for 24 h in flowing O2. Gd3TaO7 was introduced in the target by adding the desired amount of 99.99% Gd2O3 and 99.99% Ta2O5 powders to the YBa2Cu3O7-δ powder. The target mixture was pressed in a cylindrical die (diameter = 2 cm) and sintered at 950°C for 12 h in flowing O2. Pulsed laser ablation was performed using a Lambda Physik KrF excimer laser (λ = 248 nm). The substrates were single crystal SrTiO3 (100) held at 780°C, and the substrate temperature was monitored using a pyrometer. 4500 laser pulses at a 1 Hz repetition rate were used to produce 0.3 µm thick films. Films where post-annealed in situ at 520°C for 1 h in 500 mbar O2 atmosphere. Phase and orientation analysis were performed using x-ray diffraction. Cross sectional transmission electron microscopy (TEM) was used to determine the size, distribution and structure of the nanoparticles and nanorods. A conventional four-point electrical measurement on photolithographically patterned bridges of 50 µm width was used to determine the critical current density, Jc, using a 1 µVcm-1 criterion. The angular dependence of Jc was measured by rotating the applied magnetic field in a plane perpendicular to the current direction (maximum Lorentz force configuration). 3. Results and discussion Figure 1. Bragg-Brentano scan of a YBa2Cu3O7-δ + 2.5 mol% Gd3TaO7 + 2.5 mol% Ba2YNbO6 sample. Figure 1 shows the x-ray diffraction pattern of a typical film. The (00l) peaks of YBCO as well as the (002) and (004) peaks of Ba2(Y/Gd)(Nb/Ta)O6 (2θ = 21.5°, 2θ = 43.2°) are labelled. A peak at 2θ = 33.3° is identified as the (004) peak of (Y/Gd)2O3. (Y/Gd)Ba2Cu4O8 is also present as evidenced by the (008), (0010), and (0012) peaks (2θ = 25.9°, 2θ = 33.2°, 2θ = 41.4°). A balanced chemical reaction consistent with the phases observed by XRD is shown below: 0.95 YBa2Cu3O7-δ + 0.025 Ba2YNbO6 + 0.025 Gd3TaO7  0.85 (Y/Gd)Ba2Cu3O7-δ + 0.05 Ba2(Y/Gd)(Nb/Ta)O6 + 0.0375 (Y/Gd)2O3 + 0.075 (Y/Gd)Ba2Cu4O8 Gd and Y can easily cross-substitute for one another since they have similar ionic radii. For the same reason, Nb and Ta can also cross-substitute. This allows the double perovskite Ba2(Y/Gd)(Nb/Ta)O6 to grow within the (Y/Gd)Ba2Cu3O7-δ lattice. Furthermore the excess Gd and the Ba deficiency produced by the Gd3TaO7  Ba2(Y/Gd)(Nb/Ta)O6 transformation together with the Gd  Y cross-substitution is the origin of the (Y/Gd)2O3 and (Y/Gd)Ba2Cu4O8 formation. The previously reported [12] Gd3TaO7 phase does not form here because in the presence of Nb the Ba2(Y/Gd)(Nb/Ta)O6 phase is more stable. (Y/Gd)Ba2Cu4O8 forms because of the barium deficiency in the Gd3TaO7 reactant. YBa2Cu4O8 has previously been observed in coated conductors where there is barium deficiency owing to its depletion by reaction with CeO2 [18]. Also YBa2Cu4O8 has been reported in films grown by metal-organic deposition, where it is in the form of stacking fault defects [19]. Figure 2. X-ray φ scans of the (202) Ba2(Y/Gd)(Nb/Ta)O6 and the (404) (Y/Gd)2O3 peaks from a YBa2Cu3O7-δ + 2.5 mol% Gd3TaO7 + 2.5 mol% Ba2YNbO6 sample. Ba2(Y/Gd)(Nb/Ta)O6 is aligned cube-on-cube with the (Y/Gd)Ba2Cu3O(7-δ) as shown in the x-ray φ scans of figure 2. The (Y/Gd)2O3 is rotated both 45° in-plane, as previously reported [20] but also a few degrees away (with a broad range of angles) from the cube-on-cube orientation. These broader rotations are consistent with near coincidence site lattice matching to accommodate the very large strains and structural difference between (Y/Gd)2O3 and YBCO [21]. Figure 3. a) TEM image of the YBa2Cu3O7-δ + 2.5 mol% Gd3TaO7 + 2.5 mol% Ba2YNbO6 sample cross section; b) TEM image of a plate-like nanoparticle nucleated between two rod segments; c) Selected area electron diffraction pattern of the image in (a); d) TEM image of a plate-like nanoparticles of (Y/Gd)2O3. Fourier transform of the image of particle. TEM cross-sections show that there are at least two different nanoscale structural features present. Firstly, there is a set of c-axis aligned fine rods of ~7 nm in diameter, significantly smaller than the pure niobate rods of Ba2YNbO6 which are ~15 nm in diameter [14, 15], but larger than the pure tantalate rods of Gd3TaO7 which are ~5 nm in diameter [12]. Hence, the averaged growth kinetics of pure niobate and pure tantalate rods is obtained, as might be expected. Secondly, figure 3b shows a plate-like nanoparticle of RE2O3 nucleated along the ab-plane between two Ba2(Y/Gd)(Nb/Ta)O6 nanorods. Fig. 3c shows a selected area diffraction pattern of a region of the matrix and inclusions. Ba2(Y/Gd)(Nb/Ta)O6, (Y/Gd)2O3 and YBa2Cu4O8 are present, consistent with the x-ray diffraction pattern of figure 1. A Fourier transform confirming the structure of a (Y/Gd)2O3 particle of 25 nm width is shown in figure 3d inset. Other particles of (Y/Gd)2O3 were observed to be in the 25 – 30 nm width range. A fundamentally new structural feature, which has not been observed before, is the self-segmentation of the nanorods, yielding an average rod segment length of 30 nm. Kinetic effects of slower diffusion of the heavy Ta5+ ion compared to the Bb5+ ion may prevent rods from maintaining continuity as the film growth progresses. Similar to BaZrO3 and Ba2YNbO6 rods and because of the nucleation enhancing strain field generated within the YBCO, a new segment nucleates aligned along the c-axis with a segment below it [22]. The average rod length (30 nm) is shorter than both the continuous tantalate rod length (hundreds of nm) and the short niobate rods (~80-100 nm). Despite the complexity of sample composition, the possibility of varying RE concentration across samples, as well as the three different nanoinclusion phases observed, no suppression of the superconducting transition temperature was observed. The Tc of ~ 89 K for all the samples indicates no substitution of Nb or Ta into the YBCO matrix, as might be expected for these higher valence ions. Figure 4. Jc versus applied magnetic field at 77 K, H || c for samples with Nb, Ta, Nb-Ta and Zr doping compared to a pure YBCO sample. Data on a log-linear plot (a) and a log-log plot (b). The (Zr) YBCO data is taken from [23]. Jc(H) for fields applied parallel to the c-axis for the films of this study compared with the best reported films [23] of similar thickness (0.3µm) with other additions (Ta, Nb or Zr) is shown in figure 4. The sample with simultaneous addition of Nb and Ta has the highest Jc across the entire field range measured with a value of 1.8 MAcm-2 at 1 T and values above 1 MAcm-2 up to 2.5 T. Furthermore, as shown in figure 4b, the linear Jc decay range is increased from less than 1 T for pure YBCO, YBCO + 5 mol% Ba2YNbO6 ((Nb) YBCO) and YBCO + 5 mol% Gd3TaO7 ((Ta) YBCO) to ~2 T for combined Nb+Ta ((Nb-Ta) YBCO) samples, exceeding even that of YBCO + BaZrO3 ((Zr) YBCO) (2% of the YBCO target surface area made of YSZ [23]). The matching field value calculated from the nanorod spacing of ~28 nm observed in the TEM image of figure 3a is ~2.6 T, consistent with the Jc plateau up to 2.5 T of figure 4. The α values calculated for the different additions are reported in table 1. The value of 0.14 for the (Nb-Ta) YBCO samples is lower than the best previously reported value which was 0.19 in a specially grown YBCO + BaZrO3 sample where it was shown that optimisation of the pinning landscape could produce both randomly distributed nanoparticles and splayed columnar defects which together strongly reduce the depinning [24]. In fact, segmented rods were produced which gave rise to staircase vortices. Table 1. Calculated α values for the different additions. Sample α YBCO (Nb) YBCO (Ta) YBCO (Nb-Ta) YBCO (Zr) YBCO 0.47 0.35 0.26 0.14 0.25 Figure 5. Angular dependence of Jc measured at 77 K for samples with Nb, Ta, and Nb-Ta doping compared to a pure YBCO sample at 1 T (a) and 3 T (b). Superior properties were also found in the angular dependence of Jc of the (Nb-Ta) YBCO samples of this study. A strong, narrow c-axis pinning peak at low fields (1 T, figure 5a) changed to a strong, broad c-axis peak as the applied field increased (3 T, figure 5b). The broad peak is a feature already observed in (Nb) YBCO [14]. However the Jc values of the samples prepared in this work are higher than previously observed indicating more effective pinning by the finer, segmented rods compared to the coarser, shorter, non- segmented rods observed in the (Nb) YBCO samples. Figure 6. a) High field angular dependence of Jc measured at 77 K for a sample with (Nb-Ta) doping, applied fields ranging from 3 T to 6 T; b) A sketch of vortices interacting simultaneously with rod segments along the c-axis and intrinsic and extrinsic defects along the ab-planes. Figure 6a shows the evolution of the angular Jc with increasing field. At low fields, the most prominent feature is the strong c-axis pinning peak contrasted with relatively weak ab-plane peaks (figure 5a). A shoulder initially observed on the c-axis peak resolves itself at fields beyond 3 T into distinct peaks at around +/- 60 degrees, that grow in magnitude with increasing field to dominate all other types of pinning at fields above 5 T. The existence of this additional preferential pinning direction is an interesting new feature arising from the novel pinning landscape that has been established in these samples. It can be explained in terms of a vortex path model [25] in which the combination of c-axis pinning structures (segmented Ba2(Y/Gd)(Nb/Ta)O6 nanorods) and ab-plane pinning structures (plate-like (Y/Gd)2O3 particles) results in staircase vortices that experience the strongest pinning at a characteristic angle determined by the particular distribution of defect lengths present (figure 6b). This type of pinning is most effective at high fields because it utilises both available species of pinning structure and therefore has the capacity to strongly pin a greater overall vortex length. 4. Conclusion A completely new self-assembled pinning landscape was produced in laser deposited YBCO films by combining the synergistic effects from both Ta and Nb additions which together yield an optimal nanorod architecture of fine, straight, segmented rods. (Y/Gd)2O3 nanoparticles were also generated in the matrix for reasons of cation compensation, namely because the rods of Ba2(Y/Gd)(Nb/Ta)O6 formed were of different average composition (poorer in rare earth, richer in barium) than the reactants of Ba2YNbO6 and Gd3TaO7 which were added to the YBCO PLD target. Greatly enhanced current densities were achieved at 77 K for applied fields higher than 2 T compared to previously studied pinning additives. A Jc of 1 MAcm-2 was achieved at 2.5 T and 77 K in 0.3 µm thick films which is a new performance benchmark. In addition, new pinning features around 60° field angle were observed at > 3 T with the potential to extend the angular range of operation of YBCO-based coated conductors. 5. Acknowledgements Work carried out at the University of Cambridge was performed under the auspices of NESPA, Nano- Engineered Superconductors for Power Application, a framework of the Marie Curie Research Training Network, funded within the EUs 6th framework program. The effort at Texas A&M University was supported by NSF-1007969. JLM-D acknowledges ERC Grant No. ERC-2009-adG 247276. SCW is funded by The Leverhulme Trust, with supplementary funding from The Isaac Newton Trust. 6. References [1] Haugan T, Barnes P N, Wheeler R, Meisenkothen F, Sumption M. Addition of nanoparticle dispersions to enhance flux pinning of the YBa2Cu3O7-x superconductor. 2004, Nature, 430, 867-870. [2] Civale L, Marwick A D, Worthington T K, Kirk M A, Thompson J R, Krusinelbaum L, Sun Y, Clem J R, Holtzberg F. Vortex confinement by columnar defects in YBa2Cu3O7 crystals: Enhanced pinning at high fields and temperatures. 1991, Phys. Rev. Lett., 67, 648-651. [3] Gutierrez J, Llordes A, Gazquez J, Gibert M, Roma N, Ricart S, Pomar A, Sandiumenge F, Mestres N, Puig T, Obradors X. Strong isotropic flux pinning in solution-derived YBa2Cu3O7-x nanocomposites superconductor films. 2007, Nature Materials, 6, 367-373. [4] Puig T, Gutierrez J, Pomar A, Llordes A, Gazquez J, Ricart S, Sandiumenge F, Obradors X. Vortex pinning in chemical solution nanostructured YBCO films. 2008 Supercond. Sci. Technol., 21, 034008. [5] Sparing M, Backen E, Freudenberg T, Huhne R, Rellinghaus B, Schultz L, Holzapfel B. Artificial pinning centres in YBCO thin films induced by substrate decoration with gas-phase-prepared Y2O3 nanoparticles. 2007, Supercond. Sci. Technol., 20, S239-S246. [6] Mele P, Matsumoto K, Ichinose A, Mukaida M, Yoshida Y, Horii S, Kita R. Systematic study of the BaSnO3 insertion effect on the properties of YBa2Cu3O7-x films prepared by pulsed laser ablation. 2008, Supercond. Sci. Technol., 21, 125017. [7] Wimbush S C, Durrell J H, Bali R, Yu R, Wang H , Harrington S A, MacManus-Driscoll J L. Practical magnetic pinning in YBCO. 2009, IEEE Trans. Appl. Supercond., 19, 3148-3151. [8] Reich E, Thersleff T, Huhne R, Iida K, Schultz L, Holzapfel B. Structural and pinning properties of Y2Ba4CuMOy (M = Nb, Zr)/ YBa2Cu3O7-δ quasi-multilayers fabricated by off-axis pulsed laser deposition. 2009, Supercond. Sci. Technol., 22, 105004. [9] Macmanus-Driscoll J L, Foltyn S R, Jia Q X, Wang H, Serquis A, Civale L, Maiorov B, Hawley M E, Maley M P, Peterson D E. Strongly enhanced current densities in superconducting coated conductors of YBa2Cu3O7-x + BaZrO3. 2004, Nature Materials, 3, 439-443. [10] Mele P, Matsumoto K, Horide T, Ichinose A, Mukaida M, Yoshida Y, Horii S, Kita R. Incorporation of double artificial pinning centers in YBa2Cu3O7-δ films. 2008, Supercond. Sci. Technol., 21, 015019. [11] Mele P, Matsumoto K, Horide T, Ichinose A, Mukaida M, Yoshida Y, Horii S, Kita R. Ultra-high flux pinning properties of BaMO3-doped YBa2Cu3O7-x thin films (M = Zr, Sn). 2008, Supercond. Sci. Technol., 21, 032002. [12] Harrington S A, Durrell J H, Maiorov B, Wang H, Wimbush S C, Kursumovic A, Lee J H, MacManus- Driscoll J L. Self-assembled, rare earth tantalate pyrochlore nanoparticles for superior flux pinning in YBa2Cu3O7-δ films. 2009, Supercond. Sci. Technol., 22, 022001. [13] Wee S H, Goyal A, Specht E D, Cantoni C, Zuev Y L, Selvamanickam V, Cook S. Enhanced flux pinning and critical current density via incorporation of self-assembled rare-earth barium tantalate nanocolumns within YBa2Cu3O7-δ films. 2010, Phys. Rev. B, 81, 14050. [14] Ercolano G, Harrington S A, Wang H, Tsai C F, MacManus-Driscoll J L. Enhanced flux pinning in YBa2Cu3O7-δ thin films using Nb-based double perovskite additions. 2010, Supercond. Sci. Technol., 23, 022003. [15] Wee S H, Goyal A, Zuev Y L, Cantoni C, Selvamanickam V, Specht E D. Formation of Self- Assembled, Double-Perovskite, Ba2YNbO6 Nanocolumns and Their Contribution to Flux-Pinning and Jc in Nb-Doped YBa2Cu3O7-δ Films. 2010, Appl. Phys. Express, 3, 023101. [16] Kai H, Horii S, Ichinose A, Kita R, Matsumoto K, Yoshida Y, Fujiyoshi T, Teranishi R, Mori N, Mukaida M. The effects of growth temperature on c-axis-correlated pinning centers in PLD ErBa2Cu3O7-δ films with Ba(Er0.5Nb0.5)O3. 2010, Supercond. Sci. Technol., 23, 025017. [17] Horii S, Yamada K, Kai H, Ichinose A, Mukaida M, Teranishi R, Kita R, Matsumoto K, Yoshida Y, Shimoyama J, Kishio K. Introduction of c-axis-correlated 1D pinning centers and vortex Bose glass in BaNbO doped ErBa2Cu3Oy films. 2007, Supercond. Sci. Technol., 20, 1115-1119. [18] Holesinger T G, Foltyn S R, Arendt P N, Kung H, Jia Q X, Dickerson R M, Dowden P C, DePaula R F, Groves J R, Coulter J Y. The microstructure of continuosly processed YBa2Cu3Oy coated conductors with underlying CeO2 and ion 'beam' assisted yttria stabilized zirconia buffer layers. 2000, Jour. of Mat. Res., 15, 1110-1119. [19] Specht E D, Goyal A, Li J, Martin P M, Li X, Rupich M W. Stacking faults in YBa2Cu3O7-x: Measurement using x-ray diffraction and effects on critical current. 2006, Appl. Phys. Lett., 89, 162510. [20] Gapud A A, Kumar D, Viswanathan S K, Cantoni C, Varela M, Abiade J, Pennycook S J, Christen D K. Enhancement of flux pinning in YBa2Cu3O7-δ thin films embedded with epitaxially grown Y2O3 nanostructures using a multi-layering process. 2005, Supercond. Sci. Technol., 18, 1502-1505. [21] Garrison S M, Newman N, Cole B F, Char K, Barton R W. Observation of 2 in-plane epitaxial states in YBa2Cu3O7-δ films on yttria-stabilized ZrO2. 1991, Appl. Phys. Lett., 58, 2168-2170. [22] Harrington S A, Durrell J H, Wang H, Wimbush S C, Tsai C F, MacManus-Driscoll J L. Understanding nanoparticle self-assembly for a strong improvement in functionality in thin film nanocomposites. 2010, Nanotechnology, 21, 095604. [23] Mele P, Matsumoto K, Horide T, Ichinose A, Mukaida M, Yoshida Y, Horii S. Enhanced high-field performance in PLD films fabricated by ablation of YSZ-added YBa2Cu3O7-x target. 2007, Supercond. Sci. Technol., 20, 244-250. [24] Maiorov B, Baily S A, Zhou H, Ugurlu O, Kennison J A, Dowden P C, Holesinger T G, Foltyn S R, Civale L. Synergetic combination of different types of defect to optimize pinning landscape using BaZrO3-doped YBa2Cu3O7. 2009, Nature Materials, 8, 398-404. [25] Long N J. Model for the angular dependence of critical currents in technical superconductors. 2008, Supercond. Sci. Technol., 21, 025007. Pinning Stable Lattice Lattice Lattice Ref. ion Phase Parameter misfit misfit addition in YBa2Cu3O7−δ [nm] to abY BCO to cY BCO Nb+4 Ba2YNbO6 0.422 +9.4% a +8.4% [5,74] [123,124] Ta+4 Ba2YTaO6 0.422 +9.4% a +8.4% [68] Zr+4 BaZrO3 0.419 +8.8% a +7.6% [57–63] Sn+4 BaSnO3 0.411 +6.8% a +5.6% [64–66] Hf+4 BaHfO3 0.417 +8.3% a +7.1% [144] Ir+4 BaIrO3 0.410 +6.5% a +5.3% [145] Ti+4 BaTiO3 0.402 +4.4% a +3.2% [146,147] Y+3 Y2O3 1.060 -2.8% b -9.2% [78,142] [141,148] Ta+5 Gd3TaO7 1.065 -2.4% b -8.7% [67] Ta+5 Yb3TaO7 1.039 -4.8% b -10.9% [67] Table 1: Crystallographic parameters of non-superconductive second phase ma- terials used as pinning phases in YBa2Cu3O7−δ films (a cube-on-cube, b rotated 45◦ to cube-on-cube). 145 References [1] Tinkham M. Introduction to Superconductivity 2nd ed. McGraw-Hill, New York, 1996. xi, 6 [2] Cava R J et al. Structural anomalies, oxygen ordering and superconductiv- ity in oxygen deficient Ba2YCu3Ox. Physica C, 165:419–433, 1990. xi, 12, 13 [3] Singh R K et al. Pulsed laser deposition and charcterization on high-Tc Y Ba2Cu3O7−x superconducting thin films. Mat. Sci. and Eng. RR, 22:113– 185, 1998. xi, 23, 25 [4] Paulose K V et al. Ba2YNbO6: synthesis, properties and compatibility with YBa2Cu3O7−δ. Physica C, 193:273–276, 1992. xii, 41, 42, 43, 44, 46 [5] Ercolano G et al. Enhanced flux pinning in YBa2Cu3O7−δ thin films using Nb-based double perovskite additions. Supercond. Sci. Technol., 23:022003, 2010. xii, 43, 44, 46, 48, 49, 104, 109, 112, 121, 145 [6] Mele P et al. Enhanced high-field performance in PLD films fabricated by ablation of YSZ-added YBa2Cu3O7−x target. Supercond. Sci. Technol., 20, 2007. xvii, 89, 112, 113, 116 [7] Onnes H K. Investigation into the properties of sustances at low temper- atures, which have led, amongst other things to the preparation of liquid helium. Nobel lectures, pages 306–336, 1913. 1 [8] Onnes H K. Liquid helium. Comptes Rendus Hebdomadaires Des Seances de l’Academie Des Sciences, 147:421–424, 1908. 1 146 REFERENCES [9] Onnes H K. Further experiments with liquid helium D - On the change of the electrical resistance of pure metals at very low temperatures, etc V The disappearance of the resistance of mercury. Proceedings of the Koninklijke Akademie Van Wetenshappen te Amsterdam, 14:113–115, 1911. 1 [10] Ochsenfeld R Meissner W. Ein neuer Effekt bei Eintritt der Supraleitfhigkeit. Naturwissenschaften, 21:787–788, 1933. 2 [11] London H London F. The electromagnetic equations of the supraconductor. Proc. Roy. Soc., 149:71–88, 1935. 2 [12] Landau L D Ginzburg V L. To the Theory of Superconductivity. Zh. Eksp. Teor. Fiz., 20:1064, 1950. 4 [13] Abrikosov A A. On the Magnetic Properties of Superconductors of the Second Group. Soviet Phys. jetp-USSR, 5:1174–1183, 1957. 5 [14] Abrikosov A A. Type II Superconductors and the vortex lattice. Nobel Lecture, 2003. 5 [15] Kleiner W H et al. Bulk solution of ginzburg-landau equations for type II superconductors: Upper critical field region. Phys. Rev. A, 133:1226, 1964. 5 [16] Kramer L. Thermodynamical behaviour of type-II superconductors with small k near the lower critical field. Phys. Lett., 23:619, 1966. 5 [17] Schrieffer J R Bardeen J, Cooper L N. Theory of Superconductivity. Phys- ical Review, 108:1175–1204, 1957. 7 [18] Cooper L N. Bound Electron Pairs in a Degenerate Fermi Gas. Physical Review, 104:1189–1190, 1956. 7 [19] Mu¨ller K A Bednorz J G. Possible highTc superconductivity in the Ba-La- Cu-O system. Zeitschrift fr Physik B Condensed Matter, 64:189–193, 1986. 8 147 REFERENCES [20] Wu K M et al. Superconductivity at 93 K in a new mixed-phase Yb-Ba- Cu-O compound system at ambient pressure. Phys. Rev. Lett., 58:908–910, 1987. 8 [21] Maeda H et al. A New High-Tc Oxide Superconductor without a Rare Earth Element. Jpn. J. Appl. Phys, 27:209–210, 1988. 8 [22] Hermann A M Sheng Z Z. Bulk Superconductivity at 120 K in the Tl- Ca/Ba-Cu-O system. Nature, 332:138–139, 1988. 8 [23] Nagamatsu J et al. Superconductivity at 39 K in magnesium diboride. Nature, 410:63–64, 2001. 8 [24] Larbalestier D C et al. Strongly linked current flow in polycrystalline forms of the superconductor MgB2. Nature, 410:186–189, 2001. 8 [25] Buzea C and Yamashita T. Review of the superconducting properties of MgB2. Supercond. Sci. Technol., 14:R115–R146, 2001. 8 [26] Eom C B et al. High critical current density and enhanced irreversibility field in superconducting MgB2 thin films. Nature, 411:558–560, 2001. 8 [27] Jin S et al. High critical currents in iron-clad superconducting MgB2 wires. Nature, 411:553–565, 2001. 8 [28] Kamihara Y et al. Iron-Based Layered Superconductor: LaOFeP. J. Am. Chem. Soc, 128:10012–10013, 2006. 8 [29] Takahashi H et al. Superconductivity at 43 K in an iron-based layered compound LaO1−xFxFeAs. Nature, 453:376–378, 2008. 8 [30] Kamihara Y et al. Iron-based layered superconductor La[O1−xFx]FeAs (x=0.05-0.12) with Tc=26 K. J. Ame. Chem. Soc., 130:3296, 2008. 8 [31] Rotter M et al. Superconductivity at 38 K in the iron arsenide (Ba(1−x)K(x))Fe(2)As(2). Phys. Rev. Lett., 101:107006, 2008. 8 [32] Chen X H et al. Superconductivity at 43 K in SmFeAsO1−xFx. Nature, 453:761–762, 2008. 8 148 REFERENCES [33] Wu G et al. Superconductivity at 56 K in samarium-doped SrFeAsF. J. of Phys:Cond. mat., 21, 2009. 8 [34] Dinger T R et al. Direct Observation of Electronic Anisotropy in Single- Crystal YBa2Cu3O7−x. Phys. Rev. Lett., 58:2687–2690, 1987. 10 [35] Chaudhari P et al. Critical Currents Measurements in Epitaxial Films of YBa2Cu3O7−x Compound. Phys. Rev. Lett., 58:2684–2686, 1987. 10 [36] Larbalestier D C et al. Experiments concerning the connective nature of superconductivity in YBa2Cu3O7. J. Appl. Phys., 62:3308–3313, 1987. 10 [37] Dimos D et al. Orientation Dependence of Grain-Boundary Critical Cur- rents in YBa2Cu3O7−δ Bicrystals. Phys. Rev. Lett., 61:219–222, 1988. 10 [38] Mayer B et al. Superconducting transport properties of Bi2Sr2CaCu2O8+x bicrystal grain boundery junctions. Appl. Phys. Lett., 63:996–998, 1993. 10 [39] Heine K et al. High-field critical current densities inBi2Sr2Ca1Cu2O8+x/Ag wires. Appl. Phys. Lett., 55:2441–2443, 1989. 10 [40] Yamada Y et al. Critical Current Density of Wire Type Y-Ba-Cu Oxide Superconductor. Jpn. J. Appl. Phys., 26:865–866, 1987. 10 [41] Flu¨kiger R et al. Metallurgy and critical current in YBa2Cu3O7 wires. Physica C, 153:1574–1579, 1988. 10 [42] Wu X D et al. High-current YBa2Cu3O7−δ thick films on flexible nickel substrates with textured buffer layers. Appl. Phys. Lett., 65:1961–1963, 1994. 10 [43] Goyal A et al. Epitaxial superconductors on rolling-assisted biaxially- textured substrates (RABiTS): A route towards high critical current density wire. Appl. Superconductivity, 4:403–427, 1996. 10 [44] Larbalestier D et al. High-Tc superconducting materials for electric power applications. Nature, 414:368–377, 2001. 11 149 REFERENCES [45] Cava R J et al. Bulk Superconductivity at 91 K in Single-Phase Oxygen- Deficient Perovskite Ba2YCu3O9−δ. Phys. Rev. Lett., 58:1676–1679, 1987. 12 [46] Beno M A et al. Structure of the single-phase high-temperature supercon- ductor YBa2Cu3O7−δ. Appl. Phys. Lett., 51:57–59, 1987. 12 [47] David W I F et al. Structure and crystal chemistry of the high-Tc super- conductor YBa2Cu3O7−x. Nature, 327:310–312, 1987. 12 [48] Tietz L A and Carter C B. Special grain boundaries in YBa2Cu3O7−x thin films. Physica C Supercond., 182:241251, 1991. 12 [49] Linfeng Mei and Siu-Wai Chan. Enthalpy and entropy of twin boundaries in superconducting YBa2Cu3O7−x. J. Appl. Phys., 98:033908, 2005. 12 [50] Jorgenson J D et al. Structural and superconducting properties of or- thorhombic and tetragonal YBa2Cu3O7−x: The effect of oxygen stoichiom- etry and ordering on superconductivity. Phys. Rev. B, 36:5731–5734, 1987. 12 [51] Jorgenson J D et al. Oxygen ordering and the orthorhombic-to-tetragonal phase transition in YBa2Cu3O7−x. Phys. Rev. B, 36:36083616, 1987. 12 [52] Blatter G et al. Vortices in high-temperature superconductors. Rev. Mod. Phys., 66:1125–1338, 1994. 14 [53] Evetts J E. Flux Pinning by Phase Boundaries in Type-II Superconductors. Phys. Rev. B, 2:95–96, 1970. 16 [54] Coote R I et al. Flux Line Pinning by Large Normal Particles in Type-II Superconductors. Can. J. Phys., 50:421–427, 1972. 16 [55] Campbell A M and Evetts J E. Flux vortices and transport currents in type II superconductors. Adv. Phys., 21:199–428, 1972. 16 [56] Larkin A I and Ovchinnikov Y N. Pinning in type II superconductors. J. Low Temp. Phys., 34:409–428, 1979. 16 150 REFERENCES [57] MacManus-Driscoll J L et al. Strongly enhanched current densities in su- perconducting coated conductors of Y Ba2Cu3O7−x+BaZrO3. Nature Ma- terials, 3:439–443, 2004. 17, 39, 50, 59, 89, 112, 120, 145 [58] Foltyn S R et al. Materials science challenges for high-temperature super- conducting wire. Nature Materials, 6:631, 2007. 18, 145 [59] Galluzzi V et al. YBa2Cu3O7−δ films with BaZrO3 inclusions for strong- pinning in superconducting films on single crystal substrate. IEEE Trans. Appl. Supercond., 17:3628, 2007. 18, 145 [60] Peurla M et al. YBCO films prepared by PLD using nanocrystalline targets doped with BaZrO3 or Y211. IEEE Trans. Appl. Supercond., 15:3050–3053, 2005. 18, 145 [61] Ichinose A et al. Microstructures and critical current densities of YBCO films containing structure-controlled BaZrO3 nanorods. Supercond. Sci. Technol., 20:1144, 2007. 18, 145 [62] Ito M et al. Microstructure of ErBa2Cu3O7−δ films with BaZrO3 dispersion pinning centers for high Jc applications. Physica C, 426:1415, 2005. 18, 145 [63] Feldmann D M et al. 1000 Acm−1 in a 2 µm thick YBa2Cu3O7−x film with BaZrO3 and Y2O3 additions. Supercond. Sci. Technol., 23:115016, 2010. 18, 145 [64] Varanasi C V et al. Enhancement and angular dependence of transport critical current density in pulsed laser deposited Y Ba2Cu3O7−x +BaSnO3 films in applied magnetic fields. Jour. of Appl. Phys., 102, 2007. 18, 39, 50, 120, 145 [65] Mele P et al. Systematic study of the BaSnO3 insertion effect on the prop- erties of YBa2Cu3O7−x films prepared by pulsed laser ablation. Supercond. Sci. Technol., 21, 2008. 18, 39, 89, 145 [66] Mele P et al. Ultra-high flux pinning properties of BaMO3-doped YBa2Cu3O7−x thin films (M = Zr, Sn). Supercond. Sci. Technol., 21:032002, 2008. 18, 39, 145 151 REFERENCES [67] Harrington S A et al. Self-assembled, rare earth tantalate pyrochlore nanoparticles for superior flux pinning in Y Ba2Cu3O7−δ films. Supercond. Sci. Technol., 22, 2009. 18, 39, 59, 105, 108, 109, 120, 145 [68] Hun Wee S et al. Enhanced flux pinning and critical current density via incorporation of self-assembled rare-earth barium tantalate nanocolumns within Y Ba2Cu3O7−δ films. Phys. Rev. B, 81, 2010. 18, 39, 59, 108, 109, 112, 121, 145 [69] Gutierrez J et al. Strong isotropic flux pinning in solution-derived Y Ba2Cu3O7−x nanocomposite superconductor films. Nature Materials, 6:367–373, 2007. 18, 63 [70] Hougan T et al. Addition of nanoparticle dispersions to enhance flux pin- ning of the Y Ba2Cu3O7−x superconductor. Nature, 430:367–373, 2004. 18 [71] Campbell T A et al. Flux pinnig effect of Y2O3 nanoparticulate dispersion in multilayered YBCO thin films. Physica C: Supercond., 423:1–8, 2005. 18, 19 [72] Kang S et al. Strong enhancement of flux pinning in YBa2Cu3O7−δ multi- layers with columnar defects comprised of self-assembled BaZrO3 nanodots. Supercond. Sci. Technol., 20, 2007. 18 [73] Kang S et al. Flux pinning characteristic as a function of density of colum- nar defects comprised of self-assembled nanodots and nanorods in epitaxial YBa2Cu3O7−δ films for coated conductor applications. Physica C, 457:41– 46, 2007. 18 [74] Reich E et al. Structural and pinning properties of Y2Ba4CuMOy (M = Nb, Zr) / YBa2Cu3O7−δ quasi-multilayers fabricated by off-axis pulsed laser deposition. Supercond. Sci. Technol., 22, 2009. 19, 43, 89, 145 [75] Matsumoto K et al. Enhancement of critical current density of YBCO films by introduction of artificial pinning centers due to the distributed nano- scaled Y2O3 islands on substrates. Physica C Supercond., 412:1267–1271, 2004. 19 152 REFERENCES [76] Matsumoto K et al. Effects of artificial pinning centers on vortex pinning in high-temperature superconducting films. Physica C Supercond., 426- 431:1091–1095, 2005. 19 [77] Mele P et al. Critical current enhancement in PLD YBa2Cu3O7−x films using artificial pinning centers. Physica C Supercond., 445-448:648–651, 2006. 19 [78] Mele P et al. Tuning of the critical current in YBa2Cu3O7−x thin films by controlling the size and density of Y2O3 nanoislands on annealed SrTiO3 substrates. Supercond. Sci. Technol., 19:44, 2006. 19, 145 [79] Li A H et al. Effect of substrate surface modification using Ag nano-dots on the improvement of Jc and microstructures in YBa2Cu3O7 thin films grown on LaAlO3 (100) by pulsed laser deposition. Journ. Electroceram., 16:605–609, 2006. 19 [80] Dang V S et al. Combination of Ag Substrate Decoration with Introduc- tion of BaZrO3 Nano-Inclusions for Enhancing Critical Current Density of YBa2Cu3O7 Films. Journ. Electroceram., 24:505–509, 2011. 19 [81] Aytug T et al. Enhancement of flux pinning and critical currents in YBa2Cu3O7−δ films by nanoscale iridium pretreatment of substrate sur- faces. Jour. of Appl. Phys., 98, 2005. 19 [82] Aytug T et al. Analysis of flux pinning in YBa2Cu3O7−δ films by nanoparticle-modified substrate surfaces. Phys. Rev. B, 74:184505, 2006. 19 [83] Aytug T et al. Substrate Surface Decoration With CeO2 Nanoparticles: An Effective Method for Improving Flux Pinning in YBa2Cu3O7−δ Films. IEEE Trans. Appl. Supercond., 17:3720–3723, 2007. 19 [84] Crisan A et al. Sputtered nanodots: A costless method for inducing effective pinning centers in superconducting thin films. Appl. Phys. Lett., 79:4547– 4549, 2001. 19 153 REFERENCES [85] Sparing M et al. Artificial pinning centres in YBCO thin films induced by substrate decoration with gas-phase-prepared Y2O3 nanoparticles. Super- cond. Sci. Technol., 20:S239–S246, 2007. 19 [86] Cui X M et al. Enhancement of critical current density of YBa2Cu3O7−δ thin films by nanoscale CeO2 pretreatment of substrate surfaces. Physica C Supercond., 466:1–4, 2007. 19 [87] MacManus-Driscoll J L et al. Guidelines for optimizing random and cor- related pinning in rare-earth based superconducting films. Supercond. Sci. Technol., 19:S55–S59, 2006. 19 [88] Civale L et al. Vortex confinement by columnar defects in YBa2Cu3O7 crystals: Enhanced pinning at high fields and temperatures. Phys. Rev. Lett., 67, 1991. 19 [89] Weber H W et al. Critical currents in neutron-irradiated YBCO and BIS- CCO single-crystals. Supercond. Sci. Technol., 4:S103–S105, 1991. 19 [90] Jirong J et al. Irradiation effect of slow neutrons on melt-textured YBa2Cu3O7−δ. Phys. Stat. Sol., 125:289–294, 1991. 19 [91] Wisniewski A et al. Comparison of neutron-irradiation effects in the 90-K and 60-K phases of YBCO ceramics. Physica C, 197:365–370, 1992. 19 [92] Kato T et al. Radiation effect of YBa2Cu3O7−y irradiated by γ-Rays and 14 MeV neutrons. Jpn. J. Appl. Phys., 27:L2097–L2099, 1988. 19 [93] Vasek P et al. Gamma-irradiation of YBa2Cu3O7−x ceramics. Solid State Com., 69:23–25, 1989. 19 [94] Kim D H et al. Effect of Heavy-Ion Irradiation on Transport Properties of Y Ba2Cu3Ox Films. Trans. on appl. Supercond., 7, 1997. 20 [95] Nakashima K et al. Effect of ion-irradiation and annealing on supercon- ductive property of PLD prepared YBCO tapes. Physica C Supercond., 463:665–668, 2007. 20 154 REFERENCES [96] Eisterer M et al. Neutron irradiation of coated conductors. Supercond. Sci. Technol., 23:014009, 2010. 20 [97] Chudy M et al. Asymmetric angular dependence of J(c) in coated con- ductors prior to and after fast neutron irradiation. Physica C Supercond., 470:1300–1303, 2010. 20 [98] Chudy M et al. Characterization of Commercial YBCO Coated Conductors After Neutron Irradiation. IEEE Trans. Appl. Supercond., 21:3162–3165, 2011. 20 [99] Fuger R et al. Influence of neutron irradiation on high temperature super- conducting coated conductors. Physica C Supercond., 468:1647–1651, 2008. 20 [100] Yamada H et al. Flux pinning centres correlated along the c-axis in PLD- YBCO films. Supercond. Sci. Technol., 17:58, 2004. 22 [101] Civale L et al. Influence of crystalline texture on vortex pinning near the ab-plane in YBa2Cu3O7 thin films and coated conductors. Physica C Su- percond., 412:976–982, 2004. 22 [102] Takahashi K et al. Investigation of thick PLD-GdBCO and ZrO2 doped GdBCO coated conductors with high critical current on PLD-CeO2 capped IBAD-GZO substrate tapes. Supercond. Sci. Technol., 19:924–929, 2006. 22 [103] Traito K et al. Magnetic field dependence of the critical current and the flux pinning mechanism in YBa2Cu3O6+x films doped with BaZrO3. Phys. Rev. B, 73:224522, 2006. 22 [104] Horide T et al. Flux pinning properties of YBCO thin films deposited on SrTiO3(100) and MgO(100) substrates. Physica C Supercond., 412:1291– 1295, 2004. 22 [105] Douglas B C and Graham K H. Pulsed Laser Deposition of Thin Films. Jhon Wiley & Sons, New York, 1994. 22, 51 155 REFERENCES [106] Kwok H S et al. Laser induced deposition of YBa2Cu3O7−x thin films. In High-Temperature Superconductors, pages 273–277, 1988. 22 [107] Horcas I et al. WSXM: A software for scanning probe microscopy and a tool for nanotechnology. Rev. Sci. Instrum., 78:013705, 2007. 35 [108] Kuwabara M et al. Microstructure and Superconducting Properties in YBa2Cu3O7−x ceramics Doped with Nb2O5 and WO3. Jap. Jour. of Appl. Phys., 27:L1504–L1506, 1988. 40 [109] Greaves C and Slater PR. Nb and Ta Substitutions in YBa2Cu3O7−x and related phases − Structural Characterization of La1.1Ba1.9Cu2.1M0.9O8 (M=Nb, Ta). Physica C, 161:245–251, 1989. 40, 108, 120 [110] Strukova G K et al. Effect of Nb doping on properties of Y-Ba-Cu ceramics. Supercond. Sci. Technol., 6:589–592, 1993. 40 [111] Strukova G K et al. A cubic phase in ceramic Y-Ba-Cu-O doped with Nb. Physica C, 267:67–73, 1996. 40 [112] Babu N H et al. New chemically stable, nano-size artificial flux pinning cen- tres in (RE)-Ba-Cu-O superconductors. Supercond. Sci. Technol., 16:L44– L45, 2003. 41 [113] Babu N H et al. Artificial flux pinning centers in large, single-grain (RE)- Ba-Cu-O superconductors. Appl. Phys. Lett., 83:4806–4808, 2003. 41 [114] Yeoh W K et al. Improved Flux Pinning in Y-Ba-Cu-O Superconductors Containing Niobium Oxide. IEEE Trans. Appl. Supercond., 19:2970–2973, 2009. 41 [115] Satiraju S et al. Studies on Ba2YNbO6 buffer layers for subsequent YBa2Cu3O7−x film growth. IEEE Trans. Appl. Supercond., 15:3009–3012, 2005. 41 [116] Grekhov I et al. A new buffer layer for high-quality HTSC ultrathin film fabrication. Physica C, 276:18–24, 1997. 41 156 REFERENCES [117] Jia J H et al. Effect of the Second Phase on Critical Current Density in Nb-Added YBCO Films. Mod. Phys. Lett., 9:439–443, 1995. 41 [118] Hori S et al. Introduction of c-axis-correlated 1D pinning centers and vortex Bose glass in BaNbO-doped ErBa2Cu3Oy films. Supercond. Sci. Technol., 20:1115–1119, 2007. 41, 59, 89, 104, 112, 120 [119] Kai H et al. Superconducting properties and microstructure of PLD- ErBa2Cu3O7−δ film with BaNb2O6. Physica C: Supercond., 463-465:895– 899, 2007. 41, 59, 112, 120 [120] Kai H et al. Influence of Growth Temperature on Microstructure and Su- perconducting Properties of ErB2Cu3O7−δ Films With Ba(Er0.5Nb0.5)O3 Nanorods. IEEE Trans. Appl. Supercond., 19:3435–3438, 2009. 41 [121] Kai H et al. The effects of growth temperature on c-axis-correlated pinning centers in PLD-ErB2Cu3O7−δ films with Ba(Er0.5Nb0.5)O3. Supercond. Sci. Technol., 23:025017, 2010. 41 [122] Yamada K et al. Transmission electron microscopy characterization of nanorods in BaNb2O6-doped ErB2Cu3O7−δ films. Appl. Phys. Lett., 92:112503, 2008. 41 [123] Feldmann D M et al. Improved flux pinning in YBa2Cu3O7 with nanorods of the double perovskite Ba2YNbO6. Supercond. Sci. Technol., 23:095004, 2010. 43, 145 [124] Hun Wee S et al. Formation of Self-Assembled, Double-Perovskite, Ba2YNbO6 Nanocolumns and Their Contribution to Flux-Pinning and Jc in Nb-Doped YBa2Cu3O7−δ Films. Appl. Phys. Express, 3:023101, 2010. 43, 104, 109, 145 [125] Barnes P W et al. Structure determination of A2M 3+TaO6 and A2M 3+NbO6 ordered perovskites: octahedral tilting and pseudosymmetry. Acta Crystal- logr., B62:384–396, 2006. 45 157 REFERENCES [126] Goyal A et al. Irradiation-free, columnar defects comprised of self-assembled nanodots and nanorods resulting in strongly enhanced flux-pinning in YBa2Cu3O7−δ films. Supercond. Sci. Technol., 18:1533–1538, 2005. 50 [127] MacManus-Driscoll et al. Systematic enhancement of in-field critical cur- rent density with rare-earth ion size variance in superconducting rare-earth barium cuprate films. Appl. Phys. Lett., 84:5329–5331, 2004. 50 [128] Singh R K et al. Control of surface particle density in pulsed laser deposition of superconducting YBa2Cu307 and diamondlike carbon thin films. Appl. Phys. Lett., 61:483–485, 1992. 51 [129] Wu W B et al. Particulate-free a-YBa2Cu307−x/La0.5Sr0.5CoO3/a- YBa2Cu307−x sandwiched thin film on (100)SrTiO3 fabricated by pulsed laser deposition with a shadow mask. Physica C, 282:701–702, 1997. 51 [130] Boffa V et al. High-quality surface YBCO thin films prepared by off-axis pulsed laser deposition technique. Physica C, 276:218–224, 1997. 51 [131] Backen E et al. Improved pinning in YBCO based quasi-multilayers pre- pared by on- and off-axis pulsed laser deposition. IEEE Trans. Appl. Su- percond., 17:3733–3736, 2007. 51 [132] Cillessen J F M et al. Improved uniformity of multielement thin films prepared by off-axis pulsed laser deposition using a new heater design. Rev. Sci. Instr., 67:3229–3237, 1996. 51 [133] Harrington S A et al. Understandig nanoparticle self-assembly for a strong impovement in functionality in thin film nanocomposites. Nanotechnology, 21, 2010. 56, 112 [134] Maiorov B et al. Synergetic combination of different types of defect to optimize pinning landscape using BaZrO3-doped YBa2Cu3O7−δ. Nature Materials, 8, 2009. 56, 63, 89, 90, 114 [135] Liu L F et al. Effect of Deposition Temperature on the Epitaxial Growth of YBCO Thin Films on RABiTS Substrates by Pulsed Laser Deposition Method. IEEE trans. on appl. supercond., 20:1553–1553, 2010. 96 158 REFERENCES [136] Norton M G et al. Surface outgrowths on laser-deposited YBa2Cu3O7 thin films. Physica C, 233:321–326, 1994. 96 [137] Guo L P et al. Microstructure of outgrowths on the surface of laser-ablated YBa2Cu3O7 thin films. Physica C, 241:30–36, 1995. 96 [138] Kai H et al. The effects of growth temperature on c-axis-correlated pinning centers in PLD-ErBa2Cu3O7−δ films with Ba(Er0.5Nb0.5)O3. Supercond. Sci. Technol., 23, 2010. 104 [139] Holesinger T G et al. The microstructure of continuosly processed YBa2Cu3Oy coated conductors with undelying CeO2 and ion’beam’assisted yttria’stabilized zirconia buffer layers. Jour. of Mat. Res., 15:1110–1119, 2000. 108 [140] Specht E D et al. Stacking faults in YBa2Cu3O7−x: Measurement using x- ray diffraction and effects on critical current. Appl. Phys. Lett., 89:162510, 2006. 108 [141] Gapud A A et al. Enhancement of flux pinning in YBa2Cu3O7−δ thin films embedded with epitaxially grown Y2O3 nanostructures using a multi- layering process. Supercond. Sci. Technol., 18:1502–1505, 2005. 108, 145 [142] Garrison S M et al. Observation of 2 in-plane epitaxial states in YBa2Cu3O7−δ films on yttria-stabilized ZrO2. Appl. Phys. Lett., 58:2168– 2170, 1991. 109, 145 [143] Long N J. Model for the angular dependence of critical currents in technical superconductors. Supercond. Sci. Technol., 21, 2008. 117 [144] Zhang J L and Evetts J E. BaZrO3 and BaHfO3 - preparation, properties and compatibility with YB2Cu3O7−x. J. Mater. Sci., 29:778–781, 1994. 145 [145] Hanish J et al. Formation of BaIrO3 precipitates and their contribution to flux pinning in Ir-doped YB2Cu3O7−x quasi-multilayers. Appl. Phys. Lett., 86:122508, 2005. 145 159 REFERENCES [146] Alok K J and Khare N. Strongly enhanced pinning force density in YBCOBaTiO3 nanocomposite superconductor. Physica C Supercond., 469:810–813, 2009. 145 [147] Shingai Y et al. Improvement of superconducting properties by BaTiO3 doping into ErB2Cu3O7−δ films. Physica C Supercond., 445-448:841–844, 2006. 145 [148] Mele P et al. Incorporation of double artificial pinning centers in YB2Cu3O7−δ films. Physica C Supercond., 468:1631–1634, 2008. 145 160