magnesium diboride thin films and devices

163
The Pennsylvania State University The Graduate School Department of Physics MAGNESIUM DIBORIDE THIN FILMS AND DEVICES A Thesis in Physics by Yi Cui © 2007 Yi Cui Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy December 2007

Upload: others

Post on 21-Nov-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

The Pennsylvania State University

The Graduate School

Department of Physics

MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

A Thesis in

Physics

by

Yi Cui

© 2007 Yi Cui

Submitted in Partial Fulfillment of the Requirements

for the Degree of

Doctor of Philosophy

December 2007

Page 2: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

ii

The thesis of Yi Cui was reviewed and approved* by the following:

Xiaoxing Xi Professor of Physics and Materials Science and Engineering Thesis Advisor Chair of Committee

Julian D. Maynard Distinguished Professor of Physics

Peter E. Schiffer Professor of Physics

Suzanne E. Mohney Professor of Materials Science and Engineering

Jayanth R. Banavar Distinguished Professor of Physics Head of the Department of Physics

*Signatures are on file in the Graduate School

Page 3: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

iii

ABSTRACT

Magnesium diboride (MgB2) is a binary compound superconductor with a

superconducting transition temperature Tc of ~ 40 K. MgB2 has two conduction bands: a

two-dimensional σ band and a three-dimensional π band with weak interband scattering.

The two gap superconductivity in MgB2 gives rise to many interesting physical properties

not possible in other superconductors. The relatively high Tc combined with phonon

mediated superconductivity and relatively long coherence length makes MgB2 promising

for electronics applications like rapid single flux quantum (RSFQ) logics and

superconducting quantum interference devices (SQUID). The high current density and

record-high upper critical field in pure or alloyed MgB2 are also attractive to a variety of

high field applications including cryogen-free Magnetic Resonance Imaging (MRI)

systems. MgB2 may also be used in RF cavity coatings due to its low surface resistance

and in photo detection due to its fast photoresponse coupled with relatively high Tc.

High quality epitaxial thin films are produced by the hybrid physical-chemical

vapor deposition (HPCVD) technique. The HPCVD MgB2 thin films have the highest Tc

and lowest resistivity with sharp transition of all MgB2 materials reported. The HPCVD

MgB2 material is free of dendritic flux jumps due to its low resistivity. The root-mean-

square (RMS) surface roughness of HPCVD MgB2 films can be ~1 nm when ~1% of

nitrogen is added to the hydrogen carrier gas during the growth. The stability of MgB2

films in water is studied; it is found that degradation can be prevented by a thin (10 nm)

MgO layer deposited on the film surface. The Tc is enhanced by tensile strain due to the

Page 4: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

iv

Volmer-Weber growth mode and the mismatches between MgB2 and the substrate in the

lattice constants and the coefficients of thermal expansion.

High quality superconductor-insulator-superconductor Josephson tunnel junctions

were made with various barrier formation techniques. The junction critical current

densities and IcRn products are high with clear gap characteristics and low subgap

currents. The Fraunhofer-pattern of Josephson supercurrent modulation in magnetic fields

demonstrates excellent junction uniformity. The barrier thickness and height were

estimated, and the barrier composition was studied by X-ray Photoelectron Spectroscopy

(XPS). Josephson tunnel junctions with non-c-axis-oriented MgB2 were made which

clearly exhibit tunneling spectra from both MgB2 superconducting gaps. The two-band

superconductivity and its effect on vortices were studied by tunneling spectroscopy in

magnetic fields.

Planar all-MgB2 Josephson junctions were made by creating a weak-link through

a TiB2 underlayer or an ion damaged MgB2. Junctions exhibited Josephson critical

current and RSJ-like characteristics and Shapiro steps under microwave radiation.

Uniform ion damage MgB2 Josephson junction array was also demonstrated.

Page 5: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

v

TABLE OF CONTENTS

LIST OF FIGURES ..................................................................................................... vii

ACKNOWLEDGEMENTS ......................................................................................... xiv

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

1.1 Overview ................................................................................................................ 1

1.2 Superconductivity .................................................................................................. 2

1.3 Magnesium Diboride ............................................................................................. 9

1.4 Josephson Junctions and Their Applications ......................................................... 15

Chapter 2 Experimental Methods ............................................................................... 20

2.1 Hybrid Physical-Chemical Vapor Deposition ....................................................... 20

2.2 Sputter Deposition of TiB2, MgO, Al, AlN, Au, and Cr........................................ 26

2.3 Evaporation of SiOx, Pb, and Ag ........................................................................... 29

2.4 X-ray Diffraction (XRD) ....................................................................................... 30

2.5 AFM and SEM ....................................................................................................... 30

2.6 Electrical and Magnetic Characterizations ............................................................ 31

2.7 Photolithograpic Device Processing ...................................................................... 34

2.8 Device Measurements ............................................................................................ 38

Chapter 3 MgB2 Thin Films and Heterostructures ..................................................... 41

3.1 Structures of MgB2 Films Grown by HPCVD ....................................................... 41

3.2 Transport and AC susceptibility properties of MgB2 Films .................................. 44

3.3 Electron Scattering Dependence of Dendritic Magnetic Instability in MgB2 Films ..................................................................................................................... 46

3.4 Photoresponse of MgB2 Thin Bridges ................................................................... 51

3.6 Substrate, Growth Mode and Thermal Expansion Issues ...................................... 56

Page 6: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

vi

3.7 MgB2 Film Morphology and Improvement for Tunnel Junctions ......................... 61

3.8 Material Stability Study of MgB2 Films and Protection ........................................ 66

3.9 TiB2/MgB2 Hetero-structures ................................................................................. 71

Chapter 4 Superconductor-Insulator-Superconductor MgB2 Josephson Junctions .... 74

4.1 Quasiparticle Tunneling and Cooper Pair Tunneling through Insulator ................ 75

4.2 Fabrication of MgB2 SIS Josephson Junctions ...................................................... 81

4.3 Electrical Properties of MgB2 SIS Josephson Junctions ........................................ 84

4.4 XPS Study of Barrier Properties ............................................................................ 90

4.5 Barrier Height and Thickness Estimation by Transport Measurements ................ 95

4.6 Fraunhofer Pattern in SIS MgB2 Josephson Junctions and Penetration Depth in MgB2 Films ....................................................................................................... 97

4.7 Spectroscopy Study of Two Bands of MgB2 ......................................................... 101

4.8 Tunneling Study of Mixed State in MgB2 ............................................................. 107

Chapter 5 Planar MgB2 Josephson Junctions and Circuits ......................................... 119

5.1 Planar MgB2-TiB2-MgB2 SNS Josephson Junctions .............................................. 119

5.2 Ion Damage MgB2 Josephson Junctions and Series Array .................................... 125

Chapter 6 Conclusions and Future Plan ...................................................................... 128

REFERENCES ............................................................................................................ 131

Page 7: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

vii

LIST OF FIGURES

Figure 1. The timeline of superconductivity discovery in different materials. [12]…3

Figure 2. Illustration of the Meissner effect. Magnetic field lines are excluded from a superconductor when it is below its critical temperature. [12]……………………….……………………3

Figure 3. The density of state as a function of energy around the Fermi surface in a superconductor. ………….……………………………….……………………………….……………………………….………7

Figure 4. (a) Hexagonal structure of MgB2 consisting of honeycomb B (blue) layers with close-packed Mg (yellow) layers between them. The bulk lattice constants are determined by XRD to be a = 3.086Å and c = 3.524 Å. (b) The 2-D covalent σ bonds (brown) are formed within B sheets by overlapping of sp2 hybrid B orbitals and the 3-D metallic π bonds (green) are formed by the p orbital electrons perpendicular to the layers. [39] …….9

Figure 5. The B-B vibrational mode of the E2g phonon strongly couples to σ-bonding states. As B atoms move in the arrow directions, elongated bonds (marked with ‘R’) become repulsive to electrons, whereas shortened bonds (marked with ‘A’) become attractive. The σ-bonding states strongly couple to the E2g phonon mode because they are mainly located in either the attractive or the repulsive bondings of the mode. The π-bonding states do not couple strongly to this mode. [9] …….……………………………….…………11

Figure 6. Electronic band structure (left) and Fermi surface (right) of MgB2. [9,37]……12

Figure 7. (a) The calculated superconducting energy gap of MgB2 on the Fermi surface at 4 K. (b) the distribution of gap values at 4 K. (c) local distribution of the energy gap is plotted on planes at 0.05, 0.10, and 0.18 nm above a B plane, respectively. [9] …….………13

Figure 8. Calculated temperature dependence of (a) the superconducting gaps and (b) the quasiparticle density of states. [9] …….……………………….………………………….……………………………13

Figure 9. The current – voltage characteristic of a superconducting tunnel junction originally predicted by Josephson. [49] …….……………………….………………………….…………………16

Figure 10. An ideal Fraunhofer pattern: supercurrent in a Josephson junction can be modulated by an external magnetic field perpendicular to the current flow direction. …18

Figure 11 The pressure-temperature phase diagram calculated for the Mg:B atomic ratio xMg:xB ≥ 1/2. The region marked by “Gas + MgB2” represents the thermodynamic stable window for MgB2 film growth. [78] …….……………………….………………………….……………………21

Page 8: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

viii

Figure 12 (a) A Schematic of the HPCVD setup. (b) Computer simulated gas velocity profile in the reactor near the susceptor. [77] …….……………………….………………………….………23

Figure 13. A photo of the original HPCVD MgB2 film deposition system. …….……………24

Figure 14. A photograph of the integrated in situ HPCVD system consisting of 2 computer controlled CVD modules and a sputtering module with an interlocked transfer station. …….……………………….………………………….………………….………………………….…………………………25

Figure 15 Milling rate of MgB2 with a beam voltage and current of of 270 V and 5.5 mA, respectively. Red line is a linear fit. .……………………………….………………………….………………….……37

Figure 16 A high-resolution transmission electron microscope (HRTEM) image for the interface between a MgB2 film and (0001) 6H-SiC substrate. The insets are selected area electron diffraction (SAED) patterns from the film (top) and substrate (bottom). [77] …42

Figure 17. X-ray diffraction of θ-2θ scan and φ scan of HPCVD grown MgB2films. From [87] ………………………….………………………….………………….……………………………….…………………………43

Figure 18. Resistivity vs temperature curve for a 300 nm MgB2 thin film on SiC substrate. The inset shows details near the superconducting transition. ………………………….………………44

Figure 19. Real (red) and imaginary (blue) component of the AC magnetic susceptibility measurement of a MgB2 film as a function of temperature. ………………………….………………45

Figure 20 . Magneto-optical images of the zero-field-cooled pure and C-alloyed MgB2 thin film (5x5 mm ) at T = 4:2 K. The perpendicular applied field B = (a) 10 mT, (b) 20 mT, (c) 40 mT, and (d) 0 (reduced from 0.1 T), respectively. [92] ………………………………48

Figure 21 . Magnetization curves of the ultra-pure MgB2 thin film and the carbon-doped MgB thin film. (a) The hysteresis loop of the pure MgB2 thin film at T = 5, 10, and 15 K, respectively. (b) The hysteresis loops of the carbon-alloyed MgB2 thin film at the same temperatures. [92] ………………………….………………………….………………….…….…………………………49

Figure 22. (a) Experimental waveform (circles) at low optical excitation with an excitation power of 400 μW at 20 K. The solid line is a fitting with kinetic inductive response model. (b) Experimental waveform (circles) with an excitation power of 4 mW at 20 K. The data is fitted with a kinetic inductive response (dotted line) and a resistive response (dashed line). The combined kinetic-inductive fit is shown as a solid line. [100]

………………………….………………………….………………….……………………………….…………………………53

Figure 23. The dependence of the photoresponse-signal rise time (a) and amplitude (b) on absorbed optical power of a photoexcited microbridge biased at Ib = 62 mA and at T0 =

Page 9: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

ix

20 K. Two regimes can be identified from the plots with the transition point at the excitation power P ≈ 500 μW. [100] ………………………….………………………….………………55

Figure 24. 1 μm x 1 μm AFM images of an intrinsic 6H SiC substrate before and after heat treatment in H2 at 720 ºC. ………………………….…………………………………….…………………………57

Figure 25. When crystallites coalesce, they spontaneously snap together and generate a tensile strain. [77] ………………………….………………………….………………….……………………………58

Figure 26. SEM images of MgB2 films. (a) 900 nm on SiC deposited at 720 ºC with 2% N2 added. (b) 1.3 μm on Al2O3 deposited at 720 ºC. (c) 700 nm on SiC deposited at 620 ºC with 1.5% N2 added. (d) 500 nm film on SiC deposited at 720 ºC.………….…………………60

Figure 27. (a)-(f): 1 μm x 1 μm AFM scans of ~ 100 nm MgB2 films without and with 5, 10, 20, 30, 50, and 100 sccm N2 flow added during the deposition. The total flow and pressure was maintained at 700 sccm and 80 Torr, respectively. …………………………………63

Figure 28. RMS roughness and transport properties of MgB2 films grown with 5, 10, 20, 30 50, 100 sccm flow of N2 added to the carrier gas. ………………………….……………………………64

Figure 29. In magnetic field transport properties of a patterned MgB2 with 50 sccm N2 added during deposition. The patterned bridges are 15 μm or 30 μm in width and 690 μm in length. Fields were applied parallel and perpendicular to the ab plane of MgB2. ………65

Figure 30 (a) Resistance vs. time curve of 2000Å-thick MgB2 films submerged in water at room temperature and 0 °C. (b) Thickness change of a MgB2 film as a function of time in water at room temperature. (c) Resistance vs. temperature curves of a MgB2 film after consecutive exposures to water at room temperature. (d) Time dependence of Tc(0) and normalized resistance summarized from the results in (c) [120] ……………………………………67

Figure 31. Room temperature resistance of 1000Å-thick MgB2 films as a function of time submerged in water, methanol, acetone, and isopropanol. [120] …………….……………………69

Figure 32. The resistance versus temperature curves of a MgB2 film with a sputtered 10 nm MgO protection layer before and after 45 hours exposure to saturated water vapor at 23˚C. ………………………….………………………….………………….……………………………….…………………………69

Figure 33. (a) Bright-field TEM image of a TiB2/MgB2 heterostructure. (b) SAED pattern collected from an area containing all the layers. (c) HRTEM image of the TiB2/SiC substrate interface. (d) HRTEM image of the TiB2/MgB2 interface. The arrows indicate the interface. [158] ………………………….………………………….………………….………………….…72

Page 10: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

x

Figure 34. The resistivity versus temperature curve of a TiB2 film and the resistance of a planar MgB2-TiB2-MgB2 junction as a function of temperature. [158] …………………….73

Figure 35. Illustration of particle tunneling through barriers. (a) Macroscopic object like soccer ball can not tunneling through barriers like wall. (b) and (c) Microscopic particles like electrons in metal can only if the distance between the two metals is brought close enough to make work function low enough to provide practical tunneling probability. (d) A bias voltage V is needed to provide empty states for electrons to tunnel through and occupy. ………………………….………………………….………………….……………………………….………………………76

Figure 36. Illustration of quasiparticle tunneling between two superconductors with different energy gaps. (a) density of states of two superconductors with a bias of (Δ1+Δ2)/e. (b) density of states of two superconductors with a bias of (Δ1+Δ2)/e. (c) a schematic of the ideal current-voltage characteristics between two superconductors with different energy gaps. ………………………….………………………….………………….………………………………80

Figure 37. Fabrication process for sandwich type SIS Josephson tunnel junctions. ………82

Figure 38. Fabrication process for sandwich-type SIS Josephson tunnel junctions starting from a trilayer. ………………………….………………………….………………….……………………………….………83

Figure 39. (a) I–V curve of a MgB2/insulator /Pb junction made from a c-axis oriented MgB2 thin film taken at 4.4 K. Barrier is formed by exposing MgB2 to N2 at ~400 °C (Process B). ………………………….………………………….………………….……………………………….…85

Figure 40. (a) I–V characteristics of a MgB2/insulator/Pb junction measured at 4.3 K with high temperature barriers by process A (annealing MgB2 film at 710 °C in H2 for 20 seconds.) (b) Differential conductance, dI/dV, as a function of voltage, V, of the same junction in (a). (c) I–V characteristics of a MgB2/insulator/Pb junction measured at 4.3 K with barriers by process A (annealing MgB2 film at 710 °C in H2 for 15 seconds). (d) Differential conductance, dI/dV, as a function of voltage, V, of the same junction in (c). An additional peak appeared at 1.3 mV, indicating normal metal regions in MgB2 barrier interface. [133] ………………………….………………………….………………….……………………………….……88

Figure 41. (a) I–V characteristics for a MgB2/insulator/Pb junction measured at 4.3 K with a barrier by Process C. (b) Differential conductance, dI/dV, as a function of voltage, V, of the same junction in (a). ………………………….………………………….………………….…………………89

Figure 42. Surface and near surface XPS scans for (a) a HPCVD MgB2 film as an control sample (b) a MgB2 film with Process A barrier. (c) a MgB2 film with Process B barrier. …………………………………….………………………….………………….……………………………….…………………………91

Page 11: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

xi

Figure 43. Composition fittings of Surface and near surface XPS scans for (a) a HPCVD MgB2 film as an control sample (b) a MgB2 film with Process A barrier. (c) a MgB2 film with Process B barrier. ………………………….………………………….………………….……………………………….93

Figure 44. I-V curve of a MgB2/I/Pb junction measured at 42 K (left) to high voltages and the differential conductance (right) of the same junction with a parabolic fit (purple). The barrier for this junction is estimated to have a thickness of ~1.8 nm and a barrier height of 0.7 eV. ……………………………………….………………………….…………………….…………96

Figure 45. Magnetic field dependence of Josephson critical current, Ic. The filled circles are experimental data and the solid line is the calculated ideal Fraunhofer pattern. [134]

……….…………………………….……………………………………………….………………………….………………….…………99

Figure 46. (a) Normal metal – MgB2 tunneling in the ab plane direction for several interface transparencies, ranging from Z = 0 (Andreev contacts) to Z >>1 (tunnel juctions). The barrier parameter Z is determined by the barrier potential φ and the Fermi velocity vF by Z = φ/ħvF. (b) Normal metal – MgB2 tunneling in the c axis direction for several interface transparencies, ranging from Z = 0 (Andreev contacts) to Z >>1 (tunnel junctions). [142] ……….…………………………….……………………………………………….………………………102

Figure 47. (a) I–V characteristics for a MgB2/insulator/Pb junction measured at 4.3 K with barrier B formed by venting the reactor with nitrogen at 350 ºC and taking the sample out at 280 ºC. (b) Differential conductance, dI/dV, as a function of voltage, V, of the same junction in (a). [133] ……….…………………………….……………………………………….…………103

Figure 48. Left: a schematic of crystal orientation relation ship between MgO (211) substrate and MgB2 films. MgB2 is grown with c axis tilted by 19.5º, exposing the a-b plane. Right: a SEM picture of a MgB2 film grown on MgO (211) substrate. ….…………104

Figure 49. Temperature dependence of dI/dV versus V for a MgB2/insulator/Pb junction on (211) MgO substrate. The results for temperatures higher than 4.4 K are vertically shifted and multiplied by 5 for clarity. [134] ……….…………………………….……………………………105

Figure 50. Temperature dependence of the two gaps of MgB2 from a MgB2/insulator/Pb junction on (211) MgO substrate. ……….…………………………….………………………………106

Figure 51. Vortices in single crystal MgB2 with Hc at 2 K. 250 x 250 nm2 spectroscopic images of a single vortex induced by an applied field of 0.05 T (a), and the vortex lattice at 0.2 T (b). (c) Normalized zero bias conductance versus distance from the center, for the isolated vortex shown in (a). (d) Vortices in single crystal MgB2 with H

ab at 2 K. The bars indicate zero bias conductivity (ZBC). [151,155] ……….…………………109

Page 12: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

xii

Figure 52. Superconducting gaps (a) and density of states (b) of MgB2 as a function of distance from the center of vortex in magnetic field with D1=0.2D2. Maximum gap values (c) and averaged density of states (d) of MgB2 as a function of applied magnetic field. [156] ……….…………………………….………….…………………….…………………………………….…………111

Figure 53. dI/dV curves of a MgB2/insulator/Ag tunnel junctions on SiC at 4.2 K with magnetic fields of 0, 0.04, 0.16, 0.3, 0.5, 0.7, 1, 1.5, 2, 2.5, 3, 4, 5 T applied along the c axis of MgB2. ……….…………………………….………….…………………….…………………………………113

Figure 54. (a) Simulated ZBC profile around a vortex with ξπ= 40, 50 and 60 nm in a magnetic field of H=0.2 T (b) Calculated magnetic field dependence of bulk ZBC with ξπ= 30, 40 and 50 nm. (c) Simulated ZBC profile around a vortex with ξπ= 30 nm in a magnetic field of H=0.05 and 0.1 T (d) Calculated error for substituting the integrated conductance over the vortex area with a field of H with the integrated conductance with the same area in area the vortex with a field of 1.5H, for with ξπ= 20, 35 and 50 nm after scaling according vortex numbers. ……….…………………………….………….…………………….…………115

Figure 55. Extracted magnetic field dependence of bulk ZBC from thin film NIS tunnel junctions and comparison with STS direction measurement with a fitting simulated for a vortex core size of 30 nm. ……….…………………………….………….…………………….………………………116

Figure 56. dI/dV curves of a MgB2/insulator/Pb tunnel junctions on MgO (211) at 4.2 K with different magnetic field applied to the parallel to the c and ab axis of MgB2. ……117

Figure 57. dI/dV curves in magnetic field with fitting and the extracted gap values as a function of the applied field. ……….…………………………….………….…………………….…………………118

Figure 58. A schematic structure and a SEM picture of the planar SNS MgB2-TiB2-MgB2 Josephson junctions. [158] ……….…………………………………….………….……………………120

Figure 59. I-V characteristics of a MgB2/TiB2/MgB2 junction at 5, 15, 24, and 31 K. [158] ……….…………………………….………….…………………….……………………………………………….…………121

Figure 60. Temperature dependence of Ic (squares) and fit (solid line), and Rn (dashed line). [158] ……….…………………………….………….…………………….……………………………………….…………121

Figure 61. (a) I-V characteristics of an MgB2/TiB2/MgB2 junction with and without applied 29.5 GHz microwave radiation at 28 K. (b) Microwave voltage dependences of the Josephson supercurrent and the first and second Shapiro step heights (squares) with a simulated fit (lines). [158] ……….…………………………….………….…………………….………………………122

Figure 62. Josephson supercurrent modulation of a MgB2/TiB2/MgB2 junction at 28 K with both increasing (open squares) and decreasing (solid squares) field. [158] .…………124

Page 13: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

xiii

Figure 63. (a) I-V characteristics for a single junction at 37.2 K, with and without 12 GHz microwave radiation. (b) Microwave power dependence of the Josephson supercurrent and first-order Shapiro steps. (c) Junction critical current (circles) and resistance (triangles) versus temperature. The dashed and solid lines are fits with and 3, respectively. (d) critical current versus temperature near Tc. [164]

… … … . … … … … … … … … … … … . … … … … . … … … … … … … … . … … … … … … … … . … … … … 1 2 6

Figure 64. (a) I-V characteristics of a 20-junction array at 37.5 K, with and without 12 GHz microwave radiation. The inset is a SEM image of an ion implantation mask after etching used to create a multijunction array. (b) dV/dI vs V for the array. [164] …………127

Page 14: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

xiv

ACKNOWLEDGEMENTS

I would like to thank all people who contributed to this project in different ways

and whose support and help made this thesis possible. First, I would like to express my

greatest gratitude to my thesis advisor Professor Xiaoxing Xi for his valuable guidance,

encouragements, and support in the course of my studies at the Pennsylvania State

University. I am also sincerely thankful to Professor Julian D. Maynard, Professor Peter

E. Schiffer, and Professor Suzanne E. Mohney for serving on my thesis supervisory

committee. I want to thank Professor John M. Rowell for his insightful comments and

suggestions. I am grateful to Professor Qi Li and Professor Joan M. Redwing for the

guidance and encouragement. I am in debt to all other colleagues in Professor Xiaoxing

Xi’s group, Professor Qi Li’s group, and Professor Joan M. Redwing’s group at the

Pennsylvania State University for their assistance throughout the years. Dr. Ke Chen and

Dr. Alexej Pogrebnyakov provided me with a lot of important insight to the

superconducting thin films and Josephson devices. I want to thank all other colleagues for

their support and discussion. These members include: Dr. Rudeger H. T. Wilke, Mr. Dan

Lamborn, Dr. Pasquale Orgiani, Dr. Valeria Ferrando, Dr. Jun Chen, Dr. Eric T. Wertz,

Dr. Shengyong Xu, Dr. Yufeng Hu, Mr. Arsen Soukiassian, Dr. Venimadhav Adyam, and

Dr. Dmitri Tenne. My thesis would not be possible without collaborations with Professor

Robert C. Dynes’ group at University of California at Berkeley, Professor Robert A.

Buhrman’s group at Cornell, and Professor Roman Sobolewski’s group at University of

Rochester. It has been pleasure and great learning opportunity to work with Dr. Shane A.

Cybart, Mr. John Read, and Mr. Marat Khafizov. I would like to specially thank my

Page 15: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

xv

parent, Mr. Wenshi Cui and Mrs. Zhimei Zhang, and my brother Dr. He Cui. Without

them, I would never be the one I am today. Last but not least, I want to thank all my close

friends for their encouragement and support.

Page 16: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

1

Chapter 1

Introduction

1.1 Overview

Superconductive digital circuits based on Josephson junctions are desirable for

applications requiring ultrahigh speeds unachievable by semiconductors. Niobium (Nb)

based small asynchronous circuits of rapid single flux quantum (RSFQ) logic have

already been demonstrated at 770 GHz [1], and clocked RSFQ circuits are expected to

exceed 100 GHz [2]. However, the operation temperature of Nb based circuits is limited

to about 4.2 Kelvin (K) due to the low transition temperature Tc of Nb (~ 9 K), which

requires use of either costly liquid helium (He) or bulky and expensive cryocoolers with

high power consumption. [3,4] The high temperature superconductors (HTS) discovered

in the 1980s [5,6] would have advanced the field, but a reproducible uniform HTS

Josephson junction technology has not been developed after 20 years of extensive

research.

The superconductivity discovered in magnesium diboride (MgB2) [7] gives great

promise for Josephson junctions and circuits partly because of its relatively high

transition temperature of ~ 40 K. Unlike HTS, MgB2 is a phonon-mediated conventional

metallic superconductor. Low resistivity of less than 1 μΩ-cm gives MgB2 advantageous

Page 17: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

2

noise properties in superconducting quantum interference devices (SQUID), and a longer

coherence length of about 5 nm [8] would reduce the effect of the degraded layer at the

superconductor-barrier interface. MgB2-based circuits are expected to operate at over 20

K, achievable by a compact cryocooler with roughly one-fifth of the mass and one-tenth

of the power consumption for a 4.2 K cooler of the same cooling capacity [3,4].

Moreover, the larger energy gap in MgB2 [9,10] could lead to even higher clock speeds

(at very high values of critical current density) than Nb-based circuits, because the

ultimate limit of the superconductive circuit speed depends on the IcRn product, which is

proportional to the energy gap of the superconductor.

My research has been focused on fabricating high quality HPCVD MgB2 thin

films and utilizing them to make high quality and reproducible Josephson junctions to

demonstrate the feasibility of an MgB2 Josephson technology. In this process, I have also

been involved in research on the unique electrical, magnetic, and optical properties

associated with the two-band superconductor of MgB2.

1.2 Superconductivity

Superconductivity was discovered by Heike K. Onnes in 1911, when he observed

that the resistivity of mercury abruptly disappeared at 4.2 K using the then recently

discovered liquid helium as a refrigerant. [11] Since then, many elements (Sn, Pb, In, Al,

Nb…) in the Periodic Table were found to be superconductors, with all of their Tc less

than 10 K. Some metallic compounds, for example NbN, Nb3Sn, Nb3Ge, and the most

Page 18: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

3

recent one, MgB2, have also been discovered to be superconducting below Tc of 15 K, 18

Figure 1. The timeline of superconductivity discovery in different materials. [12]

Figure 2. Illustration of the Meissner effect. Magnetic field lines are excluded from a

superconductor when it is below its critical temperature. [12]

Page 19: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

4

K, 23 K, and 39 K, respectively. These superconductors are classified as low temperature

superconductors (LTS). In 1986, a family of cuprate–peroskite ceramics was discovered

to be another class of superconductors, known as high temperature superconductors

(HTS), with Tc as high as 164 K. Figure 1 shows a timeline of superconductors and their

discovery time. [12]

The Meissner effect (or Meissner-Ochsenfeld effect) [13] is another defining

characteristic of superconductivity, in addition to zero dc electrical resistance. As

illustrated in Figure 2, when a superconductor is placed in a weak external magnetic field

H, the field penetrates the superconductor for only a short distance λ, which is called the

London penetration depth, after which it decays quickly to zero [14]. If the applied

magnetic field is too large the Meissner effect does not apply any more. Superconductors

are classified into two groups according to how the magnetic field affects the

superconductor. If the superconductivity is destroyed abruptly as the applied field

exceeds the critical value of Hc, it is called a type I superconductor. In contrast, a

superconductor is called a type II superconductor if a mixed state occurs, in which an

increasing amount of magnetic flux penetrates through the material, but there remains no

resistance to the flow of electrical current when the applied field is larger than a critical

value Hc1 but less than a critical value Hc2. The mixed state is known as the vortex state,

where the flux lines run through narrow non superconducting regions, surrounded by

vortices of screening supercurrents maintaining the superconductivity inside the rest of

the superconductor. The coherence length ξ, which is a measure of the shortest distance

over which superconductivity may be established, is typically smaller than the London

Page 20: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

5

penetration depth, λ, in type II superconductors. Ginsburg-Landau parameter, κ, is

defined by λ/ξ. The vortices can self-arrange in a lattice-like structure known as the

Abrikosov (vortex) lattice. Almost all the pure elemental superconductors (except

niobium, technetium, vanadium and carbon nanotubes) are of Type I, whereas most

impure and compound superconductors are of Type II.

The underlying mechanism of superconductivity was not clearly understood until

1950s, when theoretical condensed matter physicists arrived at a solid understanding of

"conventional" superconductivity, through a pair of remarkable and important theories:

the phenomenological Ginzburg-Landau theory [15], and the microscopic BCS theory

[ 16 ]. However, a satisfactory theoretical explanation about high temperature

superconductivity remains elusive.

Ginzburg-Landau theory of superconductivity was developed by Landau and

Ginzburg in 1950. [15] The theory combines Landau's theory of second-order phase

transitions with a Schrödinger-like wave equation, and explains superconductivity in

macroscopic scale with great success. Abrikosov later showed that Ginzburg-Landau

theory predicts the division of superconductors into the two groups: type I and type II.

[17] Abrikosov and Ginzburg were awarded the 2003 Nobel Prize for their work together

with Leggett.

BCS theory was proposed by Bardeen, Cooper, and Schrieffer in 1957. The

theory assumes the existence of an attractive potential between electrons in the

neighborhood of the Fermi surface. Electrons under such a potential forms electron pairs

(called Cooper pairs), which has a total spin S=0, a total angular momentum L=0, and a

Page 21: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

6

net-zero momentum at the center of mass. A simplified explanation of this electron pair

formation is as the following: the electron attracts the positive ions inside the lattice and

this attraction can distort the positively charged ions in such a way as to attract other

electrons. This process if often referred as the electron-phonon coupling interaction. This

attraction due to the displaced ions can overcome the Coulomb repulsion between

electrons and cause them to pair-up. Usually, the pairing occurs only at low temperatures

and is quite weak, while the paired electrons may be hundreds of nanometers apart. The

distance between electrons in a Cooper pair is the coherence length ξ, which is also a

measure of the shortest distance over which superconductivity may be established.

The Cooper pairs are not independent at all. On the contrary, sufficient

experimental evidence indicates they are strictly correlated and possess the same

quantum ground state. Superconductivity is sometimes called quantum phenomenon on a

macroscopic length scale because, in the superconducting state, all the electron pairs

occupy the same quantum state and their ensemble can be described by one single

macroscopic wave function:

| | (1.2.1)

where is the phase of the Cooper pairs. Many remarkable properties of

superconductors originate from such a macroscopic ground state occupation.

Page 22: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

7

The formation of electron pairs by the attractive potential lowers the total energy

of the whole electron system. Such a negative potential energy associated with an

electron pair is the binding energy of the Cooper pair, which is often called the energy

gap, or the pairing potential. According to the BCS theory, the energy gap is given by [18]

ħ ħ (1.2.2)

where D is the Debye frequency, 0 is density of state at the Fermi surface in the

normal state, and is the attractive energy forming Cooper pairs. The energy gap is

highest at low temperatures and decreases with temperature as it is increased toward Tc.

Another important prediction from BCS theory is that the density of states in a

superconductor at 0 K can be described by

| |/√ | | | | (1.2.3)

Figure 3. The density of state as a function of energy around the Fermi surface in a

superconductor.

Page 23: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

8

where is the density of state in the superconducting state, Δ is the half width of the

energy gap, and is the energy relative to Fermi energy . Figure 3 illustrates

the density of states as a function of energy around the Fermi surface in a superconductor.

The filled area under the curve depicts occupied states while the blank area shows the

empty states. There are no states inside the energy gap for a good superconductor. For

T>0 K, some quasiparticles of electrons or holes are thermally excited outside the

superconducting gap, as shown in Figure 3.

BCS theory also gives the superconducting transition temperature Tc in terms of

characteristic phonon energy ħ and the electron-phonon coupling strength 0 :

. ħ (1.2.4)

where is the Boltzmann constant. In strong electron-phonon coupling, can be

expressed as [19]:

ħ. . . (1.2.5)

Where is the electron phonon coupling constant, and is the Coulomb pseudopotential.

Currently, superconductors are found in a wide variety of applications. Besides

large scale applications of powerful electromagnets as in MRI machines and particle

accelerators, superconductors are found in electronics applications of ultrafast digital

circuits, voltage standards, and sensitive magnetometers.

Page 24: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

9

1.3 Magnesium Diboride

Magnesium diboride (MgB2) is an inexpensive and simple binary compound

material first synthesized in 1953, but its superconductivity was not discovered until 2001

[7]. The bulk critical temperature (39 K) is the highest in conventional phonon-mediated

superconductors. The crystal structure of MgB2 is illustrated in Figure 4 (a), it is a

hexagonal (space group P6/mmm or no.191 [7]) structure consisting of honeycomb boron

(B) layers with close-packed magnesium (Mg) layers between them. A lot of interesting

properties make MgB2 a very special binary compound, such as the simultaneous

Figure 4. (a) Hexagonal structure of MgB2 consisting of honeycomb B (blue) layers

with close-packed Mg (yellow) layers between them. The bulk lattice constants are

determined by XRD to be a = 3.086Å and c = 3.524 Å. (b) The 2-D covalent σ bonds

(brown) are formed within B sheets by overlapping of sp2 hybrid B orbitals and the 3-

D metallic π bonds (green) are formed by the p orbital electrons perpendicular to the

layers. [39]

(b)(a)

Page 25: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

10

existence of two energy gaps, relatively long coherence length (~5 nm), and high critical

current density. The simultaneous existence of two bands or two gaps not only changes a

variety of physical properties, [9,39] it also provides opportunities to study new physical

phenomena not possible in conventional single gap superconductors [20,21,22,23,24,25].

The possibility of multiple gaps in one superconducting state was first predicted

in Ref. [26,27]. Indication of two band superconductivity was reported in Nb doped

SrTiO3 [28] without further confirmation. Since the discovery of superconductivity in

MgB2 [7], first principle calculations [9,10] predicted coexistance of two distinct energy

gaps in MgB2: the larger σ gap is two dimensional and confined to the crystallographic ab

plane, and the smaller π gap is three dimensional, as shown in Figure 4 (b). Experimental

evidences of two band superconductivity in MgB2 includes point-contact spectroscopy

[ 29 , 30 , 31 , 32 ], high-resolution photo-emission spectroscopy [ 33 ], specific heat

measurements[34,35], and Raman spectroscopy[36].

MgB2 has a relatively high of 39 K, compared to conventional superconductors.

According BCS theory, the transition temperature obeys:

. ħ (1.3.1)

The phonon energy in MgB2 is not much higher than other borides or light element

binary compounds. The density of states is not high as well, since MgB2 has no d

electrons. Therefore, the relatively high is believed to be the result of the strong

electron-phonon interaction as expressed through the pairing potential . There is only

one high frequency optical phonon mode (515 cm-1) in MgB2, the B-B stretch mode of

Page 26: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

11

the E2g phonon [9] as shown in Figure 5. The electronic states dominated by orbitals in

the B plane strongly couple to the E2g phonon [37], which makes pair formation favorable.

The electron phonon coupling to this E2g phonon mode in the σ-bonding states is found to

be strong with electron phonon coupling constant 1 [37], whereas the coupling for

π-bonding states is in the weak coupling regime of 0.44 [38]. Different strength of

coupling between the two bands and the phonon mode leads to different superconducting

gaps. The strong coupling strength between σ-bonding states and the E2g phonon mode of

leads to the high transition temperature. [9]

The electrical band structure of MgB2 is quite simple by containing electrons only

from s and p orbitals. Electronic band structure calculations show that the Mg donates

both its 3s electrons to the B layer. [9] Inside each B layer, overlapping of sp2 hybrid B

orbitals forms strong covalent 2D σ bonds, like the σ bonds between C atoms in benzene,

Figure 5. The B-B vibrational mode of the E2g phonon strongly couples to σ-bonding

states. As B atoms move in the arrow directions, elongated bonds (marked with ‘R’)

become repulsive to electrons, whereas shortened bonds (marked with ‘A’) become

attractive. The σ-bonding states strongly couple to the E2g phonon mode because they

are mainly located in either the attractive or the repulsive bondings of the mode. The

π-bonding states do not couple strongly to this mode. [9]

Page 27: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

12

and these σ electrons are restricted to the B layer and conduct only in this plane. Between

the B layers, the pz orbitals extend above and below the plane and form metallic-type 3D

π bonds, which connect the adjacent B sheets through inert Mg ions and allow metallic

conduction perpendicular to the B layers as well as parallel to the B layers. [39] The

geometry schematic of the electron densities associated with these bands is showed in

Figure 4 (b).

The electronic band structure and Fermi surface of MgB2 is shown in Figure 6

from first principle calculations [9,37]. In the reciprocal space, the MgB2 band structure

can be compactly represented as Fermi surface, the energy contour separating filled

electronic states from empty states. The two vertical cylindrical sheets of the Fermi

surface along Γ-Α line are associated with the σ bands, which are incompletely filled or

hole-like. The π bands form an electron-like 3 dimensional tubular network around H and

L (red), and a hole-like tubular network around K and M (blue).

Figure 6. Electronic band structure (left) and Fermi surface (right) of MgB2. [9,37]

Page 28: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

13

The superconducting energy gap on the Fermi surface from first principle

Figure 7. (a) The calculated superconducting energy gap of MgB2 on the Fermi surface at

4 K. (b) the distribution of gap values at 4 K. (c) local distribution of the energy gap is

plotted on planes at 0.05, 0.10, and 0.18 nm above a B plane, respectively. [9]

Figure 8. Calculated temperature dependence of (a) the superconducting gaps and (b) the

quasiparticle density of states. [9]

(b)(a)

(c)

(b)(a)

Page 29: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

14

calculation is shown in Figure 7 [9]. Without considering the functional shape on the

Fermi surface, the energy gap on the π sheets was calculated to be 1.2 – 3.7 meV with an

average of 1.8 meV, while the energy gap on the σ sheets was calculated to be 6.4 – 7.2

with an average of 6.8 meV. A local distribution of the energy gaps is shown in Figure 7

(c). Figure 8 shows the distribution of the energy gaps at different temperatures

calculated by Choi et al. [9]. The temperature dependence of the gaps is of the form

1 (1.3.2)

where 1.8 for the π band and 2.9 for the σ band. Both gaps disappear at the

same transition temperature because finite electron-phonon coupling between the two

bands. Figure 8 (b) shows a calculated quasiparticle density of states as a function of

energy in MgB2 at different temperatures. The quasiparticle spectrum shows two

thresholds because of different magnitude of energy gaps. The quasiparticle density of

states can be deduced experimentally from tunneling experiments and will be discussed

later in detail.

MgB2 has many important practical electronics applications beyond scientific

research. As we discussed at the beginning, it could be used in high speed RSFQ digital

circuits with the benefit of significant reduction in cryogenic requirement and, possibly,

with even higher operation speed. Superconducting quantum interference devices

(SQUIDs) [40] with excellent properties well over 20 K have also been demonstrated. It

is further realized that MgB2 is also a promising material for radio frequency (RF) cavity

applications or photodetection because of its high Tc and low resistivity or its fast

Page 30: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

15

photoresponse. [41,42,100] MgB2 can also play very important roles in large scale

applications. As of now, most large current-carrying superconducting coils are made

either from Nb-Ti, which is a 9-K superconductor with an Hc2 of 10 T at 4.2 K, or from

Nb3Sn, which is an 18-K superconductor with an Hc2 of 28 T at 4.2 K. Some high

temperature superconductors have high Hc2 values of over 50 T, but the high anisotropy

(Hc2()/Hc2()>20) and weak-link nature of grain boundaries make it very difficult to

use them for practical applications. [43], [44] Carbon alloyed MgB2 has demonstrated

higher upper critical field (Hc2 > 60 T) [45] than Nb-Ti or Nb3Sn at all temperatures [46].

Record high Hc2 together with low material cost and relatively low anisotropy makes

MgB2 an ideal candidate in high magnetic field applications, in particular, in cryogen-free

Magnetic Resonance Imanging (MRI) systems [47].

1.4 Josephson Junctions and Their Applications

In 1962, Brian D. Josephson predicted the quantum phenomenon of the passage of

Cooper pairs through a weak connection between two superconductors, even via a thin

insulator layer. [48] Josephson also predicted that if the current did not exceed a limiting

value, so called Josephson supercurrent or critical current, there would be no voltage drop

needed across the tunnel barrier. This phenomenon is called the Josephson effect, and the

weak connection structured device is called a Josephson junction (JJ). Figure 9 shows the

current–voltage characteristic of a superconducting tunnel junction originally predicted

by Josephson. [49] The predicted supercurrent in a Josephson junction obeys the weak-

link current-phase relationship [50], or the dc Josephson effect, of

Page 31: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

16

sin (1.4.1)

where I(t) is the current across the junction and Ic is a constant, the critical current of the

junction, and Δφ(t) is the phase difference between the wave function inside the two

superconductors comprising the Josephson junction. The physics of the derivation of

Josephson supercurrent will be discussed further in detail later. Basically, it predicts that

the Josephson supercurrent is proportional to the sine of the phase difference across the

insulator, and may take values between -Ic and Ic. The constant, Ic, is an important

phenomenological parameter of the device that can be affected by temperature as well as

by an external magnetic field.

The advantage of such a device is that it has two stable states, and switching

between them can occur in picoseconds (100 GHz). The zero voltage state is stable until

the critical current Ic is exceeded, when a voltage is developed across the barrier. Such a

non-zero voltage state can trace back to the zero voltage state along the I – V curve with

Figure 9. The current – voltage characteristic of a superconducting tunnel junction

originally predicted by Josephson. [49]

Page 32: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

17

decreased current. The fast switch between two stable state make Josephson tunnel

junction an ideal candidate for ultrafast digital electronics applications.

Josephson further predicted that an alternating supercurrent would be generated

across the junction barrier in addition to the dc current if a constant voltage were

maintained. This phenomenon is often called the ac Josephson Effect. This can be

generated from the superconducting phase evolution equation [50]

∆ (1.4.2)

where is the voltage across the junction, e is the elementary charge and h is Planck’s

constant. The physical constant often used, , is the magnetic flux quantum

(2.07x10-11 T cm2), the inverse of which is also called the Josephson constant, . With a

fixed voltage V across the junctions, the phase will vary linearly with time and the current

will be a periodic current with amplitude Ic and frequency 2eV/h. Conversely, one can

infer that a Josephson junction can act as a quantum dc voltage–to–frequency converter,

to an unparalleled accuracy. Based on this principle, Josephson junction arrays are now

widely used for voltage standards in national standards laboratories including the

National Institute of Standards and Technology (NIST) and the Josephson constant,

, is currently recommended by the Committee on Data for Science and

Technology (CODATA) to be [51]

483 597.891 (1.4.3)

Page 33: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

18

A unique signature of Josephson junction is the modulation of the junction critical

current by an external magnetic field perpendicular to the current flow direction. Surface

currents circulate the tunnel barrier to screen an applied magnetic field from the interior

of the Josephson junction just as if the tunnel barrier itself were a local weak

superconductor region. Such a screen process produces a reduced effective area of the

transport current, which varies with field magnitude. The result is that the supercurrent

goes through maxima and minima with increased magnetic field as shown in Figure 10.

The underlying mechanism will be described and compared with our MgB2 Josephson

junction results later.

-4 -2 0 2 40.0

0.2

0.4

0.6

0.8

1.0

Ic/Ic

(0)

Φ/Φ0

Figure 10. An ideal Fraunhofer pattern: supercurrent in a Josephson junction can be

modulated by an external magnetic field perpendicular to the current flow direction.

Page 34: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

19

Today, Josephson junctions are found in many applications, such as high speed

electronic switches and memories, sensitive magnetic field detection of brain or heart,

voltage standard, and radiation emission and detection in radioastronomy. In particular,

many superconductive electronic applications use Josephson junctions as their key

component. Josephson junctions can be divided into types of superconductor-insulator-

superconductor junctions (SIS), superconductor-normal metal- superconductor junctions

(SNS), thin bridge junctions, point contact junctions, grain boundary junctions, and

intrinsic junctions.

The most successful and widely used Josephson junctions for these digital circuits

are SIS Josephson tunnel junctions by a trilayer Nb/AlOx/Nb process, developed by John

M. Rowell and his colleagues in the early 1980s. [52], [53] Integrated circuits consist of

over 10,000 Josephson junctions using this process have been fabricated, demonstrating

unmatched uniformity, reproducibility and reliability.

Page 35: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

20

Chapter 2

Experimental Methods

In this chapter, I first briefly review the general principles and key elements of

MgB2 growth. Then, I discuss the HPCVD technique used for the growth of high quality

MgB2 films, as well as other thin film techniques, and a number of analysis tools for thin

film property characterizations. Finally, I describe lithographic processing and the device

measurement apparatus and method.

2.1 Hybrid Physical-Chemical Vapor Deposition

High quality MgB2 films are very important for electronics, electromagnets, and

RF cavity applications as well as fundamental scientific research. Extensive research

efforts have been devoted to MgB2 thin film development by various techniques [54].

The techniques include high-temperature ex situ annealing of B or Mg-B precursor films

in Mg vapor [55,56,57,58,59,60,61], intermediate-temperature in situ annealing of Mg-B

precursor films [ 62 , 63 , 64 , 65 , 66 , 67 , 68 ], low-temperature in situ deposition

[68, 69 , 70 , 71 , 72 ], and high- and intermediate-temperature in situ deposition

[73,74,75].The pros and cons of each technique have been analyzed in Ref. [76,77].

Page 36: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

21

There are certain requirements that must be satisfied in order to grow high quality

MgB2 films. The most important one is an exceptionally high magnesium partial vapor

pressure necessary for the thermodynamic phase stability at the elevated growth

temperatures. A pressure-temperature phase diagram calculated for the Mg-B system by

Liu et al [78] is showed in Figure 11. In order to achieve thermodynamically stable

growth, the pressure and temperature must fall in the window marked by “Gas + MgB2”,

where Mg gas and MgB2 are the stable phases. The magnesium pressure required is very

high with this growth window for high temperatures necessary for good crystallinity of

Figure 11 The pressure-temperature phase diagram calculated for the Mg:B atomic

ratio xMg:xB ≥ 1/2. The region marked by “Gas + MgB2” represents the thermodynamic

stable window for MgB2 film growth. [78]

Page 37: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

22

MgB2. For example, more than 40 mTorr is required at a growth temperature of 750 °C,

which is prohibitive to most high vacuum techniques.

In addition to the high magnesium pressure, high purity sources of Mg and B as

well as a clean deposition environment are other key requirements to grow high quality

MgB2 films. Oxygen (O) and carbon (C) contaminations are often observed in MgB2

films grown by various deposition techniques. Oxygen contamination is believed to be

associated with background pressure [79] while carbon is often from the impurities in the

deposition targets.

Fortunately, besides all this difficulties, the stochiometry control is not a critical

problem for the MgB2 deposition. As long as the stoichiometric ratio of Mg/B is above

1:2, the pressure-temperature phase diagram is identical. [78] As a result, MgB2 is the

only solid deposit on the substrate and extra Mg will be in the gas phase and evacuated

from the system [78]. In addition, MgB2 is thermally stable once it is formed because a

significant kinetic barrier has to be overcome before it is thermally decomposed. [80]

This makes maintaining a magnesium vapor pressure around the film during cooling after

growth easy to achieve.

Chemical vapor deposition (CVD) is a chemical process of one or more volatile

precursors reacting and/or decomposing on the heated substrate surface to produce high-

purity, high performance solid materials. [81] The advantage of CVD technique is

capable of growing a large variety of thin films with controllable stoichiometry and

desired properties.

Page 38: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

23

A variant of CVD process, hybrid physical-chemical vapor deposition (HPCVD)

technique [65,75], was developed for MgB2 film growth. This hybrid process combines

the feature of both physical vapor deposition and chemical vapor deposition by

depositing films from vaporized solid state material, in this case, Mg, as well as from a

gaseous precursor as in regular CVD. A schematic of HPCVD is illustrated in Figure 12.

It consists of a vertical water-cooled quartz tube and an inductively heated susceptor

made of stainless steel or iron (Fe). The B source, diborane (B2H6) precursor gas (0.1 – 5

at% in hydrogen) and purified ultrahigh purity (UHP) hydrogen carrier gas flow from the

top into the quartz reactor where the total pressure is controlled to be 80 – 100 Torr. The

Mg vapor is locally evaporated from bulk Mg pieces placed next to the substrate on the

susceptor. The Mg vapor pressure and the substrate temperature are controlled by

inductive heating. A photo of the HPCVD MgB2 film growth system is shown in Figure

Figure 12 (a) A Schematic of the HPCVD setup. (b) Computer simulated gas velocity

profile in the reactor near the susceptor. [77]

Page 39: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

24

13. HPCVD has been proven to be the most effective technique for MgB2 thin film

growth. [77]

A modeling study of HPCVD growth was carried out for chemical and

transportation process. [82] First, diborane precursor gas is thermally decomposed into

gas phase BH3 and adsorbed on the substrate surface. This is believed to be the limiting

step for the growth rate of MgB2, as the deposition rate is found to be linearly dependent

on the diborane flow [88]. When magnesium is evaporated into vapor and reacts with

surface-adsorbed BH3, MgB2 starts to grow onto the substrate and excess hydrogen gas is

released.

2 (2.1.1) (2.1.2)

Figure 13. A photo of the original HPCVD MgB2 film deposition system.

Page 40: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

25

(2.1.3) 2 3 (2.1.4)

Josephson tunnel junction is very sensitive to the quality of tunnel barrier material

and the barrier-superconductor interfaces. It is of great advantage for multilayer to be

grown in situ. An integrated in situ deposition system has been recently setup to fulfill

this purpose. The system consists of two computer-controlled HPCVD modules and a

sputtering chamber with an interlocked high vacuum (HV) transfer station. Both HPCVD

modules have started to produce superconducting MgB2 films with Tc of over 40 K.

Figure 14. A photograph of the integrated in situ HPCVD system consisting of 2

computer controlled CVD modules and a sputtering module with an interlocked transfer

station.

Page 41: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

26

HPCVD module #1 is designated to grow MgB2 films for in situ multilayer

deposition. It is equipped with stepping motors to control vertical and rotational

movement of substrates so that substrates can be transferred into and out of the HPCVD

chamber. The Mg source and substrate are heated independently with resistive heaters. A

reaction zone is defined by a quartz cylinder placed around the heaters for Mg and the

substrate.

HPCVD module #2 has a rotating pocket heater originally designed for MgB2 co-

evaporation at STI. [74] A rotating platter can hold one 4-inch wafer or 3 2-inch

substrates simultaneously, and spin them at several hundred rpm. During about one third

of a rotation, substrates are exposed to diborane gas flow through an opening of the

pocket heater. Mg is evaporated from an oven and transferred into the heater pocket, and

is largely confined in the pocket by means of a small gap between the platter and the

pocket. It is not possible for substrates to be transferred between HPCVD module #2 and

other chambers.

2.2 Sputter Deposition of TiB2, MgO, Al, AlN, Au, and Cr

Magnetron sputtering was done either in sputter deposition systems or in a

sputtering chamber which is part of the recently installed integrated HPCVD system.

Both dc and pulsed dc sputtering are available in the sputter deposition system, whereas

dc and RF sputtering are available in the sputtering chamber of the integrated HPCVD

system.

Page 42: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

27

There are two sputtering deposition systems in our lab capable of dc or pulsed dc

sputtering. Both are 12-inch spherical 304 stainless steel vacuum chambers containing

multiple 3¾, 6, and 10 inch inch ports. Three sputtering guns can be connected to either

of the dc or pulsed dc sputtering power supplies through the interchangeable cable

connection. Both chambers are equipped with a compact full range gauge (either Balzer

PKR 251 or Pfeifer PKR 251) for pressure control and MKS mass flow controllers for

gas delivery.

One chamber is dedicated to sputtering with a 2-inch gun from AJA corp. The

base pressure can reach lower than 1 x 10-7 mbarr after baking or pumping for extended

period. It has a fixed sample stage which can be heated to 900 °C, loading from the 10-

inch port opposite to the sputtering gun. It was used to sputtering TiB2, Nb, or NbN.

Substrates were glued to the sample stage by silver paint for good thermal conductivity.

The sample stage was usually cleaned by sand paper, and then acetone and isopropanol.

The distance between the sample and the sputtering target is 4 cm and the deposition rate

of TiB2 is 400 Å/minute at an Ar background pressure of 1.5 x 10-2 mbarr with a dc

power of 200 watts. Samples are usually heated to 750 °C during TiB2 deposition to

enhance crystallinity.

Cr/Au (or Ti/Au) is usually used to make electrical contact for device

characterizations. Two sputtering guns share the same chamber with a 3-cm ion source,

so that Cr/Au (or Ti/Au or Al/Au) can be deposited in situ after the sample surface is

cleaned by a lower energy Ar ion beam (typically 100 eV). One sputtering gun is a

rotatable 2-inch gun and the other is fixed head 1-in gun, both from Kurt. J. Lesker. The

Page 43: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

28

two guns and the ion source are located in the horizontal plane with the 1-inch gun in the

middle and the other two 45 degree to it, all facing the center of the chamber where a

rotatable aluminum sample stage can be loaded from the top 10-inch port. The sample

stage was machined from without heating or cooling capability, which is 4-5 inches away

from the sputtering guns and the ion source and can rotate 360 degrees. The sample stage

was rotated 180 degree back away from the sputtering guns to protect the sample during

presputtering (typically 3 minutes), which are usually done to remove target surface

contaminates. A one-quarter-inch thick Au target from Kurt. J. Lesker is typically in the

2-inch gun while a one-quarter-inch Cr (or Al) target in the 1-inch gun for Cr/Au or

Al/Au combination. The Au adhesion to the MgB2 films as well as SiOx was poor without

a thin layer of Cr or Al. The chamber was pumped to a background pressure ~ 6 x 10 -7

Torr (8 x 10 -7 mbar) before usual deposition. The deposition rate of 300 Å/minute for

gold was obtained when the sample directly faced the gun with a dc power of 30 watts in

an Ar background pressure of 3 mTorr (4 x 10 -3 mbar). The Cr deposition rate was ~200

Å/minute with the same power and pressure from the 1-inch gun.

In the sputtering chamber of the integrated HPCVD system, both dc and RF

sputtering are available. Samples up to 2 inch in diameter can be transferred between the

sputtering chamber and the HPCVD module #1 in situ through the transfer station. MgO

is often RF sputtered as a tunnel barrier material, or as a protection layer on top of MgB2

films, which was proved to be very effective against degradation due to moisture in air.

The distance between the sample and the target is 15 cm, which yields a deposition rate

of 0.08 Å/s with an RF power of 200 watts at an Ar background pressure of 1 mTorr.

Page 44: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

29

Sample stage can accommodate samples as large as a 3-inch wafer and is with a radiation

heater and a speed-adjustable rotating motor to improve film uniformity during sputtering.

In some occations, Al is sputtered by dc and later oxidized thermally or assisted by

plasma to form a barrier layer. The distance between the Al target and the sample is ~30

cm and the deposition rate can be monitored by an Inficon XTC/2 crystal thickness

monitor (CTM).

2.3 Evaporation of SiOx, Pb, and Ag

A thermal evaporation system is used for SiOx insulator film deposition as well as

the deposition of various metals. It is a stainless steel 6-inch six-way cross with a 12-inch

extension equipped with a Pfeifer PKR 251 wide range pressure gauge. A baffle box was

used as the evaporation source for SiOx. SiO powder was loaded into the container about

the half of the size of the baffle box through an opening on the top. The opening is

covered with a cap after loading. When passing a current of 300~350 A through the

baffle box, the vapor circumvents to the other side of the box and is released outside

through the opening on top of that half. This is designed to only release vapor phase

while containing the melted materials. An Inficon XTC/2 crystal thickness monitor was

mounted next to the sample stage to monitor the deposition. The source is heated to

evaporation temperature to outgas and burn off any contaminates from the source before

a shutter is opened for film deposition. [83] Because a large amount of current is used

during the evaporation of SiOx, the chamber is water cooled and the vacuum feedthrough

is fan cooled.

Page 45: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

30

Pb and Ag are usually evaporated through a shadow mask for counterelectrodes.

Evaporation is chosen over sputtering for its better directionality in stripe edges definition

through shadow masks and less destruction for the thin barriers on film because of lower

energy comparing with sputtering. The evaporation process is similar to SiOx deposition

except a tungsten evaporation boat (Kurt. J. Lesker, part EVSME5005W) was used

instead of a baffle box and lower operation current ~55 A is typically used.

2.4 X-ray Diffraction (XRD)

A Philips X’Pert Pro MRD four-circle x-ray diffraction system was used to do θ-

2θ scan to determine film orientation, and to do ω scan (rocking curve measurement) for

crystallinity analysis. It is a fix-anode CuKα system with a high resolution and relatively

faster scan speed. A four-circle x-ray diffraction system was used for most of thin film

characterization tasks with the help from Prof. Schlom’s Group. It has the advantage of

easy alignment in φ scan since it has a lower angular resolution.

2.5 AFM and SEM

Two techniques were used for morphology study: atomic force microscopy (AFM)

and scanning electron microscopy (SEM). AFM has the advantage in small area

quantitative profile mapping and SEM can image larger area with quick feedbacks and

the flexibility of zooming in and out. A DI Nanoscope III was used for film surface

Page 46: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

31

imaging. Tapping mode was usually used in ambient conditions with either Mikromasch

silicon tips of NSC 15 (ULTRASHARP) or DP 15 (HI’RES). Both cantilevers have a

force constant of (40 N/m) and Al backside-coating for laser reflection signal

enhancement. The HI’RES cantilever has a sharper tip for better lateral resolution. Films

are often scanned right after deposition to minimize surface reconstruction or degradation

due to the moisture in the air.

A JEOL 6700F field emission SEM and a FEI Quanta 200 environmental SEM

were used for SEM study. The JOEL FE-SEM has a higher resolution of up to 1 nm

while the FEI ESEM can load more samples simultaneously and is equipped with an

INCA Energy Dispersive X-Ray Spectroscopy (EDS or EDX) for identifying and

quantifying elemental composition of sample areas as small as a few cubic micrometers.

One disadvantage of the EDS technique is that the detection of light elements, in

particular B, is not very accurate and reliable.

2.6 Electrical and Magnetic Characterizations

Low temperature dip probes are typically used to do general electrical transport as

well as ac magnetic susceptibility measurements. The probe consisted of a long ½-inch

stainless steel tube with one end connected to a copper sample stage and the other end

attached with 19-pin military connector (Newark, part. MS3112E14-19P for probe and

MS3116F14-19S for cable). Twisted pairs of bronze wire run through the stainless steel

tube connecting the connector to the copper sample stage for signal noise reduction. A

Page 47: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

32

thermometer (Lakeshore, part DT-470-BO-11) was mounted half an inch lower than the

sample location. A copper cap enclosed the sample stage to prevent unregulated

temperature fluctuation as a result of unexpected air flow, with a small opening at the end

to expedite the cooling process. During the measurements, the probe was lowered down

towards the liquid helium surface for cooling or pulled up away from the liquid helium

surface for heating. The temperature can be well controlled with a temperature resolution

of less than 0.1 K by moving the probe slowly. A Lakeshore 330 temperature controller

was used to acquire the temperature readings from the thermometer.

The sensor reading of the thermometer was manually calibrated by preloading

voltage-temperature conversion curve for the sensor type and comparing it with a

manufacture-calibrated thermometer placed at the sample spot. The reading of the

thermometers were mapped slowly through the temperature range of 4.2 K to 320 K and

fitted by polynomials within different temperature sections and loaded into the Labview

program. Low temperature grease (Apiezon grease N) was typically used to glue the

sample onto the copper stage for its good thermal conductivity and easy sample

attachment and removal. Various metal wires or thin tapes of Au, Ag, Al, were used as

electrical leads and were usually pressed on to the film with indium.

For dip probe transport measurement, a Keithley 2400 source meter and a

Keithley 182 volmeter were used to supply the currents and measure the corresponding

voltages, respectively. Both meters were controlled by the Labview programs via a GPIB

interface to the computer. The bias current of 1 μA~5 mA, depending on the sample

resistance, was triggered from the current source and supplied as pulses of length as short

Page 48: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

33

as 50 msec. The voltages were taken and averaged over the pulse and after usually 3

msec delay for the current to reach equilibrium. The averaging of voltages would reduce

the noise from fluctuation of the signal. Each voltage point was recorded as an average of

two voltages for a plus and a minus pulse with the same current amplitude to cancel the

voltage draft. Current-voltage characteristics or resistance-temperature dependence can

be measured by the 2400-182 combined with a Lakeshore 330 temperature controller.

An ac magnetometry setup is also available with a low temperature measurement

dip probe to characterize dynamic susceptibility of films. The setup consisted of an

EG&G Instruments 7260 DSP Lock-in amplifier, and a primary excitation field coil and a

secondary pick-up coil. During measurement, an alternating magnetic field, Hacsin(ωt), is

generated from the primary excitation field coil, and the slope, χ=dM/dH, of the induced

ac moment, Mac= χHacsin(ωt), is picked up through the secondary pick-up coil. One of

the advantages of ac measurement over dc measurement is that it is very sensitive to the

slope of the moment instead of the absolute value. A small magnetic shift can still be

detected even with a large magnitude of the moment. The dynamic effect of the ac

susceptibility occurs when the ac frenquency is high (7170 Hz was typically used for our

setup), as the ac moment lags behind the drive field. In this case, the additional phase

shift term φ arises, to describe the lag to the drive signal

(2.6.2)

" (2.6.1)

Page 49: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

34

In the low frequency limit, the real component of the susceptibility (the in-phase

component) is just the slope of the moment and the imaginary component of the

susceptibility (the out-of-phase component), indicates the energy dissipation processes in

the sample, in the case of superconductivity, when magnetic irreversibility occurs. Both

real and imaginary component of the susceptibility was recorded as a function of

temperature by the computer through the lock in amplifier and the Lakeshore 330

temperature controller.

2.7 Photolithograpic Device Processing

MgB2 films and most devices based on trilayers or heterostructures were

patterned using contact photolithographic techniques. Most of the processing work was

done at the clean room (class 100) facility at the W. M. Keck Smart Materials Integration

Laboratory. Shipley 1811 photoresist was used in most general structure definitions

because it provides good resolution and appropriate thickness for ion milling. Shipley

1827 (~2.7 μm) photoresist has lower resolution and was only used for the lift-off process.

Initially, a layer of Shipley 1811 photoresist about ~1.1 μm thick was spun on the

samples using 4000 RPM for 40 seconds. Note the “11” in 1811 corresponds to the resist

thickness of 1.1 μm, and, similarly, 2.7 μm for 1827. Samples were soft baked at 100 °C

for a minute on a hot plate to remove any excess solvent in the resist, and then cooled

down to room temperature.

Page 50: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

35

Glass lithography masks with a chrome feature layer were used for the patterning

alignment. Masks were processed either by the MRI Nanofab facility or by a commercial

company Photo Sciences, Inc. Blank masks were purchased from Nanofilm for backup

copies. Masks need to be cleaned by acetone, isopropanol and rinsed by running DI water

and blown dried before exposure in order to clean off contaminates or residue left on the

mask prior to each use. A Suss MJB3 mask aligner at the Keck facility was used to

expose the samples aligned with the mask to UV light. The alignment procedures and

equipment controls are available at the Keck cleanroom facility. The constant-intensity

(CI) mode was used most of the time and the exposure time needed to be adjusted since

the UV lamp intensity varies with time.

For device patterning, the samples were developed in a Shipley 351 developer

(buffered solium hydroxide), usually diluted with 5 times of DI water. Samples were

typically developed for about 45 seconds and rinsed in a DI water bath and blown dry by

nitrogen gas. The water rinse time is kept to a short period, usually just a few seconds,

only enough to dissolve chemicals on the samples to minimize further film degradation.

Because of the sensitivity of MgB2 films to water and chemicals, the developing and

water rinsing are closely monitored. Sometimes, a lift-off is needed. A layer of SiOx was

evaporated after the photoresist was patterned, and then the part of the photoresist

leftover was removed with acetone to lift off the unwanted SiOx to open a window for

metal wiring layer with the desired insulation layer left untouched. The lift-off can be

done through either a single layer resist process or a bilayer or even more complicated

trilayer resist process. A generally used single layer processing need an additional step of

Page 51: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

36

chlorobenzene soak after soft bake to create the overhanging profile needed for lift-off.

The chlorobenzene diffuses into the photoresist causing it to swell and forms a gel into

some depth of distance, and this process makes the top layer of the photoresist having a

slower developing rate so that the developer can undercut the structure and produce a

desired profile. Because of the excellent directionality of thermally evaporated SiOx over

a sputtered film, good lift-offs can often be done with ease without this additional soak

step for our device development. The slower developing speed with the chlorobenzene

soak would very likely degrade part of the MgB2 film where the junction device will be

located. The edges of the features defined by thick resist of 1827 are not as sharp as those

by 1811.

A 3-cm ion source (Veeco/Commonwealth Scientific) was used as the etch tool

for device processing. Ion milling is a similar process to sputtering, with the sample now

as the target. During operation, local Ar+ ion plasma was generated and accelerated

toward the sample through standard collimated grid optics inside a uniform ion beam of 3

cm in diameter. In practical, samples were usually placed within a 2 cm diameter area for

better reproducibility of milling rate. The beam has a neutral charge when arriving at the

samples, with a filament supplying the electrons to the positively charged beam along the

ion beam path. This neutralized current is automatic controlled and the operation mode is

called total beam neutralization (TBN) mode. The cathode filament and the neutralizer

filament need to be replaced every several months depending on the usage.

Page 52: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

37

For ion milling, the base pressure was first pumped down to below 2x10-6 mbarr.

A 5 sccm flow of UHP Ar was supplied to the ion gun with a turbo pump running at a full

speed of 1000 Hz, which yields a Ar pressure of 5 x 10-4 Torr. The sample stage was

tilted with an ion beam incident angle of about 57 degree, which provides less

interference with the reflected beam. The samples are usually mounted with double-sided

conductive copper tapes on the stage during milling at a low voltage (270 V or less) with

water cooling. For high beam voltages of 400 V or above, silver paint was used for

sample mounting together with liquid nitrogen cooling. For most MgB2 film etches, a

beam voltage of 270 V was used with a beam current of 5.5 mA, which is the highest

current it can sustain under such a voltage. The milling rate is ~5.5 nm/min. This

relatively low Ar ion energy and current generate less heat and thus minimal material

0 20 40 60 80 1000

200

400

Mill

ed d

epth

(nm

)

Time (min)

Figure 15 Milling rate of MgB2 with a beam voltage and current of of 270 V and 5.5 mA,

respectively. Red line is a linear fit.

Page 53: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

38

damage from the ion milling, and it also gives a milling rate convenient for controlling

the mill depth in the order of a hundred angstroms. TiB2 was found to be much harder to

mill, with a milling rate of 7 times slower than MgB2 at 270 V and 5.5 mA, and hence

was usually milled using 400 V.

An in situ ion mill and Cr/Au sputtering was usually combined when doing the

top metal wiring layer in the trilayer device fabrication process. A low energy Ar ion

beam of 100 eV was used to clean off surface contaminates after the lithography steps.

The sample holder was also tilted about 45 degree during this milling process and MgB2

milling rate is fairly low at this voltage. The rotatable sample stage allows direct

sputtering deposition after ion milling surface cleaning without breaking the vacuum.

2.8 Device Measurements

Device transport and magnetic property measurements were carried out using a

liquid helium flow cryostat (Cryo Industrials, model BHRCUD). Samples were held in

vacuum and a turbo pumping station was used to pump the vacuum down to below 1

mTorr for heat isolation during cooling. A flexible stainless steel transfer tube was used

to flow liquid helium from a storage dewar to the cryostat for sample cooling and the

flexible hose jacket was also frequently pumped. A precision needle valve at the dewar

end was adjusted to control the helium flow. Samples were mounted using silver paint on

a copper stage with a thermometer to the side. Samples can be cooled down to 4.2 K

within 30 minutes.

Page 54: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

39

A wire bonder (Westbond, model 7400A) was used to make electrical leads from

the sample chip to the sample in the cryostat stage via either Al or Au wires. The

electrical wires were connected directly to a switch box via a 19-pin military connector to

minimize noises from the ambient environment. The switch box is used to select the

device to be measured and can be switched between the “quick scan” mode and the

“digital data requisition” mode. In the quick scan mode, an oscilloscope (either Tektronix

2465B or Tektronix TDS 2014) was used to track the current-voltage relationship across

the junction device through two channels. A triangular current waveform is provided by

an arbitrary waveform generator (HP 33120) through a BNC cable with a resistor

connected in series with the junction device. The current signal through the junction

device is actually monitored by the voltage across the resistor, whereas the voltage signal

across the junction is measured directly. Both the current and voltage signals were

amplified by preamplifiers before being sent to the oscilloscope for fast signal processing

and analysis. When detailed data need to be scanned and recorded by the computer, the

switch box was switched to the “digital data requisition” mode and the current source was

replaced by a Keithley 2400 programmable source meter. The voltage signal from the

junction, after amplification, was sent back to the 2400 meter and recorded by the

computer together with the current signal provided directly from the 2400 meter. The

current-voltage trace can be monitored by the oscilloscope during data requisition.

Josephson junctions were typically measured in shielded room to eliminate

external magnetic fields from the outside environment including the earth's magnetic

field. It was found that by placing a metal cap made of μ-metal (Mumetal) over the

Page 55: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

40

cryostat, the outside signals and magnetic fields were quite effectively shielded from the

device measurements. The Josephson supercurrent observed was significantly suppressed

when the shielding was removed during the measurement. The measurement setup is

equipped with two μ-metal caps for better shielding of the external magnetic fields.

A homebuilt magnetic coil placed inside the vacuum jacket and outside of a heat

isolation shield insider the cryostat. Magnetic field at the center of the space that the coil

enclosed, where the device chips was mounted, was calibrated by a guassmeter (Magnetic

Instrumentation Inc. model 907). The coil was placed in such a way that the magnetic

field direction is in the plane of the device chip and along the MgB2 strip direction for

Frauhofer pattern measurement.

Page 56: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

41

Chapter 3

MgB2 Thin Films and Heterostructures

In this chapter, I discuss the structures electrical, magnetic, and optical properties,

and material stability of HPCVD MgB2 films. The growth and optimization are reviewed.

The growth and characterization of TiB2/MgB2 hetero structures for SNS junction devices

are also discussed.

3.1 Structures of MgB2 Films Grown by HPCVD

Bulk MgB2 has a hexagonal structure with lattice constants of a = 3.086Å and c =

3.524 Å. [7] The c-cut SiC offers a less than 0.5% lattice mismatch with c-axis oriented

MgB2. (for the 4H polytype SiC a = 3.073 Å and for the 6H polytype a = 3.081 Å). The

(0001) sapphire also has the same hexagonal structure as MgB2 but with over 30% lattice

mismatch (a = 4.765 Å). MgB2 films are able to grow on c-cut sapphire with MgB2

lattice rotated by 30° to match the substrate [65]. For tunnel junctions, it is desirable to

grow non c-axis MgB2 films for σ-σ band tunneling to take advantage of the large σ gap.

[84] While it is a challenge to grow ab oriented films, we are able to growth MgB2 films

with tilted c axis on some substrates with surfaces of rectangular lattice. For example,

tilted c-axis epitaxial MgB2 films have been shown on (110) YSZ [85] and (211) MgO

substrate [86].

Page 57: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

42

High quality epitaxial MgB2 films are able to grow on single crystal substrates of

(0001) oriented SiC and sapphire by HPCVD [75], [87]. High deposition temperature

results in excellent crystallinity and epitaxy, while high local Mg vapor pressure ensures

the ideal stochoimetry. Figure 16 shows a high resolution transmission electron

microscopy (TEM) of the film/substrate interface in a MgB2 film on (0001) 6H-SiC

substrate. Top and bottom inset show, respectively, a SAED patterns from the film and

the substrate. The result shows that the MgB2 grows epitaxially on the SiC substrate with

MgB2 orientation (0001) [1120] parallel to (0001) [1120] orientation of SiC. Excellent

crystallinity and epitaxy have been confirmed by x-ray diffraction with θ-2θ and φ scan

and sharp rocking curves. [75], [87].

Figure 16 A high-resolution transmission electron microscope (HRTEM) image for the

interface between a MgB2 film and (0001) 6H-SiC substrate. The insets are selected area

electron diffraction (SAED) patterns from the film (top) and substrate (bottom). [77]

Page 58: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

43

Figure 17. X-ray diffraction of θ-2θ scan and φ scan of HPCVD grown MgB2films.

From [87]

Page 59: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

44

3.2 Transport and AC susceptibility properties of MgB2 Films

High Mg partial pressure and high purity B source together with the highly

reducing hydrogen ambient used in HPCVD result in very clean MgB2 thin films [88].

Figure 18 illustrates a resistivity versus temperature curve for a 300 nm MgB2 thin film

on SiC substrate. The room temperature resistivity of MgB2 was 8.6 μΩ-cm, and it

0 50 100 150 200 250 3000

4

8

40.5 41.0 41.5 42.00.0

0.2

0.4

ρ (μ

Ω−

cm)

T (K)

ρ (μ

Ω−c

m)

T (K)

Figure 18. Resistivity vs temperature curve for a 300 nm MgB2 thin film on SiC

substrate. The inset shows details near the superconducting transition.

Page 60: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

45

decreased quickly with temperatures. As the 41.6 K, the film had a residual resistivity ρ0

of ~ 0.42 μΩ-cm, which makes the residue resistivity ratio (RRR) of 20.5. The transition

was slowly measured during the heating for better accuracy of temperature reading. The

resistivity remained zero at 41.48 K and started to increase abruptly. The transition

completed at 41.57 with ΔTc of less than 0.1 K. The typical zero resistance transition

temperature Tc(0) for HPCVD grown MgB2 films is 40-41.5 K varying with substrates

and film thickness, clearly higher than bulk MgB2 values. Some sub-micron thick MgB2

films on sapphire can have residual resistivity ρ0 as low as 0.1 μΩ-cm and RRR as high

36 39 42-1.0

-0.5

0.0

0.5

(Real Part) (Imaginary Part)

AC

Sus

cept

ibilit

y

T (K)

Figure 19. Real (red) and imaginary (blue) component of the AC magnetic

susceptibility measurement of a MgB2 film as a function of temperature.

Page 61: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

46

as 80 with Tc over 40 K. [77] Thick films on SiC tend to crack during sample cooling due

to larger coefficient of thermal expansion in MgB2 than in SiC.

As MgB2 is cooled below its transition temperature, like many superconductors, it

becomes a perfect diamagnet with a susceptibility of χ=-1, from a normal metal, which

usually has a very small susceptibility. Figure 19 shows the temperature dependence of

the real and imaginary component of the ac susceptibility of a 50 nm thick MgB2 film.

The real part of the susceptibility stayed flat up to 40.65 K, indicating an excellent

diamagnetic property, and started a sharp upturn and finished the transition at 40.8 K,

with a normal metal state. The transition was sharp and stayed flat below the transistion,

indicating good Tc distribution inside the film. The imaginary susceptibility of the MgB2

film shows a peak at 40.8, indicating intense energy dissipation as the magnetic

irreversibility occurs. The peak location and the sharp feature also confirm high Tc and

clean transitions in HPCVD MgB2 films.

3.3 Electron Scattering Dependence of Dendritic Magnetic Instability in MgB2 Films

Most MgB2 bulk and thin film samples exhibit dendritic magnetic flux

instabilities in their superconducting states, as shown in suppressed hysteresis loops of

magnetization [89] and dendritic flux structures observed in Magneto-Optical Imaging

(MOI) experiments [90]. The instability starts as an external magnetic field applied, the

flux penetration dissipate heat and leads to a local temperature rise, which reduces flux

Page 62: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

47

pinning and assists further flux motion. This process is, therefore, called flux jump, and

can be observed using a magneto-optical imaging technique [90-92].

Such low temperature dentritic flux jumps are, however, absent in clean HPCVD

MgB2 films. [91,92] Figure 20 shows MOI images of a zero-field-cooled (ZFC) MgB2

film with regular flux propagation and trapping patterns from a perpendicular external

field at 4.2 K. The 330 nm MgB2 film was measured by transport to have a Tc of 41.2 K

and RRR of 27. When an external magnetic field applied, the flux propagates gradually

and uniformly into the film from the edges. As the field is decreased, the flux exits the

film in the same manner. In contrast, an HPCVD C-alloyed MgB2 films show dendritic

flux jump behavior, shown in Figure 20, just like in many other MgB2 samples reported

in the literature. [91, 92] Direct magnetization measurement of pure and C-alloyed MgB2

films is shown in Figure 21. Magnetic flux instability in C-alloyed MgB2 film suppressed

the 5-K magnetization hysteresis loop and caused it to cross with the 10-K and 15-K

curves. The magnetization hysteresis loops of the pure MgB2 film developed with the

temperature with no inconsistency of crossover. These all suggest that the film

cleanliness and low resistivity enhance the low temperature magnetic field stability and

distinguish the pure HPCVD MgB2 films from other samples.

Page 63: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

48

The reason for the difference between the pure HPCVD MgB2 films, where the

dendritic magnetic instability is absent, and other MgB2 samples including C-alloyed

Figure 20 . Magneto-optical images of the zero-field-cooled pure and C-alloyed MgB2

thin film (5x5 mm ) at T = 4:2 K. The perpendicular applied field B = (a) 10 mT, (b) 20

mT, (c) 40 mT, and (d) 0 (reduced from 0.1 T), respectively. [92]

(b)(a)

(c) (d)

(f)(e)

(g) (h)

Page 64: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

49

HPCVD films, where the dendritic magnetic instability is observed, was found to be the

difference between the ratios of the magnetic diffusivity to the thermal diffusivity

Figure 21 . Magnetization curves of the ultra-pure MgB2 thin film and the carbon-doped

MgB thin film. (a) The hysteresis loop of the pure MgB2 thin film at T = 5, 10, and 15 K,

respectively. (b) The hysteresis loops of the carbon-alloyed MgB2 thin film at the same

temperatures. [92]

Page 65: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

50

[91]. Dendritic flux jumps occur when there is not enough time for the heat, generated by

the flux diffusion, to be redistributed or released. The ratio between magnetic diffusion

time constant, , and thermal diffusion time constant, , [93,94,95] is determined by

(3.3.1)

where the thermal diffusivity , and the magnetic diffusivity , and , ,

, are the superconductor’s thermal conductivity, heat capacity, flux flow resistivity,

and the permittivity in vacuum, respectively. When a superconductor has 1

(local adiabatic condition), the magnetic flux motion is much faster than thermal

diffusionand the dendritic flux jumps occur. The difference in the thermal conductivity

and heat capacity between pure and C-alloyed MgB2 samples are believed to be not as

significant [96]. Therefore, the flux flow resistivity is the key parameter for the local

adiabatic condition. The flux flow resistivity can be approximated as [97]:

(3.3.2)

where is the electrical resistivity in the normal state, and are the applied field

and the upper critical field. In conventional superconductors, often increases with the

resistivity with little change in ratio . However, the two band superconductivity in

MgB2 provides the possibility of enhancement of by introducing impurity scattering

in one band, while keep the electrical resistivity in the clean band. [98]

Page 66: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

51

The pure HPCVD MgB2 films has lowest electrical resistivity of 0.4

and comparable upper critical field 7 . These provide a much lower flux flow

resistivity of 0.006 at 0.1 T, comparing with 0.16 in C-alloyed

HPCVD film ( 40 and 25 ) and 0.1 in PLD films

( 7 and 7 ) [99]. The drastically lower flux flow resistivity in pure

HPCVD MgB2 films substantially slows magnetic flux motion, preventing the dendritic

flux jumps from forming.

3.4 Photoresponse of MgB2 Thin Bridges

Time-resolved photoimpedance measurements were done on 20 x 40 μm

microbridged HPCVD grown MgB2 structures at Rochester. [100] A large and fast (130

ps) photoresponse transient was observed at low optical excitation powers below 500 μW.

This is the kinetic inductive response due to Cooper pair breaking. The picosecond

kinetic photoresponse make MgB2 thin bridges a very attractive candidate for fast and

efficient optical detectors and photon counters.

The quick and large kinetic inductive response is originated from the Cooper pair

breaking, which can be explained by the kinetic inductive model developed by Lindgren

et al. [101,102]. The kinetic inductance of a superconducting thin bridge is

(3.4.1)

Page 67: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

52

where and are electron mass and charge, is the density of superconducting

electrons, and l, w, d are length, width, and thickness of the bridge. The absorbed photon

breaks a Cooper pair and transfer the energy to one of the electrons. Therefore,

concentration of the superconducting electrons is disturbed when Cooper pairs are

photoexcited. As a result, the kinetic inductance changes and gives an induced voltage

signal:

(3.4.2)

where is the constant current bias. The photoexcited electron will excite other electrons

via electron-electron scattering and create more quasiparticles. Once the excited

quasiparticles have energy over the energy gap, they will recombine and form pairs, by

releasing a high energy phonon ħ 2∆, at a recombination rate of . The phonon

either breaks another Cooper pair at a rate of , or escape to the substrate, or decays into

lower phonons. This dynamic process can be described using the Rothwarf-Tayor model

[103] by two coupled differential equations:

2 (3.4.3)

(3.4.4)

Page 68: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

53

where is the shape of the external excitation optical pulse, , , and

are the concentration of the quasiparticles, phonons, and equilibrium phonons, and

is the escape time. Note that the concentration of the superconducting electrons obeys

Figure 22. (a) Experimental waveform (circles) at low optical excitation with an excitation

power of 400 μW at 20 K. The solid line is a fitting with kinetic inductive response model.

(b) Experimental waveform (circles) with an excitation power of 4 mW at 20 K. The data

is fitted with a kinetic inductive response (dotted line) and a resistive response (dashed

line). The combined kinetic-inductive fit is shown as a solid line. [100]

(b)

(a)

Page 69: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

54

, where is the total density of electrons. One can use above

formulas to fit an experimental curve.

Figure 22 illustrates typical time-resolved transient photoimpedance waveforms

when microbridged MgB2 structure was optically excited with powers of 400 μW and

4mW, respectively. At a low excitation power of 400 μW, the initial main

photoimpedance peak corresponds to the Cooper pair breaking process, followed by a

negative recovery process of quasiparticle recombination. The data fitted well with the

recombination rate of 12 4 10 , Cooper pair breaking rate of 10 2 , and pulse time constant of 100 .

At a high excitation power of 4 mW, another slower response arised, with the

decay time of over 500 ps, which is attributed to be resistive response from the hot-

electron heating effect [100]. This is because at high excitation power, the microbridge

absorbs enough energy to reach above Tc, with all Cooper pairs broken. This gives an

induced voltage response [104,105]:

(3.4.5)

where is the electron temperature, is the resistivity, and is geometrical factor

depending on the bridge dimensions and bias current. can be modeled as:

∆∆ (3.4.6)

where is the residue resistivity, and and ∆ are onset transition temperature and

width of the transition. can be modeled using a two temperature model [106],

Page 70: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

55

which uses two coupled differential equations to describe the energy balance between

electron and phonon subsystems [107,108]:

(3.4.7)

(3.4.8)

Where and are the specific heat of electrons and phonons, and and are the

phonon temperature and the electron temperature. The time constant of and are

related by

(3.4.9)

Figure 23. The dependence of the photoresponse-signal rise time (a) and amplitude (b) on

absorbed optical power of a photoexcited microbridge biased at Ib = 62 mA and at T0 = 20

K. Two regimes can be identified from the plots with the transition point at the excitation

power P ≈ 500 μW. [100]

Page 71: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

56

As shown in Figure 22 (b), the photoimpedance data with powers of 4mW fitted

well with the combined responses from the kinetic inductive model and electron resistive

heating model. The decay time was obtained from the fitting to be 0.5 ns.

Figure 23 shows the incident power dependence of (a) photoresponse rise time

and (b) photoresponse amplitude of HPCVD grown MgB2 structures. The low and high

excitation ranges can be clearly distinguished from the plot. Both the rise time and

amplitude increase steadily and rapidly in the low excitation power regime. However, a

transition to the high power excitation regime characterized by a nearly constant rise time

and slowly increasing photoresponse was observed at the excitation power of

approximately 500 μW. The transition point is consistent with the value of the excitation

power where we started to observe the slow photoresponse component. This is the

transition point between the kinetic inductive photoimpedance response and the mixed

kinetic-resistive photoresponse.

3.6 Substrate, Growth Mode and Thermal Expansion Issues

SiC substrates were usually heated to ~720 ºC in 100 Torr of UHP H2 before

deposition and this is found to be a very effective way to clean the substrate surface. In

Figure 24 (a), a 1 μm x 1 μm AFM image of a 6H SiC substrate from Intrinsic

Semiconductor. The image showed extra spot features in addition to the step edges on

SiC substrate surface, even after the substrates was ultrasonically cleaned by acetone,

trichloroethylene (TCE), and isopropanol as standard procedure before deposition. After

Page 72: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

57

the heat treatment in hydrogen, the image showed clean surface without those features, as

show in the Figure 24 (b).

HPCVD MgB2 films grow in the Volmer-Weber growth mode [109], or island

growth mode, at the usual deposition temperature of 700 – 750 ºC. Figure 25 shows the

HPCVD growth process of MgB2 films. Hexagonal disconnected islands of MgB2

crystallites are initially nucleated on the substrate. As the film grows thicker, the islands

eventually coalesce. Coalescence tensile strain were observed in metallic films grown in

the Volmer-Weber mode, and the tensile strain increases with the film thickness

[110,111]. This strain is believed to be the reason for Tc enhancement of HPCVD MgB2

films. HPCVD MgB2 films have substantially higher than bulk values (~39 K), including

the best single crystal samples. The transition temperatures of the HPCVD MgB2 films

were found to increase with a lattice parameter of MgB2 and film thickness on both

Figure 24. 1 μm x 1 μm AFM images of an intrinsic 6H SiC substrate before and after heat

treatment in H2 at 720 ºC.

(b)(a)

Page 73: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

58

sapphire and SiC substrates [88,109]. This was explained by the strain-induced E2g

phonon softening, which enhances the electron-phonon interaction. The softening of the

E2g phonon was observed with Raman scattering measurement [109] and agrees well with

first-principle calculations of Tc enhancement by E2g phonon softening.

It is generally believed that layer-by-layer mode or step-flow mode could be

achieved with improved the mobility by increasing the deposition temperature. In

addition, higher deposition temperature would also enhance the crystallinity and epitaxy

with the substrates. Typically, optimal temperature for expitaxy is about half of the

Figure 25. When crystallites coalesce, they spontaneously snap together and generate a

tensile strain. [77]

Page 74: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

59

melting temperature of the material being grown [112], in the case of MgB2, ~1080 ºC.

However, thermodynamic calculation shows that, to maintain a thermodynamically stable

phase of MgB2, the growth at 1080 ºC would require 11 Torr of Mg partial pressure,

which translates into Mg flux of 2x1021 atoms/cm2 s, or 0.5 mm/s. The current HPCVD

setup certainly can not achieve this enormous Mg flux due to limited Mg supply in the

reactor.

It was found that thick HPCVD MgB2 films on SiC tend to crack, while those on

sapphire do not. This is due to the much larger mismatch in coefficients of thermal

expansion between MgB2 films and the SiC substrates. The coefficients of thermal

expansion in the (0001) plane at room temperature is 5.5x10-6 /K for MgB2 [113], 6.

7x10-6 /K for Al2O3 [114], and 3.0x10-6 /K for SiC [115]. When sample is cooled from

growth temperature of 720 ºC to room temperature, MgB2 film shrinks more than it can

sustain to be in-plane aligned with SiC substrate due to its higher coefficient of thermal

expansion. Figure 26 shows SEM images of MgB2 films on SiC and sapphire substrates

with various thicknesses. Thick MgB2 films of 900 nm on SiC deposited at 720 ºC (a)

and 700 nm on SiC deposited at 620 ºC (c) showed cracks whereas thick 1.3 μm film on

sapphire (b) showed no sign of cracks. Less significant cracks showed in film (c) than in

(a) because of lower deposition temperature. Film (a) and (c) was deposited with a small

amount N2 added to the carrier gas of H2. This was found to enhance the in plane

coalescence between MgB2 islands and to be very effective in reducing the surface

roughness, which is very important for successful Josephson junction fabrication and will

be discussed later. A 500 nm film on SiC without this additional N2 treatment showed

Page 75: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

60

poor connectivity between grains as in Figure 26 (d). Cracked films typically have much

high effective resistivity, lower RRR and broadened Tc in transport measurements. The

film shown in Figure 26 (a) has an effective resistivity of 21.4 μΩ-cm at 42 K, RRR of

2.0, and a 1.2 K wide transition with Tc(onset) of 41.0 K. The high Δρ(300 K – 42 K)

value of 22.6 μΩ-cm, about three times of ideal value of 7.5, indicating only about one

Figure 26. SEM images of MgB2 films. (a) 900 nm on SiC deposited at 720 ºC with 2%

N2 added. (b) 1.3 μm on Al2O3 deposited at 720 ºC. (c) 700 nm on SiC deposited at 620

ºC with 1.5% N2 added. (d) 500 nm film on SiC deposited at 720 ºC.

(b)(a)

(c) (d)

Page 76: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

61

third of MgB2 is connected to effectively conduct currents according to Rowell’s analysis

[116].

Even though high deposition temperature of over 700 °C enhances crystallinity

and superconducting properties, deposition at lower temperature are also desirable, for

example, for better thermal stability of barrier layer when growing multilayers for

junction devices. In addition, polycrystalline, or even amorphous, MgB2 films are also

expected to grow if deposition is at enough low temperature. Such films would

presumably have grains of both ab and c orientations and it would be very interesting to

see σ-σ and σ-π band tunneling in addition to π- π tunneling. For current HPCVD system,

the deposition temperature is limited by the fact that both magnesium and substrate are

heated by the same inductive heater, while thermodynamic phase stability requires high

magnesium vapor pressure that cannot be provided by the low Mg temperature.

3.7 MgB2 Film Morphology and Improvement for Tunnel Junctions

Because of a thin insulator thickness of only 1 – 2 nm is required for an ideal

Josephson tunnel junction, surface roughness is a critical issue. Consequently, the surface

morphology study of MgB2 thin films and an improved process to reduce surface

roughness become a crucial step toward reliable MgB2 Josephson junction fabrication.

We choose MgB2 films with thickness of around 100 nm on SiC for bottom

electrode layer for sandwiched-typed Josephson junctions. SiC substrates are preferred

for junction purpose because of its excellent lattice match resulting in superior MgB2 film

Page 77: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

62

properties. Initial AFM study suggested that MgB2 films around 1000 Å thick tend to

have lower root-mean-square (RMS) roughness. As discussed previously, thick films on

SiC tend to crack due to large mismatch in coefficient of thermal expansions and more

non-c-axis grains occur with increased film thickness. Thinner films have island

connectivity issue due to the Volmer-Weber growth mode. Upon study of 1000Å-thick

MgB2 films on different substrates, there appears to be a general trend of lower RMS

roughness on smoother SiC substrates. The best RMS roughness of around ~4 nm is

obtained on SiC substrates with RMS roughness of about 0.5 nm, the highest grade

commercially available. This roughness is much more than the thickness of 1 – 2 nm for

a typical junction barrier. Therefore, it is believed that an optimized process to further

reduce HPCVD-grown MgB2 film roughness must be established for reliable Josephson

junction fabrication.

Page 78: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

63

It was found that by adding a small amount of ultrahigh purity (UHP) nitrogen to

the hydrogen carrier gas during the deposition, the RMS roughness of the MgB2 films can

be significantly reduced. MgB2 films are regularly deposited with 5 sccm flow of a 5 at%

B2H6 in H2 gas mixture in UHP hydrogen carrier gas, and the total flow and pressure was

Figure 27. (a)-(f): 1 μm x 1 μm AFM scans of ~ 100 nm MgB2 films without and with 5,

10, 20, 30, 50, and 100 sccm N2 flow added during the deposition. The total flow and

pressure was maintained at 700 sccm and 80 Torr, respectively.

(b)(a)

(c) (d)

(f)(e)

Page 79: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

64

maintained at 700 sccm and 80 Torr, respectively. Figure 27 shows 1 μm x 1 μm AFM

scans of ~ 100 nm MgB2 films without and with 5, 10, 20, 30, 50, and 100 sccm N2 flow

added during the deposition. From the images, we found that MgB2 films with a low N2

flow of 5 – 10 sccm (~1% of the total carrier gas) added to the carrier gas have the lowest

RMS roughness of ~ 1 nm. This is a significant reduction from ~ 4nm and the RMS value

0 10 20 30 40 50 60 70 80 90 1000

1

2

3

4

5

6

7

8

9

10

RM

S R

ough

ness

(nm

)

N2 (sccm)

B2H6: 5 sccm Total: 700 sccm

0 10 20 30 40 50 60 70 80 90 100

39.2

39.4

39.6

39.8

40.0

40.2

40.4

40.6

40.8

41.0

Tc(0

) (K

)

N2 (sccm)

0 10 20 30 40 50 60 70 80 90 1000

1

2

3

4

5

6

7

8

9

10

11

12

RR

R

N2 (sccm)0 10 20 30 40 50 60 70 80 90 100

0

2

4

6

8

10

12

14

ρ 42K (μ

Ω−c

m)

N2 (sccm)

Figure 28. RMS roughness and transport properties of MgB2 films grown with 5, 10, 20,

30 50, 100 sccm flow of N2 added to the carrier gas.

Page 80: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

65

of ~ 1 nm is less than the typical insulating barrier thickness, makeing MgB2 film based

Josephson junctions more practical. The RMS roughness values and the electrical

transport properties of the series of films are shown in Figure 28. The typical RMS

roughness without N2 in carrier gas is 3.8 – 5 nm. The RMS roughness is reduced to ~ 1

nm when 5 or 10 sccm of N2 flow was added in the carrier gas. The RMS roughness

20 25 30 35 400

5

10

15

20

25

30

35

40Narrow Strip: H || ab

Res

ista

nce

(Ω)

Temperature (K)

9 T 7 T 5 T 3 T 1 T 0 T

20 25 30 35 400

5

10

15

20

Res

ista

nce

Temperature (K)

9 T 7 T 5 T 3 T 1 T 0 T

Wide Strip: H || ab

0 5 10 15 20 25 30 35 400

5

10

15

20

25

30

35

40Narrow Strip: H||c

Res

ista

nce

(Ω)

Temperature (K)

7 T 6 T 5 T 4 T 3 T 2 T 1 T

0 5 10 15 20 25 30 35 400

5

10

15

20

Res

ista

nce

(Ω)

Temperature (K)

7 T 6 T 5 T 4 T 3 T 2 T 1 T

Wide Strip: H||c

Figure 29. In magnetic field transport properties of a patterned MgB2 with 50 sccm N2

added during deposition. The patterned bridges are 15 μm or 30 μm in width and 690 μm

in length. Fields were applied parallel and perpendicular to the ab plane of MgB2.

Page 81: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

66

increases as more nitrogen flow is added. The transport superconducting properties of the

films with small flow of nitrogen added are excellent: Tc ~ 41 K, ρ ~ 1μΩ-cm, and RRR >

10, which are very close to pure MgB2 films of the same thickness. The Tc decreases with

further increase of nitrogen flow, but remain above 39 K of the bulk value even when a

large nitrogen flow of 100 sccm was added to the carrier gas. The ρ increases and RRR

decreases as well. We conclude that a low flow of 5 – 10 sccm is the optimal nitrogen

flow for maintaining good quality superconductivity and significantly reducing the film

roughness for reliable device fabrication.

The transport measurement in magnetic fields for a patterned MgB2 sample with

50 sccm N2 added to the carrier gas during deposition was also conducted using a

Quantum Design physical properties measurement system (PPMS). The patterned bridges

are 15 μm or 30 μm in width and 690 μm in length. Fields were applied parallel and

perpendicular to the ab plane of MgB2. The upper critical field Hc2 was slightly enhanced

from the pure MgB2 films but not as significant as in C-alloyed samples [117]. This also

suggests that the superconducting properties of MgB2 films with a small amount of N2

added during growth are close to those in the pure MgB2 films.

3.8 Material Stability Study of MgB2 Films and Protection

Material stability problem of MgB2 arises as samples, typically with good

crystallinity and superconducting properties, were found to be very sensitive to water and

Page 82: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

67

moisture. MgB2 films by HPCVD are noticed to degrade much faster than samples by

other techniques, possibly due to the cleaner surface. Some MgB2 samples with large

amounts of C and O impurities by other techniques show no noticeable degradation with

exposure to atmosphere for many months. Consequently, it is very important to

0 1 2 3 4 5 6 70

5

10

15

20

R/R

(0)

Time (hours)

Room Temperature

0oC

0 100 200 300 4000

1000

2000

3000

4000

Thic

knes

s (?

)

Time (minutes)

36 38 40 42 44

0.01

0.1

1

10

Res

ista

nce

(Ω)

Temperature (K)

150 min 120 min 90 min 60 min 30 min 0 min

0 50 100 15030

32

34

36

38

40

42

0

100

200

300

400

500

600

R/R

(0)

T c(0) (

K)

Time (minutes)

Figure 30 (a) Resistance vs. time curve of 2000Å-thick MgB2 films submerged in water

at room temperature and 0 °C. (b) Thickness change of a MgB2 film as a function of time

in water at room temperature. (c) Resistance vs. temperature curves of a MgB2 film after

consecutive exposures to water at room temperature. (d) Time dependence of Tc(0) and

normalized resistance summarized from the results in (c) [120]

Page 83: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

68

understand the stability of cleaner HPCVD MgB2 films under various environments and

find a way to protect the films from degradation.

There are a few reported degradation studies of MgB2 materials. Degradation of

films by e-beam evaporation and ex situ post-annealing were studies by exposing the

films to water and it was observed that Tc(0) decreases rapidly while Tc(onset) of the

films remained unchanged throughout the degradation process [ 118 ]. MgB2 bulk

degradation with exposure to ambient air was also reported [119]. We have carried out a

degradation study of clean films by HPCVD with the exposure to solvents, water,

saturated water vapor, and air, and a successful protection method was also established.

Degradation experiment of HPCVD MgB2 films was carried out by transport

measurement on films submerged in de-ionized (DI) water (ρ ≥ 18.2 MΩcm). [120]

MgB2 films were found to degrade much faster in regular water, and to degrade much

faster at room temperature than at 0°C in de-ionized water, as shown in Figure 30 (a).

The film thickness and superconducting properties were also measured as a function of

time in de-ionized water. The thickness of a MgB2 film as a function of the exposure time

in water at room temperature is plotted in Figure 30 (b). The film thickness decreased

quickly initially and then the speed of the thickness reduction decreased and the thickness

saturated at about 1200 Å. Figure 30 (c) shows a transport property after consecutive time

intervals in water. Figure 30 (d) is the extracted the dependence of Tc(0) and normalized

resistance from Figure 30 (c). The original film has a Tc of 40.5 K with a transition width

of 0.4 K, and an RRR of 11.7. The film became insulating after 180 min. in water, but

before that Tc remained relatively high. It is remarkable that after 165 min. in water when

Page 84: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

69

the resistance increased by over 500 times, the Tc of the film was still over 37 K. The

1E-3 0.01 0.1 1 10 100

1.0

1.2

1.4

1.6

1.8

2.0

Isopropanol

Acetone

Methanol

R/R

(0)

Time (hours)

Water

Figure 31. Room temperature resistance of 1000Å-thick MgB2 films as a function of time

submerged in water, methanol, acetone, and isopropanol. [120]

0 50 100 150 200 250 3000.0

0.1

0.2

0.3

Res

ista

nce

(Ω)

Temperature (K)

Figure 32. The resistance versus temperature curves of a MgB2 film with a sputtered 10

nm MgO protection layer before and after 45 hours exposure to saturated water vapor at

23˚C.

Page 85: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

70

transition width became 1.8 K and RRR became 3.0 after 150 min. Mg(OH)x, MgO, and

Mg2B2O5 are likely to be in the insulating material left after degradation of MgB2.

[121,122]

Besides water, we have also studied the effects of exposure to other fluids

necessary for the devices processing, such as methanol, acetone, isopropanol, photoresist

and developer. For photoresist model Shipley 1811, the room temperature resistance of a

1000 Å MgB2 film only increased by 1% after 2 hours’ exposure. For developer model

Shipley 351, the resistance of a film with the same thickness shows an over 10% increase

during the initial 5 minutes. The faster degradation than in the photoresist can be

explained by the existence of water in the developer. Figure 32 shows the room

temperature film resistance as a function of time submerged in methanol, acetone, and

isopropanol. For comparison, the result for water is also plotted. The thickness of all the

films is 1000 Å. It can be seen that the MgB2 films degrade much more slowly in these

solvents, in particular, in isopropanol. The resistance only increased by about 2% after

being submerged in isopropanol for 2 weeks. It demonstrates that isopropanol does not

have noticeable degrading effect on MgB2 and is a preferred solvent for MgB2. The

results suggest that for the MgB2 device processing, one needs to limit the exposure of

MgB2 films to water, photoresist, and various solvents in order to minimize the

degradation of the films.

An RF sputtered MgO layer of ~ 10 nm thick was found to be very effective in

protecting MgB2 from moisture. Figure 32 depicts the resistance versus temperature

Page 86: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

71

curves of a MgB2 film with a sputtered 10 nm MgO protection layer before and after 45

hours exposure to saturated water vapor at 23˚C.

3.9 TiB2/MgB2 Hetero-structures

Epitaxial TiB2/MgB2 heterostructures were grown by dc magnetron sputtering a

20 nm layer of TiB2 on SiC substrate and subsequently growing a 50 nm MgB2 film by

HPCVD. [158] TiB2 films were dc magnetron sputtered at 750 ˚C under an argon

pressure of 1.5 Pa. The TiB2/MgB2 heterostructures were characterized by x-ray

diffraction and TEM. Figure 33 (a) shows a selected area diffraction (SAD) TEM image

collected from the heterostructure and the Cr/Au contact layer. TiB2 and MgB2 epitaxial

c-axis layers are found to be in-plane aligned with a relationship [1 100] MgB2 [1 100]

TiB2 [1 100] SiC. Diffraction pattern from TiB2 area suggests that the TiB2 has

columlar grains with tilt and rotation. MgB2 layer showed better crystallinity overall

without following columlar grains of TiB2. Figure 33 (c) shows a high-resolution TEM

(HRTEM) image of the TiB2/MgB2 interface. The arrows marked the interface between

TiB2 and MgB2. Excellent epitaxy quality was observed around the interface. Some area

of amorphous phase due to growth or TEM sample preparation were also seen in some

interface areas.

The resistivity versus temperature curve of a TiB2 film is plotted in Figure 34.

The room temperature resistivity is about 8 μΩ-cm, which is close to 6 μΩ-cm reported

for single crystal TiB2 [123]. The resistivity increased as temperature decreasing until it

Page 87: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

72

saturated at about 290 μΩ-cm below 75 K. The same figure shows the resistance of a

MgB2/TiB2/MgB2 planar junction. From the plot, the junction resistance first went

Figure 33. (a) Bright-field TEM image of a TiB2/MgB2 heterostructure. (b) SAED pattern

collected from an area containing all the layers. (c) HRTEM image of the TiB2/SiC

substrate interface. (d) HRTEM image of the TiB2/MgB2 interface. The arrows indicate

the interface. [158]

Page 88: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

73

through a transition at 39 K, where the MgB2 electrodes is believed to be

superconducting, and it decreased further to zero at about 30 K.

Figure 34. The resistivity versus temperature curve of a TiB2 film and the resistance of a

planar MgB2-TiB2-MgB2 junction as a function of temperature. [158]

0 100 200 3000

10

20

30

0

100

200

300

Res

ista

nce (

Ω)

T (K)

Res

istiv

ity (

μΩcm

)

Page 89: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

74

Chapter 4

Superconductor-Insulator-Superconductor MgB2 Josephson Junctions

Sandwich-typed Josephson tunnel junctions (JJs), or superconductor-insulator-

superconductor (SIS) Josephson junctions, are preferred for many applications due to

their compatibility with well-established multilayer technology. The possible large IcRn

product of MgB2, implying a higher ultimate circuit operation speed, is also desirable. As

we have discussed before, the relatively high Tc of MgB2 makes it attractive for

Josephson junction applications.

SIS tunnel junctions with both electrodes being MgB2 [124,125,126], or with the

counter-electrode being a conventional superconductor [127,128,129,130], have been

reported. However, almost all of them have very small Josephson supercurrent, and some

of them do not show Josephson effect at all. In addition, the available current – voltage

characteristics from MgB2 junctions show large sub-gap current leakage, poorly defined

gap features, and low IcRn products. The widely smeared tunneling spectrum and large

current leakage also prevent good direct tunneling study of MgB2 material, the first

superconductor with two bands and two gaps with many interesting physical properties.

In this chapter, I review three processes of forming insulating barrier layers,

successful Josephson tunnel junctions based on such barrier layers, and the junction

device properties. Afterward, I will discuss spectroscopy study of two gap

superconductivity in MgB2 using SIS tunnel junctions.

Page 90: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

75

4.1 Quasiparticle Tunneling and Cooper Pair Tunneling through Insulator

It is a striking feature of quantum physics that a particle can go through a barrier

with less energy than the height. From quantum mechanics [131], the transmission

amplitude t (or an tunnel transparency) for a particle to tunnel through a square potential

barrier with a width of s and height of φ0, is (4.1.1)

where and are wave numbers outside the barrier and inside the barrier, respectively,

according to

√ ħ 0 (4.1.2)

ħ (4.1.3)

The surprising results coming out is that the probability for the particle to tunnel

through the barrier is non zero: | | (4.1.4)

Theoretically, any particle has some probability of going through some energy

barrier. As shown in Figure 35, a soccer ball can be kicked through a wall without any

damage, even though practically the possibility is infinitesimally small. The physical

reason for this small probability is that the energy barrier of the wall is high and the

soccer ball is too big a macroscopic object for this quantum mechanics phenomenon to

apply. For electrons in a conductor, even though they are microscopic particles, they still

can not go into another metal object apart far away through the high energy barrier of

ambient air. Only when the distance between the two metals is close enough can there be

Page 91: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

76

practical probability for electrons to tunnel through an insulating barrier. Both electrodes

can exchange electrons by tunneling and the actually tunneling currents in the opposite

directions cancels each other. Electrons are able to tunnel to the other side and remain

there if empty states are available on that side. Because electrons are fermions, according

to the Pauli exclusion principle, a fermion cannot reside in the quantum states which is

Figure 35. Illustration of particle tunneling through barriers. (a) Macroscopic object like

soccer ball can not tunneling through barriers like wall. (b) and (c) Microscopic particles

like electrons in metal can only if the distance between the two metals is brought close

enough to make work function low enough to provide practical tunneling probability. (d)

A bias voltage V is needed to provide empty states for electrons to tunnel through and

occupy.

Page 92: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

77

already occupied by another fermion. Therefore, for realistic tunneling current, a bias

voltage V must be applied across the thin insulator barrier. The probability of electrons

penetrating the barrier is proportional to eV.

The semi-phenomenological theory by Giaever et al. [132] is very successful to

explain the quasiparticle tunneling. The tunneling probability at a given energy is

proportional to the occupied density of state in electrode 1 and the unoccupied density of

state in electrode 2. The current tunneling from 1 to 2 is the integration over different

energies: | | (4.1.5)

where V is the bias, M is the matrix element between the states of equal energy on both

sides, and f(ϵ)is the Fermi-Dirac occupation probability: (4.1.6)

In a normal metal, the density of state is assumed to be a constant, N(0), near the Fermi

surface, while an energy gap, , exists in the density of state in a superconductor:

| | | |/√ | | | | (4.1.7)

Hence, the current tunneling from 2 to 1 is | | 1 (4.1.8)

and the net current from the quasiparticle tunneling is | | (4.1.9)

When both electrodes are normal metals,

| | (4.2.10)

Page 93: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

78

In the case of low bias (eV<<EF) and low temperature (ϵ<<kT), the relationship is the

linear Ohm’s law:

| | (4.2.11)

For a superconductor-insulator-normal metal (SIN) tunnel junction, the net tunneling

current is

| | (4.2.12)

By using the density of state in a superconductor at 0 K:

| | | | | | (4.2.13)

Similarly, for a superconductor-insulator-superconductor (SIS) tunnel junction, the net

tunneling current is

| | | | | | (4.2.14)

The quasiparticle tunneling is an incoherent process. In superconductor-

superconductor junctions, coherent transfer of electron wave function may occur if order

parameters of the two superconducting electrodes overlap. In this case, a finite

supercurrent can flow across the junctions almost like through a bulk superconductor.

The Josephson supercurrent can be calculated in a simple way. The coupling

contribution to the energy up to the lowest order can be expressed as

Page 94: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

79 ∆ ∆ ∆ ∆ 2 |∆ ∆ |cos (4.2.15)

where C is a constant and (x,y) is the barrier plane. With a magnetic field in (x,y) plane,

According to the gauge invariance, all quantities depend on ħ . We can choose

out of the barrier plane and the integration of the relation between point 1 and 2 across

the barrier along z direction is ħ . Therefore,

2 |∆ ∆ |cos ħ (4.2.16)

The variation with respect to is

ħ |∆ ∆ |sin ħ (4.2.17)

Considering the electrodynamics formula of

(4.2.18)

We have

ħ |∆ ∆ |sin ħ (4.2.19)

or

sin ħ (4.2.20)

where ħ |∆ ∆ |. In the case of no magnetic field, it becomes the Josephson

supercurrent - phase difference relationship:

Page 95: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

80

sin (4.2.20)

Figure 36 shows the schematics of the density of state of two superconductors

having different energy gaps in a SIS tunnel junction under different bias voltage and the

ideal I-V characteristics of the junctions as a result of both the quasiparticle tunneling and

Josephson tunneling.

Figure 36. Illustration of quasiparticle tunneling between two superconductors with

different energy gaps. (a) density of states of two superconductors with a bias of

(Δ1+Δ2)/e. (b) density of states of two superconductors with a bias of (Δ1+Δ2)/e. (c) a

schematic of the ideal current-voltage characteristics between two superconductors with

different energy gaps.

Page 96: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

81

4.2 Fabrication of MgB2 SIS Josephson Junctions

In the successful Josephson tunnel junction fabrications, it is a necessity to have a

deposition process for high quality superconductor films with smooth surfaces, and it is

equally important to have a reliable reproducible process of forming an insulation barrier

which covers the whole surface of the junction area. The barrier formation or growth

must be able to produce a uniform thickness of 1~2 nm without pinholes and must not

degrade Tc and gap properties of the surface layer on the superconductor. Because of so

many difficulties, a reliable process to make high quality Josephson junctions with high

temperature superconductors has not been found.

For low temperature superconductor Nb Josephson junctions, the thermal

oxidation of a 2~7 nm aluminum on the bottom electrode has been the most reliable and

widely used barrier processes. Artificial barriers formed by directly depositing an

additional insulating layer have not been very successful. Naturally, processes that can

form a uniform oxidation layer on MgB2 surface are the first approach to investigate.

Several in situ post growth processes were developed to form the insulating

barriers after depositing MgB2 by HPCVD. The high temperature post growth process

(Process A) [133] was the first successful one to make Josephson tunnel junctions with

HPCVD grown MgB2 films. The barriers were formed at a high temperature of 710 °C,

the same temperature as during the deposition. After the MgB2 growth was stopped by

switching off the boron precursor gas (B2H6 in H2), the MgB2 film was held at 710 °C

under the same environment as during the deposition (in the UHP hydrogen flow with

Page 97: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

82

high Mg vapor pressure) for 15 seconds to 3 minutes. The sample was then cooled down

to room temperature and removed from the reactor for junction fabrication.

The second barrier forming process takes place at an intermediate temperature of

200 – 450 °C (Process B). When the MgB2 film was cooled to 300 – 450 °C from the

growth temperature, the hydrogen flow was replaced by a UHP nitrogen flow. It took

about 0.5 – 1.5 minutes for the nitrogen pressure to reach the atmosphere pressure, at

which time the temperature of the film was about 220 – 320 °C. The film was then taken

out of the reactor and exposed to ambient air. We found no systematic change of barrier

properties when the temperature of replacing hydrogen by nitrogen was between 300 –

450 °C.

Figure 37. Fabrication process for sandwich type SIS Josephson tunnel junctions.

Page 98: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

83

The third process includes a prolonged active heating of film in UHP nitrogen

flow, which forms a thicker insulating barrier (Process C). When the MgB2 film was

cooled to 400 °C from the growth temperature, the hydrogen flow was replaced by a UHP

nitrogen flow and the sample was held at 400 °C in 100 Torr nitrogen for 30 minutes. The

sample was then cooled in nitrogen to room temperature and removed from the reactor.

MgB2 Josephson tunnel junctions with conventional superconductor counter

electrodes were fabricated by a large area cross-stripe process, which has been used since

the early researches in Josephson tunnel junctions. Briefly, immediately after the barrier

Figure 38. Fabrication process for sandwich-type SIS Josephson tunnel junctions starting

from a trilayer.

Substratebarrier

Cr/Au

Photo resist

MgB2

MgB2

Ion milling

SiO2

Lift offCr/Au

Photo resist Ion milling Remove PR

Page 99: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

84

was formed, the base electrode strip was defined by applying a thin layer of Duco cement

(diluted with 20 – 30 times acetone) on the film leaving a narrow stripe of uncovered

MgB2 film with a dimension of 0.5 mm or less. The top electrode stripe was then

evaporated through a shadow mask. The junction area was typically 0.1~0.3 mm2.

Josephson tunnel junctions with MgB2 as counter electrode were usually

fabricated through a photolithography based microfabrication process discussed

previously.

4.3 Electrical Properties of MgB2 SIS Josephson Junctions

Process B (exposing MgB2 surface to N2 at ~400 °C) appears to produce the most

reliable barriers with best junction properties. [134] Figure 39 (a) shows a typical four-

terminal I–V characteristic of a MgB2/I/Pb junction with barrier formed by Process B

measured at 4.4 K. The junciton was made from a c-axis oriented film. The junction has a

Josephson critical current Ic = 2.8 mA, corresponding to a Jc of approximately 3 kA/cm2.

A large hysteresis with well defined superconducting gap feature and small subgap

leakage current is observed, indicating an SIS tunnel junction behavior. The junction

normal resistance RN is about 0.63 Ω, and the IcRN value is about 1.9 mV. The IcRN

product for tunneling from c axis MgB2 to Nb was theoretically predicted to be ~2.8 meV.

[135] Because the gap values of Pb (ΔPb~1.3 meV) and Nb (ΔNb~1.4 meV) are similar,

one expects a slightly lower IcRN value in Pb junctions than in Nb junctions. However,

the IcRN products in this work are considerably smaller than 2.8 meV.

Page 100: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

85

The diffrential conductance dI/dV calculated from the I–V curve in Figure 39 (a)

is plotted in Figure 39 (b). Prominent dI/dV peaks are observed at (ΔMgB2(π)+ΔPb)/e of ~

3.3 mV and weak but discernible peaks at (ΔMgB2(π)-ΔPb)/e of ~ 0.65 mV, where ΔMgB2(π)

-200 0 2001.5

2.0

2.5

-5.0 -2.5 0.0 2.5 5.00

5

10

Con

duct

ance

(Ω−1

)

Voltage (mV)

-7

0

7

(b)

Cur

rent

(mA

)

(a)

T = 42 K

dI/d

V (Ω

-1)

V (mV)

T = 4.4 K

Figure 39. (a) I–V curve of a MgB2/insulator /Pb junction made from a c-axis oriented

MgB2 thin film taken at 4.4 K. Barrier is formed by exposing MgB2 to N2 at ~400 °C

(Process B). (b) Differential conductance dI/dV as a function of voltage V of the same

junction. [134]

Page 101: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

86

and ΔPb are the superconducting energy gaps of MgB2 (the π band) and Pb, respectively.

This leads to ΔMgB2(π) ~ 2.0 meV and ΔPb ~ 1.3 meV. For junctions in this work, ΔMgB2(π)

ranges from 1.9 ~ 2.0 meV.

Detailed properties of junctions with barriers formed by Process B at intermediate

temperature are shown in Table I. As of now, this process produces junctions with largest

Josephson critical current density and highest IcRN products for MgB2 Josephson

junctions. This process is very reproducible and junction resistance is quite consistant.

From the table, Ic ranges from 0.7 to 5.7 mA, RN from 0.2 to 0.6 Ω, and IcRN from 1.1 to

1.9 mV.

TABLE I PROCESS B BARRIER JUNCTION PROPERTIES

Tvent (oC)

RN (Ω)

Ic (mA)

IcRN (mV)

ΔMgB2(π)+ΔPb (meV)

sbarrire (nm)

φbarrier (eV)

250 0.24 5.7 1.38 3.25 - - 250 0.25 5 1.25 3.25 - -250 0.45 2.5 1.13 3.25 - -350 0.63 3 1.89 3.3 1.8 0.7350 2.2 0.7 1.55 3.3 1.7 0.68

400 0.59 2.3 1.38 3.25 1.9 0.57400 0.69 2.1 1.45 3.25 - -450 0.3 5.5 1.65 3.35 1.9 0.45 450 0.3 4.8 1.44 3.35 - -450 0.3 5.2 1.56 3.35 - -350a 14.8 - - 3.25 1.7 1.03 350a 8.6 - - 3.25 1.7 0.93a Junction made on MgB2 films on MgO (211) substrates.

Page 102: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

87

Process A (annealling MgB2 in hydrogen at ~ 700 °C) also reproducibly forms an

insulating barrier which can be used to make good Josephson junctions. Figure 40 (a)

shows a typical four-terminal I–V characteristic for a Process A barrier MgB2

/insulator/Pb junction measured at 4.3 K. The barrier was formed by holding an MgB2

film for 20 seconds at 710 °C. The junction has a critical current Ic = 1 mA,

corresponding to a Jc of approximately 1 kA/cm2. A large hysteresis with small subgap

leakage current is observed, indicating an SIS tunneling behavior. The junction normal

resistance RN is about 1.6 Ω, and the IcRN value is about 1.6 mV. There is a large scatter

in the properties for junctions with high temperature barriers in this work: Ic ranges from

0.007 to 5 mA, RN from 0.065 to 21 Ω, and IcRN from 0.14 to 1.6 mV.

TABLE II PROCESS A BARRIER JUNCTION PROPERTIES

thold (s)

RN (Ω)

Ic (mA)

IcRN (mV)

ΔMgB2(π)+ΔPb (meV)

sbarrire (nm)

φbarrier (eV)

15 0.30 4.5 1.35 3.05 2.3 0.34 15 0.42 2 0.82 3.05 2.2 0.5915 0.27 5 1.35 3.05 2.3 0.3020 1.6 1 1.6 3.3 2.1 0.4820 3.7 0.12 0.45 3.3 1.8 0.79 20 0.24 4.5 1.08 3.05 - -20 0.065 5 0.33 3.05 - -20 0.30 2 0.6 3.05 - - 25 21 0.007 0.14 3.15 2.2 0.59

180 0.72 1 0.72 3.05 2.4 0.32180 0.46 1.5 0.69 3.05 - -

Page 103: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

88

The differential conductance dI/dV calculated from the I–V curve in Figure 40 (a)

is plotted in Figure 40 (b). Two strong peaks are observed at ~ 3.3 mV and two weak but

discernible peaks are observed at ~ 0.67 mV. The strong peaks are due to the sum of the

superconducting energy gaps of MgB2 (the π band), ΔMgB2(π), and Pb, ΔPb, and the

weak peaks are due to their difference. From (ΔMgB2(π)+ΔPb)/e ~ 3.3 mV and

-7.5 -5.0 -2.5 0.0 2.5 5.0 7.50

2

Con

duct

ance

(Ω−1

)

Voltage (mV)

-4

0

4

(b)

Cur

rent

(mA)

(a)

-7.5 -5.0 -2.5 0.0 2.5 5.0 7.50

10

Con

duct

ance

(Ω−1

)

Voltage (mV)

-20

0

20

(d)

Cur

rent

(mA

)

(c)

Figure 40. (a) I–V characteristics of a MgB2/insulator/Pb junction measured at 4.3 K with

high temperature barriers by process A (annealing MgB2 film at 710 °C in H2 for 20

seconds.) (b) Differential conductance, dI/dV, as a function of voltage, V, of the same

junction in (a). (c) I–V characteristics of a MgB2/insulator/Pb junction measured at 4.3 K

with barriers by process A (annealing MgB2 film at 710 °C in H2 for 15 seconds). (d)

Differential conductance, dI/dV, as a function of voltage, V, of the same junction in (c).

An additional peak appeared at 1.3 mV, indicating normal metal regions in MgB2 barrier

interface. [133]

Page 104: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

89

(ΔMgB2(π)-ΔPb)/e ~ 0.67 mV, we found that ΔMgB2(π) ~ 2.0 meV and ΔPb ~ 1.3 meV

for the junction in Fig. 1. For junctions with high temperature barriers in this work,

ΔMgB2(π) ranges from 1.75 ~ 2.0 meV. The holding time, thold, and the properties of

junctions with high temperature barriers are summarized in Table I.

-12 -6 0 6 12-3

-2

-1

0

1

2

3

I (m

A)

V (mV)

J2 4.2 Kbar. 400C 30'Rn~4.3Ω

-12 -6 0 6 120.0

0.5

1.0

1.5

J2 4.2 Kbar. 400C 30'

dI/d

V (Ω

-1)

V (mV)

Figure 41. (a) I–V characteristics for a MgB2/insulator/Pb junction measured at 4.3 K

with a barrier by Process C. (b) Differential conductance, dI/dV, as a function of voltage,

V, of the same junction in (a).

Page 105: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

90

In one junction with high temperature barrier Process A (annealing MgB2 film at

710 °C in H2 for 15 seconds), we have also observed dI/dV peaks at 1.3 mV, as shown in

Figure 40 (d), which are due to the Pb gap. The presence of the Pb peaks suggests that

there is a non-superconducting normal metal region at the interface between the MgB2

film and the barrier and there is no proximity effect with MgB2 in this region. This kind

of peak has not been seen in any other junctions with all types of barriers. It appears that

this region can be removed by longer annealing time.

Tunnel juncitons with barrier formed by Process C (heating the MgB2 film for ~

30 min at 400 °C), were found to be not very uniform. Junction resistances ranges from 4

Ω – 200 Ω. Some junction show good tunneling charateristics, as shown in Figure 41.

Barriers were estimated to be ~2.8 nm, thicker than barriers by Process A and Process B.

4.4 XPS Study of Barrier Properties

X-ray Photoelectron Spectroscopy (XPS) studies on HPCVD MgB2 films without

or with different barriers were carried out at Cornell. Samples were scanned with 2 take-

off angles at 20 degree and 90 degree, which is the angle between the sample surface and

the detector, to study the sample to various depth. The former is more sensitive to the

surface while the latter includes more information from the film interior near the surface.

Figure 42 (a) shows the spectra of a standard HPCVD MgB2 film without any

barrier as a control sample. The film peaks are clearly distinguishable at ~ 49.5 eV for the

Mg 2p spectra and at ~ 186.7 eV for the B 1s spectra. It is important to note that the B 1s

film peak is at a lower binding energy than measured in metallic boron reference

Page 106: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

91

samples. We attribute this chemical shift to the increased negative charge on the boron

Figure 42. Surface and near surface XPS scans for (a) a HPCVD MgB2 film as an control

sample (b) a MgB2 film with Process A barrier. (c) a MgB2 film with Process B barrier.

Page 107: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

92

donated by the magnesium atoms in the MgB2 structure. The Mg and B peaks at higher

binding energies have larger intensities for the surface scans compared to the interior

scans thus we can attribute these peaks to surface species. The oxygen spectra show no

change here for interior and surface scans indicating that the oxide is present to the

maximum depth probed with O 1s photoelectrons which is roughly 60 Å. The same data

is shown with offset and with fits in Figure 43 (a). The two film peaks are clearly labeled

and we can confirm that these are the bottom most peaks by comparing peak ratios for

surface and interior scans. We attribute the peak at ~ 51.5 eV to magnesium cations in the

two plus oxidation state in the native MgB2 oxide. We also attribute the peaks at ~ 193

and ~ 189 eV to B cations in the three plus and one plus oxidation states respectively.

Since the sample has been exposed to atmosphere it is likely that the oxide layer may

include MgCO3, Mg(OH)2, B2O3, and B(OH)3 in addition to the desired MgO. We

attribute the broad O 1s peak in the spectra shown here primarily to MgB2 native oxide.

In comparison to reference samples, the remaining boron peak at ~ 188 eV has the same

binding energy as amorphous boron metal. We attribute this peak to a boron rich,

magnesium depleted Mg-B material that forms when Mg is pulled from MgB2. It is

possible that this material forms at grain boundaries. It is also possible that this peak

represents polycrystalline or disordered MgB2. If we estimate film stoichiometry using

just MgB2 peaks we yield an Mg:B ratio of roughly 1:1.3. However, if we include both

the B 1s MgB2 and Mg-B peaks we yield a better ratio of 1:1.9. We consider the spectra

shown here to be typical of MgB2 native oxide formation with additional surface carbon

contamination due to atmospheric exposure.

Page 108: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

93

The data from a HPCVD MgB2 film with barrier Process A is shown in Figure 42

Figure 43. Composition fittings of Surface and near surface XPS scans for (a) a HPCVD

MgB2 film as an control sample (b) a MgB2 film with Process A barrier. (c) a MgB2 film

with Process B barrier.

Page 109: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

94

(b) and Figure 43 (b). The sample was annealed at 700 °C for 15 s. It is clear that the

surface boron oxides are reduced in comparison to the control sample. The Mg 2p and O

1s spectra indicate the formation of an MgO species which appears to be on top of the

MgB2 film. In addition, the Mg-B material is reduced slightly in comparison to the

HPCVD control sample. The O 1s spectra display the clear presence of the MgO

material. If we estimate film stoichiometry using just MgB2 peaks we yield an Mg:B

ratio of roughly 1:1.4. However, if we include both the B 1s MgB2 and Mg-B peaks we

yield a better ratio of 1:1.8. We can also estimate the thickness of the MgO layer using

the relative intensities of the Mg metal and oxide species using standard XPS analysis

techniques. For this sample we estimate that the MgO layer is ~ 25 angstroms thick.

The data from a HPCVD MgB2 film with barrier Process B is shown in Figure 42

(c) and Figure 43 (c). The data shown here are from a HPCVD grown sample that was

exposed to ultra high purity nitrogen at a temperature of 400 °C. We believe that the

sample was oxidized by the impurity oxygen in the nitrogen gas. Comparison of the

surface and interior scans show increased peak intensity for the magnesium oxide peak

both in the Mg 2p and in the O 1s spectral regions. In addition, the boron peaks

representative of surface oxides are also increased in intensity. The peak attributable to

Mg-B material is substantially increased and we suspect this is due to increased sample

oxidation. As the sample is oxidized further, more magnesium is pulled from the film,

more MgO is formed and more Mg-B material forms as well. The film has an estimated

Mg:B ratio of 1:1.4 if we use just the film peaks and a ratio of 1:2.5 if we include the Mg-

B material peak. For this process we estimate that the MgO thickness is roughly 33

angstroms.

Page 110: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

95

4.5 Barrier Height and Thickness Estimation by Transport Measurements

Tunnel barrier height and thickness of each junction were estimated using the

Simmons model [136,137]. Considering the simplest case of a tunnel junction with

symmetric rectangular barrier when V=0. The tunneling current density from one

electrode to the other is given explicitly as

, , (4.5.1)

where J is the current density, V is the voltage applied across the tunnel barrier, and φ and

s are, respectively, the barrier height and thickness, and 4 s 2m /h, where m is the

electron mass. The intermediate voltage case has been considered by Simmon by

expanding the exponentials and dropping terms of V4 or higher: [136]: , , (4.5.2)

where

, (4.5.3)

and , (4.5.4)

A more sensitive way of studying this nonlinear current-voltage dependence is to

measure differential conductance beyond ohmic region, usually up to severaly hundreds

of millivolts, 1 3 (4.5.5)

Page 111: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

96

Where is the area of the junction. Here we have a familiar constant conductance of expected at a ohmic region and a parabolic dependence term of 3 V2. Assuming s ~

1 nm and φ ~ 1 eV, as in tunnel junctions, the second term in equation (4.5.4) can be

ignored , (4.5.6)

Combining equation (4.5.3) and (4.5.6), one has

ln (4.5.7)

-150 -100 -50 0 50 100 150-0.3

-0.2

-0.1

0.0

0.1

0.2

0.3

I (m

A)

V (mV)

1.4

1.6

1.8

2.0

2.2

dI/d

V (Ω

−1)

V (mV)

Figure 44. I-V curve of a MgB2/I/Pb junction measured at 42 K (left) to high voltages

and the differential conductance (right) of the same junction with a parabolic fit (purple).

The barrier for this junction is estimated to have a thickness of ~1.8 nm and a barrier

height of 0.7 eV.

Page 112: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

97

(4.5.8)

Once and are known, one can derive the barrier thickness, s, and barrier hieght, φ, by

solving equation (4.5.7), and (4.5.8) numerically.

Figure 44 shows a junction I-V (a) and dI/dV-V (b) curves measured slightly

above the transition temperature of MgB2. at T = 42 K. The data fits well to the Simmons

model, from which the average barrier height φ was determined to be 0.7 eV and the

barrier thickness s to be 1.8 nm in this junction. For junctions in this work, the barrier

height ranges from 0.5 to 1.0 eV and the barrier thickness from 1.7 to 1.9 nm.

4.6 Fraunhofer Pattern in SIS MgB2 Josephson Junctions

and Penetration Depth in MgB2 Films

A magnetic field applied perpendicular to the current flow direction in a

Josephson junction can cause magnetic flux to penetrate into the barrier up to a

penetration depth λ into each superconducting electrode on both sides. The total flux

penetrated is given by

(4.6.1)

where and are, respectively, the width and the thickness of the barrier, and and

are the effective London penetration depth for the two superconductors. Analysis

[138] based on Ginzburg-Landau theory shows the gauge invariant phase, , follows

Page 113: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

98

ħ (4.6.2)

The current density in the x-y plane can be rewritten using dc Josephson effect equation

as

, , , ħ (4.6.3)

Assuming the current density is uniformly distributed along the x direction, the current

across the junctions, , can be integrated to be

0 ħ⁄ħ⁄ (4.6.4)

Equation (4.6.4) can be further simplified as

0 // (4.6.5)

where is the magnetic flux penetrated into the barrier region and =πħ/e is the

magnetic flux quantum (2.07x10-15 T m2). This implies that in an ideal Josephson tunnel

junction with uniform barrier and current distribution, junction critical current is

modulated by the applied magnetic field or flux in a pattern resembling a single-slit

diffraction pattern, the Fraunhofer pattern. By measuring Ic(Φ) pattern of a junction and

comparing it with an ideal Fraunhofer pattern, information about the uniformity of the

junction current distribution can be obtained. An Ic(Φ) pattern which fits a Fraunhofer

pattern is usually used as a direct proof of a uniform junction.

The magnetic field modulation of Josephson supercurrent was observed for our

MgB2 junctions. Figure 45 shows an pattern of a MgB2-insulator-Pb junction

Page 114: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

99

measured at 4.4 K. A theoretic calculation of an ideal Fraunhofer pattern is also plotted

for comparison. The suppression of Josephson supercurrent is over 99% at the first

minima, indicating a near ideal tunneling current uniformity. Good to the second minima,

the experimental data fitted the theoretical curve well, convincingly demonstrating the dc

Josephson effect. The deviation at higher fields may be due to the irregular junction area

or imperfect field alignment.

The Josephson penetration depth is defined as

-1.5 -1.0 -0.5 0.0 0.5 1.0 1.50

2

4

6

Crit

ical

cur

rent

(mA)

Magnetic field (Guass)

Figure 45. Magnetic field dependence of Josephson critical current, Ic. The filled circles

are experimental data and the solid line is the calculated ideal Fraunhofer pattern. [134]

Page 115: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

100 (4.6.6)

where Jc(0)= Ic(0)/(wL) is the critical current density for the junction. From formula

(4.6.1) and taking into account that the first minima position of B0 corresponds to Φ0, the

Josephson penetration depth for MgB2 can be calculated to be

0.3 (4.6.7)

which is comparable to the junction size. We did not see tilted peaks in the pattern,

indicating the absence of the self-field effect in the junction or self-screening of the

supercurrent. It is consistent with the short junction (λJ≥w) behavior.

The effective London penetration depth of MgB2 can also be calculated from the

supercurrent modulation plot. From formula (30) and using λPb=46nm [139] and barrier

thickness s=2 nm as estimated previously, the effective London penetration depth of

MgB2 can be calculated to be 56 (4.6.8)

By numerically solving the penetration depth correction formula for thin films

coth (4.6.9)

the London penetration depth of MgB2 is calculated from our tunneling experiment to be

53 nm at 4.4 K. This value is in good agreement with the microwave measurement of

HPCVD MgB2 films and significantly lowers than penetration depths (over 100 nm) from

Page 116: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

101

other MgB2 samples [140]. The short penetration depth also confirms the excellent

superconductivity of MgB2 at the superconductor-barrier interface.

4.7 Spectroscopy Study of Two Bands of MgB2

Theoretical calculations regarding tunneling from two bands of MgB2 have been

carried out assuming spherical Fermi surfaces in MgB2, in which case the density of

states around the Fermi surface can be solely described by energy around the Fermi

surface. [84] In the clean case of an normal metal-insulator-superconductor junction, the

tunneling conductance along the c axis and in the ab plane is calculated to be 0.67 0.33 (4.7.1) 0.99 0.01 (4.7.2)

hence, the contribution of the σ band tunneling along the c axis is negligibly small and

even along the ab plane the σ band only contribute about one third of the tunneling

current.

Models of normal metal-insulator-superconductor tunneling were later refined by

taking into account of the specific Fermi surface shape in MgB2. [141] Because the

Fermi-velocity components in the direction normal to the barrier interface contribute, the

transport depending greatly on direcitonality [142]. A simplified case of normal metal-

insulator-superconductor tunneling is illustrated in Figure 46. Approximated Fermi

surfaces of π and σ bands [143] were used for MgB2 and the Fermi surface for the normal

metal was taken spherically. The model predicts two distinct peaks in the dI/dV versus V

Page 117: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

102

curve for both bands in tunneling direction into the ab plane of MgB2, and only a very

weak conductance peak responible for σ band in the case of tunneling into the c axis

direction.

In most MgB2/insulator/Pb thin film tunneling experiments on c axis HPCVD

MgB2 films, only features due to the π gap were observed. In some junctions, we can also

observe features from the σ gap at ~8.8 mV, as shown in Figure 48. We attribute this to

Figure 46. (a) Normal metal – MgB2 tunneling in the ab plane direction for several

interface transparencies, ranging from Z = 0 (Andreev contacts) to Z >>1 (tunnel

juctions). The barrier parameter Z is determined by the barrier potential φ and the Fermi

velocity vF by Z = φ/ħvF. (b) Normal metal – MgB2 tunneling in the c axis direction for

several interface transparencies, ranging from Z = 0 (Andreev contacts) to Z >>1 (tunnel

junctions). [142]

Page 118: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

103

be some non c axis MgB2 crystallite on the film surfaces. It is later found that features

from the σ gap and the π gap were consistently observed from junctions using films on

(211) MgO substrate using barrier process B. A schematic of crystal orientation

relationship between MgB2 and MgO (211) substrate is determined by x-ray diffraction

and illustrated in Figure 48. MgB2 grows on MgO (211) substrate with the c axis tilted

Figure 47. (a) I–V characteristics for a MgB2/insulator/Pb junction measured at 4.3 K

with barrier B formed by venting the reactor with nitrogen at 350 ºC and taking the

sample out at 280 ºC. (b) Differential conductance, dI/dV, as a function of voltage, V, of

the same junction in (a). [133]

-10 -5 0 5 100

1

Con

duct

ance

(Ω−1

)

Voltage (mV)

-3

0

3

(b)

Cur

rent

(mA)

(a)

Page 119: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

an

M

te

fe

g

w

ab

n

3

on

av

F

su

p

nd the ab pl

MgO (211) su

dI/dV-

emperatures

eatures due t

ap with ΔPb

we find that Δ

bove the Tc

ormal (SIN)

5.3 K.

First-p

n the Fermi

verage of 1.

igure 48. Le

ubstrate and

lane. Right:

lane exposed

ubstrate is sh

V-V curves fo

is shown in

to the σ gap

. This result

ΔMgB2(π) ~ 2.

of Pb, the dI

) junctions. T

principles ca

i surfaces in

8 meV, and

eft: a schema

MgB2 films

a SEM pictu

d on the film

hown in Figu

or a junction

n Figure 49.

at ~ 8.65 m

ts from the t

0 meV and

I/dV - V char

The nonlinea

alculations [

n MgB2: for

for the σ ba

atic of crysta

s. MgB2 is gr

ure of a MgB

m surface. A

ure 48.

n on a (211)

At 4.4 K, b

V are also c

tunneling in

ΔMgB2(σ) ~ 7

racteristic be

arity attribut

9] have pred

the π band

and Δ ranges

al orientation

rown with c

B2 film grow

A SEM imag

MgO substr

besides peak

clearly obser

nto the ab pl

.4 meV. As

ecomes that

ed to the gap

dicted the di

d Δ ranges f

s from 6.4 –

n relation shi

axis tilted b

wn on MgO (

ge of a MgB

rate are mea

ks due to the

rved, in both

lane of MgB

the tempera

of supercon

ps of MgB2

istributions o

from 1.2 – 3

– 7.2 meV w

ip between M

by 19.5º, exp

(211) substra

B2 film grow

sured at diff

e π gap of M

h cases at the

B2. From the

ature increas

nductor-insul

can be seen

of the gap v

3.7 meV wi

with an avera

MgO (211)

posing the a-

ate.

104

wn on

ferent

MgB2,

e sum

e data

ses to

lator-

up to

values

th an

age of

b

Page 120: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

105

6.8 meV [9]. The gap values from the MgB2/insulator/Pb junctions are in excellent

agreement with the theoretical prediction. The slightly higher ΔMgB2(σ) may be due to the

higher quality of HPCVD MgB2 materials [144]. It has been shown previously [145,146]

Figure 49. Temperature dependence of dI/dV versus V for a MgB2/insulator/Pb junction

on (211) MgO substrate. The results for temperatures higher than 4.4 K are vertically

shifted and multiplied by 5 for clarity. [134]

-15 -10 -5 0 5 10 150

1

2

3

4

5

6

7.3K

4.4K

8.3K

12.3K

20.3K30.3K

Nor

mal

ized

Con

duct

ance

Voltage (mV)

35.3K

X5

Page 121: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

106

that when defects or impurities enhance interband scattering, the π gap may increase or

stay flat and the σ gap decreases, and they eventually merge into one single gap. The fact

that the π gap is small while at the same time the σ gap is large implies that the MgB2

films are very clean and the interband scattering is weak.

Figure 50 shows temperature dependences of the gap values extracted from a

tunnel junction conductance curves. At temperature below the Tc of Pb (7.2 K), voltages

of conductance peaks are identified to be the gap values. As the temperature increases to

above 7.2 K, the junction becomes a normal metal-insulator-superconductor junction and

the dI/dV - V characteristics can be fitted using NIS tunneling discussed previously. We

use a linear combination of two gaps for current:

Figure 50. Temperature dependence of the two gaps of MgB2 from a MgB2/insulator/Pb

junction on (211) MgO substrate.

0 5 10 15 20 25 30 35 400

1

2

3

4

5

6

7

8

2.3 meV

Δσ

Δ (m

V)

T (K)

Δπ

7.4 meV

Page 122: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

107 1 (4.7.3)

and for each gap, we use a smeared distribution of the density of states in MgB2, in which

an imaginary part was added to the energy [157]:

, (4.7.4)

The temperature behavior of both gaps in the thin film tunnel junction is consistant with

experimental values from other techniques [85,147,148,149], with slightly higher σ gap

values at low temperatures.

4.8 Tunneling Study of Mixed State in MgB2

A Type II superconductor enters a mixed state or a vortex state when a magnetic

field is applied above the lower critical field, Hc1 and below the upper critical field, Hc2.

As discussed in Chapter 1, vortices are formed with normal (nonsuperconducting) vortex

cores surrounded by encircling supercurrents to maintain superconductivity outside the

vortices, as shown in Figure 2. The circulating currents confine the flux of the applied

magnetic field inside the vortices in such a way that each vortex carries exactly one

magnetic flux quantum, Φ0=πħ/e (2.07x10-15 Tm2). The vortices are usually arranged in a

periodic array: the vortex lattice. The London penetration depth λ describes the radial

extent of the circulating currents, and the coherence length ξ is roughly the dimension of

the vortex core.

MgB2 has two weakly coupled superconducting bands. This leads to a composite

structure of vortex core, which consists of concentric regions of radius ξπ and ξσ where

Page 123: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

108

the π gap and the σ gap are suppressed [150,151,152,156,153,154]. The σ band is

believed to be clean, which means the mean free path l is much larger than the coherence

length ξ, while the π band is always in the dirty limit. The dimension of the coherence

lengths can be estimated from the BCS expression ħ∆ . From our thin film

tunneling spectra: ∆σ=7.4 meV and ∆π=2.3 meV, and by using 4.4 10 /

and 5.35 10 / [135], the coherence lengths in the ab plane can be

calculated to be 12 and 50 . Similarly, we have the c direction 1.3 and 22 .

Figure 51 shows falsed color spectroscopic images of vertices in single MgB2

crystals with Hc and Hab by scanning tunneling spectroscopy (STS) [151,155]. The

vortices are elliptical with Hab with relatively low anisotropy of ~1.19 [155]. Only π

band can be tunneled into by STS with Hc, and the magnetic field dependence of the σ

gap and the π gap has not been reported. Furthermore, vortex imaging by STS has not

been achieved on HPCVD MgB2 films. This may be due to nonuniform surface oxide

coverage, as indicated by our thin film tunnel junction study.

Using thin film tunnel junction, the effects of magnetic field on both the σ gap and

the π gap can be probed by tunneling through the c axis surfaces and the area where the

ab plane is exposed. In addition, since the spectrum is the average of tunneling spectra

over the whole junction area including vortices area and bulk area between vortices, the

measurement noise is low.

To describe the vortex state in two band superconductivity with Hc, Koshelev

and Golubov have presented a model for MgB2 assuming a strong intraband scattering

Page 124: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

109

and weak impurity scattering between π band and σ band. [156] With both bands in the

dirty limit, the quasiclassical Usadel equation can be extended to describe the system of

two superconducting bands:

Figure 51. Vortices in single crystal MgB2 with Hc at 2 K. 250 x 250 nm2

spectroscopic images of a single vortex induced by an applied field of 0.05 T (a), and the

vortex lattice at 0.2 T (b). (c) Normalized zero bias conductance versus distance from the

center, for the isolated vortex shown in (a). (d) Vortices in single crystal MgB2 with H

ab at 2 K. The bars indicate zero bias conductivity (ZBC). [151,155]

(b)(a)

(c) (d)

Page 125: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

110 ∆ (4.8.1)

∆ 2 ∑ , (4.8.2)

where 1,2 is the band index, are the diffusion constants with a rationship with the

coherence length of 2 , which only applies in the dirty limit, , and are

normal and anomalous Green’s functions.

Under the assumption of weak interband scattering, the Green’s functions in π and

σ band only indirectly coupled through the self-consistency equation. Further, when

considering field along the c axis and negleting the in-plane anisotropy, a circular cell

approximation can be used. Previously, we have derived the Ginzburg-Landau parameter

κ=λ/ξ=7~15>>1, so that the magnetic field can be considered uniform when the magnetic

field is much larger than the lower critical field. Using reduced variables of length ,

temperature , and energy: , , , where r is the distance

from the vortex center. The Usadel equation and the self-consistency equation can be

rewritten as cos sin ∆ cos sin 0 (4.8.3)

∆ ∆ 2 ∑ sin ∆ ∆ (4.8.4)

∆ ∆ 2 ∑ sin ∆ ∆ (4.8.5)

where 1 , , and 2 1 . The matrix is related to coupling

constants via

(4.8.6)

Page 126: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

111

(4.8.7)

(4.8.8)

(4.8.9)

Therefore partial local density of state (DOS) can be obtained assuming analytic

continuation: , cos , (4.8.10)

From the first principle calculations [9], the coupling constants were determined to be 0.81, 0.278, 0.115, and 0.091, so that are 0.088, 2.56, 0.535, and 0.424. Once the coupling constants are fixed, the

defining parameter for the system is . Realizing that , Koschelev further

Figure 52. Superconducting gaps (a) and density of states (b) of MgB2 as a function of

distance from the center of vortex in magnetic field with D1=0.2D2. Maximum gap values

(c) and averaged density of states (d) of MgB2 as a function of applied magnetic field.

[156]

Page 127: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

112

claimed that the Hc2 in MgB2 is mainly decided by the σ band and gave a slope of Hc2

near Tc and 0 K as 1 (4.8.11)

0 0 1 (4.8.12)

By solving the Usadel equation numerically, using a very small field of

0.002 at 0.1, the spatial distribution of gaps and density of states are shown in

Figure 52 (a) and (b) for 0.2, which was found to agree with single crystal STM

data, which is the case of equal transport contribution from the two bands. One can see

that the gap of the π band is suppressed to ∆ at 3.44, while the gap of the σ

band is suppressed to ∆ at 2.15. In contrast, the quasiparticle density of states

of the π band increases quickly to half the peak value at 6.35 while the density of

states of the σ band increases much slower to half the peak value at 2. Figure 52 (c)

and (d) shows modeling results of the maximum gap values at the boundary of the vortex

unit cell as a function of applied field and the averaged density of states average over the

unit cell at the Fermi surface ( 0 . One importance feature for this is that the density

of states of the π band reaches its normal value at a much lower magnetic field than the σ

band.

Page 128: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

113

Using NIS thin film tunnel junctions, the effect of magnetic field on density of

states of MgB2 can be studied by measuring the conductance across the junciton. Because

the DOS in the normal metal can be treated as a constant near the Fermi surface, the

conductance is then depending on the DOS in MgB2. Evaporated Pb or Ag were used as

the normal metal counter electrode, as the superconductivity of Pb can be quickly

suppressed by field because of its low upper critical field. HPCVD grown c axis MgB2

films were used to study the DOS only in the π band and non c axis MgB2 were used to

study both bands at the same time.

First, we consider tunneling from a c axis MgB2 surface with Hc, in which case

only DOS in the π band can be probed. Figure 53 shows dI/dV-V curves of a

-10 0 100.00

0.05

0.10

0.15

0.20

0.25

0.30

dI/d

V (Ω

−1)

V (mV)

Figure 53. dI/dV curves of a MgB2/insulator/Ag tunnel junctions on SiC at 4.2 K with

magnetic fields of 0, 0.04, 0.16, 0.3, 0.5, 0.7, 1, 1.5, 2, 2.5, 3, 4, 5 T applied along the c

axis of MgB2.

Page 129: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

114

MgB2/insulator/Ag tunnel junctions on SiC at 4.2 K with different magnetic field applied

to the parallel to the c axis of MgB2. As the magnetic field increases, more and more

vortices are created inside the MgB2. The ZBC profile around a vortex can be modeled by

one minus the GL expression for the superconducting order parameter of the π band

[151]: , 0 1 1 ∏ tanh | | (4.8.13)

where is the normalized ZBC measured in zero field, and are the vortex positions

for a hexagonal lattice. The distance between the vortices corresponding to the magnetic

field is √ . Figure 54 (a) shows simulated ZBC profile around a vortex with

40, 50 and 60 nm in a magnetic field of 0.05 , taking into account of the

primary vortex and the 6 second closest neighbors. It can be seen that at this field,

50 nm fits the STS ZBC profile in Figure 51 C well for the single crystal MgB2

material.

The tunneling current can be modeled as the result of paralleled tunnelings, one

from normal metal into superconducting MgB2 outside of the vortices (bulk) and the

other from normal metal into vortex area. We developed a method to extract bulk ZBC

data from the conductance spectra by substract the integrated conductance over the vortex

area with a smaller field with the integrated conductance with the same area in the vortex

area with a slightly larger field, after scaling according vortex numbers. Figure 54 (c)

shows the ZBC profile around a vortex with ξπ= 35 nm in a magnetic field of H=0.05 and

0.1 T. Using this method, the conductance contribution with H=0.05 from the area within

63 nm from the vortex center will be substituted by the conductance with the same area in

Page 130: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

115

the vortex area with with H=0.1. The error for this substitution is calculated to be 1.9% .

Figure 54 (d) shows calculated errors for substituting the integrated conductance over the

vortex area with a field of H with the integrated conductance with the same area in area

the vortex with a field of 1.5H, after scaling according vortex numbers.

Figure 54. (a) Simulated ZBC profile around a vortex with ξπ= 40, 50 and 60 nm in a

magnetic field of H=0.2 T (b) Calculated magnetic field dependence of bulk ZBC with

ξπ= 30, 40 and 50 nm. (c) Simulated ZBC profile around a vortex with ξπ= 30 nm in a

magnetic field of H=0.05 and 0.1 T (d) Calculated error for substituting the integrated

conductance over the vortex area with a field of H with the integrated conductance with

the same area in area the vortex with a field of 1.5H, for with ξπ= 20, 35 and 50 nm after

scaling according vortex numbers.

1 .10 7 5 .10 8 0 5 .10 8 1 .10 70

0.2

0.4

0.6

0.8

11

0

ZBCd r .05, 40 10 9−⋅,( )ZBCd r .05, 50 10 9−⋅,( )ZBCd r .05, 60 10 9−⋅,( )

d .05( )

2

d .05( )

2−

r 0 0.5 1 1.5 20

0.5

11

0

ZBCdd hh( )

2hh, 30 10 9−⋅,⎛⎜

⎝⎞⎟⎠

ZBCdd hh( )

2hh, 40 10 9−⋅,⎛⎜

⎝⎞⎟⎠

ZBCdd hh( )

2hh, 50 10 9−⋅,⎛⎜

⎝⎞⎟⎠

20 hh

1.5 .10 7 1 .10 7 5 .10 8 0 5 .10 8 1 .10 7 1.5 .10 70

0.2

0.4

0.6

0.8

11

0

ZBCd r .05, 30 10 9−⋅,( )ZBCd r .1, 30 10 9−⋅,( )

15.5 10 8−⋅15.5− 10 8−⋅ r0 1 2 3 4 5 6

0

0.2

0.4

0.6

0.8

11

4.959 10 4−×

VortexOff hh 20 10 9−⋅,( )VortexOff hh 35 10 9−⋅,( )VortexOff hh 50 10 9−⋅,( )

60 hh

Page 131: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

116

The bulk ZBCs of a c axis MgB2 on SiC and a non c axis MgB2 on MgO (211)

with different magnetic field Hc are extracted and shown in Figure 55. It can be seen

that the extracted ZBC data from our thin film tunnel junctions follows the same trend as

the STS data, which were directly taken at bulk areas between vortices. The c axis MgB2

data can be fitted with formula (4.8.13) with ξπ=30 nm, which is lower than the coherence

length obtained through STS profile of 49.6 nm. This is consistent with the Hc2 values of

3.1 T for the STS single crystal sample and 6 T for the HPCVD c axis MgB2 film, since 1/ . The extracted bulk ZBC from the non c axis MgB2 (Hc2~4 T) on MgO

(211), therefore, have a is lower than the single crystal sample but higher than the c axis

0.0 0.1 0.2 0.3 0.4 0.5 0.6

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

ξπ=30nm

SiC: ∼100%σπ (Tc~41K)

MgO(211): 89%σπ+11%σ

σ (Tc~39K)

Single Crystal STM: ∼100%σπ (Tc~37.7K)

Nor

mal

ized

Bul

k ZB

C

Normalized Magnetic Field

Figure 55. Extracted magnetic field dependence of bulk ZBC from thin film NIS tunnel

junctions and comparison with STS direction measurement with a fitting simulated for a

vortex core size of 30 nm.

Page 132: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

117

MgB2 film at low magnetic fields. Since the tunneling has ~11% conductance

contribution from σ band as we discussed before and this leads to lower ZBC at higher

fields where the π band superconductivity is presumably largly suppressed.

Figure 56 shows the diffrential conductance curves of a MgB2/insulator/Pb tunnel

junctions with different magnetic field applied parallel to the c axis of the MgB2 film at

4.2 K. The junction was made on MgO (211) substrate, which shows well defined

conductance spectrum with tunneling from both the π band and the σ band. One can see

that the π band gap feature is suppressed quickly by magnetic field while the σ band

feature survive to high magnetic field. Conductance curves were fitted with a linear

-20 -15 -10 -5 0 5 10 15 20 250.2

0.3

0.4

0.5

0.6

3421.5

1

0.7

0.5

0.3

0.16

G (S

)

V (mV)

B (T)

Pb/I/MgB2 (211) MgO

B // c T = 4.2K

0.08

-20 -15 -10 -5 0 5 10 15 20 25

0.2

0.3

0.4

0.5

0.6

87

65

4321.51

0.5

0.3

B (T)

Pb/I/MgB2 (211) MgOB // abT = 4.2K

G (S

)V (mV)

0.16

Figure 56. dI/dV curves of a MgB2/insulator/Pb tunnel junctions on MgO (211) at 4.2 K

with different magnetic field applied to the parallel to the c and ab axis of MgB2.

Page 133: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

118

combination of 2 gap contribution, with each gap smeared around the peak value. The

smearing facter was introduced to describe quasiparticle life-broadening [157], but here it

is used only to describe the broadening of the gap features. The fitted curves and the

extracted gap values as a function of the applied field is shown in Figure 57. The

extracted gap values clearly show that the π gap is suppressed much faster with

increasing magnetic field. This is consistent with the two band dirty limit vortex theory

[156], and also suggest at high magnetic fields the superconductivity in the π band is

induced from the σ band.

Figure 57. dI/dV curves in magnetic field with fitting and the extracted gap values as a

function of the applied field.

0

1

2

3

4

5

6

7

8

9

Δ (m

V)

Δ(π) (B//ab) Δ(σ) (B//ab) Δ(π) (B//c) Δ(σ) (B//c)

Page 134: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

119

Chapter 5

Planar MgB2 Josephson Junctions and Circuits

In this chapter, I discuss planar all-MgB2 Josephson junctions made by creating a

weak-link through TiB2 underlayer or ion damaged. Junctions exhibited Josephson

critical current and RSJ-like characteristics and Shapiro steps under microwave radiation.

Uniform ion damage MgB2 Josephson junction array was also demonstrated.

5.1 Planar MgB2-TiB2-MgB2 SNS Josephson Junctions

Figure 58 shows a schematic structure of planar MgB2-TiB2-MgB2 junctions

[158]. The growth of TiB2/MgB2 heterostructures has been described previously. After a

Cr (5 nm)/Au (150 nm) contact layer was deposited by dc magnetron sputtering, thin

bridges of 1-4 μm wide were patterned by contact lithography and ion milling.

Nanofabrication techniques of either electron beam lithography or focused ion beam were

used to etch a 50 nm slit on the bridge down to the TiB2 layer. In the e-beam lighographic

approach, a trilayer of photoresist S1808 (800 nm)/Ge (25 nm)/polymethyl methacrylate

(PMMA, 100 nm) was used as the mask. [159] Reactive ion etching (RIE) with CCl2F2

and O2 was used to transfer the e-beam pattern on PMMA to S1808. Finally, Ar ion

milling at 270 eV and 0.7 mA/cm2 with normal incidence on a liquid nitrogen-cooled

stage was used to make the defining etch for the gap in the MgB2 film. In the FIB

approach, MgB2 was etched by a focused Ga ion beam at 30 keV and 10 pA using an FEI

Page 135: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

120

Quanta 200 3D FIB system directly. The etching depth control was monitored by the

stage current. No systematic difference of junction properties was observed between the 2

fabrication processes.

The current-voltage characteritics at T = 5, 15, 24, and 31 K of a MgB2-TiB2-

MgB2 SNS junction are shown in Figure 59. They can be well fitted with the resistively

shunted junction (RSJ) model. The temperature dependence of the Josephson

supercurrent (Ic) and the junction normal resistance (Rn) are shown in Figure 60. junction

normal resistance roughly stays constant due to the flat resistivity of TiB2 in the

temperature range. The temperature dependence of IcRn fits very well to Likharev’s rigid

boundary condition model for the dirty normal metal proximity effect, [160]

1 (5.1.1)

Figure 58. A schematic structure and a SEM picture of the planar SNS MgB2-TiB2-MgB2

Josephson junctions. [158]

Page 136: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

121

where L is the dimension of the gap, and ξn is the coherence length in the TiB2 layer. The

Figure 59. I-V characteristics of a MgB2/TiB2/MgB2 junction at 5, 15, 24, and 31 K. [158]

Figure 60. Temperature dependence of Ic (squares) and fit (solid line), and Rn (dashed

line). [158]

-2

-1

0

1

2

-3 -2 -1 0 1 2 3

34K 24K 15K 5K

Voltage (mV)

Cur

rent

(mA

)

Page 137: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

122

fitting from the curve gives 2.6. Using de Gennes’ dirty metal proximity

-1.0

-0.5

0.0

0.5

1.0

-0.4 -0.2 0.0 0.2 0.4

-2 dBm-9 dBm

Voltage (mV)

Cur

rent

(mA

)

(a)

no RF

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.40.0

0.5

VRF / IcRn

(b)

0.0

0.5

1.0

n = 2

n = 1

n = 0 29.5 GHzT = 28 K

0.0

0.5

I step

/ I c

Figure 61. (a) I-V characteristics of an MgB2/TiB2/MgB2 junction with and without

applied 29.5 GHz microwave radiation at 28 K. (b) Microwave voltage dependences of

the Josephson supercurrent and the first and second Shapiro step heights (squares) with a

simulated fit (lines). [158]

Page 138: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

123

effect model,

ħ (5.1.2)

where vF is the Fermi velocity and ln is the mean free path in the normal metal, [161]

was calculated to be 3.9, assuming ~ 3.8 nm at 30 K for a resistivity of 290 μΩ-

cm, and L ~ 15 nm determined from SEM images. Considering the spatial accuracy of

SEM, it is a good agreement with the data. The coupling between the two

superconducting electrodes are believed to be the smaller π gap of MgB2 (Δπ(0) = 1.8

meV [9]), by fitting the IcRn(T) with the Likharev’s model. This can be explained by the

current flow direction at the MgB2/TiB2 interface, which is perpendicular to the 2-

dimensional conduction band of the larger σ gap in MgB2.

Figure 61 shows an I-V characteristic of a MgB2/TiB2/MgB2 junction with and

without 29.5 GHz microwave radiation of different powers. The junction clearly exhibits

the ac Josephson effect with Shapiro steps at voltages Vn=nhf/2e, where f is the

microwave frequency, h is the plank constant, and n = 0, 1, 2, … is the order of the steps.

The step-height as a function of microwave voltage across the junction is plotted for the

Josephson current, the 1st, and 2nd order Shapiro steps in Figure 61 (b). The data fit well

with Bessel functions, as predicted for the ac Josephson effect. [162]

Page 139: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

124

Figure 62 shows a Josephson supercurrent modulation of a MgB2/TiB2/MgB2

junction by an external magnetic field applied normal to the film surface. The critical

current data was taken at 28 K using both increasing and decreasing field magnitude. A

hysteresis due to the flux penetration into the electrodes, [163] commonly seen in planar

junctions, is observed. The Josephson supercurrent was only partially suppressed (~ 90%

of the maximum value) at the minima, instead of complete critical current suppression as

in an ideal Fraunhofer pattern. Also, the second order and third order peaks were much

larger than in a successful modulation.

Figure 62. Josephson supercurrent modulation of a MgB2/TiB2/MgB2 junction at 28 K

with both increasing (open squares) and decreasing (solid squares) field. [158]

-40 -20 0 20 400.0

0.1

0.2

0.3

0.4

0.5B Inc. B Dec.

Crit

ical

cur

rent

(mA)

Applied magnetic field (Gauss)

Page 140: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

125

5.2 Ion Damage MgB2 Josephson Junctions and Series Array

Planar MgB2 Josephson junctions and 20-junction series arrays were fabricated

using ion damaged MgB2 as weak-link between superconducting electrodes. [164] First 4

μm wide HPCVD MgB2 (100 nm, with 200 nm Cr/Au on top for contact pads) bridges

were fabricated using optical lithography. A 80 nm gap of photoresist S1808 pattern was

fabricated on each bridge through the same S1808/Ge/PMMA e-beam process as in

MgB2/TiB2/MgB2 junctions. An ion damaged MgB2 region at the gap was formed by

using 200-keV Ne+ ion implantation.

RSJ-like I-V characteristics were observed at 34–38 K for such ion damaged

MgB2 junctions. Figure 63 (a) shows I-V measurements for a single junction at 37.2 K,

with and without 12 GHz microwave radiation. The IcRn product is 75 μV and the normal

state resistance is about 0.1 Ω. Shapiro steps are visible under microwave radiation at the

expected voltages of Vn=nhf/2e. The step-heights versus microwave power for 0 and 1 are

shown in Figure 63 (b). The step-heights had Bessel-like dependence on RF powers,

indicating a good ac Josephson effect. The temperature dependence of the Josephson

supercurrent was shown in Figure 63 (c). Similar to ion damage YBCO junctions [165],

the interface between MgB2 and ion damaged MgB2 is not well defined and will spatially

move as temperature or bias current changes. This is indicated in the temperature

dependence close to Tc, as shown in Figure 63 (d). In the low

temperature range, , which can be described by de Gennes’ model for

superconductor-normal metal junctions with fixed interface. Junction resistance showed

that Tc of MgB2 and ion damaged MgB2 were, respectively, 38.8 K and 38.2 K.

Page 141: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

126

A 20-junction series array was successfully made and the I-V characteristics with

and without 12 GHz microwave radiation at 37.5 K is shown in Figure 64 (a). A flat giant

Shapiro step is observed at 20 times the value of a single junction. This suggests good

junction uniformity with a small spread in IcRn. dV/dI for the same array is shown in

Figure 63. (a) I-V characteristics for a single junction at 37.2 K, with and without 12 GHz

microwave radiation. (b) Microwave power dependence of the Josephson supercurrent

and first-order Shapiro steps. (c) Junction critical current (circles) and resistance

(triangles) versus temperature. The dashed and solid lines are fits with

and , respectively. (d) critical current versus temperature near Tc.

[164]

Page 142: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

127

Figure 64 (b). Differential resistance reaching zero confirms that the step is flat and all

junctions are locked to the 12 GHz drive signal.

Figure 64. (a) I-V characteristics of a 20-junction array at 37.5 K, with and without 12

GHz microwave radiation. The inset is a SEM image of an ion implantation mask after

etching used to create a multijunction array. (b) dV/dI vs V for the array. [164]

Page 143: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

128

Chapter 6

Conclusions and Future Plan

The hybrid physical-chemical vapor deposition (HPCVD) technique can produce

clean epitaxial MgB2 films. High Mg partial pressure in HPCVD maintains the

thermodynamic stability and excellent stoichiometry. The reducing hydrogen

environment prevents oxidation during the deposition and the high purity Mg and B

sources prevent other impurities. A relatively high deposition temperature of ~700 °C

enhances the film crystallinity. As a result, HPCVD MgB2 films have high transition

temperature, sharp transition and low residue resistivity. Magneto-optical imaging study

shows pure HPCVD MgB2 films are free of dendritic magnetic instability at low

temperature due to their low flux flow resistivity. HPCVD MgB2 films also have long

mean free path and short penetration depth.

Under current deposition conditions, MgB2 films grow in the Volmer-Weber

growth mode because of low mobility due to the high deposition rate and (~10 Å/s).

Higher deposition temperature and/or lower deposition rate could make layer-by-layer or

even step-flow growth mode possible, possibly yielding better crystallinity with smaller

full width at half maxima (FWHM) in φ and ω scans. It is also possible that at higher

deposition temperature the tensile strain in the film could become even higher because of

the larger effect of the mismatch in the coefficients of thermal expansion between the

film and the SiC substrate, and the Tc could be further enhanced. The study of different

growth modes for MgB2 film is of great interest.

Page 144: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

129

The Volmer-Weber growth mode in HPCVD MgB2 growth also introduces rough

surfaces. Nevertheless, this problem can be alleviated by adding ~1% of nitrogen into the

hydrogen carrier gas during the growth, which can enhance the ab plane connectivity

between the islands. The RMS surface roughness of c-axis HPCVD MgB2 films can be

reduced to ~1 nm from ~4 nm, with excellent superconducting properties. Roughness due

to off c-axis grains at lower deposition temperatures or on TiB2 buffer layer can not be

reduced by this method. The smoothness enhancement mechanism still needs further

investigation.

By using post growth barrier formation techniques, superconductor-insulator-

superconductor Josephson tunnel junctions have been made with excellent tunneling

characteristics, large Josephson supercurrent, and large IcRn products. Junctions show

well-defined gaps and low subgap currents. Fraunhofer pattern of the Josephson

supercurrent modulation in magnetic field demonstrates excellent junction uniformity.

The barrier thickness and height have been estimated and the MgOx formation has been

confirmed by XPS study. Near ideal tunneling properties, low π gap and large σ gap

values, and short penetration depth confirm that the superconductivity of MgB2 at the

barrier interface is excellent. However, the excellent barrier interfaces could not survive

at high temperature in the hydrogen environment during the second HPCVD MgB2

deposition. A new HPCVD process with lower deposition temperature or a new barrier

approach is needed in order to achieve an all-MgB2 Josephson tunnel junction technology.

Both the π gap and the σ gap have been observed using Josephson tunnel

junctions with non c-axis oriented MgB2 films. The two-band superconductivity and the

vortex state have also been studied by tunneling spectroscopy in magnetic fields. It would

Page 145: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

130

be very interesting to do STM vortex imaging on HPCVD MgB2 films. Combining STS

study of local density of state and averaged tunneling on a large area thin film tunnel

junction could lead to better understanding of the two band superconductivity in magnetic

field in MgB2.

Planar all-MgB2 Josephson junctions were also demonstrated by weak-links

through TiB2 underlayer or ion damaged MgB2. Junctions exhibited RSJ-like

characteristics, Josephson supercurrent, and Shapiro steps under microwave radiation. Ion

damage MgB2 Josephson junction array show the excellent uniformity of 20 junctions.

However, reproducibility and controllability remain challenging.

A new in situ HPCVD system with independent control of substrate temperature

and Mg source temperature is expected to remove the limitation on deposition

temperature and deposition time. Lower growth temperature and thick films have become

possible in the new system. The deposition process still needs to be further developed and

optimized so that an artificial barrier can be deposited in situ and trilayer junctions can be

fabricated completely in a controlled environment. This could become a viable approach

for MgB2 Josephson junction and circuit technologies.

Page 146: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

131

REFERENCES

1 W. Chen, V. Patel, A. Rylyakov, and K. Likharev. “Rapid single-flux-quantum T flip-

flop operating at 770 GHz,” IEEE Trans. on Appl. Supercond. 3212-3215 (1999).

2 K.K. Likharev and V.K. Semenov, “RSFQ logic/memory family: A new Josephson-

junction technology for sub-terahertz-clock-frequency digital system," IEEE. Trans.

Applied Superconductivity, vol. 1, no.1, pp. 3-28, March 1991.

3 H. J. M. ter Brake and G. F. M. Wiegerinck, “Low-power cryocooler survey,”

Cryogenics, vol. 42, pp. 705-718, 2003.

4 H. J. M. ter Brake, F.-Im. Buchholz, G. Burnell, T. Claeson, D. Cre´te´, P. Febvre, G. J.

Gerritsma, H. Hilgenkamp, R. Humphreys, Z. Ivanov, W. Jutzi, M. I. Khabipov, J.

Mannhart, H.-G. Meyer, J. Niemeyer, A. Ravex, H. Rogalla, M. Russo, J. Satchell, M.

Siegel, H. To¨pfer, F. H. Uhlmann, J.-C. Ville´gier, E. Wikborg, D. Winkler, A. B. Zorin,

“SCENET roadmap for superconductor digital electronics,” Physica C, vol. 439, pp. 1-41,

2006.

5 J.G. Bednorz and K.A. Mueller (1986). Z. Phys. B64 (189).

6 M. K. Wu, J. R. Ashburn, C. J. Torng, P. H. Hor, R. L. Meng, L. Gao, Z. J. Huang, Y. Q.

Wang, and C. W. Chu (1987). "Superconductivity at 93 K in a New Mixed-Phase Y-Ba-

Cu-O Compound System at Ambient Pressure". Physical Review Letters 58: 908–910.

7 J. Nagamatsu, N. Nakagawa, T. Muranaka, Y. Zenitani, and J. Akimitsu,

“Superconductivity at 39K in magnesium diboride,” Nature, vol. 410, pp. 63-64, Mar.

2001.

Page 147: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

132

8 S. L. Bud’ko, G. Lapertot, C. Petrovic, C. E. Cunningham, N. Anderson, P. C. Canfield,

“Boron isotope effect in superconducting MgB2”, Phys. Rev. Lett. 86 (2001) 1877-1880.

9 H. J. Choi, D. Roundy, H. Sun, M. L. Cohen, and S. G. Louie, “The origin of the

anomalous superconducting properties of MgB2,” Nature (London), 418, 758-760, (2002).

10 A. Y. Liu, I. I. Mazin, and J. Kortus, “Beyond Eliashberg Superconductivity in MgB2:

Anharmonicity, Two-Phonon Scattering, and Multiple Gaps,” Phys. Rev. Lett. 87,

087005, (2001).

11 H. K. Onnes, Commun. Phys. Lab. 12, 120 (1911).

12 Basic Research Needs for Superconductivity, Report on the Basic Energy Sciences

Workshop on Superconductivity, 2006,

http://www.sc.doe.gov/bes/reports/files/SC_rpt.pdf.

13 W. Meissner and R. Oschenfeld, Naturwiss. 21, 787 (1933).

14 http://en.wikipedia.org/wiki/Meissner_effect.

15 V.L. Ginzburg and L.D. Landau, Zh. Eksp. Teor. Fiz. 20, 1064 (1950).

16 J. Bardeen, L. N. Cooper, and J. R. Schrieffer, "Theory of Superconductivity", Phys.

Rev. 108 (5), 1175 (1957).

17 Soviet Physics JETP 5, 1174 (1957)

18 C. Kittel, “Introduction to Solid State Physics," 7th Ed., John Wiley & Sons, Inc., 1996.

19 R. C. Dynes, and J. M. Rowell, Phys. Rev. B 11, 1884 (1975).

20 W. Pickett, Mind the double gap, Nature (London) 418 (2002) 733-734.

Page 148: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

133

21 Q. Li, B. T. Liu, Y. F. Hu, J. Chen, H. Gao, L. Shan, H. H. Wen, A. Pogrebnyakov, J.

M. Redwing, X. X. Xi, Large anisotropic normal-state magnetoresistance in clean MgB2

thin films, Phys. Rev. Lett. 96 (2006) 167003.

22 A. Gurevich, V. Vinokur, Interband phase modes and nonequilibrium soliton structures

in two-gap superconductors, Phys. Rev. Lett. 90 (2003) 047004.

23 A. Gurevich, V. Vinokur, Phase textures induced by dc-current pair breaking in weakly

coupled multilayer structures and two-gap superconductors, Phys. Rev. Lett. 97 (2006)

137003.

24 S. G. Sharapov, V. Gusynin, H. Beck, Effective action approach to the Leggetts mode

in two-band superconductors, Eur. Phys. J. B 30 (2002) 45-51.

25 D. F. Agterberg, E. Demler, B. Janko, Josephson effects between multigap and single-

gap superconductors, Phys. Rev. B 66 (2002) 214507.

26 H. Suhl, B.T. Matthias, and L. R. Walker, Phys. Rev. Lett. 3, 552 (1959).

27 V. A. Moskalenko, Fiz. Met. Metalloved. 4, 503 (1959).

28 G. Binnig, A. Baratoff, H.E. Hoenig, and J.G. Bednorz, Phys. Rev. Lett. 45, 1352

(1980).

29 M. Iavarone, G. Karapetrov, A. E. Koshelev, W. K. Kwok, G. W. Crabtree, D. G.

Hinks, W. N. Kang, E. M. Choi, H. J. Kim, H. J. Kim, and S. I. Lee, Phys. Rev. Lett. 89

187002 (2002).

30 F. Giubileo, D. Roditchev, W. Sacks, R. Lamy, D.X. Thanh, J. Klein, S. Miraglia, D.

Fruchart, J. Marcus, and Ph. Monod, Phys. Rev. Lett. 87, 177008 (2001).

Page 149: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

134

31 Y. Bugoslavsky, Y. Miyoshi, G.K. Perkins, A.V. Berenov, Z. Lockman, J.L.

MacManus-Driscoll, L.F. Cohen, A.D. Caplin, H.Y. Zhai, M.P. Paranthaman, H.M.

Christen, and M. Blamire, Supercond. Sci. Technol. 15, 526 (2002).

32 P. Szabo´, P. Samuely, J. Kacmarcı´k, T. Klein, J. Marcus, D. Fruchart, S. Miraglia, C.

Marcenat, and A.G.M. Jansen, Phys. Rev. Lett. 87, 137005 (2001).

33 S. Tsuda, T. Yokoya, T. Kiss, Y. Takano, K. Togano, H. Kito, H. Ihara, and S. Shin,

Phys. Rev. Lett. 87, 177006 (2001).

34 Y. Wang, T. Plackowski, and A. Junod, Physica C 355, 179 (2001).

35 F. Bouquet, R.A. Fisher, N.E. Phillips, D.G. Hinks, and J.D. Jorgensen, Phys. Rev. Lett.

87, 047001 ~2001!

36 X.K. Chen, M.J. Konstantinovic´, J.C. Irwin, D.D. Lawrie, and J.P. Franck, Phys. Rev.

Lett. 87, 157002 (2001).

37 J. Kortus et al., Phys. Rev. Lett. 86, 4656 ~2001!.

38 A. A. Golubov et al., J. Phys.: Condens. Matter. 14, 1353 (2002).

39 P. C. Canfield and G. W. Crabtree, Phys. Today 56, 34 (2003).

40 D. Mijatovic, A. Brinkman, D. Veldhuis, H. Hilgenkamp, H. Rogalla, G. Rijnders, D.

H. A. Blank, A. V. Pogrebnyakov, J. M. Redwing, S. Y. Xu, Q. Li, X. X. Xi, SQUID

magnetometer operating at 37 K based on nanobridges in epitaxial MgB2 thin films, Appl.

Phys. Lett. 87 (2005) 192505.

41 R. Vaglio, Alternative superconducting materials for RF cavity applications, Particle

Accelerators 61 (1998) 391-400.

Page 150: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

135

42 E. W. Collings, M. D. Sumption, T. Tajima, Magnesium diboride superconducting RF

resonant cavities for high energy particle acceleration, Supercond. Sci. Technol. 17 (2004)

S595-S601.

43 D. Dimos, P. Chaudhari, J. Mannhart and F. K. LeGoues, Phys. Rev. Lett. 61,219

(1988).

44 A. Gurevich et al., Supercond. Sci. and Technol. 17, 278 (2004).

45 A. V. Pogrebnyakov, X. X. Xi, J. M. Redwing, V. Vaithyanathan, D. G. Schlom, and A.

Soukiassian, S. B. Mi and C. L. Jia, J. E. Giencke and C. B. Eom, J. Chen, Y. F. Hu, Y.

Cui, and Qi Li, “Properties of MgB2 thin films with carbon doping,” Appl. Phys. Lett. 85,

2017 (2004).

46 Y. Iwasa, D. C. Larbalestier, M. Okada, R. Penco, M. Sumption, X. Xi, A round table

discussion on MgB2: Towards a wide market or a niche production?, IEEE Trans. on

Appl. Supercond. 16 (2006) 1457-1464.

47 J. Bascu~n¶an, H. Lee, E. S. Bobrov, S. Hahn, Y. Iwasa, M. Tomsic, M. Rind°eisch, A

0.6 T/650 mm RT bore solid nitrogen cooled MgB2 demonstration coil for MRI - a status

report, IEEE Trans. Appl. Supercond. 16, 1427-1430 (2006).

48 B.D. Josephson, “Possible new effects in superconductive tunneling," Phys. Lett., vol.1,

pp.251-253, July 1962.

49 B. D. Josephson. “The discovery of tunnelling supercurrents,” Rev. Mod. Phys. 1974;

46(2): 251-254.

50 Barone A, Paterno G. Physics and Applications of the Josephson Effect. New York:

John Wiley & Sons; 1982.

Page 151: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

136

51 http://physics.nist.gov/cuu/Constants/.

52 J.M. Rowell, M. Gurvitch, and J. Geerk, “Modification of tunneling barriers on Nb by

a few monolayers of Al," Phys. Rev. B, vol. 24, pp. 2278-2281, 1981.

53 D.K. Brock, E.K. Track, and J.M. Rowell, “Superconductor ICs: the 100-GHz second

generation," IEEE Spectrum, December 2000.

54 X. X. Xi, A. V. Pogrebnyakov, X. H. Zeng, J. M. Redwing, S. Y. Xu, Q. Li, Z.-K. Liu,

J. Lettieri, V. Vaithyanathan, D. G. Schlom, H. M. Christen, H. Y. Zhai, A. Goyal,

“Progress in the deposition of MgB2 thin films,” Supercond. Sci. Technol. 17 (2004)

S196-S201.

55 W. N. Kang, H.-J. Kim, E.-M. Choi, C. U. Jung, S.-I. Lee, Superconducting MgB2

thin °ms with a transition temperature of 39 Kelvin, Science 292 (2001) 1521-1523.

56 C. B. Eom, M. K. Lee, J. H. Choi, L. Belenky, X. Song, L. D. Cooley, M. T. Naus, S.

Patnaik, J. Jiang, M. O. Rikel, A. A. Polyanskii, A. Gurevich, X. Y. Cai, S. D. Bu, S. E.

Babcock, E. E. Hellstrom, D. C. Larbalestier, N. Rogado, K. A. Regan, M. A. Hayward,

T. He, J. S. Slusky, K. Inumaru, M. Haas, R. J. Cava, High critical current density and

enhanced irreversibility field in superconducting MgB2 thin films, Nature (London) 411

(2001) 558-560.

57 C. Ferdeghini, V. Ferrando, G. Grassano, W. Ramadan, E. Bellingeri, V. Braccini, D.

Marr¶e, P. Manfrinetti, A. Palenzona, F. Borgatti, R. Felici, T.-L. Lee, Growth of c-

oriented MgB2 thin films by pulsed laser deposition: structural characterization and

electronic anisotropy, Supercond. Sci. Technol. 14 (2001) 952-957.

Page 152: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

137

58 A. Berenov, Z. Lockman, X. Qi, J. L. MacManus-Driscoll, Y. Bugoslavsky, L. F.

Cohen, M.-H. Jo, N. A. Stelmashenko, V. N. Tsaneva, M. Kambara, N. H. Babu, D. A.

Cardwell, M. G. Blamire, Growth of strongly biaxially aligned MgB2 thin films on

sapphire by postannealing of amorphous precursors, Appl. Phys. Lett. 79 (2001) 4001-

4003.

59 R. Vaglio, M. G. Maglione, R. Di Capua, High-quality MgB2 thin films in situ grown

by dc magnetron sputtering, Supercond. Sci. Technol. 15 (2002) 1236-1239.

60 S. H. Moon, J. H. Yun, H. N. Lee, J. I. Kye, H. G. Kim, W. Chung, B. Oh, High critical

current densities in superconducting MgB2 thin films, Appl. Phys. Lett. 79 (2001) 2429-

2431.

61 X. H. Fu, D. S. Wang, Z. P. Zhang, J. Yang, Superconducting MgB2 thin films

prepared by chemical vapor deposition from diborane, Physica C 377 (2002) 407-410.

62 D. H. A. Blank, H. Hilgenkamp, A. Brinkman, D. Mijatovic, G. Rijnders, H. Rogalla,

Superconducting Mg-B films by pulsed laser deposition in an in-situ two-step process

using multi-component targets, Appl. Phys. Lett. 79 (2001) 394-396.

63 S. R. Shinde, S. B. Ogale, R. L. Greene, T. Venkatesan, P. C. Canfield, S. Bud0ko, G.

Lapertot, C. Petrovic, Superconducting MgB2 thin films by pulsed laser deposition, Appl.

Phys. Lett. 79 (2001) 227-229.

64 H. Christen, H. Zhai, C. Cantoni, M. Paranthaman, B. Sales, C. Rouleau, D. Norton, D.

Christen, D. Lowndes, Superconducting magnesium diboride films with Tc ¼ 24K grown

by pulsed laser deposition with in-situ anneal, Physica C 353 (2001) 157-161.

Page 153: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

138

65 X. H. Zeng, A. Sukiasyan, X. X. Xi, Y. F. Hu, E. Wertz, Q. Li, W. Tian, H. P. Sun, X.

Q. Pan, J. Lettieri, D. G. Schlom, C. O. Brubaker, Z.-K. Liu, Q. Li, Superconducting

properties of nanocrystalline MgB2 thin films made by an in situ annealing process, Appl.

Phys. Lett. 79 (2001) 1840-1842.

66 S. N. Ermolov, M. V. Indenbom, A. N. Rossolenko, I. K. Bdikin, L. S. Uspenskaya, N.

S. Stepakov, V. G. Glebovskii, Superconducting MgB2 films obtained by magnetron

sputtering, JETP Lett. 73 (2001) 557-561.

67 A. Plecenik, L. Satrapinsky, P. K¶u·s, ·S. Ga·zi, ·S. Be·na·cka, I. V¶avra, I. Kosti·c,

MgB2 superconducting thin films on Si and Al2O3 substrates, Physica C 363 (2001) 224-

230.

68 J. Kim, R. K. Singh, N. Newman, J. M. Rowell, Surface electronic structures of

superconducting thin film MgB2 (0001), IEEE Trans. Appl. Supercond. 13 (2003) 3238-

3241.

69 K. Ueda, M. Naito, As-grown superconducting MgB2 thin films prepared by molecular

beam epitaxy, Appl. Phys. Lett. 79 (2001) 2046-2048.

70 W. Jo, J.-U. Huh, T. Ohnishi, A. F. Marshall, M. R. Beasley, R. H. Hammond, Thin

film superconducting MgB2 as-grown by MBE without post-anneal, Appl. Phys. Lett. 80

(2002) 3563-3565.

71 A. J. M. van Erven, T. H. Kim, M. Muenzenberg, J. S. Moodera, Highly crystallized

as-grown smooth and superconducting MgB2 films by molecular beam epitaxy, Appl.

Phys. Lett. 81 (2002) 4982-4984.

Page 154: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

139

72 A. Saito, A. Kawakami, H. Shimakage, Z. Wang, As-grown deposition of

superconducting MgB2 thin films by multiple-target sputtering system, Jpn. J. Appl. Phys.

41 (2002) L127-L129.

73 R. Schneider, J. Geerk, F. Ratzel, G. Linker, A. Zaitsev, In situ synthesis of MgB2 thin

films for tunnel junctions, Appl. Phys. Lett. 85 (2004) 5290-5292.

74 B. H. Moeckly, W. S. Ruby, Growth of high-quality large-area MgB2 thin films by

reactive evaporation, Supercond. Sci. Tech. 19 (2006) L21-L24.

75 X. H. Zeng, A. V. Pogrebnyakov, A. Kotcharov, J. E. Jones, X. X. Xi, E. M. Lysczek, J.

M. Redwing, S. Y. Xu, Q. Li, J. Lettieri, D. G. Schlom, W. Tian, X. Q. Pan, Z. K. Liu, In

situ epitaxial MgB2 thin films for superconducting electronics, Nature Materials 1 (2002)

35-38.

76 X. X. Xi, X. H. Zeng, A. V. Pogrebnyakov, S. Y. Xu, Q. Li, Y. Zhong, C. O. Brubaker,

Z.-K. Liu, E. M. Lysczek, J. M. Redwing, J. Lettieri, D. G. Schlom, W. Tian, X. Q. Pan,

In situ growth of MgB2 thin films by hybrid physical-chemical vapor deposition, IEEE

Trans. on Appl. Supercond. 13 (2003) 3233-3237.

77 X. X. Xi, A. V. Pogrebnyakova, S. Y. Xu, K. Chen, Y. Cui, E. C. Maertza, C. G.

Zhuang, Qi Li, D. R. Lamborn, J. M. Redwing, Z. K. Liu, A. Soukiassian, D. G. Schlom,

X. J. Weng, E. C. Dickey, Y. B. Chen, W. Tian, X. Q. Pan, S. A. Cybart, R. C. Dynes,

“MgB2 thin flms by hybrid physical-chemical vapor deposition,” Physica C. 456 (2007)

22-37.

Page 155: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

140

78 Z. K. Liu, D. G. Schlom, Q. Li, X. X. Xi, Thermodynamics of the Mg-B system:

Implications for the deposition of MgB2 thin films, Appl. Phys. Lett. 78 (2001) 3678-

3680.

79 R. A. Ribeiro, S. L. Bud0ko, C. Petrovic, P. C. Canfield, Carbon doping of

superconducting magnesium diboride, Physica C 384 (2003) 227-236.

80 Z. Y. Fan, D. G. Hinks, N. Newman, J. M. Rowell, Experimental study of MgB2

decomposition, Appl. Phys. Lett. 79 (2001) 87-89.

81 M. Ohring, The Materials Science of Thin Films, Academic Press (1992).

82 Daniel R. Lamborn, David W. Snyder, X.X. Xi and Joan M. Redwing, Journal of

Crystal Growth, Volume 299, Issue 2, 15 February 2007, Pages 358-364

83 Y. P. Liu, K. Warner, and C. Chan, 1989, J. App. Phys. 66 (11), 5514.

84 A. Brinkman, A. A. Golubov, H. Rogalla, O. V. Dolgov, J. Kortus, Y. Kong, O. Jepsen,

O. K. Andersen, Multiband model for tunneling in MgB2 junctions, Phys. Rev B 65

(2002) 180517.

85 M. Iavarone, G. Karapetrov, A. Menzel, V. Komanicky, H. You, W. K. Kwok, P.

Orgiani, V. Ferrando, X. X. Xi, Characterization of off-axis MgB2 epitaxial thin films for

planar junctions, Appl. Phys. Lett. 87 (2005) 242506.

86 Y. Cui, K. Chen, Q. Li, X. X. Xi, J. M. Rowell, Degradation-free interfaces in

MgB2/insulator/Pb Josephson tunnel junctions, Appl. Phys. Lett. 89 (2006) 202513.

87 X. H. Zeng, A. V. Pogrebnyakov, M. H. Zhu, J. E. Jones, X. X. Xi, S. Y. Xu, E. Wertz,

Q. Li, J. M. Redwing, J. Lettieri, V. Vaithyanathan, D. G. Schlom, Z. K. Liu, O.

Page 156: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

141

Trithaveesak, J. Schubert, Superconducting MgB2 thin films on silicon carbide substrates

by hybrid physical-chemical vapor deposition, Appl. Phys. Lett. 82 (2003) 2097-2099.

88 A. V. Pogrebnyakov, J. M. Redwing, J. E. Jones, X. X. Xi, S. Y. Xu, Q. Li, V.

Vaithyanathan, D. G. Schlom, Thickness dependence of the properties of epitaxial MgB2

thin films grown by hybrid physical-chemical vapor deposition, Appl. Phys. Lett. 82

(2003) 4319-4321.

89 S. Jin, H. Mavoori, C. Bower, R.B. van Dover, Nature (London) 411 (2001) 563.

90 T.H. Johansen, M. Baziljevich, D.V. Shantsev, P.E. Goa, Y.M. Galperin, W.N. Kang,

H.J. Kim, E.M. Choi, M.-S. Kim, S.I. Lee, Europhys. Lett. 59 (2002) 599.

91 Z.X. Ye, Q. Li, Y.F. Hu, A.V. Pogrebnyakov, Y. Cui, X.X. Xi, J.M. Redwing, Q. Li,

Appl. Phys. Lett. 85 (2004) 5284.

92 Z.X. Ye, Q. Li, Y.F. Hu, A.V. Pogrebnyakov, Y. Cui, X.X. Xi, J.M. Redwing, Q. Li,

IEEE Trans. Appl. Supercond. 15 (2005) 3273.

93 R. G. Mints and A. L. Rakhmanov, Rev. Mod. Phys. 53, 551 (1981), and references

therein.

94 S. L. Wipf, Cryogenics 31, 936 (1991).

95 I. Aranson, A. Gurevich, and V. Vinokur, Phys. Rev. Lett. 87, 067003 (2001).

96 R. A. Ribeiro, S. L. Bud’ko, C. Petrovic, and P. C. Canfield, Physica C 384, 227

(2003), and references therein.

97 P. G. de Gennes, Phys.: Condens. Matter 3, 79 (1964); N. R. Werthamer, E. Helfand,

and P. C. Hohenberg, Phys. Rev. 147, 295 (1966); K. Maki, ibid. 148, 362 (1966).

98 A. Gurevich, Phys. Rev. B 67, 184515 (2003).

Page 157: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

142

99 H.-J. Kim, W. N. Kang, E.-M. Choi, M.-S. Kim, K. H. P. Kim, and S.-I. Lee, Phys.

Rev. Lett. 87, 087002 (2001).

100 Khafizov et al. IEEE Trans. on Appl. Supercond, accepted for publication.

101 M. Lindgren et al., “Intrinsic picosecond response times of Y-Ba-Cu-O

superconducting photodetectors,” Appl. Phys. Lett., vol. 74, pp. 853–855, 1999.

102 M. Lindgren et al., “Y Ba2Cu3O7−x thin-film picosecond photoresponse in the

resistive state,” IEEE Trans. Appl. Supercon., vol. 7, pp. 3422–3425, 1997.

103 A. Rothwarf and B. N. Taylor, Phys. Rev. Lett. 19, 27 (1967).

104 F. A. Hegmann et al., “Electro-optic sampling of 1.5-ps photoresponse signals from Y

Ba2Cu3O7− thin films,” Appl. Phys. Lett, vol. 67, pp. 285–287, 1995.

105 F. A. Hegmann and J. S. Preston, “Origin of the fast photoresponse of epitaxial Y

Ba2Cu3O7 thin films,” Phys. Rev. B, vol. 48, pp. 16 023–16 039, 1993.

106 S. I. Anisimov, B. L. Kapeliovich, and T. L. Perelman, “Electron emission from metal

surfaces exposed to ultrashort laser pulses,” Zh. Eksp. Teor. Fiz., vol. 66, pp. 776–781,

1974.

107 F. A. Hegmann et al., “Electro-optic sampling of 1.5-ps photoresponse signals from Y

Ba2Cu3O7− thin films,” Appl. Phys. Lett, vol. 67, pp. 285–287, 1995.

108 F. A. Hegmann and J. S. Preston, “Origin of the fast photoresponse of epitaxial Y

Ba2Cu3O7 thin films,” Phys. Rev. B, vol. 48, pp. 16 023–16 039, 1993.

109 A.V. Pogrebnyakov, D.A. Tenne, A. Soukiassian, X.X. Xi, J.M. Redwing, V.

Vaithyanathan, D.G. Schlom, S.Y. Xu, Q. Li, M.D. Johannes, D. Kasinathan, W.E.

Pickett, Phys. Rev. Lett. 93 (2004) 147006.

Page 158: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

143

110 J. A. Floro, E. Chason, R. C. Cammarata, D. J. Srolovitz, Physical origins of intrinsic

stresses in Volmer-Weber thin films, MRS Bulletin 27(1) (2002) 19-25.

111 R. Koch, The intrinsic stress of polycrystalline and epitaxial thin metal films, J. Phys.:

Condens. Matter 6 (1994) 9519-9550.

112 Yang and Flynn, PRL, 62, 2476 (1989).

113 C. Buzea and T. Yamashita, Supercond. Sci. Technol. 14, R115 (2001).

114 Y. S. Touloukian et al., Thermophysical Properties of Matter (Plenum, New York,

1977), Vol. 13.

115 Z. Li and R. Brandt, J. Am. Chem. Soc. 69, 863 (1986).

116 J. M. Rowell, Supercond.Sci.Technol. 16 R17-R27(2003).

117 A. V. Pogrebnyakov, X. X. Xi, J. M. Redwing, V. Vaithyanathan, D. G. Schlom, and

A. Soukiassian, S. B. Mi and C. L. Jia, J. E. Giencke and C. B. Eom, J. Chen, Y. F. Hu, Y.

Cui, and Qi Li, “Properties of MgB2 thin films with carbon doping,” Appl. Phys. Lett. 85,

2017 (2004).

118 H.Y. Zhai, H.M. Christen, L. Zhang, M. Paranthaman, P.H. Fleming, D.H. Lowndes,

“Degradation of superconducting properties in MgB2 films by exposure to water’,

Supercond. Sci. Technol., vol. 14, pp. 425-428, 2001.

119 Serquis A, Zhu Y T, Peterson D E, Mueller F M, Schulze R K, Nesterenko V F,

Indrakanti, S S, Appl. Phys. Lett., 80 (2002) 4401.

120 Y. Cui, J. E. Jones, A. Beckley, R. Donovan, D. Lishego, E. Maertz, A. V.

Pogrebnyakov, P. Orgiani, J. M. Redwing, and X. X. Xi, “Degradation of MgB2 Thin

Films in Water,” IEEE Trans. Appl. Supercond. 15, 224 (2005).

Page 159: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

144

121 Serquis A, Zhu Y T, Peterson D E, Mueller F M, Schulze R K, Nesterenko V F,

Indrakanti, S S, Appl. Phys. Lett., 80 (2002) 4401.

122 W. Wong-Ng, private communication.

123 A. D. McLeod, J. S. Haggerty, and D. R. Sadoway, J. Am. Ceram. Soc. 67, 705

(1984).

124 Hisashi Shimakage, Kazuya Tsujimoto, Zhen Wang, Masayoshi Tonouchi, Appl.

Phys. Lett. 86, 072512 (2005).

125 Ueda K, Saito S, Sernba K, and T. Makimoto, Appl. Phys. Lett. 86, 172502 (2005).

126 R. K. Singh, R. Gandikota, J. Kim, N. Newman, and J. M. Rowell, Appl. Phys. Lett.

89, 042512 (2006).

127 G. Carapella, N. Martucciello, G. Costabile, C. Ferdeghini, V. Ferrando, and G.

Grassano, Appl. Phys. Lett. 80, 2949 (2002).

128 A. Saito, A. Kawakami, H. Shimakage, H. Terai, and Z. Wang, J. Appl. Phys. 92,

7369 (2002).

129 T. H. Kim and J. S. Moodera, Appl. Phys. Lett. 85, 434 (2004).

130 J.Geerk, R. Schneider, G. Linker, A.G. Zaitsev, R. Heid, K.-P. Bohnen, H. v.

Lohneysen, Phys. Rev. Lett. 94, 227005 (2005).

131 Griffiths, David J. (2004). Introduction to Quantum Mechanics (2nd ed.). Prentice

Hall.

132 I. Giaever and K Megerle, Phys. Rev. 122, 1101 (1961).

133 Y. Cui, Ke Chen, Qi Li, X. X. Xi, and J. M. Rowell, “MgB2/insulator/Pb Josephson

Junctions with Different Tunnel Barriers,” in press, IEEE Trans. Appl. Supercond.

Page 160: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

145

134 Y. Cui, Ke Chen, Qi Li, X. X. Xi, and J. M. Rowell, “Degradation-free interfaces in

MgB2/insulator/Pb Josephson tunnel junctions,” Appl. Phys. Lett. 89, 202513 (2006).

135 A. Brinkman, A. A. Golubov, H. Rogalla, O. V. Dolgov, J. Kortus, Y. Kong, O.

Jepsen, and O. K. Andersen, Phys. Rev. B, 65, 180517 (2002)

136 J. G. Simmons, J. Appl. Phys. 34, 238 (1963).

137 J. M. Rowell, in Tunneling Phenomena in Solids, edited by E. Burstein and S.

Lundqvist (Plenum, New York, 1969), p. 385.

138 T. Van Duzer and C. W. Turner, \Principles of Superconductive Devices and

Circuits," 2nd Ed., Prentice Hall, 1999.

139 A. S. Katz, A. G. Sun, R. C. Dynes, and K. Char, Appl. Phys. Lett., 66, 105 (1995).

140 B B Jin, T Dahm, C Iniotakis, A I Gubin, Eun-Mi Choi, Hyun Jung Kim, Sung-IK Lee,

W N Kang, S F Wang, Y L Zhou, A V Pogrebnyakov, J M Redwing, X X Xi and N Klein,

Supercond. Sci. Technol. 18 No 1 (2005) L1-L4.

141 A. Brinkman, S.H.W. Van der Ploeg, A.A. Golubov, H. Rogalla, T.H. Kim and J.S.

Moodera, J. Phys. Chem. Solids 67 (2006), p. 407.

142 A. Brinkman and J.M. Rowell, “MgB2 tunnel junctions and SQUIDs,” Physica C, 456,

188-195 (2007).

143 T. Dahm and N. Schopol, Phys. Rev. Lett. 91 (2003), p. 17001.

144 A.V. Pogrebnyakov, J.M. Redwing, S. Raghavan, V. Vaithyanathan, D.G. Schlom,

S.Y. Xu, Qi Li, D. A. Tenne, A. Soukiassian, X. X. Xi, M. D. Johannes, D. Kasinathan,

W. E. Pickett, J. S.Wu, and J. C. H. Spence, Phys. Rev. Lett. 93, 147006 (2004).

Page 161: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

146

145 R. S. Gonnelli, D. Daghero, A. Calzolari, G. A. Ummarino, Valeria Dellarocca, V. A.

Stepanov, S. M. Kazakov, N. Zhigadlo, and J. Karpinski, Phys. Rev. B, 71, 060503(R)

(2005).

146 M. Putti, M. Affronte, C. Ferdeghini, P. Manfrinetti, C. Tarantini, and E. Lehmann,

Phys. Rev. Lett. 96, 077003 (2006).

147 P. Szabo et al., Phys. Rev. Lett. 87 (2002), p. 137005.

148 H. Schmidt et al., Phys. Rev. Lett. 88 (2002), p. 127002.

149 R.S. Gonnelli et al., Phys. Rev. Lett. 89 (2002), p. 247004.

150 E. Babaev, Phys. Rev. Lett. 89 (2002), p. 0670011.

151 Eskildsen et al. Phys. Rev. Lett. 89, 187003 (2002).

152 S. Serventi et al., Phys. Rev. Lett. 93 (2004), p. 217003.

153 M. Ichioka, K. Machida, N. Nakai and P. Miranovic, Phys. Rev. B 70 (2004), p.

1445081.

154 A. Guman, S. Graser, T. Dahm and N. Schopohl, Phys. Rev. B 73 (2006), p. 104506.

155 Eskildsen et al. Phys. Rev. B. 68, 100508(R) (2003).

156 A. E. Koshelev, and A. A. Golubov, Phys. Rev. Lett. 90, 177002 (2003).

157 Dynes et al, Phys. Rev. Lett. 41, 1509 (1978).

158 Ke Chen, Y. Cui, Qi Li, X. X. Xi, Shane A. Cybart, R. C. Dynes, X. Wen, and J. M.

Redwing, “Planar MgB2 superconductor-normal metal-superconductor Josephson

junctions fabricated using epitaxial MgB2/TiB2 bilayers,” Appl. Phys. Lett. 88, 222511

(2006).

159 Ke Chen, Shane A. Cybart, and R. C. Dynes, Appl. Phys. Lett. 85, 2863 (2004).

Page 162: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

147

160 K. K. Likharev, Rev. Mod. Phys. 51, 101 (1979).

161 K. A. Delin, and A. W. Kleinsasser, Supercond. Sci. Tech. 9, 227 (1996).

162 A. Barone, and G. Paterno, Physics and Applications of the Josephson Effect, (Wiley,

New York) pp.305-309 (1982).

163 E. E. Mitchell, C. P. Foley, K. H. Muller, and K. E. Leslie, Physica C 321, 219 (1999).

164 Shane A. Cybart, Ke Chen, Y. Cui, Qi Li, X. X. Xi, and R. C. Dynes, “Planar MgB2

Josephson junctions and series arrays via nanolithography and ion damage,” Appl. Phys.

Lett. 85, 2017 (2006).

165 A. S. Katz, S. I. Woods, and R. C. Dynes, J. Appl. Phys. 87, 2978 (2000).

Page 163: MAGNESIUM DIBORIDE THIN FILMS AND DEVICES

148

Vita

Yi Cui

Education

• B.S., Physics, 2001, Nanjing University, Nanjing, China.

• Ph.D., Physics, 2007, The Pennsylvania State University, University Park, PA.

Honors and Awards

• Duncan Fellowship, The Pennsylvania State University, 2005, 2006

• Braddock Fellowship, The Pennsylvania State University, 2002

Professional Experience

• Research assistant, The Pennsylvania State University, 2004-2006

• Teaching assistant, The Pennsylvania State University, 2002-2003