2003 05 30 PHD Dissertation Gomez

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 130

UNIVERSITY OF CINCINNATI

DATE: May 30, 2003

I, Luis Beltran Gmez

hereby submit this as part of the requirements for the degree of:

Doctor of Philosophy
in:

Physics
It is entitled:

High Frequency Electrical Transport Measurements Of


Niobium SNS Josephson Junction Arrays And Niobium
Thin Films With Nanoscale Size Magnetic Dot Array
Approved by:
Prof. David Mast
Prof. Howard Jackson
Prof. Mark Jarrell
Prof. Brian Meadows

High Frequency Electrical Transport Measurements Of Niobium


SNS Josephson Junction Arrays And Niobium Thin Films With
Nanoscale Size Magnetic Dot Array
A dissertation submitted to the

Division of Research and Advanced Studies


of the University of Cincinnati
in partial fulfillment of the
requirements for the degree of
DOCTORATE OF PHILOSOPHY (Ph.D.)
in the Department of Physics
of the College of Arts and Sciences
2003
by
Luis B. Gmez
Licenciado en Fsica, Universidad Central de Venezuela, 1992
M.S., University of Cincinnati, 1993
Committee Chair: Professor David B. Mast

ABSTRACT
High Frequency Electrical Transport Measurements Of Niobium SNS Josephson Junction
Arrays And Niobium Thin Films With Nanoscale Size Magnetic Dot Array
By
Luis B. Gmez
Doctor of Philosophy in Physics
University of Cincinnati, 2003
Professor David. B. Mast. Chair
In this dissertation, measurements on two different niobium (Nb) based systems,
namely superconductor-normal-metal-superconductor (SNS) Josephson junction arrays
(JJA), and thin Nb films with nanoscale size magnetic dot arrays, are presented.
Without high frequency (rf) signals at nearly zero magnetic field, these
measurements show similarities for the two systems. These similarities are explained by
proposing that very close to the critical temperature (Tc) of the Nb film sample, the stray
magnetic field of the dots reduces the superconductivity in the film to make it a
superconductor-weaker-superconductor-superconductor (SSS) JJA.
With rf signals, their dc-voltage-current characteristics (VIs) show the
appearance of constant voltage steps known as Shapiro steps.

For SNS-JJA, the

dependence of the width of the steps ( In) on the amplitude of the rf current (Irf) is
Bessel-function-like. The maximum width of the first step ([ I1]MAX) was found to
increase as a function of rf frequency ( rf) reaching a maximum for rf~1.8 times the
arrays Josephson frequency ( c). This maximum was maintained for rf>4.9 c. Also,
in JJA with missing junctions, [ In]MAX was smaller, disappearing for rf~ c.

For Nb films near Tc, no Bessel-function-like oscillations were observed for In.
Also, I0 (or critical current, Ic) showed a maximum for Irf 0, contrary to what was
found in SNS-JJA, falling to zero for higher Irf. This enhancement of Ic with Irf was
attributed to non-equilibrium superconductivity mechanisms near Tc. Observation of
higher order steps (n=1,2,3,etc) became evident only from the dynamic resistance
(dV/dI)vs.V plots (where a step corresponds to a minima for this curve). The first step
was most evident for Irf values that made I0=0. At these values, V1 (the voltage that
corresponded to the first minimum in the dV/dIvs.V curve) satisfied the Josephson
relation V1=N(h/2e) rf, where N is the number of junctions in the direction of the
current. At fixed rf, V1 was found to increase with Irf. This result was interpreted as
having more junctions participating in the steps formation, which was attributed to
magnetic interactions among the dots magnetic moments and with the Irf to minimize the
total energy of the system.

Copyright by
Luis Beltran Gmez
2003

DEDICATION
To my wife and kids.

Nothing is more powerful than an idea whose time has come


Victor Hugo

ACKNOWLEDGMENTS
Many people have helped me throughout my graduate school life. It is impossible
for me to mention them all, but I would like to acknowledge a few of them here.
First of all, the completion of this work has been possible with the incredible help
and love of my wife, Kate and my kids; Florence, Evan and Vera, without which I would
not be writing these lines.
Second, I am grateful to my advisor, Dr. David Mast, for his guidance and support
through my many years of graduate school. Also, I am grateful to the members of my
committee; Dr. Howard Jackson, Dr. Mark Jarrell, and Dr. Brian Meadows, for their help
and guidance. I would like to give special thanks to Dr. Mike Sokollof for his many
hours of mentoring in data interpretation and his personal help in seeing this work
completed.
I would like to thank Dr. Ivan Schuller, Dr. Yvan Jaccard, and Axel Hoffmann for
their help providing the magnetic dots samples used in this dissertation and for their
useful discussions and letters of encouragement. Also, I would like to thank Maribel
Montero, Charles Reichhart, and Lieve Van Look for the useful discussions about their
work. Also, I would like to thank Dr. George Crabtree for his opinions regarding my
work.
I would like to thank Dr. Miguel Octavio and the late Dr. Juan Aponte from IVIC,
who helped me to graduate school and for initiating me into low temperature physics and

superconductivity. Also, I would like to give special thanks to Dr. Lisseta DOnofrio and
all my professors from UCV, for their teachings, and support.
I thank the rest of the faculty and staff at UCs Physics department who helped
and encouraged my continuation in graduate school.
Finally, I would like to thank my family and friends, for their love and for
stopping asking me when are you getting done? For all of you a BIG thanks!

TABLE OF CONTENTS
CHAPTER 1 : INTRODUCTION .............................................................................................................. 9
CHAPTER 2 : THEORY ........................................................................................................................... 12
2.1 INTRODUCTION TO SUPERCONDUCTIVITY ............................................................................................ 12
2.2 TYPE II SUPERCONDUCTIVITY AND THE MIXED STATE ......................................................................... 22
2.3 THE JOSEPHSON JUNCTION EFFECT ...................................................................................................... 24
2.4 SHAPIRO STEPS .................................................................................................................................... 27
2.5 2D JOSEPHSON JUNCTION ARRAYS ...................................................................................................... 29
CHAPTER 3 : EXPERIMENTS ............................................................................................................... 32
3.1 DESCRIPTION OF SAMPLES ................................................................................................................... 32
3.2 EXPERIMENTAL SETUP......................................................................................................................... 36
3.3 MEASUREMENTS TECHNIQUES AND DATA ACQUISITION ...................................................................... 41
CHAPTER 4 : MEASUREMENTS IN SNS JJA..................................................................................... 47
4.1 MEASUREMENTS ................................................................................................................................. 47
4.1.1 R vs. T .......................................................................................................................................... 47
4.1.2 R vs. B.......................................................................................................................................... 49
4.1.3 VIs .............................................................................................................................................. 53
4.1.4 VIs with rf................................................................................................................................... 58
4.1.5 Dependence of VIs with rf amplitude ......................................................................................... 63
4.1.6 Dependence of VIs with rf frequency.......................................................................................... 65
4.1.7 Effects of site disorder ................................................................................................................. 66
4.2 DISCUSSION......................................................................................................................................... 71
4.3 SUMMARY ........................................................................................................................................... 72
CHAPTER 5 : MEASUREMENTS IN NB FILMS WITH MAGNETIC DOT ARRAY .................... 74
5.1 MEASUREMENTS ................................................................................................................................. 74
5.1.1 R vs. T .......................................................................................................................................... 74
5.1.2 R vs. B.......................................................................................................................................... 79
5.1.3 VIs .............................................................................................................................................. 81
5.1.4 VIs with rf................................................................................................................................... 84
5.1.5 Dependence of VIs with rf amplitude ......................................................................................... 87
5.1.6 Dependence of VIs with rf frequency.......................................................................................... 95
5.2 DISCUSSION......................................................................................................................................... 98
5.3 SUMMARY ......................................................................................................................................... 106
CHAPTER 6 : CONCLUSIONS ............................................................................................................. 107
BIBLIOGRAPHY..........111
APPENDIX.........118

LIST OF TABLES
Table 5.1: Table showing the values for the voltage location for the n=1 minimum found in Figure 5.12.
The last column to the right is a calculation from Equation 2.31 of the number of junctions in the
current direction (N) using V1 and rf=2GHz. .....................................................................................91
Table 5.2: Table showing the values for the voltage location for the n=1 minimum found in Figure 5.14.
The last column to the right is a calculation from Equation 2.31 of the number of junctions in the
current direction (N) using V1 and rf=2GHz. .....................................................................................94

TABLE OF FIGURES
Figure 2.1: H-T phase diagram for a bulk superconducting slab in a uniform magnetic field (using Equation
2.1 and parameters found for bulk Al [25]) ...........................................................................................14
Figure 2.2: Graphical representation of Equation 2.4 for the superconducting half-space (using values found
for bulk Nb [25] at zero temperature)....................................................................................................16
Figure 2.3: Graphical representation of Equation 2.6 (using values for bulk Nb [25]) .................................17
Figure 2.4: H-T phase diagram for a type II superconducting slab in a uniform magnetic field. ..................23
Figure 2.5: RCSJ model representation of a real JJ. The total current i into the junction is divided into three
channels: the supercurrent (ic) through the ideal junction, the current through the junction capacitor
(C), and the quasiparticle current through the resistor (R0). ..................................................................26
Figure 3.1: SNS Josephson junction array (not to scale) of 53 junctions. Current is injected via the Nb
bus bars to the sides of the array and the voltage is measured across the whole array. .........................33
Figure 3.2: Measuring bridge prototype for the film samples. The contact lead names are assigned as
shown. In this photograph, the brown square in the center is where the magnetic dots array is located.
The clear pattern is the Nb bridge for transport measurements. The gray area is the Si substrate. ......35
Figure 3.3: Picture of the lower part of the probe without vacuum can, showing the main parts.................37
Figure 3.4: Probes lower part, showing vacuum can and superconducting magnet. ...................................38
Figure 3.5: Sample holder chamber showing the electrical connection for current, voltage and high
frequency signals (showing Nb film sample ready to be measured). ....................................................39
Figure 3.6 Picture showing in detail the temperature controlling stage and sample holder...........................41
Figure 3.7: Block diagram for the experimental set up showing dc and rf electrical connections. ...............45
Figure 4.1: Resistive transition for a typical SNS JJA sample. Measurements were taken with a bias current
of 10A in ambient magnetic field. The resistance error bars are shown in red and the line connecting
the dots is a guide to the eye..................................................................................................................48
Figure 4.2: Resistance vs. current passed through the solenoid. Measurements were taken at a temperature
T < Tc0 of 3.000K. The sample was biased with I < Ic of 200A. The error bars for the resistance are
shown in red, and the line connecting the dots is a guide to the eye. ....................................................50
Figure 4.3: VI curve at ambient magnetic field. The curve was taken for T < Tc0 of 3.000K. The error bars
for V are shown in red, and the line connecting the dots is a guide to the eye. .....................................54
Figure 4.4: Same data plotted in Figure 4.3, but on a log-log scale. The error bars for V are shown in red,
and the line connecting the dots is a guide to the eye............................................................................54
Figure 4.5: Dynamic resistance curve calculated from Figure 4.3. The error bars for dV/dI are shown in
red, and the line connecting the dots is a guide to the eye.....................................................................56
Figure 4.6: Current vs. normalized-voltage graph for a 100% 300200 JJA at T=1.8K and zero magnetic
field. The black curve is for no rf signal. The red curve shows Shapiro steps when rf signal is applied.
The voltage separation between steps is proportional to the rf frequency and the step width is
proportional to the rf power...................................................................................................................59
Figure 4.7: dV/dI vs. V curve shows a series of minima located at the steps voltage. The plot shows
dV/dI0 at the step location meaning that the step has non zero spread. ..............................................62
Figure 4.8: Detail from Figure 4.7 illustrating how to find the spread of the first step by using the dynamic
resistance of the VI, to indicate the center of the step, and the derivative of dV/dI as a function
voltage, to find its value. .......................................................................................................................63
Figure 4.9: Widths for steps n=0 and n=1 as a function of applied rf amplitude. Graph shows Bessel
function like oscillations for the step widths. Notice how the oscillations are approximately 180 out
of phase. The n=0 and n=1 steps first maxima ([ I0]MAX and [ I1]MAX) for small Irf, are later used
to further characterize the JJA irradiated by rf signal............................................................................64
Figure 4.10: Maximum step width for steps n=1 as a function of reduced frequency for a 100% array
(Figure 1 in reference [12]. Data points are obtained from similar curves as shown in Figure 4.9. In
here T=1.91 K and f 0.........................................................................................................................65

Figure 4.11: Normalized voltage vs. current for a 90% sample (300300). The first Shapiro step presents
considerable reduction in its width and increase in its spread when compared to 100% arrays............67
Figure 4.12: Normalized maximum Shapiro step width for the first step vs. normalized frequencies, for two
samples with site disorder (90% and 80% samples) (from reference [12], inset in Figure 1). The lines
are a guide the eye. ................................................................................................................................68
Figure 4.13: Voltage spread as a function of normalized frequencies for site disorder samples (from
reference [12], Figure 3). Results obtained for a 100% sample are shown for comparison. The spread
increases linearly with frequency as can be seen from the linear fit to the data for the 90% and 80%
samples. The line linking the 100% data points is a guide to the eye...................................................70
Figure 5.1: Resistive transition for the Nb/Ni square sample. Measurements were taken with a bias current
of 10A in ambient magnetic field. The resistance error bars are shown in red and the line connecting
the dots is a guide to the eye..................................................................................................................75
Figure 5.2: Low temperature detail from Figure 5.1, showing the local resistance minimum at ST=7.860K
followed by an increase in resistance until it reaches a maximum to drop to zero resistance at
Tc0=7.766K. The resistance error bars are shown in red and the line connecting the dots is a guide to
the eye....................................................................................................................................................76
Figure 5.3: Drawing for the Nb film sample grown on top a square array of magnetic dots. The magnetic
dots are the gray circles shown in the drawing. The Nb film covers the whole array of dots (the Nb
film right on top of the dots is not shown for clarity). This drawing is to scale, where the diameter of
the circles is 200nm. ..............................................................................................................................77
Figure 5.4: Resistance vs. magnetic field perpendicular to the sample. The graph shows data for two
slightly different temperatures below TRMAX. The current bias to the sample (627A) is the same in
both cases. .............................................................................................................................................79
Figure 5.5: VI curve at ambient magnetic field. The curve was taken T ~7.8K > Tc0. The error bars for V
are shown in black. The red curve is the temperature of the sample at each voltage point. The error
bars for ST are shown in red. The line connecting the dots is a guide to the eye. ................................81
Figure 5.6: Same data plotted in Figure 5.5, but on a log-log scale. The blue line here represents the
arbitrary value of voltage chosen to define Ic using the threshold method. The line connecting the dots
is a guide to the eye. ..............................................................................................................................82
Figure 5.7: Dynamic resistance curve calculated from Figure 5.5. The error bars for dV/dI are shown in red
(calculated in the same way as in Figure 4.5)........................................................................................83
Figure 5.8: VI curve for the Nb/Ni square sample at T~7.8K and ambient magnetic field. The black curve
is the VI with rf signal. The red curve is the temperature of the sample. No Shapiro steps are apparent
for rf=2GHz and Irf=123A..................................................................................................................85
Figure 5.9: dV/dI vs. V curve from VI with rf curve in Figure 5.8. The dynamic resistance is shown in blue
and the IV, in the black curve, is plotted for reference..........................................................................86
Figure 5.10: Plot of the voltage location for the minima found in the dynamic resistance curve in Figure
5.9. These voltages are plotted vs. the number of the minima staring with the first minimum at n=1, a
linear fit of this data in shown. For the calculations of the linear fit, the (0,0) point is included. ........87
Figure 5.11: Plot of the critical current (defined by the threshold method) vs. Irf. The critical current is the
black curve. A plot (in red) of the sample temperature is shown to indicate the increase in temperature
produced by the applied Irf.....................................................................................................................88
Figure 5.12: Plot of the dynamic resistance vs. voltage for five different Irf. The green curve is the one
plotted in Figure 5.8. The curves have been shifted a fixed amount in the y-axis for clarity. The line
connecting the dots is a guide for the eyes. ...........................................................................................90
Figure 5.13: Similar plot to Figure 5.12. The range [40-70] dB in the attenuator box was found to deliver
the same amount of Irf as in Figure 5.11. A maximum in the critical current as a function of Irf is more
evident. ..................................................................................................................................................92
Figure 5.14: Plot of the dynamic resistance vs. voltage for four different Irf. The curves have been shifted a
fixed amount along the y-axis for clarity. The line connecting the dots is a guide for the eyes. ..........93
Figure 5.15: Dynamic resistance vs. voltage plot. Here the dV/dI was calculated from five different VIs
all taken with an attenuation of 51dB. The first curve on top (51D) is the same curve plotted in Figure
5.13 for 51dB attenuation. The curve at the bottom (51H) was measured by allowing the current to

decrease after reaching the maximum, then increase back again from zero and so on until completing 3
round trips. In this plot the x-axis was limited to show data only up to 1.5 mV. The curves were
shifted a fixed amount in the y-axis for clarity. .....................................................................................95
Figure 5.16: dV/dI vs. V curves as a function of rf frequency. The curves were shifted a fixed amount in
the y-axis for clarity...............................................................................................................................96
Figure 5.17: Plot of the voltage points label V1 from Figure 5.16 as a function of rf frequency. The (0,0)
point has been included in the calculation. ............................................................................................97
Figure 5.18: Transverse voltage vs. current in the Nb/Fe square sample. .....................................................99
Figure 5.19: Magneto resistance curve for the Nb/Ni square sample. The curve in black is the magneto
resistance of the sample (the error bars are shown in blue). The curve in red is the temperature of the
sample (the error bars are shown in dark yellow). The error bars in this curve are the standard
deviation of N averages taken at each data point. The error of the mean value is at least 10 times
smaller. An increase of sample temperature is observed at f =2. This warming is thought to be
associated with a ferromagnetic alignment of the magnetic dots in the sample. .................................101
Figure 5.20: Magneto resistance curve for the Nb/Ni square sample. The curve in black is the magneto
resistance of the sample (the error bars are shown in blue). The curve in red is the temperature of the
sample (the error bars are shown in dark yellow). The error bars in this curve are defined as
( / N , the standard deviation of N averages divided by the square root of N). A lowering in
sample temperature is observed at f =1. This cooling is thought to be associated with an antiferromagnetic alignment of the magnetic dots in the sample. .............................................................103

LIST OF EQUATIONS
Equation 2.1

Equation 2.2
Equation 2.3

T 2
H c (T ) H c (0 )1 ........................................................................................13
Tc
v
* 2
J s ns e* v
=
E ..............................................................................................................15
t
m*
v
v
2v
B = 0L J s .........................................................................................................15

Equation 2.4

x
...................................................................................................15
B ( x) = Ba exp
L

Equation 2.5

T
ns
= 1
nn
Tc

..............................................................................................................16

Equation 2.6

Equation 2.7

Equation 2.8
Equation 2.9
Equation 2.10

T 4 2
(T ) 0 1 ............................................................................................17
Tc
1 1
1
=
+
................................................................................................................18
0 l e

= L

0
le

....................................................................................................................18

v
v
v
(r ) = (r ) e i (r ) ........................................................................................................18
v
v 1/ 2
(r ) = ns* (r ) ......................................................................................................19

Equation 2.11
Equation 2.12
Equation 2.13
Equation 2.14

r
r 2 r
r
1 h

e * A (r ) + (r ) (r ) = (T ) (r ) ..............................19
*
2m i

*
r
r
h r
e
J s = * Re * e * A .........................................................................19
m

i
*
r
e hn s r r
(r ) .................................................................................................19
Js =
2m *
h
(T ) =
................................................................................................19
1/ 2
*
2m (T )

Equation 2.15

(0) = 0le

Equation 2.16

(T ) = 0.855

..............................................................................................................20

( 0)
T
1
Tc

..............................................................................................20

Equation 2.17

Equation 2.18

m*
' = + *2
e

Equation 2.19
Equation 2.20
Equation 2.21
Equation 2.22
Equation 2.23
Equation 2.24
Equation 2.25

..........................................................................................................................20

r r
J s dl

..............................................................................................21

h
.........................................................................................................................21
2e
is = ic sin ( 21 ) ............................................................................................................24

0 =

h d ( 21 )
..........................................................................................................24
2e dt
2
2e v v
21 = 2 1 A dr ........................................................................................24
h 1
2e
c = V ...................................................................................................................25
h
h
E s = ic cos( 21 ) = E J cos( 21 ) .................................................................25
2e
V
dV
i = ic sin ( 21 ) +
+C
.....................................................................................26
R0
dt
V =

h d 21 hC d 2 21
.............................................................26
+
2eR0 dt
2e dt 2

Equation 2.26

i = ic sin ( 21 ) +

Equation 2.27

p =

Equation 2.28

V (t ) = Vdc + Vrf cos(2 rf t ) ..................................................................................28

Equation 2.29

i s = ic (1) n J n (

Equation 2.30

[in ] MAX = ic J n (

Equation 2.31
Equation 3.1
Equation 4.1
Equation 4.2

2ic
.............................................................................................................26
0C

Vrf
0 rf
Vrf

0 rf

) sin[2 ( c n rf )t + 0 ] .......................................28

) .........................................................................................28

h
Vn = nN rf .......................................................................................................30
2e
V (I +, T , B, I rf ) V (I , T , B, I rf )
V (I , T , B, I rf ) =
.................................................43
2
V
R=
..........................................................................................................................48
I bias

B 0 = b 20 ................................................................................................................50
a

Equation 4.3

Equation 4.4
Equation 4.5
Equation 4.6
Equation 4.7
Equation 4.8

Ba a 2
...................................................................................................................51
f =


.......................................................................................................51
I m = 2 I c cos
0
I c , v = 0.1I c ......................................................................................................................56
Ic
............................................................................................................................57
M
M
R0 =
RN ....................................................................................................................57
N
1 I c RN
c =
................................................................................................................57
0 N
ic =

RN I c2 NM
=
=
R0ic2 ...........................................................................................58
8
8

Equation 4.9

PMAX

Equation 5.1

B(t ) = B(0) sin(t )e

................................................................................................99

Chapter 1 : INTRODUCTION
The main focus of this dissertation is to investigate the dynamic response to an
external radio frequency (rf) signal (in nearly zero external magnetic field) of two
different niobium (Nb) based systems. The systems investigated were: superconductornormal-metal-superconductor (SNS) [1] Josephson junction arrays (JJA) [2, 3], and thin Nb
films deposited on top of a nanoscale size magnetic dot arrays [4].
This study is motivated by the similarities observed in the resistive transition
curves (RTs) of these two systems that allow us to propose a model in which the Nb film
samples are viewed as a superconductor-weaker-superconductor-superconductor (SSS)
[1]

JJA. This model is based on the assumption that the superconductivity in the film is

locally destroyed by the stray magnetic field of the dots [5, 6] and weak links or junctions
are created in the film, in the direction of the current.
In JJA, the coherent response to an rf signal is observed as a series of steps in
their dc voltage-current characteristics (VIs) known as Shapiro steps

[7]

. The study of

the conditions that give rise to the disruption of this coherence (i.e. the disappearance of
Shapiro steps from the VIs) is an important issue that can be used to sharpen our
understanding of the dynamic processes occurring in these systems. Also, in the case of
classical JJA, these studies will help us in the utilization of these systems in
applications [8, 9, 10].
When ideal classical JJA (i.e. JJA in which superconducting islands are
described by a single order parameter, with a pure sinusoidal current-phase-relationship

(CPR) among their junctions and minimal disorder), in a zero external magnetic field,
are irradiated with rf signals, well defined Shapiro steps are observed in their VIs for
large ranges of external rf signal amplitudes and frequencies [12]. In these circumstances,
the width of the steps (the range of currents in the VIs over which the voltage is a
constant, or I) can be used as a measure of the strength of the coherent response of
the JJA to the rf signal.
The conditions that disrupt the coherent response of JJA to an external rf signal
include: unavoidable sample fabrication defects, known as bond disorder

[11]

imperfections in the array (the removal or disappearance of junctions from the array)
known as site disorder [12], and having a non-sinusoidal CPR [13, 14] between the junctions.
Non-sinusoidal CRPs in junctions occur when the coupling between the superconducting
islands is not-weak

[15,

16,

17]

and/or the system is operated in conditions of

nonequilibrium superconductivity [18, 19, 20].


The consequence of disorder in JJA is to create a random distribution of critical
currents (Ic) and shunted resistances (RN) among the junctions in the array that will
diminish the strength of the Shapiro steps

[21]

(i.e. how well they are formed in the

VIs). More fundamentally, the nonequilibrium processes that can occur in these systems
affects the superconducting state in the islands themselves (described by a order
parameter = n s e i ), changing both the number of superconducting electrons (ns) as
well as their phases ( )

[22, 23]

. When this occurs, the macroscopic variables Ic and RN,

which are derived from ns and , will also change, consequently impairing the ability of
theses systems to respond coherently to the external rf signal.

10

This dissertation is organized as follows:

Chapter 2 covers the theoretical

background of superconductivity, Josephson Effect and Shapiro steps in junctions and


arrays. Chapter 3 covers the experimental aspects of this work. Chapter 4 presents the
experimental results found for SNS JJA. Chapter 5 presents the experimental results
found in the superconducting film samples. Conclusions and suggestions for future
research are given in Chapter 6.

11

Chapter 2 : THEORY
In this chapter, a historical overview of the experimental discoveries in the field
of superconductivity will be given. Also, a summary of relevant concepts and theories
necessary to understand the experimental part of this dissertation is presented. The reader
is referred to the literature for in depth derivations of the equations presented here.
In section 2.1, the fundamental experiments and key concepts are summarized. In
section 2.2, the ideas that are directly applied to superconducting thin films are
summarized. In section 2.3 the Josephson junction effect is introduced. In section 2.4,
the coherent response of a Josephson junction (JJ) to an external radio frequency signal is
developed with an emphasis on the concept of Shapiro steps. Finally, in section 2.5 two
dimensional (2D) Josephson junction arrays (JJA) are introduced as well as the concept
of a Shapiro step for arrays is developed. Also, a summary of current experiments in
which coherence response is observed in other superconducting systems (i. e.
superconducting wire networks, and superconducting films with thickness variations and
array of holes) is given.

2.1 Introduction to superconductivity


In 1911 Kamerlingh Onnes [24] discovered superconductivity while measuring the
resistivity of a highly pure mercury bulk sample as a function of temperature. As the
temperature of the sample reached slightly below 4.2K, its dc-resistivity dropped abruptly
to zero. The temperature at which this occurs is known as the critical temperature Tc.

12

In pure bulk samples the change from resistive to resistive-less takes place over a
temperature range of only a few hundredths of a degreei [25].
In 1914 Onnes

[26]

also found that if a superconductor (at T<Tc) is placed in a

strong magnetic field, the zero dc-resistance state is destroyed, to be later regained
when the magnetic field is removed. The minimum magnetic field required to destroy
superconductivity is called the critical (magnetic) field Hc and experimentally its value
depends on the shape of the sample and its orientation to the external magnetic fieldii.
The temperature dependence of Hc is represented in bulk superconductors by the
parabolic relationship:

Equation 2.1

T
H c (T ) H c (0 )1
Tc

In 1933 Meissner and Ochsenfield [27] discovered that a superconducting material


was more than merely a perfect conductor. By systematically applying a weak magnetic
field to a lead sample, they discovered that the effective relative permeability of the
sample was zero when the sample was in the superconducting state, concluding that the
magnetic field inside the superconductor must be zeroiii. This perfect diamagnetism, or
Meissner effect, observed in superconductors cannot be derived from its perfect
conductivity and therefore is a unique characteristic of the superconducting state.

For bulk niobium (Nb), Tc =9.25 K. See reference [25]


For bulk Nb, 0Hc(0)=2060 G. See reference [25]
iii
Later it was found that this is only true in bulk samples, except for a thin layer of material near the
surface of the sample of ~50nm
ii

13

The magnetic field-temperature phase diagram for a sample exhibiting the


Meissner effect is shown in Figure 2.1. The curve Hc(T) separates the normal from the
superconducting state.

110

0Hc(T)

Magnetic Induction (G)

100
90

Normal
State

80
70

B=0
Superconducting State
or
Meissner state

60
50
40
30
0.0

0.2

0.4

0.6

0.8

1.0

T/Tc

Figure 2.1: H-T phase diagram for a bulk superconducting slab in a uniform
magnetic field (using Equation 2.1 and parameters found for bulk Al [25])

Soon after the Meissner effect was discovered, the London

[28]

brothers in 1935

used this experimental evidence to develop a phenomenological theory to accommodate


the electric and magnetic properties of superconductors. Starting from a simple model
for the electromagnetic behavior of a perfect conducting gas, H. and F. London
concluded that the magnetic behavior of a superconductor might be correctly described
v
v
B
according to the Meissner effect if not only
= 0 but also B decreases exponentially
t

inside the superconductor. This theory, which is still recognized as valid, is summarized
by the following two equations:

14

Equation 2.2

v
* 2
J s ns e* v
=
E
t
m*

Equation 2.3

v
v
2v
B = 0 L J s

Where ns* is the number of super-electrons per unit volume, m* and e* are
respectively the effective mass and charge of these super-electronsi.

L =

The factor

m*
is known as the London penetration depthii. A one-dimensional solution
* *2
ns e 0

of the Londons second equation for a superconducting half-space in a static magnetic


induction is given by:

Equation 2.4

B ( x) = Ba exp

It was later found in the BCS theory that the super electrons are really a pair of electrons called a
Cooper pair, so that e* = 2e and m* = 2me.
ii
For bulk Nb, 0 = 39 nm.

15

2500

Vacuum

Superconductor

2000

B(X) (G)

1500

1000

B(0)

500

0
0
0

39 50

100

150

X (nm)

Figure 2.2: Graphical representation of Equation 2.4 for the superconducting halfspace (using values found for bulk Nb [25] at zero temperature)

A two-fluid model for superconductors was developed in 1934 by Gorter and


Casimir

[29]

originally to try to explain the observed thermodynamic properties of

superconductors.

The model postulates that below Tc the conduction electrons are

divided into two distinct groups: a superconducting aggregate of super electrons ns


(=2ns*) and a fraction of normal state electrons nn.

The best agreement between

experiments and the theory is found if:

Equation 2.5

T
ns
= 1
nn
Tc

This model, combined with a formulation of the ac electrodynamics of


superconductors based on the London equations, gives accurate predictions for the
London penetration depth L (T) i.e.,

16

T
(T ) 0 1
Tc

Equation 2.6

200
180

( T/Tc ) (nm)

160

( 0.98 Tc )

140
120
100
80
60

40
0.0

0.2

0.4

0.6

0.8

1.0

T / Tc

Figure 2.3: Graphical representation of Equation 2.6 (using values for bulk Nb [25])

The similarities between Londons first equation and Ohms law allowed Pippard
[30]

in 1953 to propose a non-local generalization of Equation 2.6, in order to explain

some experimental observations based on his study of the penetration depth in various
types of superconductors.
Chambers

[31]

Following a non-local expression of Ohms law, as per

to explain the anomalous skin effect, Pippard introduced two new

parameters 0 and . The parameter 0 i is called the intrinsic coherence length and is
independent of impurities. The parameter is called the effective coherence length and
may be given by the empirical relation:

For bulk Nb

0 = 38 nm.

17

Equation 2.7

1
l e

where le is the electron mean free path and is a constant on the order of unity.
There are two extreme cases in which Pippards non-local relationship reduces to
a simpler local form. The case for >> , which is known as the dirty limit, is
applicable for very impure specimens, alloys, or thin films. For this limit, le is much
smaller than 0 and le and the penetration depth is given by:

= L

Equation 2.8

0
le

The other extreme case for << 0 or the clean limit, is applicable for most
pure bulk superconductors at temperatures not too close to Tc.
The Ginzburg-Landau (G-L) theory [32] is a generalization of the London theory to
deal with situations in which ns varies in space, and also to deal with the nonlinear
response to fields that are strong enough to change ns. In this theory, Londons ns is
2

given by: ns = (x) , where is a complex pseudo wave-function taken as an order


parameter within Landaus general theory of second-order phase transitions.
This order parameter has a well-defined amplitude and phase angle and represents
a macroscopic many-body wave function of the superconducting electrons inside the
superconductor. It is written as:
Equation 2.9

v
v
v
(r ) = (r ) e i (r )

18

Where:

Equation 2.10

v
v 1/ 2
(r ) = ns* (r )

Assuming a series expansion of the free energy in powers of and ,


Ginzburg and Landau derived the following equation:
2

Equation 2.11

r
r 2 r
r
1 h

e * A (r ) + (r ) (r ) = (T ) (r )
*
2m i

Here, and are expansion coefficients.

Equation 2.11 is analogous to the

Schrdinger equation for a particle of charge e*and mass m*, which includes a non linear
term. Using Equation 2.10, the supercurrent is calculated as:

Equation 2.12

r
r
h r
e*
J s = * Re * e * A
m

i

Using Equation 2.9, the current inside a superconductor in the absence of applied
field, is given by:

Equation 2.13

r
e * hn s r r
Js =
( r )
2m *

This is found to vary linearly with the gradient of the phase in the superconductor.
From the G-L theory a new characteristic length is introduced as:

Equation 2.14

(T ) =

2m (T )
*

19

1/ 2

This length characterizes the distance over which can vary without an excessive
increase in energy and it is the distance of reduction of ns near a superconducting
boundary. This characteristic length is related to Pippards 0 for T<<Tc, but they are
distinct quantities. Near Tc, (T) diverges as (Tc T )

For the dirty limit at zero temperature it is found that:


Equation 2.15

(0) = 0le

And (T) is:

Equation 2.16

(T ) = 0.855

( 0)
1

T
Tc

Also from the G-L theory, another parameter known as the G-L parameter ; is
introduced. It is defined as follows:

Equation 2.17

This parameter will be used in the next section to introduce type II superconductors,
vortices, and the mixed state.
A direct consequence of the quantum nature of superconductivity (where the
effects of the wave-particle dualism are observed at macroscopic scales and the phase
and particle number are conjugate variables N 1 ) is the conservation of a quantity,

introduced by F. London, known as the fluxoid ' . It is defined as:

20

m*
= + *2
e
'

Equation 2.18

r r
J s dl

where is the ordinary magnetic flux through the integration loop.


Placing a simply connected bulk superconductor in an external weak magnetic field
r
and choosing an integration loop (deep into the superconductor, where J s = 0 ) to

evaluate Equation 2.18 implies that ' = . But the requirements of having a singlevalued order parameter implies that its phase could be the same or an integer multiple of
2 around the close loop. Therefore, the fluxoid itself is quantized by n 0 , where 0

is the quantum flux.

0 =

Equation 2.19

h
2e

The microscopic theory for the superconducting state was developed by Bardeen
Cooper and Schrieffer

[33]

in 1957 (also known as BCS theory). The basic hypothesis of

the theory is that the electrons in the superconducting state form a bound pair, known as a
Cooper pair. Cooper [34], in 1956 showed that under certain conditions the normal Fermi
gas is unstable and there is a tendency towards pairing if there is a weak attractive
interaction among the pair.

This interaction is known to be mediated by a virtual

exchange of a phonon between one electron in the pair, the crystal lattice of the
superconductor, and the other electron.

The quantum flux is; 0 = 2.0710-15 T-m2.

21

From the BCS theory it is possible to investigate the nature of the


superconducting ground state and its condensation energy relative to the normal state, the
exited quasi-particle states above the energy gap ( )i, and the temperature dependence of
the gap.

2.2 Type II superconductivity and the mixed state


Many of the superconductors studied after the discovery of superconductivity
were made of pure elements and were studied in bulk form. Their magnetic response as a
function of temperature was found to be given by Equation 2.1.
In 1930, W. J. de Hass and J Voogd discovered that for the alloy lead-bismuth
(Pb-Bi), Hc was several times larger than that of its constituents. This technologically
important discovery was explained in 1957 by Alexei Abrikosov

[35]

, when he

investigated what would happen in the case when > .


Starting from the G-L theory, Abrikosov reasoned that for large there should be
a negative surface energy at the interface between a normal and a superconducting
domain, concluding that it will be energetically favorable for the superconductor to break
into domains rather than to exclude the magnetic field completely from inside its volume.
The exact value for that separates these two distinct behaviors (i.e.,
superconductors that behaved as in the Meissner effect and those that will break up into
domains) is: =

1
.
2

Materials that have > 2 (as is the case for pure bulk

For Nb 2 (0)= 3.05 meV [25]

22

materials) are known as type I superconductors, and materials that present < 2 (such
as alloys or dirty materials are known as type II superconductors.
Figure 2.4 summarizes the magnetic behavior for type II superconductors.

Hc2 (T)

Hc2o

Normal
State
B 0
(Vortex or Mixed State)

Hc1 (T)
Hc1o

B = 0 (Meissner State)

Tc

Figure 2.4: H-T phase diagram for a type II superconducting slab in a uniform
magnetic field.

From Abrikosovs analysis it is known that in the so-called mixed state, the
magnetic field penetrates in a regular array of flux tubes, each carrying a quantum of flux
0 . Each flux tube is called a vortex and it has a circulating supercurrent concentrating
the quantum flux in the center of the vortex. The vortex array with the lowest free energy
that forms in the surface of the superconductor is a triangular array. These vortices are
also known as Abrikosovs vortices.

23

2.3 The Josephson junction effect


In 1962 B. D. Josephson

[36]

suggested that it was possible for a super-current to

flow between two superconducting electrodes separated by a thin insulating barrier at


zero voltage.

Soon afterwards, in 1963, experiments conducted by Anderson and

Rowell[37] suggested the observation of the so-called Josephson Junction Effect (JJE).
The JJE is summarized by the following two equations:
Equation 2.20

is = ic sin ( 21 )

and

Equation 2.21

V =

h d ( 21 )
2e dt

where,

Equation 2.22

2
2e v v
21 = 2 1 A dr
h 1

is the gauge-invariant phase difference of the Ginzburg-Landau wave-function in the two


r
superconductors at each side of the separation, A is the vector potential of an external
magnetic field threading the junction, ic is the zero voltage critical current, and V is the
dc voltage drop at each side of the separation.
If a dc voltage is applied to a JJ, integrating Equation 2.21 will show that the
phase difference will increase in time at a rate given by the magnitude of the voltage.

24

Then the supercurrent in the junction will oscillate at a frequency (known as the
Josephson frequency) given by:

c =

Equation 2.23

2e
V
h

The energy stored in the supercurrent in a Josephson junction is given by the time
integral of the product of Equation 2.20 and Equation 2.21. This gives:

Es =

Equation 2.24

h
ic cos( 21 ) = E J cos( 21 )
2e

where EJ is the Josephson coupling energy.


The JJE can be observed in a large number of superconducting systems that are
weakly coupled, for example: superconductor-insulator-superconductor (SIS) junctions,
superconductor-normal-metal-superconductor

(SNS)

junctions,

superconductor-

constriction-superconductor (SCS) junctions (i.e., a micro-bridges in a superconducting


film), and superconductor-weaker superconductor-superconductor (SSS) junctions [19].
A weak link is usually modeled using the resistively-capacitively-shunted
junction (RCSJ) model

[38, 39]

; in this model, a real weak link is represented by an ideal

one shunted by a resistance R and a capacitance C, as is shown in Figure 2.5.

25

i
ic

R0

Figure 2.5: RCSJ model representation of a real JJ. The total current i into the
junction is divided into three channels: the supercurrent (ic) through the ideal
junction, the current through the junction capacitor (C), and the quasiparticle
current through the resistor (R0).

Using the RCSJ model, the following loop equation can be written by applying
conservation of current in the loop:

Equation 2.25

i = ic sin ( 21 ) +

V
dV
+C
R0
dt

Substituting V from Equation 2.21 in this equation, the following second order
differential equation is obtained:

Equation 2.26

h d 21 hC d 2 21
+
i = ic sin ( 21 ) +
2eR0 dt
2e dt 2

defining;

Equation 2.27

p =

2ic
0C

26

which is known as the plasma frequency of the junction, the dimensionless time variable
is defined as: = p t , and the quality factor Q is defined as: Q = p RC . The quality
factor is identical to

c , which is a damping parameter introduced by W. Stewart[38]

and D. McCumber[39] (known as the McCumber constant).


Depending upon the value of the capacitance, a JJ can be classified as
overdamped (when C is small so that Q<< 1 as in the case of a SNS JJ and microbridges) or underdamped (when C is large so that Q>>1 as in the case of SIS JJ). The
systems studied in this dissertation belong to the overdamped category and it is
customary in simulations of these types of weak links to drop the term with the
capacitance in Equation 2.26 (in this case the RCSJ model is referred to as the RSJ
model).
Equation 2.26, is identical to the equation of a driven-pendulum where the applied
torque

[40]

is analogous to the current source, the maximum gravitational torque is

analogous to the critical current, the damping constant is analogous to the inverse of the
resistance, the momentum of inertia is analogous to the capacitance, the angular velocity
is analogous to the voltage, and the angle is analogous to the phase difference in the JJ.

2.4 Shapiro steps


The voltage-current characteristics (VIs) of a JJ, when biased with both dc and rf
signal, show the appearance of step-like structures called Shapiro steps

[7]

. These steps

occur at voltage multiples of: 0 rf , where rf is the frequency of the rf signal.

27

The occurrence of these steps can be derived analytically from the Josephson
junction equations when the junction is biased with both dc and rf voltage sources. For
simplicity it is assumed that the voltage sources are ideal (i.e., they have no internal
impedance). The following expression gives the total voltage bias for the JJ:
Equation 2.28

V (t ) = Vdc + Vrf cos(2 rf t )

Integrating Equation 2.21 to obtain 21 (t ) and substituting this value into Equation 2.20,
an expression for the supercurrent is obtained as a function of time. Then, by using the
standard mathematical expansion of the sine of a sine in terms of Bessel functions, we
have:

Equation 2.29

i s = ic (1) n J n (

Vrf
0 rf

) sin[2 ( c n rf )t + 0 ]

where 0 is a constant of integration and J n is the nth order Bessel function of the first
kind. From Equation 2.29 it can be seen that for average dc voltages such that c = n rf ,
there will be current steps in the VIs that have a maximum width given by:

Equation 2.30

[in ] MAX = ic J n (

Vrf
0 rf

This result, which is derived analytically for the case of a voltage biased JJ, has been
found

[41]

to occur in practice (when the JJ are usually current biased due to their small

impedance compared to that of the source) but the dependence of the steps widths are
different from the Bessel function dependences found above.

28

In physical terms, we can interpret the occurrence of Shapiro steps in junctions by


recalling that the supercurrent through the junction oscillates at a frequency given by
2V
for any applied dc voltage V. If an rf signal is coupled to the junction, this signal
0
will interfere with the oscillating supercurrent and the time average values of the total
supercurrent in the JJ will vanish unless c = n rf .

2.5 2D Josephson junction arrays


Consider now a set of superconducting islands at the corners of a square lattice
with lattice constant a, and that each island forms a JJ with their nearest neighbors (i. e.
there is a JJ on each vertical and horizontal bond of the square lattice). Such a structure
is known as a Josephson junction array (JJA). The characteristics of a JJA depend upon
the number of superconducting islands in the array. A square lattice with N+1 columns
and M+1 rows will have N JJ in the horizontal direction and M JJ in the vertical
direction, forming an NM JJA.
In experiments, it is customary to bias these samples with a current flowing along
the rows of the lattice and the voltage is measured between the first and last columns of
the array. If an external rf signal is superimposed on the JJA, each individual JJ will
couple (phase lock) to the external rf signal (when each individual JJs Josephson current
oscillates in phase with the external rf signal) and a voltage plateau will form in the timeaveraged VIs at a voltage which is N times greater than that of an isolated JJ. This is
given by the following equation:

29

Equation 2.31

h
Vn = nN rf
2e

In the literature the voltage steps that occurred in JJA are known as giant
Shapiro steps, to distinguishing them from the ones occurring in JJ. In this dissertation,
such distinction wont be emphasized since the samples studied here are all array types.
A JJA is not the only superconducting system on which Shapiro step structures
will occur if irradiated by an rf signal. In fact, any system with an equation of motion
that can be written in the form of the driven damped pendulum equation will present such
features in their characteristics.
For example, Martinoli et al. [42] found pronounced equidistant steps in the VIs of
a superconducting film with laterally modulated thickness for several matching magnetic
fieldsi and rf frequencies. This observation is interpreted to be caused by the coherent
motion of the Abrikosov vortex lattice in the one-dimensional periodic potential created
by the thickness modulation. Recently L. Van Look et al. [43] measured Shapiro steps in
the VI characteristics of a superconducting film with a square lattice of perforating
microholes (antidots) of lattice spacing d. They found that in order for the Shapiro steps
to occur, a magnetic field corresponding to 2

0
(the second matching field 0 H 2 ),
d2

should be applied to the sample. Under this specific condition of applied magnetic
induction and for particular temperatures

[44]

, it is possible to have a single vortex

occupying every antidot and a vortex lattice located at the center of the arrays cells.
This weakly interstitial vortex lattice can then be easily moved in a coherent fashion by

30

the applied rf currents through the potential created by the strongly pinned vortex located
at the antidot site. This weakly interstitial vortex lattice allows for the occurrence of
Shapiro steps. Finally, Shapiro steps have also been observed in superconducting wire
networks operated very close to Tc [45, 46, 47].

This is also known as commensurability.

31

Chapter 3 : EXPERIMENTS
In this chapter, a description of the samples measured in this dissertation will be
given. Also, the apparatus used to maintain the samples in the superconducting state and
the experimental techniques utilized will be described.

Finally, the measurement

techniques, including the main algorithms for data acquisition will be described.

3.1 Description of samples


Two different types of samples were measured in this dissertation. The first
samples consist of an array of niobium (Nb) crosses deposited on top of a gold (Au) thin
film underlayer

[48]

These samples are known as superconductornormal-metal-

superconductor (SNS) Josephson junction arrays (JJA). There are N+1 columns of
crosses in series in the direction of the current and M+1 rows of crosses in parallel to
form an NM JJA. Figure 3.1 is a drawing of one of these samples.
To fabricate these samples, a 1500 thick Au film was thermally evaporated onto
a Silicon (Si) substrate and without breaking vacuum a 2500 Nb film was then sputtered
at a rate of 15/sec [49]

32

Figure 3.1: SNS Josephson junction array (not to scale) of 53 junctions. Current
is injected via the Nb bus bars to the sides of the array and the voltage is measured
across the whole array.

This multilayer sample was then taken to the Cornell Nanofabrication Facility (CNF) at
Cornell University for patterning using e-beam lithography to subsequently remove the
exposed Nb by reactive ion etching (RIE) in a CF4 atmosphere. The crosses fabricated
using this method have arms of width 1.2 m and are arranged in a square lattice 10 m
center to center. Samples were made with average gap distances of 0.3 - 0.5 m
between the arms of the crosses (variation in the gap distances, determined from
measurements of SEM pictures, are approximately 10% on average, due to the limitations
of the e-beam lithography + etch process).
There are Nb contact pads in the sample for transport measurements. Leads are
attached directly to the contact pads using indium as solder.

The 4-point room

temperature resistance of these samples is on the order of 100s of milli Ohms. The
contact resistance obtained by pressing indium is usually on the order of a few Ohms at
room temperature.

33

The magnitude of the contact resistance is usually a problem when doing transport
measurements in superconductors. Self-heating effects can be observed when applying a
large current bias to a sample (as is necessary for experiments with small gap JJA or with
superconducting films below Tc).
For most of the JJA samples studied, no current heating effects were observed.
This was due to the small sample resistivity and also, because the indium contacts
become superconductive at temperatures below 3.41K.
The second samples measured were provided by Dr. Ivan K. Schuller
Yvan Jaccard, and Dr. Axel Hoffman

[50]

, Dr.

[51]

. The samples consist of a square or hexagonal

array of nickel (Ni) or iron (Fe) dots deposited directly onto a Si substrate. The dots are
200nm in diameter, 40nm thick, and are separated 400nm center to center. The dot array
was deposited on an area 5050 m2 (i.e., there are 125 dots along each side of the
square).

34

VT+

I+

VL+

I-

VT-

VL-

Figure 3.2: Measuring bridge prototype for the film samples. The contact lead
names are assigned as shown. In this photograph, the brown square in the center is
where the magnetic dots array is located. The clear pattern is the Nb bridge for
transport measurements. The gray area is the Si substrate.

After the fabrication of the dot array, a Nb thin film (65nm thick) was deposited
on top of it. To perform transport measurements on the sample, a bridge 40 m wide and
50 m long was optically imprinted and subsequently etched on the Nb film. The
measuring bridge consists of a strip of Nb 40 m wide that runs over the magnetic dot
array and is used to supply the current to the sample. Also, six Nb strips (5 m wide)
extend perpendicularly out from the main Nb strip for voltage measurements. One pair
of contacts is used for longitudinal voltage measurements along the array, another pair
could be used for transverse voltage measurements, and the final pair is a backup pair in
case any of the other contacts should fail. One of them, the top right contact, and another
contact down to the right, not shown, were used for measuring voltage drop in the Nb
film).

35

The room temperature resistance of these samples is on the order of 10s of


Ohms. Current self-heating effects played a very important role in the measurement of
these samples, limiting the current bias used to no more than 1mA

[52]

. Due to this, the

experiments where performed very close to Tc (no more than 250mK below Tc).

3.2 Experimental setup


The experiments in this dissertation were performed on a standard liquid Helium
(LHe) cryostat. Some experiments on the JJA were performed in a continuous Helium
flow cryostat (Janis ST) manufactured by Janisi. The cryostat used for most of the
experiments consists of a straight bore Dewar (Cryofabii model CSM) and a probe that
was immersed directly into the LHe bath.
The probe was designed, fabricated and assembled at the University of Cincinnati
Physics department. The main parts of the probe are: vacuum can, He pot, needle valve,
positioning oxygen-free-high-conductivity (OFHC) copper rods, cold finger stage, weak
link, temperature controlling stage, and sample holder. Figure 3.3 is a picture of the
lower part of the probe

i
ii

Janis Research Co. Inc.; 2 Jewel Dr.; P.O. Box 696; Wilmington, MA 01887-0696.
Cryofab; 540 Michigan Ave.; P.O. Box 485; Kenilworth, New Jersey 07033.

36

Positioning
Rods

Sample
Holder

Cold
Finger

He Pot

Figure 3.3: Picture of the lower part of the probe without vacuum can, showing the
main parts.

The vacuum can was designed so its lower tail fits inside the borehole of a
superconducting magnet (Cryomagneticsi 6T magnet with 0.1% homogeneity over a
10mm diameter spherical volume (DSV)). Figure 3.4 shows a picture of the lower part of
the probe, vacuum can, and superconducting magnet.
This configuration provides a homogeneous magnetic field, normal to the
samples surface.

The magnitude of the magnetic induction was determined by

measuring the current sent to the magnet and calculating the magnetic induction using the
calibration factor provided by the manufacturer (738.5 Gauss/Amp). In addition, the
applied magnetic field was measured independently using a calibrated low temperature
Hall probe (Lakeshoreii HGCT-3020) situated inside the 10mm DSV of the magnet
(4mm above the sample, located outside the sample holder).

i
ii

Cryomagnetics; 1006 Alvin Weinberg Drive; Oak Ridge, Tennessee 37830


Lake Shore Cryotronics Inc.; 575 McCorkle Blvd.; Westerville, OH 43082.

37

Figure 3.4: Probes lower part, showing vacuum can and superconducting magnet.

The sample holder (made of OFHC copper) was specially designed by Dr. Mast
and his former graduate student Dr. Hyun Cheol Lee

[53]

to guarantee proper impedance

matching with the rf signals. The sample holder consists of two pieces: a chamber and a
cover. The chamber has four female bulkhead microdot connectors by Malcoi for dc
current connections and dc voltage measurements.

Also, it has two male SMA

connectors to allow the application of rf signals (these connectors can not be seen in
Figure 3.5, they are in the other side of the chamber).
The dimensions of the sample holder (width and depth of chamber) were carefully
chosen to keep the electromagnetic field configuration in the sample approximately
equivalent to that for a microstrip line [54].

Malco Technologies LLC.; 94 County Line Rd.; Colmar, PA 18915.

38

For dc measurements, gauge # 34 insulated copper wire was attached directly to


the samples by pressing indium into Nb pads in the samples. The copper wire was
wounded into a small inductor to prevent the rf signals from leaking into the dc
equipment. The small inductors were also thermally anchored to the chamber with GE
varnish to provide an additional thermal link between the chamber and the sample.

Figure 3.5: Sample holder chamber showing the electrical connection for current,
voltage and high frequency signals (showing Nb film sample ready to be measured).

To inject the rf signal to the samples, straight pieces of bare 0.5mm diameter Au
wire were soldered to the SMAs center conductor and indium was pressed onto the
samples current contact pads to provide a connection in parallel to the current contact
pads.

39

The samples were electrically isolated from the chamber and thermally anchored
to it by placing a piece of sapphire wafer with a thin layer of Apiezoni N-grease between
the sample, the sapphire wafer, and the sample holder. In order to prevent the sample
from springing out of the camber due to the pull of the wires, and to ensure a good
thermal contact between the sample, sapphire, and the camber, a flat piece of teflon with
a copper spring on one side was placed between the samples and the sample holders
cover. Finally, the cover was fastened to the chamber by using 2-56 brass screws and the
sample holder cell was mounted in the probe.
The temperature of the sample holder was controlled upwards from the
temperature of the cold finger stage. This was accomplished by sending current to the
heater (a non-inductively wound resistive wire attached to the heater block with GE
varnish) using a temperature controller (Conductusii model LT-21), and monitoring the
temperature with a calibrated Cernox thermometer called the control thermometer
(Lakeshore model CX-1070-SD-4L) in a PID control loop.
Using this method, the temperature on the sample holder was regulated to better
than 0.5mK when using our homemade cryostat. But only 10mK (or 2mK by using a
specially designed data acquisition program) when using the continuous flow cryostat.
Figure 3.6 is a detailed photograph of the temperature controlling stage and sample
holder ready for measurement.

i
ii

Apiezon Products M&I Materials Ltd; PO Box 136; Manchester M60 1AN, UK.
Conductus, Inc.; 969 West Maude Avenue; Sunnyvale, CA 94085.

40

Cold Finger
Control
Thermometer

Weak Link

H
H
Heeeaaattteeerrr

Temperature
Controlling Stage

C
C
Cooovvveeerrr
C
C
m
Chhhaaam
mbbbeeerrr

Sample
Thermometer

Figure 3.6 Picture showing in detail the temperature controlling stage and sample
holder

The temperature of the sample was monitored independently by using a second


calibrated Cernox thermometer called the sample thermometer (Lakeshore model CX1070-SD-4L), attached to the bottom of the sample holder chamber.

3.3 Measurements techniques and data acquisition


The measurements presented in this dissertation are dc transport measurements.
Specifically, the dc voltage drop in the direction of the current (longitudinal voltage) in a
4-point configuration as a function of dc current (Idc), temperature, magnetic field, rf
current amplitude (Irf), and rf frequency ( rf). For both types of samples, the longitudinal
voltage was measured using a Keithleyi Model 182 nanovoltmeter and the dc current was
provided by a Keithley Model 220 current source.

Keithley Instruments Inc.; 38775 Aurora Rd.; Cleveland, OH 44139.

41

For the case of JJA samples, the dc current was applied through a 10.7MHz low
pass filter (Mini-circuit BLP-10.7i). The dc voltage was measured through the same type
of filters. This was done in order to further filter rf signals from coming into the dc
equipment, in addition to the small inductors described earlier.
For the JJA samples the magnetic field was provided using a homemade straight
solenoid with a second Keithley 220 current source to bias this magnet. Due to the high
sensitivity of the SNS JJA samples to external magnetic fields, the cryostat was
surrounded by a double wall -metal shield to reduce the ambient magnetic field to less
than 10 mOe inside the shield and the remaining magnetic field at the sample location
was cancelled by the solenoid for measurements at zero magnetic field.
For the case of Nb films with magnetic dot arrays, no special arrangements were
made to shield the Earths magnetic field at the samples location. The external magnetic
field was provided by a Cryomagnetics 60kG superconducting magnet biased using a
Hewlett-Packardii (HP) 6038A current supply.
All of these instruments were individually controlled by a computer via the GPIB
bus.

Individual data points were collected using a series of LabVIEWiii programs

developed specifically to respond to the demands of the samples measured, equipment


available, or data to be collected. In the appendix, one of these main programs (vis)
used for the acquisition of the data is presented.

Mini-circuits, P.O. Box 350166, Brooklyn, NY 11235.


Hewlett-Packard Co.; Components Group; 370 W. Trimble Rd.; San Jose, CA 95131.
iii
Graphical programming language develop by National Instruments.
ii

42

The basic algorithms for data collection consisted in sending current to the sample
and measuring the longitudinal voltage while the temperature, magnetic field, and rf
amplitude and frequency were kept fixed.
Some of the data presented in this dissertation (specially the measurements
performed in the SNS JJA) were taken using a technique known as the delta
measurement. In this technique, the bias current to the sample is reversed in direction
and a pair of voltages is recorded for each direction of the current. Then the average of
the two measurements is taken as follow:

Equation 3.1

V (I , T , B, I rf ) =

V (I +, T , B, I rf ) V (I , T , B, I rf )
2

This method was used to cancel slow voltage fluctuations and voltage offsets present at
the contacts, and to mathematically enhance the sensitivity of the measurements by
averaging many of these measurements.
When the delta-method was not employed, voltage measurements at a fixed
current direction were buffered into the Keithley 182s internal memory and the average
of these voltages was calculated. This method was employed to avoid fast switching of
the current direction in the samples; especially at high currents (I~Ic(T,B)).
Avoiding the fast switching of the current direction may be particularly important
when taking transport measurements in the Nb film samples, since there is the possibility
that the bias current changes the magnetic orientation of individual dots in the array

43

[55]

The possibility of changing the magnetic orientations of the dots might be a problem in
these measurements, as explained in Chapter 5.
The disadvantage of not using the delta-method when measuring the longitudinal
voltage is that the sensitivity of the measurements is reduced. The voltage noise floor
increases from ~ 10nV to ~ 100nV since the slowly varying offset voltage at the contacts
is not cancelled during the short time that the longitudinal voltage measurements take
place. This increase in the noise floor does not represent a problem when determining the
voltage location for the occurrence of Shapiro steps, since for most of the rf frequencies
studied, Shapiro steps are expected to occur above the noise floor (according to Equation
2.31).

This reduction in voltage sensitivity, though, imposes limitations on the

determination of the Ic from the VIs of the samples, as is explained in section 5.1.3.
To measure VIs with rf signals, straight bare Au wires (0.5mm diameter) were
attached in parallel to the dc current leads. The rf signals were applied to the sample
through dc blocks as is shown in Figure 3.7.

44

rf In

rf Out

dc blocks

Inductors

Indium
contacts

V
I

Figure 3.7: Block diagram for the experimental set up showing dc and rf electrical
connections.

For the measurements in JJA, the rf signal was provided using the HP8656B
(100kHz-990 MHz) Synthesized Signal Generator. The amplitude of Irf coupled to the
arrays was determined by converting the detected power through the array to current
using a 50 load.
For the measurements in the Nb film, the rf signal was provided using the
HP8671B Synthesized CW Generator (2-18 GHz). The large critical current of these
samples required the amplification of the applied rf signals power.

This was

accomplished by using the HP8349B microwave amplifier (2-20 GHz). This amplified
signal was then passed through the attenuator box, HP11713A in order to change the
amplified rf signal by one dB increments. The power of the rf signal passed through the
sample was finally measured using the HP437B power meter with the HP8485A Power
Sensor and this power was converted to current using a 50 load. Applying this high
power rf signal to the sample was found to increased the temperature of the sample stage
by about 4mK (when compared to the zero rf signal case). For this reason, and in order

45

to least affect the state (magnetic or superconducting) of the films, the measurements of
VIs with rf signals were taken from high power (less attenuation) to less power (more
attenuation) signals, as will be shown in section 5.1.4

46

Chapter 4 : MEASUREMENTS IN SNS JJA


In this chapter, a summary of the important results from the research in 2D SNS
JJA is presented. In section 4.1, the resistive transition of SNS JJA, their magneto
resistance curves (RBs), their VIs with and without rf signals, and the effects of
disorder on the VIs characteristics is described. In section 4.2, a brief discussion of the
difficulties encountered in this research is given and in section 4.3, a summary for this
chapter will be given.

4.1 Measurements
As mentioned in section 3.2, most of the measurements on these samples were
performed in a Janis ST continuous flow cryostat. Specifically, all of the VIs with rf
data presented in this section were taken using this cryostat. The temperature control and
stability achieved with this configuration was on the order of 10mk.
The rest of the data present in this chapter was taken in a similar probe and
cryostat to the one described in section 3.2.

The temperature control and stability

achieved with that system was better than 2mK. This early probe had to be replaced
because it developed a cold leak that was impossible to repair.

4.1.1 R vs. T
Figure 4.1 shows a typical resistive transition curve for a SNS JJA:

47

Sample: 300 X 50 W6
300m

10/15/97
Earth's B field
Temperature controlled better than 2 mK
Current bias= 10 A

Proximity-Effect
Model

R ()

200m

100m

LAT Theory

TcNb 9.2K
0

Tc0 3.2K
1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

10.0

11.0

12.0

Temperature (K)

Figure 4.1: Resistive transition for a typical SNS JJA sample. Measurements were
taken with a bias current of 10A in ambient magnetic field. The resistance error
bars are shown in red and the line connecting the dots is a guide to the eye.

Here, the sample resistance is calculated as:

Equation 4.1

R=

V
I bias

where V is the average of N voltage points taken at a constant temperature and fixed
current bias, as calculated with Equation 3.1.
In Figure 4.1, three distinct regions are evident first, the initial resistance drop at
9.2K, signals the temperature at which the Nb crosses become superconducting (TcNb),
the size of the drop is proportional to the amount of underlayer Au locally shorted by the
superconducting crosses (throughout this dissertation, TcNb is defined as the temperature
that corresponds to a 10% reduction in resistance, from the high temperature resistance
plateau value).

48

Second, the region were the resistance decreases slowly with temperature (known
as the proximity effect region), is basically the resistance of the remaining Au underlayer
not shorted by the crosses, which is diminished by the diffusion of Cooper pairs into the
normal metal.
Finally, the resistance drop region from ~ 6K to Tc0, is where the crosses first
become phase-coupled and long-range phase-coherence effects are observed. In this
range of temperatures, the sample displays a broad magnetic field dependent transition to
zero resistance. It is believed that in this region, a vortex-unbinding transition model is
applicable.
Vortex-unbinding ideas were first discussed by Berezinskii [56, 57]. Kosterlitz and
Thouless

[58]

predicted that in 2D neutral superfluids, bound vortex-antivortex pairs

should spontaneously dissociate at a transition temperature (TKT). These ideas (KTB)


were then modified by Halperin and Nelson [59] and Beasley, Mooij and Orlando [60] to be
applicable to superconducting films with i [61] larger than the sample width and length.
Subsequently, these theories for superconducting films were modified by Lobb,
Abraham, and Tinkham

[62]

(LAT theory) to explain the resistive transition in arrays of

weak links.

4.1.2 R vs. B
A magneto resistance curve taken on the above sample is shown in Figure 4.2.
The resistance of the sample is defined in the same fashion as in Equation 4.1.

=2/d, where d is the thickness of the film

49

40m

Sample 300 X 50 W6

30m

R ()

20m

10m
9m
8m
7m
6m
5m
4m

10/16/97
Current bias = 200 A
CT = (3.000 +/- 0.000) K

3m
-5.0m -4.0m -3.0m -2.0m -1.0m

0.0

1.0m

2.0m

3.0m

4.0m

5.0m

Magnet Current (A)

Figure 4.2: Resistance vs. current passed through the solenoid. Measurements were
taken at a temperature T < Tc0 of 3.000K. The sample was biased with I < Ic of
200A. The error bars for the resistance are shown in red, and the line connecting
the dots is a guide to the eye.

In Figure 4.2 a series of sharp minima and some other shallow minima are
observed at equidistant values of current through the solenoid. These minima correspond
to values of applied magnetic induction (Ba) which are multiples of the magnetic
induction produced by a quantum flux passing through the area of a plaquettei i.e.:

Equation 4.2


B 0 = b 20
a

where b is a constant that depends on the geometry of the array (for square arrays b =1),
and a is the arrays lattice constant.
Similar periodic oscillations in R as a function of magnetic field are also observed
in the magnetic response of other JJA properties, such as Tc and Ic.

Plaquette is the name given to the basic array of four junctions locate at the bonds of a square.

50

A useful way to characterize the amount of external magnetic induction applied to


the array is by using the concept of frustration (f) (first introduced by Toulouse [63]). For
square arrays f is given by:

Equation 4.3

B a2
f = a
0

The sharp minima in the R vs. B curve are known to occur at integer values of the
frustration while the shallow minima occur at fractional values, defined in general as:
f =

p
, with p and q being small integers. The resistance minima, associated with
q

integer frustrations, are believed to be caused by a maximum in the array critical currenti
that occurs when the array is trying to minimize its energy while keeping the phase
constraints of fluxoid quantization.
This is similar to what happens when two JJ are connected in parallelii. In this
configuration, the sum of the supercurrent at each JJ, due to their phase differences, will
produce an interference pattern (analogous to the two-slit interference pattern in optics)
given by the formula [64]:

Equation 4.4

I m = 2 I c cos

where Im is the maximum current of the combination and is the applied magnetic flux
(it is assumed here that both junctions have the same Ic). The minima observed at
i
ii

Simply, one might think that if Ic is large, there should be less resistance in the array.
This device is known as a Superconducting QUantum Interference Device or SQUID.

51

fractional frustrations is believed to be associated with the formation of a checkerboard


pattern where one array vortex

[65]

is located at every other cell of the pattern of size


[66]

qq. This configuration was found by Teitel and Jayaprakash

, and others

[67]

, to

produce a relatively low total ground energy and a relatively high Ic for the pattern
configuration studied.
The appearance (shape and depth at a given temperature) of these fractional
frustration minima (f =1/2i shown in Figure 4.2) have been found also to be dependent on
the bias current through the array, disappearing for larger bias

[68]

. A full understanding

of the dynamics of the system, even for the extensively study case of f =1/2, remains
controversial.
Finally, in Figure 4.2, the sharpest minima corresponding to f =0, is shifted to the
right of the zero-current-through-the-solenoid mark. This is due to the ambient magnetic
field present at the location of the sample. It is customary to supply a small magnetic
field, using the solenoid itself, to cancel this ambient magnetic induction when zero
magnetic field measurements are desired.
Using the sample as a magnetometer, it is easy to determine the amount of
magnetic induction present at the location of the sample, and from that, calculate the
frustration in the sample. From Figure 4.2, the current range (I) between two adjacent
sharp minima is determined to be I=3.0mAi. Knowing that the lattice constant for this
square array is a=10m; the magnetic induction for one quantum flux (from Equation
i

When the frustration f =1/2 is applied to the system, the system is known to be fully frustrated because it
can not minimize all of its junctions energies (as calculated from Equation 2.24) simultaneously.

52

4.2), is: B 0 = 20.7 T or 207mG. Since the shift from zero current (I) is 100A, the
strength of the magnetic induction at the location of the sample is determined to be
0.69mG. This residual magnetic induction represents a frustration that can alternatively
be calculated as; f =

I '
1
. This mean that for the 30050 sample measured there
=
I 300

is only 1/6 of an array vortex present.

4.1.3 VIs
Figure 4.3 shows a VI curve taken on the same sample above, minutes before
performing the R vs. B curve shown in Figure 4.2.
This curve was taken before the RB in order to determine the level of current bias
to use. It has been found that in order to observe the half integer minima in the RBs, the
current bias should be fairly small, this is a current value that will barely give a nonzero voltage above the noise floor of the VI.

The current step used to generate Figure 4.2 was 100.000A. The current is known to the implied
resolution.

53

Sample 300 X 50 W6

Voltage (V)

100

10/16/97
T=(3.0001+/-0.0002)K
Ambient magnetic field
Data: IV300KB.dat
No rf signal

RN = (0.1581 +/- 0.0006)

50

Ibias = 200 A
0
0

200

400

600

800

1m

Current (A)

Figure 4.3: VI curve at ambient magnetic field. The curve was taken for T < Tc0 of
3.000K. The error bars for V are shown in red, and the line connecting the dots is a
guide to the eye.

This voltage value is most easily determined by plotting the VI in a log-log scale, as in
Figure 4.4.

RN = (0.1581 +/- 0.0006)

Sample 300 X 50 W6
100

Voltage (V)

10

10/16/97
T=(3.0001+/-0.0002)K
Ambient magnetic field
Data: IV300KB.dat
No rf signal

Ibias = 200 A

Noise floor cutoff

100n

10n

1n
10

100

1m

Current (A)

Figure 4.4: Same data plotted in Figure 4.3, but on a log-log scale. The error bars
for V are shown in red, and the line connecting the dots is a guide to the eye.

54

In Figure 4.4, the blue line at 400nV is the cutoff between the data below and
above the noise floor. This cutoff is chosen after examining the data and determining
where it can no longer be represented by a power law V Ia, where a is the slope of the
VI curve (in the log-log graph)i, at that location. The value chosen to bias the sample
(200A), gives a voltage drop, which is above the noise floor. This voltage is determined
from the data to be: V = (0.95 0.01) V

ii.

Other important array parameters, such as the total critical current (Ic) and its
normal resistance (RN), can also be obtained from Figure 4.3. In JJA, a good estimate for
the non-fluctuating critical current Ic, is given by the peak or maximumiii,

[69]

found

from a dV/dI vs. I curve calculated from Figure 4.3. This is shown in Figure 4.5. From
Figure 4.5, Ic = 540A. Another important value that is obtained from Figure 4.5 is the
vortex depinning critical current (Ic,v)

[70]

. Ic,v is defined as the current for which the

dynamic resistance stops being zero.

By studying curves such as Figure 4.4, researchers are able to study phase transitions (like KTB) in JJA
the error here is the standard deviation of N averages of individual voltage points as calculated using
Equation 3.1
iii
This inference follows from studies of the effects of thermal fluctuations in single JJ
ii

55

250m

Sample 300 X 50 W6

200m

10/16/97
T=(3.0001+/-0.0002)K
Ambient magnetic field
Data: IV300KB.dat
f = 1/300

Ic = 540 A

dV/dI ()

150m

100m

50m

I c , v = 140 A
Noise floor cutoff

0
0.0

200.0

400.0

600.0

800.0

1.0m

Current (A)

Figure 4.5: Dynamic resistance curve calculated from Figure 4.3. The error bars for
dV/dI are shown in redi, and the line connecting the dots is a guide to the eye.

From Figure 4.5, Ic,v=140A. This value was chosen because it was the first current
value above the noise floor.
It has been estimated

[70]

that for square arrays, and in the case of very small

external magnetic induction, the depinning critical current is:


Equation 4.5

I c , v = 0.1I c

The value for Ic,v found in here, is not the value given by Equation 4.5. This discrepancy
might be associated with the lack of sensitivity of the voltmeteri.
The normal resistance of the array (RN) is determined by doing a linear fit to the
VI for current values higher than Ic. A good way of choosing what points to use in the
i

The dynamic resistance error in here was calculated as: dV * V


dI V

56

linear fit is by selecting, from Figure 4.5, data points to the right of Ic, at values for which
the dynamic resistance is fairly flatii.

From Figure 4.3, a linear fit gives:

RN = (158.1 0.6)m .iii


The normal resistance of the array and its critical current are essential parameters
that will determine the ability of a JJA to perform as a microwave oscillator. From these
two values the Josephson Frequency c (given by Equation 2.23) can be calculated.
The single junction parameters R0 and ic, required to calculate c, can be obtained
for a NM JJA, by assuming the JJA to be a lattice of resistors with resistance R0 and that
the critical current flows uniformly throughout the whole arrayi. This means:
Ic
M

Equation 4.6

ic =

Equation 4.7

R0 =

M
RN
N

Thus, the Josephson frequency, in terms of the arrays measurable parameters is


given by:

Equation 4.8

c =

1 I c RN
0 N

Using the values found for this array, its Josephson frequency is: c~138MHz.

In chapter five, a 9G ambient magnetic induction was found to be produced by the temperature controller
heater at the location of the sample. This magnetic field will produce a frustration f 43, which is enough
to be in the non dilute limit and to make Equation 4.5 not applicable in this case.
ii
In the data shown in Figure 4.5, the flat range of dV/dI has not yet been reached.
iii
In here the error in the normal resistance is the error obtained from the linear fit

57

A JJ is a frequency-to-voltage converter with the conversion factor of

[71]

2e
GHz
= 483593.420
. This oscillators performance will depend on c since at these
h
V NBS
frequencies a JJA would be capable of transmitting or coupling its greatest power.
Benz and Burroughs

[72]

found that the maximum power an NM JJA could

transmit to a matched load is:

Equation 4.9

PMAX =

RN I c2 NM
=
R0ic2
8
8

Using the values found for this array, the maximum power from Equation 4.9 is:
PMAX~6nW. This power might be too small for any applications, but from the equations
it is conceivable in principle, to increase the power transmitted or coupled by a JJA by
either increasing the arrays dimensions, its normal resistance, or its critical current up to
the energy of twice the gap for the superconductor [23], where nonequilibrium effects will
be important.

4.1.4 VIs with rf


Superimposing an external rf signal on the JJA while taking an VI curve, will
show the appearance of Shapiro steps in the VI, as seen in Figure 4.6. As mentioned in
section 2.4, Shapiro steps occur when the dc voltage biased on each individual JJ produce
a Josephson oscillation that will have the same frequency as the external rf signal. When
this happens the nearly identical JJ in an array will phase lock to the rf signal, to produce

This is a possible situation at zero temperature and zero magnetic field

58

nearly flat steps located at voltages that are N times the voltage of an individual JJ. The
width of these steps is proportional to the amplitude of the rf signal.

Voltage / V1

Sample: 300 X 200 W6

Zero rf
Irf=1.4mA

T=1.8K
rf = 50 MHz
f=0

n=3

n=2

n=1

Vn= n (N0rf )= n (31V)


0
0.0

0.5

1.0

1.5

2.0

2.5

3.0

Current (mA)

Figure 4.6: Current vs. normalized-voltage graph for a 100% 300200 JJA at
T=1.8K and zero magnetic field. The black curve is for no rf signal. The red curve
shows Shapiro steps when rf signal is applied. The voltage separation between steps
is proportional to the rf frequency and the step width is proportional to the rf
power.

For fixed external parameters such as the temperature and frustration, these
Shapiro steps can be characterized as a function of the applied rf signals amplitude and
frequency. This characterization consists in measuring the width of the step (I) as
well as its spread (V)i for different samples.

There are many factors that will

determine the shape and location of Shapiro steps. For example, for perfect JJA
(known in here as 100% arrays i.e. samples where none of their crosses have been
removed), the voltage location of the step and its width depends on the size of the
samples (NM).

59

Due to the arrays fabrication process, it is impossible to create identical JJ in any


100% array. All arrays are subject to a certain degree of difference among their R0 and ic.
These types of defects, known here as bond disorder, diminish the ability of arrays to
keep a coherent response to the external rf signal. The disruption of the phase coherence
can be seen in the VIs as steps with rounded corners and considerable spread as the rf
signals amplitude and frequency increases (as can be seen for the n=3 step in Figure 4.3,
where the spread is larger than for the case of n=1 or 2).
Other types of defects, known here as site disorder, occur when junctions are
missing, or have been intentionally removed, from the array. This type of defect will
cause a greater disruption to the phase coherence than bond disorder defects. In section
4.1.7, results showing the effects of site disorder will be presented for samples where
10% and 20% of their crosses were intentionally removed from the array (known as 90%
and 80% samples respectively).
There are other factors that affect the ability of an array to respond coherently to
an external rf signal, these include:

Performing measurements in arrays near the islands Tc.

Performing measurements in arrays for magnetic fields that are strong


enough to drive the material forming the superconducting electrodes or
banks into their mixed state.

The spread of the step refers to how horizontal the step is (V = 0), or if it has a slope (V 0). Ideally
we want zero spread at a step location so, at the voltage of the step the dynamic resistance is zero.

60

Having an array with a non-sinusoidal current-phase relationship (CPR),


which occur in other weak link structures, such as Dayem bridges, or in
SNS JJA with gaps smaller than the normal-metal coherence length n
where a large critical current can exist.

The factors listed here, affect the superconductivity in the banks that make the JJ.
Under these conditions, each bank can no longer be considered as having a single phase
and there will be a considerable amount of quasiparticles present at the banks. Also, the
possibility exists for Meissner vortices to be present in the banks.
When JJA are operated under these circumstances, they are known to be working
in the nonequilibrium regime [73].
Care was taken here to work under zero external magnetic field conditions for
which each cross could be modeled as having a strong order parameter connected by a
weak link i.e., a purely sinusoidal CPR. Under these conditions, only bond and site
disorder are the main causes for disruption to the coherent response of the arrays to
external rf signals.
To characterize the effect of disorder the width and spread of the steps were
measured directly from the VIs using a graphical method. This rudimentary method
provided a qualitative way to do this characterization.

Another more sophisticated

method that can be used for this purpose involves calculating the dynamic resistance
from the VIs and plotting it vs. the arrays voltage or current. For example, a plot of
dV/dI vs. V will show a series of minima where each minimum will be located at the
voltage for which the Shapiro steps should occur (as given by Equation 2.31).

61

I
dV/dI

Sample: 300 X 200 W6


3.0m

2.5m

T=1.8K
rf = 50 MHz
Irf=1.4mA
f=0

70m
60m
50m
40m

1.5m
30m
1.0m

20m

500.0

0.0
-20

dV/dI ()

Current (A)

2.0m

10m
0
0

20

40

60

80

100

120

Voltage (V)

Figure 4.7: dV/dI vs. V curve shows a series of minima located at the steps voltage.
The plot shows dV/dI0 at the step location meaning that the step has non zero
spread.

This can be seen in Figure 4.7, where the VI curve with rf signal from Figure 4.6
is used to illustrate the method. As can be seen from Figure 4.7, the minima are located
at the step position. Also, the minimum values for the dynamic resistance at the step are
not zero, indicating the spread of the step is not zero (having a non zero dV/dI at a step is
another way to visualized that this array has bond disorder)
The spread could then be further estimated by taking another derivative of the
dynamic resistance and finding the inflection points in dV/dI. This is illustrated in Figure
4.8 for step n=1.

62

1.2m

I
dV/dI
d[dV/dI]/dV

Sample: 300 X 200 W6


Inflection points for dV/dI

60m

30m

dV/dI ()

Current (A)

900.0

T=1.8K
rf = 50 MHz
Irf=1.4mA
f=0

600.0

Inflection points for dV/dI

n=1

300.0
20

25

30

35

40

Voltage (V)

Figure 4.8: Detail from Figure 4.7 illustrating how to find the spread of the first step
by using the dynamic resistance of the VI, to indicate the center of the step, and the
derivative of dV/dI as a function voltage, to find its value.

Similar procedures, as shown in Figure 4.8, can be followed to find the width of
the step. The next sections, will show results for this 100% sample, which was obtained
using the direct graphical method to characterize the samples in the presence of rf signals.

4.1.5 Dependence of VIs with rf amplitude


As the amplitude of the applied rf signal is changed, for fixed rf frequency, the
width of the steps oscillates as a Bessel function. Figure 4.9 shows these oscillations for
steps n=0 and n=1.

63

500

n=1
n=0

Sample: 300 X 200 W6


[ I0 ]MAX

400

n ( A)

T=1.30K + T
rf = 50 MHz
f=0

[ I1 ]MAX

300

200

100

100

200

300

400

Irf (Arb.)

Figure 4.9: Widths for steps n=0 and n=1 as a function of applied rf amplitude.
Graph shows Bessel function like oscillations for the step widths. Notice how the
oscillations are approximately 180 out of phase. The n=0 and n=1 steps first
maxima ([ I0]MAX and [ I1]MAX) for small Irf, are later used to further characterize
the JJA irradiated by rf signal.

As mentioned earlier, Figure 4.9 is obtained from a series of VI curves, like the
one shown in Figure 4.6, where the width is measured directly from the plot. From
Figure 4.9, it can be seen that the maximum width for step n=1 ([ I1]MAX) is almost as
high as the maximum for n=0.
[ I0]MAX, for the case of zero applied rf signal to the array, will represent the
critical current of the array. But, in this case [ I0]MAX is smaller because it was not
obtained as the maximum of a dV/dI vs. I curve as defined earlier from Figure 4.5. As
the rf amplitude increases the steps widths diminishes to a point where the graphical
method can no longer be used to characterize the steps from the VIs.
Curves like the one shown in Figure 4.9 are repeated for different rf frequencies
to also characterize the steps as a function of frequency. The characterization continues

64

until the steps are no longer observed in the VIs (in the low frequency range), or the
spread is two large (in the high frequency range).
In the next section, the characterization of the steps as a function of rf frequency
is presented. The maximum width value for each step is used as an indication of the
ability of the array to maintain a coherent response (i.e., wide, horizontal steps) to the
external rf signal.

4.1.6 Dependence of VIs with rf frequency


The dependence of the maximum of the step width ([ I1]MAX) for steps n=1 is
shown in Figure 4.10.

Figure 4.10: Maximum step width for steps n=1 as a function of reduced frequency
for a 100% array (Figure 1 in reference [12]. Data points are obtained from similar
curves as shown in Figure 4.9. In here T=1.91 K and f 0.

65

This Figure shows how [ I1]MAX increases as a function of reduced frequency


( =

rf
). For frequencies higher than 1.8, [ I1]MAX reaches a plateau which is
c

maintained up to =4.9 and possibly for higher . This indicates the ability of the
arrays to oscillate coherently with the external rf signal as the frequency increases. The
saturation value of the steps at high frequency is understood [12] in terms of the single JJ
behavior, where the width of the steps oscillates as a Bessel function. But in contrast to
what is found in single JJ where the saturation is reach for =1, here it is obtained at
almost twice this value. This is consistent with what has been found by others [74] and by
simulations using the RSJ model

[75,]

. Higher order steps (i. e. n=2, 3, etc) were not

characterized here because they presented considerable spread (R0), as can be seen
slightly in step n=2 and more pronouncedly in step n=3 in Figure 4.7.
Continued characterization of the steps for > 4.9 were limited by the frequency
range and power output capabilities of the rf signal generator used (HP8656B).
In the next section, the coherence response to rf signals for arrays with site
disorder are presented.

4.1.7 Effects of site disorder


The effects of site disorder on the width and spread of Shapiro steps are shown in
Figure 4.11. As mentioned in the previous section, removing crosses from the array
cause a much greater disturbance on the coherence response of arrays to an external rf
signal, compared with the response of arrays with only bond disorder.

66

In Figure 4.11, the first Shapiro step is shown for a 300300 JJA in which 10% of
the crosses have been removed at random from the sample. An array such as this is
known as a 90% sample.
These arrays with site disorder were fabricated by Dr. Donald C. Harris, former
student of Dr. J. C. Garland, from OSU. Details on the fabrication of these arrays can be
found in Dr. Harris doctoral dissertation

[76]

. These site disorder arrays had a slightly

smaller gap (0.45 m compared to 0.50 m for the 100% samples shown earlier).
Measurements on these arrays were performed at temperatures T=3.34K and near zero
magnetic field.

2.0

IV with dc + rf bias, 90% Sample

V / ( N 0 rf )

1.5

1.0

n=1

V/ V1

0.5

I1 / Ic
0.0
0.0

0.2

0.4

0.6

0.8

I / Ic

Figure 4.11: Normalized voltage vs. current for a 90% sample (300300). The first
Shapiro step presents considerable reduction in its width and increase in its spread
when compared to 100% arrays.

In order to make comparisons between measurements on samples with site


disorder and 100% samples, careful considerations need to be taken in estimating the
current for the disordered arrays (for details see reference [12]).

67

Following the same procedure described previously, the width and spread of the
steps were characterized for these site disorder samples. This characterization procedure
continued until enough points were taken to plot a graph similar to Figure 4.10. This is
shown in Figure 4.12.

Figure 4.12: Normalized maximum Shapiro step width for the first step vs.
normalized frequencies, for two samples with site disorder (90% and 80% samples)
(from reference [12], inset in Figure 1). The lines are a guide the eye.

Figure 4.12 shows the width of the first Shapiro step maximums as a function of
normalized frequencies for two 300300 samples for which 10% and 20% of the crosses
were removed at random from the arrays (these samples are known as 90% and 80%
samples respectively). It is clear from the figure that the values for the maximum width
of the steps are smaller than for the 100% sample, indicating that site disorder disrupts to
a larger extent the coherent state in the arrays, and that the greater the amount of site
disorder, the greater the disruption. Also, no saturation was observed for the maximum

68

step width at higher frequencies and the steps tended to disappear (roll off) for >0.5.
This behavior is consistent with what was found from numerical simulations

[75]

. The

disappearance of the steps at higher frequencies is understood in terms of the non uniform
redistribution of the current through the array with defects

[77, 78]

. This can be seen by

considering that the current through the array must flow around a defect and that the
junctions near the end of the defect will carry much more current than the other junctions.
This will make those junctions at the end of defects oscillate at a higher frequency than
the rest of the junctions (since the current through them is larger, the voltage drop across
the junction will be larger, and hence the Josephson oscillation will be larger). This out
of phase oscillating supercurrent will interfere destructively with the oscillations of all the
other junctions and therefore will diminish the step width. As the frequency of the rf
signal drive increases, the amount of phase difference between the junctions will be
larger and the step will tend to disappear. If the number of defects is increased the
disruption effects will be more pronounced and the steps width will be smaller and will
disappear at lower frequencies, consistent with the observations.
The effects of site disorder are also observed in the spread of the first Shapiro step
( V1) for the 90% and 80 % samples, as can be seen in Figure 4.13. This figure shows
V1 measured directly from the VIs for the 100%, 90%, and 80% samples as a function
of normalized frequencies. V1 for the 80% sample is shown to have the largest rate of
increase (160V/ ), in comparison to the 90% sample (51V/ ) and the 100% sample
(which has almost no spread)

69

Figure 4.13: Voltage spread as a function of normalized frequencies for site disorder
samples (from reference [12], Figure 3). Results obtained for a 100% sample are
shown for comparison. The spread increases linearly with frequency as can be seen
from the linear fit to the data for the 90% and 80% samples. The line linking the
100% data points is a guide to the eye.

The spread of the step at a given , gives a measurement of the amount of


frequency band spread among the oscillator (each JJ) in the array. The rate of increase
in V1 indicates that as the frequency of the rf signal increases, the amount of
destructive interference (the frequency ban spread) in the array increases, which is
consistent with the observation. This is clear again, in the frame of current redistributions
ideas in arrays with site disorder discussed in the previous paragraph.
Therefore, Figure 4.13 indicates that as the frequency of the rf signal increases
Shapiro steps become less distinguishable from the VI, and the steps in the sample with
more site disorder will tend to disappear at lower frequencies faster in comparison with
what happens in less site disordered arrays.

70

4.2 Discussion
There are two other effects that influence the global phase coherence response of
JJA that need to be addressed (their influence is greater in the presence of rf signals).
These are: thermal effect and the influence of external magnetic fields.
Thermal effects are understood

[79]

in terms of a parallel noise-current source

component added to the current channel in the context of the RSJ model. The magnitude
of the noise current source is inversely proportional to the ratio of twice the Josephson
coupling energy (Equation 2.24) to the thermal energy (KBT) (referred in here as ).
Therefore the closer the experiment is carried out to Tc (where KBT>> 2EJ and 0 )
the larger the magnitude of the noise-current source and therefore the larger the
disruption to the coherent state in the samples (based on the arguments of current
redistribution given earlier). The magnitude of was calculated for the site disordered
samples (see reference [12] for details) and agreement was found to the overall
observations that for a larger the disruption in the step is larger (further
characterizations on the influence of the thermal effects needs to be performed in 100%
samples).
The effects of an external magnetic field can also be understood in terms of
current redistributions in the arrays, but now the current is redistributed due to the
appearance in the array of array vortices that are induced by the external magnetic field
in order to maintain the overall phase difference around a closed path a multiple integer
of 2. The larger the magnitude of the magnetic field, the more array vortices will be
created around individual plaquettes in the array and the more individual junctions will

71

be out of phase with the external rf signals frequency. The disruption effect that is most
noticeable is the one produced when the magnetic frustration in the array is not an
integer. The disruption will be the greatest for the case of f =1/2.
These experiments were carried out near a zero external magnetic field
condition which was obtained as explained in section 4.1.2. During the realization of the
experiments with SNS JJA, the existence of a large source of magnetic field was not
accounted to be present at the samples location (in section 5.1.2 the magnitude of this
extra magnetic field is determined to be ~ 9G). The disruptive influences of an external
magnetic field will be more noticeable in site disordered samples as has been found [75].
In these experiments though, due to the fact that they were carried out near an
integer frustration condition, these disruptive influences can be counted to be less, but
during the performance of the experiments these magnetic field influence were noticed,
and no satisfactory explanation was given at the time.

4.3 Summary
In this chapter transport measurements in SNS JJA with and without external rf
signals were presented. In the presence of rf signals, Shapiro steps were observed in the
VIs.
In 100% samples, the width of the steps, which is related to the power coupled
from the rf signal, was found to increase as a function of frequency and to saturate for

rf1.8 c and was maintained for rf frequencies larger than 4 c.

72

In the presence of defects, the Shapiro steps width did not saturate and
diminished (rolled off) for frequencies on the order of 0.5 c.

73

Chapter 5 : MEASUREMENTS IN NB FILMS


WITH MAGNETIC DOT ARRAY
In this chapter, transport measurements performed in Nb film samples with
magnetic dot arrays will be presented. Specifically, the data presented here are obtained
from a Nb film deposited on top of a square array of nickel (Ni) dots. These results will
be compared to those for the SNS JJA in chapter 4. In section 5.1, measurements of the
resistive transition, magneto-resistance curves, and VIs with and without rf signals for
this sample are presented. In section 5.2, a discussion of the model proposed in section
5.1 is given with more detailed explanations and interpretation of the data. A summary
of these results is given in section 5.3.

5.1 Measurements
All the data collected on the sample mentioned above (known here as the Nb/Ni
square sample) were taken using the cryostat described in section 3.2. Other data was
also collected on an Nb/Fe square sample and on an Nb/Ni hexagonal sample using the
first probe mentioned in section 4.1

5.1.1 R vs. T
A typical resistive transition curve for the Nb/Ni square sample is shown in
Figure 5.1.

74

Nb/Ni Square Sample

RT-10uA-f0-Is0-BoRF-121dB-A.dat

Temperature controlled better than 0.5mK


[8.10-8.50] K, T =0.01K
9/26/01
Ambient B-field
Ibias= 10 A

R ()

TcNb= 7.926K
3

Tc0= 7.766K

ST=7.860K
7.70

7.75

7.80

7.85

7.90

7.95

8.00

8.05

8.10

Sample Temperature (K)

Figure 5.1: Resistive transition for the Nb/Ni square sample. Measurements were
taken with a bias current of 10A in ambient magnetic field. The resistance error
bars are shown in red and the line connecting the dots is a guide to the eye.

The resistance of the sample plotted in Figure 5.1 was calculated in the same way
as in Chapter 4, using Equation 4.1. For this curve, the voltage across the sample was
also measured using the delta method, since the current used to bias the sample was
10A<Ic.
Figure 5.1 shows an initial resistance drop to zero resistance starting at
TcNb=7.926K and finishing at ST=7.860K, only 66mK wide. Under close inspection of
Figure 5.1, it was found that the resistance of the sample increased from a local minimum
(at ST=7.860K), reached a maximum (at a temperature known in here as TRMAX) then
decreased again until it became zero (within the experimental errors) at a lower
temperature Tc0=7.766K, as can be seeing in Figure 5.2.

75

0.20

Nb/Ni Square Sample

R ()

0.15

RT-10uA-f0-Is0-BoRF-121dB-A.dat

Temperature controlled better than 0.5mK


[8.10-8.50] K, T =0.01K
9/26/01
Ambient B-field
Ibias= 10 A

0.10

TRMAX

Most Shapiro step


data taken here

0.05

Tc0 = 7.766K
ST=7.860K
0.00
7.75

7.80

7.85

7.90

Sample Temperature (K)

Figure 5.2: Low temperature detail from Figure 5.1, showing the local resistance
minimum at ST=7.860K followed by an increase in resistance until it reaches a
maximum to drop to zero resistance at Tc0=7.766K. The resistance error bars are
shown in red and the line connecting the dots is a guide to the eye.

Similar behavior for the resistive transitions in this sample was observed in other
magnetic dot array samples.
The resistive transition shown in Figure 5.1 could be explained using a model in
which it is assumed that the stray magnetic field of the dots suppresses the
superconductivity of the Nb film right above, and around the location of the dots [5, 6].
Figure 5.3 is a drawing of the sample (to scale), where it is visualized the state of the
sample at a temperature which is located to the left of TRMAX in Figure 5.2. More
specifically, this temperature is the one marked in Figure 5.2 as Most Shapiro step data
taken here (T~7.8K).

76

Figure 5.3: Drawing for the Nb film sample grown on top a square array of
magnetic dots. The magnetic dots are the gray circles shown in the drawing. The
Nb film covers the whole array of dots (the Nb film right on top of the dots is not
shown for clarity). This drawing is to scale, where the diameter of the circles is
200nm.

In Figure 5.3, the Nb film covers the whole square array of magnetic dots, but the
Nb film right above a dot is not shown for clarity. The gray circles in the drawing
represent the magnetic dots, the light blue regions around the circles is where the
superconducting properties of the Nb film are thought to be suppressed by the dots stray
magnetic field, and the darker blue areas represent the Nb film which is less affected by
the magnetic dots. The substrate is represented by the dark yellow areas.
With this model in mind, the first resistive drop in Figure 5.1 is associated with
the superconducting transition of the Nb filmi. At the resistive minimum (ST=7.860K),
the Nb film above the magnetic dot array is not superconducting. At this temperature the
sample is believed to be in the mixed state, where small currents could cause flux flow
resistivity.

77

At temperatures below where the local minimum occurs, the resistance of the
sample increases. An alternative explanation (other than flux flow resistivity) for this
increase in resistance could be attributed to a process similar to the Kondo

[80]

effect, in

which a source of electrical resistance is produced by the scattering of normal electrons


with scattering centers that have a magnetic momentii. This process is thought to give a
contribution to the resistivity that increases with decreasing temperature.

Similar

behavior in RTs (to Figure 5.2) has been observed in the dilute iron alloy FexNbSe2,
where for zero applied magnetic field the alloy can become superconducting [81].
As the temperature continues to decrease the resistance stops increasing when the
portion of the Nb film at the magnetic dot interstitial area becomes superconducting (this
is represented by the darker blue diamond shaped area in Figure 5.3). At this point, a
process similar to the proximity effect in JJA takes place, and the resistance starts to drop
again when Cooper pairs couple across the normal regions between the dots.
If this model is correct, these types of samples can be visualized as a JJA for
temperatures to the left of the resistive maximum in Figure 5.2. In the following sections,
other transport measurements performed on these samples will be presented which show
the similarities between the Nb film samples and the JJA studied in chapter 4.

The regions of the samples with only Nb have a lower temperature transition than expected for pure Nb
possibly due to the processing of the samples
ii
Kondo discovered, from perturbation theory, that the magnetic scattering cross section is divergent,
giving an infinite resistivity

78

5.1.2 R vs. B
For the curves shown below, the temperature of the samples is controlled at
values which are below TRMAX. At this temperature, a small current is chosen to bias the
sample and the voltage across the sample is measured using the delta method. The
resistance across the sample is then calculated using Equation 4.1.
A typical magneto-resistance curve on these samples is presented in Figure 5.4.
A series of minima are observed that correspond to locations of magnetic field induction
that correspond to integer values of the frustration in the array.

RB820KNoRF-627uA-A.txt 12/21/01
RB820KNoRF-627uA-A.txt 12/14/01

Nb/Ni Square Sample

f=2
f = 1/2
f=0

f=1

R ()

1
0.9
0.8
0.7
0.6
0.5

Ba=116 Gauss

0.4
0.3

Ibias = 627A
ST=(7.8065 +/- 0.0005) K
ST=(7.8089 +/- 0.0003) K

50

100

150

200

250

300

B (G)

Figure 5.4: Resistance vs. magnetic field perpendicular to the sample. The graph
shows data for two slightly different temperatures below TRMAX. The current bias
to the sample (627A) is the same in both cases.

For this particular array, the lattice constant is: a = (400 10)nm

[82]

. Thus, from

Equation 4.2, B 0 = (129 7)G . However, the experimental magnetic induction value
found from Figure 5.4 is Ba = 116G

[83]

. Using a calibrated Hall probe the ambient

79

magnetic induction on the sample for zero current through the superconducting magnet
was ~ 9G. This extra 9G at the samples location was found to be in the same direction
as the magnetic induction provided by the superconducting magnet, so the difference
between Ba and B 0 cannot be attributed to the ambient magnetic field found at the
samples location. Further characterization of this sample in the presence of a magnetic
field should be performed in order to find the cause for this discrepancy. A possible
explanation is that the arrays lattice constant is a little larger than 400nm [83]. Also, there
exists the possibility that the plane of the sample was not completely perpendicular to the
applied magnetic field, but this possibility should give larger values of Ba not smaller.
The source of the ambient magnetic induction was determined to be the temperature
control
f =

heater.

Using

Equation

4.3,

the

frustration

created

by

9G

is:

9
3
=
0.07 .
129 43
This level of ambient magnetic field in the sample was not considered significant

in experiments on Nb film samples, but they are an important unaccounted factor in the
measurements on JJAi.
The magneto-resistance curves taken near 7.8K were found to be very sensitive to
small changes in temperature. In Figure 5.4, the small temperature difference (<3mK)
between the two curves, resulted in a difference in resistance on the order of 200m near
the first minimum. The two curves shown in Figure 5.4 were acquired on two different
days and no special procedures were taken to control the magnetic state of the dots
i

See section 4.2 for implications of having 9G of ambient magnetic induction in the measurements of JJA.

80

between cool-downs. It is possible that the magnetic state of the dots was different
between the two days and this difference could also contribute to the resistivity measured
in the sample. In this dissertation, the magnetic response of these samples was not
studied in any detail, this types of studies can be found else were in the literature [4, 51].

5.1.3 VIs
A typical VI curve for the Nb/Ni square sample is shown in Figure 5.5.

Nb/Ni Square Sample

4.0m

ST

IV-120dBm8200K-IMag0mA-Frq2000MHz-Att121dB-OFF-A.dat
12/14/2001 11:15 PM
ST = (7.8095 +/- 0.0003) K
Ambient B-field = (9.3 +/- 0.2) G

7.814

7.812

Voltage (V)

7.810

2.0m

Ibias= 627 A

7.808

Sample Temperature (K)

RN= (11.5 +/- 0.1)

0.0

7.806
0.0

200.0

400.0

600.0

800.0

1.0m

Current (A)

Figure 5.5: VI curve at ambient magnetic field. The curve was taken T ~7.8K > Tc0.
The error bars for V are shown in black. The red curve is the temperature of the
sample at each voltage point. The error bars for ST are shown in red. The line
connecting the dots is a guide to the eye.

In Figure 5.5, the resistance of the sample for currents greater than 800A is
shown under the label RN. The value of the current bias used for measuring the RB in
Figure 5.4 is: R N = (11.5 0.1) . The plot of the sample temperature vs. I, in Figure
5.5, shows a slight increase in temperatures for currents greater than 800A; this is

81

thought to be caused by heating effects in the current contact pads of the sample (as was
mentioned in section 3.3). In order to avoid this heating, the current through the sample
was limited to no more than 1mA. These VI measurements were taken using a method
for which the direction of the current was not reversed in order to possibly avoid
disturbing the magnetic dot orientation by the current bias [55].
For thin superconducting films and JJA samples, the critical current is often
defined as the value of current that will produce a fixed arbitrary low voltage in the
voltmeter (this is known as the threshold method). This threshold voltage was chosen to
be above the noise floor of the equipment, as is shown in Figure 5.6.

Nb/Ni Square Sample

1m

ST

IV-120dBm8200K-IMag0mA-Frq2000MHz-Att121dB-OFF-A.dat
12/14/2001 11:15 PM
ST = (7.8095 +/- 0.0003) K
Ambient B-field = (9.3 +/- 0.2) G
Ic [5V] = 65.8 A

7.814

7.812

Voltage (V)

100

7.810

10

7.808

Sample Temperature (K)

RN= (11.5 +/- 0.1)

Current at 5 V
defines IC

7.806
0.0

200.0

400.0

600.0

800.0

1.0m

Current (A)

Figure 5.6: Same data plotted in Figure 5.5, but on a log-log scale. The blue line
here represents the arbitrary value of voltage chosen to define Ic using the threshold
method. The line connecting the dots is a guide to the eye.

In order to compare results found on this sample to what was found for JJA, a plot
of the dynamic resistance from the VI curve in Figure 5.5 is shown in Figure 5.7.

82

Nb/Ni Square Sample


16.0
14.0

IV-120dBm8200K-IMag0mA-Frq2000MHz-Att121dB-OFF-A.dat
12/14/2001 11:15 PM
ST = (7.8095 +/- 0.0003) K
Ambient B-field = (9.3 +/- 0.2) G

12.0

dV/dI ()

10.0
8.0

IMAX= 800 A

6.0
4.0

Imin= 413 A

2.0
0.0
0.0

200.0

400.0

600.0

800.0

1.0m

Current (A)

Figure 5.7: Dynamic resistance curve calculated from Figure 5.5. The error bars for
dV/dI are shown in red (calculated in the same way as in Figure 4.5)

The large dynamic resistance fluctuation found around 800A in Figure 5.7 is due
to a sudden variation in the slope of the VI curve. These sudden variations should be
regarded as an artifact of the measurements.
In Figure 5.7 two current values are shown: the value at the maximum of the
dV/dI curve (IMAX=800A), and a current (called Imin=413A), which is the lowest
resistance before the dynamic resistance begins to increase. Using the values found for
IMAX, RN, and knowing that in the sample there are 125125 magnetic dots (i.e. there
could be a maximum of 124 junctions in the sample), the Josephson frequency is
calculated to be: c~36GHz.
In the next section this value for c is used as a starting point to look for evidence
of Shapiro steps occurrence in the presence of rf signals.

83

5.1.4 VIs with rf


According to the dc characterization shown in the previous sections, similarities
between the RT , RB and VI curves can be established for the Nb/Ni square samples and
the JJA samples. In order to investigate the extent of these similarities the dynamic
response of the Nb/Ni square samples, in the presence of rf signals was investigated to
see if Shapiro steps would appear in the VI characteristics.
If the value for the Josephson frequency found in the previous section
(c~36GHz) was used as a frequency for this study, then according to Equation 2.31 the
first Shapiro steps should be located at a voltage V1~9mV. However, as seen from
Figure 5.5, a voltage of 9mV in the VI will be obtained at dc currents that will be larger
than 1mA, where the heating effects will adversely affect the measurements. For this
reason lower frequencies were used instead. Notice that by using frequencies below c,
according to what was found for JJA in Figure 4.10, Shapiro step widths will be smaller
that their possible saturation value (~0.6 IMAX/M, where M is the number of rows
parallel to the current), which is reached for rf > 1.8 c.

84

12/14/2001 12:13 PM
IV10dBm8200K-IMag0mA-Frq2000MHz-Att30dB-ON-B.txt
5.5m
5.0m
4.5m
4.0m

Voltage (V)

7.820
7.818
7.816
7.814
7.812

3.0m

7.810

2.5m

7.808

2.0m
1.5m

7.806

1.0m

7.804

500.0

Sample Temperature (K)

3.5m

IAC= (123.0 +/- 0.3) A


ST = (7.8095 +/- 0.0003) K
4KT = (4.01 +/- 0.00) K
B = (9.3 +/- 0.1) G
Ic [5V] = 1.43 A

V
ST

7.802

0.0

7.800

-500.0
0.0

200.0

400.0

600.0

800.0

1.0m

Current (A)

Figure 5.8: VI curve for the Nb/Ni square sample at T~7.8K and ambient magnetic
field. The black curve is the VI with rf signal. The red curve is the temperature of
the sample. No Shapiro steps are apparent for rf=2GHz and Irf=123A.

Figure 5.8 shows a VI with a 2GHz, 123A rf signal applied to the sample. At
first glance, no evidence for Shapiro steps is apparent. However, when the dynamic
resistance of the VI curve is taken, a series of minima appear in the dV/dI vs. V curve.
This is shown in Figure 5.9.

85

12/14/2001 12:13 PM
IV10dBm8200K-IMag0mA-Frq2000MHz-Att30dB-ON-B.txt
6.0 I = (123.0 +/- 0.3) A
AC

ST = (7.8095 +/- 0.0003) K


4KT = (4.01 +/- 0.00) K
5.5 B = (9.3 +/- 0.1) G
Ic [5V] = 1.43 A

1.1m
1.0m
900.0

n=6

800.0

n=5

700.0

5.0

n=4

4.5

600.0

n=3

V6 = 2090.4 V [47]; N 84
V5 = 1709.1 V [40]; N 83
V4 = 1451.8 V [35]; N 88

n=2

4.0

V3 = 1208.0 V [30]; N 97
V2 = 755.8 V [20]; N 91

3.5

V1 = 346.4 V [10]
= 2 GHz
14
2e/h = 4.836 x 10
N = V1/ * 2e/h 84

500.0
400.0
300.0

Current (A)

dV/dI ()

V'

200.0
100.0
0.0

3.0
-100.0
-500.0 0.0 500.01.0m 1.5m 2.0m 2.5m 3.0m 3.5m 4.0m 4.5m 5.0m

Voltage (V)

Figure 5.9: dV/dI vs. V curve from VI with rf curve in Figure 5.8. The dynamic
resistance is shown in blue and the IV, in the black curve, is plotted for reference.

The voltage values for the minima found in Figure 5.9 were plotted versus the
number of the minima, starting with the first minima being n=1, this plot shown in Figure
5.10, shows that the values for the voltage minima follow a linear relationship when
plotted with respect to the number of the minima. From a linear fit to this data assuming
Equation 2.31 for an array is applicable, the number of junctions in the direction of the
current can be calculated. The value found from the linear fit is: N N = 84 3 , which
is different from the value expected for this sample (N=124).
difference is discussed below.

86

The reason for this

Nb/Ni Square Sample


From: Data of 12/14/2001 12:13 PM
IV10dBm8200K-IMag0mA-Frq2000MHz-Att30dB-ON-B.txt
2.2m

Vn
Fit

Vn = [N(h/2e)] * n

2.0m
1.8m
1.6m

= 2GHz

N +/- N = 84 +/- 3

1.4m
1.2m

Vn

1.0m
800.0

Linear Fit: Vn = A + B * n

600.0
400.0

Parameter
Value Error
---------------------------------A
4.17214E-5
4.88858E-5
B
3.46164E-4
1.35585E-5

200.0
0.0
-200.0
0

Figure 5.10: Plot of the voltage location for the minima found in the dynamic
resistance curve in Figure 5.9. These voltages are plotted vs. the number of the
minima staring with the first minimum at n=1, a linear fit of this data in shown. For
the calculations of the linear fit, the (0,0) point is included.

In the next sections, more measurements in the presence of rf signals are


presented.

5.1.5 Dependence of VIs with rf amplitude


Other VIs were taken for different rf amplitudes, but for the same conditions of
temperature, ambient magnetic field, and frequency, used in taking the VI shown in
Figure 5.8. From these VIs, the critical current (defined by the threshold method) was

87

plotted as a function of Irf this is shown in Figure 5.11. In this figure, the critical current
first increases at small Irf, reaches a maximum at Irf ~60A, and then monotonically falls
to zero at larger Irf. Also in Figure 5.11 is a plot of the sample temperature as a function
of Irf, which indicates an increase in sample temperature as Irf increases. It is quite
apparent that the critical current does not show the Bessel-function like dependence that
is observed in JJA under applied rf currents, see Figure 4.9.

180
160

Nb/Ni Square Sample, 12/14/01


V-I's as a function of RF power, Attenuation = [26-39] dB
= 2GHZ, CT = 8.200K, f = 0, Ambient B field 9 Gauss

Ic
ST
7.8106

Critical current determined


using the 5V criteria

39 dB

7.8104
7.8102

120

7.8100

100

7.8098

80

7.8096

60

7.8094

40

7.8092

30 dB Attenuation

7.8090

20

26 dB

Sample Temperaure (K)

Ic [will be taken as I0 here] (A)

140

7.8088
7.8086

-20
40

60

80

100

120

140

160

180

200

Irf (A)

Figure 5.11: Plot of the critical current (defined by the threshold method) vs. Irf.
The critical current is the black curve. A plot (in red) of the sample temperature is
shown to indicate the increase in temperature produced by the applied Irf.

88

In Figure 5.11, the Irf dependence of other Shapiro steps are not shown since they
were not observed from the VIs. Only the zero-steps width was unmistakably linked to
the applied rf signals amplitude. The point marked in Figure 5.11 with the label 30 dB
Attenuation, corresponds to the critical current found from the VI curve shown in Figure
5.8.
Strictly speaking, the critical current determined using the threshold method
plotted in Figure 5.11 is not the same as I0 determined using the graphical method in
Figure 4.9. And also, it is clearly not the same as the maximum in the dV/dI vs. I curve,
The critical current (defined by the threshold method) was used for Figure 5.11 because
many of the VI curves with rf signal where almost straight lines and also because this is
the usual value reported for Ic in the literature (although for the data presented in this
dissertation, the voltage threshold value used is at least one order of magnitude higher
[84]

).
To illustrate the difficulties encountered when trying to observe Shapiro steps the

dynamic resistance calculated from VI curves with different Irf, are plotted on the same
graph in Figure 5.12. Here the green curve is the one plotted in Figure 5.8, while the
other curves are for two different Irf above and below to the curve in Figure 5.8.
For all of these VIs, the critical current is near zero as shown in Figure 5.11.
These VIs were particularly chosen because of what is known to happen in JJA; the
maximum step widths for the first Shapiro step occurs near the zero values of I0.

89

Nb/Ni Square Sample


Data of 12/14/01
rf=2.000GHz

28dB
29dB

9.0

30dB

8.0

31dB

7.0

dV/dI ()

6.0

32dB

5.0
4.0

Irf=(149.5 +/- 0.1)A, 28dB Attenuation


Irf=(132.8 +/- 0.2)A, 29dB Attenuation
Irf=(123.0 +/- 0.3)A, 30dB Attenuation
Irf=(109.4 +/- 0.5)A, 31dB Attenuation
Irf=( 97.7 +/- 0.5)A, 32dB Attenuation

3.0
2.0
1.0
0.0
0.0

1.0m

2.0m

3.0m

4.0m

5.0m

6.0m

Voltage (V)

Figure 5.12: Plot of the dynamic resistance vs. voltage for five different Irf. The
green curve is the one plotted in Figure 5.8. The curves have been shifted a fixed
amount in the y-axis for clarity. The line connecting the dots is a guide for the eyes.

In Figure 5.12, plots with more curvature represent the VIs with more of a
superconducting current component in them. As the Irf increases the level of Ic in the
sample diminishes and the curve flattens (ohmic behavior). The minima in the dV/dI vs.
V curves shown in Figure 5.12 do not seem to appear at the same voltage as would be
expected in a JJA for fixed rf frequency. The only minima that seems to be a constant in
all five curves is the n=1 from Figure 5.8. Data for this step is shown in Table 5.1.

90

ST

Ic[5mV]

Irf

Attenuation

V1

(K)

(A)

(A)

(dB)

(V)

7.8093 0.0004

5.33

97.7 0.5

32

197

48

7.8094 0.0003

2.03

109.4 0.5

31

262

63

7.8095 0.0003

1.43

123.0 0.3

30

346

84

7.8096 0.0003

1.19

132.8 0.2

29

434

105

7.8098 0.0003

1.05

149.5 0.1

28

489

118

Table 5.1: Table showing the values for the voltage location for the n=1 minimum
found in Figure 5.12. The last column to the right is a calculation from Equation
2.31 of the number of junctions in the current direction (N) using V1 and rf=2GHz.

Table 5.1 suggests that the apparent number of junctions in the current direction
seems to increase as the amplitude of the rf signal increases.

This observation is

discussed in more detail below. Figure 5.13 is a plot of Ic[5V] as a function of Irf,
similar to the one in Figure 5.11; however, in this figure, the maximum in critical current
for small Irf is more evident. Similar enhancement of the critical current as a function
of rf signal has been known [85] to occurred in weak links of the type known as a Dayem
bridgesi, and it has been found to occurred for a limited range of temperatures and
frequencies. In Figure 5.13, the VI curve taken with attenuation 50dB was approximately
the same value of Irf as in the case of Figure 5.8.

It is a uniform thickness superconducting film with a constriction of the order of in one of its
dimensions.

91

250

Nb/Ni Square Sample, 12/14/01


V-I's as a function of RF power, Attenuation = [40-70] dB
= 2GHZ, CT = 8.200K, f = 0, Ambient B field 9 Gauss

Ic
ST
7.8130

Critical current determined


using the 5V criteria

200

7.8125
7.8120

I0 (A)

7.8110
7.8105
100
7.8100

70 dB

7.8095
50

51 dB Attenuation

7.8090

40 dB

Sample Temperaure (K)

7.8115
150

7.8085
7.8080

50

100

150

200

250

300

350

400

Irf (A)

Figure 5.13: Similar plot to Figure 5.12. The range [40-70] dB in the attenuator box
was found to deliver the same amount of Irf as in Figure 5.11. A maximum in the
critical current as a function of Irf is more evident.

Four VIs, taken with attenuation of 49, 50, 51, and 52dB, are plotted in Figure
5.14. In this figure, only the dV/dI vs. V curves labeled with 50dB and 51dB show a
minima found at low voltages (i.e. V1).
In some of the curves plotted in Figure 5.14, a very sudden change in dynamic
resistance can be observed. This type of change is similar to the sudden resistive jump
found in Figure 5.7. This type of changes in the dV/dI curve is regarded here as a

92

glitch in the data, which often occurred when the data acquisition program was delayed
by the computer.

Nb/Ni Square Sample


Data of 12/14/01
rf=2.000GHz

49dB
50dB
8.0

51dB

7.0
6.0

dV/dI ()

52dB
5.0
4.0

Irf=(132.7 +/- 0.2)A, 49dB Attenuation


Irf=(122.8 +/- 0.3)A, 50dB Attenuation
Irf=(109.2 +/- 0.5)A, 51dB Attenuation
Irf=( 97.6 +/- 0.6)A, 52dB Attenuation

3.0
2.0
1.0
0.0

1.0m

2.0m

3.0m

4.0m

5.0m

6.0m

Voltage (V)

Figure 5.14: Plot of the dynamic resistance vs. voltage for four different Irf. The
curves have been shifted a fixed amount along the y-axis for clarity. The line
connecting the dots is a guide for the eyes.

Table 5.2 shows the corresponding data set for Figure 5.14 as Table 5.1 for Figure
5.12. From this data, the increase in N as Irf increases is observed again. In JJA, the only
way V1 would increase for fixed rf frequencies is by having more junctions in the current
direction participating in the formation of the Shapiro step.

93

The small depression found in the dV/dI curve at low voltages was studied further
to see if it would change locations when VI curves were taken when only small changes
were made among the different curves. For example, the current step used to generate
the VI, or the range of current investigated, etc.
ST

Ic[5mV]

Irf

Attenuation

V1

(K)

(A)

(A)

(dB)

(V)

7.8089 0.0004

5.36

97.6 0.6

52

199

48

7.8090 0.0003

2.04

109.2 0.5

51

292

71

7.8091 0.0003

1.50

122.8 0.3

50

348

84

7.8091 0.0003

1.34

132.7 0.2

49

477

115

Table 5.2: Table showing the values for the voltage location for the n=1 minimum
found in Figure 5.14. The last column to the right is a calculation from Equation
2.31 of the number of junctions in the current direction (N) using V1 and rf=2GHz.

The curves in Figure 5.15 were taken under similar conditions of rf signal and
temperature, but with different current steps and current ranges. All these VIs were
taken for 51dB attenuation. In this figure the minima in the dV/dI curve are more evident
in the first (51D), second (51E) and last (51H) curve. Using values estimated for V1 from
Figure 5.15, N for these conditions is calculated to be: N N = 70 3 .

94

51D

Nb/Ni Square Sample


Data of 12/14/01
rf=2.000GHz

51E
51F

51G
6

dV/dI ()

51H
4

Irf=(109.2 +/- 0.5)A, 51D dB Attenuation


Irf=(109.4 +/- 0.2)A, 51E dB Attenuation
Irf=(109.5 +/- 0.1)A, 51F dB Attenuation
Irf=(109.5 +/- 0.1)A, 51G dB Attenuation
Irf=(109.5 +/- 0.1)A, 51H dB Attenuation

0.0

500.0

1.0m

1.5m

Voltage (V)

Figure 5.15: Dynamic resistance vs. voltage plot. Here the dV/dI was calculated
from five different VIs all taken with an attenuation of 51dB. The first curve on
top (51D) is the same curve plotted in Figure 5.13 for 51dB attenuation. The curve
at the bottom (51H) was measured by allowing the current to decrease after
reaching the maximum, then increase back again from zero and so on until
completing 3 round trips. In this plot the x-axis was limited to show data only up to
1.5 mV. The curves were shifted a fixed amount in the y-axis for clarity.

5.1.6 Dependence of VIs with rf frequency


Using the technique described in the previous section, a series of VI curves were
taken for four different rf frequencies. The results are shown in Figure 5.16. In this
figure a series of minima were observed at different voltage positions as a function of rf

95

frequency. For the curve at 3GHz a minimum can be observed if the y-axis scale is
enlarged.

5GHz, Irf=(110.3 +/- 0.1)A, ST = (7.8058 +/- 0.0003)K


4GHz, Irf=( 76.0 +/- 0.2)A, ST = (7.8064 +/- 0.0003)K
3GHz, Irf=( 86 +/- 1 )A, ST = (7.8063 +/- 0.0003)K
2GHz, Irf=( 22.6 +/- 0.4)A, ST = (7.8061 +/- 0.0003)K

Nb/Ni Square Sample


Data of 11/02/01

5GHz 16dB [A]


V1= 965.4V [27]

4GHz 51dB [A]

8.0

dV/dI ()

7.0

V1= 758.4V [38]

3GHz 46dB [A]

6.0

2GHz 51dB [A]

V1= 562.5V [25]

5.0

4.0

V1= 363.4V [20]

3.0
0.0

200.0

400.0

600.0

800.0

1.0m

1.2m

1.4m

1.6m

1.8m

2.0m

Voltage (V)

Figure 5.16: dV/dI vs. V curves as a function of rf frequency. The curves were
shifted a fixed amount in the y-axis for clarity.

The values found in Figure 5.16 marked with the label V1 are plotted as a
function of rf frequency in Figure 5.17.

96

1m
1m
900

Nb/Ni Square Sample

From VI's taken on 11/02/01


V1
Fit

V1 = [N(h/2e)] *
N +/- N = 93 +/- 1

800
700

V1 (V)

600
500
400
300
200

Linear Fit: V1 = A + B *

100

Parameter
Value Error
----------------------------------A
-1.02973E-5
1.00958E-5
B
1.92892E-13 3.07206E-15

0
-100
0.0

1.0G

2.0G

3.0G

4.0G

5.0G

RF Frequency (Hz)

Figure 5.17: Plot of the voltage points label V1 from Figure 5.16 as a function of rf
frequency. The (0,0) point has been included in the calculation.

The data follows a linear relationship. Using Equation 2.31, the number of
junctions found in the direction of the current is: N N = 93 1 . This value is
different from previous values.
It should be mentioned that if slightly different values of rf amplitude would have
been used in taking the VIs shown in Figure 5.16, it is likely that different values for V1
would have been found and the linear fit would have given a different slope, and hence a
different N than reported here. An explanation for this rf current dependence on N is
given below.

97

5.2 Discussion
From the study in SNS JJA it is clear that any disorder in the array will tend to
smear out the Shapiro steps from the VIs. Disorder in the Nb/Ni film samples can be
attributed as one of the causes for the weakness of the steps in the VIs, but these
samples has not been fabricated to be a JJA. So, rather than focus this discussion on the
effects of disorder (which are certainly a factor for the lack of Shapiro steps from the
VIs), the focus will be given to the origins of the Josephson oscillations present in the
Nb/Ni sample.
During this research, it became obvious that the electrical response given by the
samples, between thermal cycles, depended on their magnetic history

[5, 86, 87]

. This was

most evident in the measurements of transverse voltage (VT) performed in the Nb/Fe
square sample (these measurements were performed in the original LHe probe mentioned
earlier).
Figure 5.18 shows five VTI curves taken on this Nb/Fe square sample at the same
temperature (T Tc0) and ambient magnetic field, but under different conditions that
might have affected the magnetic state of the sample. The curves in black and red were
taken after cooling down the sample but before taking an RB curve; here the curves are
very noisy. The curve in green was taken after an RB; this curve is not as noisy as
before. The curve in blue was taken the next day; this curve follows the green curve with
no apparent difference. Finally, the curve in pink was taken after the sample was
demagnetized with a perpendicular magnetic induction (which changes according to the
formula:

98

B(t ) = B(0) sin(t )e

Equation 5.1

This curve is very smooth and different from the green and blue curve.

15.0

Sample: Square Nb/Fe Dot

Transverse Voltage (V)

08/26/97
T=6.350K
10.0
Controlled
@ "Zero" B field
5.0
0.0

-5.0
-10.0
-15.0
-20.0

IVTP635A.dat, 100 pts. 2 Avg. before taking a R vs. B


IVTP635B.dat, 200 pts. 2 Avg. before taking a R vs. B
IVTP635C.dat, 100 pts. 2 Avg. after taking a R vs. B
IVTP635D.dat, 100 pts. 2 Avg. taken next day
IVTP635E.dat, 100 pts. 2 Avg. taken next day,
-t/
after demagnetizing the samplke with B=B0e Sin(t)
0.0

200.0

400.0

600.0

800.0

1.0m

Current (A)
Figure 5.18: Transverse voltage vs. current in the Nb/Fe square sample.

The curves shown in Figure 5.18 are non-linear. The non-linearity in these VT vs.
I curves is yet to be understoodi. Various models have emerged to explain similar nonlinear curves observed in superconductors.
i

One explanation might be given by

See comments in chapter 6 regarding possible explanation for this curves and future work on this samples

99

considering the motion of vortices in the flux flow regime


might be given in the context of the Kondo effect

[88]

. Another explanation

[89, 90]

. Other possibility is in the

context of nonequilibrium superconductivity processes in mesoscopic SNS Josephson


junctions

[91]

, where the dynamics of the system are understood in terms of Andreev

reflections [92].
A possible argument that might explain why the curves shown in Figure 5.18 go
from being noisy to smooth can be given in terms of magnetic moment fluctuations in the
magnetic dots. To understand the argument, the following assumptions will be made
regarding the magnetic state of the dots: the magnetic dots are single domain and their
magnetic moments lie in the plane of the dots [93], the magnetic arrangement of the dots
magnetic moments in the array (for minimum energy) can be either ferromagnetic or antiferromagnetic [94].
Knowing this, the noise (the fluctuations in resistance) observed in the figure
above could be interpreted to be produced by the flipping of individual magnetic
moments trying to achieve either minimal energy configuration. Evidence for these
ferromagnetic or anti ferromagnetic alignments has been observed while taking magneto
resistance curves (similar to the one shown in Figure 5.4) on dot arrays with magnetic
moments randomly ordered. Figure 5.19 and Figure 5.20 are examples of these curves.
Here a jump or depression in the temperature of the sample has been observed when the
perpendicular magnetic field is an integer number of the magnetic field produced by a
quantum flux. A jump in temperature (or warming of the sample) could be associated
with the magnetic moments of the dots acquiring a ferromagnetic alignment (the dots

100

minimized their energy and the excess energy warmed up the sample), as seen in Figure
5.19, and a depression in temperature (or cooling of the sample) could be associated with
an anti-ferromagnetic alignment of the dots (in this case the adjacent magnetic moments
anti-align by acquiring the energy from the surroundings producing a cooling in the
sample), as shown in Figure 5.20.

R
ST

Niobium film on top of square array of Ni dots


07/24/01
Data: RB820KE.DAT
Ibias=650 A
Avg. ST=(7.8503 +/- 0.0002) K

7.855

7.854

R=V/I ()

7.852
7.851
7.850

0.1

Sample temperature (K)

7.853

7.849

50

100

150

200

250

300

7.848
350

External magnetic field (gauss)

Figure 5.19: Magneto resistance curve for the Nb/Ni square sample. The curve in
black is the magneto resistance of the sample (the error bars are shown in blue).
The curve in red is the temperature of the sample (the error bars are shown in dark
yellow). The error bars in this curve are the standard deviation of N averages taken
at each data point. The error of the mean value is at least 10 times smaller. An
increase of sample temperature is observed at f =2. This warming is thought to be
associated with a ferromagnetic alignment of the magnetic dots in the sample.

101

Further characterization is needed, with control over the magnetic ordering of


dots. It is important to mention that in Figure 5.19 and Figure 5.20, each individual point
is the average of many voltage readings taken using the delta method, and the number of
averages (N) varied from data point to data point. N depended on the absolute error

e =

V
V desired for the run; the average was continued until e was lower than a

desired error (usually 5%). If this accuracy was not achieved before reaching a fixed
maximum number of averages (usually set < 300) the average was stopped in order to
keep the measurements of individual points from taking more than 5 minutes. Each data
point took in the order of minutes to be acquired and the whole data set was acquired in
several hours (with the new probe, He was kept in the bath for ~ 8 hours).

102

R
ST

Niobium film on top of square array of Ni dots


07/30/01
Data: RB824KC.DAT
Ibias=500 A
Avg. ST=(7.8872 +/- 0.0001) K

7.887

0.1

50

100

150

200

250

300

Sample temperature (K)

R=V/I ()

7.888

7.886
350

External magnetic field (gauss)

Figure 5.20: Magneto resistance curve for the Nb/Ni square sample. The curve in
black is the magneto resistance of the sample (the error bars are shown in blue).
The curve in red is the temperature of the sample (the error bars are shown in dark
yellow). The error bars in this curve are defined as ( / N , the standard deviation
of N averages divided by the square root of N). A lowering in sample temperature is
observed at f =1. This cooling is thought to be associated with an anti-ferromagnetic
alignment of the magnetic dots in the sample.

Another explanation for the fluctuations observed in Figure 5.18 could be


attributed to an erratic flux flow motion of vortices (where the vertices can be pinned or
de-pinned by the magnetic dots or by other imperfections in the film). However, this
possibility would not explain the absorption or liberation of energy observed in the
previous two figures, unless the dots magnetic moments undergo the lattice ordering

103

transition helped by the interactions with the moving vortices (as well as by the
interactions suggested here i.e., the interaction with the external magnetic field and the
interaction with the magnetic field produced by current through the film, acting at the
location of the dots).
Regardless of knowing what are the mechanisms promoting the ordering of the
magnetic moments, these measurements seem to indicate that the magnetic moments
interact among each other and with the external magnetic fields present at their location,
in order to achieve a lower energy state, which is finally determined by the temperature
of the sample.
The magnetic order of the dots will ultimately determine what region will be
suppressed in the superconducting film above to a large or smaller degree. The film
between the dots that are suppressed more will be where Josephson junctions will be
created. Characterizations [93] of the dipolar interactions on Ni dots arrays had shown that
the magnetic order of the dots is maintained in the middle of the samples, but at the edges
the ordering is lost. Therefore, it is clear that the model of JJA proposed earlier for these
samples depends strongly on the magnetic orientation of the dots
The fact that the ordering may be lost at the edges of the dot array can be used to
explain why the number of junctions participating in the formation of the Shapiro steps
increases as a function of rf amplitude. But, first some words regarding JJA in the
presence of external rf signals need to be said. It has been found in JJA

[95]

that the

number of phase-locked junctions can be tuned by the amplitude of an external rf signal.

104

This has been discussed by Tilley

[96]

to be analogous to what happens in an optical

system of radiators.
It can be seen that for small or no rf amplitude, N will be small (there will be JJ
only in the center of the array where the dots are more ordered). As Irf increases, more
and more junctions are formed and N increases as has been found. The discrepancy
found with the number of junctions expected for these samples (N=124) and the
maximum found to contribute from the experiments (N=118) can be attributed to the
location of the longitudinal voltage pads with respect to the dot array. It can be seen from
Figure 3.2 that the longitudinal voltage pads are not centered with respect to the dot array
and this could lead for the lost of 101 junctions.
From these arguments, it can be understood the importance of keeping the dots
magnetic order stable from measurement to measurement.

To achieve this, the

temperature of the sample needed to be controlled to a high level of accuracy. Due to the
self heating effects only a limited parameter space in temperatures, rf signal amplitude,
and frequency were accessible for experimentation.
From the study of JJA the impossibility of seeing the steps directly from the VIs
can also be understood. The experimental requirements of taking measurements close to
TcNb (in reduced temperatures t=T/TcNb~0.98) made the system operate in the nonequilibrium regime (this is also evidenced by the enhancement of Ic with Irf). In this
region more complicated CPR than sinusoidal are expected. Also, the lack of control
over the coupling strength of the junctions and the possibility of encountering a different

105

magnetic dot configuration between measurements accounts for the weakening of the
steps widths.
Finally to have encountered regions where the system behaves accordingly to the
Josephson relations is given as indication of the validity of the model.

5.3 Summary
In this chapter, transport measurements for a Nb film sample deposited on top of a
square array of Ni dots were presented. The dc RTs, RBs, and VIs measurements
performed in this sample show that there is a similarity between this sample and a JJA. A
model was developed to explain these similarities.

The model is based on an

interpretation for the features observed in the RTs of the samples. In order to test the
model, measurements with applied rf currents were performed to see the formation of
Shapiro steps in the VI characteristics. According to the model the sample will behave as
a JJA only in a small window of temperatures. Indeed, VI measurements were in
agreement with Josephsons second relation but only within a very narrow range of
experimental conditions. Verification of these results was complicated by the interaction
of the magnetic dots moments, but once greater control on the system was achieved the
sample were found to follow the Josephson relations (the sample was magnetized at room
temperature with a 0.26T permanent magnet in the direction of the current and parallel to
the surface of the sample).

106

Chapter 6 : CONCLUSIONS
In conclusion, two different patterned Nb thin film systems were studied and
found to behave similarly. The first system (an Nb/Au SNS JJA), was intentionally
fabricated like a weak type II, 2D superconductor. The second system was a continuous
Nb thin film in which its superconductive properties were weakened by an array of
magnetic dots laying underneath.

In both systems, Josephson oscillations coupling

(Shapiro steps) of their order parameters to an external rf signal drive, were observed.
The Shapiro steps observed in site disordered SNS JJA, were found to be affected
primarily by the amount of disorder present in the samples. In Nb/Ni film samples with
magnetic dots array, Shapiro steps were observed only from the dV/dI vs. V curves and
the origin for the weakness of the steps was attributed to a combination of factors that
include: disorder (like in the case of the SNS JJA samples), but also, thermal influences
(nonequilibrium processes), since the measurements were carried out very close to Tc.
The origin for the evidence of Josephson oscillations found in the Nb/Ni film
samples was attributed to the presence of the magnetic dot array and therefore, any
condition that could affect the state of the magnetic dot array will adversely affect the
appearance of Shapiro steps in the array.
From this study a more fundamental question emerges: Why do the two different
systems studied in this dissertation have similar dc responses? The answer seems to be
related to being operating the film samples near Tc for which its characteristic length
scales diverge.

107

It is difficult to use electrical transport measurement techniques on bulk


superconductors since the critical current is large. The superconductor needs to be
weakened to a degree where a voltage can be generated for small transport currents. This
means being in the region where Josephson relations are applicable and in a region were
small magnetic fields can produce great changes in the magnetic response of the system.
There are many superconducting systems that show similar magneto-resistance
curves like the one shown in sections 4.2 and 5.2. These systems include: SQUIDS, JJA,
superconductive wire networks, superconductive films with thickness modulation,
superconductive films with array of holes, and superconductive films with array of
magnetic dots.
What all these systems have in common is that they are superconductors AND
that, when characterized, they are measured at a temperature for which the coherence
length (and any other characteristic length defined for a superconductor) diverge (i.e. the
superconducting state is weaken).
For continuation of this research, Shapiro step characterizations need to be
performed in the presence of perpendicular magnetic fields.

Also, control of the

magnetic state of the magnetic dots needs to be implemented in the measurements in


order to try to elucidate the relation between the magnetic order and the absorption or
liberation of thermal energy evidence observed during the measurements of RBs.
Also, the Hall voltage measurements need to be continued in order to elucidate
the origin for the non-linearity of these curves. There is growing evidence that the Nb/Ni
samples can be understood in terms of what happens in mesoscopic SNS JJ systems and

108

that the non-linear effects observed in the transverse voltage measurements might be
related to the existence of a -junction in the direction transverse to the direction of the
current.

Such evidence has been observed in the measurements of the transverse

resistance as a function of temperature for f 0, where right before the transition, an


unexplainable increase in resistance is observed, and right after the transition a negative
resistance region is present which could be related to the dynamic processes found in
mesoscopic SNS junctions [91].
Also, the self heating effects (which in this dissertation are regarded as to be
originated at the current contacts) observed in the Nb/Ni samples, might be related to the
same process describe by Skocpol et al.[17]. This might be evident from observations of
energy gap structures

[97]

present in the dV/dI vs. V curves observed in the straight line

VI curves when the sample was irradiated with a frequency of 20GHz and a rf power of
~1mW. In this curves a series of maxima where observed at voltages close (but less) to
3.0 mV and 6.0 mV that were independent of the frequency of the signal, although where
more pronouncedly observed for the higher frequency curves taken.
Finally, from this study, the possibilities for fabricating a JJA capable of
generating high power rf signals (according to Equation 4.9 and up to the energy gap) is
visualized by the combined used of nanotechnology fabrication and magnetism. The
basic junction structure for this JJA will be the coupling of two superconducting thin film
islands through a short (length ~ (T)), but wide bridge, where a magnetic dot structure
(as wide and long as the bridge), will be placed underneath or on top of this continuous
superconducting thin film.

109

According to what was learned in this dissertation, it is in principle possible that


this short but wide bridge will be capable of maintaining a sinusoidal CPR in the junction
better than a Dayem bridge, where all three dimensions (width, length and thickness)
must be comparable to (T) [13, 19], in order to have a sinusoidal CPR. This new structure
will be cooler than a Dayem bridge

[17]

and it is also possible that the presence of the

magnetic dot will improve the self heating problems of continuous film bridges by
allowing the structure to be thicker than (T).
The width of the bridge needs be determined experimentally (in order to obtain
the maximum power output with the least amount of phase coherence disruption). The
length of the bridge can be made shorter or longer depending upon the temperature of
operation intended for the structure and still maintained a sinusoidal CPR.

If the

temperature is close to Tc, the bridge could be made longer and in this way increase the
contribution to the power output from the normal resistance, but the proximity to Tc
might affect the superconductivity of the islands (due to the nonequilibrium processes
that take place near Tc) and the power out or coupled could be smaller (due to phase
disruption effects).
The ideal JJA made out of one of these structures will operate well below Tc and
will have a plaquette area as small as possible in order for the structure to be the least
affected by ambient or self induced magnetic fields, which have a negative influence on
the phase coherence in JJA [75].

110

Bibliography
[1] See, for example, K. K. Likharev, Dynamics of Josephson Junctions and Circuits
(Gordon and Breach, New York, 1986), pp. 41
[2] H. -C. Lee, D. B. Mast, R. S. Newrock, L. Bortner, K. Brown, F. P. Esposito, D. C.
Harris, and J. C. Garland, Physics B165-166, 1571 (1990)
[3] H. -C. Lee, R. S. Newrock, and D. B. Mast, S. E. Hebboul, J. C. Garland, and C. J.
Lobb, Phys. Rev. B 44, 921 (1991)
[4] J. I. Martn, M. Vlez, J. Nogus, and I. K. Schuller, Phys. Rev. Lett. 79, 1929 (1997)
[5] Y. Nozaki, Y. Otani, K. Runge, H. Miyajima, B. Pannetier, J. P. Nozieres, and G.
Filion, J. Appl. Phys. 79, 8571 (1996)
[6] Martin Lange, Margriet J. Van Bael, Yvan Bruynseraede, and Victor V.
Moshchalkov, Phys. Rev. Lett. 90, 1929 (2003)
[7] S. Shapiro, Phys. Rev. Lett. 11, 80 (1963)
[8] A. K. Jain, K. K. Likharev, J. E. Lukens, and J. E. Sauvageau, Phys. Rep. 109, 309
(1984)
[9] S. P. Benz and C. J. Burroughs, Appl. Lett. 58, 2162 (1991)
[10] A. B. Cawthorne, P. Barbara, and C. J. Lobb, IEEE Trans. Appl. Supercon. 7, 3403
(1997)
[11] H. -C. Lee, R. S. Newrock, and D. B. Mast, Helv. Phys. Acta 65, 371 (1992)
[12] K. Ravindran, L. B. Gmez, R. R. Li, S. T. Herbert, P. Lukens, Y. Jun, S. Elhamri,
R. S. Newrock and D. B. Mast, Phys. Rev. B 53, 5141 (1996)

111

[13] A. Baratoff, J. A. Blackburn, and B. B. Schwartz, Phys. Rev. Lett. 25, 1096 (1970),
and erratum: Phys. Rev. Lett. 25, 1738 (1970)
[14] L. D. Jackel, W. H. Henkels, J. M. Warlaumont, and R. A. Buhrman, Appl. Phys.
Lett. 29, 214 (1976)
[15] H. A. Notarys, M. L. Yu*, and J. E. Mercereau, Phys. Rev. Lett. 30, 743 (1973)
[16] C. M. Falco, W. H. Parker, and S. E. Trullinger, Phys. Rev. Lett. 31, 933 (1973), and
erratum: Phys. Rev. Lett. 31, 1476 (1973)
[17] W. J. Skocpol, M. R. Beasley, and M. Tinkham, J. Appl. Phys. 45, 4054 (1974)
[18] M. Octavio, W. J. Skocpol, and M. Tinkham, Phys. Rev. B 17, 159 (1978)
[19] K. K. Likharev, Rev. Mod. Phys. 51, 101 (1979)
[20] Konrad W. Lehnert, Ph. D. dissertation (University of California Santa Barbara,
1999)
[21] Jinhyoung Lee, Sangmin Lee, Jaejun Yu, and Gwangseo Park, Phys. Rev. B 53, 3578
(1996)
[22] See, for example, M. Tinkham, Introduction to Superconductivity, 2nd Ed.
(McGraw-Hill, New York, 1996), Chapter 11
[23] J. Bindslev Hansen and P. E. Lindelof, Rev. Mod. Phys. 56, 431 (1984)
[24] H. Kamerlingh Onnes, Leiden Comm. 120b, 122b, 124c (1911)
[25] B. W. Roberts, Survey of superconductive materials and evaluation of selected
properties, J. Phys. Chem. Ref. Data, Vol. 5, No. 3, pp. 581-821
[26] H. Kamerlingh Onnes, Leiden Comm. 139f (1914)
[27] W. Meissner and R. Ochsenfield, Naturwissenschaften 21, 787 (1933)

112

[28] F. And H. London, Proc. Roy. Soc. (London) A149, 71 (1935)


[29] C. J. Gorter and H. B. G. Casimir, Physics 1, 306 (1934)
[30] A, B. Pippard, Proc. Roy. Soc. (London) A216, 547 (1953)
[31] E.g., see, J.M. Ziman, Principles of the theory of Solids, Cambridge University
Press, New York (1964), p. 242
[32] V. L. Ginzburg and L. D. Landau, Zh. Eksperim. I. Teor. Fiz. 20, 1064 (1950)
[33] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957)
[34] L. N. Cooper, Phys. Rev. 104,1189 (1956)
[35] A. A. Abrikosov, Sov. Phys.-JETP 5, 1174 (1957)
[36] B. D. Josephson, Phys. Lett. 1, 251 (1962)
[37] P. W. Anderson and J. M. Rowell, Phys. Rev. Lett. 10, 230 (1963)
[38] W. C. Stewart, J. Appl. Phys. 12, 277 (1968)
[39] D. E. McCumber, J. Appl. Phys. 39, 3113 (1968)
[40] J. R. Waldram, A. B. Pippard, and J. Clarke, Philos. Trans. R. Soc. London, Ser. A
268, 265 (1970)

[41] Y. Taur, P. L. Richards, and F. Auracher, Low Temperature Physics-LT-13. Vol. 3,


(Plenum Press, New York, 1974), pp. 276-280
[42] P. Martinoli, O. Daldini, C. Leemann, and E. Stocker, Solid State Commun. 17, 205
(1975)
[43] L. Van Look, E. Rosseel, M. J. Van Bael, K. Temst, V. V. Moshchalkov, and Y.
Bruynseraede, Phys. Rev. B 60, R6998 (1999)

113

[44] E. Rosseel, M. J. Van Bael, M. Baert, R. Jonckheere, V. V. Moshchalkov, and Y.


Bruynseraede, Phys. Rev. B 53, R2983 (1996)
[45] B. Pannetier, J. Chaussy, R. Rammal, and J. C. Villegier, Phys. Rev. Lett. 53, 1845
(1984)
[46] H. S. J. van der Zant, M. N. Webster, J. Romijn, and J. E. Mooij, Phys. Rev. B 42,
2647 (1990)
[47] H. S. J. van der Zant, M. N. Webster, J. Romijn, and J. E. Mooij, Phys. Rev. B 50,
340 (1994)
[48] Details on the fabrication process of these samples can be found in Dr. Hyun-Cheol
Lees Ph. D. dissertation; University of Cincinnati (1991)
[49] Details on the Au/Nb deposition process can be found in Dr. D. C. Harris Ph. D.
dissertation; Ohio State University (1989)
[50] J. I. Martn, Y. Jaccard, A. Hoffmann, J. Nogus, J. M. George, J. I. Vicent, and I. K.
Schuller, J. Appl. Physics 84, 411 (1998)
[51] Details on the fabrication process of these samples can be found in Dr. Hoffmans
Ph. D. dissertation, University of California, San Diego (UMI#: 9917952)
[52] See Dr. Hoffmanns dissertation (UMI # 9917952) pg. 29
[53] Hyun-Cheol Lee doctoral dissertation, University of Cincinnati (1991)
[54] Samuel Y Liao, Microwave Devices and Circuits 2nd ed., Prentice-Hall, 1985
[55] F. J. Albert, J. A. Katine, and R. A. Buhrman, Appl. Phys. Lett. 77, 3809 (2000)
[56] V. L. Berezinskii, Zh. Eksp. Teor. Fiz. 59, 907 (1970) [Sov. Phys. JETP 32, 493
(1971)]

114

[57] V. L. Berezinskii, Zh. Eksp. Teor. Fiz. 61, 1144 (1971) [Sov. Phys. JETP 34, 610
(1972)]
[58] J. M. Kosterlitz and D. J. Thouless, J. Phys. C 6, 1181 (1973)
[59] B. I. Halperin and David R. Nelson, J. Low Temp. Phys. 36, 1165 (1979)
[60] M. R. Beasley, J. E. Mooij, and T. P. Orlando, Phys. Rev. Lett. 42, 1165 (1979)
[61] J. Pearl, Appl. Phys. Lett. 5, 65 (1964)
[62] C. J. Lobb, D. W. Abraham, and M. Tinkham, Phys. Rev. B 27, 150 (1983)
[63] G. Toulouse, Commun. Phys. 2, 115 (1977)
[64] See, for example, M. Tinkham, Introduction to Superconductivity, 2nd Ed.
(McGraw-Hill, New York, 1996), pp. 213-215
[65] See, for example R. S. Newrock et. al., The Two-Dimensional Physics of Josephson
Junction Arrays, Solid State Physics 54, 263 (2000), pp. 286
[66] S. Teitel and C. Jayaprakash, Phys. Rev. Lett. 51, 1999 (1983)
[67] T. C. Halsey, Phys. Rev. B 31, 5728 (1985), W. Y. Shih and D. Stroud, Phys. Rev. B
28, 6575 (1983), W. Y. Shih and D. Stroud, Phys. Rev. B 32, 158 (1985)

[68] M. Tinkham, D. W. Abraham, and C. J. Lobb, Phys. Rev. B 28, 6578 (1983)
[69] V. U. Ambegaokar and B. I. Halperin, Phys. Rev. Lett. 22, 1364 (1969); Erratum,
Phys. Rev. Lett. 23, 274 (1969)
[70] S. P. Benz, M. S. Rzchowski, M. Tinkham, and C. J. Lobb, Phys. Rev. B 42, 6165
(1990)
[71] B. F. Field, T. F. Finnegan, and J. Toots, Metrologia 9, 155 (1973)
[72] S. P. Benz and C. J. Burroughs, Appl. Lett. 58, 2162 (1991)

115

[73] See, for example, M. Tinkham, Introduction to Superconductivity, 2nd Ed.


(McGraw-Hill, New York, 1996), pp. 403-432
[74] Y. C. Hui, E. B. Harris, and J. C. Garland (unpublished)
[75] M. Octavio, J. U. Free, S. P. Benz, R. S. Newrock, D. B. Mast, and C. J. Lobb, Phys.
Rev. B 44, 4601 (1991)
[76] D. C. Harris doctoral dissertation, The Ohio State University (1989)
[77] W. Xia and P. L. Leath, Phys. Rev. Lett. 63, 1428 (1989)
[78] P. L. Leath and W. Xia, Phys. Rev. B 44, 9619 (1991)
[79] V. Ambegaokar and B. I. Halperin, Rev. Lett. 22, 1364 (1969)
[80] J. Kondo, Prog. Theoret. Phys. 32, 37 (1964); Solid State Physics, vol. 23, F. Seitz
and D. Turnbull, eds. Academic Press, New York, 1969, pp. 183
[81] D. A. Whitney, R. M. Fleming, and R. V. Coleman, Phys. Rev. B 15, 3405 (1977)
[82] For details see Dr. Hoffmanns Ph.D. dissertation (UMI#: 9917952), pp. 35
[83] Dr. Hoffmann found for the Nb/Ni square array, magnetic field oscillations in the
order of 100 Oe with an error approximately of 10%, private communications (1997)
[84] For details see Dr. Hoffmanns Ph.D. dissertation (UMI#: 9917952), pp. 30
[85] P. W. Anderson and a. H. Dayem, Phys. Rev. Lett. 13, 195 (1964)
[86] D. J. Morgan and J. B. Ketterson, Phys. Rev. Lett. 80, 3614 (1998)
[87] M. Van Bael, Ph.D. thesis (Katholieke Universiteit Leuven, Belgium, 1998)
[88] S. J. Hagen, C. J. Lobb, R. L. Greene, M. G. Forrester, and J. H. Kang, Phys. Rev. B
41, 11630 (1990)

[89] J. F. Garvin, Jr., and R. C. Morris, Phys. Rev. B 21, 2905 (1980)

116

[90] J. C. Chen, S. C. Law, L. C. Tung, C. C. Chi, and Weiyam Guan, Phys. Rev. B 60,
12143 (1999)
[91] J. J. A. Baselmans, A. F. Morpurgo, B. J. van Wees and T. M. Klapwijk, Science
397, 43 (1999)

[92] A. F. Andreev, Zh. Eksp. Teor. Fiz. 46, 1823 (1964); 49, 655 (1965)
[93] C. A. Ross, M. Hwang, M. Shima, J. Y. Cheng, M. Farhound, T. A. Savas, Henry I.
Smith, W. Schwarzacher, F. M. Ross, M. Redjdal, and F. B. Humphrey, Phys. Rev. B 65
144417 (2002)
[94] Dr. Axel Hoffmann, private communication (2002)
[95] S. G. Lachenmann, T. Doderer, and R. P. Huebener, Phys. Rev. B 53, 14541 (1996)
[96]D. R. Tilley, Phys. Lett. 33A, 205 (1970)
[97] M. Octavio, M. Tinkham, G. E. Blonder, and T. M. Klapwijk, Phys. Rev. B 27, 6739
(1983)

117

Appendix
In this section, a summary of the main LabVIEW program (also known as virtual
instrument or vi, for short) developed for control of the system and for data acquisition
(the acquisition of the VIs), is given. With this program, the operator has control over
the temperature of the system, the magnetic field, as well as the rf signal output and
attenuation from the user interface shown below (known as Front Panel). The program
used is titled: DC-IV-Dots-New-Buffer-dVdI-LV6i.vi.
The program saves each individual data point onto a disk after each point is taken.
Also, these numerical values are presented in the front panel as well as in the four graphs
shown after they are collected. The displays are: The VI curve, the sample temperature,
and the pot and bath temperatures (which are used as quick indications of the ability of
the system to control temperature). Also, the dynamic resistance and d2V/dI2 curves are
calculated and displayed.
Front Panel

118

At the end of the run, the operator can input voltage values to calculate the critical
current of the sample using the threshold method, as well as to study the first and second
derivatives of the voltage with respect to current and calculate the expected position (Vn)
of a step by guessing a number (N) of junctions in the direction of the current (according
to Equation 2.31). After the operator is satisfied with the way the data is presented in the
front panel, a print is automatically generated before the program ends.
With this program, the user can choose to map the current in linear or in
logarithmic scales and if the current will return to Imin after reaching IMax in the
Current & Avg. Parameter box above (plus or minus IMax will be used only in the case of
a linear map of the current, Imin - IMax in the case of a logarithmic map). Also a specific
number of averages can be chosen for each data point to take or the program can decide
how many averages to take depending on an absolute error value for the measurement
(usually 5%). If the desired error cutoff is reached the averaging is finished or the
averaging continues until a set number of averages are reached.
This program could be used as a subroutine in other programs.

The input

(controls) values and outputs (indicators) are shown in the following diagrams (known as
icon and connector)
Connector Pane

119

The logic behind this program is seen in the following diagram: Three main forloops are used to control the way current is send to the sample. The inner most for-loop
is where all the data is colleted using the subroutine:
AVG-I-V-VH-CT-ST-RF-Buffer-LV6i.vi
All the data is formatted and the rows of data are saved to disk into a designated
data file created at the beginning of the run.
Block Diagram

The following is a list of the main subroutines used in this diagram:


List of SubVIs
Find First Error.vi
Write To Spreadsheet File.vi

120

Write Characters To File.vi


InitializationDCIVT8Z1-LV6i.vi
DCIV-Run-Info.vi
Iformat.vi
Errwait.vi
Beep.vi
Std Deviation and Variance.vi
Avg-I-V-VH-CT-ST-RF-Buffer-LV6i.vi
Wait for Temp stabilization.vi
I-Up-Down-LV6i.vi
Log-I-Up-Down-LV6i.vi
Log-I-LV6i.vi
LSHG10316-Cal-LV6i.vi
3D-1D-LV6.i.vi
derivative.vi
Critical-Current.vi
Shapiro step voltage.vi

121

The following diagram shows the dependence of these subroutines with respect to
the main program.
Position in Hierarchy

Some lower level subroutines are not mentioned in this appendix, only the first
row of programs is presented.

122

You might also like