Download Z - INDIGO @ UIC - University of Illinois at Chicago

Document related concepts

Superconductivity wikipedia , lookup

Lorentz force wikipedia , lookup

Aharonov–Bohm effect wikipedia , lookup

Electromagnet wikipedia , lookup

Time in physics wikipedia , lookup

RF resonant cavity thruster wikipedia , lookup

Transcript
Applications of Radio Frequency Resonant Cavities in Ultrafast
Electron Microscopy
BY
JOHN HOGAN
B.S., University of Illinois at Urbana-Champaign, 1996
M.S., University of Illinois at Chicago, 2009
THESIS
Submitted as partial fulfillment of the requirements
for the degree of Doctor of Philosophy in Physics
in the Graduate College of the
University of Illinois at Chicago, 2016
Chicago, Illinois
Defense Committee:
W. Andreas Schroeder, Chair and Advisor
Alan Nicholls, Research Resources Center
Danilo Erricolo, Electrical and Computer Engineering
Christoph Grein
Robert Klie
This thesis is dedicated to my parents, Tony and Therese Hogan.
ii
Acknowledgment
I would like to thank my advisor, W. Andreas Schroeder, for all of his
assistance in learning concepts, performing experiments, and also editing this
thesis. I thank all of my other committee members, Danilo Erricolo, Christoph
Grein, Robert Klie, and Alan Nicholls, for their help and support, with special
thanks to Professor Erricolo for allowing me to constantly borrow his extremely
expensive engineering equipment.
I would like to thank Joel Berger and Ben Rickman, members of my
research group, who provided invaluable help along the way.
Joel was
instrumental in helping me understand his research in the Ultrafast Electron
Microscopy project, which lays the foundation for much of my work completed
in this thesis. Joel and Ben both enabled me to tackle any problem involving a
computer, and Ben developed a dielectric copper coating process that I was able
to successfully use in one of my experiments.
Finally, I want to thank Kevin Lynch and Richard Frueh, who built many
of the components used in my experiments with quality and speed, and were
always willing to go out of their way to answer any of my questions and provide
any help I needed.
iii
Table of Contents
1
Introduction: UEM and RF Cavities.............................................. 1
2
Theory ............................................................................................... 6
2.1
2.1.1
Complex Notation for Time-Harmonic Fields ................................. 7
2.1.2
Cylindrical Waveguide Field Solutions ......................................... 10
2.1.3
Circular Cylindrical Cavity Field Solutions ................................... 15
2.1.4
Cavity Power Loss.......................................................................... 21
2.2
3
Electrodynamics of Resonant Cavities ................................................... 6
Power Supply ........................................................................................ 28
2.2.1
Transmission Lines and the Coaxial Cable TEM Mode ................ 30
2.2.2
General Impedance Matching ........................................................ 38
2.2.3
Inductive Coupling ......................................................................... 43
2.2.4
Impedance Matching of an Inductively Coupled Cavity ............... 50
Experimental Set-Up...................................................................... 57
3.1
Summary of Main Design Considerations ............................................ 59
3.2
Pre-existing Components ...................................................................... 64
3.2.1
Laser Oscillator and Frequency Doubling ..................................... 64
3.2.2
Laser Amplifier .............................................................................. 66
3.2.3
UEM Column Design ..................................................................... 67
3.3
Compression Resonant Cavity (CRC) .................................................. 68
3.3.1
Operation ........................................................................................ 68
3.3.2
Design ............................................................................................. 75
3.3.3
Construction ................................................................................... 80
3.3.4
Characterization.............................................................................. 91
3.4
Drift Tube and Deflection Plates......................................................... 100
3.5
Detection Resonant Cavity (DRC) ...................................................... 102
3.5.1
Operation ...................................................................................... 102
3.5.2
Double Pulse Operation and DRC Calibration ............................ 122
3.6
Synchronization and Power................................................................. 127
iv
Table of Contents
4
5
Experimental Results ................................................................... 130
4.1
Dielectric Axial Hole Effects .............................................................. 130
4.2
End Cap Hole Effects .......................................................................... 132
4.3
DRC Signal ......................................................................................... 135
4.3.1
Variation of Double Pulse Separation (Calibration) .................... 135
4.3.2
Variation of Current ..................................................................... 139
4.3.3
Compression ................................................................................. 140
Conclusion ..................................................................................... 143
5.1
Summary ............................................................................................. 143
5.2
Future Prospects .................................................................................. 146
References ............................................................................................ 148
Appendix I Mathematica Code for DRC Single Pulse Analysis ...... 151
Appendix II Mathematica Code for DRC Double Pulse Analysis ... 168
Appendix III Mathematica Code for DRC Unequal DP Analysis .... 175
v
List of Figures
Figure 1-1 Resonant Cavities in UEM .................................................................. 2
Figure 2-1 RCCC Geometry ................................................................................. 6
Figure 2-2 TM010 Mode Pattern .......................................................................... 20
Figure 2-3 TM010 Mode Current Pattern ............................................................. 21
Figure 2-4 Transmission Line Concept .............................................................. 30
Figure 2-5 Coaxial Cable Geometry ................................................................... 34
Figure 2-6 Low Frequency Network................................................................... 39
Figure 2-7 Inductive Coupling ............................................................................ 45
Figure 3-1 System Components .......................................................................... 57
Figure 3-2 Axial Hole Effect on CRC Frequency .............................................. 74
Figure 3-3 Axial Hole Effect on CRC E-Field ................................................... 75
Figure 3-4 CRC Components.............................................................................. 81
Figure 3-5 Coupling Loop Detail ........................................................................ 82
Figure 3-6 Cavity/Coupling Loop Detail ............................................................ 83
Figure 3-7 End Cap Detail .................................................................................. 84
Figure 3-8 YAG Dielectric Insert and Coated Fused Silica ............................... 86
Figure 3-9 Coated Fused Silica Cavity in DRC .................................................. 86
Figure 3-10 YAG Insert with Coupling Loop .................................................... 88
Figure 3-11 Vacuum Pass-Through Assembly ................................................... 89
Figure 3-12 Completed CRC Assembly ............................................................. 90
Figure 3-13 CRC Impedance .............................................................................. 93
Figure 3-14 Effect of Loop Area on Impedance ................................................. 95
Figure 3-15 Effect of Loop Area on Reflection Coefficient............................... 96
Figure 3-16 QZERO Output for CRC................................................................. 97
Figure 3-17 Deflector Plates ............................................................................. 101
Figure 3-18 Drift Tube Assembly ..................................................................... 102
Figure 3-19 Normalized TM010 H-Field ............................................................ 111
Figure 3-20 Pulse Transform Variation ........................................................... 113
vi
List of Figures
Figure 3-21 Single Pulse (Time-Domain) ........................................................ 115
Figure 3-22 Radial Dependence of Vectors in the Mode Projection Integral .. 116
Figure 3-23 Axial Dependence of Vectors in the Mode Projection Integral .... 117
Figure 3-24 Mode Projection and Cavity Length ............................................. 118
Figure 3-25 Mode Projection Integral and Gun Voltage .................................. 119
Figure 3-26 Mode Projection and Transfer Function ....................................... 120
Figure 3-27 Mode Projection Multiplied by the Transfer Function ................. 121
Figure 3-28 Pulse Duration Dependence of EMF ............................................ 122
Figure 3-29 Double Pulse Systems (Time-Domain) ........................................ 123
Figure 3-30 Double Pulse Transforms .............................................................. 125
Figure 3-31 Axial Dependence of DP Transform ............................................. 125
Figure 3-32 DP Separation Dependence of EMF ............................................. 126
Figure 4-1 Variation of Resonant Frequency with Axial Hole Size ................. 132
Figure 4-2 Variation of Resonant Frequency with End Cap Hole Size............ 133
Figure 4-3 Effects of End Cap Hole Diverging Field ....................................... 134
Figure 4-4 Variation of DRC Signal with DP Separation ................................ 138
Figure 4-5 Variation of DRC Signal with Current ........................................... 140
Figure 4-6 Electron Pulse Compression ........................................................... 141
vii
List of Symbols and Abbreviations
AG
Analytic Gaussian
BBO
β-Barium Borate
CCD
Charge Coupled Device
CRC
Compression Resonant Cavity
DP
Double Pulse
DRC
Detection Resonant Cavity
DTEM
Dynamic Transmission Electron Microscope/Microscopy
EM
Electromagnetic
EMF
Electromotive Force
FFT
Fast Fourier Transform
HW1/eM
Half-Width at 1/e Maximum
LBO
Lithium Triborate
PLL
Phase Locked Loop
RCCC
Right Circular Cylindrical Cavity
RF
Radio Frequency
RMS
Root Mean Square
RTP
Rubidium Titanyl Phosphate
SMA
SubMiniature Version A
TE
Transverse Electric
TEM
Transverse Electromagnetic
TM
Transverse Magnetic
UED
Ultrafast Electron Diffraction
UEM
Ultrafast Electron Microscopy
UIC
University of Illinois at Chicago
UV
Ultraviolet
YAG
Yttrium Aluminum Garnet
Yb:KGW
Ytterbium-Doped Potassium Gadolinium Tungstate
viii
Summary
Radio frequency resonant cavities are becoming important components in the
emerging field of Ultrafast Electron Microscopy, which aspires to bring Dynamic
Transmission Electron Microscopy into the realm of sub-nanometer spatial
resolution combined with sub-picosecond time resolution. To accomplish this
goal, the time-varying electromagnetic fields of resonant cavities can be utilized
for temporal focusing of laser-driven ultrafast electron pulses to be used for
imaging.
In this thesis I present the design and construction of a laser-
synchronized dielectric-filled radio frequency resonant cavity used for temporal
focusing. I also present the theoretical groundwork, design, and construction of
a second resonant cavity for the purposes of phase synchronization and direct
electron pulse duration measurement.
ix
1 Introduction: UEM and RF Cavities
Radio frequency (RF) resonant cavity oscillators have many applications
in engineering and physics, and are currently becoming valuable tools in the
specific field of Ultrafast Electron Microscopy (UEM) [1]. UEM involves
operating an ordinary electron microscope with short pulses of electrons instead
of a steady electron beam, allowing the possibility of taking “snapshots” of
dynamic processes at the nanometer scale [2]. This document details the use of
such cavities within the framework of a larger overall project that is ongoing in
the UEM group at the University of Illinois at Chicago (UIC). This project uses
a pulsed laser to drive a photoelectron gun, generating short electron pulses which
are to be focused at a specimen (perturbed immediately before the arrival of an
electron pulse by a separate, synchronized branch of the pulsed laser beam) for
the purpose of developing a single-shot Dynamic Transmission Electron
Microscope (DTEM). Such a research instrument will allow for a single pulse of
electrons to create a clear image of irreversible dynamic processes at the
nanometer scale, which is currently not possible [3-5]. The space-time resolution
goal for the ultrafast DTEM is in the sub 1 nm-ps (nanometer-picosecond) range,
with each pulse needing to contain roughly 108 electrons for proper imaging [68].
In this thesis, I detail the application of a cylindrical RF resonant cavity as
an RF pulse compression cavity in the UIC UEM system. A simple schematic
for a typical UEM system is shown in Figure 1-1. The electron pulse (shown in
blue in Figure 1-1 at different stages of propagation) generated by photoemission
in the laser-driven electron gun has intrinsic velocity dispersion (due to a variety
of excess energies inherent to the photoemission process) and, possibly, spacecharge effects that will increase its duration during propagation to the
specimen. The purpose of an RF compression cavity (labelled in Figure 1-1 as
1
the compression resonant cavity (CRC)) is to overcome this intrinsic pulse
broadening in the longitudinal dimension by reversing the internal velocities of
the electrons in the pulse, causing its duration to decrease (pulse compression) as
it propagates, with the goal of providing the shortest pulse at the sample for the
best space-time resolution in UEM. Currently, an RF resonant cavity has been
successfully used for compression in Ultrafast Electron Diffraction (UED), which
only requires up to 106 electrons per pulse instead of the 108 electrons required
for imaging [9].
Figure 1-1 Resonant Cavities in UEM
Schematic for a typical UEM system, with the addition of the DRC
which is unique to the UIC system.
Electronics are used to synchronize the CRC with the pulsed laser that is
driving the photoelectron gun, to ensure that the electromagnetic (EM) field
2
inside the cavity is at the proper phase for compression when each pulse arrives.
Depending on the phase of the field upon pulse arrival, the CRC could act to
compress, stretch, accelerate or decelerate the pulse; thus proper synchronization
is critical. The CRC described in this work is filled with low-loss dielectric
material with an axial hole to allow for the passage of the pulsed electron beam.
The dielectric material, in this case yttrium aluminum garnet (YAG), reduces the
size of the cavity which makes it much easier to mount into the UEM column.
The dielectric also allows for a polishing and copper coating process which can
significantly improve performance. The use of dielectric-filled resonant cavities
in single-shot UEM is unique to the UIC project. Power is delivered to the CRC
by a 30 W RF amplifier, operating at one of the resonant frequencies of the RF
cavity, and a low-loss coaxial cable with a small wire loop fixed to the end. The
loop is inserted into the CRC and power from the amplifier drives a large current
around the loop at the resonant frequency, which stimulates the corresponding
resonant mode EM field inside the cavity. This is known as inductively coupling
the cavity to the power source. The electric part of the stimulated EM field
applies the necessary compression force to each electron pulse passing through
the RF cavity. Proper impedance matching of the CRC and power source is
critical to efficient power delivery.
The design and application of a second cylindrical RF resonant cavity,
labelled the Detection Resonant Cavity (DRC) in Figure 1-1, will also be detailed
in this thesis. The DRC is not a standard component of any current UEM system
other than the UIC system. I developed the concept of the DRC during an attempt
to directly measure the duration of the compressed pulses at the temporal focal
point. I developed theory which shows that a traveling electron pulse can be
modelled as a current source which carries power at all frequencies, so that as a
pulse travels through an RF resonant cavity (the DRC) it will stimulate multiple
resonant modes of the cavity. In this case a coupling loop in the cavity can be
3
used as a receiving antenna, and magnetic flux from an impedance matched
stimulated mode will create an EMF in the loop. This low power signal can be
amplified and viewed on a fast oscilloscope. The theory shows that the amplitude
of this signal can determine the time duration of the pulse passing through the
DRC, without disturbing the pulse in any way. Unfortunately, the employed
GHz-level DRC can only detect pulse durations for pulses longer than about 10
ps, which is not useful for measuring the sub-picosecond compressed pulses in
either UEM or UED. However, besides having possible applications in systems
using longer duration pulses, it will be shown that the DRC can still be (and is)
used in the UIC project to determine whether the CRC EM field is at the proper
phase for compression when a pulse arrives. Moreover, the development of the
DRC lays the groundwork for higher frequency pulse detection cavities that could
potentially enter the sub-picosecond time regime.
In Chapter 2, I will provide an overview of all background theory relevant
to the application of the RF resonant cavities in a UEM system. The major topics
covered include the theory of cylindrical waveguides and cavities (which are
simply waveguides with end boundaries), and the nature of resonant eigenmodes
in ideal lossless cavities. This is followed by a discussion of power loss in real
cavities. Next, the theory of RF power delivery to a resonant cavity is detailed.
This includes basic transmission line theory and coaxial cable EM field theory,
inductive coupling and eigenmode stimulation by current sources, and impedance
matching of the power source to the cavity to prevent reflection of power back to
the source.
Chapter 3 begins by outlining the main design considerations for the CRC,
and then provides a simplified model of the operating principle. Next, it is shown
how all design criteria are met, and the details of the construction and
characterization of the CRC follow. Considerable time is then spent covering the
theory and operating principle of the DRC. A unique method for testing the
4
DRC’s ability to detect changes in larger pulse durations is then presented. It is
shown that a large duration electron pulse, for example, an 80 ps pulse (which is
too long to be generated by the UIC system), can be simulated by two shorter (~4
ps) electron pulses separated by 80 ps. It is also shown that this double pulse
method can be used to calibrate the DRC for direct long pulse duration
measurement. Chapter 3 ends by describing the process of synchronizing the
pulsed laser and the CRC, along with an analysis of system stability.
In Chapter 4, I present the results of experiments performed to test the
theory and construction of the CRC and DRC. There are two main experimental
results, the first being the successful calibration of the DRC using the double
pulse method. This type of experiment has not been performed before. Then,
results of a compression experiment using double pulses show the CRC is
properly synchronized, and achieves successful compression. The calibration of
the DRC is also shown to be accurate. The thesis concludes with Chapter 5 in
which a summary is given and future prospects are discussed.
5
2 Theory
2.1 Electrodynamics of Resonant Cavities
The goal of this section is to provide a basic understanding of how EM
fields behave inside a resonant cavity, specifically a right circular cylindrical
cavity (RCCC). Figure 2-1 shows the basic geometry of the problem.
Figure 2-1 RCCC Geometry
Relevant coordinates and origin to be used throughout this
document when discussing cylindrical cavities.
Initially, in our analysis, we will be seeking exact solutions to Maxwell’s
equations for cavities with perfect boundary conditions and material properties,
i.e. perfectly smooth metallic boundaries with infinite conductivity and lossless
linear dielectrics inside the cavity.
The actual fields are of course only
approximated by these exact solutions, and perturbation (or variational) methods
are used when more accurate solutions are needed for real-world cavities. In our
experiments with electron beams, we will be using a RCCC that has an axial hole
for the beam to pass through, an aperture and slot for the power coupling loop to
be inserted, a non-zero skin depth at the metallic boundaries, and a dielectric with
6
a non-zero loss tangent. The effect of these real-world cavity imperfections on
the exact field solutions will be discussed. We will find that the exact solutions
work well in the majority of situations and can therefore be used in estimates for
resonant frequencies, power consumption, energy transferred to the electron
pulses, DRC signal power, etc. Consequently, it is important to analyze these
exact solutions in some detail.
There are several main results from this analysis which will be summarized
here. First, a cylindrical cavity can be thought of as a cylindrical waveguide with
closed ends, so solutions to the waveguide problem that meet the end boundary
conditions are the solutions to the cavity problem. Second, the closed ends create
a completely enclosed volume which leads to a discrete spectrum of resonant
frequencies and corresponding field modes. Third, because of the geometry of
waveguides and cavities, we can assume a sinusoidal dependence of the fields in
the axial direction. As will be shown, this leads to the overall problem reducing
to that of finding a scalar function which represents each mode. The complete
vector modes and fields can then be found by taking vector derivatives of this
scalar function. Finally, the modes naturally separate into two distinct types,
transverse magnetic (TM) and transverse electric (TE). TM modes have no axial
magnetic field component, and TE modes similarly have no axial electric field
component. Any general field inside a resonant cavity can be expressed as a linear
combination of TM and TE modes. A specific TM mode, suited for compression
of an electron beam traveling along the axis of the cavity, will be used in
applications relevant to this document.
2.1.1 Complex Notation for Time-Harmonic Fields
The EM fields in the following analysis will be assumed to have sinusoidal
(harmonic) time dependence. It is much simpler to deal with such fields using
complex notation and so a brief review of this notation is given here. The
following are Maxwell’s equations in complex form for time-harmonic fields:
7
  E  i H
ˆ )E  J s .
  H  (  i
It is standard to use
H
instead of
B
2.1
for waveguide and cavity theory (it allows
for a simple analogy between waveguides and transmission lines). Here,
Js
represents what will be referred to as a source current that is applied to a system
and elicits a response from the system. This response can itself include currents
which are accounted for in other terms in the equations. All vector components
in the above equations are phasors, and include the phase and amplitude
information about each component of the fields.
For example, the phasor
Ez (r )  Ez (r ) e i where  is a phase constant. It is important to note that phasors
are a function of position only. The relationship between a phasor quantity and
an instantaneous real field quantity will be defined as the following:
Ez (r , t )  2 Re( Ez (r )eit )  2  12  Ez (r )eit  Ez* (r )eit  , where the factor of
2 means that the magnitude of a phasor is taken to be the root mean square
(rms) value of the real time-harmonic quantity it represents (simply a
convention). In Equation 2.1 we take  and

to be real and we allow ˆ to be
complex, ˆ     i  , where the imaginary part will account for dielectric loss
(the negative sign is again a convention). All of these quantities can be a function
of frequency.
Oftentimes when dealing with harmonic fields, instantaneous quantities
related to the fields will not be of interest; instead time-averaged quantities are
more useful. The most common quantity of this type (and one which will be
analyzed in detail in this document) is EM field power. For example, the
expression Ez (r , t )  Jz (r , t ) can be shown to represent instantaneous power flow
8
density in the z-direction (here Ez ( r , t ) represents the z-component of the electric
field, and Jz ( r , t ) represents the z-component of current density). The timeaverage value of this quantity is of interest for the harmonic fields that will be
used in the analysis. Time-average quantities will be represented by the following
over-bar notation: B ( r , t ) , for some quantity B . So time-average power density
is represented by the expression Ez (r , t )  Jz (r , t ) . It will be useful to represent
this type of expression (a time-average of a product of two harmonic field
quantities) in the complex phasor notation described above. From the definition
of the complex phasor it is evident that
Ez (r , t )  Jz (r , t )  2  12  Ez (r )eit  Ez* (r )eit   2  12  J z (r )e it  J z * (r )eit  .
After multiplying out all the terms and rearranging, it can be shown that the
expression on the right-hand side is the real part of another expression, and
therefore
Ez (r , t )  Jz (r , t )  2  Re  12  Ez (r )  J z* (r )  Ez (r )  J z (r )e 2it  .
It is seen that the time-average of the second term on the right side is simply zero,
so the final expression for the time-average power is
Ez (r , t )  Jz (r , t )  Re  Ez (r )  J z * (r )  .
2.2
Thus, the real part of the product of one complex phasor and the conjugate of
another is equal to the time-average value of the product of the two time-harmonic
field quantities.
9
In the remainder of this text, when a field is referred to (electric field,
magnetic field, EM field, etc.), it will be assumed that the field quantity in
question is a phasor quantity unless otherwise specified. In certain cases the term
“phasor” will be included, to avoid confusion when the distinction between
phasor quantities and time-dependent quantities is especially important (or
possibly unclear). Time-dependent quantities will always be written as E (for the
electric field, for example) and the corresponding phasor E.
2.1.2 Cylindrical Waveguide Field Solutions
In general, an EM waveguide can be thought of as any device which allows
for the propagation of EM waves along a directed path. In this section, it will be
assumed that a waveguide is filled with a homogenous lossless linear dielectric,
has perfect conducting boundaries, a constant cross-section, and allows
propagation in a straight line along the z-direction, which will always be aligned
with the central axis of the guide. The transverse plane will be defined as the
plane perpendicular to the z-direction, and the transverse field will be defined as
the field (electric, magnetic, etc.) made up only of components aligned in the
transverse plane. It is assumed that there are no current or charge sources
anywhere inside the waveguide. There are also no end boundary conditions, so
that the guide is taken to be infinitely long. Maxwell’s equations inside the guide
then have the form:
(a)   E
 i H
(b )   E
(c )   H
 i  E
(d )   H
0
 0.
2.3
A wave equation for the fields can be derived in the usual way and takes the form:
10
E
( 2    2 )    0.
H 
2.4
The assumption about the sinusoidal dependence of the fields in the z-direction is
now applied. Using the geometry of Figure 2-1, the fields can then be expressed
in the following form:
 E ( x, y, z, t )   E ( x, y )   ikz it
,


e
H
(
x
,
y
,
z
,
t
)
H
(
x
,
y
)

 

2.5
where k can be complex, and the  stands for waves traveling in the positive (+)
z-direction or negative (-) z-direction.
Having made this assumption about the z-direction dependence of the
fields, it can now be shown that all field components for a particular waveguide
mode can be derived from a single scalar function that represents that mode. The
first step is to separate Maxwell’s vector Equations 2.3(a,c) into z-component
equations and transverse component equations.
It is useful to define two
operators; the transverse gradient vector operator t    zˆ  z , and the
transverse Laplacian scalar operator t2  2   2 z 2 . Also, the expressions Et
and H t will be used to represent the transverse components of the fields. To get
the separated equations, the cross product and dot product of the unit vector ẑ is
taken with each of Maxwell’s curl equations. For example, for 2.3(a) we take
zˆ  (  E )  zˆ  (i H ) and zˆ  (  E )  zˆ  (i H ) , and similarly for 2.3(c).
Maxwell’s divergence equations can also be rewritten, separating the transverse
part, resulting in (using our specified dependence in the z-direction)
11
(a)
 ikEt  i ( zˆ  H t )  t Ez
(b )
zˆ  (t  Et )  i H z
(c )
 ikH t  i  ( zˆ  Et )  t H z
(d )
zˆ  (t  H t )  i  Ez 2.6
(e)
t  Et  
(f)
t  H t  
Ez
z
H z
,
z
where again the  stands for waves traveling in the positive (+) z-direction or
negative (-) z-direction. One can now observe that if the z-components of the
fields are known, the other components can be computed from the above
equations. If one simultaneously solves Equations 2.6(a) and 2.6(c) for Et and
H t , the result is
(a)
(b )
i
  k (t Ez )   ( zˆ  t H z )
(   2  k 2 )
i
Ht 
  k (t H z )    ( zˆ  t Ez ),
(   2  k 2 )
Et 
2.7
which shows the explicit relationship. It is important to note that these equations
are only valid if at least one of the z-components (Ez or Bz) is non-zero. (If both
are zero there can still exist what is called a TEM mode, to be discussed in a later
section). At this point a solvable equation is needed for the scalar z-components
alone. This is accomplished by substituting the expressions on the right-hand
side of Equations 2.5 into Equation 2.4, and then isolating the z-component parts.
This results in:
E 
t2  (   2  k 2 )   z   0,
H z 
12
2.8
which is the desired equation. Next, boundary conditions specifically for the zcomponents are needed. The general boundary conditions for time-varying fields
at the surface S of a perfect conductor are
(a)
nˆ  E  0
(b )
nˆ  H  0
(c )
nˆ  E   f
(d )
nˆ  H  K f ,
S
S
S
2.9
S
where n̂ is a unit normal vector at the surface,  f is the free surface charge, and
K f is the free surface current. As one generally does not have knowledge of the
surface charge or surface current, only the general boundary conditions in 2.9(a)
and 2.9(b) will be used. From Condition (a) it is clear that the boundary condition
for Ez must be
E z S  0.
2.10
The boundary condition for Hz is not as obvious. If one takes the dot product of
the unit normal n̂ with Equation 2.6(c) at the boundary surface S, one obtains
H
ik (nˆ  H t )  i  nˆ  ( zˆ  Et )   z ,
S
S 

n S
2.11
where  n is the normal derivative at a point on the surface. The first term on
the left is zero because of the general boundary Condition 2.9(b). The second
term is zero because of general boundary Condition 2.9(a). What remains is the
boundary condition for Hz :
13
H z
n
 0.
2.12
S
One now has equations and boundary conditions for Ez and Hz, from which the
complete vector fields can be solved. The problem can be simplified further,
however, if the solutions are first divided into two types. For the first type, one
sets H z  0 everywhere, which automatically satisfies boundary Condition 2.12
and Condition 2.10 remains. These are called transverse magnetic or TM modes,
since the magnetic field only has transverse components. For the second type,
one sets E z  0 everywhere, satisfying boundary Condition 2.10 and leaving
Condition 2.12. These are the transverse electric or TE modes. For TM/TE
modes, Equations 2.7 simplify nicely, becoming for TM modes
(a)
(b )
i
(TM )
  k (t Ez )
(   2  k 2 )
i
Ht 
   ( zˆ  t Ez ) (TM ),
(   2  k 2 )
Et 
2.13
and for TE modes
(a)
(b )
i
   ( zˆ  t H z ) (TE )
(   2  k 2 )
i
Ht 
(TE ).
  k (t H z )
(   2  k 2 )
Et 
2.14
From these equations it is clear that for TM(TE) modes, knowledge of a single
scalar function, namely Ez(Hz), that satisfies wave Equation 2.8 and the TM(TE)
boundary condition, will allow calculation of the full vector fields. For TM
14
 ikz it
,
modes Ez can be represented as a scalar function  TM ( x, y ) multiplied by e
where  TM ( x, y ) is a function of the transverse independent variables x and y
only. Similarly, for TE modes one has H z   TE ( x, y)e ikz it , with  TE ( x, y)
being called the scalar TE mode function. The mode functions will still solve the
wave equation
 TM ( x, y ) 
t2  (   2  k 2 )   TE
  0,
 ( x, y ) 
2.15
and when combined with the TM(TE) boundary Conditions 2.10(2.12) one has a
unique eigenvalue problem for ψTM(ψTE). The wave equation is the same, but the
different boundary conditions mean the two sets of solutions will have different
eigenvalues.
2.1.3 Circular Cylindrical Cavity Field Solutions
If a cylindrical waveguide has closed, perfectly conducting ends (the ends
being in the transverse plane), then the waveguide becomes a cylindrical cavity.
Everything about the problem remains the same, except for the ends adding an
additional boundary condition. To meet this boundary condition, for each mode
a wave traveling in the positive and negative direction will need to be combined
to create standing waves inside the cavity. The individual traveling waves remain
identical to our waveguide solutions. It will be assumed in advance via symmetry
that the contribution from the forward and backward traveling waves is of the
same magnitude. The solution will be developed for TM modes in a cavity,
noting that the TE mode analysis would be analogous. For any TM mode the
general solution in a cavity will be
15
 Et 
 Et 
 Et 
 
 
 
 Ez   A  Ez   A  Ez 
H 
H  
H  
 t
 t 
 t 
TM
cavity
,
2.16
where the +(-) sign refers to a wave propagating in the positive(negative) zdirection, and it has been assumed an equal contribution from each wave (note
there is no Hz for the TM modes). Therefore, one simply needs to determine the
constant A which allows the end boundary conditions to be met. Consider first
Et  A( Et  Et ); the waveguide solutions for the TM fields 2.13 allow this
equation to be expressed as
Et 
ik
(    k )
2
2
t TM ( x, y)eit ( Aeikz  Aeikz ),
where Ez   TM ( x, y)eikz it has been used.
2.17
To meet the end boundary
conditions, Et must be zero at z  0 and z  d (see Figure 2-1) so Et must have
a
sin(kz ) dependence with k  p  d , p  0,1,2... . This can be accomplished by
setting A  1 2 . Using this value for A also fully determines the z-dependence
for Ez and H t , which have expansions analogous to Equation 2.17. Setting
 2  (  2  k 2 ) and with k  p  d , the final expressions for the TM cavity
fields become:
16
Et  
p
 p z 
TM
sin 
 t ( x, y )
2
d
 d 
(TM cavity )
 p z 
Ez   TM ( x, y )cos 

 d 
i 
 p z 
TM
H t  2 cos 
 zˆ  t ( x, y )

 d 
(TM cavity )
2.18
(TM cavity ).
For the TE fields, the boundary condition requires Hz to equal zero at z  0 and
zd,
so it is Hz that will need the
sin(kz ) dependence with
k  p  d , p  0,1,2... . The expressions for the TE cavity fields then become
 p z 
TE
sin 
 zˆ  t ( x, y ) (TEcavity )

 d 
p
 p z 
TE
H t  2 cos 
(TEcavity )
 t ( x, y )
d
 d 
Et  
i
2
 p z 
H z   TE ( x, y )sin 

 d 
2.19
(TEcavity ).
The most important thing to note from these two sets of equations is that for any
particular mode, we only need to solve for a single scalar mode function in order
to derive all of the cavity fields, just like in the waveguide problem.
At this point we can finally investigate the nature of these scalar mode
functions. Consider the TM functions  TM ( x, y ) , which solve Equation 2.15
subject to boundary Condition 2.10, with k  p  d and Ez   TM ( x, y)eikz it .
One has:

2
t
  2  TM ( x, y)  0,  TM ( x, y)  0,
S
17
2.20
which is simply the two-dimensional scalar Helmholtz equation with Dirichlet
boundary conditions, and eigenvalues given by
 2    2   p d  .
2
2.21
The solutions to this problem are well known. In cylindrical coordinates, which
TM
will be used for the RCCC, the boundary condition becomes  (  , )
 R
0
(see Figure 2-1) and the solutions are
TM
 mn
(  , )  E0 J m ( mn  )eim ,
2.22
where E0 is a normalization phasor,  mn  xmn R , and xmn is the nth root of the
Bessel function J m ( x) . The integer m can take the values m  0,1,2,... , and the
integer n can take the values n  1,2,3,... . Solving Equation 2.21 for the TM
eigenfrequencies ωTM generates

TM
mnp

1
 
2
xmn
p 2 2
 2
R2
d
(TM ).
2.23
The corresponding TM modes are labeled TMmnp. For TM modes, the integer p
can take on values
p  0,1,2,... in order to meet the end boundary conditions. In
the experiments relevant to this document, the mode of greatest interest is the
m  0 , n  1,
TM
p  0 mode, or TM010, with eigenfrequency 010
. As x01  2.405
, the frequency for this mode becomes
18
TM
010

2.405
.

 R
2.24
TM
(  , )  E0 J 0  2.405 R  , and using
The scalar mode function simplifies to  01
Equations 2.18 one can solve for the fields:
 2.405  it
Ez  E0 J 0 
e
 R 
H  i

 2.405  it
E0 J1 
e

 R 
TM 010 
2.25
TM 010 .
Finding the TE modes and frequencies proceeds in the same manner, by solving
TE
for  mnp
(  , ) using the scalar Helmholtz equation with the appropriate boundary
condition, and then using Equation 2.19 to find the vector fields. The expression
for the eigenfrequencies becomes

TE
mnp

1
 
2
xmn
p 2 2
 2
R2
d
(TE ),
2.26
2 is the nth root of the derivative of the mth Bessel function J m ( x) .
where xmn
Also, for TE modes the integer p cannot be zero but must start at 1. The other
integers m and n take the same values as TM modes.
The field pattern for the experimentally employed TM010 mode (Equations
2.25) is shown in Figure 2-2. From this pattern, it is clear that
n̂  H is nonzero
at all surfaces of the cavity. Going back to the boundary conditions for a perfect
conductor (Equations 2.9), it is evident from Condition (d) that there must be a
surface current flowing on all surfaces of the cavity, and this surface current is
19
equal to
n̂  H at each point on each surface (due to Condition (c) there must also
be a surface charge on the end surfaces of the cavity, but this is less important).
Figure 2-2 TM010 Mode Pattern
Cylindrical resonant cavity TM010 eigenmode pattern, with the H-field lines
shown in blue and the E-field lines in red. A sinusoidal time-dependence is
assumed.
The flow pattern for the surface current is shown in Figure 2-3. It is
important to note that when the boundaries are not perfect conductors but have
finite conductivity, these currents will no longer be surface currents but will be
volume current densities that penetrate a certain distance (called the skin depth)
into the boundary material (copper, for example). These current densities will
cause ohmic losses in the boundary material and they are a major source of power
loss when operating a resonant cavity. This will be explored in detail in the next
section.
20
Figure 2-3 TM010 Mode Current Pattern
Surface current pattern for the TM010 mode. The current is
required to meet the boundary conditions for the cavity.
2.1.4 Cavity Power Loss
In UEM, the usual purpose of a resonant cavity is to affect the temporal
dynamics of an electron pulse that passes through it. The cavity accomplishes
this task by generating strong internal EM fields that then interact with the
electrons in the pulse. One of the goals of resonant cavity design is to make sure
any power that is delivered to the cavity is efficiently transformed into useful
energy in the stimulated field mode(s) that may then be transferred to the
electrons. It is therefore desirable to minimize all sources of power loss in the
system that deter from this goal. The method that will be used to investigate
power flow and loss in cavities is the well-known Poynting theorem, in complex
form for fields with sinusoidal time-dependence. To derive the relevant equation
for our specific purposes, one starts with the complex Equations 2.1, multiplying
the first equation by H * and the complex conjugate of the second equation by E.
The point of doing this is to create a term (in the second equation) that has the
form
E  J s* , the real part of which will be related to time-average power flow
21
supplied to a system by its sources, which is what will be of interest. The result
of the multiplication is
H *    E  i H *  H
ˆ  E  E *  E  J s*.
E    H *    i
Subtracting the first equation from the second, and letting ˆ     i  , one
obtains
2

2
2
E    H *  H *    E   E    E  i  H    E
2
 EJ .
s*
The left-hand side of this equation can be simplified using a vector identity,
giving
2
2

2
  ( E  H * )   E    E  i  H    E
2
 EJ .
s*
The complex Poynting vector will be defined as S  E  H , and note that E
*
2
2
and H are time-average quantities, since they are technically the real parts of
expressions involving a phasor multiplied by the complex conjugate of another
phasor. One can now integrate the equation over a volume V that is enclosed by
a surface S. A unit vector n̂ is defined to be normal to S and pointing out of S,
away from the volume V. The divergence theorem is then applied to the left-hand
side of the equation. This results in:




  S  dS (nˆ )    E  J s*   E 2    E 2  i  H 2    E 2 dV .


22
2.27
Rearranging the terms and defining the following quantities:
Ps    E  J s*dV 
 S  dS (nˆ ) 
P     E dV 
P      E dV 
W     E dV 
W    H dV 
Pf 
2
d,
2
d , 
e
1
2
m
1
2
2
2
Complex Power Supplied by Sources
Complex Power Flowing Out of S
Time  AverageConduction Loss in V
2.28
Time  Average Dielectric Loss in V
Time  Average Stored Electrical Energy in V
Time  Average Stored Magnetic Energy in V ,
allows Equation 2.27 to be rewritten in the much simpler form
Ps  Pf  Pd ,  Pd ,   2i(Wm  We ).
2.29
The real part of this equation is
Ps  Pf  Pd ,  Pd ,  ,
2.30
where it is noted from the definitions of Ps and Pf that their real parts have the
form of a time-average quantity. In words, Equation 2.30 states that the timeaverage power supplied by the sources of a system (in our case a cavity system)
is equal to the time-average power flowing out of the surface of the system, plus
the time-average power dissipated within inside the system due to conduction and
dielectric losses. The imaginary part of Equation 2.29 is
Im  Ps   Im  Pf   2 (Wm  We ).
23
2.31
This is an equation that describes the behavior of stored energy that does not
dissipate. Inside the system, energy stored in the fields switches from electric
form to magnetic form and back during every cycle. If there is an excess of one
form over the other on average, Equation 2.31 tells us that part of this excess
either is sent back to the source and returns to the system every cycle, or flows
out of the surface and returns to the system every cycle. The part that is sent back
to the source for temporary storage each cycle is Im( Ps ) . The part that flows out
of the surface for temporary storage is Im( Pf ) . A system with no sources
supplying power and no power flowing through the surface will have an equal
amount of time-averaged stored electric and magnetic energy. For example, if
one calculated the time-average stored energies in the exact cavity solutions
described earlier in Section 2.1.3 (operating at their respective eigenfrequencies),
no energy imbalance would be found, since these solutions assumed the cavity
was source-free with perfectly conducting boundaries allowing no power to flow
through them.
One of the main sources of power loss is ohmic loss due to currents flowing
in the conducting cavity boundaries. For real cavities these boundaries do not
have perfect conductivity, which was assumed previously to simplify boundary
conditions when solving for the exact field solutions. These exact solutions are
still useful, however, since the boundary materials (copper, silver, etc.) used in
practice have very high conductivity, and the change to real boundary conditions
can be viewed as a small perturbation. The exact field solutions can therefore be
used to approximate the power loss to lowest order, which is generally sufficient
for these materials.
Interestingly, in order to determine the power lost to boundary currents,
one does not need to calculate the currents directly, but can instead approximate
the (perturbed) EM fields at the boundary, on the cavity (dielectric) side, and then
24
integrate the real part of the resulting complex Poynting vector over the cavity
surface. According to the theory developed in the previous section, this integral
will give the time-averaged power flowing out of the surface of the cavity. This
must be equal to the power lost by currents in the conducting material since there
is no other place for the power to go. To determine (approximately) the EM fields
at the boundary on the cavity side, one must actually start with Maxwell’s
equations in complex form inside the conducting material surrounding the cavity.
These equations will be applied at the boundary, but on the conductor side, and
then boundary conditions will be used to show these fields must be continuous
across the boundary into the cavity. The equations inside the conductor are
  H c  i E c   E c
  E c  i H c ,
where the ‘c’ superscript means these are the fields inside the conductor. For
materials like copper at radio frequencies, 
 , so the first term on the right-
hand side of the first equation can be neglected, which after rearranging gives
Ec 
Hc 
1

 Hc
i
c
  Ec.
Next, it is noted that EM fields in a good conductor decay very quickly in the
direction normal to the surface of the conductor. Mathematically, this means that
the normal derivative of the fields will be much greater than the transverse
derivatives. If one defines a coordinate  that equals zero on the surface of the
conductor and increases in the normal direction to the surface (in the direction of
25
the surface unit vector n̂ , away from the volume), then mathematically one can
write t
  and let   n̂   in the previous set of equations. This results
in:
 c
H


i
 c
Hc 
nˆ 
E.
c

1
Ec 
nˆ 
2.32
These coupled equations can be combined into a wave equation and solved,
giving solutions of the form [10]



H ( )  H e e
c
c

E c ( ) 

1

i




i

(i  1)(nˆ  H c )e  e 
2.33
1
2

 ,


 c

 
2
where Equations 2.32 have been used to solve for the electric field in terms of the
magnetic field.
The additional subscript on the fields means, in this
approximation, they are everywhere parallel to the boundary surface S and have
no component along n̂ . The parameter  , called the skin depth, has been defined,
and is equal to the distance inside the conductor at which the field amplitude
decays to 1 e of its value at the conductor surface. Notice that the higher the
frequency and conductivity, the smaller the skin depth, which is basically a
measure of how deep EM field penetrate into a conducting surface.
The only unknown in Equations 2.33 is the value of H c , the parallel
component of the magnetic field on the conductor side of the boundary, where
26
the coordinate  is equal to zero. According to Ohm’s Law, a current density
equal to
J is
 E inside a conductor, and this may not be formulated in terms of an
infinite surface current density as this would require an infinite electric field at
the surface for a finite conductivity. Consequently, there can be no surface
current along the surface S separating the cavity volume from the conductor. As
only a surface current can cause a discontinuity in the parallel component of a
magnetic field across a boundary, the magnetic field must be continuous across
S; that is Hc  H , where the superscript has been removed to indicate we are
referring to the parallel component of the magnetic field at the boundary but on
the cavity (dielectric) side.
At this point, the solutions presented in Section 2.1.3, where infinite
conductivity was assumed, can be employed. Since the actual conductivity is not
infinite but still very large, as an approximation it will be assumed, for any
operating mode of the cavity, that the parallel component of the magnetic field at
the surface (on the cavity side) is the same as stated in the prior solutions. As this
component is continuous across the boundary, it can be used in Equations 2.33,
which apply in the conductor, to calculate the parallel component of the electric
field at the boundary on the conductor side. Parallel components of the electric
field are always continuous across a boundary, so this value will be the same on
the cavity side of the boundary. Thus, values for the parallel components of both
the electric and magnetic fields at the boundary on the cavity side are known.
Using these values of the fields, the expression for time-average power flowing
out of the surface S surrounding the cavity is
Pf  Re


 S  dS (nˆ)  Re
   E  H   dS (nˆ)   4  H

27

c
2

dS , 2.34
where H  , as previously stated, is the component of the magnetic field parallel
to the surface S on the cavity side, and taken from the eigenmode solutions of
Section 2.1.3 as an approximation.
The other main source of power loss inside the cavity is dielectric loss. For
the eigenmode solutions, in addition to assuming perfectly conducting
boundaries, it was also assumed that the dielectric material inside the cavity was
completely lossless; that is,    0 . For real cavities this is not true, although the
value of   will be very small for good dielectrics such as fused silica and YAG.
This fact will again allow the eigenmode solutions to be used as an
approximation, since the actual field will not be altered significantly by the
presence of relatively small loss. As a result, the electric field of an eigenmode
solution in Equation 2.28 is used to calculate the term ( Pd ,  ) for dielectric loss
within the cavity volume. Note that Equation 2.28 also contains a term for
conduction loss within the cavity (Pd , ) . As will be discussed in Section 2.2.3,
an operating cavity will include a small internal current carrying loop of wire that
delivers power to the cavity, and the wire will have conduction losses. This loss
is also small and the eigenmode solution can still be used for calculations.
2.2 Power Supply
So far, discussion has included the EM field modes that can exist inside a
RCCC and also the power losses associated with operation of those modes, but
how these modes are excited and driven has not been described. Power must be
delivered from a source to the RCCC in order to drive the desired eigenmode to
the proper amplitude. Once the proper amplitude is reached, power will be
continuously needed to maintain that amplitude because of cavity losses. In the
experiments described in this document, power from the source is delivered to a
RCCC along a coaxial cable. A coaxial cable can be considered a special type of
metallic waveguide that has an inner and an outer conducting surface, with the
28
volume in between capable of sustaining EM waves similar to those in the hollow
waveguide already discussed. The major difference is that the inner conductor in
a coaxial cable allows the existence of a TEM (transverse electromagnetic) wave
that is not possible in a hollow guide. A TEM wave has no field component
(electric or magnetic) in the axial direction and, as will be shown, this property
qualifies the coaxial cable as a type of transmission line. There are several
advantages to using transmission lines to transport power, the main advantage
being that the characteristic impedance of the line is independent of frequency.
This allows the line to be compatible with equipment (such as power supplies,
connectors, couplers, etc.) that operate over a range of frequencies and are
designed for use with lines having standardized characteristic impedances. This
is extremely important, since any mismatch between the impedance of the line
and the effective impedance of the equipment it is attached to will cause
reflections of power back down the line, away from its intended destination. This
phenomenon is analogous to an EM plane wave being partially reflected when it
encounters a medium with different dielectric properties.
Once power has
traveled from the source to the cavity via transmission line, the proper eigenmode
of the cavity is then, in our case, stimulated inside the cavity by a technique called
inductive coupling (other techniques are used in different applications). In order
to prevent any reflection of power back down the line away from the cavity, the
combined effective impedance of the cavity and inductive coupling system must
match the characteristic impedance of the line. This is the art-form known as
impedance matching. The theory behind these concepts is described in this
section.
29
2.2.1 Transmission Lines and the Coaxial Cable TEM Mode
Figure 2-4 Transmission Line Concept
Parallel wire transmission line model used for derivation of the
transmission line equations.
There are many different kinds of transmission lines used to deliver power;
the parallel wire transmission line is shown in Figure 2-4 and is typically used to
demonstrate the basic concepts. Here, one considers the positive z-direction to
be the direction pointing along the line towards the load.
If a sinusoidal
alternating voltage is applied by a source to one end of the line, voltage and
current waves will travel down the line carrying power with them. We are
concerned with a particular type of wave or mode that can be sustained on a
transmission line, called the transverse electromagnetic (TEM) mode or simply a
transmission line mode. For this type of mode, at any particular position or crosssection along the line, there will be a uniquely defined voltage and current related
to the transverse electric and magnetic fields surrounding the wires (again, the
fields have no z-components for the TEM mode, only transverse). For the parallel
wire set-up, the unique definitions are
30
V ( z )   E  dl
I ( z )   H  dl ,
C1
C2
2.35
where C1 is a path connecting the two wires at the particular cross-section and C2
is a path surrounding one of the wires at the cross-section. The values must be
unique regardless of the specific chosen path in order for the mode to be
considered a transmission line mode.
These voltage and current values will
change as we move in the z-direction. The fundamental postulate for transmission
lines states that the change in voltage is proportional to the current at any point,
and conversely the change in current is proportional to the voltage. From this
postulate and Figure 2-4, one can generate the well-known transmission line
equations
dV ( z )
  ZI ( z )
dz
dI ( z )
 YV ( z ).
dz
2.36
Here the proportionality constants are labeled Z, the impedance per unit length of
the line, and Y, the admittance per unit length of the line. These constants, which
can have many different forms depending on the type of transmission line, will
be assumed to be independent of z for our purposes and they are not phasor
quantities. V(z) and I(z) are phasor quantities, with a sinusoidal time-dependence
assumed for the real voltage and current values. One can derive wave equations
for the voltage and current phasors from the transmission line equations and
obtain
d 2V ( z )
 ( ZY )V ( z )  0
dz 2
d 2 I ( z)
 ( ZY ) I ( z )  0.
dz 2
2.37
These standard one-dimensional Helmholtz equations have the general solutions
31
V ( z )  V0e
ZY z
  ZY z
I ( z)  I0 e
 V0e
ZY z
  ZY z
 I0 e
2.38
,
where we see that waves traveling in the positive or negative direction are allowed
and have a sinusoidal z-direction dependence. Suppose only a wave traveling in
the positive direction is present ( V0   I 0   0 ). If one substitutes the resulting
solutions from Equations 2.38 into Equations 2.36 and divides, the ratio of the
voltage and current phasors is found to be
V0 
Z

 Z0 ,

I0
Y
2.39
where Z0 is defined to be the characteristic impedance of the line. For a wave
traveling in the negative direction only, one obtains
V0 
Z

 Z0 .

I0
Y
2.40
As long as there is only a single wave present, traveling in one direction, the
absolute value of the ratio of voltage to current will be equal to Z0 for the entire
length of the line and be independent of the z-coordinate. When waves traveling
in both directions are present, the ratio of the total voltage and current at any
particular cross-section becomes a function of the position on the line, and this
ratio Z(z) is called the impedance of the line.
One can define the complex power traveling down the line as
32
Pf ( z )  V ( z) I * ( z ),
2.41
where, as usual, the real part represents time-average power flow. For a lossless
transmission line, the time-average power flow will be independent of the zcoordinate, but for real lines there will be power losses and an attenuation per
unit length can be defined for the line. In this case the quantity
ZY will be
complex, with the real part being related to the attenuation.
As previously stated, a coaxial cable can sustain a TEM wave and therefore
can potentially be treated as a transmission line. From Equations 2.6 (b) and (e),
if Ez and Hz are set equal to zero (TEM), one finds that t  Et  t  Et  0 , and
it follows that t2 Et  0 . It is also known that the tangential electric field must be
zero at the inner and outer boundary surfaces, since they are assumed to be (near)
perfect conductors. So for any particular cross-section of the cable, it is evident
that the boundary value problem for the transverse electric field is identical to the
two-dimensional electrostatic boundary value problem. A similar analysis of the
magnetic field shows that they are solutions to the two-dimensional
magnetostatic boundary value problem. Both of these problems of course have
well-known solutions [11], whose patterns are shown in Figure 2-5. As in the
case of the RCCC, boundary conditions for these magnetic field patterns require
surface currents to flow on the inner and outer conductor surfaces, and these
currents are also shown in the figure. The current on the inner conductor will be
key to coupling the coaxial cable to the resonant cavity and delivering power, as
will be discussed later.
33
Figure 2-5 Coaxial Cable Geometry
Surface current flow and TEM mode pattern for a coaxial cable. Also shown are the
integration paths used to calculate equivalent transmission line voltages and currents.
It is now clear why a TEM mode cannot exist in a hollow waveguide; the
electrostatic solution for the electric field would be zero everywhere without the
inner conductor present. It is important to remember that although the patterns
for the field (phasors) look identical to a static field (phasor) solution, the actual
electric field still carries with it a sinusoidal time-dependence (the phasors only
contain information about the amplitude and phase of the true time-dependent
fields) as well as the usual assumed z-direction dependence. The solutions for
the fields (phasors) are
E ( z ) 
E0  ikz
e
2 r
H ( z ) 
H 0  ikz
e ,
2 r
2.42
where E0 and H0 are normalization phasors. The transverse electric and magnetic
fields ( E ( z ) and H ( z ) ) of the coaxial cable TEM mode can of course be related
to each other; from Equation 2.6 (c), with Hz set equal to zero, one finds
H ( z )  (  k ) E ( z ) . From Equation 2.8, recalling that t2 Et  0 , the result
k     is obtained. Inserting this value of k into the transverse magnetic
field expression results in
34
H ( z )  

E ( z ),
 
2.43
where the (+) stands for waves traveling in the positive z-direction and the (-)
represents waves traveling in the negative z-direction. It is now possible to
express both the electric and magnetic fields in terms of the electric normalization
phasor E0, which will turn out to be convenient later:
E ( z ) 
E0  i
e
2
  z
H ( z )  
  E0  i
e
 2
  z
.
2.44
In order to represent the coaxial cable as a transmission line, it is now necessary
to define the unique voltage and current phasors at a cross-section of the line in
terms of the transverse fields (recall that these unique phasor definitions must
exist for the cable to be considered a transmission line). One would like to do
this in such a way that the total complex power traveling down the cable will be
*
( z ) , in analogy to the transmission line complex power
equal to Vcoax ( z ) I coax
Equation 2.41. To find the total complex power flowing down the cable, the
complex Poynting vector is integrated over a cross-section of the cable. For
example, for a wave traveling in the positive z-direction we have
Pfcoax ( z ) 
 S ( z)  dA   E ( z ) H ( z ) dA[( ˆ  ˆ )  zˆ]
*
A
A
 E0     E0* 
  E0
b
 2  

d


ln   .



2    2 
 2  a 
a
2
b
35
2.45
(Note that the expression here for the complex power only has a (z-independent)
real part. This will not be true when waves traveling in both directions are
present.) If we make the following definitions, where the integration paths C1
and C2 are shown in Figure 2-5:
E0e i  z
E
Vcoax ( z )   E ( z )  
d   0 ln  b a  e  i
2
2
C1
a
b
I coax ( z )   H ( z ) 
C2
2

0
 i   z
  E0e

2
 d  
  z
   i
Ee
 0
2.46
  z
,
then the desired result for complex power can be obtained. It is important to note
that, because of the nature of the TEM field pattern, the voltage and current values
will be independent of the integration paths as long as C1 connects the inner and
outer conductors and C2 encircles the inner conductor. This is not in general true
for TM and TE modes. For those modes any type of transmission line analogy
will therefore be an approximation, but for the TEM mode the analogy is exact
and much more useful in predicting the behavior of the power flow in the coaxial
cable.
To complete the transmission line analogy it is necessary to show that Vcoax
and Icoax, as they have been defined, satisfy the transmission line Equations 2.36.
Inserting the above definitions into Equations 2.36 we obtain
dVcoax ( z )
E0e i
 i   ln  b a 
dz
2
dI coax ( z )
 i E0e i  z .
dz
If the following definitions are made:
36
  z
2.47
Z coax  i
ln  b a 
2
Ycoax  i 
2
,
ln  b a 
2.48
then Equations 2.47 will take the form
dVcoax ( z )
  Z coax I coax ( z )
dz
dI coax ( z )
 YcoaxVcoax ( z ),
dz
2.49
in exact analogy to the transmission line equations. One can also define the
characteristic impedance for the coaxial cable in analogy with Equations 2.39
and 2.40:
Z 0coax 
V0coax
I

0coax

V0coax
I

0coax
Z coax
 ln  b a 

.
Ycoax
  2

2.50
In analogy to the case of the parallel wire transmission line, Equations 2.49 will
have solutions with the general form
 ZcoaxYcoax z
 V0coax e
 ZcoaxYcoax z

0coax
Vcoax ( z )  V0coax e
I coax ( z )  I

0coax
e
I
e
 Z coaxYcoax z
 Z coaxYcoax z
2.51
.
As before, if only a positive or negative traveling wave is present the impedance
of the cable,
Z coax ( z ) 
Vcoax ( z )
,
I coax ( z )
37
2.52
will be equal to the characteristic impedance of the cable. If waves in both
directions are present, the impedance becomes a function of the z-position on the
line.
2.2.2 General Impedance Matching
The general setup for what is known as a low-frequency network is shown
in Figure 2-6. The thick black lines represent the transmission line portion of the
network. In reality this could be any type of transmission line, in our case it
represents the coaxial cable. Attached to the beginning of the line is the power
source which generates the voltage and current waves traveling down the line. It
is characterized by its impedance labeled ZS. This means that at the cross-section
where the source is attached to the line ( z  s ), the ratio of the voltage to the
current, as defined previously, will be equal to ZS (and similarly for all
impedances we define). A reference plane along the line is defined arbitrarily
and the z-coordinate is measured from this plane. ZIN is defined as the (input)
impedance at the reference plane and can be considered the impedance the source
would “see” if it were hooked up at the reference plane
( z  0) . The end of the
transmission line is hooked up to the load, located at z  l , whose impedance is
denoted by ZL. The transmission line itself has the characteristic impedance Z0.
A TEM wave traveling in the positive z-direction, towards the load, is referred to
as the incident wave. A TEM wave traveling back towards the source in the
negative z-direction is referred to as the reflected wave.
38
Figure 2-6 Low Frequency Network
Basic components of a low frequency network, showing the source, load, and
reference plane positions along a transmission line.
It is important to understand what is meant by low-frequency. Recall that
for a coaxial cable, a unique voltage and current, and therefore a unique
impedance, could be defined at a particular line cross-section at any frequency
due to the nature of the TEM mode. This was due to the fact that (for the TEM
mode) there are no z-components to the fields and there is no transverse curl in
the electric field, so the integrals used to define the unique voltage and current
along the cable were independent of integration path. However, at the end of the
line, where the load is attached, this will no longer be the case. The EM fields
surrounding the load will certainly not be perfect TEM fields. Therefore, the
exact value of ZL will be ambiguous. For example, the voltage value
V (z  l)
used in defining ZL will not be independent of integration path between points on
the inner and outer conductors of the coaxial cable in the cross-section at z  l ,
where the load is connected to the cable. A further problem also exists; it will be
shown that the amplitude of the reflected power wave depends on ZL (and
therefore the power transferred down the line depends on ZL). However, if the
frequency is low enough, the phase of the EM fields will not change significantly
in the region near the load, and the values for V ( z  l ) and
I ( z  l ) will be nearly
path independent [12]. If one can identify a specific path across the load at z  l
39
along which the voltage and current integrals in Equation 2.35 can be computed,
these values can be used to calculate an approximate ZL that can be used in further
analysis. For the experiments discussed in this work, the network is always being
operated within the low-frequency range.
For a simple network, where a source is delivering power to a load via
transmission line, generally the goal is to eliminate any reflected wave travelling
back to the source. It will be shown that the incident wave and reflected wave
carry power with them independently, so a reflected wave represents power that
is not being transferred to the load.
To eliminate the reflected wave, the
transmission line must be impedance matched with the load. This means the
effective impedance of the load must be equal to the characteristic impedance of
the line. To show this, one begins by combining Equations 2.50 and 2.51 to give
Vcoax ( z )  V0coax e
I coax ( z ) 
V0coax
Z 0coax
i Z coaxYcoax z
e
 V0coax e
i Z coaxYcoax z
 i Z coaxYcoax z
V0coax

Z 0coax
e
 i Z coaxYcoax z
2.53
.
The reflection coefficient can then be defined as
( z ) 
V0coax e
 i Z coaxYcoax z
i Z coaxYcoax z

0coax
V
e

V0coax

0coax
V
e
 i 2 Z coaxYcoax z
.
2.54
Hence, the reflection coefficient at a cross-section z is the ratio of the reflected
wave voltage to the incident wave voltage at that particular cross-section. As we
desire no reflected wave, the goal is to ensure a reflection coefficient equal to
zero everywhere on the line. One can express the impedance of the line as a
function of the reflection coefficient:
40

Vcoax ( z ) V0coax e
Z coax ( z ) 

I coax ( z ) V0coax i
e
Z 0coax
i ZcoaxYcoax z
ZcoaxYcoax z
 V0coax e


0coax
V
 i ZcoaxYcoax z
e
Z 0coax
 Z 0coax
 i ZcoaxYcoax z
1  ( z )
.
1  ( z )
2.55
Rearranging to solve for the reflection coefficient in terms of impedance gives
( z ) 
Z coax ( z )  Z 0coax
Z 0coax  Z coax ( z )
.
2.56
This expression is subject to the end boundary condition
(l ) 
V0coax
V0coax
e
 i 2 Z coaxYcoax l

Z coax (l )  Z 0coax
Z 0coax  Z coax (l )
.
2.57
One can now solve for the voltage phasor ratio to obtain
V0coax

0coax
V

Z coax (l )  Z 0coax
Z 0coax  Z coax (l )
e
i 2 Z coaxYcoax l
2.58
.
The reflection coefficient (Equation 2.54) may then be expressed as
( z ) 
Z coax (l )  Z 0coax
Z 0coax  Z coax (l )
e
i 2 Z coaxYcoax ( l  z )
.
2.59
It is now clear that in order for the reflection coefficient to be zero everywhere
(no reflected wave), the impedance at the end of the line Z coax ( z  l ) must be
equal to (match) the characteristic impedance Z 0coax of the coaxial cable.
41
So far, we have examined how power travels down the coaxial cable, but
only when a wave traveling in one direction is present (Equation 2.45). When
both incident and reflected waves are present, the complex power, like the
impedance, gains an imaginary part that will be a function of the z-position.
However, the real part, which represents time-average power flow down the line,
remains independent of position (as it must). To see this, one first expresses the
voltage and current on the line in terms of the reflection coefficient at the end of
the line
( z  l ) :
Vcoax ( z )  V0coax e
I coax ( z ) 
V0coax
Z 0coax
i Z coaxYcoax z
e
 (l )V0coax e
i Z coaxYcoax z
 (l )
i 2 Z coaxYcoax l  i Z coaxYcoax z
V0coax
Z 0coax
e
e
i 2 Z coaxYcoax l  i Z coaxYcoax z
e
2.60
.
Inserting these expressions into the definition of complex power gives
Pf ( z )  V ( z ) I * ( z )  (V0coax e
V0coax
*
(
Z 0coax
i Z coaxYcoax z
V0coax
 (l )V0coax e
i Z coaxYcoax (2 l  z )
2.61
*
e
 i Z coaxYcoax z
 (l )
Z 0coax
e
 i Z coaxYcoax (2 l  z )
)
).
The real part of Pf ( z ) is therefore given by
Re( Pf ) 
V0coax
Z 0coax
2
 (l )
2
V0coax
Z 0coax
2
.
2.62
Hence, the total time-average power flow down the line is just the difference
between the power that is carried independently by the incident and reflected
waves. The imaginary part of Pf ( z ) is
42
Im( Pf )  2 (l )
V0coax
Z 0coax
2
sin[2 ZcoaxYcoax (l  z )   ],
2.63
where we have explicitly shown the phase constant associated with the reflection
coefficient. The imaginary part represents stored energy along the line; when
there are both incident and reflected waves present a partial standing wave forms
on the line which can store energy. This fact, and the way in which this stored
energy varies with z-position, is important in many transmission line applications.
For our case, since we desire the reflection coefficient to equal zero, the
imaginary part will also equal zero when the line is impedance matched and will
not be crucial to our analysis.
2.2.3 Inductive Coupling
To this point, the characteristics of the EM field eigenmodes that can exist
inside a RCCC have been determined, and the general concepts of power flow to
a RCCC through a coaxial cable have been discussed. However, we have not
discussed the exact manner in which these modes can be excited (or driven) by a
power source. In short, some sort of current source must be present inside the
cavity to stimulate the modes. Due to some dissipation of energy in a real cavity,
this current source must remain inside the cavity, continuously replacing
dissipated energy and keeping the modes running at a constant amplitude.
Different current sources will stimulate different modes, and when a source
stimulates a particular mode we say the source has “coupled” to that mode and is
driving it. In this section, it will be shown how a properly positioned small loop
of current can couple to the TM010 mode inside a cavity. Due to the nature of this
particular coupling process, it is known as inductive coupling.
43
Thus far, it has been shown how (in general) power can be delivered to a
load via a transmission line, specifically using the TEM mode of a coaxial cable.
For our particular purposes, the load that will be receiving power from the
transmission line is the combination of the RCCC and the small loop of current
inside it. As we have also shown, this load (cavity and loop) must appear to the
transmission line to be closely matched to the characteristic impedance of the
line, otherwise power will be reflected before entering the cavity and no mode
excitation will be significantly induced. In other words, the cavity and loop
together must have an impedance equal to the 50 ohm impedance of the coaxial
transmission line. The details of how this can be accomplished will also be
discussed in this section.
The basic arrangement for inductive coupling of a coaxial cable to a RCCC
is shown in Figure 2-7. An aperture is created in the cavity wall that is roughly
the same size as the outer conductor of the cable. The outer conductor is
electrically connected flush with the surface of the aperture (and therefore the
cavity boundary). The inner conductor of the coaxial cable is extended inside the
cavity and curled into a loop, and is then electrically connected to the boundary
of the cavity. If there is dielectric material inside the cavity, a small portion must
be removed to allow space for the loop. Now recall that in order for the coaxial
cable to operate in the TEM mode, boundary conditions on the inner and outer
conductor surfaces require supporting sinusoidal surface currents to be present.
When a TEM wave travels down the line and reaches the aperture surface of the
cavity, as long as the wave is not reflected, the current on the inner conductor
surface current will continue inside the cavity and around the loop. In this way,
a source current is created inside the cavity which can couple to the different
cavity modes. It is important to remember that the TEM wave will be reflected
if the loop-cavity system is not properly impedance matched to the cable, but first
it will be assumed that matching has been achieved, and there is now an
44
oscillating current source inside the cavity following the path of the loop shown
in Figure 2-7.
Figure 2-7 Inductive Coupling
Illustration of inductive coupling using the central wire of a coaxial cable to
form a small loop inside a cylindrical resonant cavity.
We now examine the details of how the current source in the cavity will
stimulate the cavity field eigenmodes that were defined in Section 2.1.3. First,
the general case of an arbitrary current source coupling to modes of an arbitrary
cavity will be considered. It will be assumed that a set of cavity eigenmodes have
been obtained that satisfy Maxwell’s equations for the lossless (arbitrary) cavity
with no current sources, Equations 2.3, with perfect conductor boundary
Conditions 2.9. (This was accomplished for the specific case of the RCCC, with
the TM and TE modes making up the eigenmode set.) The general eigenmodes
will be labelled with index i, so that for the ith eigenmode the associated electric
field will be written Ei, the magnetic field Hi, and the eigenfrequency i . As
Equations 2.3 are satisfied by these modes, we have:
(b)  H i  i i Ei .
(a)  Ei  ii H i
45
2.64
Taking the curl of these equations then give separate wave equations for both Ei
and H i :
(a)    Ei   i 2 Ei  0
(b)    H i   i 2 H i  0.
2.65
Combined with the boundary equations, these two wave equations form an
eigenvalue problem. For the lossless cavity case, the eigenvalues will be real,
and the eigenfunctions Ei , H i form a complete orthogonal set [13]. This means
that any arbitrary electric or magnetic field inside the cavity, even when losses
and current sources are present, can be expanded in terms of the lossless
eigenfunctions as follows,
(a ) E   Ai Ei
(b) H  Bi H i ,
i
2.66
i
where the Ai and Bi are coefficients (phasors) for the ith eigenmode and must be
determined. First it is helpful to normalize the eigenfunctions, which we choose
to do in the following way:
0 i  j
,
1 i  j


  Ei  E j dV   H i  H j dV  
2.67
where V is the volume of the cavity. One can choose the Ei to be purely real
quantities, which will make the H i purely imaginary, by Equations 2.64.
Next, internal cavity current sources and dielectric losses are introduced.
Maxwell’s equations then become
46
(a)  E  i H (b)  H  i  E    E  J s ,
and the corresponding wave equation for
E
becomes
 E    2 E  i  2 E  i J s .
Inserting the eigenfunction expansion for the electric field
 A     E
i
E
2.69
gives
   2 Ei  i  2 Ei   i J s .
i
i
2.68
2.70
Using     Ei   i 2 Ei (Equation 2.65), we obtain (after simplifying and
some rearranging):
 A  (
i
i
i
2
  2 )  i  2 Ei  i J s .
2.71
One can now take the dot product of each side with the complex conjugate
eigenfunction E j * and integrate over the volume of the cavity. Due to the
orthonormal Conditions 2.67, every term in the sum except for
i j
equals zero,
leaving
Ai 
i


i 2   2   i    2
47
J s  Ei*dV .
2.72
This shows that if the current source inside the cavity is known, the coefficients
for each coupled eigenmode can be computed. The total electric field in the
cavity then becomes:
E   Ai Ei  
i
i
i Ei

i     i    2
2
2
 J
s
 Ei*dV .
2.73
The total magnetic field can also be determined using  E  i H (Equation
2.68a) and   Ei  ii H i (Equation 2.64a):
H 
i
ii H i
 
i 2   2   i    2
J s  Ei*dV ,
2.74
and it is clear that the coefficients Bi from Equation 2.66(b) become:
Bi 
ii
  2 
2
2




i
 i
  
J s  Ei dV .
2.75
It has now been demonstrated that if the current source inside a cavity is known
(such as a small current loop positioned as in Figure 2-7, or a beam of electrons
passing through the cavity (see Section 3.3.1), one can theoretically calculate the
integral
 J
s
 Ei dV which will give the amplitude of
Ei and H i for every ith
mode stimulated inside the cavity, in terms of known dielectric constants and
eigenfrequencies.
48
In the above analysis, some key approximations have been made that need
to be further examined. First, the boundary conditions that were implicitly used
required the cavity to be completely enclosed by a perfect conductor.
In
application, the conducting boundary will not be perfect but will have a finite
conductivity that will lead to power loss in the cavity walls as described in Section
2.1.4. This power loss will lessen the field amplitudes. To account for this, a
parameter called the unloaded quality factor Q0 must be introduced. This
parameter is related to the total loss in the cavity due to dielectric loss and the
conduction losses in the walls. This parameter can be measured experimentally,
as will be shown in Section 3.3.4, and the actual field amplitudes can be
calculated using the equations of this section if the quantity     (which accounts
only for the dielectric loss) is replaced by 1/Q0 [11]. For example, Equation 2.74
would become
H 
i
ii H i

i
2

2

iQ
2
 J
s
 Ei*dV .
2.76
0
Second, the cavity is not completely enclosed; there is an aperture where
the coaxial cable meets up with the cavity.
Furthermore, for our specific
experimental application, there will also be two holes (apertures) in the ends of
the cavity to allow the electron beam to pass through. These apertures come with
a boundary condition requiring the transverse field components to be continuous
across them (this condition can be quickly derived from Maxwell’s equations in
the aperture). This requirement will distort the field.
In addition, a current source was assumed to be present in the cavity, but
we did not specify that this source was traveling through a metal conducting wire,
as it will be in the case of the metal coupling loop. The conducting wire adds
49
new boundary conditions; the transverse component of the electric field must be
zero on the surface of the wire inside the cavity. These new boundary conditions
will also distort the field.
Generally, the field distortions come in the form of higher order cavity
modes being stimulated in the cavity, and these modes interfere destructively in
all regions except in the near vicinity of the aperture(s) and the current carrying
wire. In the perturbed regions, the higher order modes interfere constructively
and arrange themselves to meet the actual boundary conditions. The region of
constructive interference of the higher order modes dies off rapidly as you move
away from these regions, and the field begins to resemble the solutions to
Equations 2.73 and 2.74.
2.2.4 Impedance Matching of an Inductively Coupled Cavity
A transmission line analogy has been used to describe the system of a
coaxial cable delivering power to a resonant cavity. According to this analogy,
the transmission line will be impedance matched when it “sees” the cavity as a
load with impedance equal to the characteristic impedance of the line, in our case
50 ohms. In addition, when the line is impedance matched there will be no
reflected power and all power from the source will be delivered to the cavity as
desired.
It is important to understand the limitations of this analogy, and why it can
only be used for approximation purposes when attempting to design an actual
coupling loop to be inserted in the cavity. To do this, one needs to analyze the
EM field that exists in the plane of the aperture that connects the end of the coaxial
cable to the cavity. Since there are no surface charges or currents in the aperture,
boundary conditions require that the transverse EM field components be
continuous across the aperture. This means that one cannot simply have, on the
coaxial cable side of the aperture, a perfect coaxial TEM wave pattern with no
reflected component, and then immediately across the aperture have a perfect
50
cavity eigenmode field. The TEM pattern in the cable and the TM 010 cavity
pattern, for example, do not match up. Again, higher order modes must appear
in the aperture in order to ensure that the fields are continuous across it. Higher
order coaxial cable mode patterns have not been previously discussed, but they
do exist, and together with the coaxial TEM modes form a complete orthogonal
set of functions for the cable cross-section. This means that if the exact transverse
electric field in the aperture were known, it could be expressed as a series
expansion over the transverse coaxial cable eigenmodes as follows:
ETrans 
 E
Trans
z
ETEM
dS 
 E
Trans
z
ETEM
dS    ETrans Ei dS ,
2.77
i
where ETrans represents the actual transverse electric field in the aperture, and the
integrations are over the aperture surface. The first two integrals represent the
projected amplitudes of the TEM modes traveling in the positive and negative zdirections. The last term represents the projected amplitudes onto the higher
order modes, which will not propagate down the cable but instead will die off
exponentially away from the aperture. In order for these higher order modes to
be sustained, energy will need to be stored in the region around the aperture. As
far as the transmission line analogy is concerned, it is important to understand
that a coaxial cable can only be considered a true transmission line when only the
TEM modes are present. This is because it is the properties of the TEM modes
that allow us to uniquely define a line voltage at any point along the cable to use
in the transmission line equations. The existence of higher order modes in the
cable near the aperture means the transmission line analogy is not technically
accurate at the aperture.
Unfortunately, the aperture region is critical in
determining the reflection coefficient.
51
It is, at this point, informative to momentarily abandon our transmission
line analogy and determine what is necessary, from a field standpoint, for the
system to be impedance matched. An effective impedance of 50 ohms (real)
means the reflection coefficient in the aperture is exactly zero. Certainly this
requires the second integral in Equation 2.77 to be zero (no projection onto TEM
mode traveling in the negative z-direction). It will also require the energy stored
in the higher order modes to be equal parts electric and magnetic. One can see
from Equation 2.29, derived by analyzing EM field power flow, that an imbalance
between stored electric and magnetic energy is related to the imaginary part of
complex power; but Equation 2.63 indicates that if an imaginary part of complex
power exists, then the reflection coefficient cannot be exactly zero.
It is now possible to appreciate the difficulty of solving the impedance
problem exactly. Once a particular shape and size for the wire coupling loop has
been designed, one would need to express the field on the cable side of the
aperture as an infinite series in cable eigenmodes with unknown coefficients and,
similarly, the field on the cavity side of the aperture would be expressed as an
infinite series in cavity eigenmodes with unknown coefficients. All coefficients
would then need to be solved for while satisfying the boundary conditions
imposed by the aperture and the conducting metal wire. Once this is achieved all
integrals in Equation 2.77 would need to be calculated. The second integral
would need to evaluate to zero. The energy stored in all higher modes would
need to be calculated and the total stored electric and magnetic energies would
need to be exactly equal. If not, one would need to adjust our loop design and
start over. Needless to say, this is more or less impossible. Many approximation
methods of varying complexity have been developed [14]. These methods
involve assuming a field in the aperture, calculating the induced surface currents
on the wire loop to meet the field boundary conditions on the wire, and then
applying Green’s theorem to the surface of the aperture, wire, and cavity to
52
determine the cavity modes stimulated (the Green’s function is written as a cavity
eigenfunction expansion in the usual way). Once this is done, the fields that are
calculated in the cavity can be added to the original assumed aperture field and
an iterative process can begin, arriving at better and better approximations.
Fortunately, these more advanced approximation methods are not
necessary for our purposes. To obtain an estimate for the required size of our
wire loop, the problem can be treated in a comparatively simple way by
abandoning the transmission line definition of a load impedance for the loopcavity load. The transmission line definition would involve the ratio of a unique
voltage (calculated in the aperture) to the current flowing down the central wire
of the coaxial cable at the aperture. The presence of the higher order modes
makes this definition unusable. Instead, we will define the loop-cavity load
impedance as the ratio of the EMF around the wire loop, in the presence of the
cavity fields, to the current flowing in the central wire at the aperture. In essence,
the current flowing around the wire loop inside the cavity will stimulate fields in
the cavity according to Equations 2.73 and 2.74. These fields will in turn create
an EMF around the loop. This definition of the load impedance as the ratio of
this EMF to the input current at the aperture was developed by Condon [15] and
matches well with our loop design and experimental results.
For inductive coupling to the TM010 mode, one is specifically interested in
the magnetic field stimulated in the cavity near the loop current, because the
periodic magnetic field passing through the loop is what will generate the EMF
around the loop. Several assumptions will be made in this analysis, the first of
which is that we are operating the system at the resonance frequency of the TM010
mode. This will cause the TM010 mode to be stimulated much more strongly than
the other modes, which can then be neglected. From Equation 2.74 one can show
that the H-field stimulated in the cavity will be:
53
H

TM
TM
i010
H 010
i
010TM 

N
2
 
 J  E010
s
TM
*
N
dV 


TM
Q0 H 010
TM
010
N
 
*
s
TM
 J  E010
N
dV , 2.78
Q0
where we have replaced the quantity     with the unloaded quality factor Q0
(since there will be loss in the cavity walls), and the subscript N refers to the
normalized eigenfunctions. The next assumption is about the current in the loop.
We assume the loop is small enough that the oscillating current is in phase around
its entirety. It will also be assumed that, since the loop wire is thin, one can
approximate the current density by simply dividing the total current by the crosssectional area of the wire. The volume element in the above integral will then be
the cross-sectional area of the wire multiplied by a differential length of the wire.
Equation 2.78 then becomes
H

TM
Q0 H 010
TM
010

N

 
Iin
TM
 E010
Ac
*
N
Ac ds 

TM
Q0 H 010
TM
010

N
I in
 
TM
E010
*
N
 ds , 2.79
where we have defined Iin to be the current (phasor) in the central coaxial wire at
the aperture (which we assume continues and remains constant around the loop),
and Ac is the cross-sectional area of the wire. Next, it is assumed that the loop is
a nearly closed circle so that Stokes’ theorem can be applied to the line integral,
giving
H
 
TM
Q0 H 010
TM
010
N
Iin
 
TM
 (  E010
54
*
N
)  dAloop ,
2.80
where Aloop is the area vector for the loop. Equation 2.64(a) can now be used to
replace the curl of the electric field, resulting in
H

TM
Q0 H 010
TM
010

N
I in

TM
TM
 i010 H 010

*
N
 dAloop .
2.81
The integral now involves the dot product of the magnetic field and the loop area
vector. Because of the way the loop is positioned in the cavity, its area vector
will be pointing in the
̂ direction in our cylindrical coordinate system. The
magnetic field for this mode is also pointing in the
̂ direction and its magnitude
will remain relatively constant over the area of the loop, since the loop is small.
The loop is near the outer radius of the cylinder, so the magnetic field magnitude,

TM
which is a function of the  coordinate, will be approximately H 010

N
(   R)
over the area of the loop, where R is the radius of the cavity. The integral then
disappears to give
*
TM
TM
H  iQ0 I in  H 010
(   R ) Aloop   H 010  ˆ.

N
N


2.82
According to Faraday’s Law, this magnetic field will induce an EMF around the
loop
TM
EMFloop  i010
 H loop  Aloop ,
2.83
where Hloop is the magnetic field through the loop. The magnetic field through
the loop is just Equation 2.82 evaluated at   R , implying that
55
TM
TM
2
EMFloop  Iin  2010
Q0  H 010
.
 (  R) Aloop
2
N
2.84
With the definition of load impedance, Zload  EMFloop Iin , we finally obtain:
TM
TM
2
Zload   2010
Q0  H 010
.
 (  R) Aloop
2
N
2.85
This analysis clearly results in the impedance being real, primarily because we
ignored non-resonant modes. The unloaded quality factor Q0 must eventually be
measured experimentally but can be estimated in advance (values are typically
3000-6000 for these types of cavities), the normalized magnetic field can be
evaluated using Equation 2.25 with normalization Conditions 2.67, and the
required area of the loop can therefore be solved for with Zload  50 . As will be
shown in the following chapter, this approximation is sufficient to allow a loop
to be fabricated that achieves impedance matching (no reflected power).
56
3 Experimental Set-Up
A diagram of the system used in the pulse compression experiments is
shown in Figure 3-1. A brief description of the main components and how they
are used together will be given here; more detailed descriptions and
characterizations of certain individual components will be given in the remaining
subsections of this chapter.
Figure 3-1 System Components
Major components for the synchronized laser-driven UEM column used in experiments. Experiments can be
performed with or without the laser amplifier. The interferometer is used only in double pulse experiments. Not
shown is the vacuum system as well as control electronics.
The overall system is driven by a unique laser system that generates a
pulsed laser beam. The pulsed laser beam in turn drives a photoelectron gun,
which generates a pulsed electron beam. The electron pulses are accelerated by
a large potential difference inside the gun, and then directed through the
Compression Resonant Cavity (CRC) using horizontal and vertical deflector
57
plates. Magnetic lenses are used to control the transverse spread of the beam
pulses, ensuring they pass through the CRC without contacting any of the inner
surfaces. While inside the CRC, each electron pulse will interact with an EM
field which is resonating in the TM010 mode, causing changes to the pulse shape
and/or velocity. Depending on the phase of the EM field at the moment each
pulse enters the cavity, the pulse can compress, expand, accelerate or decelerate.
The goal of the system is to maximally compress the pulse while leaving its
velocity unchanged. This means synchronization of the phase of the EM field
with the arrival of each pulse is of the utmost importance. After passing through
the CRC, and assuming the EM field has acted to compress the pulse, the pulse
shape will continue to shorten until it reaches a maximum level of compression
at the temporal focal point of the CRC. At this point the pulse passes through the
Detection Resonant Cavity (DRC), which is capable of determining if the CRC
EM field is at the proper phase for compression, and allows for proper
adjustments to the optical delay stage. For long duration pulses (40-120 ps), the
DRC can also measure the duration of the pulse and determine the power level
when maximum compression has been achieved. A second set of magnetic lenses
and deflector plates are used to guide the pulse through the DRC. The pulse then
travels to a Faraday cup which monitors the electrons per pulse in the beam.
As previously mentioned, synchronization of the phase of the CRC EM
field with the arrival of each pulse at the CRC is critically important. To
accomplish synchronization (and stability) of the EM field phase/electron pulse
relationship, the same laser beam which drives the photoelectron gun is also used
to drive the EM field in the CRC. A small amount of energy from the laser cavity
is directed towards a fast photodiode with a sub-ns response (see Figure 3-1). The
photodiode signal contains all the harmonics of the fundamental laser cavity
frequency up to 10 GHz, and a particular harmonic matches the resonant
frequency of the TM010 mode of the CRC. This harmonic is extracted from the
58
photodiode signal by a standard filter, and then sent to a 30 watt amplifier after
pre-amplification by 20 dB. The high power signal is then sent via low-loss
coaxial cable to drive the impedance matched CRC. Since the CRC and the
photoelectron gun are being driven by the same source, there is no concern for
phase drift of the EM field relative to the generation of the electron pulses at the
photocathode of the 30 kV electron gun. The only concern is time jitter at the
CRC, which will be shown to be low enough to allow successful operation of the
system, due mainly to the high stability inherent to the mode-locked Yb:KGW
laser system. The phase of the EM field relative to the electron pulse arrival is
controlled by an adjustable optical delay stage which is part of the main laser
beam optics.
3.1 Summary of Main Design Considerations
Before describing the individual components of the system, a brief summary
of the main design considerations (and how they affect the design of the
individual components) for the system as a whole will be presented. The first
consideration is the pulse charge requirement, or the number of electrons required
in each individual electron pulse in the pulsed beam. The ultimate goal of the
UEM project at UIC is to use a single electron pulse in a DTEM process to create
an image of fast-moving phenomena (such as atomic vibration in a crystal).
According to the Rose criterion [6], 100 electrons/pixel is required for acceptable
gray-scaling of an image. For a typical 1k x 1k CCD camera with 106 pixels, this
criterion equates to about 108 electrons per pulse [8]. Resolving a diffraction
pattern requires about 106 electrons per pulse [7]. The large number of electrons
per pulse, especially for imaging, creates the need for a highly controlled electron
emission process in the electron gun, since a large percentage of the emitted
electrons need to be directed through the system and to the sample. Typical
emission processes used in traditional electron microscopy, such as thermionic
59
emission, can extract large amounts of electrons but in relatively random
directions from the cathode in the gun; the small amount of electrons that happen
to be emitted in the correct direction towards the sample are then used for imaging
while the others are screened out. This type of process will not be adequate for
imaging with DTEM however. Instead, a process is needed where the initial
direction of the emitted electrons can be much more controlled, and less electrons
are wasted. Therefore, a laser-driven photoelectric emission process must be
used.
This dictates the necessary power of the laser, the laser operating
wavelength, the types of materials that can be used as a photocathode, the
strength and shape of the gun acceleration field, as well as many other aspects of
the system [16, 17].
The pulse charge requirement discussed above leads directly to another
requirement on the system, which also affects the design criteria for multiple
system components. There is a limit to the current density which can be extracted
from a photocathode using a pulsed laser system. When the current density
exceeds this limit, the electric field created by the just emitted electrons, still
located a short distance from the cathode surface, shields the cathode from the
acceleration electric field created by the gun voltage. Since the acceleration field
is necessary for extraction of electrons from the cathode, useful extraction of
electrons is greatly diminished beyond this limit.
For long duration electron
pulses, this limit is expressed by the Child-Langmuir law [18, 19]:
J Pulse
4
 0
9
32
2qe Vg
.
me d 2
3.1
Here JPulse is the magnitude of the current density of a single electron pulse, qe
and me are the charge and mass of an electron, respectively, Vg is the acceleration
voltage of the electron gun, and d is the spacing between the photocathode and
60
anode. This law needs to be modified when the duration of the emitted electron
pulse is much less than the travel time of the pulse from cathode to anode inside
the gun. The new limit becomes [20]:
J Pulse 
 0Vg
Dp d
,
3.2
Where Dp is the electron pulse duration. When the electron pulse is short enough
that this new limit must be used, the system is considered to be operating as an
ultrafast or UEM system. This is not a distinction that is widely accepted in the
literature but it is used by the UIC UEM group. Equation 3.2 enables the
calculation of the minimum HW1/eM electron pulse parameter (wp) coming off
of the cathode which will keep the current density below this short-pulse Child’s
Law limit:
wp ,min 
ne qe d
,
 0Vg
3.3
where ne is the desired number of electrons per pulse. For typical electron guns
used in DTEM, Equation 3.3 gives a value for wp that is approximately 0.5 mm
for extraction of 108 electrons. This is considered a relatively large transverse
size for a pulsed beam, and the effects that this size requirement has on the design
of system components is detailed in Chapter 6 of [21].
A common parameter used to characterize the overall quality of an axially
symmetric Gaussian shaped pulsed electron beam is the normalized rms
transverse emittance [22]:
61
T 
wp
pT 2
mec
2
,
3.4
where c is the speed of light, and pT is the transverse momentum of an electron in
the pulse. The pulsed beam is considered to be of higher quality if the transverse
emittance is low, and it can be shown that in order to achieve a certain spatial
resolving power, the transverse emittance may not exceed a certain value [17].
In Section 3.2 of [21], Liouville’s Theorem is used to show that the transverse
emittance is conserved during electron pulse propagation. For perfect electron
optics, it is unaffected by lensing, compression, etc.; the emittance of the beam
can increase due to imperfect lensing [23], but it cannot decrease. Therefore, the
restriction on the emittance value is present the moment the pulse is generated at
the photocathode. We can combine this restriction with the minimum required
value for wp at the photocathode discussed earlier, and arrive at a maximum
allowed value for the initial standard deviation in transverse momentum for a
pulse. Rearranging Equation 3.4, and using Equation 3.3, we find:
pT 2 
2 0Vg
2 T ,max mec

  T ,max mec.
wp ,min
ne qe d
3.5
Since the initial transverse momentum of the pulse is highly dependent on the
exact process of photoemission, Equation 3.5 has led to extensive research and
experimentation involving the photoemission process and the overall electron gun
design used by the UEM group at UIC [24-28].
The design considerations discussed so far mainly correspond to the design
and construction of the laser system and the photoemission process inside the
electron gun. The last consideration involves the dynamics of a single electron
62
pulse from the moment after it leaves the photocathode to the moment it reaches
the specimen to be imaged. After emission occurs, the pulse will experience a
rapid acceleration inside the gun due to the high voltage between cathode and
anode (the gun voltage); then upon leaving the gun the pulse will be focused both
in the transverse and axial directions by magnetic lenses and the CRC,
respectively. During this time the electrons in the pulse are of course repelling
each other (the space-charge effect). A detailed understanding of the pulse
dynamics is necessary to determine the proper design criteria for the magnetic
lenses and the CRC. For example, a CRC can be designed to run at a wide range
of frequencies and power levels. It can also be operated in any of the cavity
eigenmodes. All of these characteristics will determine the exact focusing forces
that a pulse will experience as it travels through the CRC. Knowledge of the
pulse dynamics is therefore needed to determine the necessary forces for proper
focusing in both space and time.
A theoretical model of the electron pulse dynamics is utilized to determine
the above mentioned design criteria. The model used by the UIC UEM group is
based on the Analytic Gaussian (AG) theoretical model of Michalik and Sipe [29,
30], which includes a scheme to mimic the space-charge effects experienced by
the pulse but excludes external forces. The model was extended by UIC UEM
group member Joel Berger [31] to include external forces simulating gun
acceleration, magnetic lensing and RF compression, and also to add initial
conditions to the pulse propagation which simulate the photoemission process.
Computer simulations using this extended model approximate the necessary
lensing forces and proper positioning of various components. Knowledge of
necessary forces allows for determination of certain design criteria for individual
components, which will be described in the following sections.
63
3.2 Pre-existing Components
As mentioned previously, the experimental results described in this
document are part of a larger project conducted by the UEM group at UIC. Many
of the components used in these experiments were previously developed and
constructed by other members of the group and the details of their operation have
been documented in various published papers and theses. Therefore, these
components will only be discussed in this document in terms of their relevance
to the current experiments, with references given so the interested reader can
access any further detail desired about these components.
3.2.1 Laser Oscillator and Frequency Doubling
In order to produce short electron pulses to be used for imaging, the
electron gun in the UEM system is driven by an ultrafast laser system. The
requirements of the photoemission process inside the gun, and the ensuing pulse
dynamics, dictate the necessary characteristics of the short laser pulses hitting the
cathode. The pulses must be below a certain duration, have the necessary photon
energy (frequency) and the necessary pulse energy to generate the proper amount
of electrons per pulse. The laser system used by the UIC UEM group was
designed and built in-house by group members in order to meet the necessary
pulse requirements. There are three main aspects to the system which will be
briefly described in this section, they are; the crystal gain medium used in the
oscillator, the thermal lens shaping technique employed to counter astigmatism
inherent in the system, and the mode-locking process necessary for a pulsed
output. Details regarding these aspects of the system (and many others) can be
found in [16, 21, 32, 33]
The laser oscillator uses a Ytterbium-doped potassium gadolinium
tungstate (Yb:KGW) crystal gain medium. This material has a wide absorption
spectrum around 980 nm, and a wide emission spectrum around 1040 nm.
Absorption and emission efficiency are dependent on many factors, including the
64
polarization and incidence angles of the radiation relative to the crystallographic
axes of the material. By employing a particular crystal geometry, along with the
application of a 193 nm thick SiO2 anti-reflection coating, both of these properties
can be exploited. The crystal is pumped by two 35 W laser diodes at 980 nm
(about 94% absorption efficiency), so that access to the wide emission spectrum
at 1040 nm is achieved. The wide laser gain spectrum allows for generation of
ultrashort mode-locked pulses.
The oscillator system employs a thermal lens shaping (TLS) technique [32]
to account for astigmatism inherent in the particular crystal geometry. Based on
the known angles involved in the crystal geometry and the crystal material
properties, a mathematical analysis can be performed to determine the heat
distribution required in the crystal to create the proper elliptical temperature
distribution that generates the thermal lens compensating for astigmatism. This
leads to specific requirements for design of the laser diode pumping system (the
heat source), including the necessary elliptical pump spot size.
The oscillator is capable of self-starting mode-locked operation (short
pulse generation) above a certain power level. Mode-locking occurs in a laser
cavity when a wide bandwidth of radiation is emitted, and the various cavity
resonant frequencies of radiation present all oscillate with a constant phase
relationship that creates constructive interference in a small temporal region (the
pulse), and destructive interference everywhere else. The pulse then travels
through the cavity with a particular group velocity and a pulsed beam becomes
the output of the laser cavity, with a pulse repetition frequency equal to the
fundamental oscillation frequency of the laser cavity.
The output from the main oscillator cavity is a 28.5 MHz train of pulses
with an average power of 1.9 W, a spectrum centered at 1046 nm, and a pulse
duration of 250 fs. The photon energy at 1046 nm is generally not high enough
to emit electrons from most photoemission sources [34]. To increase the photon
65
energy to a level appropriate for photoemission, the radiation from the oscillator
undergoes two frequency doubling processes. The first process uses a 3 mm noncritically phase matched Lithium triborate (LBO) crystal [35]. The frequency
doubling occurs with an efficiency of 40-50%, resulting in an output of 532 nm
(green) radiation at a power of 0.8-1 W and leaving the pulse duration relatively
unaffected. The second process employs a 6 mm  -Barium borate (BBO) crystal
which generates 261 nm (UV) radiation.
This time the pulse-duration is
significantly affected, due to group velocity mismatch effects, resulting in a pulse
duration of approximately 4 ps. However, the photon energy is now 4.75 eV
which is appropriate for photoemission from many photocathode materials. It is
this 28.5 MHz UV radiation which strikes the photocathode in our experiments
that do not involve the laser amplifier.
3.2.2 Laser Amplifier
To produce the necessary pulse energy in the UV for the generation of
sufficient electrons for single-shot imaging applications, the ultrashort pulse laser
system dedicated to the UEM project at UIC includes a diode-pumped Yb:KGW
regenerative amplifier. The design of the regenerative amplifier cavity is based
on the same thermal lens shaping technology employed for the mode-locked
Yb:KGW oscillator. In this case, the cavity is a simple z-fold around the gain
medium with a Rubidium titanyl phosphate (RTP) transverse Pockels cell and
sapphire Rochon polarizer [36] in one arm to facilitate the switching in and out
of the amplified pulse. Standard chirped-pulse amplification techniques [37] are
employed to avoid optical nonlinearities in the amplification process: 1000
lines/mm transmission gratings with approximately 90% efficiency in first order
(Littrow configuration) both stretch the incident 250 fs pulse to nearly 100 ps
prior to amplification and recompress the amplified pulse. Output amplified
pulse energies of up to 0.1 mJ and compressed pulse durations of approximately
500 fs are obtained at operational repetition rates of around 600 Hz. These
66
compressed pulses are frequency doubled in a 2 mm BBO crystal and
subsequently doubled again to 261 nm in the same BBO crystal employed to
frequency up convert the Yb:KGW oscillator output; again yielding roughly 4 ps
UV pulses, but with an energy of several micro-joules.
3.2.3 UEM Column Design
All of the system components displayed in the assembly in the lower right
corner of Figure 3-1 are considered to be part of the UEM Column (the Faraday
cup is not a permanent column component but is used to terminate the column in
the experiments). Since the electron beam must be contained in a high-vacuum
environment to allow free passage of the electrons, the column is essentially a
vacuum chamber constructed from purchased vacuum hardware, with working
components contained inside the chamber. The hardware includes ports for use
with various electrical pass-throughs to power internal components. Besides the
vacuum hardware and high-vacuum system, the main pre-existing column
components are the photoelectron gun, the first set of deflector plates, and the
magnetic lenses.
The photoelectron gun design is based on Togawa [38] and modified by
Berger [17] to meet the needs of the UIC UEM group. The major consideration
specific to the UIC group is the fact that the gun must accommodate a larger beam
spot size on the photocathode, due to reasons described in Section 3.1. The gun
consists of a Wehnelt cylinder with the photocathode positioned in the aperture,
and a large aperture anode. Along with a set of deflector plates which help guide
the electron beam down the column [21], the gun is enclosed in what is called the
main UEM vacuum chamber. A valve in the bottom of this chamber opens the
entire column to the vacuum pump system. The Wehnelt/photocathode assembly
is mounted to the port aligner, which allows for precise alignment of the cathode.
Positioned behind the port aligner is the high voltage pass-through, which
provides the high voltage electrical connection to the cathode and creates the
67
acceleration field (gun voltage) between cathode and anode. The anode and the
deflector plates are mounted inside the main chamber using special mounting
hardware, which is also used to position a second set of deflector plates to be
described later.
The set of two magnetic lenses was designed and built in-house by Berger
and the details of construction and operation can be found in [21]. The design
was guided by the AG model of the pulse dynamics through the column. Again,
a major design concern was the fact that a larger beam cross-section requires
larger apertures in the lenses. Even though the CRC had not been built at the time
the magnetic lenses were designed, the AG model could determine the beam
divergence that would be caused by the CRC and show that the magnetic lenses
could compensate for the CRC lensing in the pulse compression regime.
Divergence in the beam is also caused by the gun anode aperture [21]; again the
AG model could predict the resulting effect and this was also built into the
magnetic lens design.
3.3 Compression Resonant Cavity (CRC)
The Compression Resonant Cavity (CRC) is the first of two RF cavities to
be built and installed in the UIC UEM system. The details of the operation,
construction, and characterization of the device is described in the following
subsections.
3.3.1 Operation
During the photoemission process in the electron gun, each electron pulse
will develop a certain level of temporal expansion due to velocity-induced
dispersion; individual electrons will be emitted with a range of initial velocities
due to the nature of the photoemission process, and electrons with higher
velocities will move to the front of the pulse while slower electrons lag in the
rear, as the pulse as a whole travels down the UEM column. Space-charge effects
68
can also contribute significantly to the longitudinal expansion of the electron
pulse [7]. The purpose of the CRC is to reverse the resultant velocity chirp in
each pulse, essentially slowing the faster moving electrons in the front and
speeding up the slower electrons in the rear. After the pulse leaves the CRC, the
electrons in the pulse will then begin to come together as the pulse moves through
a section of drift tube, eventually converging to a minimum time duration at the
temporal focal point of the CRC.
A simple model can be employed to demonstrate the basic operating
principle of the CRC and how it achieves velocity chirp reversal (and therefore
pulse compression). In our experiments, the cavity is operated in the TM010 mode.
The electric field of this mode has an overall sinusoidal time dependence, points
in the z-direction, and has no variation in the z-direction at any particular time.
The maximum value for the field occurs on axis where an electron pulse will
travel. Therefore, the electric field interacting with the pulse as it travels through
the CRC can be written as:
Ez (t )   E010TM 
max
sin(t   ).
3.6
Here, the variable t is assumed to be zero as the pulse enters the cavity, and φ is
therefore the phase of the electric field as the pulse enters the cavity. If the
angular frequency of the field is written in terms of the cavity length, then the
following analysis can be used to determine the ideal field phase constant and the
ideal cavity length to achieve maximum compression.
The force that any
particular electron within the pulse experiences, due to this electric field, is
69
TM
Fz  qe  E010

max
sin(t   )
3.7

2
2
2 v


,
T
Ad v Ad
where d is the cavity length, v is the average electron velocity, and A is a unitless
parameter (to be determined) which relates the period of the field to the cavity
length.
Since the field in the cavity is changing in time, as the pulse passes through
any particular point along the CRC axis, the electrons at the front of the pulse will
feel a different force at that point than the electrons at the back, which arrive at
that point a pulse duration of time later. As the pulse duration is short (~4 ps)
compared to the field period (~330 ps), it can be assumed the change in the field
strength is linear, at any point, over the pulse duration. As a result, a simple
expression for the differential force, between front and back, that the pulse (of
duration τ) experiences at any point in the cavity may be written:
Fz ,diff  qe


d
TM
TM
E010
sin(t   )  qe  E010


max cos(t   ). 3.8
max
dt
This expression can be integrated over the time the pulse is inside the cavity, to
give the total compression impulse delivered to the pulse. This impulse is what
causes reversal of the velocity chirp. To perform the integration we change
variables from time to the cavity z-coordinate and then integrate over the length
of the cavity:
70
I z ,comp   qe  E010
TM
t t d
  cos(t   )dt
TM
  qe  E010

max
t 0
z d
max
z
dz
 cos( v   ) v
z 0
d
z
  qe  E010  sin(   )
max
v
0
TM
3.9
2 v d


sin(


)

sin(

)

max 
Ad v

TM
  qe  E010
max sin( 2A   )  sin( )  .
TM
  qe  E010

For maximum compression impulse (which must be positive for compression),
the quantity in square brackets must equal its maximum possible value of -2,
which is achieved with A=2 and φ=π/2. The value of 2 for the parameter A means
the cavity length should be set to allow for a pulse travel time through the cavity
equal to one-half of the field period. The pulse dynamics of the AG model can
predict the value for velocity chirp, and therefore the necessary compression
impulse to reverse the chirp, allowing a calculation of the necessary field strength.
It should be pointed out that, in addition to compression (or expansion), the
CRC has the ability to accelerate (or decelerate) the center of mass of the pulse.
To determine this acceleration one simply has to integrate the force experienced
by the center of mass of the pulse over the time in the cavity, instead of the
differential force experienced by the whole pulse. It is found that the phase of
the EM field is what determines the effect of the CRC on the pulse. If the pulse
arrives at the CRC when the phase is exactly π/2, the center of mass will not be
accelerated and only compression occurs.
Before attempting to determine the design parameters (dimensions,
operating frequency, etc.) for the CRC, the effect of the cavity’s 2 mm axial hole
must be investigated. Up to this point, the cavity eigenmodes that have been
71
studied were the solutions to a RCCC geometry that was assumed to be
completely filled with a single dielectric material (or vacuum). When an axial
hole is added, the problem is separated into two regions; the hole and the
dielectric surrounding it. A general form for the solution to the field in each
region is applied, and then boundary conditions are enforced at the outer
conducting wall as well as at the interface between the hole (vacuum) and
dielectric. It will now be shown that a 2 mm axial hole does not significantly
affect the field pattern or resonant frequency of the TM010 mode of the CRC.
For the region of the axial hole, the general solution for the field is simply
 2 f   i 2 ft
E0 J 0 
( zˆ),
e
 c 
3.10
which is Equation 2.25 expressed in terms of frequency. For the dielectric region,
the general solution becomes:

 2 f  r  
 2 f  r    i 2 ft

BE
Y
.
 AE0 J 0 

  e
0 0

c
c




 
3.11
Here the Bessel function of the second kind is included. This function has a
singularity at   0 , but since the dielectric region no longer contains the axis,
this second solution to the second order differential equation in Equation 2.20
cannot be discarded as it was previously. A and B are constants to be determined
by boundary conditions. The first boundary condition requires the tangential Efield to be zero at the outer conducting wall
72
(   R) . This results in:
 2 f  r R 
 2 f  r R 
AJ 0 
 BY0 

  0.



c
c




3.12
Next, the tangential E-field must be continuous across the interface between
dielectric and vacuum
(   a) :
 2 f  r a 
 2 f  r a 
 2 fa 
AJ 0 
 BY0 
 J0 


.




c
c
c






3.13
Finally, the tangential H-field (Equation 2.25) must also be continuous across the
interface:
 2 f  r a 
 2 f  r a 
 2 fa 
A  r J1 
 B  r Y1 
 J1 


.




c
c
c






3.14
The above three boundary condition equations will have a solution
provided
 2 f  r R 
J0 


c


 2 f  r R 
Y0 


c


 2 f  r a 
J0 


c


 2 f  r a 
 r J1 


c


 2 f  r a 
Y0 


c


 2 f  r a 
 r Y1 


c


73
0
 2 fa 
J0 
  0.
 c 
 2 fa 
J1 

 c 
3.15
Solving this equation provides the modified resonant frequency value for a
particular axial hole radius a, and this value can be inserted into the boundary
condition equations to solve for constants A and B, giving the field solution.
A Mathematica program was written to solve for the eigenfrequencies of
the YAG-filled CRC with axial holes of varying radius, with the results shown in
Figure 3-2. From the plot it is clear that for small axial holes there is little effect
on the resonant frequency. An experiment was performed to test the results of
this theory (using a different dielectric material); the results are presented in
Chapter 4, showing good agreement.
Figure 3-2 Axial Hole Effect on CRC Frequency
The effect of the axial hole, drilled through the YAG-filled CRC (radius
13.25 mm) to allow passage of electrons, on the cavity’s resonant frequency.
For holes of small radius the change in frequency is negligible. The actual
CRC axial hole radius is 2 mm.
The Mathematica program was also used to solve for the unknown
coefficients A and B and subsequently the field pattern with the 2 mm axial hole
present. These results (with E0  1) are shown in Figure 3-3. It can be seen that
the overall field pattern is only slightly disturbed by the presence of the axial hole.
In particular, inside the hole the field magnitude is essentially the same, meaning
74
the electric force felt by the electron pulses passing centrally through the hole will
remain the same. For these reasons, when doing any analysis involving the CRC
field, we can replace the more complicated exact solution for the field with the
solution for a fully filled cavity with no hole, using an effective dielectric constant
that corresponds to the same resonant frequency as the exact solution. The
dielectric constant for YAG at 3 GHz is estimated to be about 10.4 [39], and the
effective dielectric constant for the CRC will be discussed in Section 3.3.4.
Figure 3-3 Axial Hole Effect on CRC E-Field
The electric field pattern for the CRC TM010 mode with and without the 2 mm axial hole,
showing that the overall field pattern is not greatly affected, and the field magnitude in the hole
(where the electrons interact with the field) is nearly identical.
3.3.2 Design
There are many design parameters and constraints related to construction
of the CRC. Parameters are entered into the AG model to determine if the
requirements for pulse compression are met, and then various relations are used
to check if these parameters allow successful operation of the CRC based on the
various constraints. A process of trial and error eventually led to a set of
parameters that meets all of the constraints and achieves proper pulse
compression.
75
The main constraints on the CRC design will be discussed first. There are
two size constraints; the CRC radius must be small enough to allow it to be easily
mounted to the UEM column, and the temporal focal length of the CRC must be
short enough to fit the length of the column. There is a constraint involving the
velocity of the accelerated pulse as it enters the CRC; a high velocity is desired
to reduce the time duration of the pulse for a given compressed length (improving
time resolution), however, according to pulse dynamics simulations, a higher
pulse velocity will involve a larger velocity chirp for each pulse, which in turn
requires a stronger CRC E-field to reverse the chirp. This leads to power
considerations; a stronger E-field increases the average stored electromagnetic
energy in the cavity which will require extra power to maintain due to losses in
the cavity. High power sources at RF frequencies are extremely expensive, and
they become more expensive as the operating frequency increases for a given
power level; this creates a constraint on the operating frequency. The velocity of
the pulse itself is constrained by the maximum available voltage creating the
acceleration field in the photoelectron gun. At extremely high gun voltages there
will be problems with stability and arcing.
The design parameters and their values will now be given, and it will then
be shown that these values allow the CRC to operate within the aforementioned
constraints. The first parameter involves the dielectric constant and loss tangent
of the dielectric material chosen to fill the cavity. When the CRC is being run in
the TM010 mode, the main purpose of a dielectric insert (with an axial hole to
allow free passage of the pulses) is to reduce the radius of the cavity to more
easily fit in the UEM column. This is because for a given frequency, the radius
is proportional to 1
 r . Any size reduction achieved must be weighed against
increased power losses in the dielectric, so a low loss tangent is also desired. A
crystal material with low loss as well as a large dielectric constant of
76
approximately 10.4 at 3 GHz is yttrium-aluminum-garnet or YAG [39], and this
was the material selected for the CRC.
Additional parameters that need assigned values include the operating
frequency of the CRC and the photoelectron gun acceleration voltage. A higher
frequency allows for additional size reduction and lower power needs, as will be
shown. A frequency of 3 GHz was chosen, as high stability power supplies at
this frequency are available and (relatively) affordable. The gun voltage was set
at 25 kV, which determines the pulse velocity. The length of the CRC is chosen
to correspond to a time of flight inside the cavity equal to a half period of the
CRC EM field. This allows the pulse to enter the cavity when the field is at a
maximum in the direction of the pulse velocity, and leave the cavity when the
field is at a maximum in the opposite direction, allowing for maximum
compression.
With these chosen parameter values, AG model simulations showed a
maximum CRC (time-dependent) E-field strength of approximately 0.2 MV/m
would be necessary to completely reverse the expected velocity chirp (from the
photoemission process combined with a gun voltage of 25 kV) with a temporal
focal distance of 28.8 cm, an acceptable focal distance for the UEM column. What
remains to be shown is that the CRC cavity size and the necessary power level
are also acceptable within the constraints.
The cavity dimensions are
straightforward to evaluate. From Equation 2.24 the radius of the cavity is given
by
R
2.405c
.
2 f  r
3.16
For an operating frequency of 3 GHz and a dielectric constant of 10.4, the
resulting cavity radius is 11.9 mm. However, since the axial hole was expected
77
to lower the effective dielectric constant of the CRC, a slightly larger radius of
13.25 mm was chosen. Since the time of flight for a pulse inside the cavity should
ideally be equal to a half-period, the cavity length (denoted by d) is simply related
to the pulse velocity (vp) and the frequency, which can be further related to the
gun voltage:
d
vp
2f

1
2f
2Vg qe
me
 2.963  105
Vg
f
.
3.17
For a gun voltage of 25 kV and a frequency of 3 GHz, this results in a CRC length
of 15.6 mm. These dimensions (radius and length) for the CRC allow the cavity
to easily fit in the UEM column.
One can now show that the CRC will operate under the stated parameters
at an acceptable power level. First, the unloaded quality factor Q0 must be defined
for a resonant cavity in terms of the cavity’s stored energy and power loss. In
general, a quality factor for any oscillating physical system is defined as the ratio
of time-average stored energy, to time-average energy lost per radian of a cycle.
There can be different quality factors defined for a system depending on which
type of stored energy, and which type of energy loss, are being referred to in the
ratio. The unloaded quality factor for the CRC refers to the total time-average
stored EM energy in the CRC and the total time-average energy loss per radian:
Q0 
2 f (We  Wm )
.
Pf  Pd , 
3.18
Here energy and power terms are used that were defined previously in Equations
2.28 and 2.34. For the CRC, the power terms represent the time-average power
loss in the conducting walls and the dielectric loss. The total time-average cavity
78
power loss PT can be defined as the sum of these two terms. Equation 3.18 can
then be rearranged to give
PT 
2 f (We  Wm ) 4 f We

,
Q0
Q0
3.19
where we have used the fact that stored electric and magnetic energies are equal
at resonance. This expression now needs to be expressed in terms of the design
parameters. First, we replace the stored energy term:
4 f  12    E010 (  )

PT 
Q0
TM
2
R
2


TM
2
4

f


d
E
(   0)
J

d




0
r
010
0
0
dV
 


Q0
3.20
2

TM
2
4 2 f  0 r d E010
(   0) (0.135 R )
Q0
.
Next, Equations 3.16 and 3.17 are used to replace R and d in terms of the design
parameters, and the electric field phasor magnitude is replaced with the timedependent field magnitude (which is the actual field design parameter):


 2.405c
Vg   1 TM
2
5
4 f  0 r  2.963  10
  E010 (   0)   0.135 
f  2


 2 f  r


PT 
Q0
2



2




3.21
 9.195  1010
E010TM (   0)
2
Q0 f 2
Vg
.
79
With the values given for the design parameters, this equates to a power level of
12.9 W if the CRC unloaded quality factor is 5000. Factors of 5000 are easily
attainable for resonant cavities, and the CRC in fact has an unloaded quality factor
of 5080 as will be shown in Section 3.3.4. A 30 W power source at 3 GHz is
available to the UEM lab, so the design constraints have been met.
3.3.3 Construction
The components and subassemblies that make up the CRC are shown in
Figure 3-4. First a general description of the construction and assembly of the
major components will be given, and more detail will then be provided as needed.
The cavity is located in the center of a steel housing piece which allows the CRC
to be mounted to the UEM column. There is a copper insert in the center of the
housing, and the inside surface of this insert makes up the cylindrical boundary
of the cavity. The YAG dielectric insert fits into the cavity and is secured by two
end caps which are screwed onto the main housing. The center circular surfaces
of the end caps act as the end boundaries of the cavity. To deliver power to the
cavity, a hole is drilled through the steel housing and copper insert to allow
insertion of the coupling loop assembly. The loop assembly is attached to a
vacuum pass-through with SMA (SubMiniature version A) coaxial connectors,
and mounted to a vacuum flange at the end of a steel tubular extension that is
welded to the housing. When mounted, the small coupling loop at the end of the
loop assembly will enter the cavity and fit into a notch drilled into the side of the
YAG insert. The vacuum pass-through has a female SMA connector on the nonvacuum side which allows connection to the power source.
80
Figure 3-4 CRC Components
Individual components of the Compression Resonant Cavity. The end caps and steel housing with
copper insert, along with the loop assembly, were designed and constructed in-house.
Figure 3-5 shows detail of the coupling loop construction. A small length
of coaxial cable is fitted with an SMA female connector at one end. The other
end is stripped so that a small length of the central copper conducting wire is left
exposed. A machined brass end piece is press-fit over the end of the cable, and
the central wire is bent into a loop. The end of the wire is then press-fit into a
hole drilled into the brass end piece, which is electrically connected to the cable’s
outer conductor. As the loop assembly is being inserted into the cavity, the wider
section of the brass end piece acts as a stopper so the loop is inserted the proper
distance. Once the loop is inserted into the cavity and properly oriented, a set
screw (shown in the figure on the right) is used to tighten the end piece in place.
81
Figure 3-5 Coupling Loop Detail
Detail of the coupling loop used to deliver power to the CRC. The
loop can be rotated while inside the CRC, and fixed in place using the
set screw once impedance matching has been achieved.
Figure 3-6 shows detail of the copper insert, the through-hole for the
coupling loop, the through-hole for the set screw, and how the coupling loop fits
into the cavity and is tightened by the set screw (shown without the YAG insert
present, so the loop can be clearly seen). The copper insert is machined so its
outer diameter is slightly larger than the diameter of the central hole in the steel
housing. The housing is then heated, and a cooled copper insert is fitted into the
hole in the housing. When the assembly reaches room temperature the copper is
securely fit into the housing, as shown in the picture. The loop assembly throughhole is then drilled through both the steel and copper and into the cavity. Another
small through-hole is drilled into the edge of the copper insert and tapped to allow
for a set screw as shown. When the loop is inserted, the end surface of the brass
end piece actually becomes part of the cavity boundary. The set screw can then
be tightened to hold the loop in place (although during actual assembly this is
done after the end caps have been attached). The inside surface of the copper is
highly polished to decrease the effective skin depth of the copper, which
minimizes conduction losses in the cavity walls.
82
Figure 3-6 Cavity/Coupling Loop Detail
Top: steel housing, cavity, and coupling loop through-hole are shown.
Bottom Right: close up of the highly polished inner copper surface, loop
through-hole and the through-hole for the loop tightening set screw.
Bottom Left: close up of the inserted coupling loop and set screw.
Figure 3-7 shows the detail of the copper end caps and the inside of the
cavity with the back end cap attached. In the top picture in the figure, the “back”
end cap is shown on the right. The only difference between the two pieces is a
through-hole (seen in the picture) drilled into the “front” end cap to allow
tightening of the set screw. Both caps have a circular groove milled into their
surfaces to allow for a strip of indium wire to be placed (the front cap is shown
without the wire so the groove can be seen clearly). The surface of each cap that
is inside the groove will act as a boundary for the cavity when attached, and these
inner surfaces are highly polished. In the bottom right picture in the figure, the
back end cap has been screwed onto the housing, and the close up picture on the
83
bottom left shows how the indium wire, which is extremely malleable, fills in any
gaps between the end cap and the central copper insert. This is extremely
important due to the current pattern of the TM010 mode shown in Figure 2-3. Any
gap between end cap and cylindrical copper insert will cause major disruptions
in the current flow, and this greatly affects the field pattern and resonant
frequency (as well as impendence matching and the total quality factor). The
indium wire effectively stabilizes the current flow and allows for consistent
performance of the CRC.
Figure 3-7 End Cap Detail
Detail of the CRC end caps showing how the indium wire is used
to aid in current flow around the end boundaries of the cavity.
Figure 3-8 shows a close-up picture of the YAG dielectric insert on the left.
The axial hole as well as the clearance notch for the coupling loop can both be
clearly seen. The surface of the YAG insert is highly polished which allows for
the possibility of a copper coating process, which was successfully attempted
using a fused silica dielectric insert shown on the right side of the figure. This
84
process leaves the dielectric with about a 60 micron thick coating of copper,
which is much greater than the copper skin depth at 3 GHz and therefore serves
as an appropriate cavity boundary surface. The copper coating also allows for
improved current flow in the TM010 mode due to the indium wire (and end caps)
no longer being involved in the flow. While the indium allows for more stable
current flow for an uncoated dielectric-filled cavity with end caps, it still causes
minor disruptions in the flow pattern due to varying thickness and slight
protrusions into the cavity when the end caps are tightened down. This creates a
less efficient production of magnetic flux by the current surrounding the cavity,
essentially reducing the effective inductance of the cavity. The effect is similar
to reducing the inductance in an LC circuit, and it is well known that a resonant
cavity can be modeled as an equivalent LC circuit for purposes such as calculating
the resonant frequency [11]. In an LC circuit, a reduction in inductance will cause
the resonant frequency to increase. This is indeed the case for the fused silica
insert. The resonant frequency of a cavity consisting of an uncoated fused silica
insert, end caps and indium wire was measured to be 6% higher than the coated
fused silica cavity. The unloaded quality factor of the coated cavity was also 20%
higher signaling less loss in the coated walls. These same results would be
expected for a coated YAG insert.
85
Figure 3-8 YAG Dielectric Insert and Coated Fused Silica
Left: YAG dielectric insert showing axial hole and clearance hole for the
coupling loop. Right: Fused silica dielectric coated with a 60 micron thick layer
of copper, forming a stand-alone resonant cavity.
For implementing a coated cavity in compression experiments, the
mounting process would still be the same, and the same housing hardware can be
used. The end caps and copper insert would simply be holding the coated insert
in place, and would no longer be an actual part of the cavity. The coated fused
silica cavity is shown inserted into the DRC housing in Figure 3-9.
Figure 3-9 Coated Fused Silica Cavity in DRC
Copper coated fused silica cavity inserted into the DRC housing to
allow for testing and characterization.
86
For CRC experiments, the YAG insert was left uncoated and the indium
wire was used to create stable current flow. This was due to the fact that an RF
filter had already been selected that matched the uncoated cavity’s higher
frequency. While a 20% increase in unloaded quality factor would be desired and
a coated cavity is the better option, it was not necessary to successfully perform
experiments.
Figure 3-10 shows the next step of CRC assembly, the insertion of the
YAG dielectric and the coupling loop. In the bottom right photo, the dielectric
has been fit into the cavity with the notch for the coupling loop aligned with the
loop through-hole. In the bottom left, the loop assembly has been inserted (the
loop can be seen under close inspection) and the set screw is being tightened. The
top photo shows the entire housing with YAG and loop assembly inserted.
87
Figure 3-10 YAG Insert with Coupling Loop
CRC housing with YAG insert in place and coupling loop assembly
inserted into the housing. In the lower left picture the loop can be seen
inside the YAG clearance hole (upon close inspection), ready to be fixed
in place by the set screw.
Figure 3-11 shows detail of the loop assembly and vacuum pass-through
being inserted into the housing. On the left the loop assembly is shown connected
via SMA connectors to the purchased vacuum pass-through component. The top
right photo shows the top of the tubular extension with a purchased vacuum
flange welded on. The extension piece itself is welded onto a flattened section of
the main steel housing as can be seen in the photo. In the bottom right photo a
copper gasket has been placed into the groove of the vacuum flange. Next, in the
middle top photo, the loop assembly is inserted into the cavity until the brass
stopper prevents any further insertion. Finally, the middle bottom photo shows
the top vacuum flange screwed down creating the vacuum seal. As the top flange
is tightened down, the copper gasket will flatten a small amount and push the loop
assembly further into the housing. Since the brass stopper prevents the loop from
88
moving further into the cavity, the coaxial cable will be forced to bow to the side
slightly to account for this extra insertion. Thus, a minimum length of cable must
be used in the loop assembly.
Figure 3-11 Vacuum Pass-Through Assembly
Left: vacuum pass-through connected to the coupling loop assembly via SMA connector.
Right: various stages of insertion into CRC housing with vacuum seal.
Once the top vacuum flange has been put into position, but before the
screws have been tightened, the loop assembly can still be rotated, changing the
orientation of the loop inside the YAG insert (the notch in the insert allows for
90 degree rotation). This is important for impedance matching purposes. During
the actual impedance matching process the CRC will be hooked up to a network
analyzer (after the front end cap has also been attached) to determine when the
CRC is properly matched. The exact loop area necessary for impedance matching
can only be estimated beforehand, as mentioned in Section 2.2.4. Therefore, in
order to match the cavity, the loop is rotated slightly (changing the effective area
of the loop receiving magnetic flux) until impedance matching is seen on the
89
analyzer. At this point the set screw is tightened (there is a through-hole in the
front end cap allowing access to the set screw), fixing the loop in place. Lastly,
the screws of the top vacuum flange are tightened down and the CRC is ready for
insertion into the vacuum system.
Figure 3-12 shows the fully assembled CRC. In the bottom picture the
small through-hole in the front end cap is shown with a wrench tightening the set
screw. The top photo shows how a copper gasket fits into a machined groove in
the steel housing (an identical gasket is fitted to the other side). The CRC can
now be mounted to the UEM column using the large threaded outer holes of the
housing, which compress the gaskets and create a vacuum seal.
Figure 3-12 Completed CRC Assembly
Completed CRC assembly ready to be mounted to the UEM
column using large copper vacuum gaskets. The end cap
through hole allows the coupling loop to still be rotated for
impedance matching.
90
3.3.4 Characterization
A resonant cavity (operating in a particular mode) can be fully characterized
by identifying three parameters: the resonant frequency, the coupling constant κ,
and the total quality factor. The coupling constant is related to how well the
cavity is impedance matched, and will be equal to unity if the cavity is perfectly
matched, meaning the impedance of the cavity and coupling loop combined is 50
ohms at the resonant frequency. To identify these parameters, the frequency
response of the coupled cavity in the vicinity of the resonant frequency must be
analyzed. To perform this analysis, a resonant cavity is typically connected to a
network analyzer, which sends out a swept frequency reference signal. The signal
(magnitude and phase) that is reflected back to the analyzer by the cavity is then
recorded at each frequency. A computer program is then used to analyze this data
and obtain the parameters.
An expression for the frequency response of a resonant cavity (operating in
the TM010 mode) can be developed from the theory already discussed, in order to
gain an understanding of how the computer program analyzes the data, but more
importantly to show how the size of the coupling loop has a critical effect on
impedance matching. First, a frequency-dependent expression must be developed
for the load impedance of the CRC. In Section 2.2.4 a simplified version was
provided to help approximate the necessary coupling loop size. There, the cavity
was assumed to be operating at exactly the resonant frequency, which eliminated
the frequency-dependent part of the expression. The frequency dependence
simply needs to be put back in and the identical development can be repeated.
Starting with Equation 2.76, repeated here specifically for the TM010 mode (the
other modes can still be neglected since we are interested in the frequency range
very near the TM010 resonance);
91
H 010 
TM
 

TM
TM
i010
H 010
TM
010

2

2

N
iQ
2

s
TM
 J  E010

*
N
dV ,
0
the volume integral can be evaluated the same way as before giving,
TM
H 010

 
TM
 010

TM
010

2

2
2

  I  H TM * (   R) A   H TM  ˆ.
loop 
 010 N
  in 010 N
i
Q0
2
3.22
After repeating the remainder of the derivation, an expression is obtained for the
load impedance as a function of frequency near the TM010 resonance:
TM
i 010

2
Z load 
 
TM
010

2

2

  2 H TM (   R ) 2 A2  .
 010 N
loop 
 2 

i
Q0
3.23
The real and imaginary parts of the complex quantity Zload are plotted in Figure
3-13.
92
Figure 3-13 CRC Impedance
Real and imaginary components of Zload for a CRC with a 3 GHz TM010
resonant frequency, a loop area of 1.85×10-6 mm2, and Q0 of 5080.
This frequency dependence of the CRC impedance is immediately
recognized as having the same general form as the frequency dependence of a
lumped element resonant RLC circuit.
In fact, resonant cavities are often
modeled as equivalent RLC circuits, and the computer program used to determine
the cavity parameters from the network analyzer data is based on this type of
modelling. Equation 3.23 can be re-written in the following form:
TM 
2
TM
Q0010

H 010

 N (   R)


TM
  010

 iQ0  TM 
TM
010

 010
2
Zload
2 
Aloop

.



3.24
The expression in the square brackets has units of inductance (due to the
normalization conditions on the magnetic field eigenfunctions), and can be
thought of as the inductance of the equivalent RLC circuit. Using the following
relation from analysis of RLC resonant circuits (Q being the quality factor of the
circuit),
93
R  Qres L,
TM
and the fact that  010
3.25
1 near the resonance, Equation 3.24 can be simplified
as follows:
Z load 
R
TM
 

010
1  iQ0  TM 
 
 010
3.26
.
This is not the exact impedance as seen by the network analyzer, however: there
will be a certain length of transmission line connecting the analyzer to the CRC,
and a shorted transmission line (which is what the coaxial cable terminated in a
loop becomes as soon as you move away from resonance) has a reactance that is
a function of the length of the line as shown in Equation 2.55. This means a series
reactance must be added to the expression for the load impedance to match what
the network analyzer “sees”;
Z load  iX ( ) 
R
TM
 

010
1  iQ0  TM 
 
 010
.
3.27
This expression matches the form of Equation (3) in reference [40], and this paper
(along with [41]) details the process of determining the CRC parameters starting
with Equation (3), using a network analyzer and the equivalent circuit model.
To investigate the effects of coupling loop size on the CRC parameters, the
reactance associated with the transmission line can be neglected and Equation
3.24 can be used. In Figure 3-14, the load impedance is plotted in the complex
94
plane, as the frequency is varied across the resonance, for three different loop
sizes.
Figure 3-14 Effect of Loop Area on Impedance
Zload for a CRC with a 3 GHz TM010 resonant frequency, and Q0 of 5080,
plotted in the complex plane as frequency is varied across the resonance. The
typical circular pattern is shown for three different coupling loop sizes.
As can be seen from the figure, each plot forms a near perfect circle. As
we have neglected the transmission line reactance, the points corresponding to
the resonant frequency fall on the real axis on the right-hand side of each circle.
The strong sensitivity of the impedance to loop size is clear. As a result, the exact
loop size needed for impedance matching can only be approximated beforehand,
and the CRC construction must allow the loop to be rotated while it is connected
to the network analyzer. As the loop is rotated the effective area and therefore
the impedance varies rapidly, and careful tuning is required. Figure 3-14 shows
that for the parameters entered into Equation 3.24, an effective loop area of
1.9×10-6 m2 will result in an impedance of approximately 50 ohms at the resonant
frequency, which represents successful impedance matching.
Again neglecting the effects of the transmission line, the load impedance
can be transformed into the CRC reflection coefficient according to
95
load ( ) 
Zload ( )  50
.
Zload ( )  50
3.28
The reflection coefficient is plotted in the complex plane (as frequency is varied
across the resonance) in Figure 3-15 for the same three loop areas. Again the
plots are near perfect circles, and again the effect of loop area is clear.
Figure 3-15 Effect of Loop Area on Reflection Coefficient
The reflection coefficient, for a CRC with a 3 GHz TM010 resonant frequency,
and Q0 of 5080, plotted in the complex plane as frequency is varied across the
resonance. The typical circular pattern is shown for three different coupling
loop sizes.
There is no new information contained in the reflection coefficient plots;
they are shown here because the computer program, called QZERO, which
determines the CRC parameters, uses the reflection coefficient data. The output
from this program for the YAG-filled CRC is shown in Figure 3-16. The program
plots the reflection coefficient in the complex plane for 201 different frequencies
near the resonance. The output gives five parameters for the CRC, (only four are
independent): QL, Q0, κ, κs, and fL. The loaded resonant frequency, fL, is defined
96
as the frequency corresponding to the smallest magnitude reflection coefficient.
It is technically different from the unloaded resonant frequency, which is defined
as the frequency the cavity would resonate at if the coupling loop were not present
(since the loop has a small inductance of its own, it detunes the cavity slightly).
The difference is so small that it can be neglected, and the value of fL will be
considered the operating resonant frequency of the cavity. The QZERO program
gives a value of 3.0597 GHz for fL.
Figure 3-16 QZERO Output for CRC
Output of the QZERO program for the CRC. This program takes as its input the reflection
coefficient data from a network analyzer, and gives as its output five characterization parameters
for the CRC, most important being the unloaded quality factor Q0.
The coupling coefficient κ is defined as the ratio of power loss in the
“cavity” to the power loss external to the “cavity”. In the circuit model, the
“cavity” is the RLC circuit only and is considered the load impedance, and the
rest of the set-up- the coupling loop, transmission line, and internal network
analyzer impedance, is modelled as an “external” circuit. It is well known that in
order to deliver maximum power to a load from a source, exactly half of the total
power loss must occur at the load, and the other half occurs in the external circuit.
97
It has also been shown that maximum power is delivered to a resonant cavity
when the line is perfectly impedance matched. This means that a coupling
coefficient κ equal unity in the circuit model corresponds to perfect impedance
matching. The QZERO program gives a value of 0.996 for κ. There is a second
coupling coefficient included in the output of the program, κs, which is an attempt
by the program to determine power loss occurring in the coupling loop alone.
The value is masked by the transmission line losses in our case and can be
neglected. Typically, losses in the coupling loop are much less than 1% of total
loss and keeping track of this small loss is only important in the highest precision
measurements, which is not necessary for the CRC design or operation.
The loaded quality factor, QL, related to power loss in the entire circuit, is
defined as QL  Q0 (1   ) . The analyzer actually measures QL and κ directly and
then uses this relation to arrive at Q0 for the cavity. The ability to accurately
measure Q0 is the main advantage of the QZERO program (an acceptable
approximation for fL and κ can be reached simply by viewing the network
analyzer screen without need for additional computer analysis). The program is
also able to approximate the error in its calculation for Q0 (as well as κ), which is
included in the output. QZERO gives a value of 5084.8 ± 187.0 for Q0.
The three parameters necessary for successful operation of the CRC have
now been determined, but further characterization is possible since the CRC is
filled with a dielectric material. The parameter Q0 provides information on the
total power being dissipated in the cavity, which by itself allows for determination
of the appropriate power source. However, it is also desirable to know how much
of that power is being dissipated in the cavity walls versus the dielectric losses.
There are many dielectric materials to choose from, and there are different ways
of constructing the cavity walls, so this information is helpful for design purposes.
Specifically, it is desired to compare loss in the walls for the CRC to loss in the
98
walls for a copper coated dielectric cavity. As mentioned earlier, it is found that
the copper coating process reduces wall loss by 20%.
The calculations are straight-forward. Equation 3.19 can be used to
calculate the total power loss in the cavity (now that Q0 has been determined),
and Equation 2.34 can be used to calculate the power loss in the copper walls
only. The difference is considered the power loss in the dielectric (although there
may be other small sources of loss, such as the coupling loop and indium wire).
The results give a total power loss of 12.37 W for the CRC, with a 10.47 W loss
in the copper walls and a 1.9 W loss in the dielectric.
It should be pointed out that the above calculations used a value for the
effective dielectric constant of the CRC, as calculated by Equation 3.16 using
3.0597 GHz for f and 13.25 mm for R. The value for the effective dielectric
constant is 8.01, significantly different from the approximated value for the
dielectric constant of YAG at 3 GHz, which is 10.4 (this value has never been
directly measured at 3 GHz but is extrapolated from data at other frequencies).
Some discrepancy is to be expected, since the CRC has the vacuum-filled axial
hole as well as a notch to fit the coupling loop, so it is not completely filled with
YAG to begin with. Also, as mentioned earlier, the indium wire and end cap
construction increases the resonant frequency by about 6% compared to a copper
coated cavity construction. This difference is also absorbed into the effective
dielectric constant value despite being unrelated to the characteristics of the YAG
material.
An effective dielectric loss tangent can also be calculated for the CRC
cavity. For a dielectric material, the loss tangent is defined as the ratio     ,
which from Equation 2.28 can be seen to be equal to Pd ,  2We using the actual
value for  r for the dielectric. To get the effective value for the CRC we again
perform the same calculation but using the effective  r . The result is a CRC loss
99
tangent of 3.03×10-5, which would be considered extremely low loss for a pure
dielectric material.
All of the parameters for the CRC which were determined by the network
analyzer data only hold for operation under low power, since the reference signal
of the analyzer is a low power signal. During experiments, the CRC will
experience up to 20 W of RF power, which will cause significant heating. This
heating will of course cause the cavity to expand, changing the dimensions and
therefore also the resonant frequency. From the network analyzer data, it is clear
that even the smallest change in the resonant frequency will have a large effect
on the impedance matching. Since the coupling loop cannot be adjusted during
operation, the impedance matching will continue to worsen as the CRC heats up
and eventually power will be completely reflected by the cavity. To eliminate
this problem, a cooling system was built and installed surrounding the cavity.
Unfortunately, the network analyzer cannot be used to monitor the resonant
frequency during operation in order to test the effectiveness of this cooling
system, since the CRC must be connected instead to the high power source.
However, the high power source does give a read-out of forward and reflected
power from the CRC during operation. With the cooling system on, the power
was slowly increased up to 20 W while the reflected power was monitored. The
result showed that the reflected power varied slightly during operation, meaning
the impedance matching was varying slightly, but the effect was small and so the
CRC was considered sufficiently stabilized. The success of the high power
compression experiment described in the next chapter is also evidence that the
CRC is stable under high power operation.
3.4 Drift Tube and Deflection Plates
The temporal focal distance for the CRC is approximately 30 cm, so there
needs to be this amount of distance separating the CRC (where the pulse velocity
chirp is reversed) and the DRC (where the pulse is focused). A drift tube is
100
installed in the UEM column to provide this distance. In order to help aim the
beam through this extra distance and subsequently down the central axis of the
DRC, a second set of deflection plates are included inside the drift tube. The tube
is also wrapped in a layer of impermeable material, shielding the inside of the
tube from the earth’s magnetic field, which will otherwise deflect the electron
pulses off of the central axis at 25 keV. In Figure 3-17, the deflection plates are
shown attached to a vacuum flange which will be bolted to the end of the drift
tube. The electrical leads which allow the plates to be charges can be seen in the
photo, and vacuum pass-throughs will be necessary so the leads can be connected
to a voltage source.
Figure 3-17 Deflector Plates
Deflector plate assembly specially designed for insertion into the drift tube.
Figure 3-18 shows the drift tube assembly. The tube itself is made of steel
and has vacuum flanges welded to each end. The deflector plate assembly has
been flipped over and inserted into the top of the drift tube. On the right, the
detail of the electrical lead pass-throughs is shown (in these photos the
impermeable cover for the tube is not shown).
101
Figure 3-18 Drift Tube Assembly
Detail of the drift tube assembly with deflector plates and vacuum pass-throughs. On the right, the
alignment of the pass-throughs and the electrical leads for the deflector plates is shown.
3.5 Detection Resonant Cavity (DRC)
The construction and characterization processes for the DRC are basically
identical to the CRC and do not need to be repeated (actually the DRC processes
are simplified by the absence of a dielectric insert). The design parameters are
also similar, with the differences related to the operation which is described in
the following sections. Therefore, the QZERO parameters characterizing the
DRC will simply be given here: fL=5.5408 GHz, κ=0.853, and Q0=9824.8±274.4.
3.5.1 Operation
The DRC operates in a reverse manner compared to the CRC. Operation
of the CRC is based on the coupling loop, carrying a single frequency current
source, driving the cavity field which then acts on the electron pulses. In the
DRC, the pulses themselves act as a current source and drive the resonant mode
(at a much weaker power level) as they travel through the cavity, and this field
creates an EMF around the coupling loop. The loop is connected via vacuum
102
pass-through and coaxial cable to an oscilloscope, where the loop EMF can be
directly observed. The oscilloscope performs a FFT on the time-base data, and
shows the voltage level around any driven resonant frequency of the cavity. This
voltage is directly related to the duration of each pulse in the pulse train passing
through the cavity, as will be shown below.
Equation 2.74, repeated here for convenience, was used previously to
determine the magnetic field inside the cavity when driven by current in the
coupling loop:
H 
i
ii H i

i
2
 2   i

2
 J
s
 Ei*dV ,
3.29
Q0
where Q0 here is related to the power loss in the cavity walls only, since the DRC
is a vacuum-filled cavity. The same equation may be used to determine the field
inside the cavity when it is driven by the pulses passing through it. From the
derivation of Equation 3.29 it is evident that the current density source in this
equation,
J s , is a function of position and frequency. It is therefore the Fourier
transform of the same source represented as a function of position and time. The
source itself is a train of evenly spaced electron pulses traveling down the central
axis of the cavity. As a first step, a mathematical expression for a single traveling
pulse will be developed as a function of position and time. The Fourier transform
of this expression will then be taken, and the result can be inserted into Equation
3.29. Since the transform of a traveling pulse contains a distribution over many
frequencies, Equation 3.29 will need to be integrated over frequency to determine
the total field:
103
H    d
i
ii H i

i
2

2

iQ
2
 J
s
(r ,  )  Ei*dV .
3.30
0
The single pulse current source density is modelled as a continuous charge
density distribution moving with a constant velocity down the z-axis, and the
shape of the distribution can be approximated by a Gaussian function for both the
z and transverse  directions. The shape will be assumed to remain constant
during the relatively short time it is inside the DRC and acting as a current source.
Before an expression for the current density can be written down, the following
parameters are defined:
ne  number of electrons per pulse
qe  electron charge
Vg  gun voltage
me  electron mass
D p  pulse duration parameter
3.31
wp  pulse width parameter
Q p  pulse charge  ne qe
S p  pulse velocity 
2Vg qe
me
(non-relativistic)
Lp  pulse length parameter  S p D p .
In addition, a pulse normalization constant is defined as:
Np 
Qp S p
 z2
2 
2   exp   2  2  d  dz
wp 
 Lp
 0
 
104
.
3.32
Any waveform of constant shape traveling in the positive z-direction will
be a function of ( z  vt ) , where v is the velocity of the waveform. Thus, using
the above defined parameters, a Gaussian shaped current density pulse traveling
down the positive z-axis can be expressed in the following form:
  z  S t 2  2 
 

p
JPulse (r , t )  zˆ N p exp   
    .

  Lp   wp  

 
3.33
The constants in front of the exponential (specifically the normalization constant)
have been chosen so that the integral of the above expression, over all space at a
particular time, will be equal to Q p S p . This is somewhat arbitrary, but this
quantity will remain constant if the pulse is compressed (or stretched) while the
total charge remains constant. Since this is precisely what will happen to the
pulse in the CRC, this choice of constants is appropriate.
The Fourier transform is now evaluated to obtain a function of space and
frequency:
  z  S t 2
   2  

1


p
J Pulse (r ,  )  zˆ
N p exp       exp   
 it dt. 3.34

  Lp 

  wp   
2



 
The factor of 1 2 is included to be consistent with the definition of a phasor. To
compute the time integral, the expression inside the curly brackets must be
modified in order to complete the square, giving
105
J Pulse ( r ,  ) 
zˆ
 
N exp  
2
 w
1
2
p
p
2

z
2
Lp
2

Sp
2
Lp
2
 S  iL   2 S z  
3.35
 iL   2 S z  
exp

    t 
  dt.
2S
 2S
 
 
 L 
2
2
p

p
2
p
2
2
2
p
p
2

p
2
p
p
The integral is now a standard Gaussian integral and is equal to  Lp S p . After
some simplification, the current density transform can be written in the form:
 Dp 2 2 i z  2 
1
J Pulse ( r ,  )  zˆ
 Dp N p exp  

 2 .
4
S
wp 
2
p

3.36
In order to match the equations in Appendix I used for calculations and plotting,
this expression will be re-written in term of the ordinary frequency f rather than
angular frequency ω (all expressions for the remainder of this section will also be
expressed in terms of ordinary frequency):
 Dp 2 (2 f )2 i 2 fz  2 
1
J Pulse ( r , f )  zˆ
 Dp N p exp  

 2 .
4
S
wp 
2
p

3.37
At this point, Equation 3.30, which is the general expression for the
magnetic field in a RCCC stimulated by a current source, can be developed
specifically to represent the TM010 mode field in the vacuum-filled DRC
stimulated by a Gaussian current density:
H (  , z )   df
010

i 2 f 010 H 010 (  )
TM
TM
TM

N
2


f
2
2
2
4  ( f )  f   i

Q



0 
TM
010
TM
010
106

 J Pulse (  , z, f )  E (  )
TM
010

*
N
dV ,
3.38
where the explicit spatial coordinate dependences of the various functions (for
this particular mode) have been included (there is no  dependence). In the above
equation the subscript “N” again refers to the normalized eigenmode field
patterns defined earlier. Their expressions are repeated here for convenience
along with the expression for the eigenfrequency:
H 
TM
010
N
 ˆi
 0 TM
2.405  
E010  J1 


0
0
 R 
E 
TM
 zˆ  E010
J 0 

010
0
N

2.405
TM
f 010 
,
2  0 R
TM
2.405 
R



3.39
where R is the radius of the cavity. The normalization phasor on the right side of
the first two equations is derived from the normalization Conditions 2.67.
Solving for this phasor for the TM010 mode gives:
E 
TM
010
0

d R
 0 2  J 0 2
0 0

1
2.405 
R
 d dz
,
3.40
where again R is the cavity radius and d is the cavity length.
Equation 3.38 allows calculation of the TM010 mode magnetic field
everywhere inside the cavity volume, when the cavity is stimulated by the
electron pulse passing through it. Of particular interest is the magnetic field
inside the cavity at the location of the loop antenna, which will allow calculation
of the induced EMF in the loop that can then be directly measured by an
oscilloscope. The loop is considered to be small enough, and the field constant
enough in the loop region, that the loop location can be treated as a single point
107
on the outer edge of the cavity. The relevant coordinates for the loop, according
to the previously defined cavity coordinate system, are then   R and z  d 2 .
These coordinates can be inserted into Equation 3.38 to give a value for
TM
H 010
( R, d 2) , which can then be used to calculate the magnetic flux through the
loop. With the assumption that the field is constant over the small area of the
loop, the flux is:
TM
TM
010 (t )  0 Aloop  H010 ( R, d 2, t ).
3.41
Although there are a small range of frequencies around the resonant frequency
that contribute to the magnetic field, to simplify the analysis the field will be
assumed to oscillate at the resonant frequency only. According to the previously
defined complex notation, the above expression is then equal to (both vectors
point in the ̂ direction):
TM
TM
TM
(t )  0 Aloop 2 Re  H 010
( R, d 2, t ) exp i(2 f 010 t   2)   . 3.42
010

Faraday’s Law then gives the induced EMF from the TM010 mode:
TM
EMF010
(t )  
d TM
 010 (t )
dt
TM
TM
 2 f 010TM 0 Aloop 2  Re H 010
( R, d 2) exp i 2 f 010 t  

3.43
TM
TM
 2 f 010TM 0 Aloop 2 H 010
( R, d 2) cos(2 f 010 t ).
Finally, the amplitude of the EMF is the quantity that is of experimental interest:
108
TM
TM
TM
EMF010
(t )  2 f 010
0 Aloop 2 H 010
( R, d 2) .
3.44
At this point, it should be discussed exactly how the above quantity relates to the
actual time-domain signal that will be observed during experiment. By including
only the resonant frequency in Equation 3.42, the DRC response is represented
by a pure sinusoid, which of course it will not be. When a single electron pulse
passes through the cavity it acts like an impulse source on a resonant system, and
the response signal includes a short period of amplitude build up while the pulse
is inside the cavity, followed by oscillation at the resonant frequency that
gradually decreases in amplitude at a rate inversely proportional to the total
quality factor of the system. The part of the response that is of interest is the
amplitude of the oscillation at the resonant frequency that occurs just after the
pulse leaves the cavity. Because of the large Q0 of the system, this amplitude
(which corresponds to the EMF quantity in Equation 3.44) will remain relatively
constant for many cycles and the signal can be approximated as a sinusoid during
this time. This EMF corresponds to the stimulated magnetic field inside the DRC
that is represented by the magnetic field (phasor) in Equation 3.38. This equation
was derived from an analysis that assumed time-dependent current sources were
pure sinusoids, and therefore produced a pure sinusoidal response. That original
analysis has basically been modified to include non-sinusoidal sources, but then
“pick out” the sinusoidal part of the response only.
The EMF quantity in Equation 3.44 ends up being dependent on the time
duration of the electron pulse passing through the DRC, which is why it is the
quantity of interest. This fact allows the DRC to “detect” the pulse duration in
situ without disturbing the pulse in any significant way (there will of course be a
tiny reaction force on the pulse from the field it generates inside the DRC, but
this effect is vanishingly small). For this reason, it is a useful tool for experiments
involving pulse compression. While the DRC can theoretically detect any change
109
in pulse duration, there are two main conditions that must be met for practical
use. First, the EMF voltage signal, or DRC signal, must be large enough to be
detectable. Second, this signal must vary rapidly with pulse duration so the effect
is noticeable above any random variation. To understand how these conditions
can be met, the equations of this section must be examined in greater detail.
When designing the DRC, the first choice that needs to be made is the mode
of operation which will create changing flux through the loop antenna, and thus
a large, detectable EMF. Due to the location and orientation of the loop antenna
in the above analysis, Equation 3.41 requires the magnetic field for the chosen
mode to have a significant relative amplitude at the cavity coordinates   R ,
z  d 2 and be pointing in the ̂ -direction. A quick look at the TM010 mode
pattern shows it is a suitable choice. As Equation 3.39 shows, the magnetic field
for the TM010 mode points only in the ̂ -direction and has no z-dependence (so
there is no zero in the field to worry about at z  d 2 as with many other modes).
A plot of the  -dependence of the field at z  d 2 is shown in Figure 3-19. Here,
it is important to remember that the normalization conditions are those given in
Equation 2.67, which give the normalized field its strange looking units.
110
Figure 3-19 Normalized TM010 H-Field
Radial dependence of the normalized TM010 H-field in a cylindrical cavity
with radius 20.61 mm.
With the TM010 mode chosen as a suitable mode for operation, the
relationship between the loop EMF and the pulse duration can be further
understood by examining the connection between the magnetic field (which
directly generates the EMF) and the pulse duration. This relation is given by
Equation 3.38, which can be separated into several “components” to be analyzed
individually. Specifically for the TM010 mode, the main components (with units)
are defined as follows:
Mode Projection   J Pulse (  , z , f )   E (  )  dV
TM
010
Transfer Function 
*
N
i 2 f 010
TM

f2 
2
2
2
4  ( f )  f   i

Q




0 
 N m


 s
TM
010
TM
010

Normalized H -Field  H (  )
TM
010

N
Equation 3.38 can now be rewritten as:
111
 A 
.

3 
N

m


3.45
H (  , z )  [Normalized H -Field]   df [Transfer Function]  [Mode Projection].
TM
010
3.46
The mode projection can be represented as the volume integral of two additional
sub-components;
Mode Projection   [Pulse Transform]  [Normalized E -Field] *dVcavity ,
3.47
where the sub-components are defined as
Pulse Transform  J Pulse (  , z, f )

TM
Normalized E -Field  E010
( )

N
 A 
 m2  s 
 1

 A s
N
.
m
3.48
Here again, the normalization conditions are responsible for the units of the
normalized E-field. The mode projection can be thought of as the “overlap” of
the pulse transform (at a particular frequency) and the electric field of the
operating mode (in this case the TM010 mode), integrated over the volume of the
cavity. According to Equation 3.46, the mode projection is then multiplied at
each frequency according to the transfer function, and the integral over all
frequencies gives a total amplitude which then multiplies the normalized H-field
pattern for the operating mode. This finally gives the value for the magnetic field
everywhere in the cavity (in standard SI units), stimulated by the electron pulse
passing through the cavity.
As stated previously, in order to be of practical use, the DRC must produce
a signal large enough to be detected, and this signal must vary rapidly with pulse
duration. The pulse transform is the only component of Equations 3.46 and 3.47
112
which affects the variation with pulse duration. All other components affect the
strength of the signal. Therefore, to understand the variation with pulse duration,
the pulse transform (Equation 3.36) will need to be examined in detail. In Figure
3-20, the pulse transform amplitudes for various pulse durations are plotted as a
function of frequency. A single pulse is assumed to have 175,360 electrons
(matching one of the values that will be used in experiments), and a velocity of
0.313c (corresponding to a gun voltage of 25 kV).
Figure 3-20 Pulse Transform Variation
Variation in a single pulse transform for different pulse durations. A single
pulse contains 175,360 electrons and has a velocity of 0.313c. The transform
functions were evaluated at cavity coordinates ρ=0 and z=0.
The entire operating principle of the DRC can be derived from this one
plot, along with its limitations for operation. As will be shown, because of the
transfer function (related to the resonating nature of the cavity), the DRC only
“sees” signals that are near its resonant frequency. For example, the DRC used
in the experiments described in this document has a resonant frequency of 5.54
GHz. From the plot it is easily seen that at this frequency, the amplitude of a 20
picosecond pulse transform will be approximately seven times greater than the
amplitude for a 160 picosecond pulse transform. This ratio will (almost) directly
translate to a change in the detected EMF voltage signal. So, if pulses passing
through the DRC are originally 160 picoseconds and are then compressed by the
113
CRC to 20 picoseconds, the magnitude of the DRC signal seen on the
oscilloscope will increase by (approximately) a factor of seven, a ratio that is
easily detectable.
Figure 3-20 also shows, however, that at 2 GHz the same signal ratio would
only be 1.2, which may not be detectable above the noise involved in the
experiment. A 2 GHz DRC would be more useful for working with longer pulse
durations. It can also be seen from the plot that at 25 GHz, a DRC would be most
effective at detecting changes in pulses in the 10-40 ps range, but would be
useless for larger pulses. So for a DRC at any fixed resonant frequency, there
will be an optimal range of pulse durations to work with, and at higher resonant
frequencies this range will shift to shorter pulses.
Before continuing, it is necessary to investigate whether Parseval’s
theorem holds for the pulse current density functions and their transforms as
they’ve been defined. The theorem is as follows:


2
f (t ) dt 



2
F ( f ) df

where
F( f ) 
3.49

 f (t )exp[i2 ft ]dt.
Parseval’s theorem roughly states that the square of a function of time, integrated
over time, must equal the square of its frequency transform integrated over
frequency. The pulse current density function, as it’s defined, is a function of
space and time, and the pulse transform is a function of space and frequency.
Thus, to check Parseval’s theorem, the current density function (for any pulse
duration) and its transform must be evaluated at the same point in space and then
114
the integrations over time and frequency carried out. Parseval’s theorem must
then hold at all points in space. The pulse transform functions in Figure 3-20
were evaluated at coordinates   0 and z  0 . The corresponding pulse charge
density functions, evaluated at the same point, are shown for comparison in
Figure 3-21. In Appendix I, the integrations for Parseval’s theorem are carried
out at several random points, for various pulse durations, and the theorem is
shown to hold. This is strong evidence that the form of the pulse transform is
valid.
Figure 3-21 Single Pulse (Time-Domain)
Time-domain functions for the pulse durations corresponding to the
transforms from the previous figure.
It has just been shown that the pulse transform contains within it all the
information about the pulse duration. To be able to detect this information, it
needs to be translated to a signal significantly large enough to be seen, for
example, on an oscilloscope. Maximizing the mode projection created by the
transform must therefore be the first step in creating this signal. Along with the
pulse transform, the normalized TM010 E-field constitutes the other component of
the mode projection according to Equation 3.47. This equation shows that the
dot product of the transform and E-field vectors is integrated over the volume of
the cavity to form the mode projection. Both of the vectors have only z115
components, as can be seen on their respective definitions, so the dot product is
simply the product of the (complex) amplitudes of the vectors. The transform has
a real and imaginary part, while the normalized E-field, as it’s been defined, is
purely real. So the mode projection depends on the overlap of the z-component
amplitudes of the two vectors integrated over the volume of the cavity. This
integral can be separated into two parts, the integration over the radial direction (
 ) and the integral over the axial direction (z). First, integration over the radial
direction will be analyzed. Figure 3-22 shows the radial dependence of the two
vector components that will be multiplied together.
Figure 3-22 Radial Dependence of Vectors in the Mode Projection Integral
Radial dependence and overlap of the pulse transform function and the z-component of the Efield. The z-coordinate of the plots was fixed at z=16 mm to show non-zero real and imaginary
parts of the transform. The transform is of a 40 ps pulse. A single pulse contains 175,360
electrons and has a velocity of 0.313c.
For these plots, the z-coordinate was fixed at a specific value ( z  16mm )
to show the real and imaginary parts of the transform. The gun voltage was set
to 25 kV. As can be easily seen, the transform is only non-zero near the cavity
axis, so for overlap to occur the E-field must also be non-zero near the axis. The
TM010 mode has its maximum value for the E-field on the axis, so good overlap
is ensured. This is another reason the TM010 mode is a proper choice for DRC
operation.
116
Figure 3-23 Axial Dependence of Vectors in the Mode Projection Integral
Axial dependence and overlap of the pulse transform function and the z-component of the Efield. The ρ-coordinate of the plots was fixed at ρ=0 where the amplitudes are maximum. The
transform is of a 40 ps pulse. A single pulse contains 175,360 electrons and has a velocity of
0.313c.
Next, integration over the axial direction will be analyzed. Figure 3-23
shows the axial dependence of the two vector components that will be multiplied
together. For these plots the  -coordinate was fixed at   0 , where the vector
amplitudes are at a maximum. Since the value of Ez remains constant while the
transform goes from positive to negative values sinusoidally, it is possible that
the integral over this direction could have maxima and zeroes depending on the
chosen cavity length, given a particular pulse velocity. This is in fact the case, as
can be seen in Figure 3-24. Here, the absolute value for the entire volume integral
is shown as a function of total cavity length. It is important to note that pulse
transform will vary with the velocity of the pulse, which is determined by the gun
voltage value, and for the above plots the gun voltage was fixed at 25 kV. At
fixed velocity, various cavity lengths will correspond to various amounts of time
the pulse spends inside the cavity, and this time will correspond to a certain
number of periods of the oscillating field. In Figure 3-24, the orange dashed lines
correspond to cavity lengths where, for a 25 kV pulse and a 5.54 GHz field, the
pulse is inside the cavity for an even number of field periods. The gray dashed
lines correspond to odd half-periods.
117
Figure 3-24 Mode Projection and Cavity Length
Variation in the mode projection amplitude with cavity length, for a fixed
pulse velocity of 0.313c. Each pulse contains 175,360 electrons and has a 40
ps duration.
It is clear that maxima in the mode projection closely correspond to even
periods, while zeroes in the mode projection correspond to odd half-periods. This
suggests that for a fixed cavity length, variations in the gun voltage will also affect
the mode projection, since a change in pulse velocity will change the transit time
through the cavity. This turns out to be the case, as shown in Figure 3-25. In this
plot, the cavity length was fixed at 24 mm, and the orange lines correspond again
to even periods while the gray lines correspond to odd half-periods. The same
relationship is seen between the mode projection and the number of RF oscillation
periods the pulse spends inside the cavity. The above analysis is important since
it identifies two design parameters to maximize the mode projection, namely the
cavity length and the acceleration voltage (i.e., pulse velocity).
118
Figure 3-25 Mode Projection Integral and Gun Voltage
Variation in the mode projection amplitude with gun voltage (pulse velocity),
for a fixed cavity length of 24 mm. Each pulse contains 175,360 electrons and
has a 40 ps duration.
Next, it will be shown how the transfer function “picks out” a narrow range
of the mode projection integral. Figure 3-26 shows the how the mode projection
and transfer function vary with frequency. Here, we have shown both the real
and imaginary parts of both functions, since they will be multiplied together
before integrating over frequency and cross-terms will appear. The situation is
simplified by the fact that over the narrow frequency range where the transfer
function is non-zero (about the width of the dashed line in the mode projection
plot), the values for the real and imaginary parts of the mode projection are
effectively constant.
119
Figure 3-26 Mode Projection and Transfer Function
Variation of the mode projection (real and imaginary parts) and the transfer
function with frequency. Each pulse contains 175,360 electrons, has a
velocity of 0.313c, and a duration of 40 ps. The cavity length was fixed at 24
mm. An unloaded quality factor of 9824 was assumed.
The result of the multiplication gives a function (the integrand in Equation
3.46) that resembles a scaled transfer function, still narrowly centered around the
resonant frequency as shown in Figure 3-27. This function is then integrated over
frequency and multiplied by the normalized H-field pattern for the TM010 mode,
in accordance with Equation 3.46, to give the expression for the magnetic field
everywhere in the cavity.
120
Figure 3-27 Mode Projection Multiplied by the Transfer Function
Mode projection and transfer function multiplied together with all parameters
retaining the same values.
The EMF in the loop, which will be directly measured by an oscilloscope
(minus losses) can then be calculated as explained earlier. The theoretical
variation of this EMF with pulse duration is shown in Figure 3-28. This largest
slope in this plot is in the 40-120 ps pulse range. This means, for a 5.54 GHz
cavity, it is easiest to experimentally detect changes to the duration of pulses
within this range. The same conclusion was reached by analyzing the pulse
transform plot (Figure 3-20) at 5.54 GHz, showing the variation in the transform
is the main operating principle of the DRC.
121
Figure 3-28 Pulse Duration Dependence of EMF
Variation in coupling loop EMF with pulse duration. Each pulse contains
175,360 electrons and has a velocity of 0.313c, and the cavity length is 24
mm.
3.5.2 Double Pulse Operation and DRC Calibration
It has been shown that at 5.54 GHz, a DRC is best suited for detecting
duration change in pulses in the 40-120 ps range. Unfortunately, the laser system
used in the UEM lab produces electron pulses that are approximately 4 ps (or
shorter) in duration, too short to test the DRC. In an attempt to simulate longer
pulses, a Michelson interferometer set-up was introduced into the harmonic
generation scheme employed to produce the UV laser pulses incident on the
photocathode. An optical delay stage in one arm of the interferometer allows for
the controlled generation of two time separated pulses. In this way, a 40 ps pulse
can be simulated by two 4 ps pulses, their centers 40 ps apart. This “double
electron pulse” can be sent through the CRC, where electromagnetic forces will
change the distance between the pulse centers, either “compressing” or
“expanding” the double pulse system. Figure 3-29 shows several double pulse
systems used for simulating longer pulses. To determine if this type of simulation
will actually work for testing the DRC, the Fourier transform for the double pulse
system will need to be examined. It was established in the last section that the
122
behavior of the transform is the main operating principle of the DRC, but the
transform for the double pulse system will be significantly different than that for
a single pulse. Consequently, it is not at all obvious that this simulation should
work.
Figure 3-29 Double Pulse Systems (Time-Domain)
Time-domain functions for various double pulse durations. A double pulse
contains 175,360 electrons total, and has a velocity of 0.313c. Each individual
pulse has a 4 ps duration.
A lengthy mathematical calculation was performed in the previous section
to derive the transform for a single traveling pulse. Luckily, this analysis will not
need to be repeated, and a few simple properties of Fourier transforms can be
utilized to achieve the desired result. The current density function for a single
traveling pulse has been denoted by J (r , t ) , and its transform, calculated in the
previous section, has been denoted by Ff [ J (r , t )]  J (r , f ) . The current density
function for an identical double pulse system can be represented by the following
expression: J (r , t  T 2)  J (r , t  T 2) , where T is the separation time between
123
the centers of the pulses. The transform can then be readily evaluated using the
shift properties theorem of Fourier transforms:
Ff [ J (r , t  T 2)  J (r , t  T 2)]
 Ff [ J (r , t  T 2)]  Ff [ J (r , t  T 2)]
 exp[i 2 f T 2] Ff [ J (r , t )]  exp[ i 2 f T 2] Ff [ J (r , t )] 3.50
 Ff [ J (r , t )]  exp[i 2 f T 2]  exp[i 2 f T 2] 
 J (r , f )  2cos(2 f T 2)  .
The end result is the original single pulse transform (for a short 4 ps pulse),
modulated by a cosine function that depends on the double pulse separation time
T. It is this modulation in the transform that allows the double pulse simulation
to work; as the parameter T increases and decreases, simulating pulse expansion
and compression, the cosine function will vary quickly enough to allow detection
by the DRC. Figure 3-30 shows the transforms for several pulse separation times.
It is evident that at 5.54 GHz, there is an easily detectable difference in the
transforms for separation times in the 20 ps to 100 ps range.
124
Figure 3-30 Double Pulse Transforms
Variation in a double pulse transform for different pulse durations. A double
pulse contains 175,360 electrons total and has a velocity of 0.313c. The
transform functions were evaluated at cavity coordinates ρ=0 and z=0.
The rest of the analysis from the previous section still holds for the double
pulses. For example, in Figure 3-31 the axial dependence of a 10 ps separation
double pulse is shown.
The double pulse transform has the same axial
dependence as before. This means that the optimal cavity length and gun voltage
values will remain the same as for the single pulse case.
Figure 3-31 Axial Dependence of DP Transform
Axial dependence of the double pulse transform function. The ρ-coordinate
was fixed at ρ=0 where the amplitude is a maximum. The transform is of a
40 ps double pulse. A double pulse contains 175,360 electrons total and has
a velocity of 0.313c.
125
In Figure 3-32, the theoretical variation of the EMF with separation
distance is shown at 5.54 GHz, and it is evident that this variation closely follows
the transform variation. Again, the transform is the main operating principle for
the DRC. Due to the cosine function modulation in the DP transform, the EMF
response of the cavity will now have a zero near a separation distance of 90 ps.
This characteristic will actually make it much easier to test the DRC than is the
case for a single pulse.
Figure 3-32 DP Separation Dependence of EMF
Variation in coupling loop EMF with double pulse duration. Each double
pulse contains 175,360 electrons total and has a velocity of 0.313c, and the
cavity length is 24 mm.
Additionally, due to the fact that the interferometer can set the pulse
separation distance with a high level of accuracy, an experiment that records the
EMF response at various known separation distances can be used to calibrate the
DRC for double pulse and even single pulse compression experiments. A
calibration and a compression experiment for the double pulse set-up were
completed and the results discussed in the following chapter.
126
3.6 Synchronization and Power
When an electron pulse enters the CRC, it is critical that the EM field inside
the CRC is at the proper phase for compression to occur. While the system is
running, an optical delay stage is used to adjust the pulse arrival time relative to
the EM field phase, and the DRC can then theoretically detect when maximum
compression is being achieved. In practice, it is easier for the delay stage to be
adjusted until the DRC is showing maximum expansion of the pulse (since the
DRC is more sensitive to changes in larger pulses). Once the position for
maximum expansion has been found, the delay stage can be moved a distance
equivalent to a 180 degree shift in the phase relationship between pulse arrival
and EM field.
Once the optical delay stage is in the proper position for maximum
compression of the pulse, the concern becomes the stability of the relationship
between pulse arrival and CRC EM field phase. In some similar systems using a
laser-driven electron gun and an RF compression cavity, a separate oscillator is
used to drive the RF cavity. In these types of set-ups, to avoid relative drift
between the oscillator driving the RF cavity and the laser oscillator, a phase
locked loop (PLL) system is used to continually adjust the frequency of one of
the oscillators to match the phase of the other [42]. A main advantage of the UIC
UEM system is the fact that the laser oscillator not only drives the electron gun,
but it also drives the RF compression cavity (CRC). Therefore, there is no
concern for long-term drift between the separate oscillators and no need for a PLL
to control relative drift. There remains an issue of overall drift of the laser
operating frequency due to heating of the laser cavity. While this was not a
problem for successful completion of the experiments discussed in this document,
in the future a feedback system could be employed to continuously adjust the
laser cavity length (and therefore the operating frequency) to account for the
heating effects.
127
Even though long-term relative drift between the EM field phase and the
pulse arrival at the CRC is not an issue, certain factors remain which cause
instability to be present in the system, creating variations in the EM phase
relationship from one pulse to the next. This variation causes several problems.
First, some pulses may be compressed more than others, creating uncertainty in
the pulse duration when it arrives at the specimen. Also, a net acceleration or
deceleration can be applied to a pulse by the CRC when the phase is not perfectly
matched. This will cause a variation in the location of the temporal focal point
from pulse to pulse, along with a variation in arrival time of the pulse at the
specimen. These types of variation mentioned are clearly detrimental to the
imaging synchronization capability of the system.
The main sources of instability in the UIC UEM system will now be
described. The first source is the phase noise inherent in the laser oscillator. All
oscillators exhibit some level of phase noise, which is closely related to the
variation in an oscillator’s frequency around its nominal value. A variation in an
oscillator’s frequency, of course, can be related to a variation in the time period
of the oscillator. This is known as time jitter, which can be calculated from the
phase noise of the oscillator and is defined as the standard deviation of the
oscillator’s period. Since we are interested in the timing mismatch between the
CRC EM field phase and the arrival of a pulse at the CRC, time jitter is the
quantity that is of interest. A complete time jitter analysis has not been performed
on the UIC UEM laser, but similar lasers have shown a level of time jitter less
than 0.5 ps [43].
The second source of instability in the system, which leads to additional
time jitter, is the slight fluctuation over time of the acceleration voltage in the
photoelectron gun. The UIC system uses a Spellman High Voltage SL series
power supply which has a voltage stability rating of 100 ppm/hour. Since the
velocity of the pulse is related to the square root of the voltage and the variation
128
can be considered small, the variation in velocity will be approximately half the
variation in voltage. This equates to a range in travel time from the gun to the
CRC of about 0.6 ps.
The third source of time jitter is from the 30 Watt RF Amplifier. As
mentioned previously, a small amount of energy from the main laser oscillator,
operating at 28.5 MHz, is directed towards a fast photodiode. The photodiode
signal contains all of the harmonics of the laser cavity frequency up to 10 GHz.
The 107th harmonic (3.055 GHz) is selected by a narrow band-pass filter, and
undergoes a pre-amplification of 20 dB in order to reach the minimum input
power of the 30 Watt amplifier. The UIC system uses an Ophir RF Model 5182
amplifier. The rms time jitter rating for this amplifier is 0.5 ps at 3 GHz and 24
Watts.
Adding the time jitter from the three main sources of instability in the
system gives a total timing mismatch of approximately 1.6 ps. It should be
pointed out that for the experiments completed and discussed in this document,
this level of time jitter is insignificant due to averaging over many pulses.
Minimizing the effects of time jitter is more critical for imaging purposes, when
a single pulse is taking an image and there is no averaging process. A detailed
analysis of the effects of time jitter on imaging are beyond the scope of this
document.
129
4 Experimental Results
The results of six experiments related to the theory presented in this document
will now be examined. The first three experiments were conducted to investigate
the effects of the axial hole and the end cap holes on the characteristics of resonant
cavities. The final three experiments involve a double pulse calibration of the
DRC and a pulse compression experiment simulated by a double pulse set-up.
The theoretical linear relationship between electrons per pulse and DRC signal
strength was also verified by experiment.
4.1 Dielectric Axial Hole Effects
Throughout this document, when discussing the theory of resonant cavities
filled with a dielectric material, it has been suggested that the axial hole drilled
through the dielectric does not significantly change the characteristics of the
cavity. For this reason, the theoretical eigenmodes for a fully filled cavity with
no hole have been used in calculations concerning resonant frequencies, power
loss, etc. To support this critical assumption, an experiment was completed where
the axial hole radius of a dielectric-filled cavity was varied, and the resulting shift
in the resonant frequency was measured with a network analyzer. The goal of the
experiment was to replicate the theoretical curve shown in Figure 3-2. A
successful result would be a strong indication that the theory presented in Section
3.3.1 concerning the axial hole was indeed correct, and this theory then leads to
the curves of Figure 3-3, which show the cavity EM field to be mostly unaffected
by the axial hole. The results of the experiment are shown in Figure 4-1. To
perform this experiment, a vacuum-filled aluminum cavity with end caps (using
indium wire) was constructed. The cavity was designed to have a resonant
frequency near 4 GHz, which corresponds to a radius of 28.71 mm. An actual
YAG dielectric insert could not be used for this experiment, due to the high cost
of purchasing an extra YAG insert along with the difficulty in drilling through
130
the YAG material. Instead, a much more cost effective and machining-friendly
material called Rexolite was used. By comparing the resonant frequency of a
filled cavity with an unfilled cavity, the Rexolite was found to have an effective
dielectric constant of 2.8 at 2.5 GHz, and the Mathematica program using this
parameter generated the new theoretical curve plotted in the figure. Starting with
no hole at all, various hole diameters were drilled through the Rexolite, which
was then inserted into the cavity housing secured with end caps and indium wire,
and finally hooked up to the network analyzer to measure the resonant frequency.
Each hole size tested required new indium wire and reattachment of the end caps,
which by itself will cause a slight variation in the resonant frequency. The
experiment was repeated 10 times at one particular hole size to provide an
estimate of the variation in resonant frequency caused by using new indium wire
and reattaching the end caps. This variation was found to be approximately ±
0.1%, too small to be visible at the scale of the plot in Figure 4-1. As can be seen
from the figure, the experimental data closely matched the theory and clearly
supports the assumption that eigenmodes for a completely filled cavity can be
used in calculations.
131
Figure 4-1 Variation of Resonant Frequency with Axial Hole Size
Observed network analyzer data for resonant frequency compared to a theoretical
curve calculated for the dielectric material Rexolite. The resonant frequency is
normalized to that of an unfilled cavity, which is 3.99 GHz. The hole radius is
normalized to the cavity radius, which is 28.71 mm. Uncertainty in the data is
caused by the need for new indium wire and reattachment of the cavity end caps
before each measurement, but this uncertainty is too small to be seen on this scale.
4.2 End Cap Hole Effects
Another possible problem with using eigenmodes for calculations is the fact
that the eigenmode boundary conditions are violated by the holes in the cavity
end caps. Unlike the axial hole in the dielectric, there is no exact theoretical
solution for a cavity with holes in the conducting boundary to which experimental
data can be compared. The end cap holes are assumed to be a small perturbation,
which affect the cavity EM field in the near vicinity of the holes but have only a
small effect on the overall field pattern throughout the cavity. To test this, the
(vacuum-filled) aluminum cavity from the previous experiment was again used,
and various hole sizes were drilled into the end caps with the resonant frequency
being determined by a network analyzer for each hole size. The results of the
132
experiment are shown in Figure 4-2. The results show that there is no significant
change in resonant frequency for hole radii less than 20% of the total cavity radius
(the end cap hole radius for the CRC is 15% of the total cavity radius and for the
DRC it is 10%). Since the resonant frequency of the cavities are not significantly
changed, this supports the assumption that the end cap holes affect the field
pattern only near the holes and do not have a significant effect on the overall
eigenmode field pattern.
Figure 4-2 Variation of Resonant Frequency with End Cap Hole Size
Variation in the cavity resonant frequency when the radius of the end cap
holes are varied. The radius values are normalized to the total cavity radius,
28.71 mm. Uncertainty in the data is caused by the need for new indium wire
and reattachment of the cavity end caps before each measurement.
This, however, does not mean that the end cap holes do not have an effect
on an electron pulse passing through the cavity. The holes distort the EM field
in their near vicinity, which is the region an electron pulse will pass through.
Theoretical modelling [21] indicates that the distortion results in a diverging
transverse electric field near the cavity entrance and exit that will affect the
divergence of the pulse in the transverse direction. This change in transverse
133
pulse size will need to be compensated for by the magnetic lenses. To test the
effect of the distorted fields, and the ability of the magnetic lenses to provide the
appropriate correction, an experiment was conducted and the results shown in
Figure 4-3.
Figure 4-3 Effects of End Cap Hole Diverging Field
Variation in the required magnetic lens focusing current, when an electron
pulse enters the CRC at various phases (strengths) of the end cap diverging
field. Uncertainty in the data is caused by the difficulty in exactly
determining when the pulsed beam is focused.
This experiment was performed using the CRC alone, operating at a
relatively low power level of 5 W. The methodology of the experiment is based
on the fact that the diverging field near the cap holes will have a sinusoidal time
dependence just like the eigenmode field. Using a delay stage to control the phase
of the cavity EM field when a pulse enters, the effect of the diverging field on the
pulse will vary from a minimum to a maximum and back. At any particular phase,
the magnetic lenses will need a specific adjustment to their operating current in
order to refocus the beam of pulses. At phases where the diverging field is strong,
the magnetic lenses will require maximum additional current. The AG model
134
simulations can predict this refocusing current at each phase, and a theoretical
curve based on the AG model is shown in the figure.
To perform the experiment, the beam was first focused to the smallest
observable spot size on a YAG scintillation screen placed at the position of the
Faraday cup (Figure 3-1), at one particular delay stage position, by adjusting the
current in the magnetic lenses. The delay stage was then moved a distance
equivalent to a particular change in EM field phase, and the beam refocused with
a new magnetic lens current level. As can be seen in the figure, the experimental
results agree well with the curve predicted by the AG model. The error bars of ±
2.5% are due to the obvious uncertainty involved in determining exactly when
the spot size was a minimum. The results show that the effect of the end cap
holes can be predicted and controlled, and therefore they do not pose an issue for
successful compression experiments. It should also be pointed out that this was
the first experimental verification of the predictions of the AG model developed
by Berger.
4.3 DRC Signal
The following experiments were conducted to test the validity of the DRC
operation theory described in Section 3.5.
4.3.1 Variation of Double Pulse Separation (Calibration)
The purpose of this experiment is to test the validity of the theory presented
in Section 3.5.2 regarding the double pulse operation of the DRC. Verifying this
theory is important since it will also verify the theory for a single pulse, and it
allows calibration of the DRC for compression. In this experiment, the time
separation of the two pulses in each double pulse was varied from 10 ps to 130
ps in increments of 10 ps. A low-loss 24 inch coaxial cable connected the DRC
vacuum pass-through to an Agilent model N5222A microwave amplifier. The
output of the amplifier was connected via another identical 24 inch cable to a
135
Tektronix model TDS6154C digital oscilloscope, where the voltage of the DRC
loop EMF signal could be observed.
To run the experiment, the laser amplifier was used at a pulse repetition
rate of 570 Hz. At this repetition rate, after each pulse has passed through the
DRC there will be enough time for the ringing signal to die off completely before
another pulse enters. Therefore, the experiment can be considered a single-shot
experiment and the ringing signal should match the theory (which was modeled
by a single pulse passing through the cavity). A Faraday cup was used to measure
the total current for each pulse train that passed through the DRC. From this the
average total number of electrons in each pulse can be calculated, which is needed
as an input for the theoretical model. The gun voltage was set to 25 kV. A
network analyzer was used to determine the exact amplification magnitude of the
microwave amplifier at the resonant frequency of the DRC, and the result was a
voltage signal increase by a factor of 49.93 (33.97 dB power increase). The
oscilloscope sampled the amplified voltage signal every 12.5 ps, which
corresponds to approximately 14 data points per period of oscillation for a 5.54
GHz signal and 4000 total data points for the duration of the signal sampled.
Because of variation in the current being recorded by the Faraday cup, averaging
was employed over 5000 sets of “single-shot” data at a time. More importantly,
the scope is capable of performing a FFT on the displayed and averaged timedomain data, and this information is also displayed on the screen and sent to an
output file. The data from this file was used as the “observed” amplified loop
EMF at the resonant frequency of the DRC.
Figure 4-4 shows the results of the experiment. In the plot, four sets of
data are shown. There are two sets of experimental data and two sets of
theoretical values taken from the output of the Mathematica program. The first
set of experimental data, referred to as “Corrected Scope Data”, is the raw data
taken from the FFT calculated by the oscilloscope, with an adjustment made to
136
account for a current fluctuation that occurred during the experiment. There was
a distinct drop in the mean value of the current towards the end of the experiment,
from 0.015 nA to 0.013 nA. Therefore, the voltage values for DP separations of
90 ps and greater were scaled upwards by a factor of 15/13 (the relationship
between current and voltage is linear). The second set of experimental data
(“Cable Loss Added”) adds the cable and coupling loss in the system measured
by the network analyzer as described in Section 3.3.4. In that measurement, a
single 24 inch low-loss cable was connected to the DRC for analysis. The
analyzer measured the return loss after the signal traveled through the cable twice,
once on the way towards the DRC and once on the way back. In the current
experiment, two identical 24 inch low-loss cables were used, each traversed a
single time. Therefore, the loss measured previously should be identical to the
loss seen in this experiment for the cables and DRC coupling. However, due to
the general difficulty in accurately determining loss in microwave systems, error
bars of ± 5% are added to this data set.
137
Figure 4-4 Variation of DRC Signal with DP Separation
Observed oscilloscope data compared to a range for theoretical values.
Cable and coupling loss measured with a network analyzer was added to the
raw scope data to represent the true amplified loop EMF signal. Uncertainty
in charges per pulse, loop area, and pulse charge ratio were used to create
the range of theoretical values.
For the theoretical data, three main sources of uncertainty were identified.
While there exists some level of uncertainty in many of the parameters used in
the theoretical model, the effects are insignificant compared to the uncertainty in
current, measured loop area, and the DP charge ratio. The current, as measured
by the Faraday cup, showed a continuous variation of ± 0.001 nA in addition to
the drop in mean value towards the end of the experiment that has already been
accounted for. Measurements for the loop area were made with digital calipers,
but a determination of the exact size was difficult due to the irregular shape of the
loop. It should be noted that since no rapidly changing magnetic field can exist
inside a conductor, the only area receiving flux is the area up to the inside surface
of the wire, and this is the area that was measured. The nominal figure for the
loop area is 4.75×10-6 m2, with an estimated uncertainty of ± 0.1×10-6 m2. To
measure the DP charge ratio, which is the ratio of charge between the two single
pulses making up each double pulse system (in experiment the pulses will not be
138
exactly equal), each arm of the Michelson interferometer was blocked separately,
and the resulting current generated by the single pulse coming from each arm was
measured. The determined charge ratio between the pulses was estimated to be
1.2 ± 0.1. The fact that the pulses do not have equal charge causes the theoretical
EMF response to have a minimum that is non-zero. A low/high theoretical
boundary was then calculated by running the Mathematica program using the
low/high end of the three parameter ranges (see Appendix III).
The results of the experiment show excellent agreement between theory
and experiment. Of the 13 experimental data points taken, 9 fall within the
theoretical range when error bars are included.
The (near) zero in the
experimental data occurs at approximately 90-95 ps pulse separation, in
agreement with the theory.
4.3.2 Variation of Current
According to the DRC operation theory, the relationship between the
current passing through the cavity (or, the number of electrons per pulse) and the
loop EMF signal should be linear. Testing this relationship is important not only
because it helps verify the theory, but also because data in the calibration
experiment needed to be scaled due to current variations, and it was scaled
linearly according to the theory. A simple experiment was conducted where the
number of electrons per pulse (and therefore the current) was varied using beam
optics, and the resulting current passing through the DRC was monitored with the
Faraday cup. The corresponding loop EMF seen on the oscilloscope, adjusted for
cable loss, was then plotted for each current level. Error was again estimated at
± 5%. The results are shown in Figure 4-5. The results of the experiment show
that the relationship between current and signal is indeed nearly linear, as
expected from the theory. The signal amplitude is also close to the theoretical
values, which further supports the validity of the theory.
139
Figure 4-5 Variation of DRC Signal with Current
Variation in the amplified DRC signal when the number of electrons per
pulse, or equivalently the pulsed beam current, is varied. The theoretical
relationship is linear. Uncertainty in the data is again caused by uncertainty
in cable losses.
4.3.3 Compression
In this experiment the CRC and DRC are used in tandem. First, double
electron pulses are sent through the CRC at a power level of 5 W, with an initial
separation set at 60 ps, and the optical delay stage is adjusted until the DRC signal
is a maximum (signaling maximum compression). This means the pulses are
entering the CRC at the correct phase. The compression power levels are then
varied and the resulting scope signal from the DRC is recorded. The purpose of
the experiment it to simultaneously test the ability of the CRC to compress an
electron pulse, and to test the previous calibration of the DRC.
The experiment was conducted at two different current levels, or,
equivalently, at two different values of electrons per pulse, one high and one low.
The purpose of this was to test if any space-charge effects could be noticed, with
the assumption that the higher charge electron pulses would repel more strongly
and therefore require extra power to reach maximum compression. The low
140
current value was 0.003 ± 0.0002 nA and the high value was 0.016 ± 0.002 nA.
The forward power being delivered to the CRC was slightly unstable, and varied
by approximately ± 0.5 W at any given data point. Error bars were added to the
data to include these variations. After the initial data was taken, it was scaled
linearly to match the DRC calibration current level of 0.015 nA. The results of
the experiment are shown in Figure 4-6.
Figure 4-6 Electron Pulse Compression
Variation in the amplified DRC signal after electron pulses have passed
through the CRC at various compression power levels. The experiment was
conducted at two current levels to test for space-charge effects. Uncertainty
in the data is caused by uncertainty in cable losses, instability in power levels,
and variation in current levels.
The results show that the CRC is able to successfully compress the pulses,
and the DRC signal indeed matches up well with the previous calibration. At
zero power the double pulses (with an initial 60 ps separation) pass through the
CRC unaffected, and therefore should match the 60 ps separation DRC signal
level of approximately 0.003 V. This is indeed the case. Maximum compression
should correspond to approximately 0.0055, which again is what is observed.
There was no significant evidence of space-charge effects, as pulses of both high
141
and low current levels reached maximum compression around 6 W of power. As
the power level was raised above 6 W, the DRC signal decreases (as it should
since too much power will push the pulses past each other before reaching the
DRC), but the signal varies significantly more than at low power levels. This is
most likely due to the individual pulses deflecting each other off axis (a “spacecharge pulse collision”) and not passing directly through each other.
142
5 Conclusion
5.1 Summary
The primary goal of this research project was to explore the applicability of
RF cylindrical resonant cavities to the ongoing UEM project at UIC. The initial
plan for the project involved only one cavity (the CRC), which was to be added
to the existing UEM column, and used to temporally focus electron pulses that
were emitted from a pulsed laser-driven photocathode, and accelerated down the
column through the CRC. Along the way, I proposed that a resonant cavity (the
DRC) could also be used to detect the duration of pulses passing through it,
without disturbing the pulse itself; in this case, for pulses of about 20 ps duration
and longer. While this duration of pulse is longer than the sub-ps resolution goal
of the UIC UEM project, the DRC is still useful in determining the correct phase
relationship between pulse arrival and CRC EM field, and the approximate power
level at which maximum compression is achieved.
The underlying theory of cylindrical resonant cavity eigenmode EM fields,
along with the theory of power flow to resonant cavities and sources of loss is
covered in Chapter 2. Since the CRC needs to be driven by a high power RF
source, the general theory of RF transmission lines and specifically coaxial cables
is also presented. The process of eigenmode stimulation in a cavity by current
sources, specifically current in an inductive coupling loop, is described. The
concepts (and difficulties) of impedance matching a resonant cavity to a
transmission line are detailed, and a theory developed from the cavity eigenmode
stimulation process is used to determine the approximate size of a coupling loop
that will achieve proper matching.
Chapter 3 first presents the design criteria for the CRC, driven by the spacetime resolution goals of the project, with specific parameters given by the output
of a previously developed computer simulation using a theoretical AG electron
143
pulse propagation model. This model simulates the electron pulse dynamics as
the pulses travel down the UEM column, including the effects of RF compression.
The basic operating principle of the CRC is described, and it is shown that a lowloss dielectric insert with an axial hole can be used for size reduction without
significantly affecting operation. The design of the CRC is then detailed, and the
process of selecting parameter values for operating frequency, acceleration
voltage, field strength, etc. in order to meet the design criteria is described. Next,
the construction process for the CRC is explained, including a unique indium wire
sealing technique to aid in current flow and stability, as well as an alternate copper
coated dielectric cavity which decreases power loss by 20%. Results of the CRC
characterization process are presented, and this process was accomplished using
a network analyzer to determine the resonant frequency, coupling constant, and
total quality factor of the cavity. The cavity was shown to be impedance matched,
and to have a quality factor high enough to allow use with a 30 W RF source.
The theory of operation for the DRC is also presented in Chapter 3. This
theory is basically a modification of the well-known theory of eigenmode
stimulation by current sources inside a resonant cavity that have a sinusoidal time
dependence. The sinusoidal time-dependent current source (such as a 3 GHz
current in a coupling loop) is replaced by the current density equivalent to a single
electron pulse traveling down the axis of the cavity. The Fourier transform of
this current density is calculated, and the amplitude of the transform at the
resonant frequency of the DRC is found to vary with the duration of the pulse,
with longer duration pulses having a smaller resonant frequency amplitude. It is
postulated that the driven eigenmode field amplitude will vary with the Fourier
transform amplitude (at the resonant frequency) in the same way the field
amplitude varies with the amplitude of a pure sinusoidal source at the resonant
frequency. The DRC field amplitude can then be sampled by using the same type
of coupling loop that drives the CRC. This time, however, the loop would act as
144
a receiving antenna, picking up the magnetic flux from the field inside the DRC
and sending a signal through a transmission line that can be amplified and viewed
on a fast oscilloscope. This signal would be directly proportional to the DRC
stimulated eigenmode field amplitude. A Mathematica program was developed,
based on the new theory, which could calculate the coupling loop signal given a
certain pulse duration.
As the electron pulses generated by the laser-driven DC electron gun have
durations that are too short to effectively test and calibrate the DRC, an original
double pulse experiment was designed and the additional theory for it developed.
Two individual short electron pulses with a certain separation time between them
have a Fourier transform amplitude that varies with the separation time. In this
way, for example, a 60 ps pulse could be modelled by two 4 ps pulses with a 60
ps separation between their centers. The 5.54 GHz DRC would then be capable
of detecting this separation time. Again, another Mathematica program was
written to calculate the expected DRC signal given a particular double pulse
separation time.
Chapter 3 ends with a discussion concerning the synchronization process
between the pulsed laser driving the photocathode and the EM field phase in the
CRC. The electron pulses must arrive at the CRC when the EM field is at the
correct phase in order for compression to occur.
This phase/pulse arrival
relationship must then remain stable so each pulse is compressed in the same way.
Therefore, there are two basic issues to be addressed. First, there is the possibility
of long term drift between pulse arrival and EM field phase. This potential
problem is solved by using the laser to drive both the photocathode and the CRC
EM field. The other issue is instability in the system causing random time jitter,
which needs to be minimized. It was found that the high stability of the laser
itself was adequate to successfully perform the necessary experiments with an
acceptable level of jitter.
145
In Chapter 4, the results of experiments performed to test the theory for the
CRC and DRC are presented. It was found that assumptions made about effects
of end cap holes, and the axial hole in the CRC dielectric, were correct. One of
the experiments, concerning the diverging electric field caused by the end cap
holes, also provided some verification of the AG model computer simulation. A
calibration experiment for the DRC was successfully performed that strongly
supported the double pulse theory. Using the calibration data, a compression
experiment using the CRC and DRC together was performed, which showed
maximum compression occurring at 6 W of RF power. The linearly scaled DRC
signal data for this experiment matched the previous calibration data.
5.2 Future Prospects
There are many possible applications for RF and microwave resonant
cavities in the field of UEM which warrant further study. The addition of
dielectric material not only allows for size reduction as shown in this thesis, but
also creates the possibility to operate cavities in modes which have a z-dependent
field, such as the TM011 mode. The typical problem that occurs when using these
types of modes is that the pulse must travel through the cavity at the phase
velocity of the z-dependent EM field in to achieve proper compression. In a
vacuum-filled cavity this phase velocity is the velocity of light, but in a dielectricfilled cavity the phase velocity is lowered by a factor equal to the square root of
the dielectric constant. A YAG-filled cavity, for example, would have a phase
velocity of roughly c 3 , or about the velocity the pulses are accelerated to with
the UIC gun at 25 kV. Also, by using the TM011 mode there would be a reduction
in operating power, which reduces any heating related effects. In addition, the
issue of the diverging electric fields at the end cap holes would be greatly reduced,
since the TM011 eigenmode field is zero at the cavity ends.
Another area for possible research involves the copper coating process and
machinable dielectrics such as Rexolite. These dielectrics could be formed into
146
shapes whose eigenmodes can’t be directly solved for, but could be approximated
(and designed) with computer software.
The possibilities for creating
advantageous field patterns to reduce power while increasing compression (or
acceleration) would be endless. The copper coating process, which involves
vaporizing copper in the presence of the dielectric insert and then condensing the
copper on the dielectric surface, would allow relatively complex shapes to be
made into operating resonant cavities with specially designed field patterns.
As far as the DRC is concerned, the ability to successfully design, construct,
and impedance match higher frequency cavities (which would be capable of
measuring smaller pulses) could be explored. A 50 GHz cavity, which could be
a small vacuum-filled cylindrical cavity operating in a higher order mode, would
have the ability to detect compression of a 10 ps pulse down to 1 ps.
147
References
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
Veisz, L., et al., Hybrid dc-ac electron gun for fs-electron pulse generation.
New Journal of Physics, 2007. 9.
Luiten, O.J., et al., How to realize uniform three-dimensional ellipsoidal
electron bunches. Physical Review Letters, 2004. 93(9).
LaGrange, T., et al., Single-shot dynamic transmission electron
microscopy. Applied Physics Letters, 2006. 89(4).
Domer, H. and O. Bostanjoglo, High-speed transmission electron
microscope. Review of Scientific Instruments, 2003. 74(10): p. 4369-4372.
Bostanjoglo, O. and R. Liedtke, Tracing Fast Phase-Transitions by
Electron-Microscopy. Physica Status Solidi a-Applied Research, 1980.
60(2): p. 451-455.
Rose, A., Television pickup tubes and the problem of vision. Advances in
Electronics and Electron Physics, 1948. 1: p. 131-166.
Armstrong, M.R., et al., Practical considerations for high spatial and
temporal resolution dynamic transmission electron microscopy.
Ultramicroscopy, 2007. 107(4-5): p. 356-367.
Armstrong, M.R., et al., Prospects for electron imaging with ultrafast time
resolution. Applied Physics Letters, 2007. 90(11).
van Oudheusden, T., et al., Electron source concept for single-shot sub100 fs electron diffraction in the 100 keV range. Journal of Applied
Physics, 2007. 102(9): p. 093501.
Jackson, J.D., Classical electrodynamics. 3rd ed. 1999, New York: Wiley.
xxi, 808 p.
Harrington, R.F., Time-harmonic electromagnetic fields. McGraw-Hill
texts in electrical engineering. 1961, New York,: McGraw-Hill. 480 p.
Montgomery, C.G., R.H. Dicke, and E.M. Purcell, Principles of microwave
circuits. 1st ed. 1948, New York,: McGraw-Hill Book Co. xvi, 486 p.
Morse, P.M. and H. Feshbach, Methods of theoretical physics.
International series in pure and applied physics. 1953, New York,:
McGraw-Hill.
Collin, R.E., Field theory of guided waves. 2nd ed. 1991, New York: IEEE
Press. xii, 852 p.
Condon, E.U., Forced Oscillations in Cavity Resonators. Journal of
Applied Physics, 1941. 12(2): p. 129-132.
Berger, J.A., M.J. Greco, and W.A. Schroeder, High-power, femtosecond,
thermal-lens-shaped Yb: KGW oscillator. Optics express, 2008. 16(12): p.
8629-8640.
Berger, J.A., et al., DC photoelectron gun parameters for ultrafast electron
microscopy. Microscopy and Microanalysis, 2009. 15(04): p. 298-313.
148
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
Child, C.D., Discharge From Hot CaO. Physical Review (Series I), 1911.
32(5): p. 492-511.
Langmuir, I., The Effect of Space Charge and Initial Velocities on the
Potential Distribution and Thermionic Current between Parallel Plane
Electrodes. Physical Review, 1923. 21(4): p. 419-435.
Valfells, Á., et al., Effects of pulse-length and emitter area on virtual
cathode formation in electron guns. Physics of Plasmas, 2002. 9(5): p.
2377-2382.
Berger, J.A., Ultrafast Electron Microscopes: Design Criteria, Electron
Sources, and Column Modeling. 2013, Argonne National Laboratory.
Jensen, K.L., et al., Emittance of a field emission electron source. Journal
of Applied Physics, 2010. 107(1): p. 014903.
O’Shea, P.G., Reversible and irreversible emittance growth. Physical
Review E, 1998. 57(1): p. 1081-1087.
Berger, J.A., et al., Excited-state thermionic emission in III-antimonides:
Low emittance ultrafast photocathodes. Applied Physics Letters, 2012.
101(19): p. 194103.
Schroeder, W., et al., High Brightness Nano-Patterned Photocathode
Electron Sources for UEM. Microscopy and Microanalysis, 2010. 16(S2):
p. 492-493.
Rickman, B., et al., Intrinsic Electron Beam Emittance from Metal
Photocathodes: The Effect of the Electron Effective Mass. Physical review
letters, 2013. 111(23): p. 237401.
Schroeder, W., et al., Photoemission Processes for Ultrafast TEM.
Microscopy and Microanalysis, 2012. 18(S2): p. 582-583.
Rickman, B., et al., High Brightness Photocathodes for Ultrafast TEM: A
New Paradigm. Microscopy and Microanalysis, 2013. 19(S2): p. 11621163.
Michalik, A.M. and J.E. Sipe, Analytic model of electron pulse
propagation in ultrafast electron diffraction experiments. Journal of
Applied Physics, 2006. 99(5): p. 054908.
Michalik, A.M. and J.E. Sipe, Evolution of non-Gaussian electron bunches
in ultrafast electron diffraction experiments: Comparison to analytic
model. Journal of Applied Physics, 2009. 105(8): p. 084913.
Berger, J.A. and W.A. Schroeder, Semianalytic model of electron pulse
propagation: Magnetic lenses and rf pulse compression cavities. Journal
of Applied Physics, 2010. 108(12): p. 124905.
Rimington, N.W., et al., Thermal lens shaping in Brewster gain media: A
high-power, diode-pumped Nd:GdVO4 laser. Optics Express, 2004. 12(7):
p. 1426-1436.
Schieffer, S.L., et al., Low-threshold, dual-passive mode locking of a large
mode area Nd:GdVO4 laser. Optics Express, 2006. 14(15): p. 6694-6704.
149
34.
35.
36.
37.
38.
39.
40.
41.
42.
43.
Eastman, D.E., Photoelectric Work Functions of Transition, Rare-Earth,
and Noble Metals. Physical Review B, 1970. 2(1): p. 1-2.
Shen, Y.R., The principles of nonlinear optics. Wiley classics library ed.
Wiley classics library. 2003, Hoboken, N.J.: Wiley-Interscience. xii, 563
p.
Rimington, N.W., et al., Femtosecond Ti:sapphire oscillator electrooptically cavity dumped at 50 kHz. Applied Optics, 2001. 40(27): p. 48314835.
Strickland, D. and G. Mourou, Compression of amplified chirped optical
pulses. Optics Communications, 1985. 56(3): p. 219-221.
Togawa, K., et al., CeB6 electron gun for low-emittance injector. Physical
Review Special Topics - Accelerators and Beams, 2007. 10(2): p. 020703.
Zhan, Y. and P.D. Coleman, Far-IR optical properties of Y3Al5O12,
LiYF4, Cs2NaDyCl6, and Rb2NaYF6. Applied Optics, 1984. 23(4): p. 548551.
KAJFEZ, et al., Q-factor measurement with network analyzer. Vol. 32.
1984, New York, NY, ETATS-UNIS: Institute of Electrical and
Electronics Engineers.
Kajfez, D., Q‐Factor. Encyclopedia of RF and Microwave Engineering,
2005.
Kiewiet, F.B., et al., Femtosecond synchronization of a 3 GHz RF
oscillator to a mode-locked Ti:sapphire laser. Nuclear Instruments and
Methods in Physics Research Section A: Accelerators, Spectrometers,
Detectors and Associated Equipment, 2002. 484(1–3): p. 619-624.
Poppe, A., et al., Noise characterization of sub-10-fs Ti:sapphire
oscillators. IEEE Journal of Selected Topics in Quantum Electronics,
1998. 4(2): p. 179-184.
150
Appendix I Mathematica Code for DRC Single Pulse Analysis
Presented here is the Mathematica code that generated the plots used in
Section 3.5.1, showing the theoretical analysis for the EMF response of a single
pulse traveling through the DRC.
151
152
153
154
155
156
157
158
159
160
161
162
163
164
165
166
167
Appendix II Mathematica Code for DRC Double Pulse Analysis
Presented here is the Mathematica code that generated the plots used in
Section 3.5.2, showing the theoretical analysis for the EMF response of a double
pulse system (individual pulses identical) traveling through the DRC.
168
169
170
171
172
173
174
Appendix III
Mathematica Code for DRC Unequal DP Analysis
Presented here is the Mathematica code that generated the plots used in
Section 4.3.1, showing the theoretical range of values for the double pulse
(individual pulses not identical) calibration experiment for the DRC.
175
176
177
178
179