RF Design Lecture Notes
RF Design Lecture Notes
1 Introduction
This document is a short summary of the theory covered in the USPAS class Applied
Electromagnetism for Accelerator Design. This is a living document, and will be expanded
at a future date, please excuse small errors and sudden transitions. This document is
Version 1.
2 Resonator Theory
This section will provide insights for use in further sections by producing a generic formal-
ism to describe resonators with the explicit goal of treating resonating cavities with the
same mathematical treatment. Starting from the simple harmonic oscillator, damping and
driving terms will be added, and their effects derived. The treatment of electromagnetic
resonators by this formalism with then be justified, and a special case of interest will be
presented.
d2 x
+ ω02 x = 0 (1)
dt2
where x(t) is the oscillating position at time t, and ω0 is the oscillation frequency. The
general solution for this is characterized by an amplitude A and a phase ϕ
1
2.2 The Driven, Damped Harmonic Oscillator
For this to be a useful model we must consider damping. The form of damping that is of
interest to us (because it is the form of the losses in a resonator), is damping proportional
to the change in “position”, which has a form of
d2 x dx
+γ + ω02 x = 0, (3)
dt2 dt
where the damping coefficient γ has the dimension of frequency. This type of equation has
different forms of solution depending on the strength of the damping. If we choose the
damping to be weak, the general solution to this equation has the form
where √
γ2
ω1 = ω0 1− (5)
4ω02
and
1
γ1 = γ. (6)
2
Note that adding damping shifts the resonator frequency based on how strong the damp-
ing is. From this we can define a “quality factor” Q which is related to the rate at which
the resonator loses energy. Q is defined by the equation Q = ω0 /2γ with weak damping
characterized by Q ≫ 1. For a typical superconducting cavity, Q is generally ∼ 5 × 109 ,
justifying the assumption of weak damping.
The effects of a driving term must also be considered. For an arbitrary driving term of
the form f (t), the differential equation becomes
d2 x dx
2
+γ + ω02 x = f (t), (7)
dt dt
with the solution (for γ = 0)
∫
ẋ0 1 t ( ( ))
x(t) = x0 cos ω0 t + sin ω0 t + sin ω0 t − t′ f (t′ )dt′ (8)
ω0 ω0 0
where x0 and ẋ0 are the initial position and velocity. The third term of this solution gives
the contribution from the driving term, and it is worthwhile to notice that a harmonic
driving term with frequency equal to the resonant frequency of the oscillator will produce
the largest oscillations, as expected.
2
2.3 Oscillator Behavior Near Resonance
It is most useful to examine the response of the weakly damped oscillator to driving near
its resonant frequency because this is the desired case for resonator operation. Assuming
a driving term that is purely harmonic (f (t) = f0 cos ωt) and ignoring transient behav-
ior yields a solution with a relatively simple form. These assumptions are very justified
in almost every accelerator application; most changes made to the driving term of the
cavity have a time scale much larger than the RF period (which is usually on the order
of nanoseconds). This type of differential equation is solved quite simply by assuming a
complex solution and writing
d2 Ξ dΞ
+γ + ω02 Ξ = f0 e−iωt (9)
dt2 dt
with x(t) = ℜΞ(t). We are seeking a solution of the form Ξ(t) = Ξ0 e−iωt where Ξ0 is also a
complex number of the form Ξ0 = |Ξ0 |eiϕ . Solving for the real variable of interest x gives
x(t) = ℜΞ(t) = ℜ(|Ξ0 |e−iωt+iϕ ) = |Ξ0 | cos (ωt − ϕ). Plugging this form of the solution into
the differential equation gives
( 2 )
−ω − iωγ + ω02 Ξ0 = f0 (10)
simplifying to
f0
Ξ0 = (11)
ω02 − ω 2 − iωγ
with squared amplitude of
f02
|Ξ0 |2 = . (12)
(ω02 − ω 2 )2 + ω 2 γ 2
The resulting behavior can be see in Figure 1 for a variety of damping coefficients γ [1].
Again, it is easy to see that the maximum response will be shifted slightly depending on
the strength of the damping. This shift can be neglected for γ ≪ ω0 , and as demonstrated
earlier, this is a very good approximation for superconducting resonators. Another impor-
tant feature of these curves is the characteristic width of each curve. This width (∆ω),
defined as the width at the level that is 3 dB below the maximum response, is equal to 2γ
where γ is the damping parameter. An alternative and equivalent definition of the Quality
Factor is Q = ω/∆ω.
The phase response of Ξ is also of interest. This phase ϕ can be interpreted as the
difference in phase between the driving term and the response of the resonator, and is an
important quantity for resonator control.
( ) ( )
ℑΞ0 ωγ ω 1 γ
tan (ϕ) = = 2 = ( )2 (13)
ℜΞ0 ω0 − ω 2 ω0 ω0
1 − ωω0
3
Figure 1: Resonant curves for damped, driven harmonic oscillator with γ = 0.5, 0.2, 0.1, 0
A plot of ϕ can be seen in Figure 2. The most important feature of this behavior is the
nearly linear region near resonance. Most cavity control systems treat the cavity response
as linear and must operate in this region to remain stable.
d2 x dx
+γ + ω02 x + αx3 = f0 cos ωt. (14)
dt2 dt
For the purposes of this application, it can be assumed both the damping, driving, and
non-linear terms are small compared to the frequency ω0 . Additionally, we will only look
for solutions where ω ≃ ω0 . Using the standard van der Pol transformation [2], seen in
Figure 3, we transform into a rotating coordinate frame. Using the transformations
ẋ
u = x cos ωt − sin ωt (15)
ω
4
Figure 2: Oscillator phase shift compared to driving term versus detuning. Note the nearly
linear region near resonance.
and
ẋ
v = −x sin ωt − cos ωt, (16)
ω
we arrive at the following differential equations for u and v:
1
u̇ = [−(ω 2 − ω02 )(u cos ωt − v sin ωt) − ωγ(u sin ωt + v cos ωt)
ω
+ α(u cos ωt − v sin ωt)3 − f0 cosωt] sin ωt (17)
1
v̇ = [−(ω 2 − ω02 )(u cos ωt − v sin ωt) − ωγ(u sin ωt + v cos ωt)
ω
+ α(u cos ωt − v sin ωt)3 − f0 cosωt] cos ωt. (18)
Because we are assuming small non-linearities and constant frequency (ω), we are only
interested in the average behavior of these functions. Averaging over a period of 2π/ω, we
get ( )
1 3
u̇ = −ωγu + (ω − ω0 )v − α(u + v )v
2 2 2 2
(19)
2ω 4
and ( )
1 3
v̇ = ωγv − (ω − ω0 )u + α(u + v )u − f0 .
2 2 2 2
(20)
2ω 4
5
Figure 3: Van der Pol coordinate transformation into a frame rotating at ω.
In reality, these are not the values of interest for our resonator. For √ more meaningful
results, we again transform these quantities into the magnitude (r = u2 + v 2 ) and phase
(ϕ = arctan (v/u)) in the rotating frame. This results in the following differential equations:
1
ṙ = (−ωγr − f0 sin ϕ) (21)
2ω
and ( )
1 3 3
ϕ̇ = −(ω − ω0 )r + αr − f0 cos ϕ .
2 2
(22)
2ω 4
Again, we can simplify by assuming a steady state solution: ṙ = ϕ̇ = 0. The resulting
amplitude and phase plots can be seen in Figures 4 and 5, and the non-linear response
is evident. Because of this detuning, both the amplitude and phase become multi-valued
for certain ranges of detuning, and if the effect is severe enough, controlling the resonator
behavior can become prohibitively challenging.
6
Figure 4: Oscillator amplitude versus detuning. Circles represent α = −0.002, crosses
represent α = 0.
∇·B =0 (23)
∂B
∇×E+ =0 (24)
∂t
∇·D =0 (25)
∂D
∇×H − =0 (26)
∂t
7
Figure 5: Oscillator phase versus detuning. Circles represent α = −0.002, crosses represent
α = 0.
where E is the electric field, B is the magnetic field, and D (equal to ϵ0 E in this case) is
the electric displacement field. Assuming a harmonic time dependance of the form e−iωt
gives
∇·B =0 (27)
∇ × E − iωB = 0 (28)
∇·D =0 (29)
∇ × H + iωD = 0 . (30)
From this form, we can construct an arbitrary solution using Fourier superposition. As-
suming that we are in a region where there are no losses (µ and ϵ are real and positive),
Maxwell’s Equations reduce to
∇ × E − iωB = 0 (31)
∇ × B + iωµϵE = 0 (32)
8
where the divergence equations can be found from taking the divergences of these two
equations. Combining these two equations we get the Helmholtz wave equations:
∇2 E + µϵω 2 E = 0 (33)
∇ B + µϵω B = 0 .
2 2
(34)
√
A solution to equations of this form is a plane wave of the form eikx−iωt , where k = µϵω is
the wave number. While solutions of this form are discussed many places at great length,
the harmonic behavior allows us to use the same differential formulations to treat the
resonator. For the remainder of this document, we will only consider harmonic behavior
in vacuum where ϵ = ϵ0 and µ = µ0 with c2 ϵ0 µ0 = 1.
n̂ × E
⃗ = 0, n̂ · H
⃗ =0 (35)
where n̂ is the conductor surface’s normal vector. Let us consider a cylindrical waveguide of
radius R that is infinite in length. Defining our cylindrical coordinate system in the usual
way, we can use the generalized plane wave solution to the Helmholtz equation derived in
the previous section, we can assume the fields take the form
⃗ x, t) = E(ρ,
E(⃗ ⃗ ϕ)eikz−iωt (36)
⃗ x, t) = H(ρ,
H(⃗ ⃗ ϕ)eikz−iωt . (37)
For the purposes of an accelerating cavity, it is desired to have longitudinal electric fields
with no longitudinal magnetic fields, i.e., Bz = 0. Additionally, we know that the depen-
dence of the fields on ϕ must be periodic, so we assume a dependence of the form eimϕ
where m is an integer. This gives us a form for the electric field of
9
which we can plug into the wave equation:
( )
1 ∂2z 1 d df m2 ω2
= ∇2
z =⇒ ρ − f+ −k f =0
2
(39)
c2 ∂t2 ρ dρ dρ r2 c2
to get the form of f (ρ). The solution of this equation is given by standard Bessel functions
f (ρ) = E0 Jm (k⊥ ρ) where m is the order of the Bessel function and
1√ 2
k⊥ = ω − c2 k 2 . (40)
c
To satisfy the boundary conditions at ρ = R, Ez must be zero. This requires that k⊥ R
is equal to one of the Bessel zeros, jm,n (where n is an integer indicating the zero), giving
k⊥ = jm,n /R. Eliminating k⊥ from our equations gives our definition of the wave number
k for this geometry:
( )1/2
2
ω 2 jm,n
km,n = ± − 2 . (41)
c2 R
It is of interest to note that this gives a lower bound on the frequency for a traveling
wave solution. For ωjm,n c/R, the wave number becomes imaginary, resulting in a solution
known as an “evanescent” wave. This frequency is known as the “cut-off” frequency, with
larger diameter waveguides able to support lower frequencies. This wave exponentially
decays/increases as it propagates, and is of interest in the study of resonators only for
study of mode propagation into and out of the resonator through beam pipes or other
ports. For the remainder of this section, we will only consider traveling wave solutions.
Now that the Ez has been defined and remembering that Bz = 0, we can use ∇ × E = iωB
and c2 ∇ × B = −iωE to find the remaining components of both fields: Eϕ , Eρ , Bϕ , Bρ .
These component are
ikm,n R ′ ( ρ ) ikm,n z−iωt−imϕ
Eρ = E0 Jm jm,n e (42)
jm,n R
10
Eϕ (ρ = R) = 0 is automatically satisfied because this component has the same radial
dependence as Ez .
This derivation for T M modes is based on the condition that Bz = 0. One can follow
the same approach starting with the condition of Ez = 0 to obtain another family of modes,
T ransverse Electric (T E) modes. As these are not useful for accelerator applications, they
are beyond the scope of this work.
11
This shift to a discrete set of modes (conventionally referred to as T Mmnl modes) can be
repeated for T E modes as well.
E = E0 e−iωt (59)
12
and defining the electrical conductivity σ such that
⃗
⃗j = σ E (60)
where ⃗j is the surface current, we can show that, for a good conductor,
∇2 E
⃗ = τn2 E
⃗ (61)
√
where τn2 = iωσµ0 . For a simple geometry, such as an infinite half-plane of conductor
at x > 0 with an applied electric field that is only in the z direction, the solution to this
equation is
−x −ix
Ez = E0 e−τn x = E0 e δ e δ (62)
where
1
δ=√ . (63)
πf µ0 σ
This quantity, δ, is called the “skin depth”, and gives a characteristic length for how deeply
the electromagnetic fields penetrate the conductor. As an example of this, a copper surface
with σ = 6 × 107 S/m at a frequency of 1 GHz has a skin depth of approximately 2 µm.
Also note that as the frequency increases, this skin depth decreases. Because this skin
depth is so small compared to the wavelength of the resonator, it can be ignored in the
treatment of the cavity properties. In the geometry above, the current density and losses
can also be quantified. The current and field of this configuration are
13
of the cavity. There are three major scaling figures of merit. First, and perhaps most
important, is the Accelerating Voltage (Vacc ). This is the measure of how much energy a
charged particle passing through a cavity will gain, and it is defined as
∫ +∞ ∫ +∞ ( )
ωz
Vacc = Eacc (z) cos (ωt + ϕ)dz = Eacc (z) cos + ϕ dz (67)
−∞ −∞ βc
where z axis is the beam axis, β = v/c, ϕ is a phase offset that maximizes Vacc and
assuming that β doesn’t change. This last approximation is equivalent to saying that the
energy gained in the cavity is small compared to its totally energy. This approximation
is not valid if a high degree of accuracy is required, say, for particle tracking. Practically,
this can be calculated as
√( )2 (∫ +∞ )2
∫ +∞
Vacc = Eacc sin (ωt)dz + Eacc cos (ωt)dz . (68)
−∞ −∞
For the pillbox cavity, this integral has an analytic solution because the accelerating electric
field is constant in z.
( )
∫ +∞
∫ L
ω0 L
sin 2βc
Vacc =
Eacc (z)eiωt+iϕ dt
= E0
eiω0 z/βc dz
= LE0 ( ) (69)
−∞ 0 ω0 L
2βc
The length of the cavity, at this stage, remains unspecified. An important consideration of
the length is that this type of cavity is often operated in series as can be seen in Figure 6,
with each cavity being π ahead of the previous one in phase. It is obvious to see that it is
desirable to have the time it takes the particle to transit the cavity be equal to the time it
takes the fields to reverse:
L π βλ
= ;L = . (70)
βc ω0 2
14
Figure 6: An example of pillbox type cavities coupled together.
It is important to note that, in reality, a given length of cavity L gives an optimum particle
velocity βopt . This means that the T T F varies with β for a given cavity geometry:
( ) ( )
πβ
sin ω2βc
0L
sin 2βopt
T T F (β) = ( ) = ( ) . (73)
ω0 L πβopt
2βc 2β
Figure 7 shows this curve for a cavity designed for βopt = 0.50 for 2, 4, and 8 cavities in
series. Notice that as the number of synchronized cavities (“gaps”) increases, the range
of velocities that can be efficiently accelerated shrinks. This is of considerable importance
for an accelerator that has to accelerate a beam with a velocity that changes significantly.
While larger numbers of synchronized gaps mean a higher amount of acceleration per
meter of constructed accelerator, this may require more cavities designed for a variety of
different βopt . Two extreme examples of this are heavy ion and electron accelerators. An
electron is such a light particle that it very rapidly approaches the speed of light, thus only
requiring one resonator geometry. To give an example of how fast this happens, remember
that the International Linear Collider has a resonator designed to give roughly 30 MeV of
acceleration with each resonator having 9 cells. This means that, from rest, an electron
would be going 99%(!!!) the speed of light at the exit of the first resonator (even after
including the reduction in acceleration efficiency when the electron’s speed is mismatched
relative to the optimium beta). For a heavy ion accelerator, the situation is much different.
For a beam of uranium (common in nuclear physics), over 500,000 times as much energy
(almost 2 TeV!!!) is required to reach the same velocity, even in its highest charge state.
Even using the most flexible accelerating resonators (2-gap structures), this would take at
15
Figure 7: The Transit Time Factor (TTF) for 2, 4, and 8 cavities in series with βopt = 0.50.
3.7 Cavity Figures of Merit - Peak Surface Electric and Magnetic Field
The remaining scaling figures of merit are the peak surface electric and magnetic fields.
For the pillbox cavities, these values
Epk = E0 (74)
E0 E0
Hpk = J1 (1.84) = (75)
η 647 Ω
are located on the end walls of the cavity.
16
3.8 Cavity Figures of Merit - Geometry and Quality Factor
The efficiency with which the resonator stores energy is of principal importance. For
superconducting cavities, this is especially true because all dissipated power is exhausted
into liquid helium. The figure of merit for this efficiency is the Quality Factor:
ω0 U
Q= . (76)
Pd
From our study of the harmonic oscillator, remember that we defined the quality factor as
the ratio of the energy stored in the resonator to the energy lost per radian of oscillation.
Inspection shows that this definition is consistent with the resonator definition. In addition
to the Quality Factor, there is a figure of merit that does not depend on the surface
resistance (Rs ) called the Geometry Factor. This is defined as
ωU
G = Rs Q = ( ) (77)
Pd
Rs
and is solely dependent on the geometry. This is a more useful figure of merit for cavity
design because the Rs depends on many factors outside of the electromagnetic design such
as material purity, surface quality, heat treatment of the cavity, and many more. This
is also a useful figure to compare different geometries and cavities of different frequencies
because it is independent of frequency [4, pp. 43-44]. To calculate the Geometry Factor
for a given geometry, we must first calculate the stored energy (U ) and dissipated power
(Pd ). Remembering that
dPd 1
H⃗
2
= Rs
(78)
dA 2
⃗
where H
is the magnetic field at dA, and that the Geometry Factor requires P /R
d s
the equation for the total dissipated power becomes
∫
Pd 1
⃗
2
= H
dA
(79)
Rs 2 S
where S includes all RF surfaces in the cavity. Note that we are assuming that Rs is
constant over the surface, which is unlikely to be strictly true. While it is expected that its
value will not vary much over the surface, accurate measurements of Rs and its variation
over the surface of the cavity are extremely difficult. In any case, the variation will not
likely be predictable, and should not affect cavity design. In addition to Pd , we need the
stored energy in the cavity. In general, the time averaged stored energy density in the
cavity is [5] ( )
1 ⃗ 2 + 1 |B|
⃗ 2 .
u= ϵ|E| (80)
2 µ
17
At any given time, the energy is stored in the electric and magnetic fields of the cav-
ity. However, the total energy is fixed, and in this time-averaged form, both electric and
magnetic contributions are equal. This means we can simplify this density to
ϵ 2 1 2
u= E
=
B
(81)
2 2µ
which yields ∫ ∫
ϵ0
2 1 2
U= E0
dV =
B0
dV. (82)
2 V 2µ0 V
This gives us
∫
⃗
ω0 µ0 V
H
2 dV
G = Rs Q = ∫ . (83)
S
H dA
2
With some calculation, these integrals can be done for the pillbox cavity as described [4,
p. 46]:
πϵ0 E02 2
U= J1 (2.405)LR2 (84)
2
πRs E02 2
Pd = J1 (2.405)R(R + L) (85)
η2
L
ω0 µ0 LR2 2.405L 453 R
G= = η = L
Ω. (86)
2(R2 + RL) 2(R + L) 1+ R
By remembering that we desire
βλ
L= (87)
2
we find that, for a pillbox cavity,
L βπ
= (88)
R 2.405
which gives a Geometry Factor for ALL optimized pillbox cavities of
G = 257β Ω. (89)
As a comparison, the Facility for Rare Isotope Beams (FRIB) calls for a cavity (of a
totally different geometry) designed for β = 0.29. An equivalent pillbox cavity would have
G ≈ 74 Ω with the current FRIB cavity design having G = 76 Ω. While a pillbox cavity is
impractical at these velocities (it would require a cavity that has a diameter over 5 times
its length), this demonstrates the usefulness of this figure of merit for comparison between
cavities.
18
3.9 Cavity Figures of Merit - R/Q
The third major efficiency figure of merit is the R/Q. This is a measure of how efficiently
the cavity transfers its stored energy to the beam passing through it. This is the ratio of
the effective shunt impedance to the cavity’s quality factor, and this is a point that needs
to be stressed. There are many definitions of the shunt impedance depending on context
or even convention, and I will be using the common accelerator conventional definition [4,
p. 47]:
V2
Ra = acc (90)
Pd
in units of Ω per cell. This is further complicated by the fact the the majority of this
document will be spent discussing cavities that, while one “cell”, have two accelerating
gaps. From all this, the definition of the R/Q is
2
Vacc
R/Q = . (91)
ω0 U
One final note is that this is the effective shunt impedance. While I will omit the eff
subscript from R, I will be using Vacc instead of V0 , which is implied by “shunt impedance”
in some circles. For a pillbox cavity, this can be readily calculated
L
R/Q = 150 Ω = 196β Ω. (92)
R
19
and a cavity radius of R = 0.534 m. This cavity has very poor mechanical qualities from
such a large ratio of radius to gap size, in addition to the loss in accelerating efficiency due
to the small gap size.
4 Coaxial Resonators
In the previous section, we derived the mode structures and properties of resonators con-
structed from a terminated section of cylindrical waveguide. The scaling of these structures
means they become undesirable for acceleration of very low velocity particles. Coaxial res-
onators originate from the mode structure of coaxial transmission lines. These resonators
offer a practical option for lower velocity particle acceleration because they can achieve the
lower fundamental frequency required for a lower optimum velocity of beam while main-
taining efficiency of acceleration. In payment for this, these cavities have more complex
geometries that require more material, construction, and processing cost and time. Their
mode structures, scaling parameters, and figures of merit will be derived and compared
with the pillbox cavity type, and their use in heavy ion accelerators will be justified.
20
4.2 The Ideal Quarter Wave Resonator
First, we will discuss a coaxial resonator with one short termination (at z = L) and one
ideal open termination (at z = 0), seen in Figure 8. Coaxial lines are typified by a radial
electric field and an azimuthal magnetic field, with a 1/ρ dependence in amplitude. The
traveling wave formulation of this gives
I0 µ0 c iωt+ikz
Eρ = − e (94)
2πρ
I0 µ0 iωt+ikz
Bϕ = e . (95)
2πr
21
Again, summing two identical traveling waves with opposite directions of propagation
eikz + e−ikz = 2 cos (kz) (96)
eikz − e−ikz = 2i sin (kz) (97)
we can satisfy the boundary conditions to get
E0 a ( pπz )
Eρ = cos sin (ωt) (98)
ρ 2L
E0 a ( pπz )
Bϕ = − sin cos (ωt) (99)
ρc 2L
where p is an integer giving the order of the mode in z, E0 is the peak surface electric
field (located at z = 0 and ρ = a), a and b are the inner and outer conductor radii, and
ω = pcπ/2L. From this, it is trivial to see that the peak surface magnetic fields is E0 /c
and is located at z = L and ρ = a. Rearranging this last equation gives us L = λ/4 for
the lowest mode, p = 1, giving the cavity its name: Quarter Wave Resonator (QWR). The
accelerating axis for this cavity is at z = ϕ = 0, giving an ideal QWR accelerating field seen
in Figure 9. Using these field distributions, it is straight-forward to find an ideal QWRs
figures of merit. The stored energy, U , is
∫
ϵ0
2
U = E dV (100)
2 V
∫ ∫ 2π ∫ L ( πz )
ϵ0 E02 a2 b 1
= dρ dϕ cos2 dz (101)
2 a ρ 0 0 2L
( )
ϵπE02 λ b
= ln (102)
8 a
remembering that L = λ/4. The dissipated power, Pd , is also straight-forward to calculate:
∫ [ ( ) ( )]
Pd 1
πE02 a2 ϵ0 λ 1 1 b
= 2 B 2
dA = + + 2 ln (103)
Rs 2µ0 S 2µ 0 4 a b a
remembering to sum the contributions from the inner conductor, outer conductor, and the
short plate. The dissipated power on the open termination is zero because the magnetic
field there is zero. These numbers allow us to calculate the geometry factor G:
( )
cπµ0 ln ab
G = [λ (1 1) ( b )] . (104)
2 a + b + 4 ln a
The calculation of the optimum beta is less straight-forward for this geometry compared to
the pillbox cavity. The integral is made complicated by the form of the accelerating field
which gives an equation for the accelerating voltage
∫ b ( )
E0 a ωx
Vacc = 2 sin dx (105)
a x βc
22
Figure 9: An example of the accelerating electric field for an ideal coaxial resonator. The
beam axis passes radially through the cavity at z = 0.
that does not have an analytic solution. We can sidestep this problem by remembering
that we can find the optimum velocity by solving for when the derivative of Vacc is zero.
( )
∫ b cos ωx ωx [∫ a ( ) ]
∂Vacc βc β 2 c 2E0 aω ωx
= 2E0 a − dx = 2
cos dx (106)
∂β a x β c b βc
which gives [ ( ) ( )]
∂Vacc 2E0 a ωa ωb
= sin − sin = 0. (107)
∂β β βc βc
Using the trigonometric identity
( ) ( )
u+v u−v
sin u − sin v = 2 cos sin (108)
2 2
23
we obtain ( ) ( )
ω(a + b) ω(b − a)
cos sin = 0. (109)
2βc 2βc
This gives conditions of
nβλ
nβλ = b − a; =b+a (110)
2
where n is an integer. The first equation give solutions where the gap size (b − a) is
extremely large, large enough for the particle to see at least one full RF period. Thus,
these solutions are likely to be minima, and will be disregarded. The second equation gives
the relationship that we expected for n = 1, that
βλ
=b+a (111)
2
which is identical to the relationship assumed for the pillbox cavity (remembering that
b + a is the gap-to-gap distance). For an ideal QWR, knowing the frequency and optimum
particle velocity desired specifies two of the three variables for the cavity, with the gap size
b − a being the only unspecified value. A plot of R/Q and G versus gap size (b − a) can
be seen in Figure 10. In this plot, it is easy to see that a large gap size results in a more
efficient cavity. This is tempered, however, with the knowledge that the larger the gap
for a given βopt , the smaller the inner conductor diameter. The remaining figures of merit
for the ideal QWR are the peak surface electric and magnetic fields. These are both on
the inner conductor, and their magnitudes scale like 1/a. The limit to how much we can
reduce a as we increase R/Q and G is driven by the requirement of minimizing these peak
surface fields.
From Figure 11, we can see that the ideal open termination is not truly representative
of the electric field region of a practical QWR. While a detailed treatment of QWR design
is outside the scope of this document, an issue of some importance is the asymmetry of this
geometry not present in the ideal QWR fields. This asymmetry causes steering electric and
magnetic fields on the beam axis, especially when the frequency of the cavity is increased
above ∼ 200 MHz. This steering can be corrected to a degree [6], but QWR usage is
generally limited to applications at lower frequencies, and therefore lower velocities.
24
Figure 10: R/Q and G versus gap size for an ideal 80.5 MHz, β = 0.041 QWR.
E0 a ( pπz )
Bϕ = − sin cos (ωt) (113)
ρc L
where p is an integer giving the order of the mode in z, E0 is the peak surface electric
field, a is the inner conductor radius, and ω = pcπ/L. Rearranging this last equation gives
us L = λ/2 for the lowest mode, p = 1, giving the cavity its name: Half Wave Resonator
(HWR). These equations are all very similar to a QWR, and in fact are only different
because of the change in definition of the length, L. An ideal HWR is effectively an ideal
QWR, but reflected about the QWR’s ideal open boundary condition. This means that
almost all of the derived quantities for the QWR can be used here with minor modification.
25
Figure 11: FRIB Quarter Wave Resonator designs.
26
Figure 12: Schematic of an ideal HWR.
βλ
= b + a. (120)
2
remain identically valid for the HWR. It should be noted that, for a given length, the
frequency of the structure has doubled. In addition, because the losses are driven by the
magnetic fields, the losses for a given acceleration will double going from a QWR to a
HWR. This will double the HWR’s R/Q, but because the stored energy also doubles, G
remains the same. In addition, the location of the peak surface electric and magnetic fields
remain the same, with the caveat that there are not two peak surface magnetic fields, at
27
z = L/2 and z = −L/2. The benefit of the HWR geometry is the symmetry along the long
axis, meaning that steering fields are only generated from manufacturing errors as opposed
to the inherent geometry itself. This means that the frequency limitations of the QWR do
not apply to HWR, allowing acceleration of higher velocity particles.
28
Factor of ∼ 80 Ω, where the realistic FRIB design for a similar cavity has a Geometry
Factor of ∼ 100 Ω. This is because the ideal HWR geometry is not very efficient; a more
complex inner and outer conductor geometry provides this increase in efficiency.
Figure 13: Geometry Factor plotted versus β for the ideal cavity geometries.
29
the 3QWR will have the same G while the R/Q will be 1/3 lower.
30
as a good fit for the BCS component of the surface resistance of niobium where RBCS
is in ohms, T is the temperature in Kelvin, and f is the frequency in GHz. CRRR is
a factor based on the material purity, 1 for reactor grade niobium, and approximately
1.5 for RRR = 300 niobium. From this, it is easy to see that it is desirable to operate
superconducting cavities at the lowest temperature possible. In practice, RBCS isn’t the
only contribution to the surface resistance. The remaining resistance is a material/cavity
processing dependent resistance called the “residual resistance” Rres that does not depend
on temperature or frequency. While residual resistances of below 1 nΩ have been achieved,
10 nΩ is more reliably reproducible for large scale cavity fabrication [13]. If the cavity is
operated at or below 2 K, the residual losses tend to dominate performance. An approach
for estimating the skin depth of a superconducting material can be derived [4, pp. 85-88]
by using a two-fluid model for the electrons in the material, one fluid composed of the
Cooper pairs, and the other the unbound electrons. This gives a penetration depth (called
the London Length [14]) of
m
λ2L = (130)
ns e2 µ0
where m and e are the mass and charge of the electron, and ns is the density of the
superconducting electron pairs. For niobium, this penetration depth is on the order of tens
of nanometers, meaning that the RF performance is dominated by the quality of a very
thin layer on the surface of the material. The temperature dependence for λL has been
empirically found as [15]
[ ( )4 ]−1/2
T
λL (T ) ≈ λL (0) 1 − . (131)
Tc
While it becomes energetically favorable for electrons to form Cooper pairs below Tc , this is
offset by the application of a magnetic field. Magnetic flux is expelled in the transition from
normal to superconducting (called the Meissner Effect [16]), but the fields still penetrate
the material on the order of the London Length. BCS theory approximates the temperature
dependence of this critical magnetic field as
[ ( )2 ]
T
Hc (T ) = Hc (0) 1 − . (132)
Tc
Finally, niobium is a Type-2 superconductor. This means that the transition from su-
perconducting to normal conducting in the presence of a magnetic field actually has two
critical magnetic field levels, Hc1 and Hc2 , in contrast to a Type-1 superconductor in which
the superconductivity breaks completely at one field level. This behavior can be see in Fig-
ure 14. One of the reasons niobium is used for superconducting accelerators is the relatively
high Tc1 , one of the highest for any Type-2 superconductor.
31
Figure 14: Magnetic field just inside superconducting material around transition for Type
1 and 2 materials. Note the sudden penetration for Type 1 and the gradual break down
for Type 2.
32
5.3 The Challenges of Superconducting Materials
While the decrease in required power is substantial, there are drawbacks to using super-
conducting cavities. The cryogenic engineering talent and facilities required is a dramatic
investment, and can be a significant portion of the overall cost of an accelerator. Addi-
tionally, the ultimate performance and achievable accelerating gradient is limited by the
peak surface magnetic field. Even for an extremely well prepared cavity, the peak sur-
face magnetic field is ∼ 220 mT, and for low velocity cavities, even with state of the art
processing, this number is closer to 120 mT. Because the losses in a superconductor also
increase strongly with frequency, this means that superconducting cavities are limited to
operation under 2 GHz for practical reasons. Even the International Linear Collider which
is designed to operate at 1.3 GHz and 2 K has moved to a pulsed design to limit the load
on the cryogenic systems. Additionally, there are many processing and fabrication consid-
erations for these cavities to achieve a low surface resistance with little to no radiation.
These considerations will be discussed later in this document.
σ(4.2K)
RRR = ∝κ (134)
σ(300K)
where σ is the electrical conductivity and κ is the thermal conductivity. Material can be
readily obtained from suppliers with RRR of over 250, considered good enough to give
good cavity performance. This material is produced by repeated electron beam melting in
vacuum to drive off impurities, and can also be improved by baking over 1200◦ C in vacuum
with either titanium or Yttrium as a getter to remove impurities.
33
of different applications. These topics will be expanded upon with more specific design and
theory for couplers used for low beta superconducting resonators like those being designed
for the FRIB driver linac.
34
which gives
Q0 = QL (1 + β1 + β2 ). (139)
It is desired to measure Q0 directly, but in steady state, only the scattering parameters
(S11 and S21 ) can be measured. The scattering parameters are defined as
Pr
|S11 |2 = (140)
Pf
Pt
|S21 |2 = (141)
Pf
with Pf being the forward power and Pr being the total reverse power from the cavity, the
vector sum of the reflected and emitted traveling waves. β1 and β2 can be expressed in
terms of these scattering parameters [4, p. 48]
|S21 |2
β2 = (142)
1 − |S11 |2 − |S21 |2
1 − |S11 |
β1 = (143)
1 + |S11 |
or
1 + |S11 |
β1 = (144)
1 − |S11 |
depending on coupling strength. For β1 < 1 (Equation 143), the coupling is considered
“undercoupled,” for β1 > 1 (Equation 144), the coupling is considered “overcoupled.” The
remaining case, β = 1, is the “matched” condition. In this configuration, there is no reverse
power because the emitted and reflected power from port 1 are equal in magnitude and
opposite in phase. These equations are derived assuming the measurement is being done in
a steady-state and that β1 ≫ β2 so that Pt can be dropped while considering the behavior
of Port 1. It can be see from these equations that some additional knowledge is required
for accurate interpretation of the measured values, any measurement of |S11 | can result
in two different β1 results. This information can be found by changing the parameters of
the measurement. Observing the change in |S11 | because of adjusting the coupling is a
known way to determine if the cavity is over or under coupled. For example, increasing
the length of a probe coupler should increase the coupling strength β. If |S11 | increases
because of this change, the coupler is now overcoupled. If |S11 | drops, the coupler started
as undercoupled. This type of change must be done with care to avoid confusion from
crossing β = 1. Additionally, the Pf can be modulated, and the behavior of Pr and Pt can
be used to determine the coupling. This method is only realistic when the time constant
of the cavity
QL
τ= (145)
ω
35
is large enough to be measured easily. Because QL is so small when the cavity is normal
conducting (∼ 1000), it is only possible to measure the coupling by modulation of the
forward power when the cavity is superconducting, with QL ≈ 1 × 107 or higher. For a
realistic cavity, two couplers are used. The “input” coupler is a high power coupler designed
to be either close to matched or overcoupled depending on the situation. The second
coupler, called the “pickup” coupler, is used as a field probe, allowing simple measurement
of the cavity stored energy. This coupler is significantly undercoupled so that the power
radiating out of this coupler (Pt ) is very low, so it can be treated as a perturbation, greatly
simplifying the mathematical analysis of the measured results. A useful rule of thumb
when choosing the pickup coupling is to ensure that Pt is always below 1 mW, even at the
extreme limit of cavity performance.
36
effect on cavity frequency and performance must be taken into account during standard
cavity design. Additionally, the coupling is chosen based on an assumption of Q0 , but
in reality, this can vary from cavity to cavity. The ability to vary the input coupling in
situ adds mechanical complexity to coupler design, but allows fine tuning of each cavity
to either minimize power requirements on an overperforming cavity or allow operation of
a poorly performing cavity. This benefit extends to Dewar testing, where Q0 is varying
more widely than in real operation. A variable coupler allows for testing or conditioning
at a higher field by keeping the cavity close to matched for a much wider range of cavity
performance.
This equation is then discretized and applied to each element. Elements “interact” at fixed
places, generally along the outside edges, called nodes. The parameters of each element are
iterated with the goal of minimizing the global stored energy. The study and development
of these solvers is an active and growing field, but for the basic methods and mathematical
foundations of these solvers, “Electro-magnetics and Calculations of Fields” by Ida and
37
Figure 15: A cavity vacuum space meshed with tetrahedral elements, a coupler probe is in
the foreground with a beam port cup visible in the background.
Bastos [26] or “Finite-element Methods for Electromagnetics” by Humphries [27] are rec-
ommended as starting places. What this procedure allows is the accurate representation
of the fields in a complex geometry with evaluation of the electromagnetic figures of merit.
38
Figure 16: Structure of a first order (linear) tetrahedral element. Image credit to F.
Krawczyk.
39
7.4 An Aside - Computer Resource Management
Computer resource management is an important setup for large scale simulations. This
includes storing solutions files on the fastest available disk drive available. Using solid
state drives in place of traditional palette drives can decrease simulation time by 30 -
40%, depending on the type of solution being done, assuming a large enough solid state
drive can be acquired. Eigenmode simulations tend to be memory limited, as opposed to
processor speed limited, meaning that investments in upgrading available memory speed
and size can speed simulations considerably. Additionally, the temporary files generated
during simulation and results files can be extremely large. If the disk drive you are using
is 32-bit formatted, it will be unable to store these files. Some simulation software have
a feature that ensures no files are larger than the 4 Gigabyte limit for FAT32 formatted
drives. A 32-bit architecture also limits the amount of memory that can be allocated to
a single program to 2 Gigabytes as well, so moving to 64-bit architecture removes both
of these limitations. Depending on the software in use, multiple processor cores can be
used in parallel to increase solving speed. When starting a simulation, ensure that the
maximum number of processors, largest amount of memory, and fastest hard drive space
available are allocated for the job.
40
eigenmode type element, and the order of this element must be set. Material properties
are also set here, but for a simple high-frequency eigenmode solution, ϵ = ϵ0 and µ = µ0
are the only required definitions.
Figure 17: HWR vacuum model. Symmetry planes can be seen in blue, Perfect Electric
Boundaries can be seen in gray, the cavity origin is marked in red.
41
be seen in Figure 18.
Figure 18: Comparison of different meshing on a curved surface. Left: Poor meshing on a
curved surface. Right: Same surface with improved mesh density. Images generated using
ANSYS.
While most software has advanced automatic meshing algorithms, additional manual
refinement may be required based on the geometry being simulated. The first, most com-
mon, mesh optimization step is the setting of a maximum element edge length. This can
be done for both whole volumes and on areas of known high surface fields. For the vol-
ume sizing, it is important to set a global maximum value for an element edge length to
prevent drastic gradients in element size. The standard rule of thumb for electromagnetic
simulation calls for a maximum element edge length of λ/10 ≈ 9 cm for 322 MHz. For
complex geometries like the ones used for HWRs, additional surface refinement may be
required to accurately calculate surface losses and peak surface fields. For accurate surface
fields on the small curved surfaces, elements as small as 1 mm were found to be needed in
some locations of the cavity. If the transition between these element sizes is not carefully
controlled, poor mesh quality can result. Setting a global maximum on the element size
is a powerful tool for regulating this transition. A major factor in mesh quality tolerance
is the order of element chosen (i.e., what order function is used to represent the fields in
an element). As can be seen in Figure 19, increasing the mesh density reduces frequency
error down to a baseline inaccuracy that is solver dependent. For ANSYS-APDL, this
error was found to be roughly 1 part in 105 , or ∼ ±50 kHz out of 322 MHz. This level of
accuracy was achieved with second order elements at a much lower density than for first
order elements. While the solve times for the same accuracy is roughly equivalent, the
use of second order elements greatly reduces the meshing tolerances while giving similar or
42
Figure 19: Fractional frequency error for different mesh densities in identical geometries
for first and second order elements.
improved surface field accuracy. There are sometimes limitations from the meshing that
can help drive the decision to strive for first or second order meshing. Some geometries
are so complex that their mesh densities must be high for geometric fidelity. Depending on
the available computing hardware, this type of geometry can often require far more mesh
elements than could be solved with second order elements, but can be solved with first
order elements at a loss of accuracy. In this situation, all results, especially surface fields,
must be interpreted with this in mind. When the computing hardware allows, this type
of complex geometry should be solved using second order elements if possible, especially
if surface fields are of particular interest. Meshing must also be approached carefully with
geometries that are very simple. Meshing software will often generate very large elements
in these geometries if allowed, and care must be taken to ensure they do not grow large
enough to sacrifice accuracy. This is an ideal situation to use a global element edge length
limit.
Additional care must be taken when generating geometry to avoid causing meshing
problems. While this is a more serious issue for mechanical simulations, and will be dis-
cussed in greater detail in the relevant section, there are certain features that should be
avoided if possible, and these are places of optimization to generally reduce the size of simu-
lations without great sacrifice for accuracy. CAD geometries are represented in simulation
43
software as a continuous and closed collection of surface, each bounded by lines. Mesh
structure must respect the shape and boundary of these surfaces, and care should be taken
to not create very small areas unnecessarily, or otherwise collect many boundaries together
without purpose. An example of a mesh that is overly dense at surface intersections can be
see in Figure 20. This type of feature can be tolerated if the solution time isn’t excessive,
but when working on a very large simulation, it may be beneficial to find instances like
this and correct the geometry to allow for more optimal meshing.
Figure 20: The intersection of four surfaces can force excess mesh density, indicated by the
red arrow. Images generated using ANSYS.
Automatic meshing algorithms often make meshing decisions based strictly on the ge-
ometry, not on the structure of the mode being simulated. This is understandable because
the information is not available to the program before the simulation is complete. Some
programs like ANALYST have automatic iteration algorithms that use solution data to
adjust and optimize the meshing until the solution has converged to the accuracy desired.
For most simulations software, foresight can dramatically improve the quality of the sim-
ulation results. We’ve touched on this for surface fields and especially external surfaces
44
with a large curvature, but this is also quite important for the accelerating field profile.
Large elements on the beam axis result in a non-physical accelerating field because the
fields cannot be properly interpolated between elements to give a smooth shape. Improv-
ing the density of the meshing around the beam axis gives a more accurate accelerating
field profile. Both of these behaviors can be seen in Figure 21.
Figure 21: Top Left: Poorly meshed cavity, beam axis marked in red. Top Right: Resulting
accelerating electric field. Note the non-physical roughness that results from the large
element size. Bottom Left: Cavity with improved meshing on beam axis which is marked
in red. Bottom Right: Resulting accelerating electric field. Note the improved field profile
that results. Images generated in ANSYS.
45
for the cavity element matrix from the low end to the high end of the range until the re-
quested number of modes are found. For most purposes, this range should only be as large
as needed to encompass all the variation in the single frequency anticipated. Searching for
a range that is ∼ ±10% of the desired resonant frequency is often large enough to still
find the resonance even after relatively large geometry changes. For complex structures,
this range may also include unintended trapped modes in small features like couplers or
tuners, so it is often advisable to scan a larger frequency range for modes before focusing
exclusively on the mode of interest. After launching the solution process, the solver will
often list the amount of memory needed for the solution. “In core” solutions are done
with the entire simulation resident in memory. This greatly speeds simulation time and
is the desired mode of operation. “Out of core” solutions require the solver to transfer
data between memory and the hard disk, a much slower process. This is one of the prime
reasons simulation programs should be allowed access to as much memory as possible.
46
Figure 22: Left: Magnitude of the surface electric field on a beam port cup with inadequate
meshing for surface field accuracy. Right: Identical geometry with improved meshing and
more realistic surface fields.
The peak surface fields can be found in ANSYS-APDL by making an element table of the
field in question. This table can be sorted by field magnitude and the maximum value
can be extracted. Again,
√ this is the
√ field value for the nominally low stored energy of the
simulation, so Epk / U and Bpk / U should be calculated for later scaling when the proper
stored energy is known. The remaining figures of merit require Vacc , and the bulk of the
remaining post processing will revolve around its calculation. The first step is to extract
the accelerating electric field component along the beam axis. In ANSYS-APDL, this is
done via the PATH command, specifying the beginning, end, and field component desired.
This gives the value of the accelerating field at a number of discrete points along the line
specified, and this number must be large enough for this to be a good approximation of
the actual (continuous) accelerating field profile. The first calculation is
∫ +∞ ∑
N
V0 = Eacc (z)
dz = Eacc,i
∆z (150)
−∞ i
which gives the maximal acceleration possible, V0 . The next step is calculating the acceler-
ating voltage, which includes the field’s variation in time. This calculation uses the general
form √
(∫ +∞ )2 (∫ +∞ )2
Vacc = Eacc sin (ωt)dz + Eacc cos (ωt)dz (151)
−∞ −∞
47
Figure 23: Plot of transit time factor versus particle beta.
in practical terms, although this can be simplified sometimes by the choice of (z = 0) and
cavity symmetry. For instance, a HWR with the origin placed in the geometric center of the
cavity can neglect the cos term because of symmetry. This formula depends on β, although
does not directly give the optimum β. To find this, Vacc (β) is calculated for β ranging from
0.001 to 1 by steps of 0.001. Taking the maximum of this set gives the optimum β with
enough accuracy for essentially all applications. This data can be normalized to V0 to
give the transit time factor curve, an√example of which can be seen in Figure 23. Once
the optimum β has been found, Vacc / U is calculated, giving the last of the desired field
quantities. This is used to calculate R/Q:
( )
1 Vacc 2
R/Q = × √ . (153)
ω U
To find the figures of merit for the cavity at full field, some normalization must be chosen.
For the design of the FRIB cavities, it was convention to choose a Vacc desired
√ for a cavity
and normalize to that. This determined the stored energy (U ) from Vacc / U , and therefore
the peak surface fields from the equivalent ratios already calculated. A detailed ANSYS-
48
APDL script used for all of this work can be seen in Appendix A. To this point, the basic
electromagnetic figures of merit have been calculated. Sometimes, more details or complex
problems needed to be solved, and these often can be solved in post processing. This type
of problem often involves field levels or losses on certain parts of the cavity (RF joint,
plunger, flange). These problems are solved by selecting the areas, elements, and nodes
related to the area in question, and repeating the same process used to extract the fields
or losses for the full cavity. If necessary, the losses can be scaled to a different resistivity to
simulate a small normal conducting part of the cavity and the quality factor recalculated
for the new losses.
49
Figure 24: The coupler probe can be seen in the foreground connected to the RF port
by the surface indicated by the red arrow. This is a quarter of the geometry cut on the
symmetry planes.
and ∫
ωU η
Pv = = |H|
⃗ 2 ds (156)
Qv 2 S
following the standard definition of a quality factor. The form of a TEM wave in a coaxial
transmission line gives a relatively simple form for Qext :
( )
ωU η 1
Qext = ( ) + (157)
π ln ab (aEpk )2 η(aHpk )2
where a and b are the inner and outer conductor radii for the coupler port, η = 377 Ω is
the impedance of free space, and Epk and Hpk are the peak fields on the termination for
the PMC and PEC boundary conditions, respectively.
Comparison of simulation and measured data for a prototype β = 0.53 HWR can be
seen in Figure 26. The details of measuring the Qext on a real cavity will be discussed
later. The most error-prone part of this measurement is the transformation of a given
coupler tip length to penetration into the cavity volume. While this is easy to do in a CAD
model, errors in manufacturing and positioning add significant error to this part of the
50
Figure 25: Plot of the magnitude of the magnetic field on the PEC termination of a coupler
port. Simulated with planes of symmetry in x and y.
measurement. This can be see in the data plotted in Figure 26, where the measured values
(thin black trend lines) and the simulated data (large black squares) are offset from each
other by as much as 5 mm. This can even be seen in the shift between different measurement
runs which are a similar distance apart. The slope of the measured and simulated data
agrees well, however. This is because the errors in fabricating the port diameter and
measuring the coupler probe diameter are much smaller, and can be simulated accurately.
It is important to notice that at very negative penetration into the cavity (very small
probe lengths), the trend line changes significantly from both the measured and simulated
data. This is because the approximation used in the simulation breaks down significantly
for very small probe lengths, the modes no longer have enough space to transform into a
pure TEM shape before reaching the termination plane of the RF flange. To accurately
model this behavior, a more advanced theoretical method or a traveling wave simulation
is required. These additionally require a much more accurate description of the geometry
involved, including the coupler feed through and flange shape.
51
Figure 26: Plot of measured and simulated Q-externals for a 0.53 HWR prototype. The
simulated data is represented by the large black squares. The rest are repeated rounds of
measured data.
this shift can be perturbatively estimated. Because the simulation procedure requires the
volume be broken into discrete elements, this formula requires modification to:
∆f f0 ∑ ( )
= ϵ0 Ei2 − µ0 Hi2 dAi (159)
∆T 4U
i
where ∆T is the thickness of the etched layer, Ai is the area of the external face of element
i, and the fields Ei and Hi are the fields from surface element i. For an ideal HWR, this
can be calculated directly from Slater’s theorem: −1383 Hz/µm. This large negative shift
is intuitive because etching of an ideal HWR effectively makes the cavity longer, directly
52
lowering the cavity frequency. In the process of cavity optimization, the changes to the
electric field region increase its contribution to this shift greatly, producing a less negative
coefficient. For the HWR design being developed here, the simulated etching coefficient is
−206 Hz/µm. In reality, this calculation makes assumptions that are unlikely to be true.
The most important one is that the etching is uniform over the cavity surface, but etching is
a quite complex and dynamic process, involving both fluid dynamics and thermodynamical
issues. What can be said is that, in an ideal etching setup with very good temperature
and flow control, the achieved frequency shift should approach the simulated value. Of
additional interest is using etching as a way to fine-tune the frequency of a cavity. This
process is called differential etching, etching shift versus the acid fill level can be seen in
Figure 27 for an example HWR. This process has been demonstrated and used for cavity
production for QWRS at TRIUMF [37] and elsewhere. From this plot, it can be seen that
Figure 27: Plot of simulated etching coefficient for a prototype β = 0.53 HWR filled with
acid to different levels. The blue diamonds represent simulation data, and the red squares
are this data mirrored assuming symmetry.
a much more positive frequency shift can be achieved if the cavity is filled such that the
full electric field region is etched while only one of the magnetic field regions are etched.
Experience gained from fabricating a cavity design is required to determine if such a cavity
tuning process would need to be developed.
53
8 Coupled Electromagnetic and Mechanical Simulations
A mature cavity design starts with an electromagnetic model, but must account for various
practical concerns. During normal operation, the cavity walls may be subjected to small
forces, which will result in small perturbations to the shape of the cavity. This deformation
will change the cavity vacuum space, and therefore the resonant frequency. This can
have significant controls implications and must be evaluated and optimized before cavity
production. Because matching the simulation setup to the exact conditions the cavity will
see in operation is difficult, several rounds of simulation and prototyping can sometimes be
required to achieve the accuracy and performance desired. In this section, the procedure for
this type of simulation, the types of pressures simulated, and the results of these simulations
for the HWR developed in the last section will be discussed in detail.
54
Other than this caveat, the procedure for performing the electromagnetic simulation is the
same as described earlier in this document. Once the frequency is simulated, the next step
is the setup of the mechanical simulation. This includes giving all volumes the appropriate
material properties, including the vacuum space. This may seem counter-intuitive, but
treating the vacuum space like a material is essential for accuracy. This is because this
process preserves the meshing in the vacuum space, allowing it to be perturbed by the
mechanical deformation, providing a far more accurate result than remeshing after the me-
chanical simulation. The material properties used for the various mechanical simulations
can be seen in Table 1, all of which are for cryogenic temperatures.
The remaining steps before starting the simulation are applying the pressure to be
studied and the boundary conditions. The forms of the pressures will be discussed later in
this section. The boundary conditions are of critical importance for this type of simulation.
55
Table 1: Material Properties used for Simulation.
First, any faces on symmetry planes must be constrained to stay in that plane, i.e. fixing
the variable defining the plane while letting the two other degrees of freedom float. Then
the simulation must be fixed in space. If you are using three symmetry planes, then this
is automatically done. If two or less planes of symmetry are being used, at least one part
of the cavity (a point that is on all symmetry planes used) must be fixed. While this will
not change the simulation result, it will prevent the solution from diverging if the applied
force isn’t symmetric. Additionally, other parts of the geometry can be fixed, based on
the situation you are trying to simulate. An example of this is the relationship between a
HWR and its tuner. While simulating how much force the tuning will require, the beam
ports cannot be fixed, but when doing some other simulation that assumes the tuner is in
place, fixing the beam port flanges in their normal direction is a good approximation. If
more detailed studies are required an effective spring boundary condition can be applied
to simulate, for instance, a weakly compliant tuner. After solving for the deformation, the
meshing in the material space can be discarded, and the vacuum space material should
be reset to true vacuum. The deformation should be kept by updating the coordinates
of the mesh nodes. Once the simulation is switched back to modal from structural, the
standard electromagnetic simulation procedure should be used to get the new frequency.
The difference between the first and second frequency is the frequency shift for the applied
pressure.
56
example of this would be a small cut in preparation for welding of two subassemblies that
will be obliterated by the welding process, leaving a much smoother surface. This type
of feature must be modified to accurately represent the final geometry as well as possible.
In non-sensitive areas, this type of feature can be safely removed entirely, which will give
great benefit in simulation performance with almost no sacrifice in the result’s accuracy.
The second type of geometry feature that can be changed are small details that are totally
unrelated to electromagnetic performance and have little to no impact on mechanical per-
formance. Excellent examples of this type of feature are bolt holes and chamfers on parts
of the cavity vessel. These features will exist in the finished cavity, but do not affect the
cavity vacuum and do not change the mechanical properties of the cavity assembly. These
features are often quite small as well, leading to dense local mesh that is not required.
Removing these two types of features can drastically improve simulation efficiency by re-
ducing meshing requirements, an example of this process can be seenb in Figure 29. It
is, indeed, even common for these types of features to forestall any simulation because of
the complexity of meshing them. Many of these features can include overlapping volumes
or other small/non-physical features that will need to be fixed before any work can be
done. Additionally, even if they can be meshed, the density required may be so high as
to exceed the limits of the available simulation hardware. Additionally, simplifications can
be made to allow estimates of less critical figures of merit. An excellent example of this is
the removal of bellows from a mechanical simulation. Their removal greatly simplifies the
model by removing all of the associated small/delicate features while allowing relatively
accurate simulations. This is achieved by replacing the bellows with a contact body with
an effectively equivalent spring constant, or depending on the level of deformation under
study, they can be safely removed all together. This modification must include more careful
deliberation of the consequences involved, and the increased error of this modification. Ad-
ditionally, small gaps between volumes must be removed to allow “gluing” during meshing.
57
Figure 29: Left: Flange before simplification for simulation. Bolt holes are indicated by
red arrows, fillets and chamfers by green arrows, and a complex feature (vacuum seal
knife edge) is indicated by a blue arrow. Right: Simplified flange ready for inclusion in a
simulation.
This assumption is valid under almost all cavity design considerations where the time scale
of the changes to be studied are at least five orders of magnitude longer than the RF period.
For the classical harmonic oscillator, the stored energy at any time is given by
ω 2 mx2 p2
U= + (160)
2 2m
where p and x are canonical variables. This means that the area of the ellipse described
by this equation is an adiabatic invariant. Modifying this equation into the standard form
of the ellipse
x2 p2 x2 p2
1 = 2 + 2 = 2U + (161)
a b ω2 m
2mU
we find the area of the ellipse is
U
A = πab = (162)
f
where f is the resonant frequency. Given that A is constant, we find that
U df
dU = (A)df = . (163)
f
58
Figure 30: Frequency shift versus applied pressure for a mechanical simulation. Note the
linearity down to very small frequency shifts.
Using the known form of dU from Maxwell’s Equations and assuming a finite (if small)
change in volume (∆V ), we get Slater’s Theorem:
∫
∆f ∆U 1 ( )
= = ϵ0 E 2 − µ0 H 2 dV. (164)
f0 U 4U ∆V
59
Figure 31: Left: Deformation of the inner conductor due to pressure differential. Right:
Deformation of the outer conductor due to a pressure differential (5000 Pa).
60
8.6 Lorentz Force Detuning
The fields in the cavity will interact with the induced charges and currents on the RF sur-
faces. The detuning from these forces is therefore called Lorentz Force Detuning (LFD) [41].
Again using Slater’s Theorem, we find that
∫ ∫
∆f 1 ( ) 1
= ϵ0 E − µ0 H dV = −
2 2
(P )dV (166)
f0 4U ∆V U ∆V
where P is the LFD pressure. It is worth noting that the term (P )dV is always a positive
quantity, a change in sign of the pressure P changes the direction of deformation, changing
the sign of dV . This means that any contribution to this frequency shift will be negative.
This means that the compensation techniques that could be used for correcting the pressure
sensitivity do not apply here, so we must rely on straightforward stiffening of the cavity.
The LFD frequency shift scales with the stored energy in the cavity, and therefore the field
level squared. Thus, the LFD coefficient, KL , is defined as
∆f Vacc,βopt
KL = ; Eacc = . (167)
(∆Eacc )2 βopt λ
This type of detuning is well modeled by the Duffing Equation, discussed earlier along
with the basic controls implications of this effect. Because the FRIB HWRs are to be
run with a duty factor of 1, this effect is not the dominant concern it can be for pulsed
machines [42, 43], so the use of advanced compensation methods and control systems is
not anticipated for FRIB. FRIB specifications do, however, call for |KL | to be below 3
Hz/(MV/m)2 for the HWRs. The bare cavity, as designed earlier, gives an initial KL =
−3.06 Hz/(MV/m)2 . The deformation resulting from this force can be seen in Figure 32.
61
Figure 32: Deformation of the cavity with the Lorentz Force applied at full accelerating
voltage.
as opposed to fully moved, beam port cup would be slightly less sensitive. The deformation
from the applied tuning pressure can be seen in Figure 33.
62
Figure 33: Simulation of cavity deformation for an applied tuning force.
and many others. Cavity testing measures the cumulative effect of all of these factors. In
this section, the theory and setup of a cavity test will be discussed. Additionally the data
measured during a specific HWR test will be examined along with the details of translating
this data into relevant cavity parameters.
63
(“pickup” coupler). Power is coupled to the cavity with the input coupler, while the pickup
coupler is used strictly as a diagnostic tool. Cavity tests are done with a phase lock loop
matching the driving frequency to the cavity resonance. The basic testing phase lock loop
arrangement can be seen in Figures 37 and 38.
This RF circuit allows a signal generator to track the cavity resonant frequency by
mixing the transmitted power signal with the forward power signal, giving a DC signal
whose amplitude is based on the difference between the driving frequency and the cavity
resonant frequency. With the cavity driven in steady-state on resonance, only the cavity
frequency (f ), forward power (Pf ), reverse power (Pr , vector sum of the reflected and
emitted traveling waves from the input coupler), and the transmitted power (Pt , from the
pickup coupler) can be measured. Using equations from earlier
Pr
|S11 |2 = (168)
Pf
Pt
|S21 |2 = (169)
Pf
|S21 |2 Q0 Pt
β2 = = = (170)
1 − |S11 | − |S21 |
2 2 Qext,2 Pd
1 ± |S11 | Q0 Pr
β1 = = = (171)
1 ∓ |S11 | Qext,1 Pd
and
Q0 = QL (1 + β1 + β2 ) (172)
we see that we don’t have enough information to calculate Q0 and U . To do this, we must
determine the QL and the coupling (whether β1 is greater or less than 1). Both of these
can be obtained from modulating the forward power. The most straightforward method is
to suddenly shut off the forward power, so that the equation for the stored energy in the
cavity becomes
dU ωU
= −PL = − (173)
dt QL
which gives ( )
ωt
U (t) = U0 exp − . (174)
QL
For a sufficiently large QL (true for most Dewar tests), a decay time τL can be measured
for a given U/U0 (usually about a −6 dB fall), giving QL . Additionally, the transitory
response of Pr determines whether the cavity is over or under coupled. This effect can
be seen in Figure 34. In this way, we have enough information to calculate not only β1
and β2 , but also Q0 . Given this, we can calculate Qext,2 , allowing us to transform our
measurement of Pt into a measurement of U . While it is often assumed that Qext,1 and
64
Figure 34: Transient behavior of the reverse power for different coupling strengths. All
curves are in response to the shown forward power behavior. Image credit to Tom Powers,
TJNAF.
Qext,2 do not change dramatically over the course of a Dewar test, it is recommended to
periodically repeat the modulated measurements to ensure these values have not drifted
during testing. With knowledge of Qext,2 , every set of power measurements can be used to
directly calculate Q0 and U :
Qext,2
Pd = Pf − Pr − Pt ; U= Pt (175)
ω
where the first equation is the cavity energy balance, and
ωU
Q0 = . (176)
Pd
This method, known as the “direct” method, has the disadvantage of using Pr extensively.
Especially far from unity coupling, Pf − Pr is likely to be quite small, and error prone.
Fortunately, there is another method for calculating Q0 and U . While the first method
does not require the use of Qext,1 , this method uses this previously measured number in
place of Pr . This method is known as the “indirect” method, because, while measuring
Qext,1 involves Pr , this usage makes it less sensitive to the uncertainty of its measurement.
Plugging the definition of Qext,1 into our equations, and making the usual assumption that
β2 ≪ 1, we find that
Qext,1
Q0 ≈ √ . (177)
Qext,1 Pf
2 Qext,2 · Pt − 1
65
It should be noted that once Qext,2 and Qext,1 are calculated with modulated measurements,
they are assumed to be constant for following measurements. This, however, makes each
CW measurement over-specified, and this can be used to check the measurement’s accuracy.
In principle, both the direct and indirect methods should give the same answer. Starting
from
Qext,2 β1
= . (178)
Qext,1 β2
it is straight-forward to derive (assuming, as usual, that Pt is small compared to Pf )
√
Qext,2
|S21 | = 1 ± |S11 |. (179)
Qext,1
Plotting the left side of this equation versus the right side of the equation using the data
(|S21 | and |S11 |) and the assumed values for Qext,2 and Qext,1 gives an easy graphical
method of checking the consistancy of the measured valueds. Depending on the coupling,
|S11 | will vary from 1 to −1, so the plotted data should, ideally, lay on a simple triangle.
Thus, the plot of these quantities is an excellent way to test whether the values of Qext,1 ,
Qext,2 , Pr , Pf , and Pt are consistent. Any systematic deviation between the plotted data
and this “Duality Triangle” indicates a problem with the data. This generally happens
when the assumption of constant coupling is invalid, which is the case for the test presented
in Section 4.4, a dramatic shift in coupling between test day 1 and test day 2 had to be
accounted for.
66
It is important to accurately correlate this measurement with a set of CW measurements
taken immediately after the modulated measurement. Combining these modulated and
CW measurements allows the calculation of Qext,1 and Qext,2 as detailed in the previous
subsection, and these are used to interpret the following CW measurements.
The second phase of measurement is to perform CW measurements at different Pf . The
forward power is increased in steps of 1 − 3 dB, and several CW data points are recorded at
every step, after adjusting the phase shift. The higher the cavity fields, the smaller the steps
in Pf should be, to get the best resolution in this region. This is because the most dynamic
behavior happens here, and sudden shifts in Q0 can occur. This is generally continued until
limited either by cavity quench, maximum available Pf , or high radiation levels from field
emission. It should be noted that many values are recorded with each data point, not
just power levels. This includes data such as date and time, readings from temperature
and pressure sensors, liquid helium level, and more. Any interesting phenomena are then
investigated, time permitting, such as the detailed behavior of sudden drops in Q0 , field
emission onset levels, quench field level, etc. If desired, time can be spent conditioning
the cavity at high fields in an attempt to improve performance. This is most generally
done when there is high radiation levels in an attempt to destroy or otherwise shut off the
serious field emission sites causing the radiation. In cases of high radiation, staying at high
power or pulsing to even higher power can be successful in improving performance, allowing
access to higher field levels. Once all of the desired data are collected for a given cavity
configuration, then the configuration is changed and the process is repeated. The nature
of the cavity configurations depends strongly on the exact phenomenon being studied, but
a common example of this would be measurement at different temperatures as the cavity
is pumped down to 2 K.
67
Figure 35: Basic RF circuit for a phase lock loop for cavity testing with the calibration
points marked.
must be calibrated while the cavity is cold because of the significant thermal shifts. This
is generally measured directly after the completion of a cavity test, while the circuit is as
thermally stable as possible. This is done by connecting Cal Point A to Port 1 of a VNA
though a calibrated cable, then measuring the round-trip losses (S11 ), repeating for Cal
Point D. These values are called the “Cold Calibration”. The rest of the process is called
the “Warm Calibration” because it involves measurement of the rest of the RF circuit,
which is at room temperature. The first step for this is calibration of the Pt and Pr power
sensors. Because we have already measured the losses inside the Dewar, all that remains is
measuring the losses from Cal Point D to Cal Point E for Pt and from Cal Point A to Cal
Point F for Pr . These losses will include the cable losses as well as the attenuation from the
directional couplers used to sample these signals. This measurement is done by connecting
the Dewar end to Port 1 of the VNA and the power sensor end to Port 2 of the VNA, and
measuring the losses (S21). The final calibration for these power sensors is the warm losses
plus half of the appropriate round-trip cold losses. The most complex calibration is for
Pf . This is because what we desire is the forward power at Cal Point B (again, what the
cavity actually sees) as measured at Cal Point G for a given power at Cal Point H (right
68
out of the amplifier). The first step accounts for the losses from the amplifier to the cavity
by connecting Cal Point H to Port 1 of the VNA and Cal Port A to Port 2 of the VNA,
then calibrating out the through losses. Port 2 of the VNA is then moved without changing
the calibration to Cal Point G while terminating Cal Point A with a 50 Ω load, where the
power loss (S21) is measured. The power measured, though, is the forward power at Cal
Point A seen at Cal Point G. This means that, for the final calibration of Pf , the warm
calibration number must have half the appropriate cold calibration subtracted to properly
reflect the forward power at the cavity.
10.1 Multipacting
The name multipacting is a compound word used to shorten “multiple impacting”, and this
name is a good description of this phenomenon. Stray electrons in the cavity vacuum space
interact with the cavity fields and strike the cavity walls. The impact can generate more
electrons emitted from the surface, which will also have some trajectory based on their
energy, angle of emission, and RF phase. The number of electrons emitted is quantified by
the Secondary Emission Yield (SEY), which depends on many factors including electron
energy, surface preparation, and impact angle. The electron’s trajectory also extracts
energy from the cavity, depositing it into the cavity walls. Based on the surface it strikes,
the energy of impact, and the angle of impact, the average number of ejected electrons
varies, and can be over 1.4 [46], even in a well processed cavity. The energy this cascade
extracts is trivial, in principle, until it is sustained for many generations. If the trajectories
of the electrons becomes closed, i.e. return to near the original emission point (and RF
phase), the cascade continues, eventually consuming all additional RF power injected into
the cavity.
69
by regions of parallel surfaces, especially items like couplers, plungers, and stub tuners.
The strongest multipacting occurs when there is a wide range of trajectories in location,
energy, and RF phase that satisfy the resonant condition. These barriers are called “hard”
barriers because they are extremely difficult to get beyond in testing/operation. Particle
tracking simulations [46, 47] for predicting multipacting have been increasing in accuracy
in recent years, and have become reliable for predicting and avoiding strong multipacting
barriers in cavity designs. However, because of the perfection of the models used, these
simulations often find far more barriers than will present themselves in practice. Fabrication
errors are often beneficial, in this case. The random errors added to the surface often
break the resonant condition for the weaker multipacting bands, although this is not a
strong enough benefit to loosen manufacturing tolerances. Additionally, careful surface
preparation during cavity fabrication and processing can reduce the SEY, automatically
narrowing multipacting bands. For HWRs of this geometrical type, multipacting studies
have been done [46] and found no significant barriers (besides the rinse port plungers,
which were later removed from the design). This has been born out in testing where no
significant multipacting has been encountered.
70
of electrons, usually particulate (“field emitters”) on the cavity surface not removed by
processing. These field emitters locally enhance the electric field, causing a stream of
electrons to flow into the cavity at high fields. Because of this behavior, field emission
is a problem at high fields, unlike multipacting. At low fields, the fields on the emitters
aren’t enough to generate large numbers of electrons, and they don’t gain enough energy
from their trajectory to generate significant numbers of X-Rays. As the fields in the cavity
increase, both the number of electrons and the energy they gain increase until cavity
performance begins to suffer. The energy extracted by the electrons during their transit
can significantly reduce the cavity quality factor, and can limit achievable fields by requiring
more RF power, both from extracted energy and increased coupler mismatch to the cavity.
If the field emission is strong enough, the energy deposited by the electrons on the cavity
wall can be enough to cause a local thermal quench. Once the superconductivity is broken
locally, the cavity stored energy is rapidly dissipated there, turning the field emission off.
Field in the cavity can be restored once the wall has cooled and become superconducting
again. There are several common ways of mitigating this problem. During processing,
high-pressure rinsing with ultra-pure water forceably removes larger particulate from the
surface, significantly reducing the sources for electrons at high fields, while clean room
preparation and assembly reduces reintroduction of these particulate after cleaning. During
operation, the field emitters can be used to process themselves. By operating close to the
limit of operation (limited by available power or radiation limits), often pulsing to higher
fields as well, enough power can sometimes be introduced to strong emitters to cause
their destruction. This process works best for small numbers of strong emitters, and has
obvious limits. While strong emitters can be suppressed by careful fabrication, processing,
and testing, stray electrons can never be eliminated from the cavity. With a non-zero
population of electrons in the cavity from things such as residual gas, cosmic rays, and
weak field emitters, it is common to see minor field emission at very high fields. Care must
be taken during design of testing facilities to include adequate radiation shielding to allow
testing of even the worst cavities.
10.5 Q-Disease
Hydrogen is introduced into the cavity material at many points along the cavity processing.
The strongest source is etching, especially if the temperature of the acid blend used isn’t well
controlled. The hydrogen itself is not thought to degrade cavity performance. However,
between 40 K and 120 K [48], niobium reacts with hydrogen to form several kinds of
hydrides on the surface of the cavity. These hydrides are dielectric, and extremely lossy.
In operation, the quality factor of the cavity is severely reduced, and degrades further
as the field is increased. The formation process of these hydrides is relatively slow; if
the dangerous temperature region is passed through in under an hour [48], little to no
degradation is seen. This results in interesting requirements for the design of cryomodule
cryogenic systems, a balance between efficient use of cryogenics and rapid cooling of the
71
cavities through the dangerous temperature region. Additionally, a degassing bake can
be used to drive off the hydrogen from the material, significantly reducing the danger of
Q-Disease. This bake is generally done at 600◦ C for roughly 10 hours in high vacuum.
Once observed in testing or operation, there is no known remedy for Q-Disease besides a
warming to 200 K and more rapid cooling cycle.
72
in this section from Appendix A:
!Calculates Volume parameters
vsum
!Select volume 1
vsel,s,,,1
!Define cavity vacuum as "cavity"
cm,cavity,volu
!Select "cavity"
cmsel,s,cavity
!Select areas associated with cavity vacuum
aslv,s
!Unselect symmetry planes
asel,u,loc,x,0
!asel,u,loc,y,0
asel,u,loc,z,0
!asel,u,,,4
!Define PEC walls
cm,cavwall,area
aplot
This section of code defines two objects for use later in the script: the cavity vacuum volume
cavity and the cavity RF surfaces cavwall. These references are used extensively through
the rest of the scripts to set boundary conditions and for post-processing. It is important to
note the extreme flexibility of the code used to do this. For essentially any cavity type, this
code will work properly if the correct volume number is used in line 4 and the appropriate
symmetry planes are selected in lines 12-14, an almost trivial amount of work. This is
just one example of this scripting philosophy that can be seen in the Appendices, and
it allows rapid generation of the appropriate scripts for a wide variety of cavity designs,
greatly speeding simulation. This means that the time required for parameter scans used
to optimize cavity performance can be reduced to the time it takes for the computation.
Proper scripting of the meshing can also reliably reduce this time as well. Some code
systems like ACE3P are being designed to be fully scriptable, from geometry creation to
post-processing, using high-power computing clusters to speed simulation time. This suite
of software shows great promise for allowing rapid simulation and optimization of even the
most complex geometries. Simulation software that allows scripting and optimization like
this is understandably preferable for serious optimization projects.
References
[1] G. Stupakov, “Lecture notes on Classical Mechanics and Electromagnetism in Accel-
erator Physics”, United States Particle Accelerator School, Revision 20, 2011
73
Figure 36: Both days of testing data plotted on the Duality Triangle to verify self consis-
tency.
[2] B. Van der Pol, J. Van der Mark, “Frequency demultiplication”, Nature, 120, 363-364,
(1927)
[3] J.D. Jackson (1999), “Classical Electrodynamics, 3rd Edition”, John Wiley and Sons
Inc.
[5] J.D. Jackson (1999), “Classical Electrodynamics, 3rd Edition”, John Wiley and Sons
Inc., p. 298
74
[9] M. Tinkham, “Introduction to Superconductivity, 2nd Edition”, Dover Publications,
Inc., Chapter 3
[10] J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev. 108, 1175 (1957)
[13] A. Facco, “SRF Low-beta Accelerating Cavities for FRIB”, Lecture notes, Phys 905 -
The Accelerator Physics of FRIB, Michigan State University, October 2011
[24] C. Piel et. al., “Phase 1 Commissioning Status of the 40 MeV Proton/Deuteron Ac-
celerator SARAF”, Proceedings of EPAC08, Genoa, Italy, THPP038
[26] N. Ida, J.P.A. Bastos, “Electromagnetics and calculation of fields”, Second Edition,
Springer Verlag, N.Y., 1996
75
[27] S. Humphries, “Finite-Element Methods for Electromagnetics”,
www.fieldp.com/femethods.html
[35] W. Hartung, “LEP 1500 Plans Collapse and Other Stories or What I Learned on my
CERN Vacation,” Report NS/RF-92-1701, National Superconductor, 160000 Lincoln
Avenue, Brentwood, Colorado 80523 (1992).
[36] W. Hartung and E. Haebel, “In Search of Trapped Modes in the Single-Cell Cavity
Prototype for CESR-B,” Proceedings of the 1993 Particle Accelerator Conference,
Washington, D.C., May 1993, Vol. 2, pp. 898-900.
[37] V. Zvyagintsev, et. al., “SCRF Development at TRIUMF”, RUPAC 2010, Protvino,
Russia
[38] H. Padamsee (2009), “RF Superconductivity: Volume II: Science, Technology, and
Applications”, John Wiley and Sons Inc.
[40] J.C. Slater, “Microwave Electronics”, D. Van Nostrand Company, New York, 1950,
pp. 80-81.
[42] M. Grecki et. al., “Compensation of Lorentz Force Detuning for SC Linacs (with Piezo
Tuners)”, Proceedings of EPAC08, Genoa, Italy, MOPP129
[43] R. Carcagno et. al., “First Fermilab Results of SRF Cavity Lorentz Force Detuning
Compensation Using a Piezo Tuner”, Proceedings of SRF2007, Peking University,
Beijing, China, TUP57
76
[44] H. Padamsee, J. Knobloch, T. Hayes (2008), “RF Superconductivity for Accelerators,
2nd Edition”, John Wiley and Sons Inc., p. 120
[45] T. Powers, “Theory and Practice of Cavity RF Test Systems”, Technical Document,
Thomas Jefferson National Accelerator Facility
[46] L. Ge et. al., “Multipacting Simulation and Analysis for the FRIB Superconducting
Resonators Using TRACK-3P”, Proceeding of LINAC10, Tsukuba, Japan, September
12 - 17, 2010, THP092
[47] R.L. Geng, “Multipacting Simulations for Superconducting Cavities and RF Coupler
Waveguides”, Proceedings of the 2003 Particle Accelerator Conference, Portland, OR,
May 12-16, 2003
[48] K. Saito, “Cavity Preparation”, Lecture notes from ILC School 2009
77
Figure 37: One half of the RF circuit for a phase lock loop for cavity testing (circulator,
isolators, and RF diodes excluded). Letters in Red connect to the other half of the RF
circuit, seen on the next page.
78
Figure 38: One half of the RF circuit for a 79
phase lock loop for cavity testing (circulator,
isolators, and RF diodes excluded). Letters in Red connect to the other half of the RF
circuit, seen on the previous page.