You are on page 1of 20

Mode Structure in Superconducting Metamaterial Transmission Line Resonators

H. Wang,1 A.P. Zhuravel,2 S. Indrajeet,1 B.G. Taketani,3, 4 M.D. Hutchings,1, ∗ Y. Hao,1, †


F. Rouxinol,1, ‡ F.K. Wilhelm,4 M.D. LaHaye,1 A.V. Ustinov,5, 6 and B.L.T. Plourde1, §
1
Department of Physics, Syracuse University, Syracuse, NY 13244-1130
2
B. Verkin Institute for Low Temperature Physics & Engineering of
National Academy of Science of Ukraine, UA-61103 Kharkiv, Ukraine
3
Departamento de Fı́sica, Universidade Federal de Santa Catarina, 88040-900 Florianópolis, SC, Brazil
4
Theoretical Physics, Saarland University, Campus, 66123 Saarbrücken, Germany
5
Physikalisches Institut, Karlsruhe Institute of Technology, 76131 Karlsruhe, Germany
6
Russian Quantum Center, National University of Science and Technology MISIS, 119049, Moscow, Russia
arXiv:1812.02579v3 [physics.app-ph] 31 May 2019

(Dated: June 4, 2019)


Superconducting metamaterials are a promising resource for quantum information science. In the
context of circuit QED, they provide a means to engineer on-chip, novel dispersion relations and a
band structure that could ultimately be utilized for generating complex entangled states of quantum
circuitry, for quantum reservoir engineering, and as an element for quantum simulation architectures.
Here we report on the development and measurement at millikelvin temperatures of a particular type
of circuit metamaterial resonator composed of planar superconducting lumped-element reactances
in the form of a discrete left-handed transmission line (LHTL) that is compatible with circuit
QED architectures. We discuss the details of the design, fabrication, and circuit properties of this
system. As well, we provide an extensive characterization of the dense mode spectrum in these
metamaterial resonators, which we conducted using both microwave transmission measurements and
laser scanning microscopy (LSM). Results are observed to be in good quantitative agreement with
numerical simulations and also an analytical model based upon current-voltage relationships for a
discrete transmission line. In particular, we demonstrate that the metamaterial mode frequencies,
spatial profiles of current and charge densities, and damping due to external loading can be readily
modeled and understood, making this system a promising tool for future use in quantum circuit
applications and for studies of complex quantum systems.

I. INTRODUCTION processing and quantum optics in the microwave regime [5–


8].
Engineered quantum systems present a unique oppor- Recent efforts have been undertaken to achieve mul-
tunity to study many-body phenomena in a controlled timode strong coupling in cQED systems, wherein a su-
way through analog quantum simulation [1]. An intrigu- perconducting qubit is able to couple strongly to mul-
ing direction in this area involves the strong coupling tiple photonic modes. Efforts to engineer a spectrally
of atoms – artificial or natural – to a quasicontinuum dense mode structure include the use of electrically long
of modes, as is the case with quantum impurity models (approximately1 m) superconducting transmission lines
or the investigation of scattering in strongly interacting [9], arrays of coupled transmission-line resonators or
systems [2]. cQED systems [10–15], and frequency combs implemented
In the microwave regime, the field of circuit QED through a pumped superconducting nonlinearity [16].
(cQED) involves one or more artificial nonlinear few- An alternative approach, which we explore in a cQED
level systems (qubits) that are coupled to linear res- architecture here, utilizes a periodic array of supercon-
onant modes, often formed from standing waves on a ducting lumped-element reactances to engineer a one-
finite-length microwave transmission line [3] or a three- dimensional metamaterial in the microwave range [17, 18].
dimensional waveguide cavity [4]. This platform permits This metamaterial is characterized by a photonic bandgap
the attractive feature of reaching strong coupling between at low frequencies, with a band characterized by a left-
the qubit and individual photonic modes in the resonator handed dispersion relation [19–22] and a dense set of
due to both the large transition dipole of the qubit and modes at frequencies just above the bandgap, commen-
the small mode volume of the resonator, which is now surate with superconducting qubit transition energies.
utilized regularly for experiments in quantum-information A variety of superconducting metamaterials have been
studied previously [23–27], including some systems with
applications in microwave quantum optics [28, 29] and
quantum-limited amplification of microwave signals [30–
∗ Present address: SeeQC, Inc., Suite 141, 175 Clearbrook Road, 32]. In our present work, the dense mode spectrum we
Elmsford, NY 10523, USA
† Present address: SiTime Corporation 5451 Patrick Henry Drive,
implement is appealing for future applications for archi-
Santa Clara, CA 95054 USA l tectures of quantum simulators and entanglement gener-
‡ Present address: “Gleb Wataghin” Institute of Physics, University ation [17], whereby the qubits could be tuned between
of Campinas UNICAMP 13083-859, Campinas, SP, Brazil the low-frequency bandgap and the left-handed region
§ bplourde@syr.edu of the spectrum just above the bandgap. However, in
2

contrast to earlier work on multimode cQED systems, cutoff frequency below which the structure does not sup-
which involved physically long cavities or coupled dis- port propagating waves. As the wavenumber is decreased
tributed transmission lines, our use of lumped-element the frequency diverges, which is clearly unphysical as this
components provides a more compact design that can implies that the group velocity diverges in the limit of
be readily interfaced with superconducting qubits, and k → 0.
thus is inherently scalable for applications in quantum Such behavior would not occur in any practical trans-
information and communication. mission line due to the stray reactances in the constituent
Here we present a characterization of the mode struc- elements for an LHTL, as described in Refs. [22, 36, 37].
ture in superconducting metamaterial resonators in a Each capacitor Cl will have a stray series inductance

cQED architecture that is compatible with superconduct- Lr giving it a self-resonance frequency ωC = 1/ Lr Cl .
ing qubit integration through low-temperature measure- Similarly, each inductor Ll will have a stray shunt capac-
ments of transmission spectra, imaging of microwave fields itance C√ r resulting in another self-resonance frequency
with a laser-scanning microscope (LSM), as well as nu- ωL = 1/ Ll Cr . Figure 1(b) shows a schematic of a com-
merical simulations. Our results lay the foundations for posite LHTL including these stray reactances. As first
future applications of this system for quantum simula- described in Ref. [36], a Bloch-Floquet analysis can be ap-
tions and the design of superconducting resonator spectra. plied to such a composite LHTL to obtain the dispersion
In Sec. II we discuss the general structure of 1D meta- relation between wavenumber k and frequency ω. We
material transmission lines and resonators. We describe include our own derivation in Appendix A and obtain the
the fabrication and microwave characterization of our following expression [22, 37]:
superconducting metamaterial resonators in Sec. III. In
Sec. IV we present our investigation of the mode struc- 1
k(ω) = ×
ture through LSM imaging of the microwave fields, which ∆x    
we analyze in detail in Sec. V. We present numerical 1 1 1
cos−1 1 − ωLr − ωCr − .
simulations of our metamaterial design in Sec. VI. In 2 ωCl ωLl
Sec. VII we present an analysis of losses in superconduct- (2)
ing metamaterial resonators. Section VIII contains our
conclusions and possible future directions for experiments
with superconducting metamaterial resonators. Δx
(a)
Cl Cl Cl Cl Cl
II. 1D METAMATERIAL TRANSMISSION Ll Ll Ll Ll Ll
LINES

The basic properties of 1D left-handed transmission Δx


lines (LHTL) have been treated previously by several (b) Cl Lr Cl Lr Cl Lr
others [17, 19, 22, 33, 34], thus, we only review these
briefly. The simplest circuit model for a LHTL consists Cr Ll Cr Ll Cr Ll
of a chain of capacitors Cl and inductors Ll in a similar
manner to the discrete implementation of a conventional (c)
right-handed transmission line that would be described 80
by the telegrapher’s equation [35]. However, the LHTL
has the positions of the inductors and capacitors swapped
with respect to the right-handed line [Fig. 1(a)]. This 60
max(ωC, ωL)/2π
ω/2π (GHz)

rearrangement has a profound effect on the transmission


and dispersion properties of the transmission line. 40
The dispersion relation for such an LHTL is given by min(ωC, ωL)/2π

1 1
ω(k) = √ , (1) 20
2 Ll Cl sin (k∆x/2) ωIR /2π
where k is the wavevector in the LHTL and ∆x is the unit 0
cell length [17]. This expression is plotted in Fig. 1(c) and 0.0 0.2 0.4 0.6 0.8 1.0
the curve clearly exhibits a falling dispersive behavior, kΔx/π
characteristic of a left-handed medium. At the largest
wave numbers in the limit of k∆x → π, or one wavelength FIG. 1. Metamaterial transmission line schematic: (a) ideal
for every two unit cells, corresponding to the lowest fre- LHTL; (b) composite LHTL including stray reactances; (c)
quencies of the system, the band becomes flat approaching calculated dispersion relations ω(k) for ideal LHTL (dashed
the edge of the Brillouin zone. The corresponding fre- blue line) [Eq. (1)] and composite LHTL (solid black line)
√ [Eq. (2)] using circuit parameters described in the text.
quency, ωIR = 1/2 Ll Cl , defines a high-pass infrared
3

Figure 1(c) includes a plot of ω(k) obtained by inverting


Eq. (2). Similar to the corresponding curve from Eq. (1), Δx
(a) Cl
the band becomes flat as k∆x → π at ωIR /2π, which Cc Cl Lr Cl Lr Lr Cc
is chosen to be 5 GHz for this plot. However, for small
k∆x the curve does not diverge and instead turns over Cr Ll C r Ll Cr Ll
and intersects k∆x = 0 at the lower of the two stray
self-resonance frequencies, min(ωC , ωL ). Nonetheless, the
curve still has a left-handed nature over the entire range (b)
of k∆x. At the upper frequency end of the plot, it is 40
clear that there is a second band corresponding to another
solution of Eq. (2). This upper band begins at the higher 30
of the two self-resonance frequencies, max(ωC , ωL ), and

ωn/2π (GHz)
increases with k, characteristic of right-handed dispersion.
Between ωC and ωL there is a gap where there are no 20
solutions to Eq. (2) and thus no wave propagation along
the transmission line. For this plot, estimates of likely
self-resonance values in physically realizable structures 10
are used: ωC /2π = 40 GHz, ωL /2π = 50 GHz. For the
remainder of the discussion we focus on the left-handed 0
branch, as the devices we study put the high-frequency 0.0 0.2 0.4 0.6 0.8 1.0
gap and the right-handed branch well beyond the fre- knΔx/π
quency range that we are able to access in our measure-
ments. However, we note that with future devices we FIG. 2. (a) Schematic of a composite LHTL resonator with
stray reactances including coupling capacitances Cc at each
may be able to enhance the stray reactances in order to
end. (b) Plot of mode frequencies vs. wave number computed
lower the frequency of the gap and bring the right-handed for N − 1 = 41 modes in a LHTL (ωIR /2π=5 GHz, ωC /2π =
branch into our measurement range. 40 GHz, ωL /2π = 50 GHz). The solid line is the disperson
In order to generate resonances with the transmission relation obtained from Eq. (2).
line, one can impose either open-ended or short-circuited
boundary conditions at two positions along the length of
the line. We choose to focus on the case with open-ended spectrum just above ωIR , we aim for ωIR /2π ∼ 4 − 7 GHz.
boundary conditions, using input and output coupling If we target a characteristic impedance near 50 Ω, this
capacitors Cc at either end. As with a conventional right- requires Cl = 250 fF and Ll = 0.625 nH. These pa-
handed transmission line, the system will exhibit standing- rameter values can be achieved with superconducting
wave resonances given by the standard relation kn l = nπ thin-film structures consisting of interdigitated capacitors
for integer n, where l is the length of the transmission line and meander-line inductors with a single-layer fabrica-
between the two ends. With a discrete LHTL including tion process that can easily be extended to incorporate
stray reactances with N unit cells, the standing-wave superconducting qubits in future devices. Based on nu-
condition becomes kn ∆x = nπ/N for n = 0 : N − 1 merical simulations of these structures with Sonnet [39],
[21]. Thus, if we disregard the zeroth-order mode (n = 0) we expect the stray reactances Lr and Cr to result in
[21], there areN − 1 modes for frequencies between ωIR self-resonant frequencies ωC /2π and ωL /2π of 40 GHz or
and min(ωC , ωL ). Due to the shape of the dispersion higher, well beyond the frequency range of our measure-
relation given by Eq. (2), the lower frequency and higher ments. Thus, we expect our metamaterial resonators to
n modes areclosely spaced in frequency with the mode exhibit a left-handed dispersion over the entire range of
spacing increasing as one moves towards higher frequency our measurements. While some of our initial devices have
and lower n. This behavior is illustrated in the plot of been fabricated from Al thin films on Si [40], we fabricate
Fig. 2(b), which shows ωn (kn ) points for each of the N − 1 the metamaterial resonators described in this work using
left-handed modes computed for a circuit with the same 90-nm-thick Nb films on Si. The higher Tc of the Nb (ap-
parameters as that considered in Fig. 1. proximately 9 K) makes the structures compatible with
the typical operating temperature of our microwave imag-
ing technique, which is described in Sec. IV. The features
III. FABRICATION AND are defined using photolithography with a standard liftoff
CHARACTERIZATION OF METAMATERIAL process following the Nb sputter deposition.
RESONATORS The meander-line inductors have a 2-µm linewidth and
17 turns [Fig. 3(c)]. The interdigitated capacitors have 26
In order to implement metamaterial resonators that are pairs of 50-µm-long, 4-µm-wide fingers spaced by 1 µm
compatible with the integration of superconducting trans- [Figs. 3(e, f)]. The metamaterial transmission line runs
mon qubits [38] that can be tuned through resonance with across most of the width of the chip, which has an active
the various modes in the dense region of the metamaterial area of 4 × 4 mm2 [Fig. 3(a)] when mounted in one of our
4

(b) (c) (d) baseline transmission curve with no sample present is


used to normalize the S21 (f ) spectra.
CC
Figure 4 contains a plot of our measured S21 (f ) spec-
tra for one of our devices at two different temperatures:
100μm 60μm 20μm 65 mK and 3 K. Below 4.2 GHz, the transmission is below
-40 dB, characteristic of the IR cutoff expected for a LHTL.
(e) For frequencies above this cutoff, we observe densely
(a)
spaced narrow resonances that become increasingly far
apart with broader linewidths at higher frequencies. The
frequency placement of the resonances are discussed in
60μm Sec. V and the coupling-limited linewidths are analyzed
in Sec. VII. When the sample temperature is raised to
(f) 3 K, approaching the operating temperature range for the
LSM imaging, the resonances are still present, although
shifted slightly lower in frequency due to the increased
1mm
kinetic inductance, with reduced quality factors because
20μm of the presence of thermal quasiparticles.

FIG. 3. Optical micrographs of metamaterial resonator device: 0


(a) zoomed-out image of entire chip, (b) input coupling capac-
itor Cc and the first few unit cells, (c) meander-line inductor -20
of the first unit cell, (d) detail of input coupling capacitor and
connection between inductor and capacitor of the first unit
cell, (e) interdigitated capacitors in several unit cells, (f) detail |S21| (dB) -40
of interdigitated capacitor. 0

|S21| (dB)
-60
-20
-40
standard microwave chip holders. Based on the length of -80
the interdigitated capacitor fingers, this allows us to fit 5.19 5.20 5.21
f (GHz)
N = 42 unit cells on a chip. The ends of the metamaterial -100
2 4 6 8 10 12 14 16
resonator are defined with coupling-capacitor gaps at the Frequency (GHz)
first and last cells [Fig. 3(d)] to conventional coplanar
waveguide (CPW) traces carrying the output (input) sig- FIG. 4. Measurements of the magnitude of the microwave
nals to (from) launcher pads at the opposite corners of transmission |S21 (f )| on the ADR at two different tempera-
the chip. We estimate the Cc values for these structures tures: 65 mK (solid black line); 3 K (blue dashed line). The
to be 30 fF based on HFSS Q3D simulations [41]. inset shows an enlarged plot in the vicinity of n = 38 mode
near 5.21 GHz.
Because of the spatial extent of the meander-line induc-
tors, we choose to stagger the placement of the inductors
so that alternating inductors connect to the ground plane
either upwards or downwards from the path of the trans-
mission line [Fig. 3(a)]. Besides allowing the inductors to IV. LSM IMAGING OF MODE STRUCTURE
fit in the available space, this configuration also avoids
the asymmetry in the grounding that would occur if all of Low-temperature laser-scanning microscopy (LSM) has
the inductors extended to ground on only one side of the been used previously to image a wide variety of super-
transmission line. However, as we show later, the trade-off conducting structures under rf excitation [42, 43]. The
of using the staggered inductor arrangement results in a various modes of LSM operation have been described in
distortion of the dispersion relation from that described detail in Refs. [44] and [45]. The basic principle of opera-
by Eq. (2) for the lowest frequency modes. tion of the LSM is based on the local deposition of laser
For initial measurements of the mode spectrum, we cool energy onto a superconducting sample followed by the
the device on an adiabatic demagnetization refrigerator measurement of some global response of the sample, for
(ADR) with a base temperature of 65 mK. The device example S21 , to the perturbation at that particular loca-
is shielded from stray magnetic fields with a cryogenic tion. By raster scanning the laser probe over the surface
µ-metal can and microwave signals are delivered to the of the sample while modulating the beam intensity, an
metamaterial through a coaxial line with 74 dB of cold at- image of the photoresponse R(x, y, f ) can be produced
tenuation; the output signal is amplified with a cryogenic using a lock-in technique by correlating the output signal
HEMT plus a room-temperature low-noise microwave am- with the location of the laser spot.
plifier. We use a vector network analyzer to measure the For the metamaterial imaging reported here, the chip
microwave transmission S21 (f ). A separately measured is mounted inside a vacuum volume of a cryostat with
5

optical access and the temperature of the chip is measured (a) 7.79 GHz
to be about 5 K during the imaging. The laser probe
is produced by a diode laser with wavelength 640 nm
that is focused to a spot of about 12-µm diameter with
a power of approximately 10 µW at the sample, result-
ing in an increase of local temperature at this location. (b)
The laser intensity is modulated at 100 kHz while the
laser-induced modulation of the microwave transmission
|∆S21 (f )| is measured with cryogenic coaxial cables car-
rying microwave signals to and from the sample. The 3.34 GHz n=40
output signal is amplified by 60 dB, rectified with a crystal
diode, then lock-in detected at the modulation frequency (c)
to generate the photoresponse signal. The maximum fre-
quency accessible on our LSM is 20 GHz, similar to the
measurement electronics for the experiments on our ADR.
n=40 charge
(d)
(a) -80
(b)
|S21|(dB)

-90
-100
-110 n=40 current
7.0 7.2 7.4 7.6 7.8 8.0
PR(y0,f) (mV)

0.8 (c) FIG. 6. (LSM photoresponse compared with Sonnet current


0.6 and charge density simulation: (a) Reflectivity map from LSM
0.4 showing details of metamaterial layout in the imaged area.
0.2 (b) 2D distribution of LSM photoresponse R(x, y, 3.44 GHz)
0.0 for the n = 40 mode. Bright (dark) regions correspond to
7.0 7.2 7.4 7.6 7.8 8.0 large (small) PR signal. Sonnet simulations of the n = 40
Frequency (GHz) mode (c) charge density and (d) current density.
LSM PR (d)
max
x
f We then perform 2D scans of R(x, y, f ) while exciting
one of the metamaterial modes. In Fig. 6(b), we present
0 such a 2D scan for the n = 40 mode at 3.34 GHz, which
6.9 Frequency (GHz) 8.0 is the next-to-lowest resonant frequency in the device.
The LSM image exhibits a clear standing-wave pattern
FIG. 5. (a) LSM reflectivity image with the arrow indicating with two antinodes along the length of the transmission
the direction of 1D line scans. (b) Microwave transmission line. We perform numerical electromagnetic simulations
(not normalized) |S21 (f )| measured on LSM. (c) Average LSM of our metamaterial layout using Sonnet, as aredescribed
photoresponse R(y0 , f ) along 1D line scans. (d) 1D line scans in detail in Sec. VI. Figures 6(c, d) contain plots of the
R(x, y0 , f ) vs. frequency; dashed horizontal lines indicate simulated charge density and current density in the meta-
location of input and output coupling capacitors. material, also excited on the n = 40 mode.
With the application of LSM imaging to superconduct-
Figure 5(a) contains a reflectivity image of the metama- ing structures, the interaction of the laser spot with the
terial in the LSM, with an arrow indicating the orientation superconductor can produce a bolometric response, with
and location of 1D scans along the x axis of the photore- both inductive and resistive components, due to local heat-
sponse that are measured in the LSM. As the frequency is ing of the superconductor by the locally absorbed laser
varied, the photoresponse signal R(y0 , f ) averaged along power. In addition, there can be a non-thermal response
the 1D line scans exhibited sharp peaks [Fig. 5(c)], which due to the generation of nonequilibrium quasiparticles
line up with the microwave transmission resonances in in the superconductor from the deposited pair-breaking
|S21 (f )| that is measured simultaneously [Fig. 5(b)]. In laser energy. All three types of photoresponse signals are
Fig. 5(d), the density plot of R(x, y0 , f ), where y0 indi- proportional to the square of the local microwave current
cates the location of the line scan, again shows the sharp density in the superconductor, JM W (x, y) [46]. In our
features at the resonant frequencies, but also exhibits fine LSM imaging experiments on the LHTLs, we observe a
structure along the scan direction. Thus, it is clear that significant photoresponse in regions where we expect large
the LSM imaging of the device can be used to investigate microwave currents, in particular, in the vicinity of the
resonances in the metamaterial and their corresponding inductors near the current antinodes in the standing-wave
standing-wave patterns. patterns. The frequency dependence of the photoresponse
6

41 34 12

2.09 GHz 5.86 GHz 12.77 GHz


40 22 11

3.44 GHz 8.04 GHz 13.65 GHz


39 21 10

4.30 GHz 8.31 GHz 14.63 GHz


38 20 9

4.83 GHz 8.62 GHz 15.82 GHz


37 19 8

5.19 GHz 8.96 GHz 17.19 GHz


36 18 7

5.45 GHz 9.33 GHz 18.59 GHz

FIG. 7. Array of LSM images of metamaterial for different modes, labeled by mode number and frequency. Bright (dark) regions
correspond to large (small) PR signal. Amplitudes of PR signal are normalized for best contrast. Green boxes indicate the pair
of n = 34 and n = 8 modes, which both have 8 antinodes. Blue boxes indicate the pair of n = 22 and n = 20 modes, which both
have 20 antinodes.

on our LHTLs [Fig. 5(c)] and its correlation with |S21 (f )| V. ANALYSIS OF LSM IMAGES
[Fig. 5(b)] is most consistent with the nonthermal LSM
response mechanism [43]. However, we observe an even From the LSM images of the metamaterial resonator
larger photoresponse in regions where we expect large mi- (Fig. 7), we observe a stepwise change in the number of
crowave voltages, around the capacitors near the standing- antinodes in the standing-wave profile between neighbor-
wave antinodes for the charge density. This appears to ing modes. To understand these images, we analyze them
be caused by an enhancement in the loss in the capaci- separately for low and high mode numbers. For the higher
tors due to photoinduced carriers in the Si substrate in frequency modes, n ≤ 21, it is clear that the wave number
between the fingers of the interdigitated capacitors. This is decreasing with increasing frequency, consistent with
mechanism results in the largest photoresponse in regions the left-handed dispersion relation from Eq. (2). However,
with large microwave electric fields. We are not aware for the lower-frequency modes, from n = 41 to 21, the
of prior LSM imaging experiments of superconducting images show a standing-wave pattern with an increasing
interdigitated capacitors on Si. Thus, the nature of our number of antinodes for increasing mode frequency, a
sample layout and substrate material provide us with a typical behavior of right-handed media. A plot of the
photoresponse image of both the microwave current and mode frequency vs. the number of antinodes is shown
charge-density distributions in our LHTLs. in Fig. 8. For a given number of antinodes we see two
standing-wave profiles with different frequencies, one at
In the LSM image of Fig. 6(b) for the capacitors in the low-frequency part of the band, and one at high fre-
the LHTL near an antinode, we observe that the photore- quency. These partner modes can be observed at least up
sponse is highest at the edges of the capacitor where the to almost 20 GHz. Figures 7 and 8 contain colored boxes
inductors are connected. This effect can be attributed indicating two such pairs of modes.
to the enhanced reflectivity of the laser signal from the We can understand this behavior as an undersampling
regions covered with Nb compared with the bare Si re- effect due to both the discrete lumped-element nature of
gions beyond the capacitor edge. When the laser spot our metamaterial transmission line and due to the LSM
straddles the edge of the capacitor and the bare Si be- measurement itself. To demonstrate this, we consider
yond, the excess absorbed energy in the Si enhances the a simplified model of an ideal 10-cell metamaterial res-
photoresponse signal. onator with no stray reactances, sketched in Fig. 9(a). For
now, we focus on the microwave currents in the inductors,
We repeat the LSM imaging for all of the modes in our but the same analysis also applies to the voltage across
LHTL below 20 GHz. We present an array of many of the capacitors. Our open-ended boundary conditions im-
these in Fig. 7. The various features in these images are ply current standing-wave patterns of the form sin (kn x),
discussed in the next section. with kn = nπ/L. However, the LSM photoresponse is
7

proportional to the square of the local current density, (a) Cc Cl Cc


so in Fig. 9(b, c) we plot the square of this continuous
waveform. The figure contains the continuous waveforms Ll
for a low-frequency mode and its partner high-frequency
mode N − p and p, which are clearly different. However, 1
the current-square values at the unit-cell locations, indi-

I/Imax
cated by red dots in the figure, appear the same. This 0
undersampling effect is consistent with the LSM images 1 3 5 7 9
and the plot of ωn vs. the number of antinodes in Fig. 8 -1
if we extend the example circuit in Fig. 9 to 42 cells. x/Δx
(b) Cc Cl Cc
20
Ll

15 1

I/Imax
ωn/2π (GHz)

0
10 1 3 5 7 9
-1
x/Δx
5

FIG. 10. Circuit schematics for ideal ten-cell metamaterial


00 5 10 15 20
resonators with arrows indicating the relative sense of currents
flowing in each unit cell along with waveform for standing-
Number of antinodes wave current for hypothetical continuous LHTL. Red circles
correspond to current at location of each unit cell for mode
FIG. 8. Plot of mode frequency vs. number of antinodes in the (a) n = 8, (b) n = 2.
photoresponse from the corresponding LSM images. Colored
boxes indicate the partner modes shown in Fig. 7.
the continuous waveform for the standing-wave current
along with the corresponding discrete values at each of the
Cl unit cells (Fig. 10). This analysis indicates that the low-
(a)
Cc
frequency modes, with mode number n > N/2, should
Cc
Ll have currents that alternate in sign between adjacent
unit cells. As the LSM is only sensitive to the square of
(b) the local current density, we are unable to resolve these
1 relative phases.
(I/Imax)2

This undersampling effect is also the origin for a beating


pattern that is particularly prominent for modes near N/2.
0 1 3 5 7 9
Figure 11(a) shows the LSM image for mode n = 23,
(c) which exhibits such a beating pattern, along with a line-
1 cut of the square root of the photoresponse along the
(I/Imax)2

center of the image in Fig. 11(b). Appendix E details the


calculation of current and voltage values at the circuit
0 1 3 5 7 9 nodes for an ideal LHTL resonator. Although nonideal
x/Δx lumped-element effects may be important, especially at
short wavelengths, for now we assume that the currents
FIG. 9. (a) Circuit schematics for ideal ten-cell metamaterial and voltages are constant within each unit cell. The
resonator. Waveform for the square of standing-wave current calculated voltage across the capacitors from Eq (E3),
for hypothetical continuous LHTL for mode (b) n = 8, (c) shown in Fig. 11(c), displays reasonable agreement with
n = 2, including red circles corresponding to location of each the measurement.
unit cell.

Besides a comparison with the LSM images, we can


also use this approach to understand more details about VI. NUMERICAL SIMULATIONS OF
the nature of the mode patterns in the metamaterial METAMATERIALS
resonators. If we again consider the ideal ten-unit-cell
device, but plot the local current vs. position rather We simulate the microwave properties of our metama-
than the square of the current, we can again compare terial structures using the Sonnet electromagnetic field
8

(a) 23
(a) 0
line-cut scan
-20
7.79 GHz -40
(b) 1.0 -60 ADR

|S21| (dB)
0.8
PR (a.u.)

0.6 -80
0.4
0.2 -100
0.0
(c) -120 Sonnet
1.0
0.8 -140
V/Vmax

0.6 2 4 6 8 10 12 14 16
0.4
0.2 Frequency (GHz)
0.0 (b) 0
0.0 0.2 0.4 0.6 0.8 1.0
position along sample
-40

|S21| (dB)
FIG. 11. (a) LSM photoresponse for mode 23 with the arrow
indicating the location of subsequent line cut. Bright (dark) -80
regions correspond to large (small) PR signal. (b) Plot of the
square root of LSM photoresponse signal along line cut. (c) -120
10 20 30 40
Standing-wave pattern of voltage across capacitors computed
with Eq. (E3) for mode 23 of a 42-cell metamaterial resonator.
Frequency (GHz)

FIG. 12. (a) |S21 (f )| from Sonnet simulation (red) offset by


60 dB compared with measured LHTL spectrum from ADR
at 65 mK (black). (b) |S21 (f )| from Sonnet simulation plotted
solver [40]. We compute the S21 transmission spectrum, out to 45 GHz showing gap beyond n = 0 mode near 40 GHz.
as well as the current-density and charge-density distri-
butions, near the metamaterial resonance frequencies. In
Fig. 12(a) we compare the simulated |S21 (f )| from Sonnet along the length of our metamaterial resonator without
with our measured spectrum that is presented earlier in having the inductors of neighboring cells interfering with
Fig. 4. The simulated curve is in reasonable qualitative each other. However, this arrangement for the inductors
agreement with the experimentally measured spectrum, allows for the possibility of non-lumped-element behavior
although there are notable deviations, particularly at the impacting the frequencies of the high wave-number modes.
lowest frequency modes. Both the Sonnet and experimen- To explore this further, we implement a metamaterial lay-
tal spectra exhibit the highest density of modes around out in Sonnet with the same Ll and Cl values as in our
6 GHz, in contrast to the prediction from Eq. (2) that other devices, but with all of the inductors on one side of
the mode density should be highest just above ωIR where the transmission line. This requires that the simulated
the band for the ideal circuit is flat. We also observe structure is 10% longer than our measured metamaterial
good agreement between the resonance frequencies from resonators. From the simulations of the metamaterial
the simulated spectrum, theoretical dispersion relation with nonstaggered inductors, we observe that the mode
from Eq. (2), and the experimentally measured values frequencies are consistent with the prediction from Eq. (2),
from both the ADR and LSM over much of the frequency even up to the highest wave-number mode (Fig. 13). Fur-
range of our measurements (Fig. 13). However, there ther details of the inductor staggering characterization
are discrepancies between these various methods for the are discussed in detail in Appendix G.
modes near the low-frequency end of the spectrum. Besides the frequency shifts arising from the staggered
Modes just above ωIR correspond to the highest wave inductors, another factor that impacts the frequencies
number and thus the fastest spatial oscillations, where of the high wave-number modes is the quality of the mi-
the wavelength is approaching twice the unit cell size. crowave grounding of the chip. This is evident in Fig. 13,
Thus, deviations from ideal lumped-element behavior are where the mode frequencies of the high −n resonances
most prominent in this regime. One such deviation arises measured on the LSM, where it is especially challenging to
from our choice of layout. Due to the constraints of our maintain a good microwave ground across the chip while
microwave chip holder, we chose a metamaterial design permitting the LSM imaging process, are even further
with staggered inductors, where inductors on the odd- shifted from the dispersion relation of Eq. (2) compared
numbered unit cells terminate on one side of the ground to the ADR measurements and the Sonnet values. We
plane, while those on the even-numbered cells connect to simulate the effects of degraded ground connections for
the other side. This permitted us to fit more unit cells the metamaterial using Sonnet and observed that this
9

indeed results in downward shifts of the resonance fre- 40


quencies for the modes with the highest wavenumber. 7

ωn/2π (GHz)
Appendix F contains a more detailed discussion of the
effects of ground quality on the mode spectrum. 30 5

ωn/2π (GHz)
3
In addition to studying the effects of different inductor
20
layouts and imperfect grounding, we can also use Sonnet 1.0
0.9
to explore the spectrum of our metamaterial resonators knΔx/π
for frequencies beyond the range of our experimental hard- 10 Eq. (2)
ware. In Fig. 12(b), we extend the Sonnet simulation of LSM Sonnet
our metamaterial up to 46 GHz, where it is clear the ADR Sonnet non-staggered
mode spacing continues to increase and the linewidths be- 00 0.2 0.4 0.6 0.8 1.0
come progressively larger. In addition, beyond the n = 0 knΔx/π
mode, we observe a gap of roughly 4 GHz in the simu-
lated spectrum. This is consistent with the theoretical
FIG. 13. Plot of mode frequency vs. normalized waven um-
prediction from Eq. (2) for a metamaterial with different ber for ADR measurements, LSM images, Sonnet simulations,
self-resonance frequencies for the inductor and capacitor including both staggered and non-staggered inductor configura-
elements [22]. We run separate simulations in Sonnet tions, as discussed in the text; the solid black curve corresponds
of an individual inductor and capacitor from our meta- to dispersion relation of Eq. (2). The inset shows a closeup of
material design and obtained self-resonance frequencies behavior near ωIR /2π.
of approximately 50 GHz for the capacitor and approx-
imately 60 GHz for the inductor, close to the location
of the gap in the simulated LHTL resonator spectrum VII. COUPLING AND INTERNAL LOSSES
of 40-44 GHz. The simulation of individual capacitors
and inductors may not capture the actual grounding en- The quality factor of the metamaterial resonances can
vironment or loading effects that arise when the elements be determined by standard fits to the trajectory of S21 (f )
are built into a metamaterial, thus leading to differences in the complex plane, in the same way as with a resonator
in the simulated self-resonance frequencies. Nonetheless, formed from either a conventional transmission line or
the simulations give qualitative agreement between the lumped elements [47, 48]. For a dip-style reflection mea-
capacitor and inductor self-resonances and the gap in surement of a resonator coupled to a feedline where the
the metamaterial spectrum. If we take the self-resonance baseline level self-calibrates full transmission, the fitting
frequencies from the gap frequencies, they correspond process allows for a simultaneous extraction of both inter-
to stray reactances of Lr = 59.5 pH and Cr = 21.8 fF. nal losses Q−1 and losses due to coupling to dissipation in
i
For the Sonnet simulations of the alternate metamaterial external circuitry Q−1c . However, for peak-style transmis-
layout with non-staggered inductors discussed earlier, we sion resonances, as with the metamaterial measurements
observe in Fig. 13 that the high-frequency, low−n modes presented here, one requires a separate baseline calibra-
are shifted to somewhat lower frequencies than those of tion in order to separate the values of Qi and Qc from
the layout for our measured device, consistent with a the fit. Such baseline measurements inevitably introduce
reduction in the inductor self-resonance frequency due to an extra source of variation since they must be performed
the different geometry. on a separate cooldown with a different device that may
have slight differences in the bonding and packaging.
We note that the simulation of the transmission spec- Following a separate baseline measurement on our ADR,
tra for our metamaterial design is rather time consuming. we obtain the normalized spectrum for our metamaterial
We use a mesh size of 1 µm to capture the smallest fea- resonator in Fig. 4. We then fit each resonance and extract
ture size in our layout for the interdigitated capacitors Qc and Qi , allowing us to plot the internal and external
and meander-line inductors. These relatively small fea- losses as a function of mode number n in Fig. 14. It is
ture sizes, combined with the large spatial extent of our clear that most of the modes are strongly overcoupled,
metamaterial resonator, pose a significant challenge for with Q−1 c  Q−1i , consistent with the majority of the
a numerical field solver. Using a small-scale computing resonance peaks in the spectrum of Fig. 4 reaching close
cluster with 10 slave machines, each with 16 cores and 64 to full transmission. The error bars are determined by an
GB of memory, a simulation of a complete spectrum, as estimate of the uncertainty in the baseline transmission
in Fig. 12, takes approximately 3.5 weeks. However, once level from variations in |S21 | for cooldowns of multiple
the spectrum computation is completed, we are able to nonmetamaterial devices exhibiting broadband full trans-
run simulations of the current-density and charge-density mission with the same wiring as on our metamaterial
distributions at particular frequencies, with scans at all measurements. We take this variation to be ±2 dB, tak-
42 modes requiring about 2 days to complete. We present ing care to ensure that the normalized |S21 | never exceeds
examples of such distributions in other sections of the unity.
manuscript, including Sec. IV and Appendix E. From Fig. 14, the Q−1i values for most of the modes are
10

with both right-handed (which we have not seen treated


15 previously in the literature) and left-handed dispersion re-
lations. For the simplest case of an ideal LHTL resonator,
1/Qi 1/Qc
1/Qi, 1/Qc (x 105)

we obtain
10


Cc2 cos2
QLHT
c
L
(n)−1 = 3
2N 

. (3)
2N Cl2 sin 2N
5
In Fig. 15, we plot Q−1c (n) from Eq. (3), along with
values from a numerical simulation of an ideal lumped-
0 element LHTL resonator using AWR Microwave Office
25 30 35 40 showing nearly perfect agreement. The Q−1 c values ex-
Mode number tracted from our measured resonances agree reasonably
well with Eq. (3) for lower values of n, but begin to de-
FIG. 14. Comparison of internal loss and coupling loss ex- viate significantly for higher n modes starting around
tracted from the measured S21 (f ). n ∼ 30. From fits to extract Qc from the resonances
in our Sonnet simulations, the Q−1 c values for the non-
staggered inductor configuration follow the dependence
of the order of 10−5 , consistent with dielectric loss in the of Eq. (3) rather closely. However, the values from the
interdigitated capacitors that make up the metamaterial Sonnet simulations of the same staggered-inductor layout
[49, 50]. However, a few of the lowest frequency, highest as with our experimental device also follow the trend of
n modes have an order of magnitude higher internal loss, excess coupling loss for the highest n modes. This effect
which may be related to the short wavelengths for these is likely a consequence of the short wavelengths for these
modes or non-lumped-element effects due to the stag- high-n modes, which are strongly influenced by the layout
gering of the inductors for this particular metamaterial of the inductors, as well as the integrity of the ground
layout. plane across the chip (Appendix F).
The Q−1c values generally decrease for larger n modes,
although the highest n mode deviates from this trend,
perhaps for the same reasons as the Q−1 i behavior of the VIII. CONCLUSIONS
high-n modes described above. For a conventional, dis-
tributed transmission-line resonator, formed from perhaps In conclusion, we demonstrate superconducting thin-
a coplanar waveguide, the mode-number dependence of film metamaterial resonators that produce dense mode
the coupling loss is quite different: Q−1 c ∝ n [51]. In spectra with left-handed dispersion in a frequency range
Appendix D, we derive expressions for Qc for resonators that is compatible with the integration of superconducting
formed from discrete lumped-element transmission lines, qubits in a cQED architecture. We perform LSM imag-
ing of the microwave field distributions on one of these
devices and compare this with both analytical modeling
of the circuit as well as numerical simulations that allow
us to predict the spectrum of new metamaterial designs.
10-4 While our current minimum mode spacing is 147 MHz,
this could be reduced by increasing the number of unit
Measurement cells N or by intentionally reducing the self-resonance of
10-5
1/Qc

Sonnet staggered the capacitive or inductive elements, perhaps by using a


high kinetic inductance film for the capacitor fingers or by
Sonnet non-staggered designing intentional capacitive shunts on the inductors.
10-6 Eq. (4) Metamaterial resonators with substantial nonlinearity and
AWR frequency tunability could also be developed by incorpo-
rating Josephson elements in place of the linear inductors
20 25 30 35 40 in our present design.
Mode number Our circuit layout allows for straightforward extensions
to future devices incorporating superconducting qubits
FIG. 15. Coupling losses for different mode numbers extracted that couple to the metamaterial resonances. This would
from experimentally measured resonances on ADR at 65 mK, potentially open up a new regime for analog quantum
AWR simulation of ideal lumped-element LHTL resonator, simulation with the spectrum of photonic modes tailored
and Sonnet simulations of both staggered and nonstaggered by the circuit parameters making up the metamaterial.
inductor configurations. the solid curve corresponds to Eq. (3). In addition to superconducting qubits, our metamaterial
design is also compatible with coupling to nanomechanical
11

devices [52], quantum-dot qubits [53, 54], or even qubits two expressions
based on hyperfine transitions in trapped ions or Rydberg
atoms in hybrid superconducting/atomic systems [55]. vm (2 + ZY ) =vm−1 + vm+1 (A5)
im (2 + ZY ) =im−1 + im+1 . (A6)
ACKNOWLEDGMENTS Assuming a plane wave solution for propagation through
the transmission line, the voltage and current for cell m
This work is supported by the U.S. Government under can be written as
ARO Grants No. W911NF-14-1-0080 andNo. W911NF-
15-1-0248. Device fabrication is performed at the Cornell vm =V0+ e−ikm∆x + V0− eikm∆x (A7)
NanoScale Facility, a member of the National Nanotech-
nology Infrastructure Network, which is supported by the im =I0+ e−ikm∆x + I0− eikm∆x , (A8)
National Science Foundation (Grant No. NNCI-1542081).
A.P.Z. acknowledges support from Volkswagen Founda- where ∆x is the unit cell length, and V0+ (V0− ) and I0+ (I0− )
tion (Grant No. 90284). B.G.T. acknowledges support are the amplitudes of the forward (reverse) propagating
from FAPESC, CNPq and the National Institute for Sci- voltage and current, respectively. Combining Eqs. (A7)-
ence and Technology - Quantum Information. A.V.U. (A8) with Eqs. (A5)-(A6), we obtain
acknowledges partial support from the German Science
Foundation (Grant No. US18/15-1), the Russian Science [V0+ e−ikm∆x + V0− eikm∆x ][2 cos (k∆x) − (2 + ZY )] = 0.
Foundation (Grant No. 16-12-00095), and the Ministry (A9)
of Education and Science of Russian Federation in the This expression must be satisfied for all values of km∆x,
framework of the Increase Competitiveness Program of therefore
the National University of Science and Technology MISIS
(Grant No. K2-2017-081). 2 cos (k∆x) = (2 + ZY ). (A10)

This relationship between k∆x and ZY leads to the dis-


Appendix A: Derivation of dispersion relation persion relation for the transmission line

1
k(ω) = ×
Cl Cl ∆x 
im-1 Lr im Lr im+1   
1 1 1
+ + + cos−1 1 − ωLr − ωCr − ,
2 ωCl ωLl
vm-1 Cr L l vm Cr Ll v (A11)
m+1
_ _ _ matching Eq. (2) in the main text. For the case of a purely
m-1 m right-handed line (i.e., 1/Cl =0 and 1/Ll =0), using stan-
dard trigonometric relationships combined with Eq. (A10),
FIG. 16. Voltage and current in a unit cell of a general discrete
transmission line. inverting Eq. (A11) yields the dispersion relation for a
right-handed line:
For a general discrete transmission line, such as the  
2 k∆x
circuit sketched in Fig. 1(b), we can study the current ωR (k) = √ sin . (A12)
and voltage supported by such a line by focusing on two Lr Cr 2
adjacent unit cells, as in Fig. 16, and applying Kirchhoff’s
laws to obtain the following relations On the other hand, if the line is purely left-handed (i.e.,
Lr =0, Cr =0), the admittance becomes 1/iωLl and the
  impedance becomes 1/iωCl . Again, inverting Eq. (A11)
1 with these conditions leads to the left-handed dispersion
vm − vm+1 =im iωLr + (A1) relation for an ideal LHTL
iωCl
 
1 1 1
im−1 − im =vm iωCr + (A2) ωL (k) = √ . (A13)
iωLl 2 Ll Cl sin k∆x
  2
1
vm−1 − vm =im−1 iωLr + (A3)
iωCl
Appendix B: Impedance of general terminated
 
1
im − im+1 =vm+1 iωCr + . (A4) discrete transmission line
iωLl

By defining the admittance Y = iωCr + 1/iωLl and In order to treat a resonator formed from a discrete
impedance Z = iωLr + 1/iωCl , we can reduce these to transmission line, we need to continue our analysis from
12

Appendix A to derive the impedance for a discrete trans-


mission line with a general terminating load. Before treat- Z-N
ing the addition of the termination, we revisit Eqs. (A7) R 0 C c C l Lr C l Lr C l Lr CcZl
and (A8) and substitute into Eqs. (A1)-(A4) to obtain
Zs Ll
Vg v-N Cr Ll C r Cr Ll R 0
V2
k∆x V0+
 
+ −ik∆x/2
I0 =2ie sin (B1)
2 Z -N -N+1 -1 0
  −
k∆x V0
I0− = − 2ieik∆x/2 sin . (B2)
2 Z
FIG. 17. Circuit diagram for a S21 measurement setup for an
Following these expressions, we define the characteristic N -cell metamaterial transmission line resonator. We assume
symmetric coupling so that the source impedance at the input
impedance of the line to be Z0 = Z/2i sin (k∆x/2) and
end Zs and the load impedance at the terminated end Zl are
we can express the current in cell m in terms of Z0 : equal. The total impedance of the LHTL resonator looking
from the left end is Z−N given by Eq. (B10) and is marked in
V0+ −ik(m+ 1 )∆x V0− ik(m+ 1 )∆x red dashed box.
im = e 2 − e 2 . (B3)
Z0 Z0
Next, we consider a finite length discrete transmission
This is a general result for a lossless discrete transmission
line with N unit cells and the last cell on the right being
line. For example, it is straightforward to compute the
terminated by a load impedance Zl , which forms an extra
input impedance for a loaded LHTL for a given mode
cell. To simplify the calculation, we choose the cell of the
number by applying the appropriate dispersion relation
terminating load to be m = 0. Introducing the reflection
from Appendix A to Eq. (B10).
coefficient Γ = V0− /V0+ , the voltage and current at cell m
can then be expressed in terms of Γ as

vm =V0+ e−ikm∆x + ΓV0+ eikm∆x (B4) Appendix C: Derivation of S21 (ω) for general
discrete transmission line resonator
V0+ −ik(m+ 1 )∆x V0+ ik(m+ 1 )∆x
im = e 2 −Γ e 2 . (B5)
Z0 Z0
In order to derive a general expression for S21 (ω)
The impedance at m = 0 can be found as through a discrete transmission line resonator, we take
the terminating load to be determined by the output
v0 1+Γ coupling capacitor Cc and a resistive load R0 such that
Zm=0 = = Z0 −ik∆x . (B6)
i0 ik∆x
e 2 − Γe 2 the impedance of the terminating cell at m = 0 is
Zl = R0 + 1/iωCc . We then add an input drive with
At cell m = 0, we require that Zm = Zl . Thus, we have source impedance R0 and input coupling capacitance Cc ,
−ik∆x
such that Zs = Zl (Fig. 17). For simplicity and to match
Zl e 2 − Z0 our experimental configuration, we assume symmetric
Γ= ik∆x , (B7)
Zl e 2 + Z0 coupling. However, our analysis could easily be extended
to the case of asymmetric coupling. For a source voltage
which, it is important to note, is not equal to one if Vg and voltage V2 across the output load R0 , the trans-
Zl → ∞, in contrast to a continuous transmission line. mission S21 is given by 2V2 /Vg . Initially we consider the
With the reflection coefficient calculated, we can then voltage at the first cell of the transmission line v−N and
calculate the impedance seen from the opposite end of the express this in terms of Vg
transmission line. Applying the convention of increasing
cell number from left to right and our choice of m = 0 for Z−N
the terminating cell on the right end of the line, the first v−N = Vg . (C1)
Zs + Z−N
cell on the left is m = −N . Thus, based on Eqs. (B4)-
(B5), we can write the voltage and current for the input Next we consider the voltage across Zl at the output end,
cell at the left end as which is (Zl /R0 )V2 . This voltage should correspond to
V0+ (1 + Γ) and V0+ can be obtained from Eq.(B8). Thus,
v−N =V0+ eikN ∆x + ΓV0+ e−ikN ∆x (B8) V2 is given by
V0+ −ik(−N + 1 )∆x V0+ ik(−N + 1 )∆x
i−N = e 2 −Γ e 2 , (B9) Z−N R0 1+Γ
Z0 Z0 V2 = Vg . (C2)
Zs + Z−N Zl eikN ∆x + Γe−ikN ∆x
leading to the following expression for the impedance at
the input cell Finally, S21 can then be written as

eikN ∆x + Γe−ikN ∆x 2Z−N R0 1+Γ


Z−N = Z0 1 1 . (B10) S21 = . (C3)
e−ik(−N + 2 )∆x − Γeik(−N + 2 )∆x Zs + Z−N Zl eikN ∆x + Γe−ikN ∆x
13

S21 can then be computed from Eq. (C3) by substituting


2.5
Eq. (B10) for Z−N and Eq. (B7) for Γ. Note that the
frequency dependence of S21 is determined by the disper-
sion relation k(ω) that one chooses. In Fig. 18, we plot 2.0

1/Qc (x 106)
1.5
1.0
1.0
0.8 Eq. (C4)
AWR 0.5 Eq. (D1)
0.6 AWR simulations
|S21|

0.0
0.4 0 5 10 15 20 25 30 35
Mode number
0.2
FIG. 19. Solid line is 1/Qc expression of a continuous trans-
mission line resonator, as described by Eq. (D1); note that this
8.246 8.248 8.250 8.252 8.254 is plotted as a continuous function for clarity, even though n is
Frequency (GHz) limited to integer values. The red points are the 1/Qc values
for a 40-cell right-handed discrete transmission line resonator
extracted from a circuit simulation using AWR Microwave
FIG. 18. S21 (f ) calculated for the n = 23 mode of a 42-cell
Office. The parameters for 1/Qc described by Eq. (D1) are:
LHRH metamaterial resonator using Eq. (C3) (solid red line)
ω0 /2π = 1 GHz, Z0 = 50 Ω, Cc = 1 fF. The parameters of the
and simulated using AWR Microwave Office (blue points). Unit
discrete RHTL resonator are: Lr = 0.625 nH, Cr = 250 fF,
cell parameters are: Cl = 266 fF, LL = 0.6 nH, Cr = 21.806 fF
Cc = 1 fF.
and Lr = 0.595 nH, chosen based on the discussion in Sec. III
and the Sonnet simulation result of the stray reactances.
equivalent LC resonant circuit for each mode of a discrete
S21 (ω) computed from Eq. (C3) for the n = 23 mode
transmission line resonator then mapping this onto the
of a LHRH resonator with the parameters given in the
expression for Qc for the simple LC circuit. We do this
caption. A numerical simulation of a circuit with these
below – first for an RHTL resonator, then for a resonator
same parameters in AWR Microwave Office yields quite
based on an LHTL.
good agreement.
For a basic LC circuit with symmetric input and output
coupling capacitors Cc to load and source resistors R0
Appendix D: Coupling quality factor for discrete and resonance frequency ω0LC , when ω0LC Cc R0  1, then
transmission line resonators Qc is given by [49, 51]

We can use the expressions derived in the previous C̃


appendices to investigate the coupling loss for a dis- Qc = , (D2)
2ω0LC Cc2 R0
crete transmission line resonator. For a resonator formed
from a continuous transmission line with characteris- where C̃ is the capacitance of the LC circuit. For a
tic impedance Z0 equal to the load impedance, when discrete RHTL or LHTL resonator, if we replace ω0LC
ωn Cc Z0  1, the coupling loss is given by [56]
with ωn for mode n and determine C̃ corresponding to
1 4Z02 Cc2 ωn 2 4nZ02 Cc2 ω0 2 mode n, we can then use this expression to compute Qc .
= = , (D1) For an ideal RHTL, we first compute the total electric
Qc nπ π
field energy stored in the RHTL resonator
where we make use of the fact that such a transmission
line will have a linear dispersion, resulting in harmonic res- −1
1 X
onances ωn = nω0 . In contrast, we use AWR Microwave EeRHT L = Cr |vm |2 . (D3)
2
Office to simulate a discrete transmission line resonator m=−N
and extracted Qc for all of the resonance modes. As
plotted in Fig. 19, there are dramatic differences in the When Cc is small, we can take the limit ZL =
variation of 1/Qc with n for all but the lowest n modes. (1/iωn Cc + R0 ) → ∞, so that Γ → e−ik∆x , then using
One approach to extract Qc for each mode of a discrete Eqs. (B4) and (B7), the voltage at each unit cell m is
transmission line resonator is to use Eq. (C3) for S21 (ω) h mnπ (m−1)nπ
i
and fit Lorentzians to each resonance peak to determine vm = V0+ e−i N + ei N . (D4)
the linewidth. Alternatively, it would be useful to derive
a closed-form expression for Qc for more efficient evalu- Next, we substitute Eq. (D4) into Eq. (D3) and equate
ation of circuits. This can be done by considering the the result to the total electric field energy stored in the
14

equivalent LC circuit for a voltage Vin across the capacitor Substituting this result for C̃LHT L in Eq. (D10) leads to
−1
2N Z0 Cl2 sin3 2N


1 X 1
Cr |vm |2 = N Cr (V0+ )2 = C̃|Vin |2 . (D5) QcLHT L
(n) = nπ
 , (D13)
2 2 R0 Cc2 cos2 2N
m=−N

In order to relate V0+ with Vin , we recognize that the which, when inverted, reduces to Eq. (3) for Z0 = R0 .
voltage at the input cell m = −N must be Vin . Then, As with the RHTL resonator analysis, we use Eq. (D13)
using Eqs. (B7)-(B8), we can write to compare the coupling loss for an LHTL resonator
 nπ  computed with this analytic expression with that obtained
|Vin |2 = |V−N |2 = 4(V0+ )2 cos2 . (D6) from an AWR Microwave Office circuit simulation and
2N linewidth fits extracted from the full expression for S21 (ω)
Thus, we extract the appropriate C̃ for mode n of an from Eq. (C3). Again, the agreement between the three
RHTL resonator, which we label C̃RHT L approaches is quite good. Also for completeness, the
inductance of the equivalent LC oscillator for mode n of
N an LHTL resonator can be found as before to be
C̃RHT L = Cr nπ
. (D7)
2 cos2 2N nπ

2Ll cos2 2N
Combining this result with Eq. (D1) and the RHTL dis- L̃LHT L = . (D14)
N
persion relation from Eq. (A12), we obtain
Cr2 Z0 N Our analysis thus far in this appendix assumes ideal
QRHT
c
L
(n) = nπ
 nπ
. (D8) RHTL or LHTL resonators with no stray reactance. For
8Cc2 R0 sin 2N cos2 2N
a realistic circuit that does include such parasitic effects,
In Fig. 20 we use the analytic expression for QRHT c
L
these stray reactances can indeed be accounted for, but
from Eq. (D8) with the 1/Qc values vs. n for an RHTL the expressions become rather unwieldy. To test the
resonator from the AWR Microwave Office circuit simula- effects of neglecting the stray reactances in this analysis,
tion presented earlier in Fig. 19, as well as linewidth fits in Fig. 22 we compare the coupling loss vs. n for an ideal
extracted from the resonance peaks using Eq. (C3) for LHTL resonator computed from Eq. (D13) with coupling
S21 (ω). There is excellent agreement between all three loss values obtained from the complete S21 (ω) expression
approaches. from Eq. (C3) using realistic values of stray reactance Lr
For completeness, the inductance of the LC oscilla- and Cr included. As the comparison shows, the analytic
tor correspondingpto mode n of an RHTL resonator is expression from Eq. (D13) for an ideal circuit agrees
determined by 1/ L̃C̃ = ωn , leading to reasonably well with the realistic LHRH resonator for
mode numbers beyond ∼ 10. Only the lowest n (highest
Lr frequency) modes have any significant deviation. Thus,
L̃RHT L = nπ
. (D9)
2N tan2 2N the simple analytic expression from Eq. (D13) can be used
for estimating coupling losses for LHTL resonators for
We apply a similar approach to compute the coupling all but the highest frequency modes, where a numerical
loss for an LHTL resonator, but with Eq. (A13) for the
appropriate dispersion relation, resulting in

1.2

LHT L Z0 Cl sin 2N C̃LHT L
Qc = . (D10)
Cc2 R0
1.0
1/Qc (x 106)

In order to compute C̃LHT L , we recognize that all of the


0.8
capacitors Cl are in series in the LHTL, while the cell
voltage vm is defined relative to ground. Thus, the voltage 0.6 1/Qc from AWR
across the capacitor in cell m is vm−1 -vm and the electric
field energy stored in the LHTL resonator is 0.4 1/Qc from Eq. (C4)

1 X
−1 0.2 1/Qc from Eq. (D10)
EeLHT L = Cl |vm−1 − vm |2
2 (D11) 0.0
m=−N
 nπ  0 10 20 30 40
=4N Cl (V0+ )2 N sin2 , Mode number
2N
using Eq. (D4) for vm . Thus, the effective capacitance of FIG. 20. Coupling loss comparison for an RHTL resonator
the equivalent LC circuit for an LHTL resonator is for analytic expression from Eq. (D10), AWR Microwave
Office circuit simulation, and linewidth extraction for S21 (ω)
2N sin2 2Nnπ

expression from Eq. (C3). The parameters for the RHTL are:
C̃LHT L = nπ
 Cl . (D12) C = 250 fF, L = 0.625 nH, Cc = 1 fF, N = 40.
cos2 2N
15

extraction of linewidths using Eq. (C3) should be used


instead.
1000

100

1/Qc (x 106)
Appendix E: Spatial variations in charge and
current distributions in LHTL resonators 10

1
As discussed in Sec. IV, when we image our metamate- Ideal LHTL resonator
rial resonator while exciting the various modes, we observe 0.1 LHRH resonator
both the microwave current and charge density distribu-
tions and these are consistent with Sonnet simulations 0 10 20 30 40
(Fig. 6). In contrast to standing waves on a continuous Mode number
transmission line resonator, where the antinodes in the
current will always match the nodes in the voltage and FIG. 22. Coupling loss comparison between calculation for an
vice versa, for our images of the n = 40 mode for our ideal LHTL resonator using Eq. (D13) with an LHRH resonator
metamaterial resonator, we observe the current density with non-zero stray reactances from linewidth extractions using
and charge density distributions to be spatially in phase. Eq. (C3). The parameters for the LHTL are: C = 250 fF,
In this Appendix, we show that this behavior is expected L = 0.625 nH, Cc = 10 fF, N = 40 Cr = 16.211389 fF, Lr =
for a discrete transmission line resonator when exciting a 0.0334947 nH, corresponding to self resonance frequencies of
high-n mode; while for a low-n mode, the two distribu- 50 GHz and 55 GHz.
tions are spatially out of phase, as with a conventional
continuous resonator. We begin by revisiting the treat- capacitor and current through the inductor for cell m are
ment in Appendix B, where Eqs. (B4) and (B5) give
the voltage and current at cell m. Here, we write these
h nπ i
vmCl
(n) = 2V0+ sin (m − 1)
expressions explicitly for mode n h  nπ  N  i (E3)

× sin + i − i cos
h n n
i N N
vm (n) = V0+ e−im N π + ei(m−1) N π (E1) 1


+
iL
m (n) = 2I0 cos
l
m−
2 N
h 1 n 1 n
i
im (n) = I0+ e−i(m+ 2 ) N π − ei(m− 2 ) N π . (E2) h  nπ   nπ i (E4)
× 1 − cos + i sin .
N N
As discussed in Appendix D, the voltage across each
capacitor of the LHTL is: vm Cl
= vm−1 − vm and the In order to compare with our Sonnet simulations of the
current flowing through each inductor is iL charge density, which is related to the voltage across the
m = im−1 − im .
l

Thus, on resonance for mode n, the voltage across the capacitors, and the current density, dominated by the
current through the inductors, we take the magnitude of
Eqs. (E3)-(E4). If we start by examining the low-n limit,
for mode n = 1, we obtain
104    
vml (1) ≈ 2V + π sin (m − 1) π
C
(E5)
103 0
N N
" #
1

102 iml (1) ≈ 2I + π cos m − 2 π .
L  
1/Qc (x 106)

0 (E6)
N N
10
From Eqs. (E5) and (E6), it is clear that the voltage and
1/Qc from AWR
1 current standing wave patterns are spatially out of phase,
1/Qc from Eq. (C4) as one would expect for a continuous transmission line
0.1 1/Qc from Eq. (D13) resonator. However, if we examine the large-n limit, for
mode n = N − 1, we obtain
0 10 20 30 40 
N −1
 
Mode number vml (N − 1) ≈ 4V + sin
C
0 (m − 1) π (E7)
N

FIG. 21. Coupling loss comparison for an LHTL resonator
 
iml (N − 1) ≈ 4I + sin
L N −1
for analytic expression from Eq. (D13), AWR Microwave Of- 0 (m − 1) π , (E8)
N
fice circuit simulation, and linewidth extraction for S21 (ω)
expression from Eq. (C3). The parameters for the LHTL are: thus, the voltage and current standing wave patterns are
C = 250 fF, L = 0.625 nH, Cc = 10 fF, N = 40. now in phase spatially. This behavior is also reflected
16

n=3 current density n=39 current density (a) inductor scan


inductor

PR (mV)
Sonnet max 0 Sonnet max 0 n=40 0
-2
capacitor scan
-4
1 1

ImLl (39)
ImLl (3)

2 capacitor

PR (mV)
inductor scan 1
0 0 0
-40 -30 -20 -10 0 -40 -30 -20 -10 0
m m -1
0 100 200 300
n=3 charge density n=39 charge density LSM PR 0 max X-scan (a.u.)
Sonnet max 0 Sonnet max 0
0.8 inductor

PR (mV)
(b) inductor scan
n=9 0.6
1 0.4
1
VmLl (39)
VmLl (3)

capacitor scan 2.5 capacitor

PR (mV)
0 0-40 1.5
-40 -30 -20 -10 0 -30 -20 -10 0
m m inductor scan 0.5
200 μm 0 100 200 300
FIG. 23. Standing wave patterns for modes n = 3 and 39 of 42- X-scan (a.u.)
cell LHTL resonator from Sonnet simulations of current density
and charge density distributions compared with expressions for FIG. 24. High-resolution LSM images of 11 cells near center
current in inductors (voltage across capacitors) from Eq. (E4) of metamaterial resonator for (a) n = 40 and (b) n = 9 modes.
[Eq. (E3)] normalized by the maximum value for each mode Bright (dark) regions correspond to large (small) PR signal.
L L
of current i0 (n) (voltage v0 (n)) so Iml (n) = |iml (n)/i0 (n)| Plots to the right of image correspond to linecuts as indicated
Cl Cl
(Vm (n) = |vm (n)/v0 (n)|). on the images. For the inductor linecuts, the traces through
the top and bottom inductors are added together.

in our Sonnet simulations. In Fig. 23, we compare the


standing wave patterns for mode n = 3 and 39, computed Thus, the voltage antinodes will line up with the current
from the magnitude of Eqs. (E3)-(E4) with the corre- nodes and vice versa, as one would observe with a contin-
sponding Sonnet simulations of the charge density and uous transmission line resonator. Next, we can examine
current density. the opposite limit, where for n = N − 1, we have
In addition to the Sonnet simulations and Eqs. (E7)-
vmr (N − 1) ≈ 2V + sin mπ
C  
(E8), we can also observe this behavior in the relative 0 (E11)
N 
positions of the standing-wave antinodes in the LSM
imr (N − 1) ≈ 2I + sin mπ .
L 
images, as mentioned earlier for Fig. 6. In Fig 24, we 0 (E12)
N
present high-resolution LSM images using a smaller laser
spot size (5 µm) of the middle 11 cells of the metamaterial As with the LHTL resonator for high n, the voltage and
resonator for modes n = 40 and 9. For each image, we current standing-wave patterns are spatially in phase. Of
include plots of linecuts across the inductors at the top course, for the RHTL resonator, this occurs for the highest
and bottom as well as the center of the capacitors. For frequency modes, while for the LHTL resonator, it is the
n = 40 (3.44 GHz), clearly the node in the LSM signal is in lowest frequency modes.
the same position for the inductor and capacitor; similarly,
the antinodes line up as well. For n = 9 (15.82 GHz),
although the PR signal in the inductors is weaker for Appendix F: Effects of imperfect grounding
this high-frequency excitation, the antinodes and nodes
no longer line up between the capacitor and inductor In our experimental setup for measuring S21 (f ), the
linescans. Nb ground plane of the chip is connected to the copper
We follow a similar analysis for a discrete RHTL res- ground plane of our sample holder through a series of
onator. Again, starting from Eqs. (E1)-(E2), we derive short aluminum wire bonds around the perimeter of the
similar expressions for mode n for the current through chip. In addition, there are multiple jumper wirebonds
the inductor m, iLm (n), and voltage across the capacitor
r across the LHTL and the CPW segments on either end to
Cr ensure that the different sections of ground plane are as
m, vm (n). For the n = 1 mode, we obtain
close to a uniform equipotential as possible. However, as
vmr (1) ≈ 2V + cos mπ is well known with superconducting thin-film microwave
C  
0 (E9)
 mπ N  circuits [57], the nonzero self-inductance of the wirebonds
L +
im (1) ≈ 2I sin can lead to imperfect grounding over the entire area of the
r
. (E10)

0
N device. Here, we attempt to simulate the effects of such
17

imperfect grounding conditions using Sonnet to study device with that for a similar metamaterial resonator,
how this affects the spectrum and coupling quality factor but with a non-staggered layout with all of the inductors
of the metamaterial resonances. on one side of the metamaterial. For this non-staggered
The typical grounding setup in Sonnet involves defining layout, we maintained the same capacitor parameters for
a grounding box and attaching the metal layer bound- the finger width, length, spacing, and number of pairs;
aries of the device to the sides of the grounding box. the meander-line inductor linewidth and number of turns
The grounding box and the device are normally both are also matched to our original layout. However, in
rectangular to fit together exactly so that the simulated order to avoid adjacent inductors from overlapping, the
metal layer is perfectly grounded. In addition to the ideal interdigitated capacitor spines are widened so that the
grounding scenario described above, which we refer to unit cell size increased by 6 µm. Thus, the total length of
as “grounded”, we consider three other levels of ground the non-staggered metamaterial resonator must increase
connection quality where we extend the ground box so by 252 µm beyond that of our measured device.
that it has a 200 − µm gap around the perimeter of the In Fig. 26 we compare the simulated S21 (f ) spectra
chip. We then add different numbers (12, 3, and 0) of for the original staggered and non-staggered inductor
6 × 200 µm metal strips to join the edges of the chip to configurations. Although both spectra follow the overall
the ground box. We note that these simulated connec- pattern expected for an LHRH resonator, as presented
tions are smaller than our actual wirebonds, which have earlier, there are clear differences between the spectra for
a diameter of 32 µm and a typical length of ∼ 500 µm in the two inductor layouts. In the case of the non-staggered
order to enhance the effects of reduced ground connection inductor configuration, the infrared cut-off frequency is
quality. about 2 GHz higher than that for the staggered inductor
In Fig. 25, we plot the results of these Sonnet simula- layout used in our experiments. In addition, the modes
tions of a 42-cell LHTL resonator for the various grounding just above ωIR /2π have the highest density, as expected
configurations where we compare the frequency of the from the theoretical treatment of the LHTL and LHRH
resulting modes as well as the coupling loss for each mode resonator spectra from Sec. II. This is in contrast to
from n = 24 − 41. We observe that the mode frequen- the staggered inductor spectrum, where the modes reach
cies are only influenced significantly for the highest ∼ 5 their highest density for a frequency about 2 GHz above
mode numbers (and hence lowest frequencies); while the ωIR /2π. This discrepancy between the spectra for the
coupling loss of only the highest ∼ 10 modes is impacted two inductor configurations can also be seen in Fig. 13,
by the ground quality. We do not plot modes n < 24 where the mode frequencies of the non-staggered inductor
(that is, higher frequency modes) here since there are no configuration follow the theoretical dispersion relation
discernible differences for the different grounding configu- closely, while the lowest frequency and highest n modes
rations. Generally, for the highest n and lowest frequency of the staggered inductor configuration fall below the
modes, the worse the grounding quality, the lower the theoretical curve. Based on the Sonnet simulations, we
frequency of the mode and the higher the coupling loss. conclude that the distribution of the currents through the
These simulations provide a qualitative explanation for staggered inductors for the shortest wavelength modes
most of the differences in our measured mode frequencies cause deviations from ideal lumped-element behavior,
and coupling losses and the Sonnet simulations with ideal leading to reduced resonance frequencies.
grounding and the calculated theoretical values based
on the circuit parameters, as presented earlier. We note
that in the case of the microwave transmission measured (a) (b)
during the LSM imaging, the downward frequency shifts 8
of the highest n and lowest frequency modes are even 100
larger than in our ADR measurements. In order to allow 7 50
Frequency (GHz)

the LSM laser spot to access all of the portions of the


1/Qc (x 106)

metamaterial, no jumper wirebonds are used across the 6


metamaterial or CPW segments, which likely degraded
the ground integrity. The reduced mode frequencies are 5
consistent with the Sonnet simulations performed here. 10
4 0 bond 0 bond
3 bonds 5 3 bonds
3 12 bonds 12 bonds
Appendix G: Effects of staggered inductor layout Grounded Grounded
2
As described in Sec. III, we chose to use a staggered 25 30 35 40 25 30 35 40
layout for the inductors in our metamaterial resonators, n n
as shown in the device images of Fig. 3. In this Appendix,
we use Sonnet simulations to examine the effects of the FIG. 25. Sonnet simulations for different grounding configura-
inductor layout configuration on the resonator spectrum. tions for a 42-cell LHTL resonator for (a) resonance frequency
We compare the simulated spectrum for our measured vs. n, (b) coupling loss 1/Qc vs. n.
18

0 Thus, although the simulations of the non-staggered con-


figuration lead to closer agreement with the theoretical
-20 dispersion relation for medium- and high-n modes, the
Staggered arrangement with all of the inductors on one side of the
-40
|S21| (dB)

metamaterial leads to excess stray reactances Lr and Cr ,


-60 resulting in the decreased gap at high frequency.
-80
Non-staggered
-100 In addition to differences in the mode frequencies, we
5 10 15 20 25 30 35 40 can also use the Sonnet simulations to study the effect
Frequency (GHz) of the inductor configuration on the coupling loss. As
shown in Fig. 15 where we plot 1/Qc vs. n for the ex-
FIG. 26. Sonnet simulations of S21 (f ) for non-staggered (or- perimental measurements, AWR circuit simulations of
ange, with -40 dB offset for clarity) and staggered (blue) an ideal LHTL resonator, and Sonnet simulations for
inductor layouts. Note that the blue curve is the same as in both inductor configurations, the Sonnet simulation for
Fig. 12. non-staggered inductors is reasonably close to the theoret-
ical dependence derived in Appendix D. In contrast, the
staggered inductor simulation from Sonnet shows excess
Besides the differences near the infrared cutoff, the coupling loss for modes beyond n ∼ 30, consistent with
spectra for the two inductor configurations also deviate our experimentally measured values. Again, this is consis-
at high frequencies. The gap beyond mode n = 1 occurs tent with deviations from ideal lumped-element behavior
about 12 GHz lower in frequency for the non-staggered for the short wavelength, high-n modes, which impact the
inductor configuration compared to the staggered layout. coupling loss dynamics as well as the mode frequencies.

[1] J. I. Cirac and P. Zoller, Goals and opportunities in [10] D. L. Underwood, W. E. Shanks, J. Koch, and A. A.
quantum simulation, Nat. Phys. 8, 264 (2012). Houck, Low-disorder microwave cavity lattices for quan-
[2] S. Mostame, J. Huh, C. Kreisbeck, A. J. Kerman, T. Fu- tum simulation with photons, Phys. Rev. A 86, 023837
jita, A. Eisfeld, and A. Aspuru-Guzik, Emulation of (2012).
complex open quantum systems using superconducting [11] M. Fitzpatrick, N. M. Sundaresan, A. C. Li, J. Koch, and
qubits, Quantum Inf. Processing 16, 44 (2016). A. A. Houck, Observation of a dissipative phase transition
[3] A. Blais, R.-S. Huang, A. Wallraff, S. M. Girvin, and in a one-dimensional circuit QED lattice, Phys. Rev. X 7,
R. J. Schoelkopf, Cavity quantum electrodynamics for 011016 (2017).
superconducting electrical circuits: An architecture for [12] D. C. McKay, R. Naik, P. Reinhold, L. S. Bishop, and
quantum computation, Phys. Rev. A 69, 062320 (2004). D. I. Schuster, High-contrast qubit interactions using mul-
[4] H. Paik, D. Schuster, L. S. Bishop, G. Kirchmair, G. Cate- timode cavity qed, Phys. Rev. Lett. 114, 080501 (2015).
lani, A. Sears, B. Johnson, M. Reagor, L. Frunzio, L. Glaz- [13] R. Naik, N. Leung, S. Chakram, P. Groszkowski, Y. Lu,
man, et al., Observation of high coherence in josephson N. Earnest, D. McKay, J. Koch, and D. Schuster, Random
junction qubits measured in a three-dimensional circuit access quantum information processors using multimode
qed architecture, Phys. Rev. Lett. 107, 240501 (2011). circuit quantum electrodynamics, Nat. Commun. 8, 1904
[5] A. Blais, J. Gambetta, A. Wallraff, D. Schuster, S. Girvin, (2017).
M. Devoret, and R. Schoelkopf, Quantum-information [14] Y. Liu and A. A. Houck, Quantum electrodynamics near
processing with circuit quantum electrodynamics, Phys. a photonic bandgap, Nat. Phys. 13, 48 (2017).
Rev. A 75, 032329 (2007). [15] J. Leppäkangas, J. Braumüller, M. Hauck, J.-M. Reiner,
[6] J. Majer, J. Chow, J. Gambetta, J. Koch, B. Johnson, I. Schwenk, S. Zanker, L. Fritz, A. V. Ustinov, M. Weides,
J. Schreier, L. Frunzio, D. Schuster, A. Houck, A. Wallraff, and M. Marthaler, Quantum simulation of the spin-boson
et al., Coupling superconducting qubits via a cavity bus, model with a microwave circuit, Phys. Rev. A 97, 052321
Nature 449, 443 (2007). (2018).
[7] X. Gu, A. F. Kockum, A. Miranowicz, Y.-x. Liu, and [16] R. Erickson, M. Vissers, M. Sandberg, S. Jefferts, and
F. Nori, Microwave photonics with superconducting quan- D. Pappas, Frequency comb generation in superconduct-
tum circuits, Phys. Rep. (2017). ing resonators, Phy. Rev. Lett. 113, 187002 (2014).
[8] J. Braumüller, M. Marthaler, A. Schneider, A. Stehli, [17] D. J. Egger and F. K. Wilhelm, Multimode circuit quan-
H. Rotzinger, M. Weides, and A. V. Ustinov, Analog tum electrodynamics with hybrid metamaterial transmis-
quantum simulation of the Rabi model in the ultra-strong sion lines, Phys. Rev. Lett. 111, 163601 (2013).
coupling regime, Nat. Commun. 8, 779 (2017). [18] A. Messinger, B. G. Taketani, and F. K. Wilhelm, Left-
[9] N. M. Sundaresan, Y. Liu, D. Sadri, L. J. Szőcs, D. L. handed superlattice metamaterials for circuit QED, Phys.
Underwood, M. Malekakhlagh, H. E. Türeci, and A. A. Rev. A 99, 032325 (2019).
Houck, Beyond strong coupling in a multimode cavity, [19] G. V. Eleftheriades, A. K. Iyer, and P. C. Kremer, Pla-
Phys. Rev. X 5, 021035 (2015). nar negative refractive index media using periodically lc
19

loaded transmission lines, IEEE Trans. Microwave Theory [36] C. Caloz and T. Itoh, Electromagnetic Metamaterials:
Tech. 50, 2702 (2002). Transmission Line Theory and Microwave Applications
[20] C. Caloz and T. Itoh, Transmission line approach of left- (John Wiley & Sons, 2005).
handed (lh) materials and microstrip implementation of [37] A. Lai, T. Itoh, and C. Caloz, Composite right/left-
an artificial lh transmission line, IEEE Trans. Antennas handed transmission line metamaterials, IEEE Microw.
Propag. 52, 1159 (2004). Mag. 5, 34 (2004).
[21] A. Sanada, C. Caloz, and T. Itoh, Novel zeroth-order [38] J. Koch, M. Y. Terri, J. Gambetta, A. A. Houck, D. Schus-
resonance in composite right/left-handed transmission ter, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin, and
line resonators, in Proc. 2003 Asia-Pacific Microwave R. J. Schoelkopf, Charge-insensitive qubit design derived
Conf., Vol. 3 (2003) pp. 1588–1591. from the cooper pair box, Phys. Rev. A 76, 042319 (2007).
[22] C. Caloz, A. Sanada, and T. Itoh, A novel composite [39] Sonnet, Sonnet suites, Version 16.52.
right-/left-handed coupled-line directional coupler with ar- [40] B. Plourde, H. Wang, F. Rouxinol, and M. LaHaye,
bitrary coupling level and broad bandwidth, IEEE Trans. Superconducting metamaterials and qubits, in Quantum
Microwave Theory Techn. 52, 980 (2004). Information and Computation XIII, Vol. 9500 (Interna-
[23] Y. Wang and M. J. Lancaster, High-temperature supercon- tional Society for Optics and Photonics, 2015) p. 95000M.
ducting coplanar left-handed transmission lines and res- [41] Ansoft, ANSYS HFSS, 3D Full-wave Electromagnetic
onators, IEEE Trans. Appl. Supercond. 16, 1893 (2006). Field Simulation.
[24] P. Jung, S. Butz, M. Marthaler, M. Fistul, J. Leppäkangas, [42] C. Kurter, A. P. Zhuravel, A. V. Ustinov, and S. M.
V. Koshelets, and A. Ustinov, Multistability and switch- Anlage, Microscopic examination of hot spots giving rise
ing in a superconducting metamaterial, Nat. Commun. 5, to nonlinearity in superconducting resonators, Phys. Rev.
3730 (2014). B 84, 104515 (2011).
[25] A. M. Zagoskin, D. Felbacq, and E. Rousseau, Quantum [43] A. P. Zhuravel, S. M. Anlage, and A. V. Ustinov, Mea-
metamaterials in the microwave and optical ranges, EPJ surement of local reactive and resistive photoresponse of
Quantum Techn. 3, 2 (2016). a superconducting microwave device, Appl. Phys. Lett.
[26] C. Hutter, E. A. Tholén, K. Stannigel, J. Lidmar, and 88, 212503 (2006).
D. B. Haviland, Josephson junction transmission lines [44] A. Zhuravel, A. Sivakov, O. Turutanov, A. Omelyan-
as tunable artificial crystals, Phys. Rev. B 83, 014511 chouk, S. M. Anlage, A. Lukashenko, A. Ustinov, and
(2011). D. Abraimov, Laser scanning microscopy of hts films and
[27] P. Jung, A. V. Ustinov, and S. M. Anlage, Progress in devices, Low Temperat. Phys. 32, 592 (2006).
superconducting metamaterials, Supercond. Sci. Techn. [45] J. C. Culbertson, H. S. Newman, and C. Wilker, Optical
27, 073001 (2014). probe of microwave current distributions in high temper-
[28] J. P. Martinez, S. Leger, N. Gheeraert, R. Dassonneville, ature superconducting transmission lines, J. Appl. Phys.
L. Planat, F. Foroughi, Y. Krupko, O. Buisson, C. Naud, 84, 2768 (1998).
W. Guichard, S. Florens, I. Snyman, and N. Roch, A tun- [46] A. Zhuravel, S. M. Anlage, S. K. Remillard,
able Josephson platform to explore many-body quantum A. Lukashenko, and A. Ustinov, Effect of LaAlO3 twin-
optics in circuit-QED, npj Quantum Inform. (2018). domain topology on local dc and microwave properties of
[29] M. Mirhosseini, E. Kim, V. S. Ferreira, M. Kalaee, cuprate films, J. Appl. Phys. 108, 033920 (2010).
A. Sipahigil, A. J. Keller, and O. Painter, Supercon- [47] D. P. Pappas, M. R. Vissers, D. S. Wisbey, J. S. Kline,
ducting metamaterials for waveguide quantum electrody- and J. Gao, Two level system loss in superconducting
namics, Nat. Commun. 9, 3706 (2018). microwave resonators, IEEE Trans. Appl. Supercond. 21,
[30] M. Castellanos-Beltran, K. Irwin, G. Hilton, L. Vale, and 871 (2011).
K. Lehnert, Amplification and squeezing of quantum noise [48] J. M. Sage, V. Bolkhovsky, W. D. Oliver, B. Turek, and
with a tunable josephson metamaterial, Nat. Phys. 4, 929 P. B. Welander, Study of loss in superconducting coplanar
(2008). waveguide resonators, J. Appl. Phys. 109, 063915 (2011).
[31] C. Macklin, K. OBrien, D. Hover, M. Schwartz, [49] A. D. OConnell, M. Ansmann, R. C. Bialczak,
V. Bolkhovsky, X. Zhang, W. Oliver, and I. Siddiqi, M. Hofheinz, N. Katz, E. Lucero, C. McKenney, M. Nee-
A near–quantum-limited Josephson traveling-wave para- ley, H. Wang, E. M. Weig, et al., Microwave dielectric loss
metric amplifier, Science 350, 307 (2015). at single photon energies and millikelvin temperatures,
[32] W. Zhang, W. Huang, M. Gershenson, and M. Bell, Appl. Phys. Lett. 92, 112903 (2008).
Josephson metamaterial with a widely tunable positive [50] J. M. Martinis, K. B. Cooper, R. McDermott, M. Steffen,
or negative kerr constant, Phys. Rev. Appl. 8, 051001 M. Ansmann, K. Osborn, K. Cicak, S. Oh, D. P. Pappas,
(2017). R. W. Simmonds, et al., Decoherence in Josephson qubits
[33] A. K. Iyer and G. V. Eleftheriades, Negative-refractive- from dielectric loss, Phys. Rev. Lett. 95, 210503 (2005).
index transmission-line (nri-tl) metamaterial lenses and [51] M. Göppl, A. Fragner, M. Baur, R. Bianchetti, S. Filipp,
superlenses, in Applications of Metamaterials (Metamate- J. Fink, P. Leek, G. Puebla, L. Steffen, and A. Wall-
rials Handbook Vol. 2), edited by F. Capolino (Taylor & raff, Coplanar waveguide resonators for circuit quantum
Francis-CRC Press, Boca Raton, FL, USA, 2009). electrodynamics, J. Appl. Phys. 104, 113904 (2008).
[34] E. Ovchinnikova, S. Butz, P. Jung, V. Koshelets, L. Fil- [52] F. Rouxinol, Y. Hao, F. Brito, A. O. Caldeira, E. K.
ippenko, A. Averkin, S. Shitov, and A. Ustinov, Design Irish, and M. D. Lahaye, Measurements of nanoresonator-
and experimental study of superconducting left-handed qubit interactions in a hybrid quantum electromechanical
transmission lines with tunable dispersion, Supercond. Sci. system, Nanotechn. 27, 364003 (2016).
Techn. 26, 114003 (2013). [53] T. Frey, P. J. Leek, M. Beck, A. Blais, T. Ihn, K. Ensslin,
[35] D. M. Pozar, Microwave Engineering (John Wiley & Sons, and A. Wallraff, Dipole coupling of a double quantum dot
2009). to a microwave eesonator, Phys. Rev. Lett. 108, 1 (2012).
20

[54] K. D. Petersson, L. W. McFaul, M. D. Schroer, M. Jung,


J. M. Taylor, A. A. Houck, and J. R. Petta, Circuit
quantum electrodynamics with a spin qubit, Nature 490,
380 (2012).
[55] J. D. Pritchard, J. A. Isaacs, M. A. Beck, R. McDermott,
and M. Saffman, Hybrid atom-photon quantum gate in a
superconducting microwave resonator, Phys. Rev. A 89,
010301 (2014).
[56] D. I. Schuster, Circuit Quantum Electrodynamics, Ph.D.
thesis, Yale University (2007).
[57] J. Wenner, M. Neeley, R. C. Bialczak, M. Lenander,
E. Lucero, A. D. OConnell, D. Sank, H. Wang, M. Weides,
A. N. Cleland, and J. M. Martinis, Wirebond crosstalk
and cavity modes in large chip mounts for superconduct-
ing qubits, Supercond. Sci. Technol. 24, 065001 (2011).

You might also like