Nonlinear resonators of superconducting thin-film kinetic inductance have attracted considerable research interest in the fields of detectors, qubits, parametric amplifiers, and more. By tuning the deposition parameters, niobium titanium nitride (NbTiN) films (∼12 nm in thickness) present high resistivity(∼2000 μΩ cm) and large sheet kinetic inductance (∼0.7 nH/□). We designed resonators with NbTiN micronwire as the inductor and aluminum as the capacitor, which result in a high internal quality factor and a Kerr nonlinearity of a few hertz. The radio frequency response of the resonators demonstrates nonlinear behavior similar to that of the cubic Duffing oscillator, including frequency shift and hysteresis region. The kinetic inductance resonator is a promising candidate for high saturation power parametric amplifiers.

Superconducting microwave resonators are fundamental and versatile devices. With the advancement of superconducting micro-nano-processes in the last decades, high performance superconducting microwave resonators have been gradually realized.1 Due to the ultra-low noise provided by the superconductivity and low temperature environment, a range of phenomena in superconducting microwave resonators can be harnessed. This enables advancements in fields such as quantum computing,2 sensing,3 and parametric amplifying,4 with the quantum noise being less than or equal to the limit.

In a superconducting microwave resonator, the introduction of a nonlinear element such as a Josephson junction, or the kinetic inductance (KI) from a superconductor, gives rise to distinguishable energy levels, with the lowest energy gap described by the so-called Kerr nonlinearity. Josephson junctions can typically provide Kerr nonlinearities in excess of 1 MHz,5 which is important for qubit state differentiation. However, the dynamic range of Josephson junction based parametric amplifiers is limited by the undesired higher order nonlinear effects.6 In addition, Josephson junctions are sensitive to magnetic fields, and the Josephson inductance may be unexpectedly tuned by the complex electromagnetic environment, which limit the application scenarios for Josephson junction-based devices. Devices using the kinetic inductance of superconducting films demonstrate magnetic resistance, and the Kerr nonlinearity can be varied from7,8 10−7 Hz to 100 kHz by the choice of materials and dimension design. Moreover, the nonlinearity effects caused by the material properties are less sensitive to the background magnetic noise.9 Thus, devices based on kinetic inductance have attracted increasing research interest, ranging from qubits10 to amplifiers4 and detectors.11 High kinetic inductance superconducting films of NbN,12 NbTiN,13 granular Al,10 and TiN8 were studied recently. As a ternary compound, NbTiN presents rich tunability in normal resistivity, critical temperature, crystallographic directions, and so on.14 These features make it achieve a high quality factor in resonators15, (>106), high saturation input power in parametric amplifiers16  (>40dBm), and high detection efficiency in single-photon detectors17 (nearly 95%).

For TTc, the kinetic inductance of the superconducting films can be obtained4 as Lk(I)=Lk(0)(1+I2I2+), where I sets the scale of the quadratic nonlinearity and is expected to be of the same order as the critical current. It shares the same form as the kinetic inductance of Josephson junctions. For the Josephson junction parallel circuit, its RCSJ18 (resistively and capacitively shunted junction) equation has the form of that of a driving cubic Duffing oscillator, in which the energy of nonlinear inductance corresponds to the potential energy. A driving cubic Duffing oscillator with mean-field approximation19 has been well-studied mathematically, and the bifurcation phenomenon is quite generic in nonlinear oscillators provided certain conditions are satisfied.20 Moreover, the basic model resulted in a lot of interesting research,21–25 such as the JPA (Josephson parametric amplifier), JBA (Josephson bifurcation amplifier), and JCA (Josephson chirped amplifier). Therefore, a pivotal inquiry pertaining to the strategic design and effective utilization of the nonlinear properties inherent in thin-film superconductors is carried out.

In this study, we characterized NbTiN films with large sheet kinetic inductance and experimentally demonstrated a high quality nonlinear resonator with a Kerr coefficient of around 10 Hz. Our experimental results suggest that by modulating the sputtering parameters, the NbTiN films exhibit high normal state resistivity and their kinetic inductors can be converted to super-inductors for microwave resonators.26 Furthermore, the kinetic inductance of the NbTiN films can introduce sufficient Kerr-type nonlinearity to our resonators, which displayed the bifurcation and jump phenomenon precisely as seen in the cubic Duffing oscillator.

We deposited the NbTiN films by DC magnetron sputtering onto intrinsic, high-resistivity (20kΩcm), ⟨100⟩-oriented 100 mm Si wafers in a high vacuum chamber with the base pressure down to 1.0 × 10−7 Pa. By changing the DC current flowing through the Nb target and the Ti target separately, we can control the deposition rate and the components of the thin films. The Ar/N2 mixture ratio can significantly influence the characterization of the final films as well. To obtain a prominent kinetic inductance, the films are deposited at a thickness of ∼15 nm.

The transport properties of the thin films were measured by the standard four-probe method in a Quantum Design Physical Property Measurement System, down to T = 20 mK with a DC excitation of 1μA, and the thickness of the thin films were measured by x-ray reflection in a Bruker D8 Discover. Table I shows the deposition parameters and the characterization of the NbTiN films. In the resistivity and temperature curve, we define the normal resistance at 20 K and the superconducting critical temperature Tc at the half-height points of superconducting and normal transition, as shown in Fig. 1. The results indicate that increasing the flow of N2 with a fixed flow of Ar can decrease the critical temperature of the NbTiN thin films and conversely increase the resistivity. For high nitrogen contents of NbTiN films, the same trends have been observed in many studies.27,28 Similarly, the reduction of current through the targets can increase the resistivity and decrease the critical temperature, especially for a comparison group with a high nitrogen component, and result in a larger kinetic inductance. The injection of nitrogen atoms alters the lattice and electronic structure of the material, increasing lattice distortion and grain boundary scattering, which affects its superconductivity.14 

TABLE I.

Deposition parameters and characterization of NbTiN films.

RecipeINb (A)ITi (A)Ar (sccm)N2 (sccm)P (Pa)Time (min)Thickness (nm)Tc (K)ρ (μΩ cm)Lk□,BCS (nH/□)
W1 0.36 0.18 13 0.29 10 12.46 5.9 1250 0.249 
W2 0.36 0.18 13 0.25 10 15.23 8.1 826 0.098 
W3 0.24 0.12 13 0.26 15 12.4 7.5 847 0.133 
W4 0.24 0.12 13 0.29 15 11.52 4.1 2340 0.726 
RecipeINb (A)ITi (A)Ar (sccm)N2 (sccm)P (Pa)Time (min)Thickness (nm)Tc (K)ρ (μΩ cm)Lk□,BCS (nH/□)
W1 0.36 0.18 13 0.29 10 12.46 5.9 1250 0.249 
W2 0.36 0.18 13 0.25 10 15.23 8.1 826 0.098 
W3 0.24 0.12 13 0.26 15 12.4 7.5 847 0.133 
W4 0.24 0.12 13 0.29 15 11.52 4.1 2340 0.726 
FIG. 1.

Resistivity as a function of temperature for the films grown at different parameters corresponding to Table I.

FIG. 1.

Resistivity as a function of temperature for the films grown at different parameters corresponding to Table I.

Close modal

According to the Bardeen–Cooper–Schrieffer (BCS) theory,29 the kinetic inductance per square Lk□ of the film is Lk□(T = 0) = ℏR/πΔ(0), where R is the normal state resistance per square when the film is in the normal state and Δ(0) ≃ 1.764kBTc is the superconducting energy gap at T = 0 K. We estimated the kinetic inductance per square as shown in Table I, and it is reasonable to let Δ(T) ≈ Δ(0) for30  Δ(T)=Δ(0)tanh[1.74(1T2Tc2)0.5], for which the temperature in the dilution refrigerators can reach 10 mK. The changes in the quasiparticle density of states would deviate from the BCS prediction,31 which are induced by the short elastic scattering length of the high-resistive electromagnetic response of NbTiN. However, this significant deviation occurs at higher temperatures (almost 0.2Tc), and it still agrees well with the BCS prediction in the low-temperature range.

We chose the parameters of recipe W3 shown in Table I, for the reason that thin films can provide large kinetic inductance but avoid superconducting phase transitions due to extreme disorder,32 to design the micronwire inductors. For a lumped resonator circuit, we can design the frequency f=12πLC and the impedance z=LC, where L and C are the inductance and capacitance of the resonator, respectively. Conversely, L and C can be designed from the given frequency and impedance for such a system of quadratic equations. As shown in the circuit schematic diagram in Fig. 2(a), the total inductance of the resonators can be separated into geometric inductance Lm and kinetic inductance Lk. The former is determined solely by the geometry of the inductors, while the latter is determined by the material characterization of the superconducting film and the geometry of the design. To simplify our design, we estimated the geometric inductance of the wire-type inductors33 using
where l is the wire length of the inductor and a and b are the thickness and width of the micronwire, respectively. For a 1 μm wide wire, the geometric inductance is about 1 pH/μm, which is approximately two orders of magnitude smaller than the kinetic inductance for the wire inductor in our device, and hence, we neglected Lm in our design. Then, we can change the length of the 1 μm fixed width micronwire to match the inductance of different frequencies by26  Lk=(lw)RπΔ. In order to match the impedance of the resonator as close as possible to 50 Ω, a 410 fF capacitor, shown in Fig. 2(c), is designed. The distance between the coplanar waveguide (CPW) and the capacitors is 100 μm, which determines the coupling between the resonator and the transmission line. To avoid unexpected coupling, we designed only two resonators on each chip and placed them apart.
FIG. 2.

(a) Circuit schematic diagram. (b) Fabrication diagram. (c) Microscopic photo of the device. The hanger resonators are capacitively coupled to a 50 Ω CPW feedline made of Al. The NbTiN micronwire is 1 μm in width and 13.5 μm in length, corresponding to sample R1. The upper layer of the Al capacitor is a 530 × 530 μm2 square connected to the lower layer of the inductor by two 5 × 5 μm2 pads. (d) Schematic of the measurement setup.

FIG. 2.

(a) Circuit schematic diagram. (b) Fabrication diagram. (c) Microscopic photo of the device. The hanger resonators are capacitively coupled to a 50 Ω CPW feedline made of Al. The NbTiN micronwire is 1 μm in width and 13.5 μm in length, corresponding to sample R1. The upper layer of the Al capacitor is a 530 × 530 μm2 square connected to the lower layer of the inductor by two 5 × 5 μm2 pads. (d) Schematic of the measurement setup.

Close modal

The fabrication process for our device mainly consists of three steps. First, we deposited an almost 12-nm-thick NbTiN film onto an intrinsic, high-resistivity silicon wafer. Second, the NbTiN micronwires and the marks were defined using a direct-write laser lithography system (Heidelberg DWL66+) using a positive AZ703 photoresist, and the pattern was then transferred by dry etching using reactive ions of CF4 plasma. Finally, the capacitors and CPW transmission line were delineated by a second round of photolithography. Following this, a 120-nm-thick aluminum film was deposited using ultra high vacuum e-beam evaporation systems for the lift-off process. The upper and lower layers, shown in Fig. 2(b), are connected to each other by the 5 × 5 μm2 NbTiN pads on either side of the micronwire. Figure 2(c) shows the microscopic photo of the device, in which the inductor is 13.5 μm in length and 1 μm in width. The 530 × 530 μm2 capacitors are separated from the ground by a 10 μm wide gap (see the supplementary material for the SEM and AFM photos). After the fabrication, the wafer was diced into 5 × 5 mm2 chips and then bonded into a copper sample box. As shown in Fig. 2(d), the sample was mounted on a cold finger installed at the mixing chamber stage of a Bluefors LD400 dilution refrigerator at a base temperature of 10 mK.

In order to find the resonant frequency over the entire bandwidth of the VNA, we first compare the transmission characteristics of the device at different power levels. For a nonlinear resonator with a negative Kerr coefficient, its resonant frequency decreases with increasing power, as shown in Fig. 3(b). Subsequently, we performed more detailed scans to fit the data. For a linear resonant system in a circuit, there are well-established models to describe it.34,35 In this paper, we use the following model35 to fit the experimental data:
where η is a complex constant accounting for the gain and phase shift through the system, the constant τ accounts for the cable delay related with the path length of the cables, Qi (Qc) is the internal (external) quality factor, fr is the resonant frequency, Q is the total quality factor, and ϕ0 quantifies the impedance mismatch.
FIG. 3.

Measurement of the micronwave resonators. (a) Fitting the measurement data into the Lorentz curve. (b) Nonlinear response of a micronwave resonator. With the drive power swept from −105.5 to −81.5 dBm, the resonator shows characteristic behavior of the cubic Duffing oscillator. (c) Extracted Kerr shift per photon (K/2π = −7.32 Hz) from the scattering parameter measurement via increasing drive powers.

FIG. 3.

Measurement of the micronwave resonators. (a) Fitting the measurement data into the Lorentz curve. (b) Nonlinear response of a micronwave resonator. With the drive power swept from −105.5 to −81.5 dBm, the resonator shows characteristic behavior of the cubic Duffing oscillator. (c) Extracted Kerr shift per photon (K/2π = −7.32 Hz) from the scattering parameter measurement via increasing drive powers.

Close modal

Figure 3(a) shows that the measurement data for the nonlinear resonators at low on-chip power are in very good agreement with the Lorentz curve. As shown in Table II, Lk,theory is the kinetic inductance estimated by Lk=(lw)RπΔ and Lk,VNA is the kinetic inductance extrapolated from VNA measurement data by substituting the capacitance simulation value. The internal quality factor of the two resonators is greater than 104 despite the wafer not undergoing any treatment. For the external quality factor, the design of our resonators exhibits weak coupling. The average photon number36 in the resonators can be calculated by n=4Q2PinQcωr2, where Pin is the on-chip power and ωr is the resonator angular frequency.

TABLE II.

Properties of NbTiN KI-resonators.

Samplefr (GHz)Lk,theory (nH)Lk,VNA (nH)K/2π (Hz)Qi (×104)Qc (×104)
R1 7.0365 1.796 1.248 −7.32 1.4 11.2 
R2 8.0844 1.530 0.965 −16.9 1.1 5.3 
R1a 7.0260 1.796 1.252 −11.65 2.1 11.2 
Samplefr (GHz)Lk,theory (nH)Lk,VNA (nH)K/2π (Hz)Qi (×104)Qc (×104)
R1 7.0365 1.796 1.248 −7.32 1.4 11.2 
R2 8.0844 1.530 0.965 −16.9 1.1 5.3 
R1a 7.0260 1.796 1.252 −11.65 2.1 11.2 
a

The sample was aged in the air atmosphere for three months, and the temperature of the cabinet was kept at 25 °C.

In circuit electrodynamics, the Duffing-type nonlinear resonator can be described by the Hamiltonian36 
where â (â) is the bosonic creation (annihilation) operator and K is the Kerr nonlinearity. For our device, the Kerr nonlinearity originates in kinetic inductance of the superconducting films, and it can be written8 as K=34ω2N0Δ02α2V, where N0 is the density of electron states at the Fermi energy, Δ0 is the superconducting energy gap at 0 K, the fraction α = Lk/L refers to the ratio of kinetic inductance to total inductance, and V accounts for the volume of the micronwire. This formula shows that we can enhance the Kerr nonlinearity by reducing the volume of the wire shape inductors for the selected material, and it is important for the regulation of Kerr nonlinearity in devices such as kinetic inductance parametric amplifiers, qubits, and sensors. For a selected film with a thickness of tens of nanometers, the Kerr nonlinearity of the wire-type inductor can increase by a factor of two orders when its wire width and length both reduced ten-fold to maintain the given inductance. This is because the small wire volume leads to a higher energy density while the Kerr nonlinearity scales the frequency shift due to a single photon. In particular, for two-dimensional and one-dimensional superconductors, the superconducting physical properties will also change, typically with a decrease in Tc and an increase in resistivity, which both lead to an enhancement of the Kerr nonlinearity.

In order to extract the Kerr nonlinearity coefficients, we performed power scans of the devices to be able to estimate the average microwave photon number in the resonator n and resonant frequency shift Δf, and the linear fitting result of the Kerr nonlinearity coefficients is shown in Fig. 3(c). Furthermore, we took the same measurement after three months of aging in the open air, keeping the cabinet temperature at 25 °C. We observed a systematic frequency shift dip of about 10 MHz (see Table II) toward lower frequency for resonator R1. The frequency shift could result from the oxidation on the surface of the NbTiN films, which would lead to a reduction in the thickness of the superconducting layer and hence an increase in the kinetic inductance. After exposure to air for three months, the oxide layer on the surface of the aluminum film underwent full oxidation and had reduced lattice defects that account for the increase in Qi.

The normalized frequency ratio is shown in Fig. 4(a), defined as Ω = ωd/ωr, where ωd represents the driving frequency. This graph depicts the bifurcation response, which is theoretically derived from the cubic Duffing oscillator model.20 The direction of the sweep, either forward or backward, influences the distinct dynamical trajectories observed. The dotted line corresponds to unstable solutions, and the red (blue) solid line occurs during jumping when sweeping forward (backward). We performed forward and backward continuous wave sweep in the device. The transmission response curve in Fig. 4(b) shows the hysteresis region at the −64dBm on-chip power.

FIG. 4.

(a) Analytical solution of amplitude-frequency response to the cubic Duffing equation in a hysteresis region. (b) The bifurcation response curve was measured at −64 dBm on-chip power; the inset shows the phase–frequency curve.

FIG. 4.

(a) Analytical solution of amplitude-frequency response to the cubic Duffing equation in a hysteresis region. (b) The bifurcation response curve was measured at −64 dBm on-chip power; the inset shows the phase–frequency curve.

Close modal

In conclusion, we first study the superconducting characterization of the NbTiN films. Reducing the target current and increasing the N2 flow rate can result in a higher resistivity, which corresponds to higher kinetic inductance. Then, we designed and fabricated nonlinear superconducting lumped resonators with strong coupling. The nonlinear resonators demonstrated a high internal quality factor (Qi ≥ 104) and Kerr coefficients (K/2π ∼10 Hz). By reducing the inductor linewidth to tens of nanometers, the Kerr nonlinearity is expected to reach several hundred kHz. The response curve displays the bifurcation hysteresis region at high power. This can help us understand more about the nonlinear mechanism of the KI-resonator. Furthermore, the KI-resonator is a promising candidate for high saturation power parametric amplifiers due to the high intrinsic quality factor and wide Kerr nonlinearity range.

See supplementary material for the AFM images of the roughness of the NbTiN films and the SEM images of the patterning of the microwires.

This work was supported by the Strategic Priority Research Program of the Chinese Academy of Sciences, Grant No. XDB0670301, and the Shanghai Technology Innovation Action Plan Integrated Circuit Technology Support Program (Grant No. 22DZ1100200).

The authors have no conflicts to disclose.

M.Y. and X.H. contributed equally to this work.

Ming Yang: Conceptualization (lead); Data curation (equal); Formal analysis (equal); Investigation (lead); Methodology (equal); Validation (lead); Visualization (lead); Writing – original draft (lead); Writing – review & editing (equal). XiaoLiang He: Investigation (supporting); Methodology (supporting); Software (equal); Validation (equal); Visualization (equal); Writing – review & editing (equal). WanPeng Gao: Data curation (equal); Formal analysis (equal); Investigation (equal). JunFeng Chen: Investigation (supporting); Methodology (supporting). Yu Wu: Data curation (supporting); Investigation (equal); Methodology (supporting). XiaoNi Wang: Data curation (supporting); Investigation (supporting). Gang Mu: Conceptualization (supporting); Formal analysis (equal); Investigation (supporting); Project administration (supporting). Wei Peng: Funding acquisition (equal); Project administration (equal). ZhiRong Lin: Funding acquisition (equal); Project administration (equal); Supervision (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
J.
Zmuidzinas
, “
Superconducting microresonators: Physics and applications
,”
Annu. Rev. Condens. Matter Phys.
3
,
169
214
(
2012
).
2.
A.
Wallraff
,
D. I.
Schuster
,
A.
Blais
,
L.
Frunzio
,
R.-S.
Huang
,
J.
Majer
,
S.
Kumar
,
S. M.
Girvin
, and
R. J.
Schoelkopf
, “
Strong coupling of a single photon to a superconducting qubit using circuit quantum electrodynamics
,”
Nature
431
,
162
167
(
2004
).
3.
S.
Danilin
and
M.
Weides
, “
Quantum sensing with superconducting circuits
,” arXiv:2103.11022 (
2021
).
4.
B.
Ho Eom
,
P. K.
Day
,
H. G.
LeDuc
, and
J.
Zmuidzinas
, “
A wideband, low-noise superconducting amplifier with high dynamic range
,”
Nat. Phys.
8
,
623
627
(
2012
).
5.
Y.
Krupko
,
V.
Nguyen
,
T.
Weißl
,
É.
Dumur
,
J.
Puertas
,
R.
Dassonneville
,
C.
Naud
,
F.
Hekking
,
D.
Basko
,
O.
Buisson
et al, “
Kerr nonlinearity in a superconducting Josephson metamaterial
,”
Phys. Rev. B
98
,
094516
(
2018
).
6.
C.
Eichler
and
A.
Wallraff
, “
Controlling the dynamic range of a Josephson parametric amplifier
,”
EPJ Quantum Technol.
1
,
2
19
(
2014
).
7.
N.
Kirsh
,
E.
Svetitsky
,
S.
Goldstein
,
G.
Pardo
,
O.
Hachmo
, and
N.
Katz
, “
Linear and nonlinear properties of a compact high-kinetic-inductance WSi multimode resonator
,”
Phys. Rev. Appl.
16
,
044017
(
2021
).
8.
C.
Joshi
,
W.
Chen
,
H. G.
LeDuc
,
P. K.
Day
, and
M.
Mirhosseini
, “
Strong kinetic-inductance Kerr nonlinearity with titanium nitride nanowires
,”
Phys. Rev. Appl.
18
,
064088
(
2022
).
9.
M.
Xu
,
R.
Cheng
,
Y.
Wu
,
G.
Liu
, and
H. X.
Tang
, “
Magnetic field-resilient quantum-limited parametric amplifier
,”
PRX Quantum
4
,
010322
(
2023
).
10.
P.
Winkel
,
K.
Borisov
,
L.
Grünhaupt
,
D.
Rieger
,
M.
Spiecker
,
F.
Valenti
,
A. V.
Ustinov
,
W.
Wernsdorfer
, and
I. M.
Pop
, “
Implementation of a transmon qubit using superconducting granular aluminum
,”
Phys.Rev. X
10
,
031032
(
2020
).
11.
P. K.
Day
,
H. G.
LeDuc
,
B. A.
Mazin
,
A.
Vayonakis
, and
J.
Zmuidzinas
, “
A broadband superconducting detector suitable for use in large arrays
,”
Nature
425
,
817
821
(
2003
).
12.
S.
Frasca
,
I. N.
Arabadzhiev
,
S. B.
de Puechredon
,
F.
Oppliger
,
V.
Jouanny
,
R.
Musio
,
M.
Scigliuzzo
,
F.
Minganti
,
P.
Scarlino
, and
E.
Charbon
, “
NbN films with high kinetic inductance for high-quality compact superconducting resonators
,”
Phys. Rev. Appl.
20
,
044021
(
2023
).
13.
T. M.
Bretz-Sullivan
,
R. M.
Lewis
,
A. L.
Lima-Sharma
,
D.
Lidsky
,
C. M.
Smyth
,
C. T.
Harris
,
M.
Venuti
,
S.
Eley
, and
T.-M.
Lu
, “
High kinetic inductance NbTiN superconducting transmission line resonators in the very thin film limit
,”
Appl. Phys. Lett.
121
,
052602
(
2022
).
14.
N.
Iossad
,
Metal Nitrides for Superconducting Tunnel Detectors
(
Delft University of Technology
,
2002
).
15.
R.
Barends
,
N.
Vercruyssen
,
A.
Endo
,
P.
De Visser
,
T.
Zijlstra
,
T.
Klapwijk
,
P.
Diener
,
S.
Yates
, and
J.
Baselmans
, “
Minimal resonator loss for circuit quantum electrodynamics
,”
Appl. Phys. Lett.
97
,
023508
(
2010
).
16.
M. R.
Vissers
,
R. P.
Erickson
,
H.-S.
Ku
,
L.
Vale
,
X.
Wu
,
G.
Hilton
, and
D. P.
Pappas
, “
Low-noise kinetic inductance traveling-wave amplifier using three-wave mixing
,”
Appl. Phys. Lett.
108
,
012601
(
2016
).
17.
P.
Hu
,
H.
Li
,
L.
You
,
H.
Wang
,
Y.
Xiao
,
J.
Huang
,
X.
Yang
,
W.
Zhang
,
Z.
Wang
, and
X.
Xie
, “
Detecting single infrared photons toward optimal system detection efficiency
,”
Opt. Express
28
,
36884
36891
(
2020
).
18.
V.
Manucharyan
,
E.
Boaknin
,
M.
Metcalfe
,
R.
Vijay
,
I.
Siddiqi
, and
M.
Devoret
, “
Microwave bifurcation of a Josephson junction: Embedding-circuit requirements
,”
Phys. Rev. B
76
,
014524
(
2007
).
19.
A. H.
Nayfeh
and
D. T.
Mook
,
Nonlinear Oscillations
(
John Wiley & Sons
,
2008
).
20.
I.
Kovacic
and
M. J.
Brennan
,
The Duffing Equation: Nonlinear Oscillators and Their Behaviour
(
John Wiley & Sons
,
2011
).
21.
R.
Vijay
,
M.
Devoret
, and
I.
Siddiqi
, “
Invited review article: The Josephson bifurcation amplifier
,”
Rev. Sci. Instrum.
80
,
111101
(
2009
).
22.
M.
Mück
,
C.
Welzel
, and
J.
Clarke
, “
Superconducting quantum interference device amplifiers at gigahertz frequencies
,”
Appl. Phys. Lett.
82
,
3266
3268
(
2003
).
23.
R.
Vijayaraghavan
, “
Josephson bifurcation amplifier: Amplifying quantum signals using a dynamical bifurcation
,” (
2008
) https://api.semanticscholar.org/CorpusID:118581731.
24.
I.
Barth
,
L.
Friedland
,
O.
Gat
, and
A.
Shagalov
, “
Quantum versus classical phase-locking transition in a frequency-chirped nonlinear oscillator
,”
Phys. Rev. A
84
,
013837
(
2011
).
25.
O.
Naaman
,
J.
Aumentado
,
L.
Friedland
,
J.
Wurtele
, and
I.
Siddiqi
, “
Phase-locking transition in a chirped superconducting Josephson resonator
,”
Phys. Rev. Lett.
101
,
117005
(
2008
).
26.
A. J.
Annunziata
,
D. F.
Santavicca
,
L.
Frunzio
,
G.
Catelani
,
M. J.
Rooks
,
A.
Frydman
, and
D. E.
Prober
, “
Tunable superconducting nanoinductors
,”
Nanotechnology
21
,
445202
(
2010
).
27.
S. S.
Yeram
,
S.
Bhakat
,
S. S.
Dash
, and
A.
Pal
, “
Structural transitions in superconducting NbTiN thin films
,” arXiv:2311.14154 (
2023
).
28.
P.
Pratap
,
L.
Nanda
,
K.
Senapati
,
R. P.
Aloysius
, and
V.
Achanta
, “
Optimization of the superconducting properties of NbTiN thin films by variation of the N2 partial pressure during sputter deposition
,”
Supercond. Sci. Technol.
36
,
085017
(
2023
).
29.
J.
Bardeen
,
L. N.
Cooper
, and
J. R.
Schrieffer
, “
Theory of superconductivity
,”
Phys. Rev.
108
,
1175
(
1957
).
30.
M.
Tinkham
,
Introduction to Superconductivity
(
Courier Corporation
,
2004
).
31.
E. F.
Driessen
,
P.
Coumou
,
R.
Tromp
,
P.
De Visser
, and
T.
Klapwijk
, “
Strongly disordered TiN and NbTiN s-wave superconductors probed by microwave electrodynamics
,”
Phys. Rev. Lett.
109
,
107003
(
2012
).
32.
M.
Burdastyh
,
S.
Postolova
,
T.
Proslier
,
S.
Ustavshikov
,
A.
Antonov
,
V.
Vinokur
, and
A. Y.
Mironov
, “
Superconducting phase transitions in disordered NbTiN films
,”
Sci. Rep.
10
,
1471
(
2020
).
33.
E. B.
Rosa
,
The Self and Mutual Inductances of Linear Conductors, 80
(
US Department of Commerce and Labor Bureau of Standards
,
1908
).
34.
S.
Probst
,
F.
Song
,
P. A.
Bushev
,
A. V.
Ustinov
, and
M.
Weides
, “
Efficient and robust analysis of complex scattering data under noise in microwave resonators
,”
Rev. Sci. Instrum.
86
,
024706
(
2015
).
35.
J.
Gao
,
The Physics of Superconducting Microwave Resonators
(
California Institute of Technology
,
2008
).
36.
M. J.
Reagor
,
Superconducting Cavities for Circuit Quantum Electrodynamics
(
Yale University
,
2016
).