Next Article in Journal
Pulse-Height and 2-D Charge-Spread Single-Pixels Studies on a LuYAP:Ce Scintillation Array
Next Article in Special Issue
Impact of Superconductors’ Properties on the Measurement Sensitivity of Resonant-Based Axion Detectors
Previous Article in Journal
An Active Plasma Beam Dump for EuPRAXIA Beams
Previous Article in Special Issue
New Projections for Dark Matter Searches with Paleo-Detectors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Josephson Junctions as Single Microwave Photon Counters: Simulation and Characterization

by
Alessio Rettaroli
1,2,*,
David Alesini
1,
Danilo Babusci
1,
Carlo Barone
3,4,
Bruno Buonomo
1,
Matteo Mario Beretta
1,
Gabriella Castellano
5,
Fabio Chiarello
1,5,
Daniele Di Gioacchino
1,
Giulietto Felici
1,
Giovanni Filatrella
4,6,
Luca Gennaro Foggetta
1,
Alessandro Gallo
1,
Claudio Gatti
1,
Carlo Ligi
1,
Giovanni Maccarrone
1,
Francesco Mattioli
1,5,
Sergio Pagano
3,4,
Simone Tocci
1 and
Guido Torrioli
1,5
1
INFN, Laboratori Nazionali di Frascati, 00044 Frascati, Italy
2
Dipartimento di Matematica e Fisica, Università di Roma Tre, 00154 Roma, Italy
3
Dipartimento di Fisica E.R. Caianiello, Università di Salerno, 84084 Fisciano, Italy
4
INFN, Gruppo Collegato di Salerno, 84084 Fisciano, Italy
5
Istituto di Fotonica e Nanotecnologie CNR, 00156 Roma, Italy
6
Dipartimento di Scienze e Tecnologie, Università del Sannio, 82100 Benevento, Italy
*
Author to whom correspondence should be addressed.
Submission received: 17 May 2021 / Revised: 7 July 2021 / Accepted: 12 July 2021 / Published: 16 July 2021
(This article belongs to the Special Issue Innovative Experimental Techniques for Direct Dark Matter Detection)

Abstract

:
Detection of light dark matter, such as axion-like particles, puts stringent requirements on the efficiency and dark-count rates of microwave-photon detectors. The possibility of operating a current-biased Josephson junction as a single-microwave photon-detector was investigated through numerical simulations, and through an initial characterization of two Al junctions fabricated by shadow mask evaporation, done in a dilution refrigerator by measuring escape currents at different temperatures, from 40 mK up to the Al transition temperature. The escape dynamics of the junctions were reproduced in the simulation, including the dissipative effects. Inhibition of thermal activation was observed, leaving the macroscopic quantum tunneling as the dominant effect well beyond the crossover temperature.

1. Introduction

1.1. Detection of Axions

In the 70s, an extension of the standard model of particle physics was advanced to explain the absence of CP violation in the strong interaction (strong CP problem) [1,2]. The theory predicts the existence of additional particles called axions [3,4]—sub-eV particles with feeble interactions with ordinary matter, which could be abundantly produced non-thermally in the early universe [5]. These characteristics make them good candidates for explaining the composition of cold dark matter (CDM) in the universe [6,7,8], an issue that modern cosmology and particle physics are still tackling.
Most of the experiments are sensitive to the axion–photon coupling constant g a γ γ , which is proportional to the axion mass m a . The region of interest for QCD axions is the yellow band in the m a - g a γ γ plane shown in Figure 1.
Axions are an example of weakly interacting slim particles (WISPs), a category of particles emerging from many extensions of the standard model. WISPs as axion-like particles (ALPs) have a more general relationship between m a and g a γ γ .
The most exploited detection principle of axions and ALPs is the haloscope proposed by Sikivie [14,15], whose description is addressed in a review: [16]. Examples of running experiments, also shown by the vertical lines in Figure 1, are ADMX [17,18], HAYSTAC [19], CAPP-8T [20], CAPP-9T [21], ORGAN [22], and QUAX [23,24,25,26]; other proposed experiments include RADES [27,28], MADMAX [29], BRASS [30], and KLASH [31,32,33].
Since very tiny signals are involved in the game, searches of axions and ALPs will benefit from the latest developments in microwave quantum technologies, such as single microwave-photon detectors based on superconducting elements. In fact, as discussed in the next section, photon detectors may have better signal-to-noise ratios than quantum amplifiers, especially if they have low dark-count rates.

1.2. Single Photon Counters as Detectors

In the last two decades, attention to single photon detection has grown due to the demanding applications in quantum computing. Several techniques have been developed to detect single photons, such as quantum non-demolition measurements [34,35,36,37,38,39], switching detectors [40,41,42], hot electron detectors [43,44], and quantum dot detectors [45]. Belonging to switching detectors in the microwave range of radiation are Josephson junction (JJ)-based photon counters [41] and transmon qubits [35,46].
The best performances of currently running axion experiments were achieved with quantum-limited devices, such as superconducting quantum interference devices (SQUIDs) and Josephson parametric amplifiers (JPAs) [17,18,19,26,47], but efforts have been put toward counting single photons from axion interactions [42,46,48,49]. The reason is clear from [50]: working at higher frequencies while having a good scan rate at the same time requires single photon counters, which have better noise performances with respect to linear amplifiers above about 10 GHz. As discussed in [42], dark counts of single photon counters may be a limiting factor: in dark matter axion searches, a single photon counter will be competitive with a quantum-limited amplifier if the dark count ν d c is below the intrinsic axion width Δ ν a ν a / 10 6 corresponding to 10 kHz at 10 GHz. Devices based on qubits have dark-counts rates ν d c p e r r / t q u b i t where p e r r 10 % is a typical qubit readout error and t q u b i t 10 μ s is a typical qubit decoherence time, corresponding to ν d c 10 kHz. On the contrary, switching devices, such as those based on current-biased Josephson junctions (CBJJ) [41,42,48], have the potential to reach dark counts at mHz rates. However, previous realization of a photon counter based on CBJJ [41] was operated in a regime with far higher dark counts, and no clear evidence exists that dark counts can be lowered without affecting the device efficiency. Within the SIMP project [51], we are developing a microwave single-photon detector based on CBJJ to couple to a haloscope, with the aim of probing the axion existence. In this paper we describe the device and its response to photon excitations obtained from simulations of different experimental conditions, showing that single photons can induce the switch of a JJ. We describe the first step in the characterization of a JJ with parameters optimized to detect photons originating from an axion conversion and derive the dark counts expected for a photon detector. We finally discuss the improvement needed to keep both high efficiency and a low dark-count rate.

2. A CBJJ as a Photon Detector

The electrodynamics of a CBJJ is accurately described in terms of the resistively shunted junction (RSJ) model [52], an equivalent electrical model whose circuit components are related to specific junction physical characteristics. With reference to Figure 2, the capacitor C J represents the capacitance between the junction electrodes, the resistor R J represents the tunneling of normal electrons (quasiparticles), and the component J J represents the tunneling of superconductive electrons (Cooper pairs). The components I s , I b , and I n represent current sources that will be detailed below. The current I J through the component J J and the corresponding voltage drop V J are related to φ , the phase difference between the macroscopic wavefunctions of the two superconductors, by the Josephson equations:
I J = I 0 sin φ ,
V J = 2 e d φ d t .
I 0 is the maximum Cooper pair current that can flow through the J J component, and and e are the reduced Planck constant and electron charge, respectively. A stationary phase, and correspondingly, a zero average voltage is possible, according to the Josephson equations, Equations (1) and (2), when a dc current below I 0 flows in the J J element. Depending on the values of the junction parameters ( I 0 , R J , C J ), a finite-voltage state is also possible. If the dc bias current overcomes I 0 , the phase cannot be stationary and a finite-voltage state is the only possibility. From the experimental point of view, this behavior is reflected in the occurrence of hysteresis in the current–voltage characteristic of the JJ, for certain dc bias current intervals, where two voltage states (zero and finite-valued) can be observed. Such hysteresis is at the basis of the use of a JJ as a detector. It is worth noting that in a typical tunnel-type JJ, the resistance of R J strongly depends both on voltage and temperature. However, as the overall effect of the resistor is to introduce dissipation in the system, its nonlinearity is often not considered in the presence of moderate or weak damping. Another important effect of the quasiparticle current, modeled by R J , is the presence of random charge fluctuations, which can be represented, using the fluctuation dissipation theorem, by a noise current source, indicated with I n in Figure 2, whose spectral power density is assumed to be frequency independent (Johnson noise) [53].
In order to use a JJ as a detector, it is biased with a dc current just below I 0 through a suitable source, indicated with I b in Figure 2. The occurrence of an external additional current (i.e., the signal to be detected) can induce the switching of the junction from the zero to the finite-voltage state. Of course, current noise can induce a switching, contributing to false alarms or dark counts in detector language. In the case of a simple CBJJ, the phase difference φ can be considered as a particle moving in a one-dimensional tilted cosine potential [54]; see Figure 3. The tilt of the potential corresponds to the normalized bias current flowing through the junction. The zero-voltage state corresponds to the confinement of the particle in a potential well of barrier height Δ U [54], where it can oscillate at the characteristic plasma frequency, ω J . Given enough energy, the particle can escape from this metastable state and roll down the potential slope, giving rise to a finite-voltage state [54]. At temperatures such that k B T ω J , the escape rate is dominated by the thermal activation (TA) process [55]:
Γ T ω J 2 π exp Δ U k B T ,
whereas at lower temperatures, it is dominated by macroscopic quantum tunneling (MQT) [55]:
Γ q ω J 2 π exp 7.2 Δ U ω J
which provides an irreducible contribution to the dark-count rate.

3. Device Simulation

We investigated the dynamics of a CBJJ in the presence of a microwave pulse, representing the process of single-photon absorption, by means of numerical simulations of the model equations [56]. We estimated in this way the parameters needed to make the junction switch in the presence of a signal.

3.1. Isolated CBJJ

To model a weak microwave field (weak because carrying few photons) coupled to the junction, we consider a deterministic current source, I s in Figure 2, generating a properly shaped current pulse, representing the absorption of a single microwave photon. The following equations refer to the circuit of Figure 2a and describe the dynamics of an isolated junction where we imagine the current source directly coupled to the junction.
By denoting I J as the current through the J J element, I R J as the current through the resistor, and I C J as the current through the capacitor, we write the following current balance equation:
I C J + I R J + I J = I b + I s t + I n t .
By using the constitutive relations of the resistor and capacitor, and the Josephson relations, we obtain the following second-order differential equation:
C J 2 e d 2 φ d t 2 + 1 R J 2 e d φ d t + I 0 sin φ = I b + I s t + I n t ,
which, due to the presence of the stochastic noise term, is a Langevin equation [57]. By defining a normalized time τ = ω J t , where
ω J = 2 e I 0 / C J
is the Josephson plasma frequency, Equation (6) is re-written as:
d 2 φ d τ 2 + α d φ d τ + sin φ = γ b + γ s τ + γ n ( τ ) ,
where the parameters are defined as
α = 1 R J C J ω J , γ b = I b I 0 , γ s = I s I 0 .
The statistical properties of the noise term γ n are:
γ n ( τ ) = 0 , γ n ( τ ) , γ n ( τ ) = 4 D δ ( τ τ ) ,
where D = k B T ω J / R J I 0 2 is the normalized noise intensity ( k B is the Boltzmann constant and T the physical temperature), δ ( ) is the Dirac delta function, and the parentheses represent ensemble averages.

3.2. CBJJ Coupled to an RC Circuit

If the JJ is small in size, its capacitance and conductance can be very small too, thereby making it possible for external parasitic effects to modify its dynamics. To investigate these effects, we considered a model of a JJ loaded with a series parasitic RC circuit; see Figure 2c. In this case, Equation (5) becomes:
I C J + I R J + I J + I R p = I b + I s t + I n t ,
where I R p is the current flowing in the parasitic resistor R p . The JJ voltage V J is related to I R p by:
V J = R p I R p + 1 C p I R p d t .
By repeating the same procedure as before, we obtain the following two normalized differential equations:
d 2 φ d τ 2 + α d φ d τ + sin φ = γ b + γ s τ + γ n ( τ ) + γ R ,
d γ R d τ + α R C γ R + α i n t d 2 φ d τ 2 = 0 ,
where
α R C = 1 R p C p ω J , α i n t = 1 R p C J ω J , γ R = I R p I 0 .

3.3. CBJJ Coupled to a Transmission Line

The simplest design for a photon detector operating at microwave frequency based on a JJ is a transmission line (TL), such as a coplanar waveguide, terminated with a JJ. When the JJ is coupled to a TL with a characteristic impedance Z T L , as shown in Figure 2b, signal reflection back to the TL has to be considered. The overall effect is modeled by a modified effective junction resistance, given by the values of the parallel resistors R J and Z T L . In Figure 2b, the current source I s is given by an ideal generator I g with an impedance R g in parallel. The latter has to be equal to Z T L to ensure impedance matching. To relate the simulation results to this particular case, we compare Equation (6) with the equation for the flux variable ϕ = φ ( ϕ 0 / 2 π ) in a TL terminated by a parallel LC (a linearized JJ) [58]:
C J ϕ ¨ + ϕ L J + 1 Z T L ϕ ˙ = 2 1 Z T L ϕ ˙ i n = 2 I i n ,
where L J is the Josephson inductance and the term 1 / Z T L models the aforementioned signal reflection on the TL. Moreover, the current I s has to be interpreted as twice the input current I i n . The input peak-current due to a single photon on the waveguide with a Gaussian wavepacket of time duration σ t is
I peak 2 = ω J Z T L 2 2 π σ t .
Then, the amplitude of the signal current I s in Equation (6) corresponding to a single photon is:
I s photon = 2 ω J Z T L 2 2 π σ t .
For a 10 GHz photon with σ t = 600 ps on a 50 Ω TL, this corresponds to about 26 nA.
We ran several simulations with the CBJJ either isolated or coupled to a TL excited by Gaussian current pulses of about 600 ps in length. We considered values of the critical current ranging from few hundred nanoamperes to a few microamperes, and junction capacitance ranging from a fraction of a picoFarad to a few picoFarads. The ratio I b / I 0 , about 0.8, was set in such a way as to keep the estimated rate of the MQT from the ground level by a few hertz at most. With the isolated junction we observed switching currents I s w i t c h between 20 and 50 nA corresponding to about 1–4 photons, estimated as N γ = ( I s w i t c h / I s photon ) 2 . On the contrary, with the junction coupled to a 50 Ω TL, these values increased to a range between 200 to 300 nA corresponding to about 100 photons. The increase of number of photons is directly proportional to the relaxation rate of the junction γ T L = ω J / Q = ω J Z J / Z T L = 2 π / Z T L C J , where Z J = L J / C J , highlighting the need for a proper circuit for matching the CBJJ to the TL. Matching of a device to a TL is discussed in [39,59] where the detection efficiency is expressed as
P R = 4 γ T L γ sw ( γ T L + γ sw ) 2 .
γ sw is the switching rate of the excited JJ to the resistive state. Perfect matching condition γ T L = γ sw implies that the escape to the resistive state must be as fast as the relaxation of the junction to the ground level. This prevents us from using MQT transitions from the first excited level as a detection mechanism: an escape caused by MQT from the first excited level, E 1 in Figure 3, equal to γ T L 1–10 GHz, would induce a dark-count rate due to escape from the ground level of about γ 0 10 3 γ sw 1–10 MHz. On the contrary, while keeping low dark-count rates, two or more photons may induce the escape if absorbed by the junction within the relaxation time 1 / γ T L .

4. Measurements

The study of the escape mechanisms of a CBJJ allow understanding and characterizing the escape rates in the absence of excitation signals, and therefore, of the dark counts. Much literature on measurements and interpretations of the escape processes has been published and underlines different dynamics, such as the TA regime, MQT, and phase diffusion (PD) with multiple retrapping processes [54,60,61,62,63,64,65,66,67,68]. The experimental conditions to highlight the diverse regimes and transitions between them have been defined. The regimes will depend on the comparison between the Josephson energies ( E J ), related to the Josephson critical current I 0 ; the Coulomb energy E C , depending on the capacity C J ; the plasma frequency ω J proportional to I 0 / C J ratio; the quality factor Q = ω J R J C J , proportional to the losses; the potential tilt E J I / I 0 , controlled by the bias current I; and temperature [54,60,61,62,63,64,65,66,67,68]. This rich phenomenology of the escape process makes it ideal for a DC characterization of a JJ. In general, the typical escape dynamic’s experimental setup consists of a slow ramping of the bias current across the junction up to the value of the critical current [61]. This ensures both that the JJ stays at the temperature of the cryostat thermal bath at milli-Kelvins, and that the superconducting state does not dissipate during the escape measurements before the final switching. Then, the data analyzed are the switching current distributions produced after countless repetitions of the process. Finally, measurements as a function of temperature can be a diagnostic tool for determining the processes of TA, tunneling, and phase diffusion regimes that regulate the escape process.

4.1. Fabrication Parameters from Simulations

The JJs we tested were fabricated at IFN-CNR by shadow mask evaporation (Figure 4), with electron beam lithography on a copolymer/PMMA bilayer (thickness of about 1 μ m for each layer) over a Si substrate, with two evaporation angles of Al at 25 and 90 (thickness of about 30 nm), and with 5 min of oxidation at 5 mbar (expected thickness in the order of 1 nm). Junctions with different areas were fabricated: 2 μ m × 2 μ m ( I 0 300 nA and C 200 fF, expected from design parameters) and 2 μ m × 4 μ m ( I 0 600 nA and C 400 fF).

4.2. Setup

We tested the devices in a range of temperatures between 50 mK and about 1 K in a Leiden Cryogenics MCK 50–100 dilution refrigerator. We used a 4-terminal measurement scheme (Figure 5), with a bias resistance R = 14.63 k Ω at room temperature, twisted/shielded phosphor bronze cryogenic lines (about 25 Ω per line), and EMI low-pass filters at room temperature (bandpass 600 kHz). An Agilent 33220A waveform generator was used, together with two low noise preamplifiers EGG 5113 and a NI USB 6366 acquisition board.
This experimental setup allowed us to acquire current-voltage characteristics. We measured a gap Δ 200 μ V and critical currents within a factor 2 with respect to the estimated ones (Figure 6).
In order to measure the escape probabilities, we applied a sawtooth waveform with frequency f = 314 Hz, amplitude I p p = 670 nA and offset I o f f = 250 nA. In each period, during the slow growth (with slope d I / d t = 210 μ A / s ), sudden transition to the voltage state occurred at random values of current I c . In the fast decrease region, the junction resets to the superconducting state. We acquired a series of N = 5000 transition values I c by using the voltage jumps as trigger signals. These values were arranged in a histogram to evaluate the escape probability P ( I ) . In Figure 7 we report the experimental escape probabilities at different temperatures ranging from about 40 to 850 mK. Data analysis is discussed in the next section.

5. Results and Discussion

5.1. Results

From the 5000 transition values I c we constructed escape-probability densities. Before doing this, we applied a fast Fourier transform to the measured I c series and identified a narrow peak at 100 Hz and minor peaks at lower frequencies, attributed to electronic noise in the current-measurement setup. We calculated the peaks’ frequencies, phases, and amplitudes from the Fourier analysis, and filtered out the noise from the original I c series. We thereby obtained a 20% improvement in resolution. We show the experimental switching distributions P ( I ) obtained after filtering the data for each temperature in Figure 7a,b—the smallest and the largest junctions, respectively.
We extracted the mean I c and width σ from each distribution as an arithmetic mean and a standard deviation. These are plotted as a function of temperature in Figure 8a,b, for the two junctions.
The mean critical current values (empty blue squares) follow qualitatively the typical BCS behavior [52] as a function of temperature. The widths (full red squares) present a peculiar feature. The constant behavior of σ at low temperatures was expected from quantum tunneling processes (Equation (4)), whereas its decrease at higher temperatures, above about 400 mK, suggests that TA (Equation (3)) is suppressed. We attribute the decreasing regime partially to the decrease of the critical current with increasing temperature, although a further contribution due to retrapping is not excluded.

5.2. Interpretations of Results

To investigate the effect of a dissipative environment on the escape by TA, we ran simulations of the JJ in parallel to the dissipative RC circuit shown in Figure 2c and described in Section 3.2. In our experimental setup, the parasitic circuit is represented by the dc lines carrying the bias current to the junction. The driving equations of the simulations were obtained from Equations (11)–(15) by setting the current source I s ( t ) to zero. Qualitatively, in the RCSJ model, energy dissipation induced a relaxation rate
Γ rel = 1 τ = 1 R C J = ω J Q ,
where R is the parallel between R J and R p . If relaxation is faster than the TA, the latter is suppressed:
ω J 2 π exp Δ U k B T   <   ω J Q ,
corresponding to the condition
R < 2 π Z J exp Δ U k B T .
Thus, even for temperatures such that k B T Δ U , TA is suppressed if R < 2 π e Z J 20 Z J . For the described junction, Z J is 50   Ω ; therefore, an external resistance equal to about 1 k Ω would dominate in parallel to R J = 100 k Ω , causing dissipation. On the contrary, the MQT process, which is not affected by dissipation, would become the dominant switching effect even at temperatures higher than the expected crossing temperature. This was confirmed by our simulations, whose results for different simulation parameters are shown in Figure 9. The inclusion of the RC circuit strongly modifies the qualitative behavior of I c and σ as a function of the temperature. In Figure 10 we show the comparison of the experimental results obtained for the 2 μ m × 2 μ m junction and results of a simulation run with parameters I 0 = 143 nA, C J = 0.2 pF, R J = 100 k Ω , C = 2 pF, and R p = 1 k Ω , where only MQT and dissipation were included. Both the simulated distributions of I c and σ are in good agreement with the measured ones. As anticipated, a further contribution from a retrapping process at a higher temperature could take into account the observed residual difference between the experimental and simulated values of σ .

5.3. Dark-Count Rate

Finally, we investigate the dark-count rate performances of the 2 μ m × 2 μ m junction to predict how it would behave if used as a single photon counter. Following [69], we calculated the escape rate from the current distribution of the JJ taken at the lowest temperature (50 mK). The result as a function of the bias current is shown in Figure 11, where the blue dots represent the escape rate values extracted from data. The dashed red line is the escape rate expected from MQT only. It reproduced the data for currents up to 120 nA, where the current distribution at 50 mK has its maximum (Figure 7a). The deviation at higher currents is attributable to a Gaussian resolution effect with standard deviation σ = 1.1 nA, as confirmed by the data’s agreement with the green solid line obtained by applying such a smearing to the theoretical current distribution of the MQT before deriving the escape rate.
To determine the working point for a photon detector based on a similar JJ ( I 0 = 140 nA, C J = 0.2 pF), we relied on the simulation of an “isolated” junction with R J = 5 k Ω driven by a resonant rf Gaussian pulse of frequency 6.5 GHz, duration σ t = 370 ps, and peak current I m a x = 21 nA corresponding to one photon. The simulated junction switched for a bias current I b 91 nA ( γ b 0.65 ) when the junction plasma frequency was about 6.5 GHz. By extrapolating the escape rate measured in our data (Figure 11) to the value γ b 0.65 , we estimated for the photon detector a dark-count rate of about 1 mHz.
It should be noted that in simulations with R J < 1.5 k Ω , dissipation inhibits junction switches, as observed in the TA process in our data and as expected from the discussion in Section 5.2. As already noted in Section 3.3, when a junction is coupled to a TL, the 50 Ω line impedance acts as a dissipative medium, preventing the junction from switching. Finally, when γ b 0.65 , the height of the potential well is estimated to be 28 GHz, corresponding to about four energy levels, showing that the current pulse must induce a strong modification of the potential to cause the junction switch.

6. Conclusions

We have investigated the escape mechanism of a CBJJ both with simulations and through measurements of the escape currents in two Al JJs. We highlighted the effect of the dissipative mechanism introduced by the current leads—in particular, the suppression of escape induced by TA. While this suppression leads to a reduction in the escape rates and to dark counts above the crossover temperature, it also implies a low quality factor with consecutive reductions in signal efficiency, and this must be taken into account in the design of a microwave photon counter. We in fact showed with simulations that a current pulse with peak current I m a x = 21 nA, corresponding to a single photon, induces the switch of an isolated junction of critical current I 0 = 140 nA when biased with a current I b 91 nA, and that the expected dark-count rate would be at the mHz-level. This value was confirmed by the escape rates extracted from our data. However, when the junction was not isolated but coupled to a TL, the simulation showed that switches were inhibited, and larger current pulses were required. The JJ must then be resonantly matched with the TL with a bandwith estimated to be in the order of 1 / R j C j 1 GHz.

Author Contributions

Formal analysis, A.R., C.B., F.C., D.D.G., G.F. (Giovanni Filatrella), C.G., C.L., G.M., S.P. and S.T.; investigation, A.R., D.A., D.B., B.B., M.M.B., G.C., F.C., D.D.G., G.F. (Giulietto Felici), L.G.F., A.G., C.G., C.L., G.M., F.M., S.P., S.T. and G.T.; resources, G.C., F.C., F.M. and G.T.; software, C.B., F.C., G.F. (Giovanni Filatrella) and S.P.; supervision, C.G.; writing—original draft, A.R., F.C., D.D.G., C.G., C.L., F.M. and S.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

We are grateful to G. Papalino and M. Iannarelli (LNF) for help with electronics and F. Tabacchioni (INAF) for cryogenics support. We would like to thank also A. Tudorache for his help during the short time spent at LNF.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Peccei, R.D.; Quinn, H.R. CP Conservation in the Presence of Pseudoparticles. Phys. Rev. Lett. 1977, 38, 1440–1443. [Google Scholar] [CrossRef] [Green Version]
  2. Peccei, R.D.; Quinn, H.R. Constraints imposed by CP conservation in the presence of pseudoparticles. Phys. Rev. D 1977, 16, 1791–1797. [Google Scholar] [CrossRef]
  3. Weinberg, S. A New Light Boson? Phys. Rev. Lett. 1978, 40, 223–226. [Google Scholar] [CrossRef]
  4. Wilczek, F. Problem of Strong P and T Invariance in the Presence of Instantons. Phys. Rev. Lett. 1978, 40, 279–282. [Google Scholar] [CrossRef]
  5. Sikivie, P. Axions. In Particle Dark Matter: Observations, Models and Searches; Bertone, G., Ed.; Cambridge University Press: Cambridge, UK, 2010. [Google Scholar]
  6. Preskill, J.; Wise, M.B.; Wilczek, F. Cosmology of the invisible axion. Phys. Lett. B 1983, 120, 127–132. [Google Scholar] [CrossRef] [Green Version]
  7. Abbott, L.; Sikivie, P. A cosmological bound on the invisible axion. Phys. Lett. B 1983, 120, 133–136. [Google Scholar] [CrossRef]
  8. Dine, M.; Fischler, W. The not-so-harmless axion. Phys. Lett. B 1983, 120, 137–141. [Google Scholar] [CrossRef]
  9. Kim, J.E. Weak-Interaction Singlet and Strong CP Invariance. Phys. Rev. Lett. 1979, 43, 103–107. [Google Scholar] [CrossRef]
  10. Shifman, M.A.; Vainshtein, A.I.; Zakharov, V.I. Can confinement ensure natural CP invariance of strong interactions? Nucl. Phys. B 1980, 166, 493–506. [Google Scholar] [CrossRef]
  11. Dine, M.; Fischler, W.; Srednicki, M. A simple solution to the strong CP problem with a harmless axion. Phys. Lett. B 1981, 104, 199–202. [Google Scholar] [CrossRef]
  12. Zhitnitsky, A.R. On Possible Suppression of the Axion Hadron Interactions. (In Russian). Sov. J. Nucl. Phys. 1980, 31, 260. [Google Scholar]
  13. O’Hare, C. cajohare/AxionLimits: AxionLimits. 2020. Available online: https://zenodo.org/record/3932430#.YO-ZTPkzZPY (accessed on 15 July 2021).
  14. Sikivie, P. Experimental Tests of the “Invisible” Axion. Phys. Rev. Lett. 1983, 51, 1415–1417. [Google Scholar] [CrossRef]
  15. Sikivie, P. Detection rates for “invisible”-axion searches. Phys. Rev. D 1985, 32, 2988–2991. [Google Scholar] [CrossRef]
  16. Irastorza, I.G.; Redondo, J. New experimental approaches in the search for axion-like particles. Prog. Part. Nucl. Phys. 2018, 102, 89–159. [Google Scholar] [CrossRef] [Green Version]
  17. Du, N.; Force, N.; Khatiwada, R.; Lentz, E.; Ottens, R.; Rosenberg, L.J.; Rybka, G.; Carosi, G.; Woollett, N.; Bowring, D.; et al. Search for Invisible Axion Dark Matter with the Axion Dark Matter Experiment. Phys. Rev. Lett. 2018, 120, 151301. [Google Scholar] [CrossRef] [Green Version]
  18. Braine, T.; Cervantes, R.; Crisosto, N.; Du, N.; Kimes, S.; Rosenberg, L.J.; Rybka, G.; Yang, J.; Bowring, D.; Chou, A.S.; et al. Extended Search for the Invisible Axion with the Axion Dark Matter Experiment. Phys. Rev. Lett. 2020, 124, 101303. [Google Scholar] [CrossRef] [Green Version]
  19. Backes, K.M.; Palken, D.A.; Al Kenany, S.; Brubaker, B.M.; Cahn, S.B.; Droster, A.; Hilton, G.C.; Ghosh, S.; Jackson, H.; Lamoreaux, S.K.; et al. A quantum enhanced search for dark matter axions. Nature 2021, 590, 238. [Google Scholar] [CrossRef]
  20. Choi, J.; Ahn, S.; Ko, B.; Lee, S.; Semertzidis, Y. CAPP-8TB: Axion Dark Matter Search Experiment around 6.7 μeV. arXiv 2020, arXiv:2007.07468. [Google Scholar]
  21. Jeong, J.; Youn, S.; Bae, S.; Kim, J.; Seong, T.; Kim, J.E.; Semertzidis, Y.K. Search for Invisible Axion Dark Matter with a Multiple-Cell Haloscope. Phys. Rev. Lett. 2020, 125, 221302. [Google Scholar] [CrossRef] [PubMed]
  22. McAllister, B.T.; Flower, G.; Ivanov, E.N.; Goryachev, M.; Bourhill, J.; Tobar, M.E. The ORGAN experiment: An axion haloscope above 15 GHz. Phys. Dark Universe 2017, 18, 67–72. [Google Scholar] [CrossRef] [Green Version]
  23. Crescini, N.; Alesini, D.; Braggio, C.; Carugno, G.; Di Gioacchino, D.; Gallo, C.S.; Gambardella, U.; Gatti, C.; Iannone, G.; Lamanna, G.; et al. Operation A Ferromagn. Axion Haloscope ma=58μeV. Eur. Phys. J. C 2018, 78, 703. [Google Scholar] [CrossRef] [Green Version]
  24. Crescini, N.; Alesini, D.; Braggio, C.; Carugno, G.; D’Agostino, D.; Di Gioacchino, D.; Falferi, P.; Gambardella, U.; Gatti, C.; Iannone, G.; et al. Axion Search with a Quantum-Limited Ferromagnetic Haloscope. Phys. Rev. Lett. 2020, 124, 171801. [Google Scholar] [CrossRef]
  25. Alesini, D.; Braggio, C.; Carugno, G.; Crescini, N.; D’Agostino, D.; Di Gioacchino, D.; Di Vora, R.; Falferi, P.; Gallo, S.; Gambardella, U.; et al. Galactic axions search with a superconducting resonant cavity. Phys. Rev. D 2019, 99, 101101. [Google Scholar] [CrossRef] [Green Version]
  26. Alesini, D.; Braggio, C.; Carugno, G.; Crescini, N.; D’Agostino, D.; Di Gioacchino, D.; Di Vora, R.; Falferi, P.; Gambardella, U.; Gatti, C.; et al. Search for invisible axion dark matter of mass ma = 43 μeV with the QUAX–aγ experiment. Phys. Rev. D 2021, 103, 102004. [Google Scholar] [CrossRef]
  27. Melcón, A.Á.; Cuendis, S.A.; Cogollos, C.; Díaz-Morcillo, A.; Döbrich, B.; Gallego, J.D.; Gimeno, B.; Irastorza, I.G.; Lozano-Guerrero, A.J.; Malbrunot, C.; et al. Axion searches with microwave filters: The RADES project. J. Cosmol. Astropart. Phys. 2018, 2018, 040. [Google Scholar] [CrossRef] [Green Version]
  28. Arguedas, C.S.; Álvarez, M.A.; Cogollos, C.; Díaz-Morcillo, A.; Döbrich, B.; Gallego, J.D.; Gimeno, B.; Irastorza, I.G.; Lozano-Guerrero, A.J.; Malbrunot, C.; et al. The 3 cavity prototypes of RADES, an axion detector using microwave filters at CAST. arXiv 2019, arXiv:1903.04323. [Google Scholar]
  29. Caldwell, A.; Dvali, G.; Majorovits, B.; Millar, A.; Raffelt, G.; Redondo, J.; Reimann, O.; Simon, F.; Steffen, F. Dielectric Haloscopes: A New Way to Detect Axion Dark Matter. Phys. Rev. Lett. 2017, 118, 091801. [Google Scholar] [CrossRef] [Green Version]
  30. BRASS: Broadband Radiometric Axion SearcheS. Available online: http://www.iexp.uni-hamburg.de/groups/astroparticle/brass/brassweb.htm (accessed on 15 July 2021).
  31. Alesini, D.; Babusci, D.; Di Gioacchino, D.; Gatti, C.; Lamanna, G.; Ligi, C. The KLASH Proposal. arXiv 2017, arXiv:1707.06010. [Google Scholar]
  32. Gatti, C.; Alesini, D.; Babusci, D.; Braggio, C.; Carugno, G.; Crescini, N.; Di Gioacchino, D.; Falferi, P.; Lamanna, G.; Ligi, C.; et al. The Klash Proposal: Status and Perspectives. In Proceedings of the 14th Patras Workshop on Axions, WIMPs and WISPs (AXION-WIMP 2018) (PATRAS 2018), Hamburg, Germany, 18–22 June 2018. [Google Scholar]
  33. Alesini, D.; Babusci, D.; Björkeroth, F.; Bossi, F.; Ciambrone, P.; Monache, G.D.; Di Gioacchino, D.; Falferi, P.; Gallo, A.; Gatti, C.; et al. KLASH Conceptual Design Report. arXiv 2019, arXiv:1911.02427. [Google Scholar]
  34. Gleyzes, S.; Kuhr, S.; Guerlin, C.; Bernu, J.; Deleglise, S.; Hoff, U.B.; Brune, M.; Raimond, J.-M.; Haroche, S. Quantum jumps of light recording the birth and death of a photon in a cavity. Nature 2007, 446, 297–300. [Google Scholar] [CrossRef] [Green Version]
  35. Schuster, D.I.; Houck, A.A.; Schreiner, J.A.; Wallraff, A.; Gambetta, J.M.; Blais, A.; Frunzion, A.; Majer, J.; Johnson, B.; Devoret, H.; et al. Resolving photon number states in a superconducting circuit. Nature 2007, 445, 515–518. [Google Scholar] [CrossRef] [Green Version]
  36. Johnson, B.R.; Reed, M.D.; Houck, A.A.; Schuster, D.I.; Bishop, L.S.; Ginossar, E.; Gambetta, J.M.; DiCarlo, L.; Frunzio, L.; Girvin, S.M.; et al. Quantum Non-Demolition Detection Single Microwave Photons A Circuit. Nat. Phys. 2010, 6, 663–667. [Google Scholar] [CrossRef] [Green Version]
  37. Besse, J.C.; Gasparinetti, S.; Collodo, M.C.; Walter, T.; Kurpiers, P.; Pechal, M.; Eichler, C.; Wallraff, A. Single-Shot Quantum Nondemolition Detection of Individual Itinerant Microwave Photons. Phys. Rev. X 2018, 8, 021003. [Google Scholar] [CrossRef] [Green Version]
  38. Kono, S.; Koshino, K.; Tabuchi, Y.; Noguchi, A.; Nakamura, Y. Quantum non-demolition detection of an itinerant microwave photon. Nat. Phys. 2018, 14, 546–549. [Google Scholar] [CrossRef]
  39. Lescanne, R.; Deléglise, S.; Albertinale, E.; Réglade, U.; Capelle, T.; Ivanov, E.; Jacqmin, T.; Leghtas, Z.; Flurin, E. Irreversible Qubit-Photon Coupling for the Detection of Itinerant Microwave Photons. Phys. Rev. X 2020, 10, 021038. [Google Scholar] [CrossRef]
  40. Inomata, K.; Lin, Z.; Koshino, K.; Oliver, W.D.; Tsai, J.-S.; Yamamoto, T.; Nakamura, Y. Single microwave-photon detector using an artificial Λ-type three-level system. Nat. Commun. 2016, 7, 1–7. [Google Scholar] [CrossRef] [PubMed]
  41. Chen, Y.F.; Hover, D.; Sendelbach, S.; Maurer, L.; Merkel, S.T.; Pritchett, E.J.; Wilhelm, F.K.; McDermott, R. Microwave Photon Counter Based on Josephson Junctions. Phys. Rev. Lett. 2011, 107, 217401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Kuzmin, L.S.; Sobolev, A.S.; Gatti, C.; Di Gioacchino, D.; Crescini, N.; Gordeeva, A.; Il’ichev, E. Single Photon Counter Based on a Josephson Junction at 14 GHz for Searching Galactic Axions. IEEE Trans. Appl. Supercond. 2018, 28, 1–5. [Google Scholar] [CrossRef]
  43. Semenov, A.D.; Gol’tsman, G.N.; Sobolewski, R. Hot-electron effect in superconductors and its applications for radiation sensors. Supercond. Sci. Technol. 2002, 15. [Google Scholar] [CrossRef]
  44. Paolucci, F.; Buccheri, V.; Germanese, G.; Ligato, N.; Paoletti, R.; Signorelli, G.; Bitossi, M.; Spagnolo, P.; Falferi, P.; Rajteri, M.; et al. Development of highly sensitive nanoscale transition edge sensors for gigahertz astronomy and dark matter search. J. Appl. Phys. 2020, 128, 194502. [Google Scholar] [CrossRef]
  45. Komiyama, S. Single-Photon Detectors in the Terahertz Range. IEEE J. Sel. Top. Quantum Electron. 2011, 17, 54–66. [Google Scholar] [CrossRef]
  46. Dixit, A.V.; Chakram, S.; He, K.; Agrawal, A.; Naik, R.K.; Schuster, D.I.; Chou, A. Searching for Dark Matter with a Superconducting Qubit. Phys. Rev. Lett. 2021, 126, 141302. [Google Scholar] [CrossRef] [PubMed]
  47. Kutlu, Ç.; van Loo, A.F.; Uchaikin, S.V.; Matlashov, A.N.; Lee, D.; Oh, S.; Kim, J.; Chung, W.; Nakamura, Y.; Semertzidis, Y.K. Characterization of a flux-driven Josephson parametric amplifier with near quantum-limited added noise for axion search experiments. Supercond. Sci. Technol. 2021. [Google Scholar] [CrossRef]
  48. Revin, L.S.; Pankratov, A.L.; Gordeeva, A.V.; Yablokov, A.A.; Rakut, I.V.; Zbrozhek, V.O.; Kuzmin, L.S. Microwave photon detection by an Al Josephson junction. Beilstein J. Nanotechnol. 2020, 11, 960–965. [Google Scholar] [CrossRef] [PubMed]
  49. Alesini, D.; Babusci, D.; Barone, C.; Buonomo, B.; Beretta, M.M.; Bianchini, L.; Castellano, G.; Chiarello, F.; Di Gioacchino, D.; Falferi, P.; et al. Development of a Josephson junction based single photon microwave detector for axion detection experiments. J. Phys. Conf. Ser. 2020, 1559, 012020. [Google Scholar] [CrossRef]
  50. Lamoreaux, S.K.; van Bibber, K.A.; Lehnert, K.W.; Carosi, G. Analysis of single-photon and linear amplifier detectors for microwave cavity dark matter axion searches. Phys. Rev. D 2013, 88, 035020. [Google Scholar] [CrossRef] [Green Version]
  51. Alesini, D.; Babusci, D.; Barone, C.; Buonomo, B.; Beretta, M.M.; Bianchini, L.; Castellano, G.; Chiarello, F.; Di Gioacchino, D.; Falferi, P.; et al. Status of the SIMP Project: Toward the Single Microwave Photon Detection. J. Low Temp. Phys. 2020, 199, 348–354. [Google Scholar] [CrossRef]
  52. Barone, A.; Paternò, G. Physics and Applications of the Josephson Effect; Wiley: New York, NY, USA, 1982. [Google Scholar]
  53. Kogan, S. Electronic Noise and Fluctuations in Solids; Cambridge University Press: Cambridge, UK, 1996. [Google Scholar]
  54. Martinis, J.M.; Devoret, M.H.; Clarke, J. Energy-Level Quantization in the Zero-Voltage State of a Current-Biased Josephson Junction. Phys. Rev. Lett. 1985, 55, 1543–1546. [Google Scholar] [CrossRef] [Green Version]
  55. Devoret, M.H.; Martinis, J.M.; Clarke, J. Measurement of Macroscopic Quantum Tunneling out of the Zero-Voltage State of a Current-Biased Josephson Junction. Phys. Rev. Lett. 1985, 55, 1908. [Google Scholar] [CrossRef] [PubMed]
  56. Piedjou Komnang, A.; Guarcello, C.; Barone, C.; Gatti, C.; Pagano, S.; Pierro, V.; Rettaroli, A.; Filatrella, G. Analysis of Josephson junctions switching time distributions for the detection of single microwave photons. Chaos Solitons Fractals 2020, 142, 110496. [Google Scholar] [CrossRef]
  57. Ben-Jacob, E.; Bergman, D. Thermal noise effects on the microwave-induced steps of a current-driven josephson junction. Phys. Rev. A 1984, 29, 2021–2028. [Google Scholar] [CrossRef]
  58. Yurke, B.; Denker, J. Quantum Network Theory. Phys. Rev. A 1984, 29, 1419. [Google Scholar] [CrossRef]
  59. Schondorf, M.; Govia, L.; Vavilov, M.; McDermott, R.; Wilhelm, F. Optimizing microwave photodetection: Input—Output theory. Quantum Sci. Technol. 2018, 3, 024009. [Google Scholar] [CrossRef] [Green Version]
  60. Kivioja, J.M.; Nieminen, T.E.; Claudon, J.; Buisson, O.; Hekking, F.W.J.; Pekola, J.P. Observation of Transition from Escape Dynamics to Underdamped Phase Diffusion in a Josephson Junction. Phys. Rev. Lett. 2005, 94, 247002. [Google Scholar] [CrossRef] [Green Version]
  61. Kivioja, J.M.; Nieminen, T.E.; Claudon, J.; Buisson, O.; Hekking, F.W.J.; Pekola, J.P. Weak coupling Josephson junction as a current probe: Effect of dissipation on escape dynamics. New J. Phys. 2005, 7, 179. [Google Scholar] [CrossRef]
  62. Longobardi, L.; Massarotti, D.; Stornaiuolo, D.; Galletti, L.; Rotoli, G.; Lombardi, F.; Tafuri, F. Direct Transition from Quantum Escape to a Phase Diffusion Regime in YBaCuO Biepitaxial Josephson Junctions. Phys. Rev. Lett. 2012, 109, 050601. [Google Scholar] [CrossRef] [Green Version]
  63. Bae, M.H.; Sahu, M.; Lee, H.J.; Bezryadin, A. Multiple-retrapping processes in the phase-diffusion regime of high-Tc intrinsic Josephson junctions. Phys. Rev. B 2009, 79, 104509. [Google Scholar] [CrossRef] [Green Version]
  64. Yu, H.F.; Zhu, X.B.; Peng, Z.H.; Tian, Y.; Cui, D.J.; Chen, G.H.; Zheng, D.N.; Jing, X.N.; Lu, L.; Zhao, S.P.; et al. Quantum Phase Diffusion in a Small Underdamped Josephson Junction. Phys. Rev. Lett. 2011, 107, 067004. [Google Scholar] [CrossRef] [Green Version]
  65. Massarotti, D.; Longobardi, L.; Galletti, L.; Stornaiuolo, D.; Montemurro, D.; Pepe, G.; Rotoli, G.; Barone, A.; Tafuri, F. Escape dynamics in moderately damped Josephson junctions. Low Temp. Phys. Fiz. Nizk. Temp. 2012, 38, 336. [Google Scholar]
  66. Castellano, G.; Torrioli, G.; Chiarello, F.; Cosmelli, C.; Carelli, P. Return current in hysteretic Josephson junctions: Experimental distribution in the thermal activation regime. J. Appl. Phys. 1999, 86, 6405. [Google Scholar] [CrossRef]
  67. Fenton, J.C.; Warburton, P.A. Monte Carlo simulations of thermal fluctuations in moderately damped Josephson junctions: Multiple escape and retrapping, switching- and return-current distributions, and hysteresis. Phys. Rev. B 2008, 78, 054526. [Google Scholar] [CrossRef] [Green Version]
  68. Yoon, Y.; Gasparinetti, S.; Pekola, M.M.J.P. Capacitively Enhanced Thermal Escape in Underdamped Josephson Junctions. J. Low Temp. Phys. 2011, 163, 164. [Google Scholar] [CrossRef]
  69. Martinis, J.M.; Devoret, M.H.; Clarke, J. Experimental tests for the quantum behavior of a macroscopic degree of freedom: The phase difference across a Josephson junction. Phys. Rev. B 1987, 35, 4682–4698. [Google Scholar] [CrossRef]
Figure 1. Axion parameter space showing the exclusion limits of experiments. The yellow band is representative of the QCD axions, together with the brown lines showing theoretical predictions of the KSVZ [9,10] and DFSZ [11,12] models. The plot is freely available from the online repository of [13].
Figure 1. Axion parameter space showing the exclusion limits of experiments. The yellow band is representative of the QCD axions, together with the brown lines showing theoretical predictions of the KSVZ [9,10] and DFSZ [11,12] models. The plot is freely available from the online repository of [13].
Instruments 05 00025 g001
Figure 2. (a) An electrical model of a JJ with intrinsic and external current sources. (b) An electrical model of a JJ attached to a transmission line. (c) An electrical model of a JJ with a parasitic RC load.
Figure 2. (a) An electrical model of a JJ with intrinsic and external current sources. (b) An electrical model of a JJ attached to a transmission line. (c) An electrical model of a JJ with a parasitic RC load.
Instruments 05 00025 g002
Figure 3. Equivalent potential of a JJ. The phase value is represented by the green particle which can overcame the energy barrier after the absorption of a suitable stimulus.
Figure 3. Equivalent potential of a JJ. The phase value is represented by the green particle which can overcame the energy barrier after the absorption of a suitable stimulus.
Instruments 05 00025 g003
Figure 4. (a) Schematic image of the shadow mask evaporation technique. The copolymer/PMMA bilayer was exposed by EBL and developed, in order to obtain a self-standing bridge of the desired geometry, as shown by the scanning electron microscope (SEM) micrograph in (b). A two-angle evaporation with an oxidation step in between defines the junction geometry (light and dark gray in (a). (c) SEM micrograph of the junction after the lift-off process used to remove the bilayer mask. (d) An atomic force microscope (AFM) characterization of a typical junction.
Figure 4. (a) Schematic image of the shadow mask evaporation technique. The copolymer/PMMA bilayer was exposed by EBL and developed, in order to obtain a self-standing bridge of the desired geometry, as shown by the scanning electron microscope (SEM) micrograph in (b). A two-angle evaporation with an oxidation step in between defines the junction geometry (light and dark gray in (a). (c) SEM micrograph of the junction after the lift-off process used to remove the bilayer mask. (d) An atomic force microscope (AFM) characterization of a typical junction.
Instruments 05 00025 g004
Figure 5. A scheme of the experimental setup.
Figure 5. A scheme of the experimental setup.
Instruments 05 00025 g005
Figure 6. Measured voltage-current characteristic of a Josephson junction (2 μ m × 2 μ m ).
Figure 6. Measured voltage-current characteristic of a Josephson junction (2 μ m × 2 μ m ).
Instruments 05 00025 g006
Figure 7. Filtered escape current distributions for 2 μ m × 2 μ m junction (a) and for 2 μ m × 4 μ m junction (b). The ranges of temperatures are slightly different and are shown in the legends. Rightmost curves are taken at the lowest temperature and leftmost curves are taken at highest temperature. The effect of the filter was a reduction of σ up to 20% with respect to the unfiltered data (not shown); I c was left unchanged.
Figure 7. Filtered escape current distributions for 2 μ m × 2 μ m junction (a) and for 2 μ m × 4 μ m junction (b). The ranges of temperatures are slightly different and are shown in the legends. Rightmost curves are taken at the lowest temperature and leftmost curves are taken at highest temperature. The effect of the filter was a reduction of σ up to 20% with respect to the unfiltered data (not shown); I c was left unchanged.
Instruments 05 00025 g007
Figure 8. I c and σ for 2 μ m × 2 μ m junction (a) and for 2 μ m × 4 μ m junction (b). In both plots, the red squares are relative to the left axis (widths) and the empty blue squares are relative to the right axis (mean currents), as indicated by the arrows.
Figure 8. I c and σ for 2 μ m × 2 μ m junction (a) and for 2 μ m × 4 μ m junction (b). In both plots, the red squares are relative to the left axis (widths) and the empty blue squares are relative to the right axis (mean currents), as indicated by the arrows.
Instruments 05 00025 g008
Figure 9. Mean switching currents (a) and widths (b) of the simulated data for different simulation parameters. In the legend, TA refers to thermal activation processes, QT to quantum tunneling, and RC to the dissipation effect due to the external circuit.
Figure 9. Mean switching currents (a) and widths (b) of the simulated data for different simulation parameters. In the legend, TA refers to thermal activation processes, QT to quantum tunneling, and RC to the dissipation effect due to the external circuit.
Instruments 05 00025 g009
Figure 10. Comparison of the mean switching currents and widths of the simulated data with R p = 1 k Ω , C = 2 pF, R J = 100 k Ω , I 0 = 143 nA, and C J = 0.2 pF with respect to experimental results of the 2 μ m × 2 μ m junction, shown in empty squares.
Figure 10. Comparison of the mean switching currents and widths of the simulated data with R p = 1 k Ω , C = 2 pF, R J = 100 k Ω , I 0 = 143 nA, and C J = 0.2 pF with respect to experimental results of the 2 μ m × 2 μ m junction, shown in empty squares.
Instruments 05 00025 g010
Figure 11. Experimental dark-count rate for the 2 μ m × 2 μ m junction at 50 mK, compared with quantum tunneling escape rate (red dashed line) and quantum tunneling plus the effect of 1.1 nA of Gaussian smearing (green solid line).
Figure 11. Experimental dark-count rate for the 2 μ m × 2 μ m junction at 50 mK, compared with quantum tunneling escape rate (red dashed line) and quantum tunneling plus the effect of 1.1 nA of Gaussian smearing (green solid line).
Instruments 05 00025 g011
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rettaroli, A.; Alesini, D.; Babusci, D.; Barone, C.; Buonomo, B.; Beretta, M.M.; Castellano, G.; Chiarello, F.; Di Gioacchino, D.; Felici, G.; et al. Josephson Junctions as Single Microwave Photon Counters: Simulation and Characterization. Instruments 2021, 5, 25. https://0-doi-org.brum.beds.ac.uk/10.3390/instruments5030025

AMA Style

Rettaroli A, Alesini D, Babusci D, Barone C, Buonomo B, Beretta MM, Castellano G, Chiarello F, Di Gioacchino D, Felici G, et al. Josephson Junctions as Single Microwave Photon Counters: Simulation and Characterization. Instruments. 2021; 5(3):25. https://0-doi-org.brum.beds.ac.uk/10.3390/instruments5030025

Chicago/Turabian Style

Rettaroli, Alessio, David Alesini, Danilo Babusci, Carlo Barone, Bruno Buonomo, Matteo Mario Beretta, Gabriella Castellano, Fabio Chiarello, Daniele Di Gioacchino, Giulietto Felici, and et al. 2021. "Josephson Junctions as Single Microwave Photon Counters: Simulation and Characterization" Instruments 5, no. 3: 25. https://0-doi-org.brum.beds.ac.uk/10.3390/instruments5030025

Article Metrics

Back to TopTop