Advertisement

An on-chip microwave source

The active elements of superconducting quantum circuits are typically addressed and controlled using pulses of microwave radiation. The microwaves are usually generated externally and coupled into the circuitry, resulting in rather bulky systems. Cassidy et al. developed an on-chip source of microwaves by using a superconducting Josephson junction inserted in a high-quality microwave cavity. The integrated version should enhance the control capability for manipulating miniaturized quantum circuits.
Science, this issue p. 939

Abstract

Superconducting electronic devices have reemerged as contenders for both classical and quantum computing due to their fast operation speeds, low dissipation, and long coherence times. An ultimate demonstration of coherence is lasing. We use one of the fundamental aspects of superconductivity, the ac Josephson effect, to demonstrate a laser made from a Josephson junction strongly coupled to a multimode superconducting cavity. A dc voltage bias applied across the junction provides a source of microwave photons, and the circuit’s nonlinearity allows for efficient down-conversion of higher-order Josephson frequencies to the cavity’s fundamental mode. The simple fabrication and operation allows for easy integration with a range of quantum devices, allowing for efficient on-chip generation of coherent microwave photons at low temperatures.
Josephson junctions are natural voltage-to-frequency converters via the ac Josephson effect. For a Josephson junction without an applied dc voltage bias, Cooper pairs tunnel coherently from one superconducting condensate to the other, resulting in a supercurrent flowing without dissipation. However, for a nonzero dc voltage bias (Vb) less than the superconducting energy gap, transport is prohibited unless the excess energy (hf = 2eVb, where h is Planck’s constant, f is frequency, and 2e is the charge on the Cooper pair) can be dissipated into the environment.
The analogy between a single Josephson junction and a two-level atom was first proposed theoretically in the 1970s (1). The voltage difference across the junction provides the two energy levels for the Cooper pairs, and spontaneous as well as stimulated emission and absorption were predicted to occur (1, 2). Emission from Josephson junctions into low-quality cavities (either constructed artificially or intrinsic to the junction’s environment) in the so-called weak-coupling regime has been studied extensively (36). However, because the total output power of these systems is low, coherent radiation has not been directly demonstrated. By using a tightly confined cavity mode, coherent interaction of a single Josephson junction and the cavity can be achieved. Lasing results when the transfer rate of Cooper pairs across the junction, ΓCP, exceeds the cavity decay rate, κ, of the microwave photons (Fig. 1A). Photon emission from alternative single emitters coupled to superconducting resonators has been the subject of several recent investigations (79).
Fig. 1 An ac Josephson laser.
(A) Illustration of the operating principle of the device. A dc voltage bias Vb applied across the Josephson junction results in photon emission into the cavity when twice the bias voltage is equal to a multiple of the cavity frequency. If the emission rate ΓCP into the cavity exceeds the cavity lifetime κ, these photons can then be reabsorbed and reemitted by the junction, a process akin to stimulated emission in atomic laser systems. Dashed lines depict the superconducting condensate; solid lines represent the superconducting gap, ∆. (B) Scanning electron microscopy (false color) and (C) optical microscopy images of the device. A dc SQUID (red), acting as a tunable Josephson junction, is strongly coupled to the electric-field antinode of a half-wave superconducting coplanar waveguide resonator (yellow).
We demonstrate lasing in the microwave frequency domain from a dc voltage–biased Josephson junction strongly coupled to a superconducting coplanar waveguide resonator. Our device obeys several properties present in conventional optical lasers, including injection locking and frequency-comb generation, with an injection-locked linewidth of ≤1 Hz, which exceeds the performance of other state-of-the-art laser systems.
The laser consists of a half-wave coplanar waveguide (CPW) resonator with resonant frequency f0 ≈ 5.6 GHz made from thin (20-nm) NbTiN (Fig. 1, B and C). A dc superconducting quantum interference device (SQUID), located at the electric field antinode of the cavity, effectively acts as a single junction with Josephson energy tunable via the magnetic flux ϕ threading its loop: E J = E J 0 | cos ( π ϕ / ϕ 0 ) | , where E J 0 ~ 78  GHz , and ϕ 0 = h / 2 e is the superconducting flux quantum. One side of the SQUID is tied to the central conductor of the CPW, with the other end attached directly to the ground plane to enhance the coupling to the cavity. An on-chip inductor positioned at the electric-field node of the cavity allows for a dc voltage bias to be applied across the SQUID (10). Coupling capacitors at each end of the cavity provide an input and output for microwave photons at a rate of κin and κout, respectively, as in standard circuit–quantum electrodynamics experiments (11). The device is mounted in a dilution refrigerator with base temperature T = 15 mK, and the magnetic flux through the SQUID is tuned via a superconducting vector magnet (12).
We first examine the response of the device without applying any microwave power to the cavity input. At the output, we measure the power spectral density, S(f), of the emitted microwave radiation as a function of voltage bias Vb. Simultaneously, we record the corresponding current flowing through the device, ID = 2eΓCP. The coupling between a dc voltage–biased Josephson junction and a cavity λ =   E J / ϕ 0 2 L is set by the junction’s Josephson energy, EJ, together with the cavity inductance, L = 1 / C ω 0 2 , where C is the cavity capacitance and ω0 = 2πf0 (12). When the device is configured in the weak–Josephson coupling regime ( λ 1 ) (Fig. 2A), by tuning the external flux close to ϕ =   ϕ 0 , a series of discrete microwave emission peaks are visible at bias voltages corresponding to n multiples of the bare cavity resonance, Vr: Vb = nVr = nhf0/2en × 11.62 μV. At each of these emission bursts, we observe an increase in dc current (Fig. 2B), a measure of inelastic Cooper pair transport across the Josephson junction. In this weak-coupling regime, both the current and microwave emission are dominated by linear effects, with the rate of photon emission determined by the environmental impedance (6, 13) (Fig. 2C).
Fig. 2 Microwave emission from the Josephson laser.
(A) Power spectrum of the emitted radiation S(f), (B) integrated emission (gray), and corresponding current flowing through the Josephson junction ID (red) as a function of Vb when the coupling strength λ 1 . (C) Cooper pair transport can occur at discrete voltage biases corresponding to multiples of the cavity resonance frequency f0, resulting in spontaneous photon emission into the cavity. When λ > 1 , the emission (D) and corresponding current flow (E) become continuous across a range of bias values, peaking at bias voltages corresponding to multiples of the cavity frequency. (F) Cooper pair transport is accompanied by the release of multiple photons into the cavity at the fundamental frequency, as well as emission of photons into the higher-order resonator modes, resulting in a cavity photon occupancy large enough for stimulated emission to occur.
We increase the microwave emission by increasing EJ via the applied flux, to the extent that the junction and cavity become strongly coupled and the system transitions to nonlinear behavior ( λ 1 ) (14). In contrast to the discrete emission peaks seen at low Josephson energy, the emission now shifts to higher bias voltages, persisting continuously even when the voltage bias is detuned from resonance (Fig. 2D), and is accompanied by a constant flow of Cooper pairs tunneling across the junction (Fig. 2E). The emission peaks at voltages corresponding to multiples of the cavity resonance, exhibiting bifurcations common to nonlinear systems under strong driving. Between these points of instability, the emission linewidth narrows to Δf0 = 22 kHz, well below the bare cavity linewidth of ~5 MHz (12), corresponding to a phase coherence time τc = 1/πΔf0 = 15 μs. The enhancement in emission originates from stimulated emission, as a larger photon number in the cavity increases the probability of reabsorption and coherent reemission by the junction. Notably, the emission power increases here by more than three orders of magnitude, whereas the average dc power input, Pin = VbID, varies by only a factor of three. By comparing Pin with the integrated output power, we estimate a power conversion efficiency Pout/Pin > 0.3 (12), several orders of magnitude greater than that achieved for single junctions without coupling to a cavity (3) and comparable only to arrays containing several hundred synchronized junctions (4). Similar power conversion efficiencies have been seen in other strongly coupled single-emitter–cavity systems (8, 9). Application of a larger perpendicular magnetic field adjusts the cavity frequency, directly tuning the laser emission frequency by more than 50 MHz (12).
To understand the emission characteristics, we numerically simulate the time evolution of the coupled resonator–Josephson junction circuit for increasing EJ (12). If the cavity only supports a single mode, the emission power at n = 1 grows with increasing EJ; however, there is only weak emission for bias voltages corresponding to n > 1. Higher-order cavity modes allow for direct emission of higher-frequency photons that can be down-converted to the fundamental cavity frequency via the nonlinearity of the Josephson junction. Simulations for n = 20 modes show that for weak coupling ( λ 1 ) , disconnected resonant peaks are visible in the response, in agreement with the experimental data (Fig. 2A). Only when combining strong coupling ( λ 1 ) with the presence of many higher-order modes do we find a continuous narrow emission line, as observed experimentally (Fig. 2B). Simulations show that this behavior persists even if the mode spacing is inhomogeneous, as under strong driving the presence of harmonics and subharmonics for each mode means that down-conversion to the frequency of the fundamental mode of the cavity is always possible (12).
To directly confirm lasing, we measure the emission statistics in the high–Josephson coupling regime at Vb = 192.5 μV. The emitted signal is mixed with an external local oscillator and the resulting quadrature components digitized with a fast acquisition card (Fig. 3A). A time series of the demodulated free-running laser emission over a period of 100 μs (Fig. 3B) shows a clear sinusoidal behavior, never entering a subthreshold state. This is in contrast to recently demonstrated lasers made from quantum dots (9) or superconducting charge qubits (7, 8), which are strongly affected by charge noise. Instead, the coherence of our system is disrupted by occasional phase slips (Fig. 3B, inset). To quantify the effect of these phase slips, we plot the autocorrelation g(1) (Fig. 3C) and extract a phase coherence time of 14 μs, in good agreement with the coherence time extracted from the free-running linewidth.
Fig. 3 Coherence and emission statistics of the free-running Josephson laser.
(A) Real-time evaluation of the emission statistics of the free-running Josephson laser is performed with a heterodyne measurement setup. (B) Time series VI(t) of the emission over 100 μs. (Inset) Small phase slips in the emission result in a loss of coherence, resulting in an artificially broadened emission line. (C) Autocorrelation g(1)(τ) of the time series shows a phase coherence of ∼14 μs. (D) The IQ histogram acquired above the lasing threshold shows a clear donut shape, a characteristic of coherent emission. (E) The IQ histogram obtained when the device is not emitting shows a Gaussian peak centered at zero, corresponding to thermal emission. (F) The reconstructed photon number distribution pn at the cavity output (red) is well fit by a single Gaussian peak centered around an average photon number   N ¯   29 , 600 , slightly broader than what expected for an ideal coherent source (blue).
To confirm coherence over longer time scales, we plot the in-phase and quadrature components of the down-converted signal from 5 × 105 samples on a two-dimensional histogram (Fig. 3D). The donut shape of the histogram confirms lasing, with the radius A =   N ¯ = 172 (where N ¯ is the average photon number) representing the average coherent amplitude of the system, whereas the finite width σ l =   ( 2 δ A 2 + N noise ) / 2 = 6.89 is a result of amplitude fluctuations in the cavity emission δA = 2.66 broadened by the thermal noise from the amplifier chain, Nnoise. When the device is not lasing (Vb = 18 μV in Fig. 2A), we record a Gaussian peak of width σ th =   N noise / 2 = 6.36 , corresponding to thermal emission (Fig. 3E). To extract the photon number distribution at the output of the cavity, the contribution of thermal fluctuations due to the amplifier chain in Fig. 3E is subtracted from the emission data in Fig. 3D. The extracted distribution takes the form p n exp [ ( n N ¯ ) 2 / 2 N ¯ ( 1 + 4 δ A 2 ) ] and is centered at N ¯   29 , 600 (red curve in Fig. 3F). In contrast, a perfectly coherent source would show a shot-noise–limited Poissonian distribution, which tends to a Gaussian distribution of the form p n exp [ ( n N ¯ ) 2 / 2 N ¯ ] in the limit of large N ¯ (blue curve in Fig. 3F). The residual fluctuations in the cavity amplitude are most probably due to EJ fluctuations, which change the instantaneous photon emission rate into the cavity.
Emission linewidth is a key figure of merit for lasers. A narrow linewidth implies high-frequency stability and resolution, which is important for a range of technologies including spectroscopy, imaging, and sensing application. One technique commonly used for stabilizing lasers is injection locking (15, 16) (Fig. 4, A and B). The injection of a seed tone of frequency finj into the cavity generates stimulated emission in the Josephson junction at this injected frequency, narrowing the emission spectrum. Figure 4C shows S(f) as a function of input power Pinj for an injected signal with frequency finj = 5.651 GHz, well within the emission bandwidth of the free-running source [linecuts at Pinj = –127 dBm and Pinj = –90 dBm (Fig. 4D)]. For very low input power, Pinj < –140 dBm, the average photon occupation of the cavity is N ¯ < 1 , and the device remains unaffected by the input tone. Once the photon occupancy exceeds N ¯ 1 , the injected microwave photons drive stimulated emission in the device, causing the emission linewidth to narrow with increasing power, reaching an ultimate (measurement-limited) linewidth of 1 Hz (Fig. 4D, inset), which is more than three orders of magnitude narrower than the free-running emission peak and approaches the Schawlow-Townes limit of ~15 mHz (12). In this regime, our device acts as a quantum-limited amplifier, similar to other Josephson junction–based amplifiers (17, 18); however, no additional microwave pump tone is required to provide amplification. Figure 4E shows the effect when the input tone is applied at a frequency finj = 5.655 GHz, outside the cavity bandwidth. When Pinj > –130 dBm, distortion sidebands appear at both positive and negative frequencies, and the free-running emission peak is pulled toward the input tone, eventually being locked when Pinj > –85 dBm [linecuts at Pinj = –127 dBm and Pinj = –90 dBm (Fig. 4F)]. The positions and intensities of these emission sidebands are well described by the Adler theory for the synchronization of coupled oscillators (19), similar to what has been observed for both traditional and exotic laser systems (15, 16). The frequency range over which the device can be injection-locked strongly depends on the injected power. Figure 4G shows S(f) as a function of finj at an input power Pinj = –90 dBm, with an injection locking range ∆f of almost 5 MHz. Here, the distortion sidebands span more than 100 MHz (12). Measurements of ∆f as a function of Pinj are shown in the inset of Fig. 4G. Adler’s theory predicts that the injection-locking range should fit Δ f = α P inj , with a measured prefactor α = ( 3.66 ± 1.93 )  MHz / W (W, Watts).
Fig. 4 Injection locking of the Josephson laser.
(A) Schematic illustration and (B) phasor diagram of the injection-locking process. Injection of a low-power input tone into the cavity drives stimulated emission of photons synchronous with the input tone, reducing the phase fluctuations δ ϕ experienced in the free-running mode. (C) S(f) as a function of input power (Pinj) for an on-resonance input tone. (D) Linecuts of (C) at Pinj = –90 dBm (red) and –127 dBm (black). (Inset) The linewidth of the injection-locked laser is ≤1 Hz. (E) S(f) as a function of Pinj for an off-resonance input tone, demonstrating frequency pulling. (F) Linecuts of (E) at Pinj = –90 dBm (red) and –127 dBm (black). (G) S(f) at a fixed input power Pinj = –90 dBm as the frequency of the input tone finj is swept. For probe frequencies finj in the range ∆f, the laser emission frequency locks to the frequency of the input signal. (Inset) The bandwidth of frequency locking ∆f scales proportionally with the square root of the input power, in agreement with the Adler theory of coupled oscillators.
We can also use the device to generate a microwave frequency comb, an alternative to recently demonstrated four-wave–mixing methods (20). The time-frequency duality implies that a voltage modulation applied to the junction will create a comb in the frequency domain (21). By configuring the device in the on-resonance injection-locked regime (Pinj = –110 dBm in Fig. 4C) and applying a small ac excitation of frequency fmod = 111 Hz to the dc bias (12), we generate a comb around the central pump tone, with frequency separation 111 Hz. The total width of the comb is set by the amplitude of the modulation, as well as the input power of the injection lock.
Our results conclusively demonstrate lasing from a dc-biased Josephson junction in the strong-coupling regime. Analysis of the output emission statistics shows 15 μs of phase coherence, with no subthreshold behavior. The Josephson junction laser does not suffer from the charge-noise–induced linewidth broadening inherent to semiconductor gain media and, thus, reaches an injection-locked linewidth of <1 Hz. The device produces frequency tunability over >50 MHz via direct tuning of the cavity frequency and over >100 MHz through the generation of injection-locking sidebands. Additional frequency control may be achieved by using a broadband tunable resonator (22), and pulse control may be provided with a tunable coupler. The phase coherence is likely limited by fluctuations in EJ, either due to 1/f-dependent flux noise from magnetic impurities (23) or to defects within the Josephson junction, as well as thermal fluctuations in the biasing circuit that vary Vb. We anticipate that improvements to the magnetic shielding and passivating magnetic fluctuators, together with the use of a cryogenically generated voltage bias, will further stabilize the emission. In this case, the device would perform at the quantum limit, with a linewidth that would only be limited by residual fluctuations in the photon number in the cavity. Along with the high efficiency, the possibility of engineering the electromagnetic environment and guiding the emitted microwaves on demand lends this system to a versatile cryogenic source for propagating microwave radiation.

Acknowledgments

We thank D. J. van Woerkom, W. Uilhoorn, and D. de Jong for experimental assistance and C. Eichler, Y. M. Blanter, F. Hassler, J. E. Mooij, and Y. V. Nazarov for helpful discussions. This work has been supported by the Netherlands Organisation for Scientific Research (NWO/OCW), as part of the Frontiers of Nanoscience program, Foundation for Fundamental Research on Matter (FOM), European Research Council (ERC), and Microsoft Corporation Station Q. M.C.C. and L.P.K. have filed a provisional patent application that relates to the Josephson junction laser.

Supplementary Material

Summary

Materials and Methods
Supplementary Text
Figs. S1 to S14
Table S1
References (2431)

Resources

File (cassidy.sm.pdf)

References and Notes

1
D. R. Tilley, Superradiance in arrays of superconducting weak links. Phys. Lett. A 33, 205–206 (1970).
2
D. Rogovin, M. Scully, Superconductivity and macroscopic quantum phenomena. Phys. Rep. 25, 175–291 (1976).
3
N. F. Pedersen, O. H. Soerensen, J. Mygind, P. E. Lindelof, M. T. Levinsen, T. D. Clark, Direct detection of the Josephson radiation emitted from superconducting thin‐film microbridges. Appl. Phys. Lett. 28, 562–564 (1976).
4
P. Barbara, A. B. Cawthorne, S. V. Shitov, C. J. Lobb, Stimulated emission and amplification in Josephson junction arrays. Phys. Rev. Lett. 82, 1963–1966 (1999).
5
L. Ozyuzer, A. E. Koshelev, C. Kurter, N. Gopalsami, Q. Li, M. Tachiki, K. Kadowaki, T. Yamamoto, H. Minami, H. Yamaguchi, T. Tachiki, K. E. Gray, W.-K. Kwok, U. Welp, Emission of coherent THz radiation from superconductors. Science 318, 1291–1293 (2007).
6
M. Hofheinz, F. Portier, Q. Baudouin, P. Joyez, D. Vion, P. Bertet, P. Roche, D. Esteve, Bright side of the Coulomb blockade. Phys. Rev. Lett. 106, 217005 (2011).
7
F. Chen, J. Li, A. D. Armour, E. Brahimi, J. Stettenheim, A. J. Sirois, R. W. Simmonds, M. P. Blencowe, A. J. Rimberg, Realization of a single-Cooper-pair Josephson laser. Phys. Rev. B 90, 020506 (2014).
8
O. Astafiev, K. Inomata, A. O. Niskanen, T. Yamamoto, Y. A. Pashkin, Y. Nakamura, J. S. Tsai, Single artificial-atom lasing. Nature 449, 588–590 (2007).
9
Y.-Y. Liu, J. Stehlik, C. Eichler, M. J. Gullans, J. M. Taylor, J. R. Petta, Semiconductor double quantum dot micromaser. Science 347, 285–287 (2015).
10
F. Chen, A. J. Sirois, R. W. Simmonds, A. J. Rimberg, Introduction of a dc bias into a high-Q superconducting microwave cavity. Appl. Phys. Lett. 98, 132509 (2011).
11
A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, R. Huang, J. Majer, S. Kumar, S. M. Girvin, R. J. Schoelkopf, Strong coupling of a single photon to a superconducting qubit using circuit quantum electrodynamics. Nature 431, 162–167 (2004).
12
Supplementary materials are available on Science Online
13
M. H. Devoret, D. Esteve, H. Grabert, G. Ingold, H. Pothier, C. Urbina, Effect of the electromagnetic environment on the Coulomb blockade in ultrasmall tunnel junctions. Phys. Rev. Lett. 64, 1824–1827 (1990).
14
S. Meister, M. Mecklenburg, V. Gramich, J. T. Stockburger, J. Ankerhold, B. Kubala, Resonators coupled to voltage-biased Josephson junctions: From linear response to strongly driven nonlinear oscillations. Phys. Rev. B 92, 174532 (2015).
15
H. L. Stover, W. H. Steier, Locking of laser oscillators by light injection. Appl. Phys. Lett. 8, 91–93 (1966).
16
Y.-Y. Liu, J. Stehlik, M. J. Gullans, J. M. Taylor, J. R. Petta, Injection locking of a semiconductor double-quantum-dot micromaser. Phys. Rev. A 92, 053802 (2015).
17
B. Yurke, P. G. Kaminsky, R. E. Miller, E. A. Whittaker, A. D. Smith, A. H. Silver, R. W. Simon, Observation of 4.2-K equilibrium-noise squeezing via a Josephson-parametric amplifier. Phys. Rev. Lett. 60, 764–767 (1988).
18
I. Siddiqi, R. Vijay, F. Pierre, C. M. Wilson, M. Metcalfe, C. Rigetti, L. Frunzio, M. H. Devoret, RF-driven Josephson bifurcation amplifier for quantum measurement. Phys. Rev. Lett. 93, 207002 (2004).
19
R. Adler, A study of locking phenomena in oscillators. Proc. IRE 34, 351–357 (1946).
20
R. P. Erickson, M. R. Vissers, M. Sandberg, S. R. Jefferts, D. P. Pappas, Frequency comb generation in superconducting resonators. Phys. Rev. Lett. 113, 187002 (2014).
21
P. Solinas, S. Gasparinetti, D. Golubev, F. Giazotto, A Josephson radiation comb generator. Sci. Rep. 5, 12260 (2015).
22
Z. L. Wang, Y. P. Zhong, L. J. He, H. Wang, J. M. Martinis, A. N. Cleland, Q. W. Xie, Quantum state characterization of a fast tunable superconducting resonator. Appl. Phys. Lett. 102, 163503 (2013).
23
F. C. Wellstood, C. Urbina, J. Clarke, Low‐frequency noise in dc superconducting quantum interference devices below 1 K. Appl. Phys. Lett. 50, 772–774 (1987).
24
R. N. Simons, Coplanar Waveguide Circuits, Components, and Systems, Wiley Series in Microwave and Optical Engineering (Wiley, 2004).
25
L. Spietz, K. W. Lehnert, I. Siddiqi, R. J. Schoelkopf, Primary electronic thermometry using the shot noise of a tunnel junction. Science 300, 1929–1932 (2003).
26
A. L. Schawlow, C. H. Townes, Infrared and optical masers. Phys. Rev. 112, 1940–1949 (1958).
27
H. Wiseman, Light amplification without stimulated emission: Beyond the standard quantum limit to the laser linewidth. Phys. Rev. A 60, 4083–4093 (1999).
28
C. Eichler, D. Bozyigit, A. Wallraff, Characterizing quantum microwave radiation and its entanglement with superconducting qubits using linear detectors. Phys. Rev. A 86, 032106 (2012).
29
M. Armand, On the output spectrum of unlocked driven oscillators. Proc. IEEE 57, 798–799 (1969).
30
V. Gramich, B. Kubala, S. Rohrer, J. Ankerhold, From Coulomb-blockade to nonlinear quantum dynamics in a superconducting circuit with a resonator. Phys. Rev. Lett. 111, 247002 (2013).
31
A. D. Armour, M. P. Blencowe, E. Brahimi, A. J. Rimberg, Universal quantum fluctuations of a cavity mode driven by a Josephson junction. Phys. Rev. Lett. 111, 247001 (2013).

(0)eLetters

eLetters is a forum for ongoing peer review. eLetters are not edited, proofread, or indexed, but they are screened. eLetters should provide substantive and scholarly commentary on the article. Embedded figures cannot be submitted, and we discourage the use of figures within eLetters in general. If a figure is essential, please include a link to the figure within the text of the eLetter. Please read our Terms of Service before submitting an eLetter.

Log In to Submit a Response

No eLetters have been published for this article yet.

Information & Authors

Information

Published In

Science
Volume 355 | Issue 6328
3 March 2017

Submission history

Received: 28 July 2016
Accepted: 3 February 2017
Published in print: 3 March 2017

Permissions

Request permissions for this article.

Acknowledgments

We thank D. J. van Woerkom, W. Uilhoorn, and D. de Jong for experimental assistance and C. Eichler, Y. M. Blanter, F. Hassler, J. E. Mooij, and Y. V. Nazarov for helpful discussions. This work has been supported by the Netherlands Organisation for Scientific Research (NWO/OCW), as part of the Frontiers of Nanoscience program, Foundation for Fundamental Research on Matter (FOM), European Research Council (ERC), and Microsoft Corporation Station Q. M.C.C. and L.P.K. have filed a provisional patent application that relates to the Josephson junction laser.

Authors

Affiliations

QuTech, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, Netherlands.
QuTech, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, Netherlands.
Kavli Institute for Nanoscience, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, Netherlands.
Kavli Institute for Nanoscience, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, Netherlands.
QuTech, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, Netherlands.
QuTech, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, Netherlands.
Kavli Institute for Nanoscience, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, Netherlands.
Kavli Institute for Nanoscience, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, Netherlands.
QuTech, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, Netherlands.
Kavli Institute for Nanoscience, Delft University of Technology, P.O. Box 5046, 2600 GA Delft, Netherlands.

Funding Information

Microsoft: award290728

Notes

*
Corresponding author. Email: [email protected]

Metrics & Citations

Metrics

Article Usage

Altmetrics

Citations

Cite as

Export citation

Select the format you want to export the citation of this publication.

Cited by

  1. Latched detection of zeptojoule spin echoes with a kinetic inductance parametric oscillator, Science Advances, 10, 14, (2024)./doi/10.1126/sciadv.adm7624
    Abstract
Loading...

View Options

View options

PDF format

Download this article as a PDF file

Download PDF

Check Access

Log in to view the full text

AAAS ID LOGIN

AAAS login provides access to Science for AAAS Members, and access to other journals in the Science family to users who have purchased individual subscriptions.

Log in via OpenAthens.
Log in via Shibboleth.

More options

Register for free to read this article

As a service to the community, this article is available for free. Login or register for free to read this article.

Purchase this issue in print

Buy a single issue of Science for just $15 USD.

Media

Figures

Multimedia

Tables

Share

Share

Share article link

Share on social media