probing the strongly driven spin-boson model in a ... · a. lupascu2,3,4, c.m. wilson2,6 & m....

8
ARTICLE Probing the strongly driven spin-boson model in a superconducting quantum circuit L. Magazzù 1 , P. Forn-Díaz 2,3,4,5 , R. Belyansky 2,6 , J.-L. Orgiazzi 2,4,6 , M.A. Yurtalan 2,4,6 , M.R. Otto 2,3,4 , A. Lupascu 2,3,4 , C.M. Wilson 2,6 & M. Grifoni 7 Quantum two-level systems interacting with the surroundings are ubiquitous in nature. The interaction suppresses quantum coherence and forces the system towards a steady state. Such dissipative processes are captured by the paradigmatic spin-boson model, describing a two-state particle, the spin, interacting with an environment formed by harmonic oscilla- tors. A fundamental question to date is to what extent intense coherent driving impacts a strongly dissipative system. Here we investigate experimentally and theoretically a super- conducting qubit strongly coupled to an electromagnetic environment and subjected to a coherent drive. This setup realizes the driven Ohmic spin-boson model. We show that the drive reinforces environmental suppression of quantum coherence, and that a coherent-to- incoherent transition can be achieved by tuning the drive amplitude. An out-of-equilibrium detailed balance relation is demonstrated. These results advance fundamental understanding of open quantum systems and bear potential for the design of entangled light-matter states. DOI: 10.1038/s41467-018-03626-w OPEN 1 Institute of Physics, University of Augsburg, Universitätsstraße 1, D-86135 Augsburg, Germany. 2 Institute for Quantum Computing, University of Waterloo, Waterloo N2L 3G1, Canada. 3 Department of Physics and Astronomy, University of Waterloo, Waterloo N2L 3G1, Canada. 4 Waterloo Institute for Nanotechnology, University of Waterloo, Waterloo N2L 3G1, Canada. 5 Barcelona Supercomputing Center (BSC), C/Jordi Girona 29, 08034 Barcelona, Spain. 6 Department of Electrical and Computer Engineering, University of Waterloo, Waterloo N2L 3G1, Canada. 7 Institute for Theoretical Physics, University of Regensburg, 93040 Regensburg, Germany. These authors contributed equally: L. Magazzù, P. Forn-Díaz. Correspondence and requests for materials should be addressed to A.L. (email: [email protected]) or to C.M.W. (email: [email protected]) or to M.G. (email: [email protected]) NATURE COMMUNICATIONS | (2018)9:1403 | DOI: 10.1038/s41467-018-03626-w | www.nature.com/naturecommunications 1 1234567890():,;

Upload: others

Post on 28-Oct-2019

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Probing the strongly driven spin-boson model in a ... · A. Lupascu2,3,4, C.M. Wilson2,6 & M. Grifoni7 Quantum two-level systems interacting with the surroundings are ubiquitous in

ARTICLE

Probing the strongly driven spin-boson model in asuperconducting quantum circuitL. Magazzù1, P. Forn-Díaz2,3,4,5, R. Belyansky2,6, J.-L. Orgiazzi2,4,6, M.A. Yurtalan2,4,6, M.R. Otto 2,3,4,

A. Lupascu2,3,4, C.M. Wilson2,6 & M. Grifoni7

Quantum two-level systems interacting with the surroundings are ubiquitous in nature. The

interaction suppresses quantum coherence and forces the system towards a steady state.

Such dissipative processes are captured by the paradigmatic spin-boson model, describing a

two-state particle, the “spin”, interacting with an environment formed by harmonic oscilla-

tors. A fundamental question to date is to what extent intense coherent driving impacts a

strongly dissipative system. Here we investigate experimentally and theoretically a super-

conducting qubit strongly coupled to an electromagnetic environment and subjected to a

coherent drive. This setup realizes the driven Ohmic spin-boson model. We show that the

drive reinforces environmental suppression of quantum coherence, and that a coherent-to-

incoherent transition can be achieved by tuning the drive amplitude. An out-of-equilibrium

detailed balance relation is demonstrated. These results advance fundamental understanding

of open quantum systems and bear potential for the design of entangled light-matter states.

DOI: 10.1038/s41467-018-03626-w OPEN

1 Institute of Physics, University of Augsburg, Universitätsstraße 1, D-86135 Augsburg, Germany. 2 Institute for Quantum Computing, University of Waterloo,Waterloo N2L 3G1, Canada. 3 Department of Physics and Astronomy, University of Waterloo, Waterloo N2L 3G1, Canada. 4Waterloo Institute forNanotechnology, University of Waterloo, Waterloo N2L 3G1, Canada. 5 Barcelona Supercomputing Center (BSC), C/Jordi Girona 29, 08034 Barcelona, Spain.6 Department of Electrical and Computer Engineering, University of Waterloo, Waterloo N2L 3G1, Canada. 7 Institute for Theoretical Physics, University ofRegensburg, 93040 Regensburg, Germany. These authors contributed equally: L. Magazzù, P. Forn-Díaz. Correspondence and requests for materials shouldbe addressed to A.L. (email: [email protected]) or to C.M.W. (email: [email protected]) or to M.G. (email: [email protected])

NATURE COMMUNICATIONS | (2018) 9:1403 | DOI: 10.1038/s41467-018-03626-w |www.nature.com/naturecommunications 1

1234

5678

90():,;

Page 2: Probing the strongly driven spin-boson model in a ... · A. Lupascu2,3,4, C.M. Wilson2,6 & M. Grifoni7 Quantum two-level systems interacting with the surroundings are ubiquitous in

The spin-boson model has been prominent for severaldecades in the study of open quantum systems1, 2. Itdescribes a two-state quantum system (spin), interacting

with its environment. The latter is modeled as a set of harmonicoscillators (bosons) constituting a so-called heat bath. Thedynamical regimes of the spin-boson model at a given finitetemperature are essentially dictated by the coupling to theenvironment and by the low-frequency behavior of the bathspectrum. In the strong coupling regime, this model provides anaccurate representation of a variety of physical and chemicalsituations of broad interest, including incoherent tunneling ofbistable defects in metals3 and amorphous systems4, macroscopicquantum tunneling in superconducting circuits5, or electron andproton transfer in solvent environments6. Moreover, the spin-boson model is relevant in describing exciton transport in bio-logical complexes7, 8. The weak coupling regime characterizessituations where preserving quantum coherence is crucial, such asin quantum computing, whereas strong coupling can give rise tonovel entangled states of system and reservoir, for example, topolaron or Kondo clouds2.

In the Ohmic spin-boson model, the environment has a linearspectrum at low frequencies which leads to various remarkablephenomena, such as bath-induced localization or a coherent-to-incoherent transition even at zero temperature for large enoughcoupling strengths1.

Recently, a new experimental setup was implemented9 whichrealizes the Ohmic spin-boson model with an environmentalcoupling tunable from weak to ultrastrong10. This particularimplementation is formed from a superconducting flux qubitcoupled to a transmission line, which play the role of the two-state system and environment, respectively. The tunability of theinteraction allows one to test the key predictions of the spin-boson model. In11, a qubit ultrastrongly coupled to a singleoscillator mode was demonstrated.

In this article, we study the spin-boson setup from ref. 9 understrong driving, which adds a new dimension of exploration for aspin-boson system12. Previous experiments studying stronglydriven systems have reported remarkable effects, such as theformation of dressed states13–15, Landau-Zener interference16, 17,amplitude spectroscopy18, and the observation of Floquet states19.However, these experimental reports were restricted to weak ormoderate coupling to the environment. Here, we combine intensedriving and diverse dissipation strengths in a superconductingqubit circuit, with the aim of tracing out the dynamical phasediagram of a driven spin-boson system in coupling regimesranging form weak to ultrastrong.

ResultsRelation between experimental and theoretical observables. Aschematic representation of the experimental setup is shown inFig. 1a. The two-state system is a flux qubit, a superconductingcircuit consisting of a loop interrupted by four Josephson junc-tions20. The bosonic environment is formed from electromagneticmodes in the superconducting transmission line coupled to thequbit. The qubit is pumped by a strong continuous-wave driveapplied through the transmission line. Both the amplitude andthe frequency of the drive can be changed over a broad range. Thedriven system is studied spectroscopically by additionally apply-ing a weak probe field. The measured transmission T at the probefrequency ωp gives direct access to the linear response functionassociated to the weak probe signal, the so-called linear suscept-ibility χ via the relation

T ωp� �

¼ 1� iN �hωpχ ωp� �

; ð1Þ

where N is a coupling constant (see Methods). According to

Kubo’s linear response theory21, χ(ω) carries information aboutthe dispersive and absorptive properties of the qubit in theabsence of the probe, and in turn, as discussed below, about thedynamical phases of the driven spin-boson system. By measuringthe transmission also when the drive is switched off, we get areference for the effects of a coherent drive on quantum coher-ence and localization properties.

Phase diagram of the undriven spin-boson model. We firstintroduce the spin-boson model and its dynamics without driv-ing. Historically, the Ohmic spin-boson model was first studied inthe context of the tunneling of a quantum particle in a double-well potential1. At low temperatures the dynamics are effectivelyrestricted to the Hilbert space spanned by the states Lj i and Rj i,

2

1.5

1

0.5

00 0.25

T*(

�) (

h– �

/kB)

0.5 0.75

Coh

eren

t

Incoherent

IIIIII

Dig

�d

�p

�LO

�(t)

�IF

RT 3 K 10 mK

–60 dB

� = 0

h�

� > 0

k f k f

kbkb

1�

a

b

c

Fig. 1 Experimental setup and phase diagram of the symmetric spin-bosonmodel. a Measurement, driving circuit schematic and optical micrograph ofa device similar to the ones used in the experiment. A coplanar waveguiderunning across the chip plays the role of the bath coupled to the qubit. Theinset is a scanning electron micrograph showing the qubit attached to theline. The scale bar is 2 μm. Here and in panel b the red (black) arrowindicates clockwise (anticlockwise) circulating persistent currents. bSchematics of the double-well potential associated to the flux threading thequbit. In the absence of external driving sources the potential is symmetricand the forward and backward tunneling rates kf/b are equal. In thepresence of a positive bias asymmetry ε, forward tunneling dominates overbackward tunneling. c Dependence of the temperature T*(α) for thecrossover from the coherent to the incoherent tunneling regime on thecoupling α. The red curve interpolates numerical results (asterisks)obtained within the nonperturbative NIBA. The dots labeled I, II, and IIImark the positions in parameter space of the three devices used in thiswork

ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/s41467-018-03626-w

2 NATURE COMMUNICATIONS | (2018) 9:1403 | DOI: 10.1038/s41467-018-03626-w |www.nature.com/naturecommunications

Page 3: Probing the strongly driven spin-boson model in a ... · A. Lupascu2,3,4, C.M. Wilson2,6 & M. Grifoni7 Quantum two-level systems interacting with the surroundings are ubiquitous in

localized in the left and right well, respectively (see Fig. 1b).Transitions between the two localized configurations are possibledue to quantum-mechanical tunneling and are recorded in thetime evolution of the population difference PðtÞ � σzðtÞh i ¼PRðtÞ � PLðtÞ of the two localized eigenstates. The coordinateassociated with the double-well potential need not be geometrical,but it can represent other continuous variables. For the super-conducting flux qubit used in our experiment, this is the magneticflux Φ in the loop. The eigenstates Lj i and Rj i of the flux operatorare related to the currents circulating clockwise/anticlockwise inthe superconducting loop20 (see red/black arrows in Fig. 1a, b). Inthis basis, the qubit Hamiltonian is

HqbðtÞ ¼ � �h2σx þ εðtÞσz½ �; ð2Þ

where σi are the Pauli matrices. The parameter Δ accounts forinterwell tunneling and ħε(t) is the difference in energy betweenthe two wells, which is controllable. The electromagnetic field inthe transmission line can be described as a continuously dis-tributed set of propagating modes with a distribution in fre-quency given by the spectral density

GðωÞ ¼ 2αωe�ω=ωc ; ð3Þ

corresponding to Ohmic damping with the dimensionless cou-pling strength α and high frequency cutoff ωc.

Theoretical work on the spin-boson model has primarilyfocused on the temporal dynamics of the spin. Quite generally,independent of the initial state of the qubit and the form of thebath spectral density, energy exchange with the environment isresponsible for equilibration of the qubit with the bath on a timescale given by the relaxation rate γr. Furthermore, quantumfluctuations and energy exchange yield dephasing with rate γ. Inthe Ohmic spin-boson model, low frequency environmentalmodes also lead to a strong renormalization of the bare qubittunneling splitting Δ. The renormalized qubit frequency Ωdepends on the bath temperature and coupling strength α, andis always reduced with respect to Δ. This leads to three distinctdynamical regimes. Two of them, occurring for α < 1, are depicted

in Fig. 1c for the symmetric spin-boson model shown in the leftdrawing in Fig. 1b. The coherent regime corresponds to Ω > γ.This occurs for α < 1/2 and a temperature T < T*(α). In thisregime, for a spin initially localized in the right well (P(0)= 1),the qubit displays damped coherent oscillations of frequency Ω,specifically, P(t)= exp(−γt)cos(Ωt) (see insets of Fig. 2a, b). Atthe crossover temperature, the renormalized frequency Ωvanishes (see Methods and Eq. (26)). The incoherent regimecorresponds to α < 1/2 and T > T*(α) or 1/2 < α < 1. The dynamicsare characterized by incoherent tunneling transitions with rateskf/b defined in Section III of the Methods section (Fig. 2b).Correspondingly, we have PðtÞ ¼ e�γrt , where γr= kf+ kb (seeinset in Fig. 2c). In the third regime, corresponding to α > 1,localization occurs. Here, the backward and forward rates arerenormalized to zero by the low-frequency bath modes. As shownin Fig. 1c, in the Ohmic spin-boson model, the dynamics becomesfully incoherent above α= 0.5 for any value of the temperature.As the coupling approaches this value, any perturbative approachin the coupling fails to describe the physics of the system.Consistently with ref. 9, we refer to the coupling regimes α > 0.5 asultrastrong. Primary scope of this work is to understand how thedynamical phase diagram in Fig. 1c is modified by a periodicmodulation of the detuning. This is a formidable task, since thespin-boson problem with time-periodic detuning cannot besolved analytically in the whole parameter space. Exact solutionsexist for the particular value α= 1/222. Recently, an analyticalsolution was suggested for the case of a spin-boson system withtime-periodic tunneling amplitude23.

Linear susceptibility of the driven spin-boson model. To carryout our spectroscopic analysis, we describe the bias between thepotential wells in our experimental setup by means of the time-dependent function

εðtÞ ¼ ε0 þ εp cos ωpt� �

þ εd cos ωdtð Þ: ð4Þ

Here, the static component ε0 is related to the externally appliedflux Φε by ε0∝ (Φε−Φ0/2), with Φ0 the magnetic flux quantum.The remaining contributions account for the probe (p), with

� = 0.007 � = 0.21

�r

2�GME

e–�t cos (�t)

0

–1

0

0

0

1 1 1

25 50 75Time [�–1]

0 25 50Time [�–1]

0 25 50 75Time [�–1]

P (

t)

P (

t)

P (

t)

1.05

7

| |2 (dB)

6 6

5

4

4

2

8

10.95

�″

(�p)

�p/

2 (

GH

z)

�p/

2 (

GH

z)�

″ (�

p)

�″

(�p)

1.1�p/� �p/� �p/�

�* �*

e–�t cos (�t)

GME

0 00.5 0.251 1.5 0.52

e–�rt

GME

–1 –4 40 01

Theory� = 0.007

Theory� = 0.21

0–2–4–6–8–10 –10

–5

0

5

–12Exp.

Device IExp.

Device II

7

6

5

4

�p/

2 (

GH

z)

(� – 0/2) / 0

–5 0 5(×10–3)(� – 0/2) / 0 (×10–3)(� – 0/2) / 0 (×10–3)

Theory� = 0.8

Exp.Device III

–2

–3

–1

0

1

2�

a b c

d e f

2�

� = 0.8

| |2 (dB) | |2 (dB)

Fig. 2 Spin-boson dynamics and spectra at different coupling strengths in the absence of the drive. a–c Frequency dependence of the imaginary part χ″(ωp)of the linear susceptibility (a.u.) and time evolution of the population difference P(t) (insets) for the three selected combinations of coupling andtemperature shown in Fig. 1c. The position ω* and FWHM 2γ of the linear susceptibility peak in the coherent regimes (α= 0.007, α= 0.21) provide a directmeasure of the renormalized qubit frequency ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiω�ð Þ2�γ2

q. In the incoherent regime (α= 0.8), the peak position yields the relaxation rate γr. d–f

Experimental transmission spectra of three flux qubit devices with different coupling junctions are compared with spectra calculated within the NIBA. Thecharacteristic hyperbolic spectrum of the flux qubit is evident in d and recognizable in e. Its disappearance in f indicates the transition to the incoherentregime. At Φε =Φ0/2 the spin-boson system is unbiased, which is the situation of panels a–c

NATURE COMMUNICATIONS | DOI: 10.1038/s41467-018-03626-w ARTICLE

NATURE COMMUNICATIONS | (2018) 9:1403 | DOI: 10.1038/s41467-018-03626-w |www.nature.com/naturecommunications 3

Page 4: Probing the strongly driven spin-boson model in a ... · A. Lupascu2,3,4, C.M. Wilson2,6 & M. Grifoni7 Quantum two-level systems interacting with the surroundings are ubiquitous in

amplitude εp and frequency ωp, and the drive (d), with amplitudeεd and frequency ωd. For details, see the Methods section. Thecentral quantity in this work is the linear susceptibility χ(ωp),which describes the qubit’s response at the probe frequency ωp,see Eq. (1). The susceptibility measures deviations of theasymptotic population difference, Pas(t), from its value P0 in theabsence of the weak probe according to24

PasðtÞ ¼ P0 þ �hεp χ ωp� �

eiωpt þ χ �ωp� �

e�iωpt� �

: ð5Þ

In this work, the dynamical quantity P(t), and in turn thesusceptibility χ(ωp), have been calculated within the so-callednoninteracting-blip approximation (NIBA). This approximationyields a generalized master equation for P(t) with kernels that arenonperturbative in α. It becomes exact at large temperatures and/or coupling strengths2. Under the assumption that ωd is largecompared to the (renormalized) frequency scales of the spin-boson particle, closed expressions for the transient evolution of P(t), as well as for the linear susceptibility of the driven spin-bosonsystem, can be obtained (details in the Methods section).

Characterizing the dynamical regimes of the undriven devices.We first demonstrate in Fig. 2a–c the connection between theimaginary part, χ″(ωp), of the susceptibility and P(t) for thesymmetric spin-boson model in the presence of the probe only(ε0= εd= 0). We choose three distinct values of the coupling,namely α= 0.007, 0.21, situated in the coherent regime, and α=0.8 in the incoherent regime (see the three dots indicated inFig. 1c). In the coherent regime, χ″(ωp) has a peak at ω*= (Ω2+γ2)1/2 with full width at half maximum (FWHM) given by 2γ. Inthe incoherent regime, the peak is located near zero frequency, atthe value of the relaxation rate γr. According to Eq. (1), a

maximum in χ″(ωp) corresponds to a minimum in the trans-mission T ðωpÞ. By recording the evolution of the transmission asa function of ωp and of another external parameter, e.g., the staticasymmetry ε0, various dynamical regimes can be identified.

The theoretically calculated transmission is presented inFig. 2d–f as a function of the applied static bias ε0 for the threevalues of α discussed above. As expected, the qubit dispersionrelation can be traced back in the highly coherent and under-damped regimes corresponding to α= 0.007 and α= 0.21,respectively. In the overdamped regime, with α= 0.8, thetransmission is nearly independent of ωp. Finally, comparisonwith the measured transmission for three distinct tunable devices,named I, II, and III in the following, allows us to position thethree devices as shown in the phase diagram in Fig. 1c.Temperature, cutoff frequency, renormalized splitting Ω, andconversion factor N are estimated from the experiments.Deviations in the choice of these parameters can yield variationsin the estimate of the coupling strength α. The close agreementbetween the calculated and measured qubit spectra gives a strongevidence that Device III, with an estimated coupling α= 0.8 ± 0.1(see the Supplementary Note 7), is in the nonperturbativeultrastrong coupling regime, buttressing the conclusion of ref.9, 25. In a recent work26 a polaron approach, which is equivalentto the NIBA2, has been used to provide approximate expressionsfor the response of an undriven qubit coupled to a transmissionline.

Spectroscopy of the driven spin-boson model. Let us now turnto the impact of a strong coherent drive on a spin-boson particlein the intermediate and ultrastrong coupling regimes captured bydevices II and III, respectively. The experimental spectra inFig. 3a, e show the probe transmission as a function of flux bias ε0

–10

–5

0

5

10

Pum

p po

wer

(dB

m)

–10

0.510.500.49

–5

0

5

10

Pum

p po

wer

(dB

m)

Exp.Device II

Exp.Device III

Theory� = 0.21

Theory� = 0.8

0

5

10

15

0

5

10

15

20

0

5

10

15

0

0 5 10 15 20 25

5

10

15Min Zero Max

�0 = 0� = 0.21

c

d

g

a

0

0.81

e

0

5

10

15

20

P (t )

P (t )

P (t )

−0.2

0.20.40.6

0

0.81

−0.2

0.20.40.6

0

0.8

1

0.2

0.4

0.6

0.8

1

0.4

0.6

0.8

1

0.4

0.6

0.8

1

0.6

0.8

1

0.6

(�d/

�)2

(dB

)(�

d/�

)2 (d

B)

(�d/

�)2

(dB

)(�

d/�

)2 (d

B)

(�d/

�)2

(dB

)

� = 0.21�0 = �d

� = 0.8�0 = 0

Time [�−1]

0 5 10 15 20 25Time [�−1]

0 5 10 15 20 25Time [�−1]

∋/0

0.510.5–0.49

∋/0

0.600.550.500.45 0.550.5–0.450.4

∋/0 ∋/0

�0 = �d

J0(�d/�d)

J1(�d/�d)

b

f

| |2

| |2 | |2

| |2

Fig. 3 Spectral response and dynamics of the driven spin-boson system. a, b Observed and calculated transmission at the probe frequency for themoderately coupled Device II (α= 0.21) as function of static bias and pump strength. A clear structure of multi-photon resonances appears. The dashedlines indicate cuts at fixed bias where the dynamics in panels c, d are calculated. c,d Predicted dynamics of P(t) with P(0)= 1. e, f Observed and calculatedspectrum of the ultrastrongly coupled Device III (with α= 0.8). The spectrum is smoothed, as compared to panels a, b, indicating fully incoherentdynamics. The black dashed line in f corresponds to the condition εeff= 2kBT/ħ for the effective nonequilibrium bias (see Eq. (7)) and the symbols “×”markthe first two zeroes of J0(εdτenv). g Time evolution of P(t) calculated at the symmetry point, ε0= 0, in the same range of pump strengths as in panel f

ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/s41467-018-03626-w

4 NATURE COMMUNICATIONS | (2018) 9:1403 | DOI: 10.1038/s41467-018-03626-w |www.nature.com/naturecommunications

Page 5: Probing the strongly driven spin-boson model in a ... · A. Lupascu2,3,4, C.M. Wilson2,6 & M. Grifoni7 Quantum two-level systems interacting with the surroundings are ubiquitous in

and drive power (/ ε2d) for these devices. Probe and drive fre-quencies are respectively set to ωp/2π= 5.2 GHz and ωd/2π= 9GHz for Device II. For Device III we choose ωp/2π= 4 GHz andωd/2π= 3 GHz. For Device II, the probe is on-resonance with theundriven qubit at the symmetry point. For Device III, the qua-litative features of the driven spectra are largely insensitive to thechoice of ωp and ωd. The theoretical predictions, shown inFig. 3b, f, agree well with the experimental observations. Similarto the pump-only case, striking differences are observed in thetransmission of the two devices. Let us start discussing Device II.Minima in the transmission are clearly seen in Fig. 3a, b wheneverthe static bias matches a multiple of the pump frequency, ε0=nωd, as indicated by the vertical lines drawn in Fig. 3b for n= 0,1. Furthermore, the observed pattern with fixed bias at the n-thresonance results from a modulation by a prefactor proportionalto Jn(εd/ωd), where Jn is a Bessel function of the first kind. Forexample, the qubit response at the symmetry point is suppressedin correspondence with the first zero of the Bessel function J0(εd/ωd) (indicated by a circle), where the incoming probing field isfully transmitted. At larger power, as the zero order Besselfunction increases again, the transmission diminishes. Similarpatterns have already been reported in driven qubit devices in thehighly coherent regime14, 17. Those results can be interpreted as asignature of entangled light-matter states known as dressed-states13, 27, 28. Near the multiphoton resonance, ε0= nωd, two ofthese dressed states form an effective two-level system withdressed tunneling splitting Δn= ΔJn(εd/ωd). Near a zero of the n-th Bessel function, tunneling is strongly suppressed and hence thetransmission is maximal. This phenomenon has been dubbedcoherent destruction of tunneling in the literature29. Dissipationmodifies this simple coherent picture, as demonstrated for DeviceIII in Fig. 3e, f where no Bessel pattern is present and a smooth“V-shaped” transmission is observed instead.

DiscussionTo understand to what extent dissipation modifies the dressedstate picture, we have studied the transient dynamics of thepopulation difference P(t) in the presence of drive only (εp= 0).As discussed in the Methods, P(t) is governed by a generalizedmaster equation featuring the two nonequilibrium kernelsKþ=�ðtÞ, which in the absence of probe field, are symmetric/antisymmetric in the static bias ε0. In Laplace space, bysolving the pole equation λ+ K+ (λ)= 0, whereKþðλÞ ¼

R10 expð�λtÞKþðtÞ, the phase diagram of the driven

spin-boson particle can in principle be found along the linesdiscussed in the Methods. The kernel K+(λ) can be expressed asthe sum Kf(λ)+ Kb(λ) of the nonequilibrium forward and back-ward kernels

K f=bðλÞ ¼2

2

Z 1

0dt e�Q′ðtÞ�λt J0 dðtÞ½ �cos Q′′ðtÞ � ε0t½ �; ð6Þ

with dðtÞ ¼ 2εdω�1d sin ωdt=2ð Þ. The correlation function Q(t)=

Q′(t)+ iQ″(t) describes the environmental influence and itsexplicit form is discussed in the Supplementary Note 1 and inEqs. (15) and (16) of the Methods. For the present discussion, it isenough to observe that in the long-time limit t � τenv, whereτenv= (2παkBT/ħ)−1, the real part of Q(t) assumes the form Q′(t)~ t/τenv+ const. appropriate to white noise. Thus, τenv yields anestimate of the memory time of the kernels entering Eq. (6). Theimpact of the drive is encapsulated in the time-dependent argu-ment of the Bessel function of first kind J0. Depending on whetherωdτenv ≥ 1 (slow relaxation) or ωdτenv ≤ 1 (fast relaxation), twodistinct regimes corresponding to devices II and III are encoun-tered, respectively.

Let us focus on the first case, explored in Fig. 3a, b. In thisregime, one full cycle of the drive field is possible before envir-onmental effects induce a loss of coherence. Thus, we expect thatcoherent absorption and emission processes from the drive fieldtake place during a cycle. An expansion of the Bessel function inEq. (6) in a Fourier series, J0[d(t)]=

Pn J

2nðεd=ωdÞexp inωdð Þ,

shows that the channel with nωd= ε0, dominates the series12, andhence an effective two-level description with renormalized tun-neling splitting Δn applies. A solution of the pole equation in thisapproximation yields a renormalization of the crossover tem-perature T*(α) → T*(α)[Jn(εd/ωd)]1/(1−α). Because Jn < 1, the pumpfield always yields a reduction of quantum coherence. Near thezeros of Jn, quantum coherence is fully suppressed and an inco-herent decay is expected. This behavior is seen in Fig. 3c, d, wherewe show the simulated time evolution of P(t) as a function ofpump power at ε0= 0 and ε0= ωd, respectively. The color map ofP(t) displays coherent oscillations at low to moderate pumpamplitudes, where J0(εd/ωd) is still of order one. However, a fullsuppression of quantum coherence occurs near the first zero of J0,highlighted by a solid white circle. We notice that the almostcomplete standstill predicted to occur at the zeros of J0 for adissipation-free, symmetric two-level particle29, is destroyed byenvironmental relaxation processes, albeit on a very slow timescale. A similar suppression of coherence, together with a veryslow incoherent decay, is observed at the first resonance, ε0= ωd,shown in Fig. 3d, in correspondence with the first zero of J1.Independently of the initial preparation, the steady state popu-lation acquires the value P0= (Kf− Kb)/(Kf+ Kb), where Kf/b=Kf/b(λ= 0) are the nonequilibrium backward and forward rates.For the symmetric case shown in Fig. 3c, the backward andforward rates are equal and hence P0= 0. A genuine none-quilibrium behavior is observed in Fig. 3d in the region betweenthe first zeros of J0 and J1, where the steady state qubit populationP0 < 0, corresponding to a larger population of the left statedespite ε0 > 0. This phenomenon originates from the effectivedetailed balance relation

K f ¼ Kbe�hεeff =kBT ð7Þ

between the nonequilibrium backward and forward rates Kf/b.This equation implicitly defines the effective asymmetry εeff. Onlyin the absence of the drive does εeff coincide with the static bias ε0.We note that the use of an external coherent drive to tune thedirection of long-range electron chemical reactions via a drive-induced effective bias was originally proposed in refs. 30, 31.

Let us turn to the explanation of the results for Device IIIdisplayed in Fig. 3e–g, where ωdτenv � 1 applies. In this regimethe approximate result

χ ωp� �

¼ 14kBT

∂εeff=∂ε0cosh2 �hεeff=2kBTð Þ

γdγd þ iωp

ð8Þ

can be obtained from the exact expression Eq. (19) of theMethods section. This form is associated to the incoherentdynamics of the spin boson particle with nonequilibriumrelaxation rate γd≡ Kf+ Kb. At the symmetry point we have εeff= ε0= 0, with limε0!0∂εeff=∂ε0≠0. Correspondingly, the suscept-ibility χ″(ωp) has a peak at ωp= γd. An expansion in the smallparameter ωdτenv yields J0[d(t)] ≈ J0(εdt) and hence a relaxationrate γd which is independent of the driving frequency ωd, con-sistent with the experimental observation that the spectra dependweakly on ωd. The dependence on the pump amplitude εdremains, as clearly seen in Fig. 3e–g where the transmission at thesymmetry point smoothly increases for increasing drive ampli-tude. The transmission is almost complete for drive powers abovethe value (εd/Δ)2≃ 16 dB roughly corresponding to the second

NATURE COMMUNICATIONS | DOI: 10.1038/s41467-018-03626-w ARTICLE

NATURE COMMUNICATIONS | (2018) 9:1403 | DOI: 10.1038/s41467-018-03626-w |www.nature.com/naturecommunications 5

Page 6: Probing the strongly driven spin-boson model in a ... · A. Lupascu2,3,4, C.M. Wilson2,6 & M. Grifoni7 Quantum two-level systems interacting with the surroundings are ubiquitous in

zero of J0(εdτenv) (see Fig. 3f, where the black crosses highlight thefirst two zeroes). Regarding the transmission at finite static bias,we expect that no thermally assisted excitation is possible when�hεeff � kBT ; correspondingly the susceptibility vanishes, asaccounted by the term cosh�2ð�hεeff=2kBTÞ in Eq. (8). Thisbehavior is clearly seen in Fig. 3f, where the black dashed linecorresponds to the condition ħεeff= 2kBT. Below the dashed linethe effective bias is larger than the temperature and the signal isfully transmitted.

In conclusion, we have experimentally and theoreticallyexplored the paradigmatic driven spin-boson model in theunderdamped and ultrastrong dynamical regimes. Quantumcoherence is generally reduced or even destroyed by a drive fieldin a way which can be tuned by sweeping the drive amplitude andfrequency. The control of the dynamics is possible for a genericOhmic spin-boson particle, independently of its microscopicdetails. Localization and even population inversion can beattained by properly tuning the parameters of the coherent drive.Our results might find application in various physical, chemical,and quantum biology realizations of the driven spin-bosonmodel.

MethodsExperimental fabrication and measurement setup. Devices were fabricatedaccording to the procedure explained in ref. 9. Our setup was designed in such away that the reservoir (the photons in the transmission line) can still be consideredin equilibrium despite the strong pumping applied to the qubit. The response of thephotons depends on the intensity of the drive and on the coupling mechanisms. Inour experiment, the degrees of freedom of the bath are very weakly coupled to thedrive, compared to the qubit. Hence, even though the qubit is strongly driven, thebath is not. To be more quantitative, the most sensitive component of our bath isthe 50 Ohm input of our amplifier. From its data sheet, the amplifier starts tobecome nonlinear for an input power of −12 dBm (its 1 dB compression point),which is many orders of magnitude higher than what our pump power is. Theother components of our bath, which would be microwave attenuators (resistors),are linear up to energies a few orders of magnitude higher. From the theoreticalpoint of view, we expect that the transmission of the fully-driven spin-boson modelwould differ qualitatively from the one of the system-driven spin-boson modelconsidered in this work. No trivial mapping exists between the two models. Thevery good agreement between theoretical predictions and the experiment validateour conclusion that merely the system is driven.

Relation between theoretical and experimental observables. The flux operatorin the qubit basis is identified with Φ̂ ¼ f σz . The proportionality constant f is afitting parameter which, for low couplings, is estimated to be f=MIpers, withM thequbit-line mutual inductance and Ipers the persistent current in the super-conducting loop. This estimate provides values (see Table 1) which are not far fromthose obtained from fit to data for devices I and II and from qualitative analysis forDevice III. The externally applied tunable flux Φε is related to the static bias by ħε0= 2Ipers(Φε−Φ0/2), with Φ0 the magnetic flux quantum. The probe input voltage isconnected to the angular frequency εp yielding the theoretical probe amplitude, seeEq. (4), through V in

p tð Þ ¼ fZεp cos ωpt� �

, where the proportionality constant is fZ=

ħZ/f and Z is the line impedance. It follows that the constant N in Eq. (1) is givenby the ratio f/fZ.

Parameters used in the simulation. The parameters used in the numericalsimulations shown in Figs. 2 and 3 are provided in Table 1. Coupling α, baretunneling frequency Δ, and proportionality constant N are determined by fit todata of Tj j2 vs. ωp performed for the nondriven devices I and II at the symmetrypoint Φε=Φ0/2 (see Fig. 2d, e). Such fits along with their accuracy are shown inSupplementary Fig. 4. In Fig. 2, the measured value of 90 mK is used for thetemperature. Temperature values used in Fig. 3 account for a possibly highereffective temperature introduced by the drive at the qubit position. Specifically, forDevice II, in the presence of the pump drive, a better qualitative agreement betweensimulated and experimental transmission is obtained by assuming a higher tem-perature. As the qualitative features of the simulated transmission for Device III,operating at ultrastrong coupling, are weakly sensitive to variations of the tem-perature, we used the same value of temperature for the pump-probe and theprobe-only cases.

Driven spin-boson dynamics within the NIBA. The spin-boson model describesthe coupling of a two-level quantum system to a bath of harmonic oscillators32. Byassuming a coupling which linearly depends on the coordinates of the oscillators,one arrives at the famous spin-boson Hamiltonian

HðtÞ ¼ HqbðtÞ ��h2σz

Xi

ci ayi þ ai� �

þXi

�hωiayi ai; ð9Þ

where ai, ayi are bosonic annihilation and creation operators and the coefficients ci

are the amplitude of the interaction strength of the two-level system with mode i.The bosonic heat bath is fully characterized by the spectral functionGðωÞ ¼

Pi c

2i δ ω� ωið Þ. For Ohmic damping, G(ω) ∝ ω, as assumed in Eq. (3).

The Ohmic spin-boson problem owes its popularity to its ubiquity and to thevariety of parameter regimes it encompasses as the temperature T and the couplingstrength α are varied. We refer the readers to ref. 2 for an exhaustive treatment. Thedynamical properties of a driven spin-boson system in the strongly damped and inthe incoherent regimes, are well described within the so-called noninteracting-blipapproximation (NIBA). Furthermore, the NIBA captures well the dynamics of asymmetric (ε0= 0) spin-boson system in the whole parameter regime. The NIBAapproximation provides a generalized master equation (GME) for the evolution ofthe population difference P(t) with rates in second order in the bare tunnelingsplitting Δ but nonperturbative in α. Accounting for the presence of timedependent fields, the GME explicitly reads

_PðtÞ ¼Z t

t0

dt′ K�ðt; t′Þ � Kþðt; t′ÞPðt′Þ� �

: ð10Þ

The NIBA kernels K± , averaged over a pump period, are given by

Kþðt; t′Þ ¼ hþðt � t′Þcos ζðt; t′Þ� �

; ð11Þ

K�ðt; t′Þ ¼ h�ðt � t′Þsin ζðt; t′Þ� �

; ð12Þ

with

hþðtÞ ¼ 2e�Q′ðtÞcos Q}ðtÞ½ �J02εdωd

sinωdt2

� � ; ð13Þ

h�ðtÞ ¼ 2e�Q′ðtÞsin Q′′ðtÞ� �

J02εdωd

sinωdt2

� � : ð14Þ

The function Q(t)=Q′(t)+ iQ″(t) is the environmental correlation function.For the Ohmic spectral density function G(ω)= 2αω exp(−ω/ωc), α being thedimensionless coupling strength and ωc a high frequency cutoff, and in the scalinglimit �hωc � β�1 ¼ kBT , these functions have an explicit form2

Q′ðtÞ ¼ 2α lnffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1þ ω2

c t2

q sinhðπt=�hβÞπt=�hβ

; ð15Þ

Q}ðtÞ ¼ 2αarctan ωctð Þ: ð16Þ

The above formulas are accurate in all coupling regimes, provided that thecutoff frequency is large with respect to the other frequency scales involved. In thelong-time limit (t=β�h � 1) the real part of Q(t) assumes the form Q′(t) ~ t/τenv+const., where τenv= (2παkBT/ħ)−1. Thus the latter quantity determines thememory time of the kernels K± in Eqs. (11) and (12).

Table 1 Parameters used for simulations

Figures 2 and 3 Device I Device II Device III

ωc/2π (GHz) 65 65 65Ipers (nA) 600 280 250α 0.007 [fit] 0.21 [fit] 0.8a

Δ/2π (GHz) 4.04 [fit] 7.23 [fit] 8.0a

Figure 2 Device I Device II Device IIIT (mK) 90 90 90N (estimated) 0.03 [fit] (0.02) 1.1 [fit] (0.5) 8.0a (5–10)Figure 3 Device II Device IIIT (mK) 175a 90a

N 1.1a 16.0a

ωp/2π (GHz) 5.2 4.0ωd/2π (GHz) 9.0 3.0

aValue yielding qualitative agreement with the experiment, see Supplementary Note 7

ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/s41467-018-03626-w

6 NATURE COMMUNICATIONS | (2018) 9:1403 | DOI: 10.1038/s41467-018-03626-w |www.nature.com/naturecommunications

Page 7: Probing the strongly driven spin-boson model in a ... · A. Lupascu2,3,4, C.M. Wilson2,6 & M. Grifoni7 Quantum two-level systems interacting with the surroundings are ubiquitous in

The dynamical phase entering the kernels reads

ζðt; t′Þ ¼ ðt � t′Þε0 þεpωp

sin ωpt� �

� sin ωpðt′Þ� �� �

: ð17Þ

Note that in the absence of the probe field, εp= 0, the pump-averaged kernelsdepend only on the difference t− t′, i.e., K ± (t, t′)=K ± (t− t′), as in the staticcase. The latter is then recovered by additionally setting εd= 0. On the other hand,the probe-only setup is described by Eq. (10) upon setting εd= 0 in Eqs. (13) and(14). The dynamics shown in the insets of Fig. 2a–c are based on the numericalsolution of the GME (10) for ε(t)= 0, whereas in the time evolution of P(t) vs.pump power shown in panels c, d, and g of Fig. 3, only the probe field is set to zero.

The linear susceptibility. The linear susceptibility is related to the asymptoticprobability difference by

PasðtÞ ¼ P0 þ �hεp χðωpÞeiωp t þ χ �ωp� �

e�iωp t� �

; ð18Þ

where, in the NIBA, P0 reduces to the equilibrium value Peq= tanh(ħε0/2kBT) inthe absence of pump driving. The transmission T ðωpÞ and the susceptibility χ(ωp)shown in the theoretical plots of Figs. 2 and 3 are calculated by means of the exactNIBA expression12

P0 ¼K�ð0ÞKþð0Þ ; χðωpÞ ¼

Hþ ωp� �

� H� ωp� �

P0iωp þ Kþ iωp

� � ; ð19Þ

with superscripts ± denoting symmetric/antisymmetric functions of ε0. For ourpump-probe case we find

Hþ ωp� �

¼ 1�hωp

Z 1

0dt e�iωp t=2sin

ωpt

2

�h�ðtÞcos ε0tð Þ; ð20Þ

H� ωp� �

¼ �1�hωp

Z 1

0dt e�iωp t=2sin

ωpt

2

�hþðtÞsin ε0tð Þ; ð21Þ

KþðλÞ ¼Z 1

0dt e�λthþðtÞcos ε0tð Þ; ð22Þ

K�ðλÞ ¼Z 1

0dt e�λth�ðtÞsin ε0tð Þ: ð23Þ

Here K ± ðλÞ ¼R10 dτe�λτK± ðτÞ are the Laplace transforms of the pump-averaged

kernels in Eqs. (11) and (12) with εp= 0. The kernels K±(λ) are related to theforward and backward rates Kf/b(λ), introduced in Eq. (6), by K±= Kf ± Kb. Also,the incoherent rates for the static case are defined as kf/b= Kf/b(λ= 0,εd= 0). Fordevices I and II, in the absence of pump driving, we analytically evaluated theintegrals in Eqs. (20), (21), (22), and (23) and used the resulting expressions in thesusceptibility χ, Eq. (19), to perform fits to the data. In the limit ωpτenv � 1, Eq.(19) simplifies to Eq. (8) of the main text (see the Supplementary Note 4).

Coherent-to-incoherent transition. In the absence of probe driving, εp= 0, thepopulation difference P(t) is conveniently obtained by introducing the Laplacetransform P̂ðλÞ ¼

R10 dte�λtPðtÞ. From Eq. (10) one finds

P̂ðλÞ ¼ 1þ K�ðλÞ=λλþ KþðλÞ : ð24Þ

The pole in λ= 0 determines the asymptotic value P0= K− (0)/K+ (0) reachedat long times. The solution of the equation λ+ K+(λ)= 0 yields information on thetransient dynamics. In the underdamped regime, complex solutions yield therenormalized tunneling frequency with associated dephasing rate. In theincoherent regime, the long-time dynamics is ruled by a single exponential decaywith relaxation rate γd≡ K+ (λ= 0), see Eq. (22).

Let us focus exemplarily on the undriven spin-boson system at the symmetrypoint ε0= 0. Then, an expansion around λ= 0 yields a quadratic equation for thepoles of P̂ðλÞ33. In the coherent regime the roots are complex conjugated, λ1,2=−γ ± iΩ(T), while they are real in the incoherent regime (cf. insets in Fig. 2a–c).The temperature T* at which the oscillation frequency Ω(T) vanishes determinesthe transition between the coherent and incoherent regimes. For weak coupling onefinds for example Ω= Δr(1− παħΔr/kBT) with

r ¼ =ωcð Þα=ð1�αÞgðαÞ ð25Þ

and g(α)= [Γ(1− 2α)cos(πα)]1/2(1−α). This allows the estimate T*(α) ≈ ħΔr(kBα)−1

when ɑ≪ 1. For general α < 1 it is given by

T�ðαÞ �hrkB

ðαÞ=αð1� αÞ½ �1=2ð1�αÞ; ð26Þ

where Γ(x) is the Euler Gamma function. This approximate expression matcheswell the numerically calculated crossover temperature shown in Fig. 1c. Thecoherent-incoherent transition temperature T*(α) depicted there is established, forα < 0.5, by using Eq. (19), with numerically evaluated kernels, whereas the point atα= 0.5 is individuated by the exact result kBT*(α= 0.5)/ħΔ= Δ/2ωc

2. Furtherdetails are found in the Supplementary Note 9.

Data availability. The data that support the main findings of this study areavailable from the corresponding author upon request.

Received: 7 December 2017 Accepted: 28 February 2018

References1. Leggett, A. J. et al. Dynamics of the dissipative two-state system. Rev. Mod.

Phys. 59, 1–85 (1987).2. Weiss, U. Quantum dissipative systems 4th edn (World Scientific, Singapore,

2012).3. Golding, B., Zimmerman, M. N. & Coppersmith, S. N. Dissipative quantum

tunneling of a single microscopic defect in a mesoscopic metal. Phys. Rev. Lett.68, 998–1001 (1992).

4. Golding, B., Graebner, J. E., Kane, A. B. & Black, J. L. Relaxation of tunnelingsystems by conduction electrons in a metallic glass. Phys. Rev. Lett. 41,1487–1491 (1978).

5. Han, S., Lapointe, J. & Lukens, J. E. Observation of incoherent relaxation bytunneling in a macroscopic two-state system. Phys. Rev. Lett. 66, 810–813(1991).

6. Morillo, M. & Cukier, R. I. Solvent effects on proton transfer reactions. J.Chem. Phys. 91, 857–863 (1989).

7. Thorwart, M., Eckel, J., Reina, J. H., Nalbach, P. & Weiss, S. Enhancedquantum entanglement in the non-Markovian dynamics of biomolecularexcitons. Chem. Phys. Lett. 478, 234–237 (2009).

8. Huelga, S. F. & Plenio, M. B. Vibrations, quanta and biology. Contemp. Phys.54, 181–207 (2013).

9. Forn-Díaz, P. et al. Ultrastrong coupling of an artificial atom to anelectromagnetic continuum. Nat. Phys. 13, 39–43 (2017).

10. Peropadre, B., Zueco, D., Porras, D. & Garca-Ripoll, J. J. Nonequilibrium andnonperturbative dynamics of ultrastrong coupling in open lines. Phys. Rev.Lett. 111, 243602 (2013).

11. Yoshihara, F. et al. Superconducting qubit-oscillator circuit beyond theultrastrong-coupling regime. Nat. Phys. 13, 44–47 (2017).

12. Grifoni, M. & Hänggi, P. Driven quantum tunneling. Phys. Rep. 304, 229–358(1998).

13. Nakamura, Y., Pashkin, Yu. A. & Tsai, J. S. Rabi oscillations in a Josephson-Junction charge two-level system. Phys. Rev. Lett. 87, 246601 (2001).

14. Wilson, C. M. et al. Coherence times of dressed states of a superconductingqubit under extreme driving. Phys. Rev. Lett. 98, 257003 (2007).

15. Wilson, C. M. et al. Dressed relaxation and dephasing in a strongly driventwo-level system. Phys. Rev. B 81, 024520 (2010).

16. Oliver, W. D. et al. Mach-Zehnder interferometry in a strongly drivensuperconducting qubit. Science 310, 1653–1657 (2005).

17. Sillanpää, M., Lehtinen, T., Paila, A., Makhlin, Y. & Hakonen, P. Continuous-time monitoring of Landau-Zener interference in a Cooper pair box. Phys.Rev. Lett. 96, 187002 (2006).

18. Berns, D. et al. Amplitude spectroscopy of a solid-state artificial atom. Nature455, 51–57 (2008).

19. Deng., C. et al. Observation of Floquet states in a strongly driven artificialatom. Phys. Rev. Lett. 115, 133601 (2015).

20. Mooij, J. E. et al. Josephson persistent-current qubit. Science 285, 1036–1039(1999).

21. Kubo, R. Statistical mechanics of irreversible processes. J. Phys. Soc. Jpn. 12,570–586 (1957).

22. Grifoni, M., Sassetti, M., Stockburger, J. & Weiss, U. Nonlinear response of aperiodically driven damped two-level system. Phys. Rev. E 48, 3497–3509(1993).

23. Restrepo, S., Cerillo, J., Bastidas, V. M., Angelakis, D. G. & Brandes, T. Drivenopen quantum systems and Floquet stroboscopic dynamics. Phys. Rev. Lett.117, 250401 (2016).

24. Grifoni, M., Sassetti, M., Hänggi, P. & Weiss, U. Cooperative effects in thenonlinearly driven spin-boson system. Phys. Rev. E 52, 3596 (1995).

25. Daz-Camacho, G., Bermudez, A. & Garca-Ripoll, J. J. Dynamical polaronansatz: a theoretical tool for the ultra-strong coupling regime of circuit QED.Phys. Rev. A. 93, 043843 (2016).

26. Shi, T., Chang, Y. & García-Ripoll, J. J. Ultrastrong coupling few-photonscattering theory. Preprint at http://arxiv.org/abs/1701.04709 (2017).

NATURE COMMUNICATIONS | DOI: 10.1038/s41467-018-03626-w ARTICLE

NATURE COMMUNICATIONS | (2018) 9:1403 | DOI: 10.1038/s41467-018-03626-w |www.nature.com/naturecommunications 7

Page 8: Probing the strongly driven spin-boson model in a ... · A. Lupascu2,3,4, C.M. Wilson2,6 & M. Grifoni7 Quantum two-level systems interacting with the surroundings are ubiquitous in

27. Shevchenko, S. N., Ashhab, S. & Nori, F. Landau-Zener-Stückelberginterferometry. Phys. Rep. 492, 1–30 (2010).

28. Hausinger, J. & Grifoni, M. Dissipative two-level system under strong ac-driving: a combination of Floquet and Van Vleck perturbation theory. Phys.Rev. A. 81, 022117 (2010).

29. Grossmann, F., Dittrich, T., Jung, P. & Hänggi, P. Coherent destruction oftunneling. Phys. Rev. Lett. 67, 516–519 (1991).

30. Dakhnovski, Y. & Coalson, R. D. Manipulating reactant-product distributionsin electron transfer reactions with a laser field. J. Chem. Phys. 103, 2908–2916(1995).

31. Goychuk, I. A., Petrov, E. G. & May, V. Control of the dynamics of adissipative two-level system by a strong periodic field. Chem. Phys. Lett. 253,428–437 (1996).

32. Caldeira, O. & Leggett, A. J. Quantum tunneling in a dissipative system. Ann.Phys. 149, 374–456 (1987).

33. Weiss, U. & Grabert, H. Effects of temperature and bias on macroscopicquantum coherence. Europhys. Lett. 2, 667–672 (1986).

AcknowledgementsThe authors acknowledge financial support by the Deutsche Forschungsgemeinschaft viaSFB 631, NSERC of Canada, the Canadian Foundation for Innovation, the OntarioMinistry of Research and Innovation, Industry Canada and Canadian MicroelectronicsCorporation. L.M. gratefully acknowledges financial support by the Angelo Della RicciaFoundation and hospitality by the Regensburg University during the early stages of theproject. P.F.-D. is supported by the Beatriu de Pinós fellowship (2016BP00303). Theauthors thank J.J. García-Ripoll, B. Peropadre, and P. Hänggi for fruitful discussions, andS. Chang, A. M., and C. Deng for help with device fabrication and with the measurementsetups.

Author contributionsL.M. and M.G. performed the theoretical analysis, with numerical simulations carried outby L.M. The experiments were designed and performed by P.F.-D., A.L., and C.M.W. Thedevices were fabricated by P.F.-D., J.-L.O., M.A.Y., and M.R.O. contributed to devicedesign and fabrication. R.B. assisted in numerical modeling of the device. The manuscript

was mainly written by M.G. with critical comments provided by all authors. The sup-plementary information was mainly written by L.M.. The experimental work was acollaboration between the labs led by A.L. and C.M.W.

Additional informationSupplementary Information accompanies this paper at https://doi.org/10.1038/s41467-018-03626-w.

Competing interests: The authors declare no competing interests.

Reprints and permission information is available online at http://npg.nature.com/reprintsandpermissions/

Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims inpublished maps and institutional affiliations.

Open Access This article is licensed under a Creative CommonsAttribution 4.0 International License, which permits use, sharing,

adaptation, distribution and reproduction in any medium or format, as long as you giveappropriate credit to the original author(s) and the source, provide a link to the CreativeCommons license, and indicate if changes were made. The images or other third partymaterial in this article are included in the article’s Creative Commons license, unlessindicated otherwise in a credit line to the material. If material is not included in thearticle’s Creative Commons license and your intended use is not permitted by statutoryregulation or exceeds the permitted use, you will need to obtain permission directly fromthe copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

© The Author(s) 2018

ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/s41467-018-03626-w

8 NATURE COMMUNICATIONS | (2018) 9:1403 | DOI: 10.1038/s41467-018-03626-w |www.nature.com/naturecommunications