Realization of a system with metastable states of a ... · based on variants of the transmon qubit...

11
Realization of a Λ system with metastable states of a capacitively-shunted fluxonium N. Earnest, 1, * S. Chakram, 1 Y. Lu, 1 N. Irons, 2 R. K. Naik, 1 N. Leung, 1 L. Ocola, 3 D.A. Czaplewski, 3 B. Baker, 2 Jay Lawrence, 4 Jens Koch, 2 and D. I. Schuster 1, 1 The James Franck Institute and Department of Physics, University of Chicago, Chicago, Illinois 60637, USA 2 Department of Physics and Astronomy, Northwestern University, Evanston, Illinois 60208, USA 3 Argonne National Laboratories, Center for Nanoscale Materials, Argonne, Illinois 60439, USA 4 Department of Physics, Dartmouth College, Hanover, New Hampshire 03755, USA We realize a Λ system in a superconducting circuit, with metastable states exhibiting lifetimes up to 8ms. We exponentially suppress the tunneling matrix elements involved in spontaneous energy relaxation by creating a “heavy” fluxonium, realized by adding a capacitive shunt to the original circuit design. The device allows for both cavity-assisted and direct fluorescent readout, as well as state preparation schemes akin to optical pumping. Since direct transitions between the metastable states are strongly suppressed, we utilize Raman transitions for coherent manipulation of the states. PACS numbers: Quantum computation with superconducting circuits has seen rapid progress over the past decade [1–3] largely due to improvements in qubit coherence [4–6]. Perform- ing large-scale quantum computation, error correction and simulation, will require significantly longer coherence times [7, 8]. Fermi’s golden rule states that qubit life- times are governed by two factors: (1) the noise spectral density associated with environmental degrees of free- dom; (2) the transition matrix elements, which are de- termined by the qubit wavefunctions. To date, improve- ments in superconducting qubit lifetimes have primarily been achieved by modifying the noise spectral density – for example, by filtering [9, 10] and by minimizing the contributions of two-level systems [6, 11]. We demon- strate a complimentary approach, making the qubit in- sensitive to environmental noise by reducing the transi- tion matrix elements. This leads to qubit lifetimes as high as 8 ms, and realizes a Λ system analogous to those commonly found in atomic systems. Most current superconducting qubit architectures are based on variants of the transmon qubit [5, 12–14]. Transmons have large dipole matrix elements, simple se- lection rules, and a small non-linearity, sufficient to re- solve the lowest energy levels as the qubit states. In contrast, flux qubits [15–17] have a large nonlinearity, rich level structure, and selection rules that can be finely engineered to yield a smooth trade-off between decay ma- trix elements and gate fidelities. With the realization of a linear superinductance and the fluxonium qubit [18], this class of qubits has seen enhanced lifetimes and re- duced flux-noise induced decoherence [19]. These prop- erties make the fluxonium a promising system for engi- neering a Λ system. Traditionally, a Λ system is comprised of a ground and metastable excited state, coherently coupled through a third intermediate state. Λ systems are ubiquitous in atomic physics, realized using a combination of selection rules [20], relative strengths of optical-dipole and mi- crowave hyperfine matrix elements [21], and large differ- ences in frequency scales in conjunction with the 3D den- sity of states (decay rate γ ν 3 ) [22]. Superconducting qubits are typically not protected by symmetry-based se- lection rules, and possess a much smaller dynamic range of frequency scales and a 1D density of states (γ ν ), making it more challenging to realize the metastability required to explore the physics associated with Λ sys- tems. Previous work in cQED has utilized the Purcell ef- fect [23] to modify the density of states and explore multi- tone coherent interactions of three-level systems [24, 25]. In this work, we present a Λ system in a capacitively shunted fluxonium circuit: the heavy fluxonium. The added capacitance further localizes the lowest energy states, exponentially suppressing the dipole matrix ele- ments and boosting the metastable state lifetime to 8 ms. The suppressed matrix elements make controlled popu- lation transfer to this state a challenge, but we surmount this by using multi-tone Raman transitions in the Λ sys- tem to perform coherent operations with substantial im- provement in gate fidelities relative to direct transitions. The heavy fluxonium circuit (Fig. 1a,b) consists of a small-area Josephson junction connected in parallel to a capacitance (C q ) and a large superinductor (L JA ), real- ized as an array of 100 large-area Josephson junctions. To ensure idealized inductive behavior of the array, the length and individual junction size must satisfy several conditions as explained in [18]. Once these conditions are satisfied, the Hamiltonian of the fluxonium is given by: H f = -4E C d 2 2 - E J cos(ϕ - 2πΦ ext /Φ 0 )+ 1 2 E L ϕ 2 (1) where E C = e 2 /2C q is the charging energy, E J the Josephson energy of the small junction, and E L = Φ 2 0 /2L JA the inductive energy of the Josephson junction array. In contrast to earlier fluxonium devices [18], the heavy fluxonium shunts the small junction with a large arXiv:1707.00656v2 [quant-ph] 13 Apr 2018

Transcript of Realization of a system with metastable states of a ... · based on variants of the transmon qubit...

Page 1: Realization of a system with metastable states of a ... · based on variants of the transmon qubit [5, 12{14]. Transmons have large dipole matrix elements, simple se-lection rules,

Realization of a Λ system with metastable states of a capacitively-shunted fluxonium

N. Earnest,1, ∗ S. Chakram,1 Y. Lu,1 N. Irons,2 R. K. Naik,1 N. Leung,1 L. Ocola,3

D.A. Czaplewski,3 B. Baker,2 Jay Lawrence,4 Jens Koch,2 and D. I. Schuster1, †

1The James Franck Institute and Department of Physics,University of Chicago, Chicago, Illinois 60637, USA

2Department of Physics and Astronomy, Northwestern University, Evanston, Illinois 60208, USA3Argonne National Laboratories, Center for Nanoscale Materials, Argonne, Illinois 60439, USA

4Department of Physics, Dartmouth College, Hanover, New Hampshire 03755, USA

We realize a Λ system in a superconducting circuit, with metastable states exhibiting lifetimes upto 8 ms. We exponentially suppress the tunneling matrix elements involved in spontaneous energyrelaxation by creating a “heavy” fluxonium, realized by adding a capacitive shunt to the originalcircuit design. The device allows for both cavity-assisted and direct fluorescent readout, as well asstate preparation schemes akin to optical pumping. Since direct transitions between the metastablestates are strongly suppressed, we utilize Raman transitions for coherent manipulation of the states.

PACS numbers:

Quantum computation with superconducting circuitshas seen rapid progress over the past decade [1–3] largelydue to improvements in qubit coherence [4–6]. Perform-ing large-scale quantum computation, error correctionand simulation, will require significantly longer coherencetimes [7, 8]. Fermi’s golden rule states that qubit life-times are governed by two factors: (1) the noise spectraldensity associated with environmental degrees of free-dom; (2) the transition matrix elements, which are de-termined by the qubit wavefunctions. To date, improve-ments in superconducting qubit lifetimes have primarilybeen achieved by modifying the noise spectral density –for example, by filtering [9, 10] and by minimizing thecontributions of two-level systems [6, 11]. We demon-strate a complimentary approach, making the qubit in-sensitive to environmental noise by reducing the transi-tion matrix elements. This leads to qubit lifetimes ashigh as 8 ms, and realizes a Λ system analogous to thosecommonly found in atomic systems.

Most current superconducting qubit architectures arebased on variants of the transmon qubit [5, 12–14].Transmons have large dipole matrix elements, simple se-lection rules, and a small non-linearity, sufficient to re-solve the lowest energy levels as the qubit states. Incontrast, flux qubits [15–17] have a large nonlinearity,rich level structure, and selection rules that can be finelyengineered to yield a smooth trade-off between decay ma-trix elements and gate fidelities. With the realization ofa linear superinductance and the fluxonium qubit [18],this class of qubits has seen enhanced lifetimes and re-duced flux-noise induced decoherence [19]. These prop-erties make the fluxonium a promising system for engi-neering a Λ system.

Traditionally, a Λ system is comprised of a ground andmetastable excited state, coherently coupled through athird intermediate state. Λ systems are ubiquitous inatomic physics, realized using a combination of selectionrules [20], relative strengths of optical-dipole and mi-

crowave hyperfine matrix elements [21], and large differ-ences in frequency scales in conjunction with the 3D den-sity of states (decay rate γ ∝ ν3) [22]. Superconductingqubits are typically not protected by symmetry-based se-lection rules, and possess a much smaller dynamic rangeof frequency scales and a 1D density of states (γ ∝ ν),making it more challenging to realize the metastabilityrequired to explore the physics associated with Λ sys-tems. Previous work in cQED has utilized the Purcell ef-fect [23] to modify the density of states and explore multi-tone coherent interactions of three-level systems [24, 25].

In this work, we present a Λ system in a capacitivelyshunted fluxonium circuit: the heavy fluxonium. Theadded capacitance further localizes the lowest energystates, exponentially suppressing the dipole matrix ele-ments and boosting the metastable state lifetime to 8 ms.The suppressed matrix elements make controlled popu-lation transfer to this state a challenge, but we surmountthis by using multi-tone Raman transitions in the Λ sys-tem to perform coherent operations with substantial im-provement in gate fidelities relative to direct transitions.

The heavy fluxonium circuit (Fig. 1a,b) consists of asmall-area Josephson junction connected in parallel to acapacitance (Cq) and a large superinductor (LJA), real-ized as an array of 100 large-area Josephson junctions.To ensure idealized inductive behavior of the array, thelength and individual junction size must satisfy severalconditions as explained in [18]. Once these conditionsare satisfied, the Hamiltonian of the fluxonium is givenby:

Hf = −4ECd2

dϕ2−EJ cos(ϕ−2πΦext/Φ0)+

1

2ELϕ

2 (1)

where EC = e2/2Cq is the charging energy, EJ theJosephson energy of the small junction, and EL =Φ2

0/2LJA the inductive energy of the Josephson junctionarray. In contrast to earlier fluxonium devices [18], theheavy fluxonium shunts the small junction with a large

arX

iv:1

707.

0065

6v2

[qu

ant-

ph]

13

Apr

201

8

Page 2: Realization of a system with metastable states of a ... · based on variants of the transmon qubit [5, 12{14]. Transmons have large dipole matrix elements, simple se-lection rules,

2

Cin

Cout

CC

LR CR

Cq EJ ···Φext LJA

-2π -2π-4π 4π2π 2π0 0Phase, ϕ Phase, ϕ

|g0〉|g-1〉

|f0〉

|e0〉

|g1〉 |g0〉

|e0〉

|g1〉

|e1〉

|g-1〉 |g2〉

Ene

rgy

a b

c d

FIG. 1: a) Equivalent circuit diagram of the heavy fluxoniumcapacitively coupled to a readout resonator (SupplementaryInformation). b) Scanning Electron Microscope image of thedevice, with a magnified view of the 100 Josephson junc-tion array, fabricated using the bridgeless method detailed in[26]. c) Simulated Potential energy landscape/wavefunctionsat Φext = 0.02 Φ0 demonstrating localized wavefunctions inthree wells. d) Simulated Potential energy landscape atΦext = 0.51Φ0, where |g0〉 and |g1〉 are nearly degenerate.

capacitance (43 fF, dashed red squares in Fig. 1b). Thisresults in a reduced EC/h = 0.46 GHz, increases the ef-fective mass of the phase degree of freedom, and pro-duces quasi-localized states in the different wells of thepotential (see Fig. 1c,d). The other circuit parameters,EJ/h = 8.11 GHz and EL/h = 0.24 GHz, are compa-rable to those in previous fluxonium devices. We labelstates by their fluxoid number −1, 0, 1 (number of fluxquanta in the loop formed by the junctions), and by theplasmon levels within that well |g〉 , |e〉 , |f〉. The heavyfluxonium allows for two types of transitions: intra-wellplasmons (e.g., |g0〉 ↔ |e0〉), and inter-well fluxons (e.g.,|g0〉 ↔ |g1〉).

Inter-well transitions involve states with wavefunctionssuch as ψg0(ϕ) and ψg1(ϕ), which are disjoint. Accord-

ingly, matrix elements∫dϕψ∗g1(ϕ)O ψg0(ϕ) with respect

to local operators O(d/dϕ, ϕ) will be exponentially sup-pressed with ∼ exp[−π2(EJ/8EC)1/2], inferred from con-sidering the tails of displaced harmonic-oscillator wave-functions [27]. Consequently, the |g1〉 state of the heavyfluxonium is much longer lived relative to the originalfluxonium. However, the suppressed transition matrix el-ements also make coherent operations more challenging.This circuit resembles a recently reported design [28],

|g0〉 → |e0〉

|g1〉 → |e1〉

|g0〉 → |e1〉

|g1〉 → |e0〉|g0〉 → |g1〉

|g0,0r〉 → |g0,1r〉

|g1,0r〉 → |g1,1r〉 |g1,0r〉 → |g0,1r〉

|g0,0r〉 → |g1,1r〉

Flux (Φext/Φ0)

Pro

be F

requ

ency

(GH

z)

2te

2tg

FIG. 2: Single-tone spectroscopy of the fluxonium-resonatorsystem in the vicinity of the resonator and primary plasmontransition frequencies. Dashed lines indicate simulated energylevels of the coupled system based on device parameters ex-tracted from fits to single and two-tone spectra. Transitionsthat change rapidly with flux are inter-well fluxon transitions,while the flatter transitions are intra-well plasmon transitions.Left inset: interference due to coupling between the groundand excited states of neighbouring wells. Right inset: featuresassociated with fluxon transitions crossing with the resonator.The spectrum is normalized by the transmission amplitude ofthe bare resonator (Supplementary Information).

whose dipole moment (and thereby the fluxon transitionrate) is tunable through the use of a SQUID in placeof a single Josephson junction. Unlike the fluxonium in[28], our heavy fluxonium – with a fixed EJ/EC ≈ 18– is sufficiently heavy to disallow coherent direct drives.We solve this issue by realizing a Λ system between theground state |g0〉, the metastable state |g1〉, and the ex-cited state |f0〉, and perform coherent Raman transitionsbetween |g0〉 and |g1〉.

For fast readout, the heavy fluxonium is capacitivelycoupled to a lossy resonator (Q ∼ 500). The Hamiltonianof the combined system is given by [29]:

HS = Hf + hνra†a+

∑j,k

hg |j〉 〈k| 〈j| n |k〉 (a+ a†), (2)

where, νr = 4.95 GHz is the bare frequency of the res-onator and g = 76 MHz is the coupling between the res-onator and fluxonium (as extracted from fits to spectra).n is the charge operator of the fluxonium and controlsthe transition rates arising from driving on the input port(Cin in Fig. 1a).

Single-tone spectroscopy (Fig. 2) reveals both the res-onator photon and the plasmon transitions. The curva-ture of the plasmon transitions arises from flux-induceddistortion of the well (Supplementary Information), andallows one to easily distinguish between wells (blue andmagenta lines in Fig. 2). Furthermore, the strong hy-bridization of the plasmon and resonator (detuned byup to 155 MHz) allows for fluorescent readout of themetastable state, over the entire flux range, through cy-

Page 3: Realization of a system with metastable states of a ... · based on variants of the transmon qubit [5, 12{14]. Transmons have large dipole matrix elements, simple se-lection rules,

3

Δ

νpump

νprobe

|g0〉

|g1〉|e0〉

|f0〉

Flux (Φext /Φ0) Pump Detuning, δνpump (MHz)

Pro

be F

requ

ency

, νpr

obe (

GH

z)

Pum

p Fr

eque

ncy,

νpu

mp (

GH

z)

2νpump-νprobe = Eg1-Eg0

a b

FIG. 3: a) Two-tone spectroscopy showing direct fluxon transitions (orange and green lines) and two-photon transitions tothe two-excitation manifold (|f0〉, |e−1〉,|e1〉). State labeling for the transitions is valid for Φext > 0. The |f0〉 level serves asthe intermediate state in a Λ system comprising the ground state |g0〉 and the metastable |g1〉 state, and assists in Ramantransitions. b) Pump-probe spectroscopy of Raman transitions between |g0〉 and |g1〉 as a function of pump (near |g0〉 → |f0〉-2γtransition) and probe frequency (near |g1〉 → |f0〉). The Raman transition is seen when 2νpump−νprobe = Eg1−Eg0 , representedby the dashed line. The upper-left inset shows wavefunctions of the states involved in the transition. The intermediate |f0〉 statecouples to |g0〉 via a two-photon process, and has a small amplitude in the right well, with a direct dipole-allowed transitionto the metastable |g1〉 state. The dashed lines are simulated energy levels of the fluxonium-resonator system. The colorbar isnormalized by the transmission of the bare resonator.

cling the plasmon transition of the metastable state manytimes, similar to quantum non-demolition measurementsof single trapped ions and atoms [30, 31].

The tunnel splitting between the wells can be directlyobserved in the plasmon spectrum at Φext = Φ0/2. Atthis flux location, there are two identical wells with de-generate ground and first excited states. This results inthe feature shown in the left inset of Fig. 2, where theinterference of the levels results in a unique rhombus-shaped avoided crossing. The separation of the levelcrossings forming the top and bottom corners of therhombus (black arrows) is a direct measure of the tun-nel coupling of the excited states in the well (|e0〉 and|e1〉), corresponding to te ≈ 7 MHz. The tunnel split-ting between ground states is smaller than the linewidthof the plasmon and fluxon transitions, and is inferredfrom the fits to be tg ≈ 0.42 MHz, one thousand timessmaller than in previous experiments [18]. Another setof avoided crossings is visible in the resonator trans-mission peak (right inset of Fig. 2) at Φext ≈ 0.5Φ0.The outer set of crossings arise from the |g1〉 → |e0〉fluxon transition (also seen in the bottom of the left in-set), while the inner crossings are formed by compositeresonator/fluxon transitions: |g1, 0res〉 → |g0, 1res〉 and|g0, 0res〉 → |g1, 1res〉. The latter indicate that couplingbetween fluxon transitions is increased when a photon ispresent in the resonator (Supplementary Information forphoton-assisted fluxon transitions). The fluxon transi-tions are (to first order) linear in flux, with slopes givenby ∂f/∂Φext ≈ ±4π2EL/Φ0 = ±9.59 GHz/Φ0.

Fluxon transitions, not seen in single-tone spec-troscopy, can be identified via two-tone spectroscopy inwhich we monitor the transmission of the readout res-onator while sweeping the frequency of a second drivetone. The lines of largest slope are the single-photoninter-well fluxon transitions |g0〉 → |g±1〉. The rest ofthe lines are two-photon transitions to the second-excitedmanifold of the fluxonium-resonator system, with flat fea-tures corresponding to plasmons and sloped resonancescorresponding to fluxons. Of particular importance arethe two-photon features located at ∼ 4.73 GHz corre-sponding to the |g0〉 → |f0〉 two-photon transition, whichwill assist in performing coherent operations on the qubit.

The heavy fluxonium energy-level structure allows fora variety of state-preparation schemes. We can performT1 measurements from the highest fluxon transition fre-quency of 4.65 GHz down to about 3 GHz by directlydriving the fluxon transition at high powers to realizea classically mixed state (100µs pulse duration). Belowthis point, we perform T1 measurements using a processthat is similar to optical pumping [32]. Through contin-uous cycling of the bright |g0〉 → |e0〉 plasmon in Fig. 2,we take advantage of a small probability of decaying from|e0〉 to |g1〉 arising from the finite matrix element betweenthese states and incoherently “pump” the system into the|g1〉 state, and perform a typical T1 measurement.

Since direct fluxon transitions are forbidden, we re-alize faster gates by means of Raman transitions thatutilize the excited levels of the fluxonium, in analogywith atomic physics. Recently, such multi-tone transi-

Page 4: Realization of a system with metastable states of a ... · based on variants of the transmon qubit [5, 12{14]. Transmons have large dipole matrix elements, simple se-lection rules,

4

0.0 0.1 0.2 0.3 0.4 0.5Flux ( ext/ 0)

0

1

10T 1

(ms)

10

g 0|n

|g1

2

0 15 30 45Time (ms)

0.0

0.5

Popu

latio

n

0 500 1000Idle time [ns]

0.4

0.5

0.6

0.7

Excit

ed st

ate

popu

latio

n

0 1 2Rabi drive time ( s)

1

0

1

Prob

e de

tuni

ng (M

Hz)

0

1

a

b c

FIG. 4: a) Energy relaxation time (T1) as a function of mag-netic flux measured through a combination of direct-drive to amixed state, plasmon pumping, and Raman transitions. Thedashed purple line indicates the inverse square of the chargematrix element of the fluxon transition of interest. The insetshows the T1 decay curve of the point indicated by the redstar, after driving to a mixed state following a long Ramandrive. (b) Rabi chevron obtained by detuning the probe-drivetone away from the Raman transition described in Fig. 3bat Φext = 0.078 Φ0. The Raman transition is chosen to be60 MHz from the |f0〉 level and the peak π pulse fidelity is∼ 90%. (c) Ramsey experiment at Φext = 0.078 Φ0, obtainedusing π/2 pulses extracted from Rabi drive of the Ramantransition resulting in a T ∗2 of 500-550 ns.

tions have been used in superconducting qubits in thecontext of stabilization, and coherent population trap-ping [16, 24, 25, 33–35]. As tunneling is suppressed ex-ponentially by the depth of the well, it is advantageousto use higher plasmon excited states. Of particular im-portance is the |g0〉 → |f0〉 transition shown in the in-set of Fig. 3. Though the direct transition is disallowedby the symmetry of the wavefunctions, we can access itthrough a two-photon process mediated by the |e0〉 level.Further, from the inset in Fig. 3b we can see that the|f0〉 wavefunction has a noticeable amplitude in the rightwell, and |g1〉 → |f0〉 is dipole allowed. This indicatesthat we can use the |g0〉, |f0〉, and |g1〉 states to forma Λ system. We explore Raman transitions in this Λsystem by sweeping the pump and probe tone frequen-cies in the vicinity of these transitions, as shown in Fig.3b. We find a shift in the resonator transmission when2νpump − νprobe = Eg1 − Eg0 , corresponding to the in-tended transfer of population from |g0〉 → |g1〉. The Ra-man transition rate is related to the Rabi rates of the two

Raman tones, Ωprobe from |g1〉 → |f0〉, and Ωpump from|g0〉 → |f0〉 according to:

Ωg0g1 =ΩprobeΩ2

pump

∆δ2γ, (3)

where ∆ = 2νpump−Ef0 = νprobe−(Ef0−Eg1) is the de-tuning of the pump and probe tone from the two-photonresonance, while δ2γ = Ee0−νpump is the detuning of thetwo-photon |g0〉 → |f0〉 pump tone from the intermedi-ate |e0〉 state. Ωpump and Ωprobe are set by the strengthof the drive and by the charge matrix elements of the|g1〉 → |f0〉 and |g0〉 → |e0〉 transitions, respectively.

Having established the Λ system and the necessarytones required to perform a Raman transition betweenthe otherwise forbidden metastable states, we induceRabi oscillations by simultaneously switching on resonantpump and probe drives. The pump is detuned 30 MHzfrom the two-photon |f0〉 transition, and the probe fre-quency is chosen to be νprobe = 2νpump − ∆Eg1g0 . Ata flux value of 0.078 Φ0 we achieve a π pulse rate oftπ ∼ 400 ns with 90% contrast (Fig. 4b). While thisfidelity can be further optimized in future devices, itdemonstrates several orders of magnitude improvementfrom the direct drive which takes 100µs to generatea classically mixed state (Supplementary Information).The upper limit of the Raman transition rate arisesfrom off-resonant excitation of the resonator through thetwo-photon pump drive, which drives the |g0, 1res〉 →|g1, 1res〉 transition.

Using these different methods, direct driving, plasmonpumping, and a three-photon Raman transition, we mea-sure the T1 of the device over the flux range 0 ≤ |Φext| <0.45 Φ0, as shown in Fig. 4a. Plotting the T1 versus fluxshows improvement as we move toward 0.5 Φ0 and followsthe (inverse square of the) charge matrix elements. Thisindicates that the T1 is limited by a charge-based lossmechanism, as was also observed in the recent work on asimilar fluxonium device [28]. Furthermore, we success-fully measure the coherence of the fluxon transition usinga standard Ramsey sequence (Fig. 4c) with π

2 pulses ex-tracted from Rabi oscillations (Fig. 4b). The T ∗2 is mea-sured to be 500− 550 ns. Using the measured flux slopeand assuming a 1/f form, this corresponds to a flux noisespectral density Sφ(1 Hz) = 1.3µΦ0/

√Hz, comparable to

the flux noise measured for tunable transmons with sim-ilar magnetic shielding. This indicates that T ∗2 shouldbe improved by increasing the chain inductance, sincethe transition flux slope is given by ∂f/∂Φext ∼ 1/L. Aspin-echo experiment using Raman transition-based π/2and π pulses gives a T2,echo of 1.3µs with a single insertedπ pulse.

In summary, we have realized a heavy fluxonium in a2D cQED architecture, with metastable states exhibit-ing lifetimes of several milliseconds over a broad range offlux values, likely limited by a charge-based loss mech-anism. We study the coherence of the device by state

Page 5: Realization of a system with metastable states of a ... · based on variants of the transmon qubit [5, 12{14]. Transmons have large dipole matrix elements, simple se-lection rules,

5

preparation schemes that use the rich energy level struc-ture of the device, including a process analogous to op-tical pumping. We perform coherent operations on thelong-lived metastable states using a three-photon Ramantransition using an excited plasmon level as the interme-diate state, realizing single-fluxon gates (tπ ∼ 400 ns)that are several orders of magnitude faster than di-rectly driving the fluxon transition with comparable mi-crowave powers. Additionally, the relative proximity ofthe plasmon and readout resonator allow for photon andplasmon-mediated transitions, that could be useful forhigh-fidelity fluorescent readout and photon detectionschemes with cQED (Supplementary Information).

In future work, we seek to improve the speed and fi-delity of inter-well transitions, by increasing the lifetimeof the plasmon states, by using more sophisticated multi-tone techniques[36], and by increasing the inductance toreduce dephasing rates. The fabrication techniques de-veloped here will be useful for other types of protectedqubits including the 0 − π [37] and Josephson rhombuschain qubits [38].

We thank R. Ma for help with the pulsed measure-ment setup. This material is based upon work sup-ported by the Army Research Office under (W911NF-15-1-0421), and by the National Science Foundation Gradu-ate Research Fellowship under Grant No. DGE-1144082.This work was partially supported by the Universityof Chicago Materials Research Science and EngineeringCenter, which is funded by the National Science Foun-dation under award number DMR-1420709. Use of theCenter for Nanoscale Materials, an Office of Science userfacility, was supported by the U. S. Department of En-ergy, Office of Science, Office of Basic Energy Sciences,under Contract No. DE-AC02-06CH11357. We grate-fully acknowledge support from the David and LucillePackard Foundation.

∗ Electronic address: [email protected]† Electronic address: [email protected]

[1] J. Kelly, R. Barends, A. Fowler, A. Megrant, E. Jeffrey,T. White, D. Sank, J. Mutus, B. Campbell, Y. Chen,et al., Nature 519, 66 (2015).

[2] A. Kandala, A. Mezzacapo, K. Temme, M. Takita,M. Brink, J. M. Chow, and J. M. Gambetta, Nature 549,242 (2017).

[3] N. Ofek, A. Petrenko, R. Heeres, P. Reinhold, Z. Leghtas,B. Vlastakis, Y. Liu, L. Frunzio, S. Girvin, L. Jiang,et al., Nature 536, 441 (2016).

[4] A. Wallraff, D. I. Schuster, A. Blais, L. Frunzio, R.-S.Huang, J. Majer, S. Kumar, S. M. Girvin, and R. J.Schoelkopf, Nature 431, 162 (2004).

[5] J. Koch, T. M. Yu, J. Gambetta, A. A. Houck, D. I.Schuster, J. Majer, A. Blais, M. H. Devoret, S. M. Girvin,and R. J. Schoelkopf, Physical Review A 76, 042319(2007).

[6] H. Paik, I. Schuster, D., L. S. Bishop, G. Kirchmair,G. Catelani, A. Sears, B. Johnson, M. Reagor, L. Frunzio,L. Glazman, et al., Physical Review Letters 107, 240501(2011).

[7] A. G. Fowler, M. Mariantoni, J. M. Martinis, and A. N.Cleland, Physical Review A 86, 032324 (2012).

[8] M. H. Devoret and R. J. Schoelkopf, Science 339, 1169(2013).

[9] J. M. Martinis, M. H. Devoret, and J. Clarke, PhysicalReview B 35, 4682 (1987).

[10] M. Reed, B. Johnson, A. Houck, L. DiCarlo, J. Chow,D. Schuster, L. Frunzio, and R. Schoelkopf, AppliedPhysics Letters 96, 203110 (2010).

[11] J. M. Martinis and A. Megrant, arXiv preprintarXiv:1410.5793 (2014).

[12] M. Takita, A. D. Corcoles, E. Magesan, B. Abdo,M. Brink, A. Cross, J. M. Chow, and J. M. Gambetta,Physical Review Letters 117, 210505 (2016).

[13] R. Barends, J. Kelly, A. Megrant, A. Veitia, D. Sank,E. Jeffrey, T. C. White, J. Mutus, A. G. Fowler,B. Campbell, et al., Nature 508, 500 (2014).

[14] R. Versluis, S. Poletto, N. Khammassi, N. Haider,D. Michalak, A. Bruno, K. Bertels, and L. DiCarlo, arXivpreprint arXiv:1612.08208 (2016).

[15] J. Mooij, T. Orlando, L. Levitov, L. Tian, C. H. Van derWal, and S. Lloyd, Science 285, 1036 (1999).

[16] I. Chiorescu, P. Bertet, K. Semba, Y. Nakamura, C. Har-mans, and J. Mooij, Nature 431, 159 (2004).

[17] Z. Zhou, S.-I. Chu, and S. Han, Physical Review B 70,094513 (2004).

[18] V. E. Manucharyan, J. Koch, L. I. Glazman, and M. H.Devoret, Science 326, 113 (2009).

[19] I.-M. Pop, B. Doucot, L. Ioffe, I. Protopopov, F. Lecocq,I. Matei, O. Buisson, and W. Guichard, Physical ReviewB 85, 094503 (2012).

[20] T. Nicholson, S. Campbell, R. Hutson, G. Marti,B. Bloom, R. McNally, W. Zhang, M. Barrett,M. Safronova, G. Strouse, et al., Nature communications6 (2015).

[21] C. Monroe, D. M. Meekhof, B. E. King, W. M. Itano,and D. J. Wineland, Physical Review Letters 75, 4714(1995).

[22] V. F. Weisskopf and E. P. Wigner, Z. Phys. 63, 54 (1930).[23] A. Houck, J. Schreier, B. Johnson, J. Chow, J. Koch,

J. Gambetta, D. Schuster, L. Frunzio, M. Devoret,S. Girvin, et al., Physical Review Letters 101, 080502(2008).

[24] S. Novikov, T. Sweeney, J. Robinson, S. Premaratne,B. Suri, F. Wellstood, and B. Palmer, Nature Physics12, 75 (2016).

[25] A. A. Abdumalikov Jr, O. Astafiev, A. M. Zagoskin,Y. A. Pashkin, Y. Nakamura, and J. S. Tsai, PhysicalReview Letters 104, 193601 (2010).

[26] F. Lecocq, I. M. Pop, Z. Peng, I. Matei, T. Crozes,T. Fournier, C. Naud, W. Guichard, and O. Buisson,Nanotechnology 22, 315302 (2011).

[27] G. Zhu and J. Koch, Physical Review B 87, 144518(2013).

[28] Y.-H. Lin, L. B. Nguyen, N. Grabon, J. S.Miguel, N. Pankratova, and V. E. Manucharyan,arXiv:1705.07873 (2017).

[29] G. Zhu, D. G. Ferguson, V. E. Manucharyan, andJ. Koch, Physical Review B 87, 024510 (2013).

[30] W. Neuhauser, M. Hohenstatt, P. E. Toschek, and

Page 6: Realization of a system with metastable states of a ... · based on variants of the transmon qubit [5, 12{14]. Transmons have large dipole matrix elements, simple se-lection rules,

20 d

B

20 d

B

Still (700 mK) MC (10-30 mK)

4 K 50 K

300 K

YIG Filter

Micro Lambda

HEMT

Miteq

Minicircuits amplifiers ZX60-83LN-S+

N5242A PNA-X Network analyzer

YOKOGAWA GS200A

Charge + Readout drive

DC flux

Fluxonium Sample

Ecco

sorb

Fi

lter

RF

DC

Min

icirc

uits

S

HP

- 48

K &

L Fi

lter

LP-

6 G

Hz

Ecco

sorb

Fi

lter

Min

icirc

uits

VL

F-18

00M

inic

ircui

ts

SLP

-1.9

+

Flux in Out

Bias teeMinicircuits ZFBT -4R2GW+

50 Ω

50 Ω

DC Block

Circ

ulat

ors

30 d

B10

dB

20 d

B

µ metal shield

Output

+30 dB

+25 dB

+16 dB+16 dB

Charge in

RF flux

1

2

3

FIG. S1: Wiring of microwave and DC connections to thedevice.

H. Dehmelt, Physical Review A 22, 1137 (1980).[31] N. Schlosser, G. Reymond, I. Protsenko, and P. Grangier,

Nature 411, 1024 (2001).[32] C. Cohen-Tannoudji and A. Kastler, Progress in optics

5, 1 (1966).[33] E. Arimondo, Progress in Optics 35, 257 (1996).[34] Y. Lu, S. Chakram, N. Leung, N. Earnest, R. K. Naik,

Z. Huang, P. Groszkowski, E. Kapit, J. Koch, and D. I.Schuster, Physical Review Letters 119, 150502 (2017).

[35] W. R. Kelly, Z. Dutton, J. Schlafer, B. Mookerji, T. A.Ohki, J. S. Kline, and D. P. Pappas, Physical ReviewLetters 104, 163601 (2010).

[36] B. B. Zhou, A. Baksic, H. Ribeiro, C. G. Yale, F. J. Here-mans, P. C. Jerger, A. Auer, G. Burkard, A. A. Clerk,and D. D. Awschalom, Nature Physics 13, 330 (2017).

[37] J. M. Dempster, B. Fu, D. G. Ferguson, D. I. Schuster,and J. Koch, Physical Review B 90, 094518 (2014).

[38] M. T. Bell, J. Paramanandam, L. B. Ioffe, and M. E.Gershenson, Physical Review Letters 112, 167001 (2014).

Device design and fabrication

The device (shown in Fig. 1 in the main text) wasfabricated on a 430 µm thick C-plane sapphire substrate.The base layer of the device, which includes the majorityof the circuit (excluding the Josephson junctions of thetransmon), consists of 150 nm of niobium deposited viaelectron-beam evaporation at 0.9 nm/s, with features de-fined via optical lithography and reactive ion etch (RIE)

at wafer-scale. The wafer was then diced into 7x7 mmchips. The junction mask was defined via electron-beamlithography with a bi-layer resist (MMA-PMMA) in thebridgeless junction pattern [26], with overlap pads fordirect galvanic contact to the optically defined capaci-tors. Before deposition, the overlap regions on the pre-deposited capacitors were milled in-situ with an argonion mill to remove the native oxide. The junctions werethen deposited with a three step electron-beam evapora-tion and oxidation process. First, an initial 30 nm layerof aluminum was deposited at 1 nm/s at an angle of -29 relative to the normal of the substrate. Next, thejunctions were exposed to 50 mBar of high-purity O2 for45 minutes for the first layer to grow a native oxide. Fi-nally, a second 60 nm layer of aluminum was depositedat 1 nm/s at 29 relative to the normal of the substrate.Another oxidation step at 3mbar for 5minutes was doneafter to put a clean oxide layer atop the aluminum. Afterevaporation, the remaining resist was removed via liftoffin N-Methyl-2-pyrrolidone (NMP) at 80C for 6 hours,leaving only the junctions directly connected to the baselayer, as seen in the inset of Figure 1 of the main text.After both the evaporation and liftoff, the device was ex-posed to an ion-producing fan for 15 minutes, in order toavoid electrostatic discharge of the junctions.

The device is mounted and wirebonded to a multilayercopper PCB microwave-launcher board. Additional wire-bonds connect separated portions of the ground plane toeliminate spurious differential modes. The device chiprests in a pocketed OFHC copper fixture that pressesthe chip against the launcher board. Notably, the fixturecontains an additional air pocket below the chip to alter3D cavity modes resulting from the chip and enclosure,shifting their resonance frequencies well above the rele-vant band by reducing the effective dielectric constantof the cavity volume. The filtering, amplifier chain, andwiring diagram of microwave and DC and microwave con-nections to the device are as in Fig. S1.

Equivalent Circuit

The actual circuit of the heavy fluxonium device isshown in Fig. S2(a). In this appendix, we exactly treatthe realistic four-node circuit (assuming that we can re-place the Josephson chain as a pure inductor), and showthat it reduces to an effective three-node circuit shownin the main text and again here [Fig S2(b)]. This circuitis described by the Hamiltonian Hf (Eq. 2 of the maintext), with effective three-node capacitances defined interms of their four-node counterparts.

The phase variables for each of the circuits are shownat their respective nodes (small black circles). With onenode as ground, the four-node ciruit has three phase vari-ables, while the three-node has just two. The key point inthis reduction is that the potential energy in both cases

Page 7: Realization of a system with metastable states of a ... · based on variants of the transmon qubit [5, 12{14]. Transmons have large dipole matrix elements, simple se-lection rules,

7

CC

CR LRC1

C2

EJ ···Φext

Cgi

CjAi

LjAi

ϕR

ϕ1

ϕ1 − ϕQ

a

CC

CR LR CQ EJ Φext LjA

ϕR ϕQ

b

FIG. S2: (a) The complete four node circuit model of theheavy fluxonium device shown in Fig.1. The inductive chainis made up of a series of large-area Josephson junctions withinductance LjAi and capacitance CjAi , with each having aparasitic capacitance to ground, Cgi . The idealized behav-ior of the inductive chain requires that each individual chainjunction have a sufficiently large area to reduce both phaseslips in the chain and the uncontrolled offset charges on theislands. Furthermore, the chain length must be short enoughto avoid having the total stray capacitance

∑i Cgishunting

the chain inductance, but while also ensuring the total chaininductance, LjA, is substantially larger than the individualjunction’s inductance, LJ . Requirements further detailed in[18]. This circuit is equivalent to the three node circuit shown

in (b), whose effective capacitances, C, are all defined in termsof those in (a), see Eqn. S6, S7, and S8.

depends on only two variables, and we choose the label-ing of nodes in Fig. S2(a) to make the two potentialenergies equivalent. Thus, we need only show that thekinetic energy of S2(a) reduces in form to that of S2(b).The kinetic energy T is the capacitive energy, since thevoltage at each node is Vj = hϕj/(2e). For the four-nodecircuit, then,

T = 12

(h2e

)2[CRϕ

2R + CC(ϕ1 − ϕR)2

+C1ϕ21 + C2(ϕ1 − ϕQ)2] (S1)

The potential energy of the system,

V =1

2ELRϕ

2R +

1

2ELjAϕ

2Q + EJ(1− cosϕQ), (S2)

is independent of ϕ1, so that the Euler-Lagrange equationof motion, d

dt∂L∂ϕ1

= ∂L∂ϕ1

= 0, where L = T − V gives

∂L/∂ϕ1 = h2e [(C1 + C2 + CC)]ϕ1

−CC ϕQ − C2ϕQ] (S3)

is a constant of the motion, which we may set to zero.This allows us to eliminate ϕ1 from Eqn.S1, which, aftersome algebra, results in the desired form,

T =1

2

(h

2e

)2 [CRϕ

2R + CQϕ

2Q − 2CC ϕRϕQ

], (S4)

with three effective capacitances [corresponding to Fig.S2(b)] defined as follows: With the preliminary defini-tion,

CT = C1 = C2 + CC , (S5)

the effective capacitances are

CR = CR + CC(C1 + C2)/CT , (S6)

CQ = C2(C1 + CC)/CT , (S7)

CC = C2CC/CT . (S8)

Physically, CQ is the capacitance between the two

nodes of the qubit, and CR is the capacitance betweenthe resonator and ground nodes. We note that CC van-ishes as C2 → 0 , and approaches CC as C2 →∞. In oursystem, C2 ≈ C1 + CC , so CC ≈ CC/2.

Direct Drive

As mentioned in the main text, we cannot achieve co-herent operations on the metastable states by drivingdirectly on the fluxon transition. The result of such adrive at Φ = 0.078Φ0 is shown in Fig. S3. The width ofthe fluxon transition is found to saturate at high drivepowers, in Fig.S3a. Directly driving the fluxon at thelocation of largest width of the fluxon line results in thepopulation saturating to a mixed state in ∼ 100 µs.

Origin of Plasmon Dispersion

The simulations of the spectra shown in Fig. 3 of themain text account accurately for the observed plasmondispersion, and also for the smaller dispersion observed inthe resonator line. In this and the next section we providea simplified analysis to interpret and trace the origins ofthese and related phenomena. This section is devotedto estimating the change in the bare qubit levels arisingfrom Φext induced distortions of the potential well shape,with resonator interaction effects put off to the followingsection.

The strategy is to expand the potential about itsshifted minimum to characterize the distortion, and thenapply perturbation theory to obtain the spectra. The re-sulting dispersion is quadratic in Φext, with small quarticcorrections.

When Φext is applied, the central well rides up in-side the inductive parabola, making contact at the phase

Page 8: Realization of a system with metastable states of a ... · based on variants of the transmon qubit [5, 12{14]. Transmons have large dipole matrix elements, simple se-lection rules,

8

a

0 50 100 150 200 250 300Rabi Drive Time (µs)

0.0

0.1

0.2

0.3

0.4

0.5

0.6

Exci

ted p

opula

tion

80 70 60 50Power at Resonator (dBm)

4.19

4.20

4.21

4.22

4.23Pro

be F

requency

(G

Hz)

0.54

0.56

0.58

0.60

0.62

0.64

0.66

0.68

0.70

Reso

nato

r Tra

nsm

issi

on

b

FIG. S3: a) Power sweep of the probe tone while directlydriving the fluxon transition at Φ = 0.078Φ0. We draw at-tention to two features: 1) The width of fluxon transition(4.21 GHz) saturates at high powers and 2) The appearanceof other photon-assisted transitions immediately above andbelow the fluxon transition, pointing towards a source of lim-itation for the achievable Rabi rates. b) At a power of -52dB, where the fluxon transition has the largest width, a Rabidrive results in no coherent oscillations and a 50/50 mixturein ∼100 µs.

value ϕ = 2πΦext/Φ0, where the cosine term vanishes.However, clearly, this point is not the potential minimum;the actual minimum is Stark-shifted by the inductive cur-rent, ELϕ, so that ϕmin = 2π(Φext/Φ0)(1 − EL/EJ), tolowest order in EL/EJ . We define the displacement fromthis minimum as ϕ = ϕ−ϕmin, and rewrite the potentialenergy as a function of ϕ:

V = EJ[1− cos

(ϕ− 2π

ELEJ

Φext

Φ0

)]+

1

2EL(ϕ+ ϕmin)2.

(S9)Expanding and ignoring constant terms,

V = −EJ(

cos ϕ cos 2πEL

EJ

Φext

Φ0+ sin ϕ sin 2πEL

EJ

Φext

Φ0

)+ 1

2ELϕ2 + ELϕminϕ. (S10)

Making small-angle expansions and regrouping,

V = −EJ[1− 2

(πEL

EJ

)2(Φext

Φ0

)2]cos ϕ

+ 12ELϕ

2 + 2πEL(

Φext

Φ0

)(ϕ− sin ϕ). (S11)

This equation shows that the symmetric part of the dis-placed well is identical to the original well, except for areduction in the effective EJ by the small relative amountshown, while the odd part appears and cancels the linearpart of the inductive term, leaving cubic and higher oddpowers.

Both symmetric and antisymmetric parts of V con-tribute quadratic terms (∼ Φ2

ext) to the plasmon disper-sion; the even part at first order in perturbation theory,the odd part at second order. To estimate the former,note that in the harmonic approximation, the plasmon

energy EP increases as√EJ , so that the fractional re-

duction is half that of EJ , and hence

∆EP (symm) = −(πEL

EJ

)2EP(Φext/Φ0

)2≈ −45MHz

(Φext/Φ0

)2. (S12)

Regarding the antisymmetric contribution, applyingsecond-order perturbation theory to the leading cubicterm in ϕ gives:

∆EP (anti) =(π

3EL)2(Φext

Φ0

)2

·(|〈f0|ϕ3|e0〉|2

E′P− 2|〈g0|ϕ3|e0〉|2

EP

)The first term in brackets is the shift in energy of the|e0〉 state due to virtual occupation of |f0〉, with en-ergy denominator E′P = Ef0 − Ee0 = 4.39 GHz. Vir-tual occupation of |g0〉 accounts for half the second term.The other half is (minus) the shift in energy of |g0〉 dueto virtual occupation of |e0〉, with energy denominatorEP = 5.072 GHz. Evaluating the matrix elements usingϕ = 0.60(a+ a†), we find

∆EP (anti) ≈ 39MHz(Φext/Φ0

)2. (S13)

The above two contributions together predict a totalquadratic dispersion ∆EP = 84 MHz (Φext/Φ0)2. Quan-titative agreement with the observed dispersion is decep-tive - interaction with the resonator reduces the predic-tion by about 12 %. While quantitative agreement is notexpected or sought from such a calculation, it suggeststhat a predominantly quadratic plasmon dispersion arisesfrom Φext induced distortions in the well, with compara-ble contributions arising from reduced effective EJ in thesymmetric part, and induced antisymmetry dominatedby cubic anharmonicity. Quartic corrections arise at thenext contributing order in perturbation theory, secondorder for the symmetric part, and fourth order for the an-tisymmetric part. We have found these to be no greaterthan 1% of the quadratic contributions. [The small pa-rameter governing successively higher contributing or-ders in perturbation theory in both cases is (EL/EP )2.]Larger (but still small) quartic corrections will arise fromthe interaction with the resonator, as shown in the fol-lowing.

Resonator/Qubit hybridization

The most direct indication of the strong interactionbetween the principle plasmon and the resonator is pro-vided by the power sweep: At low power levels one seesboth features, separated by 155 MHz, whereas at higherpowers, one sees only the resonator line, shifted upwardby 33 MHz. Assuming that the high-power value is the

Page 9: Realization of a system with metastable states of a ... · based on variants of the transmon qubit [5, 12{14]. Transmons have large dipole matrix elements, simple se-lection rules,

9

120 110 100 90 80 70Drive power at Readout Resonator (dBm)

4.90

4.95

5.00

5.05

5.10Pro

be F

requency

(G

Hz)

ωr=4.95GHz

ωq=5.067GHz

ωr ′=4.915GHz

0.15

0.30

0.45

0.60

0.75

0.90

1.05

FIG. S4: Single Tone Spectroscopy power sweep at the in-put of the readout resonator. At high powers (> −90 dB),the heavy fluxonium qubit is saturated, and the bare reso-nance, ωr of the readout resonator is seen. At lower powers(< −100 dB) the qubit plasmon ωq and the dressed resonator,ωr′ are seen. The bare resonator transmission amplitude isused to normalize the spectra in all the figures of the maintext.

upcoupled (bare) resonator line, one can infer the inter-action strength g from these three numbers alone. Thisis found to be within the range of experimental uncer-tainty of the value g = 76 MHz (Fig. S4) determined bya global fit to all the data.

Here we introduce a simple model that describes theplasmon and resonator lines over the entire flux range,excluding the immediate neighborhood of the pointsΦext = 0 and Φ0/2, where more transitions are involved.Consider the coherent superposition of the two lowest ex-cited states of the noninteracting system, one involvingthe resonator and the other involving the qubit:

|ψ〉 = α|1R, g0〉+ β|0R, e0〉. (S14)

The effective Hamiltonian acting on the subspace of S14is

H =

(εR mm∗ εQ

)(S15)

where m is the interaction matrix element specific to theplasmon excitation, m = g〈e0|n|g0〉 = 0.062i. Diago-nalizing H, we find the two lowest excited eigenstates ofthe system. Near Φext = 0, where the bare detuning isεQ − εR ≈ 89 MHz, we find the resonator-like eigenstatedefined by (α/β)R = 1.94i, and the qubit-like eigenstate(α/β)Q = −i/1.94. In the former, there is a probabilityof |βR|2 = 0.21 that the qubit will be found in the (|e0〉state, and conversely, in the latter, the same probabilitythat the resonator will be found in its |1〉 state.

Since both resonance lines are read from the resonator,their relative intensities are given by the relative proba-bilities of finding the resonator in its |1〉 state,

IR/IQ = |αR/αQ|2 = 3.8, (S16)

and a rough comparison of resonance peak heights timeslinewidths is consistent with this estimate. As Φext is in-creased and the detuning is reduced, the eigenstates ex-hibit further hybridization. At Φext = 1/2, for example,the predicted intensity ratio is reduced to IR/IQ = 2.6.

A more obvious consequence of hybridization, quan-titatively, is the curvature of the resonator line, whichappears to be due entirely to level repulsion from thequbit. To demonstrate, we write the bare plasmon en-ergy as εQ = 5.039 GHz −a(Φext/Φ0)2, where a purelyquadratic dependence on Φext is assumed, following ar-guments of the previous section. Assuming that εR is aconstant, we may write down the eigenvalues of H andevaluate the leading quadratic dependence on Φext withsmall quartic corrections. The Φext dependent parts are

∆ER = −0.21a [Φext/Φ0]2 − 0.3a [Φext/Φ0]4, (S17)

∆EQ = −0.79a [Φext/Φ0]2 + 0.3a [Φext/Φ0]4.(S18)

While the estimate of the previous section may not be ac-curate quantitatively, the ratio of coefficients found hereshould be (of both the quadratic and quartic terms). It isinteresting to note that the ratio of quadratic coefficientsis 3.8; compare that of S16.

The calculation above supports the interpretation thatthe observed dispersion of the resonator line is a directconsequence of hybridization with the plasmon, whosedispersion is driven in turn by Φext induced distortions inthe potential well. And, like the plasmon, the resonatorline has two distinct branches visible near the half fluxquantum point, one generated by hybridization with the|g0〉 → |e0〉 branch, the other with the |g1〉 → |e1〉 branch.

Photon/Plasmon Assisted Transitions

The two-photon features of Fig. 4a of the main textare a result of transitions to the two-excitation subspace(|e1,−1〉 , |f0〉) starting with the combined fluxonium-resonator in |g0〉 ⊗ |0〉 state. The avoided crossing asso-ciated with this two-excitation space are also observedin the resonator-photon and plasmon assisted transi-tions shown in Fig. S5a, b. These transitions start withthe coupled resonator-fluxonium system in |g0, 1res〉 and|e0, 0res〉, respectively. The final states of the transitionare the same as those in Fig. 4a of the main text, butare here reached using a single photon, with the sec-ond photon acquired from the readout resonator and theplasmon, respectively. Due to the low Q of the readoutresonator (Q ∼ 500), and the resulting Purcell limitedlifetimes of the primary plasmon features, these transi-tions cannot be used for coherent transitions between thestates. Increasing the Q of the readout resonator, and it’sdetuning from the primary plasmon will allow the use ofthese transitions for coherent operations between fluxoidstates, as will be explored in future devices.

Page 10: Realization of a system with metastable states of a ... · based on variants of the transmon qubit [5, 12{14]. Transmons have large dipole matrix elements, simple se-lection rules,

10

0.04 0.02 0.00 0.02 0.04Flux (Φext/Φ0)

4.4

4.5

4.6

4.7

4.8Pro

be F

requency

(G

Hz)

|g0, 1r⟩→ |g1, 1r

⟩|g0, 1r

⟩→ |e1, 0r⟩

|g0, 1r⟩→ |f0, 0r

⟩|g0, 1r

⟩→ |g−1, 1r⟩

|g0, 1r⟩→ |e−1, 0r

⟩ 0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

Reso

nato

r Tra

nsm

issi

on

0.04 0.02 0.00 0.02 0.04Flux (Φext/Φ0)

4.0

4.1

4.2

4.3

4.4

4.5

4.6

4.7

Pro

be F

requency

(G

Hz)

|e0, 0r⟩→ |g1, 1r

⟩|e0, 0r

⟩→ |e1, 0r⟩

|e0, 0r⟩→ |f0, 0r

⟩|e0, 0r

⟩→ |g−1, 1r⟩

|e0, 0r⟩→ |e−1, 0r

⟩ 0.05

0.10

0.15

0.20

0.25

0.30

0.35

0.40

0.45

Pla

smon T

ransm

issi

on

a

b

FIG. S5: Two-tone spectroscopy as a function of flux (nearzero flux) showing (a) resonator-photon and (b) plasmon as-sisted transitions. In (a) the readout tone is placed at the res-onator frequency and the transitions start with the fluxonium-resonator system in |g0〉 ⊗ |1〉. In (b), the ‘readout’ tone isplaced at the plasmon frequency and the transitions start withthe fluxonium-resonator system in |e0〉 ⊗ |0〉. In each case,the transmission of a readout tone is monitored, while thefrequency of a second drive tone is swept (y-axis). Both setsof transitions are absent when the drive and readout tonesare pulsed and staggered, as a result of the short lifetimes ofboth the resonator and the plasmon. (State labeling for thetransitions is valid for Φext > 0.)

The resonator assisted transitions are useful for photon‘latching’ in which a single photon from the resonator canbe transferred to an excited metastable state, and subse-quently detected by cycling on the excited plasmon tran-sition. This might be potentially useful in dark photondetection schemes.

Qubit Inversion in Single Tone Spectroscopy

During the measurement of the device, there was aweek long period in which the device was in a frustratedexcited state as shown Fig. S6. We take particular noteof the switching from the |g1〉 → |e1〉 plasmon branch tothe |g0〉 → |e0〉 branch. While we were unable to repro-duce this particular spectrum, with inverted single tone

0.75 0.50 0.25 0.00 0.25 0.50Flux ( ext/ 0)

4.90

4.95

5.00

5.05

5.10

Prob

e Fr

eque

ncy

(GHz

)

|g0 |e0|g1 |e1|g0, 0r|g1, 0r|g1 |e0|g1, 0r |g0, 1r|g0 |e1|g0, 0r |g1, 1r|g0 |g1

FIG. S6: Single tone spectroscopy demonstrating qubit stateinversion for varying time durations, and appearance of newavoided crossings. Reasons for this inversion and avoidedcrossings are unknown and will be part of future investiga-tions.

spectrum and the onset of new avoided crossings, theswitching from one well branch to another seems to havebeen a common feature of this device. This switchingcan also been seen at Φext ≈ 0.6Φ0 in Fig. 2 in the maintext. A more careful analysis to address the exact natureof the metastable qubit inversion is warranted. Fig. S6,however, is a further demonstration of the large disper-sive shifts of the readout resonator, and demonstratesthe ability to readout the fluxoid state over the entireflux range.

Steady-State Simulations for Single-ToneSpectroscopy

We perform numerical simulations of single-tone spec-troscopy using the Lindblad master equation:

dρ(t)

dt=− i [H(t), ρ(t)] (S19)

+∑ω

γω(T )

(Aωρ(t)A†ω −

1

2A†ωAω, ρ(t)

).

Here, γω is the temperature-dependent decoherence ratefor a transition with frequency ω, Aω is the correspond-ing jump operator, ρ the density matrix, and H(t) theHamiltonian describing the coupled fluxonium–resonatorsystem along with a coherent drive:

H(t) =HFluxonium +HResonator (S20)

+∑n,m

gnm |n〉 〈m|(a+ a†

)+ ζ

(aeiωdt + h.c.

)The drive strength is small, i.e., ζ ωd, and the

rotating wave approximation between the driving fieldand resonator is used to drop rapidly oscillating terms.The qubit coupling strengths, gnm, are obtained fromthe fluxonium charge matrix elements. The dissipa-tion operators, Aω, are eigenoperators of the qubit-resonator system (not including the driving field). The

Page 11: Realization of a system with metastable states of a ... · based on variants of the transmon qubit [5, 12{14]. Transmons have large dipole matrix elements, simple se-lection rules,

11

0.6 0.4 0.2 0.0 0.2 0.4 0.6

Flux (Φ0)

4.90

4.95

5.00

5.05Fr

equency

(G

Hz)

0.004

0.008

0.012

0.016

0.020

0.024

0.028

|<a>

|

FIG. S7: Steady-state single tone simulation of the coupledfluxonium-resonator system at a finite temperature of 30mK.When a non-zero temperature is included in the steady-state simulation, a non-zero photon expectation value in the|g1〉 → |e1〉 branch is observed, agreeing with the single-tonemeasurements shown in Fig. 2.

.

temperature-dependent decoherence rates, γω(T ), aregiven by a resonator rate, κf(T, ω), and fluxonium rates,Γ| 〈n|N |m〉 |2f(T, ω). 〈n|N |m〉 are fluxonium chargematrix elements, f(T, ω > 0) = eβω/(eβω − 1), andf(T, ω < 0) = 1/(eβω − 1) . Γ and κ are rate con-stants that are fit from the experimental transmissiondata near Φext = 0. These rate constants are found tobe Γ = 0.0005 GHz and κ = 0.04 GHz.

The signal transmission amplitude is obtained by cal-culating the average of the resonator annihilation oper-ator, |tr(aρ)|, using the solution to S19 in the long-time

limit:dρ′ss(t)dt = 0. We perform the calculation in a rotat-

ing frame in which time dependence of S20 is effectivelyeliminated. This removes the need for time averaging,considerably speeds up computation time, and therebyallows us to investigate behavior on time scales of mil-liseconds and greater, such as the the thermally assisted|g1〉 → |e1〉 transition depicted in figure S7.