Sei sulla pagina 1di 14

Demonstration of nonstoquastic Hamiltonian in coupled superconducting flux qubits

I. Ozfidan,1, ∗ C. Deng,1, ∗ A. Y. Smirnov,1 T. Lanting,1 R. Harris,1 L. Swenson,1 J. Whittaker,1 F. Altomare,1


M. Babcock,1 C. Baron,1 A.J. Berkley,1 K. Boothby,1 H. Christiani,1 P. Bunyk,1 C. Enderud,1 B. Evert,1
M. Hager,1 A. Hajda,1 J. Hilton,1 S. Huang,1 E. Hoskinson,1 M.W. Johnson,1 K. Jooya,1 E. Ladizinsky,1
N. Ladizinsky,1 R. Li,1 A. MacDonald,1 D. Marsden,1 G. Marsden,1 T. Medina,1 R. Molavi,1 R. Neufeld,1
M. Nissen,1 M. Norouzpour,1 T. Oh,1 I. Pavlov,1 I. Perminov,1 G. Poulin-Lamarre,1 M. Reis,1 T. Prescott,1
C. Rich,1 Y. Sato,1 G. Sterling,1 N. Tsai,1 M. Volkmann,1 W. Wilkinson,1 J. Yao,1 and M. H. Amin†1, 2
1
D-Wave Systems Inc., 3033 Beta Avenue, Burnaby BC, Canada, V5G 4M9
2
Department of Physics, Simon Fraser University, Burnaby BC, Canada, V5A 1S6
(Dated: March 15, 2019)
arXiv:1903.06139v1 [quant-ph] 14 Mar 2019

Quantum annealing (QA) is a heuristic algo-


rithm for finding low-energy configurations of a
system, with applications in optimization [1–3],
machine learning [4, 5], and quantum simulation
[6, 7]. Up to now, all implementations of QA
have been limited to qubits coupled via a single
degree of freedom. This gives rise to a stoquastic
Hamiltonian that has no sign problem in quantum
Monte Carlo (QMC) simulations [5, 8, 9]. In this
paper, we report implementation and measure-
ments of two superconducting flux qubits cou-
pled via two canonically conjugate degrees of
freedom—charge and flux—to achieve a nonsto-
quastic Hamiltonian. Such coupling can enhance
performance of QA processors [11, 12], extend the
range of quantum simulations [13], and provide
a path towards annealing-based universal quan-
tum computation [14–16]. We perform microwave
spectroscopy to extract circuit parameters and
show that the charge coupling manifests itself as a
σ y σ y interaction in the computational basis. We
observe destructive interference in quantum co- FIG. 1: Schematic and potential energy of coupled rf-
herent oscillations between the computational ba- SQUIDs. a. Schematic circuit of two rf-SQUIDs, coupled
sis states of the two-qubit system. Finally, we inductively via a tunable coupler and capacitively via a fixed
capacitor. b. Effective potential energy of the circuit shown
show that the extracted Hamiltonian is nonsto-
in a. Arrows show tunneling paths between |↑↑i and |↓↓i
quastic over a wide range of parameters. states within the two-state approximation for each rf-SQUID.
As early generations of quantum annealing processors The solid arrows indicate tunneling due to two consecutive
mature, there is growing interest in novel extensions to single qubit flips facilitated by σix . The dotted arrow repre-
this technology. Currently available large-scale quantum sents direct two-qubit cotunneling due to σ1x σ2x and σ1y σ2y . The
annealers are made from a network of radio-frequency su- tunneling amplitudes may have opposite signs, thus leading
perconducting quantum interference devices (rf-SQUIDs) to destructive interference.
[1, 2, 19]. Interaction between pairs of devices is re-
alized through tunable magnetic coupling of their flux
degrees of freedom. The low-energy dynamics of indi- the transverse-field Ising Hamiltonian:
vidual rf-SQUIDs are effectively captured with their two 1X X X
lowest-energy eigenstates, allowing one to approximate H=− ∆i σix + hi σiz + Jij σiz σjz , (1)
2 i
rf-SQUIDs as qubits, described by Pauli matrices σ x,y,z . i i<j
The computational basis states |↑i and |↓i (eigenfunc-
tions of σ z ) correspond to directions of persistent current where ∆i and hi are tunneling amplitude and energy bias
in the body of the rf-SQUID. This network implements of qubit i, respectively, and Jij is the magnetic coupling
strength between qubits i and j. Quantum annealing is
performed by adjusting ∆i  hi , Jij at the beginning of
the annealing process and gradually evolving until ∆i 
† corresponding author, e-mail: amin@dwavesys.com hi , Jij at the end.
∗ These authors contributed equally to this work A Hamiltonian is stoquastic if there exists a local ba-
2

FIG. 2: Microwave spectroscopy of the coupled two-qubit system at M12 = 0 and Φxq,1 = Φxq,2 = 0. a. Pulse sequences
for two-qubit microwave spectroscopy. At Φcjj,monostable the tunneling barrier is at its lowest, resulting in a monostable potential.
Φxcjj,latched is the opposite, where the tunneling barrier is high and the tunneling amplitude is negligible. The effective single-
qubit spectroscopy has the same pulse sequence except for Φxcjj,2 = 0.5Φ0 . b-d Two-qubit spectroscopy versus Φxcjj,1 , which
controls the qubit 1 barrier height. Qubit 2 is kept at Φxcjj,2 corresponding to effective single-qubit tunneling amplitudes: b
∆2 = 1.5 GHz, c 2.0 GHz, and d 3.0 GHz. Energies are measured relative to the ground state. The solid lines are obtained by
fitting the rf-SQUID model, Eq. (2), to the experimental data. Dashed lines represent the excited state energies of uncoupled
qubits (M12 = 0, C12 = 0) given the other fitting parameters.

sis (formed by product states of individual qubits) in and Φxcjj,i are external fluxes, and Φ0 = πh̄/e is the
which all off-diagonal elements are nonpositive [5]. This flux quantum. Each rf-SQUID is characterized by its
eliminates the sign problem for algorithms like QMC. If capacitance Ci , inductance Li , and tunable Josephson
there is no such basis, the Hamiltonian is nonstoquas- energy EJi (Φxcjj,i ) ≈ (Φ0 /2π)Ici cos(πΦxcjj,i /Φ0 ) [1] (for
tic. Therefore, taking into account unitary local trans- a more detailed description, see the supplementary in-
formations, Hamiltonian (1) is stoquastic regardless of formation (SI)). The tunable mutual inductance M12 is
the sign of ∆i . To achieve nonstoquasticity, additional adjusted with the coupler control bias Φxco . The renor-
interactions such as σix σjx or σiy σjy are needed. Such in- malized capacitances are defined as C e1(2) = C1(2) +
teraction terms can be realized by coupling the charge C12 C2(1) /(C2(1) + C12 ).
degrees of freedom of qubits. Nonstoquasticity may also The potential energy of each rf-SQUID can have a
be achieved via nonadiabatic evolution [20]. double-well shape (see the two-qubit potential in Fig. 1b).
To this end, we implemented two rf-SQUIDs, coupled The barrier height of the potential is controlled by Φxcjj,i ,
both inductively through a tunable mutual inductance which tunes the tunneling amplitude ∆i , but also changes
M12 [2] as well as capacitively through a fixed capacitance the persistent current. The potential is monostable when
C12 as shown in Fig. 1a. The Hamiltonian of this system Φxcjj,i = 0.5 Φ0 . The flux bias Φxq,i adjusts the tilt of the
is approximated by potential, setting hi . Both Φxq,i and Φxcjj,i are controlled
" # by high bandwidth coaxial lines, allowing for microwave
2
X Q2i (Φq,i −Φxq,i )2 2πΦ q,i operation and fast quench of the coherent dynamics. At
H = + −EJi (Φxcjj,i ) cos the end of quench, which involves raising the tunneling
i=1 2Cei 2L i Φ0
barrier rapidly, the persistent current is measured via
M12 (Φxco )(Φq,1 −Φxq,1 )(Φq,2 −Φxq,2 ) a shift register coupled to a microwave resonant read-
+
L1 L2 out [21, 22].
C12 Q1 Q2 Before characterizing the circuit parameters, we cali-
+ , (2) brate the tunable magnetic coupler, which provides the
C1 C2 +(C1 +C2 )C12
function M12 (Φxco ) as discussed in Ref. [2]. For the rest of
where Qi and Φq,i are charge and flux variables that sat- the manuscript, we treat the coupler as a simple tunable
isfy the commutation relation [Φq,i , Qj ] = ih̄ δij , Φxq,i mutual inductance, assuming dynamics of the coupler
3

FIG. 3: Two-qubit spectroscopy at nonzero energy bias hi 6= 0. In all panels, qubit 2 is biased away from degeneracy
with Φxq,2 = 0.1 mΦ0 and Φxcjj,2 is set such that the effective single-qubit tunneling is ∆2 = 1.5 GHz. Energy is measured
relative to the ground state. Solid white lines are numerical simulations obtained using the two-qubit Hamiltonian (3) with no
fitting parameters. a. Energy spectrum as a function of Φxcjj,1 at fixed Φxq,1 = 0.1 mΦ0 and M12 = 0. b. Energy spectrum
as a function of Φxq,1 at M12 = 0.55 pH and effective single-qubit ∆1 = 1.5 GHz. The vertical dotted line goes through the
Φxq,1 = 0 to highlight the asymmetry in the spectrum due to fixed Φxq,2 = 0.1 mΦ0 . c. Energy spectrum as a function of
M12 at Φxq,i = 0.1 mΦ0 and ∆i = 1.5 GHz. The dashed lines correspond to numerical simulations at zero energy biases. The
observed avoided crossing is at M12 = 0.55 pH is due to nonzero biases. The vertical dotted line separates ferromagnetic
(FM) and antiferromagnetic (AFM) regions. The asymmetry in the data about this line is due to capacitive coupling. d.
Extracted interaction parameters in Hamiltonian (3) that provide the theoretical (solid) lines in panel c. The Hamiltonian is
nonstoquastic in the white unshaded area according to Ref. [5].

are significantly faster than single- and coupled-qubit dy- any type of coupling, the spectral lines representing the
namics. Next, we measure the persistent current of each first excited states of noninteracting qubits would cross
qubit for a range of Φxcjj in a regime where its tunneling as shown by the dashed lines [24]. The anticrossing is
amplitude is negligible and the other qubit is kept monos- therefore a signature of capacitive coupling (at M12 = 0)
table. We fit these measurements to a classical rf-SQUID and its energy gap is a measure of coupling strength. The
model [1] and obtain Ic,1 = 3.227 µA, L1 = 231.9 pH, extracted anticrossing gaps of 0.77, 1.14, and 1.78 GHz
Ic,2 = 3.157 µA, and L2 = 239.0 pH. at ∆2 = 1.5, 2.0, and 3.0 GHz, respectively, suggest a
To extract the remainder of the circuit parameters, strong capacitive coupling.
we perform microwave spectroscopy on the single- and We now reduce the continuous rf-SQUID model to a
two-qubit systems [3]. Applying a fixed Φxcjj,2 to the two-state (qubit) model, relevant for quantum computa-
second qubit, we sweep the barrier height of the first tion. The flux degree of freedom is described by σiz , with
qubit, controlled by the external flux bias Φxcjj,1 . At ev- σ1z σ2z representing the inductive coupling. The charge

operator Qi = − ih̄ ∂Φ , on the other hand, is related to
ery Φxcjj,1 , a long microwave pulse (1 µs) is applied to the y
i

first qubit to excite the two-qubit system. The energy σi since both are complex in the computational basis.
eigenstates are read out by applying an adiabatic tilt Thus, the electrostatic coupling between the rf-SQUIDs
to both qubits to transform the energy eigenstates into gives rise to a σ1y σ2y term. When reducing the contin-
persistent current states, followed by a quench to freeze uous Hamiltonian, the interaction terms in Eq. (2) mix
the dynamics of both qubits before readout. The pulse the two lowest-energy states with higher-energy states
sequence is shown in Fig. 2a. The excited state popu- of the individual rf-SQUIDs. This may introduce ad-
lation of each qubit serves as a signal for detecting the ditional coupling terms, such as σ1x σ2x . Taking into ac-
energy spectrum of the coupled system. We collect effec- count these higher energy states, we derive an effective
tive single-qubit data by removing the potential barrier two-qubit Hamiltonian, which can be represented as
of the other qubit, making it monostable, and perform ∆ 1 x ∆2 x X
two-qubit spectroscopy for various Φxcjj,2 . In both sets of H=− σ1 − σ2 +h1 σ1z +h2 σ2z + Jαβ σ1α σ2β , (3)
2 2
experiments we keep Φxq,1 = Φxq,2 = 0, M12 = 0 pH, while α,β
the capacitive coupling is always present. Jointly fitting where α, β = {x, y, z} (see SI for details). Note that
the effective single- and two-qubit spectroscopy data to Jαy = Jyα = 0, for α 6= y, as they make the Hamiltonian
the coupled rf-SQUID model, Eq. (2), we obtain the rest complex. In what follows, we show that this Hamiltonian
of the circuit parameters, C1 = 119.5 fF, C2 = 116.4 fF, can explain experimental observations.
and C12 = 132 fF. Figures 3a-c depict two-qubit spectroscopy at nonzero
Figures 2b-d show two-qubit spectroscopy data along energy bias (hi 6= 0). Solid white lines in these panels cor-
with the numerical fit using the rf-SQUID model (2). respond to numerical simulations obtained using Hamil-
One can see a clear ∆2 dependent anticrossing that is in tonian (3) with no fitting parameters. The presence of
good agreement with simulations (solid lines). Without an energy bias is necessary for nonstoquasticity, as there
4

FIG. 4: Two-qubit coherent oscillations as a function of mutual inductance. a. Population of state |↓↑i when the
system is initialized in |↑↓i. b and c. Population of state |↓↓i when the system is initialized in |↑↑i. The final flux-bias on
qubits for panel c are set to 0.1 mΦ0 corresponding to Figs. 3c,d. Panels a and b both have a final flux bias of zero. Panels
d, e, and f are numerical simulations obtained using the reduced 4-level model given in Eq. (3) corresponding to a, b, and c,
respectively.

exists a unitary transformation that can remove positive is almost constant over the whole range of M12 . The two-
off-diagonal elements at hi = 0 (see SI). The experimen- qubit cotunneling mediated by the higher energy states
tal parameters in Fig. 3a are the same as those in Fig. 2b, of the rf-SQUIDs leads to a σ1x σ2x coupling with a coeffi-
except for the flux bias Φxq,i determining hi . This nonzero cient Jxx comparable to Jyy in magnitude. Considering
bias manifests itself in Fig. 3a as an avoided level cross- rotations in x-z plane, both Jxx and Jzz terms can re-
ing between the second and third excited states. Figure duce the nonstoquastic contribution of Jyy . The Hamil-
3b shows the energy spectrum as a function of Φxq,1 while tonian becomes stoquastic if either of them exceeds Jyy
Φxq,2 = 0.1 mΦ0 and M12 = 0.55 pH. The top two energy in magnitude, as highlighted by the shaded area in the
levels cross when Φxq,1 = −Φxq,2 . Energy spectrum as a figure. Applying all possible local unitary transforma-
function of M12 is presented in Fig. 3c. As in Fig. 3a, tions outlined in Ref. [5], we confirm nonstoquasticity in
the avoided level crossing observed at M12 = 0.55 pH is the unshaded region. Both Jxx and Jyy depend on the
a result of nonzero energy bias (hi ≈ 0.15 GHz at this barrier heights. As a result, the width of the nonstoquas-
point). Zero bias simulations are shown by the dashed tic region changes with Φxcjj,i . We should mention that
lines. One can clearly see that the capacitive coupling in- finding a nonlocal transformation to cure sign problem is
troduces an asymmetry between the AFM and FM sides by itself intractable [9].
of the magnetic coupling. Without the coupling capac- Finally, we measure quantum coherent oscillations be-
itor, the energy spectrum is expected to be symmetric tween the computational basis states. Qubits are ini-
about M12 = 0. tialized in a computational basis state by applying a
To demonstrate nonstoquasticity, we extract coeffi- strong flux bias Φxq,i . The coherent oscillations are in-
cients in Hamiltonian (3) corresponding to Fig 3c. Fig- duced by pulsing down the barriers of both qubits si-
ure 3d plots interaction parameters except for Jxz and multaneously. At the same time, the flux bias Φxq,i
Jzx , which are negligibly small. Other parameters are on each qubit is changed from its value at preparation
provided in SI. We see that the electrostatic coupling be- pulse to its final value. Since the computational basis
tween the rf-SQUIDs gives rise to a pronounced Jyy that states are not the eigenstates of the total Hamiltonian,
5

the system undergoes coherent oscillations between these through sequential single qubit flips caused by σ1x and σ2x
states. After some dwell time τ , the qubits are simulta- terms in Hamiltonian (3), as schematically illustrated in
neously quenched by rapidly raising their energy barriers Fig. 1b. Moreover, we see an additional signature of Jyy ,
via Φxcjj,i , and the qubit states are read out. We repeat since without such coupling Fig. 10a would be a reflection
this process for a range of dwell times τ and coupling of Fig. 10b with respect to M12 = 0 (see SI).
strengths M12 .
In conclusion, we have fabricated two superconduct-
In Figs. 10a,b, we show the measured state popula-
ing flux qubits coupled both inductively and capacitively.
tion P↑↓ and P↓↓ for qubits initially prepared in |↓↑i and
To our knowledge, this is the first time that two qubits
|↑↑i configurations, respectively, for hi = 0. The agree-
have been coupled via two conjugate degrees of freedom.
ment between time domain experiments and numerical
Starting from an rf-SQUID model, with experimentally
simulations, shown in Fig. 10d,e, justifies the two-qubit
extracted circuit parameters, we have obtained a reduced
Hamiltonian (3) as a valid description of the circuit in
two-qubit Hamiltonian with σ1x σ2x , σ1y σ2y , and σ1z σ2z inter-
Fig. 1a. Deviation between theory and experiment can
actions. We show that the reduced Hamiltonian can ex-
be attributed to decoherence, which is absent in simula-
plain spectroscopy and coherent oscillation experiments.
tions. Single qubit measurements reveal relaxation and
Considering all local transformations, we prove that the
dephasing times of T1 = 17 ns and T2 = 16 ns, respec-
Hamiltonian is nonstoquastic in a wide range of param-
tively.
eters. Equilibrium statistics of such nonstoquastic quan-
Features of the energy spectrum as a function of
tum annealers cannot be simulated by QMC due to the
M12 (Fig. 3c) are reflected in coherent oscillations
sign problem. Implementation of conjugate couplings be-
(Figs. 10c,f). The initial configuration, |↑↑i, has signifi-
tween qubits is an important step towards the develop-
cant overlap with the second and third excited states in
ment of universal quantum annealers [14–16]. Our imple-
the AFM region (M12 > 0). The slow oscillation fre-
mentation is based on current, scalable superconducting
quency on the right half of figures is therefore related
fabrication technology that is ready to be expanded to a
to the gap between these two states. The minimum gap
large number of qubits.
at M12 ≈ 0.55 pH in Fig. 3c corresponds to the maxi-
mum slowdown at the same point in Figs. 10c,f. When We are grateful to D. Lidar and M. Marvian for point-
h1 = h2 = 0 this gap vanishes (see dashed lines in ing out unitary transformations that cure nonstoquas-
Fig. 3c), nullifying oscillations as seen in Fig. 10b and ticity at zero bias. We also thank J. Biamonte, I. Hen,
10e. This is a result of the destructive interference be- P. Love, and P. Saint-Jean for fruitful discussions, and
tween the direct two-qubit cotunneling channel due to F. Hanington and A. King for carefully reading the
σ1x σ2x and σ1y σ2y terms and the indirect tunneling channels manuscript.

[1] T. Kadowaki and H. Nishimori, Quantum annealing in the F. Altomare, K. Boothby, P. I. Bunyk, C. Enderud, A.
transverse Ising model, Phys. Rev. E 58, 5355 (1998). Frechette, E. Hoskinson, N. Ladizinsky, T. Oh, G. Poulin-
[2] E. Farhi, J. Goldstone, S. Gutmann, J. Lapan, A. Lund- Lamarre, C. Rich, Y. Sato, A. Yu. Smirnov, L. J. Swenson,
gren, and D. Preda, A quantum adiabatic evolution al- M. H. Volkmann, J. Whittaker, J. Yao, E. Ladizinsky, M.
gorithm applied to random instances of an NP-complete W. Johnson, J. Hilton, and M. H. Amin, Observation of
problem, Science 292, 472 (2001). topological phenomena in a programmable lattice of 1,800
[3] G.E. Santoro, R. Martonak, E. Tosatti, and R. Car, The- qubits, Nature 560, 456 (2018).
ory of quantum annealing of an Ising spin glass, Science [8] S. Bravyi, D. P. DiVincenzo, R. I. Oliveira, and B. M.
295, 2427 (2002). Terhal, The Complexity of Stoquastic Local Hamiltonian
[4] A. Mott, J. Job, J.-R. Vlimant, D. Lidar, and M. Spirop- Problems, Quant. Inf. Comp. 8, 0361 (2008).
ulu, Solving a Higgs optimization problem with quantum [9] M. Marvian, D.A. Lidar, and I. Hen, On the computational
annealing for machine learning, Nature 550, 375 (2017). complexity of curing the sign problem, arXiv:1802.03408
[5] M. H. Amin, E. Andriyash, J. Rolfe, B. Kulchytskyy, and [quant-ph] (2018).
R. Melko, Quantum Boltzmann Machine, Phys. Rev. X 8, [10] J. Klassen and B. M. Terhal, Two-local qubit Hamiltoni-
021050 (2018). ans: when are they stoquastic?, arXiv:1806.05405 [quant-
[6] R. Harris, Y. Sato, A. J. Berkley, M. Reis, F. Altomare, ph] (2018).
M. H. Amin, K. Boothby, P. Bunyk, C. Deng, C. Enderud, [11] L. Hormozi E. W. Brown, G. Carleo, and M. Troyer,
S. Huang, E. Hoskinson, M. W. Johnson, E. Ladizin- Nonstoquastic Hamiltonians and quantum annealing of an
sky, N. Ladizinsky, T. Lanting, R. Li, T. Medina, R. Ising spin glass, Phys. Rev. B 95, 184416 (2017).
Molavi,R. Neufeld, T. Oh, I. Pavlov, I. Perminov, G. [12] H. Nishimori and K. Takada Exponential enhancement
Poulin-Lamarre, C. Rich, A. Smirnov, L. Swenson, N. of the efficiency of quantum annealing by nonstoquastic
Tsai, M. Volkmann, J. Whittaker, J. Yao, Phase tran- Hamiltonians, Frontiers in ICT, 4, 2 (2017).
sitions in a programmable quantum spin glass simulator, [13] R. Babbush, P. J. Love, and A. Aspuru-Guzik, Adiabatic
Science 361, 162 (2018). quantum simulation of quantum chemistry, Scientific Re-
[7] A. D. King, J. Carrasquilla, J. Raymond, I. Ozfidan, E. ports 4, 6603 (2014).
Andriyash, A. Berkley, M. Reis, T. M. Lanting, R. Harris, [14] D. Aharonov, W. van Dam, J. Kempe, Z. Landau, S.
6

Lloyd, and O. Regev. Adiabatic quantum computation is


equivalent to standard quantum computation, SIAM J.
Comp., 37 166 (2007).
[15] J.D. Biamonte and P.J. Love, Realizable Hamiltonians
for universal adiabatic quantum computers, Phys. Rev. A
78, 012352 (2008).
[16] S.P. Jordan, D. Gosset, P.J. Love, Quantum-Merlin-
Arthur-complete problems for stoquastic Hamiltonians and
Markov matrices, Phys. Rev. A 81, 032331 (2010).
[17] R. Harris, J. Johansson, A. J. Berkley, M. W. Johnson, T.
Lanting, Siyuan Han, P. Bunyk, E. Ladizinsky, T. Oh, I.
Perminov, E. Tolkacheva, S. Uchaikin, E. M. Chapple, C.
Enderud, C. Rich, M. Thom, J. Wang, B. Wilson, and G.
Rose, Experimental demonstration of a robust and scalable
flux qubit, Phys. Rev. B 81 134510 (2010).
[18] R. Harris, T. Lanting, A. J. Berkley, J. Johansson, M.
W. Johnson, P. Bunyk, E. Ladizinsky, N. Ladizinsky, T.
Oh, and S. Han, Compound Josephson-junction coupler
for flux qubits with minimal crosstalk, Phys. Rev. B 80,
052506 (2009).
[19] M.W. Johnson, M. H. S. Amin, S. Gildert, T. Lanting,
F. Hamze, N. Dickson, R. Harris, A. J. Berkley, J. Jo-
hansson, P. Bunyk, E. M. Chapple, C. Enderud, J. P.
Hilton, K. Karimi, E. Ladizinsky, N. Ladizinsky, T. Oh, I.
Perminov, C. Rich, M. C. Thom, E. Tolkacheva, C. J. S.
Truncik, S. Uchaikin, J. Wang, B. Wilson, and G. Rose,
Quantum annealing with manufactured spins, Nature 473,
194 (2011).
[20] W. Vinci and D.A. Lidar, Nonstoquastic Hamiltonians in
quantum annealing via geometric phases, npj Quant. Info.
3, 38 (2017).
[21] J. D. Whittaker, L. J. Swenson, M. H. Volkmann, P.
Spear, F. Altomare, A. J. Berkley, B. Bumble, P. Bunyk,
P. K. Day, B. H. Eom, R. Harris, J. P. Hilton, E. Hoskin-
son, M. W. Johnson, A. Kleinsasser, E. Ladizinsky, T.
Lanting, T. Oh, I. Perminov, E. Tolkacheva, and J. Yao,
A frequency and sensitivity tunable microresonator array
for high-speed quantum processor readout, J. of Appl. Phys.
119, 014506 (2016).
[22] A.J. Berkley, M.W. Johnson, P. Bunyk, R. Harris, J. Jo-
hansson, T. Lanting, E. Ladizinsky, T. Tolkacheva, M.H.S.
Amin, G. Rose, A scalable readout system for a supercon-
ducting adiabatic quantum optimization system, Supercon-
ductor Science and Technology,10,23,105014,(2010)
[23] C.M. Quintana, Yu Chen, D. Sank, A.G. Petukhov, T.C.
White, Dvir Kafri, B. Chiaro, A. Megrant, R. Barends, B.
Campbell, Z. Chen, A. Dunsworth, A.G. Fowler, R. Graff,
E. Jeffrey, J. Kelly, E. Lucero, J.Y. Mutus, M. Neeley,
C. Neill, P.J.J. O0 Malley, P. Roushan, A. Shabani, V.N.
Smelyanskiy, A. Vainsencher, J. Wenner, H. Neven, and
J. M. Martinis, Observation of classical-quantum crossover
of 1/f flux noise and its paramagnetic temperature depen-
dence, Phys. Rev. Lett. 118, 057702 (2017).
[24] Although the capacitive coupling is off, the capacitive
loading on the individual qubits are kept in the simulation.
7

Supplementary Information where


X (Φκ − Φx )2
κ
U (Φq , Φcjj ) = −
I. SYSTEM HAMILTONIAN 2Lκ
κ=q,cjj
   
πΦcjj Φq
Our goal here is to derive a Hamiltonian of two rf- EJ cos cos 2π . (7)
Φ0 Φ0
SQUIDs that have both, inductive and capacitative, cou-
plings. This leads to the following effective Hamiltonian for the
rf-SQUID,

A. rf-SQUID with a cjj-loop Q2q


H= + Ueff (Φq ). (8)
2Cq
We begin with a description of a single rf-SQUID hav-
ing a symmetric compound Josephson junction (CJJ) As a rough approximation, the effective potential energy
loop, as it is outlined in details in Ref. [1] (see Fig. 1 of the rf-SQUID can be written as
in the main text). The Hamiltonian of this system has
(Φq − Φxq )2
 
three components: Φq
Ueff (Φq ) = − EJ (Φxcjj ) cos 2π , (9)
    2Lq Φ0
πΦcjj Φq
H = Hq + Hcjj − EJ cos cos 2π . (4) with a tunable Josephson energy
Φ0 Φ0
 x 
Φ0 I c Φcjj
Here the Hamiltonians Hq and Hcjj describe the main EJ (Φxcjj ) = cos π . (10)
body of the qubit and the cjj-loop, respectively. They 2π Φ0
are expressed as
However, in our simulations, we do not use the above
Q2κ (Φκ − Φxκ )2 approximation but actually do the minimization with re-
Hκ = + , κ = {q, cjj}. (5) spect to Φcjj as given in Eq. (6) and described in detail
2Cκ 2Lκ in Ref.[1].
The third term in Eq. (4) introduces an interaction be-
tween these two loops. The fluxes in the main body of
B. Two coupled rf-SQUIDs
the qubit and in the cjj-loop are denoted as Φq and Φcjj ,
their inductances are Lq and Lcjj , respectively. The ex-
ternal fluxes applied to the main body of the qubit and In this section we analyze two rf-SQUIDs connected
to the cjj-loop are Φxq and Φxcjj . The cjj-loop has two sym- by a mutual inductance M12 and by a capacitor C12 as
metric branches a and b. Each of them has one Josephson depicted in Fig. 1a of the main text. The inductive cou-
junction, with capacitance Ca (Cb ) and with a persistent pling between the main loops of the SQUIDs is given by
current Ia (Ib ). The total current flowing through both the formula
branches of the cjj-loop is denoted as Ic = Ia + Ib , with M12
the total Josephson energy EJ = Φ2π 0 Ic
, where Φ0 = πh̄/e UM = (Φq,1 − Φxq,1 ) (Φq,2 − Φxq,2 ), (11)
L1 L2
is the flux quantum, e is the electron charge. Effec-
tive capacitances of the main qubit and the cjj-loop are where Li ≡ Lqi (i = 1, 2).
Cq = Ca + Cb and Ccjj = Ca Cb /(Ca + Cb ), respectively. The kinetic energy of two electrostatically-coupled

The charge, Qκ = −ih̄ ∂Φ κ
, and flux, Φκ , are conjugate SQUIDs, with capacitances Ci ≡ Cqi , has the form
operators obeying the commutator: [Φκ , Qκ ]− = ih̄.
For qubits 1 and 2 studied in the main text of the paper X Ci Φ̇2q,i C12 (Φ̇q,2 − Φ̇q,1 )2
K= + , (12)
we have Lcjj,1 = 17.0 pH and Lcjj,2 = 17.2 pH, while the 2 2
i
body inductances are L1 = 231.9 pH and L2 = 239.0
pH. When Lcjj  Lq , the dynamics of the fast degrees of where we use the relation Vi = Φ̇q,i , between a volt-
freedom described by the operator Φcjj is determined by age Vi on the i-junction and the flux Φq,i . Charge Qi =
the ground state of the Hamiltonian Hcjj (see Ref. [1]).
∂L/∂ Φ̇q,i of the i−qubit is defined as a derivative of the
This ground state adiabatically follows the flux degrees of
two-qubit Lagrangian
freedom, Φq , in the main body of the rf-SQUID. One can
then neglect the kinetic part of the Hamiltonian Hcjj and 2
X
define an effective potential of the main loop by finding L=K− Ueff,i (Φq,i ) − U12 , (13)
the minimum potential energy of the system (4) for each i=1
given flux Φq :
where Ueff,i (Φq,i ) is the effective potential energy of the
Ueff (Φq ) = minΦcjj U (Φq , Φcjj ), (6) i−qubit (see Eqs. (6) and (9)). Using Eqs. (12, 13), we
8

obtain the relation between charge and time derivatives C. Reduction approach
of flux,
Q1 = (C1 + C12 )Φ̇q,1 − C12 Φ̇q,2 , Working with continuous models becomes computa-
tionally challenging beyond a small number of coupled
Q2 = (C2 + C12 )Φ̇q,2 − C12 Φ̇q,1 . (14) rf-SQUIDs. With the goal to use rf-SQUIDs as qubits,
We can now writePthe total Hamiltonian of two coupled one needs to reduce the continuous Hamiltonian to a dis-
rf-SQUIDs, H = i Φ̇q,i Qi − L, as crete (qubit) Hamiltonian. For uncoupled SQUIDs, we
2
choose the following superpositions of two lowest-energy
X states with the mixing angle θi :
H= Hi + UM + UC . (15)
i=1
|↓ii = cos θi |χi1 i + sin θi |χi2 i
Here
|↑ii = − sin θi |χi1 i + cos θi |χi2 i. (24)
Q2
Hi = i +Ueff,i (Φq,i ) (16)
2C̃i The basis states |↓i i and |↑i i correspond to the left and
right circulating currents, or, equivalently, to the left and
is the Hamiltonian of i-qubit. The qubits are now char-
right sides of the SQUID potential well. The mixing angle
acterized by the loaded capacitances:
θi is chosen to maximize the left-well population in the
C12 C2 state |↓i i and, thus, the right-well population when the
C̃1 = C1 + ,
C2 + C12 SQUID is in the state |↑i i. The interaction Hamiltonian
C12 C1 mixes the states |↓i i, |↑i i with higher energy states of the
C̃2 = C2 + . (17) individual rf-SQUIDs that one needs to take into account
C1 + C12
for a correct description of the coupled system.
The inductive interaction between qubits is given by (11).
Since we are interested in the low energy spectrum
The capacitative coupling is determined by the potential
of the coupled system, the number of eigenstates taken
C12 Q1 Q2 into account in Eqs. (19,20) can be truncated to just
UC = . (18)
C1 C2 +(C1 +C2 )C12 two states for each rf-SQUID. With this truncation four
eigenvectors, |ηa i (a = 1, . . . , 4), of the two-SQUID sys-
Diagonalizing the single-SQUID Hamiltonian Hi (16)
tem are approximated as
we obtain a set of energy eigenstates, |χiµ i, and eigenen-
ergies, εiµ , such that 2 2
1 XX a 1
|ηa i ' c |χ ⊗ χ2ν i, (25)
Ni
X Na µ=1 ν=1 µν µ
Hi = εiµ |χiµ ihχiµ |. (19)
µ=1
where the amplitudes caµν are given in Eq. (22) , and the
For each SQUID we take into account a large number, normalization coefficient is calculated as
Ni  1, of the energy eigenstates and write the Hamilto- v
nian (15) of two coupled rf-SQUIDs in the basis formed u 2 2
by direct products |χ1µ ⊗ χ2ν i ≡ |χ1µ i ⊗ |χ2ν i,
uX X
Na = t |caµν |2 .
N1 N2 µ=1 ν=1
X X
H= ε1µ |χ1µ ihχ1µ | + ε2ν |χ2ν ihχ2µ | + UM + UC . (20)
µ=1 ν=1
We apply the Gram-Schmidt procedure to the four states
given in Eq. (25) to obtain the orthonormalized set of the
Here charge and flux operators should be also written in two-qubit basis states obeying the relation: hηa |ηb i = δab .
the |χ1µ ⊗ χ2ν i basis. The eigenstates, |ηa i, of the Hamil- This reduction approach only works in the limit where
tonian (20) become Na ≈ 1 for a = 1, . . . , 4. In this case the most of the
N1 X
N2 population of two-SQUID system is distributed over the
tensor products |χ1µ ⊗ χ2ν i of two lowest eigenstates of the
X
|ηa i = caµν |χ1µ ⊗ χ2ν i, (21)
µ ν isolated rf-SQUIDs (µ, ν = 1, 2).
In order to derive the reduced Hamiltonian of two cou-
where the amplitudes are given as
pled SQUIDs we start with a Hamiltonian (23) trun-
caµν = hχ1µ ⊗ χ2ν |ηa i. (22) cated to the four lowest-energy states. In the energy
basis this Hamiltonian is described by the diagonal ma-
The two-SQUID Hamiltonian,
trix: H = diag(ε1 , ε2 , ε3 , ε4 ). The Hamiltonian H can
N
X be written in the χ−basis formed by the four vectors
H= εa |ηa ihηa |, (23) |χ1µ ⊗ χ2ν i (µ, ν = 1, 2) by applying the rotation matrix
a=1 R1 = (|η1 i |η2 i |η3 i |η4 i) such that Hχ = R1 HR1T . Fi-
is characterized by the energy spectrum εa , with a total nally, the unitary matrix R2 = (|1i |2i |3i |4i) rotates the
number of levels N = N1 N2 . Hamiltonian into the computational basis, formed by the
9

vectors: After this rotation the Hamiltonian H = R2 Hχ R2T can


be represented by the four-by-four matrix,
|1i = |↓1 ⊗ ↓2 i, |2i = |↓1 ⊗ ↑2 i,
|3i = |↑1 ⊗ ↓2 i, |4i = |↑1 ⊗ ↑2 i. (26)

Jzz − h1 − h2 −∆2 /2 − Jzx −∆1 /2 − Jxz Jxx − Jyy


 
 −∆2 /2 − Jzx −Jzz − h1 + h2 Jxx + Jyy −∆1 /2 + Jxz 
H= , (27)
−∆1 /2 − Jxz Jxx + Jyy −Jzz + h1 − h2 −∆2 /2 + Jzx 
Jxx − Jyy −∆1 /2 + Jxz −∆2 /2 + Jzx Jzz + h1 + h2

or by the Hamiltonian of two coupled spins (qubits): proach agrees with the two-qubit Hamiltonian given by
Eqs. (27) and (28).
∆1 x ∆ 2 x X
H=− σ1 − σ2 +h1 σ1z +h2 σ2z + Jαβ σ1α σ2β , (28)
2 2
α,β
D. Unitary rotations and stoquasticity
where α, β = x, y, z. We use the following representation
of the Pauli matrices: Hamiltonian (27) can have positive off-diagonal matrix
elements under several conditions but not all of those
σix = |↓i ih↑i | + |↑i ih↓i |, make the Hamiltonian nonstoquastic. One has to take
σiy = i(|↓i ih↑i | − |↑i ih↓i |), into account all local transformations of the basis be-
σiz = |↑i ih↑i | − |↓i ih↓i |. (29) fore deciding about nonstoquasticity of the Hamiltonian.
Using the unitary transformation H → U † HU , with
The parameters of the Hamiltonian (28) are extracted U = ⊗i (σiz )(1−sign ∆i )/2 , one can make all ∆i σix terms
from a comparison of the operator H = R2 Hχ R2T with in the Hamiltonian nonpositive regardless of the sign of
the matrix (27). We note that Jyx = Jxy = Jyz = Jyz = ∆i . Hamiltonian (27) has positive matrix elements if
0, since the Hamiltonian is real. We also find that coef- |Jyy | > |Jxx |, since matrix elements H14 and H23 have
ficients Jxz and Jzx are quite small: Jxz , Jzx  ∆1 , ∆2 . opposite signs, due to Jyy , making one of them positive.
The non-negligible parameters of the Hamiltonian H de- However, this condition is not sufficient for nonstoquas-
scribing interactions between the qubits are shown in ticity. To demonstrate this fact, we apply a Hadamard
Fig. 3d of the main text. rotation on both qubits turning x−axis to z−axis and
One can equivalently use the projection technique of vice versa, so that σix = τiz , σiy = τiy , and σiz = −τix .
Ref. [4] to derive a reduced Hamiltonian for two in- Here τix , τiy , τiz are Pauli matrices of i-qubit in the ro-
teracting subsystems, namely our SQUIDs. The re- tated frame. After the Hadamard rotation the two-qubit
duced Hamiltonian obtained with the projection ap- Hamiltonian has the form:

Jxx + ∆1 /2 + ∆2 /2 −h2 + Jxz −h1 + Jzx Jzz − Jyy


 
−h2 + Jxz −Jxx + ∆1 /2 − ∆2 /2 Jzz + Jyy −h1 − Jzx
H= . (30)
 
−h1 + Jzx Jzz + Jyy −Jxx − ∆1 /2 + ∆2 /2 −h2 − Jxz
Jzz − Jyy −h1 − Jzx −h2 − Jxz Jxx − ∆1 /2 − ∆2 /2

This Hamiltonian has positive matrix elements at |Jyy | > of the σ y σ y interaction strength: |Jzz | ∼ |Jyy |.
|Jzz |. At non-zero biases, hi 6= 0, (and at non-zero ∆i ) At zero bias, hi = 0, one can rotate (30) in the y-z
only rotations in the x-z plane are allowed [5] since ro- plane to bring Jyy to the diagonal elements of the Hamil-
tations in x-y and z-y planes introduce complex matrix tonian eliminating the nonstoquastic effect of Jyy .
elements in the Hamiltonian. In Fig. 3d of the main text, The matrix (30) also provides the intuition necessary
we search over all possible local rotations outlined in de- to see effects of Jxx and h1 , h2 terms on experimentally
tail in Ref. [5] and find, in particular, that at the borders measured spectra. When Jzz = 0 and h1 = h2 = 0,
of the nonstoquastic range the σ z σ z coupling is of order the Hamiltonian is block-diagonal as there is no inter-
10

FIG. 5: Measured Coupler M12 vs coupler bias Φxco and a fit FIG. 6: Measured persistent current Ip versus Φxcjj that con-
to the classical model of the coupler. trols the barrier height of the double-well potential and a fit
to the classical model of Qubit 1.

action between FM and AFM states. In absence of a


Jxx
q term, the eigenvalues of the Hamiltonian becomes
∆1 ±∆2 2

2
± Jyy + . Without the Jxx term, the two
2
highest eigenvalues can only cross if ∆1 ∆2 = 0. However
we see a clear level crossing in spectroscopy vs. Φxcjj,1
plots in the middle row of Fig. 8 where ∆i 6= 0. This is
a clear signature of a Jxx type interaction. Furthermore,
the second and third excited states, that cross in presence
of Jxx belong to two separate blocks of the Hamiltonian.
When hi 6= 0 the Hamiltonian is no longer block diago-
nal and the two highest excited states start interacting.
Hence, the avoided crossings visible in energy vs Φxcjj,1
and M12 in Fig. 3a,c of the main text are clear signa-
tures of non-zero h1,2 .

FIG. 7: (a) Pulse sequence and the effective qubit poten-


II. QUBIT AND COUPLER PARAMETER tial at each pulse segment for the single-qubit spectroscopy
CHARACTERIZATION experiments. (b) Color plot of the microwave spectroscopy
data versus the flux bias Φxcjj that controls the barrier height
for Qubit 1. (c) Extracted spectroscopy peak positions that
A. Coupler characterization represent the qubit frequency versus Φxcjj and a fit to the nu-
merical rf-SQUID model.
We follow Ref. [2] for the characterization of the tun-
able magnetic coupler. In Fig. 5, we show the measured
coupler M12 versus the coupler flux bias Φxco . The data B. Quasi-static qubit characterization
is fitted to a classical model:
When the tunneling barrier is high, where the single-
qubit tunneling is largely suppressed, the properties of
2 a flux qubit can be described by a classical model [1].
Mco,q β cos(ϕxco /2) (0) In this regime, we measure the qubit persistent current
M12 = + M12 , (31)
Lco 1 + β cos(ϕxco /2) versus the flux bias that control the barrier height Φxcjj
across three flux quantum Φ0 for each qubit individually.
with Mco,q the mutual inductance between the qubit Fits to the classical model (see Fig. 6) yield the following
and the coupler, Lco the coupler inductance, β = qubit parameters: Ic,1 = 3.227 µA, Ic,2 = 3.157 µA,
(0)
2πLco Ic,co /Φ0 , M12 the stray mutual inductance be- L1 = 231.6 pH, L2 = 239.0 pH.
tween the qubits, and ϕxco = 2πΦxco the normalized ex-
ternal bias of the coupler. From the fitting, we obtained
2 (0) C. Single-qubit spectroscopy
Mco,q /Lco = 10.77 pH, β = 1.416, and Mef f = 1.848 pH.
Using Eq. (31) and the above fitting parameters, we can
set the coupler to any coupling strength within the range The qubit energy eigenstates can be characterized by
|M12 | ≤ 8.145 pH. performing microwave spectroscopy. In Fig. 7(a), we
11

show the pulse sequence for measuring the qubit spec- applied to set the qubit to a given ∆ is determined using
troscopy and the effective qubit potential at each seg- the model extracted above. Note that the effective single-
ment of the pulse sequence. The qubit is first initialized qubit ∆ is influenced by the capacitive coupling that is
in its ground state by adiabatically preparing a symmet- always present. As a result the individual qubit dynamics
ric double-well potential with a relatively low tunneling are always affected by the presence of the secondary qubit
barrier. In this limit, the qubit frequency, which is set even if it is in monostable state (Φxcjj = 0.5Φ0 ).
by the tunneling amplitude, is much larger than 1GHz
(∆  kB T ). Then, we apply a microwave pulse Φxq to
excite the qubit to its excited state. We sweep the fre-
quency of the microwave pulse from 0.5GHz to 8GHz
D. Two-qubit spectroscopy
to probe all excited states in that range. After the mi-
crowave pulse, an adiabatic tilt followed by a quench,
that increases the barrier height, is applied to the qubit The coupled system energies can be probed by mi-
to project the qubit ground and excited states to the crowave spectroscopy in a way that is similar to single-
clockwise and counter-clockwise persistent current states qubit spectroscopy. In the main text we show the pulse
for readout. The same pulse sequence was used Ref. [3]. sequence of the two-qubit spectroscopy. Although a mi-
In this experiment, we use a rise time of 1 ns for both crowave pulse is only applied to Qubit 1, the energy
the adiabatic tilt and the quench on the barrier height. eigenstates of both qubits are read out by simultaneous
Throughout the single qubit spectroscopy experiments, adiabatic tilts followed by quench. We plot the spec-
we keep the magnetic coupling strength at M12 = 0 and troscopy, as the excited state population of Qubit 1, ver-
the other qubit at Φxcjj = 0.5Φ0 . In Fig. 7(b), we show sus Φxcjj,1 that controls the barrier height of Qubit 1. The
the spectroscopy versus Φxcjj,1 for Qubit 1. The extracted barrier height of the second qubit is set to an effective
qubit frequency from this figure along with the fit to rf- single-qubit tunneling amplitude of ∆2 . Using the pre-
SQUID numerical model is shown in Fig. 7(c). viously extracted circuit parameters and initializing the
Effective single-qubit ∆ corresponds to the energy gap fitting procedure by the design values for the unknown
between the ground state and the first excited state in parameters, we extract C1 = 119.5 fF, C2 = 116.4 fF,
the single-qubit spectroscopy measurements. Once we and C12 = 132.0 fF and Lcjj,1(2) = 17.016(17.175)pH.
have an accurate model that predicts the location of the At this point, we have all the circuit parameters of this
first excited state, the appropriate Φxcjj,i that needs to be system as summarized in the following table:

Qubit Ic,i (µA) Li (pH) Lcjj,i (pH) Ci (fF) C12 (fF) |M12 |(pH)
Q1 3.22697±4.1e-5 231.633±4.5e-3 17.016 119.5
132.0 ≤ 8.145
Q2 3.15711±3.6e-5 238.981±04.1e-3 17.175 116.4

We also perform similar two-qubit spectroscopy for times (≈ 200 ps). Here, we discuss our method of mea-
various ∆2 and M12 at h1 = h2 = 0 and compare the sys- suring and correcting pulse distortion in-situ.
tem energy spectra with numerical simulation using pre-
viously extracted parameters. In Fig. 8, we show the two- We first measure the single-qubit coherent oscillations
qubit spectroscopy data and calculated energy spectra of at ∆/h = 5 GHz, where the qubit population in the
the system for all the combination of ∆2 /h = 1.5 GHz, computational basis is supposed to oscillate at the fre-
2.0 GHz, and 3.0 GHz and M12 = 0 pH and ±2 pH. quency ∆. Pulse distortion is mainly caused by reflec-
Very good agreement between the numerical model and tions. As a result, the Φxcjj (t) signal reaching the qubit
the data is achieved. In all figures the solid (dashed) deviates from the ideal square pulse which may distort
lines represent excitations from the ground (first excited) the frequency of the coherent oscillations from the tar-
state. get ∆ at a given time τ . To measure the distorted
pulse in the time domain, we slice the coherent oscil-
lation data into small time windows, typically with a
length of 400 ps that contains more than one period of
III. PULSE DISTORTION COMPENSATION
FOR COHERENT OSCILLATIONS
oscillation. We then extract the coherent oscillation fre-
quency Ω(τ ) in each slice starting at time τ and treat
this frequency as a measurement of the instantaneous
Short duration pulse (≤ 10 ns) distortion imposes a ∆m (τ ) = Ω(τ ). The pulse distortion at any given time
great limitation on the fidelity of coherent qubit opera- ∂∆
is then calculated as τ by δΦxcjj (τ ) = (∆m (τ ) − ∆)/ ∂Φx
tion. In our experiment, the qubit control involves ap- ∂∆
cjj

plying fast pulses to lower and rise the tunneling barrier where ∂Φx can be evaluated numerically using the qubit
cjj
of the flux qubits. These pulses have short rise and fall model. To correct for the pulse distortion, we simply ap-
12

FIG. 8: Two-qubit spectroscopy versus Φxcjj,1 for ∆2 /h ≈ 1.5 GHz, 2.0 GHz, and 3.0 GHz and M12 = 0 pH and ±2 pH. The
solid and dashed lines represent numerical calculations of energy differences. The solid lines correspond to excitations from the
ground state and the dashed lines correspond to excitations from the first excited state.

ply the first order correction to the applied pulse with IV. COHERENT OSCILLATION PROTOCOL
Φxcjj,corr (τ ) = Φxcjj (τ ) + δΦxcjj (τ ) at the same time τ . As
the above correction at time τ may lead to distortions at
times t > τ , we iterate the entire measurement and cor- The pulse sequence and the effective two-qubit poten-
rection procedure until the corrected pulse converges. In tial at each pulse segment of a coherent oscillation pro-
practice, this pulse distortion correction procedure con- tocol is shown in Fig. 10. The qubits are initialized in
verges within 5 iterations. In Fig. 9, we show the co- two steps. A large flux-bias, Φxq,prep , is applied to tilt
herent oscillation data and the extracted instantaneous the potential in its monostable state. Next, the tun-
tunneling amplitude ∆m with uncorrected pulse (blue) neling strength is reduced, keeping the potential tilted.
and corrected(green) pulses Φxcjj , shown respectively. The Once both qubits are prepared in a computational state
measured tunneling amplitude ∆m /h is fixed to the tar- controlled by the signs of Φq,prep pulses, the tilts are
get value 5 GHz after the pulse distortion compensation removed. The coherent oscillations are induced by re-
is applied, which significantly increases the fidelity of the ducing the barriers of both qubits simultaneously to al-
coherent oscillation. low quantum fluctuations drive the coherent dynamics of
the system. As the computational basis states are not
the eigenstates of the total Hamiltonian, the system un-
dergoes coherent oscillations between the computational
basis states with near-degenerate Ising energies. After
13

In coherent oscillation experiments, the system is pre-


pared and measured in one of the four computational ba-
sis states | ↑↑i, | ↑↓i, | ↓↑i, and | ↓↓i. In Fig. 10, we show
that the coherent oscillations are indeed mostly between
the two low-energy configurations at strong magnetic
couplings. At strong FM coupling with M12 = −2 pH,
the coherent oscillations occur between | ↑↑i and | ↓↓i,
whereas the coherent oscillations occur between | ↑↓i and
| ↓↑i at strong AFM coupling with M12 = 2 pH.

V. COHERENCE

The phase coherence time T2 between the energy eigen-


FIG. 9: Characterization and compensation of pulse distor- states of each qubit can be characterized by measuring
tion. The coherent oscillation data, measured tunneling am- single-qubit coherent oscillations. As in Eq. (24), at zero
plitude ∆m , and applied barrier pulses without (blue) and bias we define the computational basis as | ↑i = |χ1 i+|χ
√ 2
2
i
with (green) pulse distortion corrections.
and | ↓i = |χ1 i−|χ i
√ 2 for each qubit. In Fig. 11, we show
2
the coherent oscillations of Qubit 1 at ∆1 = 2 GHz.
some time τ , the states are read out by measuring qubit- A fit to the decay envelop of the oscillations yields
persistent currents after simultaneously quenching both T2 = 16.2 ns. Energy relaxation time T1 at the same
qubits. The rise and fall times of the pulse, about 200 ps, operation point is characterized by measuring the qubit
are much faster than the dynamics of the qubits to snap- excited state population versus the delay time between
shot the qubit population at the end of evolution. We initialization and readout. By fitting the excited-state
repeat this process for a range of dwell times τ and mag- population decay to an exponential function, we extract
netic coupling strengths. This protocol works equally T1 = 17.5 ns. We perform the same characterization on
well for a single-qubit if the qubit potential is kept at its Qubit 2 and obtain T2 = 14.3 ns and T1 = 17.4 ns at
monostable state. ∆2 = 2 GHz.

[1] R. Harris, J. Johansson, A. J. Berkley, M. W. Johnson, T. E. Jeffrey, J. Kelly, E. Lucero, J.Y. Mutus, M. Neeley,
Lanting, Siyuan Han, P. Bunyk, E. Ladizinsky, T. Oh, I. C. Neill, P.J.J. OMalley, P. Roushan, A. Shabani, V.N.
Perminov, E. Tolkacheva, S. Uchaikin, E. M. Chapple, C. Smelyanskiy, A. Vainsencher, J. Wenner, H. Neven, and
Enderud, C. Rich, M. Thom, J. Wang, B. Wilson, and G. J. M. Martinis, Observation of classical-quantum crossover
Rose, Experimental demonstration of a robust and scalable of 1/f flux noise and its paramagnetic temperature depen-
flux qubit, Phys. Rev. B 81, 134510 (2010). dence, Phys. Rev. Lett. 118, 057702 (2017).
[2] R. Harris, T. Lanting, A. J. Berkley, J. Johansson, M. [4] M.H. Amin, A.Yu. Smirnov, N.G. Dickson, and M. Drew-
W. Johnson, P. Bunyk, E. Ladizinsky, N. Ladizinsky, T. Brook, Approximate diagonalization method for large-scale
Oh, and S. Han, Compound Josephson-junction coupler Hamiltonians, Phys. Rev. A 86, 052314 (2012).
for flux qubits with minimal crosstalk, Phys. Rev. B 80, [5] J. Klassen an B.M. Terhal, Two-local qubit Hamiltonians:
052506 (2009). when are they stoquastic? arXiv:1806.05405 [quant-ph]
[3] C.M. Quintana, Yu Chen, D. Sank, A.G. Petukhov, T.C. (2018).
White, D. Kafri, B. Chiaro, A. Megrant, R. Barends, B.
Campbell, Z. Chen, A. Dunsworth, A.G. Fowler, R. Graff,
14

FIG. 10: Two-qubit coherent oscillations. a. Pulse sequence and the effective qubit potential during each segment of
the pulse. The potential landscape represents a diagonal cut across the landscape depicted in Fig. 1a of the main text. The
direction depends on the initial tilt applied on the two qubits. b, c. Data (b) and simulations (c) of the coherent oscillations
between the low-energy states versus the magnetic coupling strength M12 . For M12 < 0, the qubits are initialized in | ↑↑i and
measure the probability of the state | ↓↓i. For M12 ≥ 0, the qbuits are initialized in | ↓↓i and the population of P↑↓ and P↑↑ are
plotted respectively.(a,b) Two-qubit coherent oscillation between the computational basis states with strong FM coupling at
M12 = −2 pH (a) and strong AFM coupling at M12 = 2 pH (b). (c,d) Color plots of the two-qubit coherent oscillation versus
the magnetic coupling strength M12 with the system initially prepared at | ↓↓i (c) and | ↓↑i (d) respectively.

FIG. 11: T1 and T2 measurements on Qubit 1 at ∆ = 2 GHz


and at zero bias, h1 = 0. By fitting the decay envelop of
the coherent oscillation to an exponential function, we obtain
T2 = 16.2 ns. By fitting the energy relaxation of the first
excited state, we obtain T1 = 17.5 ns.

Potrebbero piacerti anche