High-Fidelity Qutrit Entangling Gates For Superconducting Circuits
High-Fidelity Qutrit Entangling Gates For Superconducting Circuits
Noah Goss,1, 2, ∗ Alexis Morvan,2 Brian Marinelli,1, 2 Bradley K. Mitchell,1 Long B. Nguyen,2 Ravi K. Naik,1, 2
Larry Chen,1 Christian Jünger,2 John Mark Kreikebaum,1, 3
David I. Santiago,2 Joel J. Wallman,4 and Irfan Siddiqi1, 2, 3
1
Department of Physics, University of California, Berkeley, Berkeley CA 94720, USA.
2
Computational Research Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA.
3
Materials Science Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA.
4
Keysight Technologies Canada, Kanata, ON K2K 2W5, Canada.
(Dated: June 17, 2022)
arXiv:2206.07216v2 [quant-ph] 16 Jun 2022
Quantum error correction (QEC) [1] is necessary for gates remains a major challenge in superconducting cir-
noisy intermediate-scale quantum (NISQ) [2] computers cuits.
to realize their full potential. The surface code [3, 4] us- The most commonly used qubit in superconducting cir-
ing qubits is considered the main route to fault toler- cuits [29], the transmon [30], is well suited to be oper-
ance [5–8], though its technical challenges have led the ated as a qutrit due to its weak anharmonicity. Techni-
community to explore other approaches that could have cal advancements in microwave engineering and improved
more favorable QEC schemes, such as storing a two level fabrication techniques have increased transmon coherence
system in the large Hilbert space of quantum oscillators times [31], enabling coherent control of the full qutrit
[9, 10]. Another alternative is to use d-dimensional quan- Hilbert space. Furthermore, dispersive readout can be
tum objects, or qudits, which mobilize a larger and more used for high-fidelity single shot qutrit readout [27]. In
connected computational space than their qubit counter- addition, high-fidelity single qutrit operations [32, 33],
parts. Qutrits, the simplest form of qudits, can provide quantum information scrambling [28], compact decompo-
advantages in QEC for magic state distillation [11, 12], sitions of multi-qubit gates [34–37], and improved qubit
compactly encoding qubits [13], and can be used to en- readout [38] have all been demonstrated using transmons
code logical qutrits themselves [14–16]. There are several as qutrits. Nonetheless, past qutrit entangling gates in
proposals utilizing qutrits to improve quantum algorithms transmons have been limited by relying on either a slow,
[17–20] and in the short term, qutrits can improve quan- static interaction or an interaction that restricts the en-
tum simulation [21] and other NISQ applications [22]. Re- tanglement to only a subspace of the qutrit.
alizing multi-qudit systems, however, is challenging due to In this work, we characterize the differential AC Stark
the complexities of the larger Hilbert space. Nonetheless, shift [39–42] on two fixed frequency transmon qutrits
coherent control of qudits has been performed in several with static coupling and leverage it to generate dynamic
physical platforms [23–28]. While state of the art exper- qutrit entangling phases. With this interaction, we en-
iments have demonstrated high-fidelity qudit entangling gineer the ternary controlled-Z gate (CZ) and its inverse
gates with trapped ions [23, 24] and photonic circuits [26], (CZ† ). Both gates performed in our work are univer-
generating high-fidelity, maximally entangling two-qudit sal for ternary computation and Clifford gates needed for
QEC in qutrits. We achieve an estimated process fidelity
of 97.3(1)% and 95.2(3)% for the CZ† and CZ respec-
tively, measured using cycle benchmarking [43] and our
∗ Correspondence should be addressed to [email protected] generalization of the cross-entropy benchmarking routine
2
a b
Stark Drive Duration
Hadamard Tomography
ℋ(𝜔! , 𝜙, Ω)
Q ! |0⟩
%$ (% %$ (%
𝑋"/$ 𝑋"/$ 𝑋"/$ 𝑋"/$
Stark Drive Duration
Hadamard Tomography
ℋ(𝜔! , 𝜙, Ω)
Q " |0⟩
%$ (% %$ (%
𝑋"/$ 𝑋"/$ 𝑋"/$ 𝑋"/$
c d e
𝜔%$,'" /$"
𝜔%$,'! /$"
𝜔(%,'! /$"
𝜔(%,'" /$"
Figure 2. Characterizing the dynamical cross-Kerr entanglement. a, To study the accumulation of entangling phases
under the driven cross-Kerr interaction, we place two qutrits in a full superposition using ternary Hadamard gates (virtual Z
gates ommited in diagram), then study the evolution under the Stark drive scheme by performing state tomography. b, We
demonstrate fitting the accumulation of entangling phase found by tomography to our linear, driven cross-Kerr model, where
αij is the slope of the line and the uncertainty is from the linear fit. c-d, We match the behavior of the cross-Kerr entanglement
given relevant experimental parameters in our system to our Hamiltonian model for the relative phase of the driving, φ, and
amplitude of the driving, fixing Ω = Ωa = Ωb . e, We additionally compare the dependence of α12 on the frequency of the drive
ωd using an ab-initio master equation simulation in QuTiP [45, 46].
sical crosstalk, or parasitic two level systems (TLS) in and robust gate scheme and this “fine-tuned” approach,
our device. In an experimental setting, the flexibility of while still taking advantage of the dynamical nature of our
this entanglement allows us the freedom to set the drive cross-Kerr interaction. By employing the pulse scheme in
frequency far from any of these features. Fig. 3, where echo pulses in the {|1i , |2i} subspace shuffle
Qutrit CZ/CZ† gate entangling phases, we have the modified unitary evolution
(omitting the single-qutrit phases for brevity):
We next construct qutrit controlled-phase gates utiliz-
ing this entangling interaction. The qutrit CZ and CZ† U = exp(−i[(α11 + α22 )τ (|11ih11| + |22ih22|)+
gate are both maximally entangling and members of the (4)
(α12 + α21 )τ (|12ih21| + |21ih21|)]),
two-qutrit Clifford group making them particularly useful
gates for ternary computation. The CZ gate is defined as: Tuning the experimental knobs demonstrated in Fig. 2,
X we are able to find the relaxed conditions on our cross-
UCZ = ω ij |ijihij| , (3) Kerr evolution: (α11 + α22 ) ≈ −(α12 + α21 ) and (α11 +
i,j∈{0,1,2}2 α22 ) ≈ 2(α12 + α21 ). Under these conditions, at some
drive time τ , we will have approximately acquired the
with ω = e2iπ/3 , the third root of unity. Under simultane- desired entangling phases found in Eq. 3 to synthesize re-
ous Stark drives, the two-qutrit Hilbert space follows the spectively a CZ or CZ† unitary. We find that when using
unitary evolution U = exp(−i(H+φ1 I ⊗Z 01 +φ2 I ⊗Z 12 + flat-top cosine pulses for the cross-Kerr drive, the ramp
φ3 Z 01 ⊗I +φ4 Z 12 ⊗I)τ ) where H is given in Eq. 1. To per- features can lead to offsets on the linear accumulation
form a CZ gate with a single round of cross-Kerr driving, of entangling phase; for this reason, the highest fidelity
for example, one would need to find driving parameters implementation of our gates only uses an approximate
meeting the conditions: {α11 = α22 = 2α21 = 2α12 }, relationship on the αij terms with some adjustment nec-
a task that is not broadly feasible. Practically speak- essary. Finally, as outlined schematically in Fig. 3, one
ing, we desire a compromise between the most general can undo the local Z-like phases (found via tomography)
†
τ τ a d
×𝒎
ℋ(𝜔$ , 𝜙, Ω) ℋ(𝜔$ , 𝜙, Ω)
Q! Z!" 𝜙" |0⟩ B P P B†
CZ !
𝑋!"# 𝑋!"# Z"#(𝜙#)
τ τ |0⟩ B P P B†
CB
Q" ℋ(𝜔$ , 𝜙, Ω) ℋ(𝜔$ , 𝜙, Ω) Z!" 𝜙$
Z"#(𝜙%) b ×𝒎
𝑋!"# 𝑋!"# †
|0⟩ G G
CZ!
XEB
Figure 3. Gate schematic. For the CZ and CZ† gate, we |0⟩ G G
perform two rounds of cross-Kerr entanglement for duration τ XEB
with interleaved echo pulses in the {|1i , |2i} subspace which c
shuffle the entangling phases. For proper conditions on the αij
terms in Eq. 1, the CZ† (CZ) is compiled with a total gate time
of 580(783) ns. The local Z terms in both two level subspaces
†
CB
in both the {|0i,|1i} and {|1i,|2i} subspaces with virtual
Z gates [51]. In this work, we performed the CZ and CZ†
on two different pairs of transmon qutrits, demonstrating
the flexible nature of generating two-qutrit gates from this
driven cross-Kerr scheme. Figure 4. Benchmarking. a, Circuit schematic of cycle
benchmarking (CB). The errors of the CZ† are twirled via
Benchmarking random Weyl gates (red) to tailor errors into stochastic Weyl
channels. The initial state and measurement basis (blue) are
We first benchmark our two-qutrit gates with cycle selected to pick out the decay associated with specific Weyl
benchmarking (CB) [43] using True-Q [52]. While origi- channels. b, Circuit schematic of cross-entropy benchmarking
nally written in terms of qubits, CB naturally generalizes (XEB). The errors of the CZ† are twirled via random SU(3)
to qutrits [32]. We use CB instead of, e.g., interleaved gates (green) to tailor the noise to a simple depolarizing chan-
randomized benchmarking [53, 54], because it requires nel. c, An integrated histogram of CB for both the CZ† gate
significantly fewer multi-qutrit gates per circuit. We de- and a reference cycle, with the solid vertical lines giving the
scribe the generalization in the supplementary material. fidelities 0.936(1) and 0.966(1) respectively, yielding an esti-
mated process fidelity of 97.3(1)%. We extract an error bud-
With this technique, we estimate the Weyl (generalized-
get directly from CB, estimating a purity limited fidelity of
Pauli) error rate of the CZ† and CZ gate to be 2.7(1)% 0.973(9) and 0.989 (with negligible error) for the dressed CZ†
and 4.8(3)% respectively. By contrast, the highest fidelity, and reference cycles, yielding a purity limit 0.986(9) for the
two-qutrit gate performed previously with transmons had isolated CZ† gate. d, From XEB we estimate the depolarized
an error rate of 11.1% [28]. CB also allows us to construct fidelity as 0.933(3). Additionally, we estimate the speckle-
the Weyl-twirled error per channel of the unitary in Fig. 4. purity limited fidelity of the CZ† dressed with random SU(3)
This provides us with an estimate of the worst case sce- gates to be 0.961(3).
nario of less than 8% and demonstrates a relatively low
dispersion of our error channels.
As an added confirmation of the fidelity of the CZ† gate, Finally, we would like to be able to characterize what
we generalize the cross-entropy benchmarking (XEB) rou- fraction of the errors present in our two qutrit unitaries
tine [44, 55] to tailor all gate errors into a depolarizing are coherent on the time scale of multiple experiments,
channel, for work with qutrit unitaries. In the qutrit case, and thus could be removed by improved calibration. As
we find sufficient tailoring of our noise can be performed we show in supplementary note 6, under the depolariz-
by interleaving random SU(3) gates around our target ing unitary noise model, the variance of CB and XEB
gate. The circuit diagram for qutrit XEB is in Fig. 4b circuits both provide a robust method of estimating the
and the results can be found in Fig. 4e. We find that the purity limit [56, 57]. The corresponding estimates are
depolarized fidelity of the CZ† dressed with random SU(3) shown in Fig.4c with the CB estimate of 97.3(9)% exceed-
gates agrees with the estimate of the process fidelity from ing the speckle-purity limit of 96.1(3)% for the dressed
our Weyl twirled CB results within a standard error. Ad- CZ† gate. This disagreement is likely due to the fact
ditional discussion of the qutrit XEB method is provided that the CB data reveals that the noise is dominated by
in the supplementary materials. single-qutrit phase errors. As these errors are likely to
5
R1 Ri
and U, i.e. the infidelity 1-F(V, U).
R2 Rj
We perform this numerical investigation on 1000 Haar
Ansatz (V)
random gates and 1000 Clifford gates. We find that all
1000 Haar random gates can be synthesized at depth 6
for CZ/CZ† , 7 for Cinc , and 9 for Cex . The synthesis
U success rate for all 3 unitaries in terms of target Clifford
Target (U) circuits are shown in Fig. 4b. Notably, almost all two
b qutrit Clifford gates were successfully compiled at depth
2 in CZ/CZ† , with 100% success at depth 3. By contrast,
Haar Cinc Cex and Cinc did not demonstrate as much improvement
Haar Cex
Haar CZ
[1] Girvin, S. M. Introduction to quantum error correction damping. IEEE Transactions on Information Theory 64,
and fault tolerance (2021). URL https://arxiv.org/ 4674–4685 (2018).
abs/2111.08894. [17] Bocharov, A., Roetteler, M. & Svore, K. M. Fac-
[2] Preskill, J. Quantum computing in the NISQ era and toring with qutrits: Shor’s algorithm on ternary and
beyond. Quantum 2, 79 (2018). metaplectic quantum architectures. Phys. Rev. A 96,
[3] Bravyi, S. B. & Kitaev, A. Y. Quantum codes on a lattice 012306 (2017). URL https://link.aps.org/doi/10.
with boundary (1998). URL https://arxiv.org/abs/ 1103/PhysRevA.96.012306.
quant-ph/9811052. [18] Gedik, Z. et al. Computational speed-up with a single
[4] Dennis, E., Kitaev, A., Landahl, A. & Preskill, J. qudit. Scientific Reports 5, 14671 (2015). URL https:
Topological quantum memory. Journal of Mathematical //doi.org/10.1038/srep14671.
Physics 43, 4452–4505 (2002). [19] Bullock, S. S., O’Leary, D. P. & Brennen, G. K. Asymp-
[5] Google Quantum AI. Exponential suppression of bit totically optimal quantum circuits for d-level systems.
or phase errors with cyclic error correction. Nature Phys. Rev. Lett. 94, 230502 (2005). URL https://link.
595, 383 (2021). URL https://doi.org/10.1038/ aps.org/doi/10.1103/PhysRevLett.94.230502.
s41586-021-03588-y. [20] Pavlidis, A. & Floratos, E. Quantum-fourier-transform-
[6] Marques, J. F. et al. Logical-qubit operations in based quantum arithmetic with qudits. Phys. Rev. A
an error-detecting surface code. Nature Physics 18, 103, 032417 (2021). URL https://link.aps.org/doi/
80–86 (2021). URL http://dx.doi.org/10.1038/ 10.1103/PhysRevA.103.032417.
s41567-021-01423-9. [21] Gustafson, E. Noise improvements in quantum simula-
[7] Krinner, S. et al. Realizing repeated quantum error cor- tions of sqed using qutrits (2022). URL https://arxiv.
rection in a distance-three surface code (2021). URL org/abs/2201.04546.
https://arxiv.org/abs/2112.03708. 2112.03708. [22] Gokhale, P. et al. Asymptotic improvements to quan-
[8] Zhao, Y. et al. Realization of an error-correcting surface tum circuits via qutrits. In Proceedings of the 46th In-
code with superconducting qubits (2022). URL https: ternational Symposium on Computer Architecture, ISCA
//arxiv.org/abs/2112.13505. 2112.13505. ’19, 554–566 (Association for Computing Machinery, New
[9] Campagne-Ibarcq, P. et al. Quantum error correction of York, NY, USA, 2019). URL https://doi.org/10.1145/
a qubit encoded in grid states of an oscillator. Nature 3307650.3322253.
584, 368–372 (2020). URL https://doi.org/10.1038/ [23] Hrmo, P. et al. Native qudit entanglement in a trapped
s41586-020-2603-3. ion quantum processor (2022). URL https://arxiv.org/
[10] Grimm, A. et al. Stabilization and operation of a Kerr- abs/2206.04104.
cat qubit. Nature 584, 205–209 (2020). URL http://dx. [24] Ringbauer, M. et al. A universal qudit quantum proces-
doi.org/10.1038/s41586-020-2587-z. sor with trapped ions. arXiv preprint arXiv:2109.06903
[11] Campbell, E. T., Anwar, H. & Browne, D. E. Magic-state (2021). URL https://arxiv.org/abs/2109.06903.
distillation in all prime dimensions using quantum reed- [25] Lanyon, B. P. et al. Manipulating biphotonic qutrits.
muller codes. Phys. Rev. X 2, 041021 (2012). URL https: Phys. Rev. Lett. 100, 060504 (2008). URL https://link.
//link.aps.org/doi/10.1103/PhysRevX.2.041021. aps.org/doi/10.1103/PhysRevLett.100.060504.
[12] Campbell, E. T. Enhanced fault-tolerant quantum [26] Chi, Y. et al. A programmable qudit-based quantum pro-
computing in d-level systems. Phys. Rev. Lett. 113, cessor. Nature Communications 13, 1166 (2022). URL
230501 (2014). URL https://link.aps.org/doi/10. https://doi.org/10.1038/s41467-022-28767-x.
1103/PhysRevLett.113.230501. [27] Bianchetti, R. et al. Control and tomography of a three
[13] Kapit, E. Hardware-efficient and fully autonomous quan- level superconducting artificial atom. Phys. Rev. Lett.
tum error correction in superconducting circuits. Phys. 105, 223601 (2010). URL https://link.aps.org/doi/
Rev. Lett. 116, 150501 (2016). URL https://link.aps. 10.1103/PhysRevLett.105.223601.
org/doi/10.1103/PhysRevLett.116.150501. [28] Blok, M. S. et al. Quantum information scrambling on
[14] Majumdar, R., Basu, S., Ghosh, S. & Sur-Kolay, S. Quan- a superconducting qutrit processor. Phys. Rev. X 11,
tum error-correcting code for ternary logic. Phys. Rev. A 021010 (2021). URL https://link.aps.org/doi/10.
97, 052302 (2018). URL https://link.aps.org/doi/10. 1103/PhysRevX.11.021010.
1103/PhysRevA.97.052302. [29] Blais, A., Grimsmo, A. L., Girvin, S. M. & Wallraff,
[15] Muralidharan, S., Zou, C.-L., Li, L., Wen, J. & Jiang, A. Circuit quantum electrodynamics. Rev. Mod. Phys.
L. Overcoming erasure errors with multilevel systems. 93, 025005 (2021). URL https://link.aps.org/doi/10.
New Journal of Physics 19, 013026 (2017). URL https: 1103/RevModPhys.93.025005.
//doi.org/10.1088/1367-2630/aa573a. [30] Koch, J. et al. Charge-insensitive qubit design de-
[16] Grassl, M., Kong, L., Wei, Z., Yin, Z.-Q. & Zeng, rived from the cooper pair box. Phys. Rev. A 76,
B. Quantum error-correcting codes for qudit amplitude 042319 (2007). URL https://link.aps.org/doi/10.
7
Acknowledgements
We are grateful to A. Hashim, W. Livingston, and
I. Hincks for conversations and insights. This work
was supported by the Quantum Testbed Program of
the Advanced Scientific Computing Research for Basic
Energy Sciences program, Office of Science of the U.S.
Department of Energy under Contract No. DE-AC02-
05CH11231.
Supplementary Information: High-Fidelity Qutrit Entangling Gates for
Superconducting Circuits
Noah Goss,1, 2 Alexis Morvan,2 Brian Marinelli,1, 2 Bradley K. Mitchell,1, 2 Long B. Nguyen,2 Ravi K. Naik,1, 2
Larry Chen,1 Christian Jünger,2 John Mark Kreikebaum,1, 3
David I. Santiago,2 Joel J. Wallman,4 and Irfan Siddiqi1, 2, 3
1
Department of Physics, University of California, Berkeley, Berkeley CA 94720, USA.
2
Computational Research Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA.
3
Materials Science Division, Lawrence Berkeley National Laboratory, Berkeley, California 94720, USA.
arXiv:2206.07216v2 [quant-ph] 16 Jun 2022
4
Keysight Technologies Canada, Kanata, ON K2K 2W5, Canada.
(Dated: June 17, 2022)
Both gates and cross-Kerr characterization data was taken using an 8 transmon ring with fixed-frequency transmons
and fixed-frequency coupling mediated by a coplanar waveguide resonator. The CZ and CZ† gate are performed on
two different chips, using two different pairs of transmon qutrits. We give the relevant single-qutrit parameters and
coherences for the CZ and CZ† gate in Table S1 and Table S2 respectively. Here, T2r denotes coherence statistics
taken using a Ramsey experiment and T2e using an echo pulse.
Q3 Q4
Qubit freq. (GHz) 5.436 5.327
Anharm. (MHz) -260.20 -262.94
T101 (µs) 125(37) 78(16)
T112 (µs) 63(9) 47(5)
01
T2e (µs) 190(28) 138(25)
12
T2e (µs) 61(13) 45(7)
02
T2e (µs) 75(19) 62(6)
01
T2r (µs) 114(47) 99(24)
12
T2r (µs) 17(8) 17(9)
02
T2r (µs) 20(16) 21(9)
Table S1. Single-qutrit parameters for the pair of transmons used to perform the CZ gate.
Q5 Q6
Qubit freq. (GHz) 5.362 5.523
Anharm. (MHz) -275 -271.35
T101 (µs) 45(7) 58(7)
T112 (µs) 33(3) 28(3)
01
T2e (µs) 63(7) 84(6)
12
T2e (µs) 28(3) 30(3)
02
T2e (µs) 37(3) 35(3)
01
T2r (µs) 36(9) 76(8)
12
T2r (µs) 10(6) 18(6)
02
T2r (µs) 11(6) 21(8)
Table S2. Single-qutrit parameters for the pair of transmons used to perform the CZ† gate.
2
6* 6*
|3,1⟩ |2,2⟩ |1,3⟩
Ω" Ω!
"!
""
3* 4*
|3,0⟩
|2,1⟩ |1,2⟩ 3* |1,0⟩ 𝐽
Δ
|0,3⟩
|0,1⟩
"" +2$
Ω"
Ω!
Energy |", $⟩
"! +2$
2* |1,1⟩
|2,0⟩ 2*
|0,2⟩
"" + $
𝜔! + 𝜂
𝜔!
𝜔"
Ω"
"! + $
𝜔" + 𝜂
Ω! Ω! Ω"
|1,0⟩ *
Δ
|0,1⟩
|0,0⟩
""
""
"" + $
"! + $
Ω" Ω!
"!
"!
|0,0⟩
Figure S1. Drive scheme for conditional stark-induced cross-Kerr Hamiltonian . We present an example energy level
diagram in the rotating frame of the drive where we place the off resonant microwave drive on both qutrits between their two
respective |1i → |2i transitions. The simultaneous off resonant microwave drives induce conditional Stark shifts that generate
entangling phases on four states in our two qutrit Hilbert space.
In the frame of the drive at frequency ωd and after making a rotating wave approximation (RWA), the system
Hamiltonian is
Xh ηi i
H= (ωi − ωd ) a†i ai + a†i a†i ai ai + Ωi eiϕi ai + e−iϕi a†i + J a†c at + ac a†t (S1)
i=c,t
2
with ~ = 1 and the transmons approximated as Duffing oscillators with qubit frequency ωi , anharmonicity ηi , capacitive
coupling J, and where we define ai as the bosonic annihilation operator. The parameters of the drive are given as
amplitude Ωi , and phase ϕi . Since only the relative drive phase is physical, we choose a basis where ϕc = 0 and
ϕd = ϕt − ϕc . The detuning of transmon i from the drive is ∆i = ωi − ωd . We analyze the system perturbatively in the
limit Ωi , J |ηi |, |∆i |. In this limit the bare transmon Hamiltonians serve as the unperturbed system, H0 , and the
perturbation, V , is composed of the single qubit drive terms and the coupling term. Time independent perturbation
theory applied to Eq. S1 will yield energies Eij , the approximate diagonal elements of the Hamiltonian in the basis
labelled by the transmon occupation numbers |iji with i, j = 0, 1, 2, . . . , d − 1 the state of the “control” (c) and “target”
(t) transmons respectively. The dimension where we truncate the Hamiltonian is d. In the present case of qutrits
P2
d = 3 and H 0 ≈ i,j=0 Ẽij |iji hij| where we define Ẽij = Eij − E00 , performing a global shift of the energies to set
the energy of the |00i state to zero.
We isolate the entangling cross-Kerr terms by transforming H 0 according to the unitary
n h io
U = exp − it Ẽ01 I ⊗ |1i h1| + Ẽ10 |1i h1| ⊗ I + Ẽ02 I ⊗ |2i h2| + Ẽ20 |2i h2| ⊗ I , (S2)
where I = |0i h0| + |1i h1| + |2i h2| is the single qutrit identity operator. The transformation eliminates the single qutrit
energies Ẽi0 and Ẽ0j , i, j = 1, 2 which simply result in local phases that can be eliminated by virtual single qutrit
phase gates. The transformed Hamiltonian
H 00 = α11 |11i h11| + α21 |21i h21| + α12 |12i h12| + α22 |22i h22| (S3)
3
is written in terms of the cross-Kerr rates αij which describe the rates at which the entangling phases on the states
|iji are accumulated. These αij , explicitly given by
(2) 4ηJ 2
α11 = (S5)
(η − ∆)(η + ∆)
(2) 2ηJ 2 (5η − 4∆)
α21 = (S6)
∆(η − ∆)(2η − ∆)
(2) 2ηJ 2 (5η + 4∆)
α12 =− (S7)
∆(η + ∆)(2η + ∆)
(2) 16ηJ 2
α22 = (S8)
(η + ∆)(η − ∆)
where ∆ = ωc − ωt and we have set ηc = ηt = η in order to arrive at more compact expressions. In most systems
this is a reasonable approximation and in particular the pair of transmons used in this work have anharmonicities
ηc = 272 MHz and ηt = 270 MHz.
As in the qubit case the driven cross-Kerr interaction contributes starting at third order when both transmons are
driven (see the discussion in the Supplementary Materials of [1]) and we find rates
8 cos φη 2 ΩC ΩT J
α11 = (S9)
∆C ∆T (η − ∆C )(η − ∆T )
The expressions are complicated but we now point out some important features. In typical systems the drive
(3) (2)
strengths Ωi are larger than the coupling strength J and as a result αij > αij . For example, the CZ gate described
in the main text is performed with Ωc , Ωt ≈ 11 MHz on a pair of transmons with estimated coupling J = 2.7 MHz.
Therefore, the static cross-Kerr can in principle be cancelled by the driven cross-Kerr. The result also gives the
leading order linear dependence of the rates αij on the drive strengths Ωi and the coupling J as well as their sinusoidal
dependence on the relative drive phase ϕ (see Figure 2 in the main text).
4
When analyzing qubits, we use tensor products of the single-qubit Pauli operators P = {I, X, Y, Z}. These operators
have the following helpful properties:
Unfortunately, no set of operators with the same properties exist for higher dimensions. We thus need to separate
some of the properties when analyzing qudits.
There are two sets of operators, namely, the Weyl and Gell-Mann operators that, taken together, satisfy all four
properties and also coincide with the Pauli operators in 2D. We now review these two sets of operators.
The Gell-Mann operators are obtained by embedding the familiar Pauli matrices into two-dimensional subspaces of
a higher dimensional space. Recall that the standard single-qubit Pauli operators are
X = |1ih0| + |0ih1|
Y = i |1ih0| − i |0ih1|
Z = |0ih0| − |1ih1| . (S13)
X jk = |jihk| + |kihj|
Y jk = i |jihk| − i |kihj|
Z jk = |jihj| − |kihk| (S14)
for 0 ≤ j < k < d. The X jk and Y jk operators are trace-orthogonal Hermitian operators and also correspond to
transitions between two levels, which is the natural way to control a harmonic oscillator. Specifically, we can drive the
X 01 and X 12 Hamiltonians by applying tones at ω01 and ω12 , which generate Rabi oscillations in the corresponding
qubit subspaces of the qutrit, from which we can generate the single qutrit gates
1 1 −i 0 1 1 0 0
01
Xπ/2 = √ −i 1 0 , Xπ/2 12
= √ 0 1 −i . (S15)
2 0 0 1 2 0 −i 1
Similarly, we can perform virtual Z gates within these two qubit subspaces of the qutrit, which natively yield the
continuous gates
−iφ
e 0 0 1 0 0
Z 01 (φ) = 0 1 0 , Z 12 (φ) = 0 1 0 (S16)
0 0 1 0 0 eiφ
computational basis, however, this obscures the fact that all density operators have unit trace. We thus use the
operators
X
Dj = −j |jihj| + |kihk| (S17)
0≤k<j
for 1 ≤ j < d, together with the identity operator to obtain the Gell-Mann basis. Here the Gell-Mann matrices λi
plus the identity I3 span SU(3) and are a natural choice for qutrit Pauli transfer matrices (PTMs). For convenience,
we index these elements for a single qutrit as follows:
1 0 0 0 1 0 0 −i 0 1 0 0 0 0 1
I3 = 0 1 0 , λ1 = 1 0 0 , λ2 = i 0 0 , λ3 = 0 −1 0 , λ4 = 0 0 0 ,
0 0 1 0 0 0 0 0 0 0 0 0 1 0 0
(S18)
0 0 −i 0 0 0 0 0 0 1 1 0 0
λ5 = 0 0 0 , λ6 = 0 0 1 , λ7 = 0 0 −i , λ8 = √ 0 1 0
i 0 0 0 1 0 0 i 0 3 0 0 −2
Having defined the Gell-Mann basis, we now define the Heisenberg-Weyl operators, which are a unitary generalization
of the familiar single-qubit Pauli operators to higher dimensional spaces (qudits) that enable the Clifford group to be
generalized. Let d be a positive integer and Zd = {0, . . . , d−1} denote the set of integers modulo d. The generalizations
of the X and Z operators to qudits are
X
X= |j ⊕d 1ihj|
j∈Zd
X
2πi
Z= exp j |jihj| , (S19)
d
j∈Zd
where ⊕d denotes addition modulo d. The Weyl basis for Cd×d is the set
Wd = {Wxz = X x Z z : x, z ∈ Zd } , (S20)
tr W † V = dδW,V ∀ W, V ∈ Wd . (S21)
Let n be a positive integer and D = dn . Then we define the n-qudit Weyl basis to be the set Wd,n = W⊗n d , which
is a trace-orthogonal basis for CD×D . Note that the Weyl basis is not a proper group as it is not closed under
multiplication. However, every element of the closure of the Weyl basis is proportional to an element of the Weyl
basis up to an overall phase. This overall phase vanishes in all cases as we only consider the adjoint action of Weyl
operators and so we treat the Weyl basis as a projective group. The n-qutrit Clifford group is then defined to be the
normalizer of the extended Weyl group EWd,n = U(1)Wd,n (which is a proper group), that is,
We now define the assumption of time-dependent Markovian noise that we use to analyze the results of our charac-
terization protocols. At a high level, we assume that each operation applied to the system corresponds to a linear map
that is independent of what other maps have been applied but may depend on the number of operations that have
been applied since the system was initialized. We allow this time dependence primarily to facilitate post-processing
of time-stationary noise processes wherein the physical process does not depend on the number of operations that
have been applied since the system was initialized but we are performing a weighted average over the operation that
is applied in the jth time step.
Formally, for a vector space V let V∗ denote the dual of V and L(V) denote the set of linear maps from V to itself.
Then we assume the following.
1. The state space of a physical implementation of an n-qudit system is some fixed vector space V.
2. Preparing the system in the state ρ corresponds to setting the state of the quantum system to some Θρ ∈ V.
3. Applying some unitary operation U to the system in the jth time step after preparing the system in a state
corresponds to applying some linear map, Θ(j, U ) ∈ L(V) to the state of the system.
4. The expectation value of an observable Q is obtained by applying some fixed ΘQ ∈ V∗ . (Note that we will ignore
finite measurement statistics for now.)
With some abuse of notation, we refer to the functions Θj : N × U(dn ) → L(V → V) and the vectors Θρ and ΘQ
together as the implementation map Θ [3].
The assumption of time-dependent Markovian noise allows hidden Markovianity as the implementation map can
include a coupling to an environment. This hidden Markovianity is frequently referred to as non-Markovianity in
the quantum information community. However, we allow it in the general setting because the additional assumptions
(such as positivity and complete positivity) are more cumbersome to define and typically are only helpful in the final
steps of an analysis. Indeed, it is conceptually useful to include post-processing steps that depend only on one time
step into an “effective” implementation map.
An ideal isolated implementation is one wherein there is a linear isomorphism between V and CD×D . For concrete-
ness, we define an isomorphism |∗iiB : CD×D → V relative to a trace-orthonormal basis B ⊂ CD×D as follows. As B
is a trace-orthonormal basis, we can write any A ∈ CD×D in terms of B as
X
A= tr B † A B. (S23)
B∈B
Therefore we can choose {|BiiB : B ∈ B} to be an orthonormal basis of V and extend it to an isomorphism by defining
X
|AiiB = tr B † A |BiiB . (S24)
B∈B
We will typically suppress the subscript B as it will be clear from the context (either the normalized Weyl basis or the
normalized Gell-Mann basis). To avoid having to define normalized versions of the bases explicitly, we define
√
B̂ = B/ tr B † B. (S25)
Moreover, defining hhA| = |Aii† and hhA|Bii = hhA||Bii and using the fact that B is a trace-orthonormal basis and
{|Bii : B ∈ B} is an orthonormal basis, we have
X
tr A† T = tr A† B tr C † T tr B † C
B,C∈B
X
= tr A† B tr B † T
B∈B
= hhA|T ii. (S26)
7
With the above isomorphism, we can define the ideal implementation of a unitary operator U ∈ U(D) to be
X
φB (U ) = |U BU † iihhB|, (S27)
B∈B
As above, we will also suppress the B on φ when the basis is clear from the context. Moreover, as U BU † is a trace-
orthonormal basis, one can readily verify that φ(U )φ(V ) = φ(U V ) for all U, V ∈ U(D), that is, φ is a representation
of U(D).
8
a b c
d e f
Figure S2. Cross entropy benchmarking (XEB) for qutrit, a, For the XEB circuits, the entropy difference H(p, q)−H(p, u)
is plotted against the ideal H(p, p) − H(p, u) at each cycle depth, with 30 randomizations at each depth. The linear fit gives
the fidelity at that particular depth. b, We show the exponential decay of the depolarized CZ† fidelity obtained from the linear
fits in a. c, We plot an example of the speckle purity decay as a function of cycle depth for a representative set of four states in
the two-qutrit Hilbert space. The probabilities of measuring the given tritstring are shown as a function of cycle depth across
all 30 randomizations. The bright “speckle” pattern characteristic of the Porter-Thomas distribution at low circuit depths is
smoothed out at larger depths. d-f, We show that the distribution of tritstrings transitions from the Porter-Thomas ditribution
at low depth, to a uniform distribution at deeper depths. The CDF is emphasized for a representative tritstring, |21i, while the
CDFs of the other tritstrings are shown in grey to show typical variation of the CDF across different states.
As discussed in the main text, performing randomized benchmarking (RB) on a two qutrit gate is prohibitively
expensive [2]. We utilize cross entropy benchmarking (XEB) as a second SPAM (state preparation and measurement)
free benchmarking protocol to corroborate the fidelity obtained from cycle benchmarking (CB). Recently XEB played
a central role in the quantum supremacy experiment in [4] and was used to benchmark the non-Clifford, three qubit
iToffoli gate in [5]. XEB theory has been discussed in detail in previous works so we present only a brief review of the
method before outlining more explicitly how it can be used to benchmark the two-qutrit CZ gate demonstrated here.
For XEB purposes a convenient definition of a random quantum circuit (RQC) is a quantum circuit randomly
selected from an ensemble of circuits such that the distribution of probabilities (across the ensemble of circuits) of
observing a particular ditstring, x, follows the Porter-Thomas distribution (for all possible ditstrings x) [6]. In the
present case of quantum circuits involving two qutrits we consider the probabilities of tritstrings 00, 01, . . . , 21,
22. After averaging over an ensemble of random circuits errors are tailored to be purely depolarizing so an error
corresponds to the circuit outputting a fully mixed state. In a mixed state the tritstring distribution is uniform
with each outcome xi having equal probability of 1/3n for an n qutrit system. Thus, intuitively the circuit fidelity
(under this error model) can be thought of as the deviation of the measured tritstring distribution from the uniform
distribution. The XEB fidelity makes this relationship precise as we now outline.
Denote the possible tritstrings xi for i = 1, . . . , 3n and let p(xi ) be the ideal tritstring distribution for the output
of a particular quantum circuit. Then q(xi ) is the measured distribution. The linear cross entropy of two probability
9
where the sum runs over the full support of the probability distributions and the self entropy is written compactly as
H(p1 ) ≡ H(p1 , p1 ). Under the depolarizing error model it can be shown straightforwardly that the circuit fidelity is
H(p, q) − H(p, u)
FXEB = (S30)
H(p, p) − H(p, u)
where u(xi ) is the uniform probability distribution [4]. This is ultimately the difference in the ideal to measured and
ideal to uniform cross entropies, normalized by the difference if the measured distribution were to perfectly match the
ideal distribution (p(xi ) = q(xi ) for all i).
We follow the XEB protocol for benchmarking gates/cycles outlined in Ref. [4]. This consists of measuring two
qutrit circuits of varying cycle depths M where cycle i consists of two single qutrit gates Gi,j where i = 1, . . . , M
labels the cycle and j = 1, 2 labels the qutrit, followed by the two qutrit CZ gate† (see Figure S2a). The single qutrit
gates are randomly selected unitaries from SU (3) and decomposed according to the decomposition given in [7]. For
each cycle depth M we generate N random circuits to be run and compute the ideal probability distributions of the
output tritstrings for each circuit. The XEB fidelity at cycle depth M , FXEB,M is determined by performing a least
squares fit of the linear relationship between H(pi , qi )−H(pi , u) and H(pi , pi )−H(pi , u) with i = 1, . . . , N labeling the
random circuit at the given depth (see Figure). The cycle infidelity, cycle , is estimated from the exponential decay of
FXEB,M as a function of cycle depth M . The error rate extracted in this way agrees well with CB (see Benchmarking
section in the main text).
Next we validate our implementation of XEB by verifying that the distribution of probabilities of a given tritstring
across the ensemble of random circuits does indeed approach the Porter-Thomas distribution,
where again D is the Hilbert space dimension, D = 32 for two qutrits and the approximate equality holds in the limit
of large Hilbert space dimension, D 1. Next we consider the evolution of this distribution as a function of cycle
depth. At short to intermediate cycle depths, the circuit is sufficiently random and the distribution approaches Porter-
Thomas. At larger cycle depths, the distribution begins to converge to P(p) → δ(p−1/D) since the depolarizing errors
dominate and the tritstring distributions approach the uniform distribution for all circuits. We plot our experimental
observation of this behavior in Fig. S2d-f.
The method of Speckle Purity Benchmarking (SPB) is based on this observation. Denoting the state purity as γ, it
can be estimated at depth M from the raw results of the XEB protocol outlined above by the relationship
D2 (D + 1)
γ(M ) = Var(pM ) (S32)
(D − 1)
where pM is the set of measured probabilities of a given tritstring x across the N random circuits at cycle depth M
[4]. Thus, once we have demonstrated that the distribution does indeed converge to the Porter-Thomas distribution
we can estimate the decay of the state purity from just the variance of the distribution, without the need for state
tomography which requires an exponential number of measurements, 3n(d−1) where n is the number of qudits and
d the dimension of each qudit. For our present case of two qutrits we would need to perform 81 measurements per
circuit to determine the state purity by state tomography.
10
&