This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Quantum spherical codes

Shubham P. Jain Joint Center for Quantum Information and Computer Science, NIST/University of Maryland, College Park, Maryland 20742, USA    Joseph T. Iosue Joint Center for Quantum Information and Computer Science, NIST/University of Maryland, College Park, Maryland 20742, USA Joint Quantum Institute, NIST/University of Maryland, College Park, Maryland 20742, USA    Alexander Barg Department of ECE and ISR, University of Maryland, College Park, Maryland 20742, USA    Victor V. Albert Joint Center for Quantum Information and Computer Science, NIST/University of Maryland, College Park, Maryland 20742, USA
Abstract

We introduce a framework for constructing quantum codes defined on spheres by recasting such codes as quantum analogues of the classical spherical codes. We apply this framework to bosonic coding, obtaining multimode extensions of the cat codes that can outperform previous constructions while requiring a similar type of overhead. Our polytope-based cat codes consist of sets of points with large separation that at the same time form averaging sets known as spherical designs. We also recast concatenations of CSS codes with cat codes as quantum spherical codes, revealing a new way to autonomously protect against dephasing noise.

Bosonic (a.k.a. oscillator) codes [Uncaptioned image] offer alternative qubit blueprints that are compatible with continuous-variable (CV) quantum platforms [1, 2, 3, 4, 5, 6, 7, 8, 9, 10] and that can reduce overhead by offering an extra layer of protection [11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25]. Qubits defined on a few bosonic modes or more exotic spaces [26] are likely to prove useful as control of quantum systems improves, but the field remains relatively unexplored [27, 28] in part because structures and intuition from qubit-based coding theory need not apply.

We develop a framework that yields generalizations of a class of bosonic codes called the cat codes [Uncaptioned image] [29, 30] and unifies such codes with several others. Our key observation is that all such codes are particular instances of quantum versions of spherical codes [Uncaptioned image] [31, 32], a family well known in classical coding theory. We overview the framework and demonstrate its utility with several new multimode cat codes. A rigorous study of general features is left to a companion follow-up work.

General codes on the sphere

Codewords of qubit codes [Uncaptioned image] are quantum superpositions of bit-strings. By analogy, we start with a spherical code, which is a set, or constellation, of points on the unit sphere. To construct a quantum spherical code, or QSC, we take a collection {𝒞k}k=0K1superscriptsubscriptsubscript𝒞𝑘𝑘0𝐾1\{\mathcal{C}_{k}\}_{k=0}^{K-1} of logical constellations, each of which gives rise to a codeword of the QSC obtained by taking a quantum superposition of all points 𝐱𝒞k𝐱subscript𝒞𝑘\mathbf{x}\in\mathcal{C}_{k}. We consider uniform superpositions here, leaving more general codes to future work. Taken together, the logical constellations yield the code constellation, 𝒞=k=0K1𝒞k𝒞superscriptsubscript𝑘0𝐾1subscript𝒞𝑘\mathcal{C}=\bigcup_{k=0}^{K-1}\mathcal{C}_{k}.

In the electromagnetic setting, spherical codes protect classical information against signal fluctuations during transmission, which correspond to small shifts acting on points in the constellation. A code’s ability to protect against such errors can be quantified by the minimum (squared) Euclidean distance dEsubscript𝑑𝐸d_{E} between any pair of distinct points. QSCs naturally inherit dEsubscript𝑑𝐸d_{E} as a figure of merit for protecting against such “bit-flip” noise.

Since QSCs store quantum information, they also suffer from “phase” noise, which comes from, e.g., fiber attenuation. Such noise can be expressed in terms of “potential-energy” functions on the sphere whose evaluation can be used to distinguish logical constellations (cf. [26, Sec. VI.B; 33]). If the average of a function over points in a constellation 𝒞ksubscript𝒞𝑘{\cal C}_{k} depends on k𝑘k, then the function’s underlying physical process causes an undetectable “phase” error.

Refer to caption
Figure 1: Quantum spherical codewords are quantum superpositions of constellations on a sphere. Logical constellations can form the vertices of a polytope and unite to form a code polytope compound. Projections of polytope compounds are shown for the (𝐚)𝐚\mathbf{(a)} cat, (𝐛)𝐛\mathbf{(b)} Möbius-Kantor, and (𝐜)𝐜\mathbf{(c)} Hessian quantum spherical codes, with logical constellation points colored either green, red, or purple.

An (n,K,dE)𝑛𝐾subscript𝑑𝐸(n,K,d_{E}) spherical code contains K𝐾K points on the n𝑛n-dimensional unit sphere such that the squared Euclidean distance between any two points is at least dEsubscript𝑑𝐸d_{E}. An ((n,K,dE,))𝑛𝐾subscript𝑑𝐸(\!(n,K,d_{E},\dots)\!) QSC is a K𝐾K-dimensional subspace of a quantum system’s vector space whose states are labeled by points on an n𝑛n-dimensional (real or complex) unit sphere, and whose protection against rotations is quantified by dEsubscript𝑑𝐸d_{E}. Protection against “phase” noise is designated by the proxy “\dots” because the notion of a “phase-flip” distance depends on the physical system embedding the QSC.

In principle, the above framework applies to any quantum state space parameterized by points on a sphere. CV systems [34, 35] admit several such spaces, and there exist examples of QSCs expressed using ordinary [29, 30, 36], squeezed [37, 38, 39], and pair-coherent states [40]. Collective atomic systems described by spin-coherent states as well as rotational state spaces of diatomic molecules also admit QSCs, namely, various large-spin codes [Uncaptioned image] [41] and diatomic molecular codes [Uncaptioned image] [26, Sec. VI], respectively. We focus on coherent-state QSCs because such codes naturally generalize the cat codes, and error-correction procedures for these new multimode cat codes require a similar type of overhead as what has already been realized [1, 2, 3, 4, 5]. We note that the discussion below can be modified to apply to other manifestations of QSCs.

Coherent-state formalism

A single-mode coherent state is a quantum representation of a standing wave of a fixed-frequency signal. An n𝑛n-mode coherent state |𝜶ket𝜶|\bm{\alpha}\rangle is parameterized by a complex n𝑛n-dimensional point 𝜶𝜶\bm{\alpha}. The point’s norm 𝜶2superscriptnorm𝜶2\|\bm{\alpha}\|^{2} corresponds to the state’s energy, and points of all states with a fixed energy n¯¯n\bar{\textsc{n}} form a complex n𝑛n-sphere, Ωn={𝜶n,𝜶2=n¯}subscriptΩ𝑛formulae-sequence𝜶superscript𝑛superscriptnorm𝜶2¯n\Omega_{n}=\{\bm{\alpha}\in\mathbb{C}^{n}\,,\,\|\bm{\alpha}\|^{2}=\textnormal{$\bar{\textsc{n}}$}\}.

Coherent-state QSCs consist of disjoint logical constellations 𝒞ksubscript𝒞𝑘\mathcal{C}_{k} of |𝒞k|subscript𝒞𝑘|\mathcal{C}_{k}| points picked from the n𝑛n-sphere and superimposed to form logical codewords,

|𝒞k1|𝒞k|𝜶𝒞k|¯n𝜶,similar-toketsubscript𝒞𝑘1subscript𝒞𝑘subscript𝜶subscript𝒞𝑘ket¯n𝜶|\mathcal{C}_{k}\rangle\sim\frac{1}{\sqrt{|{\cal C}_{k}|}}\sum_{\bm{\alpha}\in\mathcal{C}_{k}}|\sqrt{\textnormal{{\hbox{\bar{{n}}}}}}\bm{\alpha}\rangle~{}, (1)

where we restrict logical constellations to lie on the unit n𝑛n-sphere and delegate the overall scaling of the sphere’s radius to n¯¯n\bar{\textsc{n}}. An example to keep in mind is the four-component cat code defined by 𝒞0={(1),(1)}subscript𝒞011\mathcal{C}_{0}=\{(1),(-1)\} and 𝒞1={(i),(i)}=i𝒞0subscript𝒞1iiisubscript𝒞0\mathcal{C}_{1}=\{(\mathrm{i}),(-\mathrm{i})\}=\mathrm{i}\mathcal{C}_{0}.

The normalization in Eq. (1) is only valid asymptotically as n¯¯n\textnormal{$\bar{\textsc{n}}$}\to\infty because coherent states are not quite orthogonal due to the uncertainty principle,

|𝜶|𝜷|2=exp(𝜶𝜷2)exp(¯ndE).superscriptinner-product𝜶𝜷2superscriptnorm𝜶𝜷2¯nsubscript𝑑𝐸|\langle\bm{\alpha}|\bm{\beta}\rangle|^{2}=\exp\left(-\|\bm{\alpha}-\bm{\beta}\|^{2}\right)\leq\exp\left(-\textnormal{{\hbox{\bar{{n}}}}}d_{E}\right)~{}. (2)

The above “quantum corrections” for two coherent states of a code are suppressed exponentially with the energy n¯¯n\bar{\textsc{n}} and the minimum distance between two points in the code’s constellation 𝒞=k𝒞k𝒞subscript𝑘subscript𝒞𝑘\mathcal{C}=\bigcup_{k}\mathcal{C}_{k},

dE=min𝜶𝜷𝒞𝜶𝜷2.subscript𝑑𝐸subscript𝜶𝜷𝒞superscriptnorm𝜶𝜷2d_{E}=\min_{\bm{\alpha}\neq\bm{\beta}\in\mathcal{C}}\|\bm{\alpha}-\bm{\beta}\|^{2}~{}. (3)

Since dEsubscript𝑑𝐸d_{E} sets the scale of resolution of the constellation points, we refer to it as the resolution from now on.

Coherent states are subjected to two essentially different types of distortion — angular dephasing due to fluctuations in a mode’s frequency and changes in the mode’s excitations [42, Sec. II.A]. These induce “bit” and “phase” noise on QSCs, respectively. The corresponding relevant noise operators are passive linear-optical transformations and products of modal ladder operators {aj,aj}j=1nsuperscriptsubscriptsubscript𝑎𝑗superscriptsubscript𝑎𝑗𝑗1𝑛\{a_{j},a_{j}^{\dagger}\}_{j=1}^{n}, whose commutator is [aj,a]=δjsubscript𝑎𝑗superscriptsubscript𝑎subscript𝛿𝑗[a_{j},a_{\ell}^{\dagger}]=\delta_{j\ell}. Products of transformations and ladder operators can be used to express any physical noise channel [43, Eq. (39)].

Transformations on n𝑛n modes are parameterized by the unitary group 𝖴(n)𝖴𝑛\mathsf{U}(n) [35, Sec. 5.1.2]. A transformation U𝑹subscript𝑈𝑹U_{\bm{R}} corresponding to the n𝑛n-dimensional unitary matrix 𝑹𝑹\bm{R} rotates a coherent state |𝜶ket𝜶|\bm{\alpha}\rangle into |𝑹𝜶ket𝑹𝜶|\bm{R}\bm{\alpha}\rangle. If the rotation satisfies 𝑹𝜶𝜶2<dEsuperscriptnorm𝑹𝜶𝜶2subscript𝑑𝐸\|\bm{R}\bm{\alpha}-\bm{\alpha}\|^{2}<d_{E}, the transformation is detectable in the n¯¯n\textnormal{$\bar{\textsc{n}}$}\to\infty limit. Codes with larger resolution protect against larger sets of transformations.

A general ladder error,

L𝒑,𝒒(𝒂,𝒂)=j=1najpjajqj,subscript𝐿𝒑𝒒superscript𝒂𝒂superscriptsubscriptproduct𝑗1𝑛superscriptsubscript𝑎𝑗absentsubscript𝑝𝑗superscriptsubscript𝑎𝑗subscript𝑞𝑗L_{\bm{p},\bm{q}}(\bm{a}^{\dagger},\bm{a})=\prod_{j=1}^{n}a_{j}^{\dagger p_{j}}a_{j}^{q_{j}}~{}, (4)

is a monomial in the operators (a1,a2,,an)=𝒂subscript𝑎1subscript𝑎2subscript𝑎𝑛𝒂(a_{1},a_{2},\cdots,a_{n})=\bm{a} and their adjoints. It is parameterized by non-negative integer vectors 𝒑=(p1,p2,,pn)𝒑subscript𝑝1subscript𝑝2subscript𝑝𝑛\bm{p}=(p_{1},p_{2},\cdots,p_{n}) and 𝒒=(q1,q2,,qn)𝒒subscript𝑞1subscript𝑞2subscript𝑞𝑛\bm{q}=(q_{1},q_{2},\cdots,q_{n}) quantifying how many energy carriers (e.g., photons or phonons) are gained and lost in each mode, respectively.

Lowering operators ajsubscript𝑎𝑗a_{j} are “diagonal” in the coherent-state basis, satisfying aj|𝜶=αj|𝜶subscript𝑎𝑗ket𝜶subscript𝛼𝑗ket𝜶a_{j}|\bm{\alpha}\rangle=\alpha_{j}|\bm{\alpha}\rangle, where αjsubscript𝛼𝑗\alpha_{j} is the j𝑗jth component of 𝜶𝜶\bm{\alpha}. This “diagonality” relation and its adjoint imply that the expectation value of a ladder error over the k𝑘kth codeword (1) reduces to the average of the operator’s corresponding monomial over 𝒞ksubscript𝒞𝑘\mathcal{C}_{k},

𝒞k|L𝒑,𝒒(𝒂,𝒂)|𝒞k¯n|𝒑+𝒒|/2|𝒞k|𝜶𝒞kL𝒑,𝒒(𝜶,𝜶),similar-toquantum-operator-productsubscript𝒞𝑘subscript𝐿𝒑𝒒superscript𝒂𝒂subscript𝒞𝑘superscript¯n𝒑𝒒2subscript𝒞𝑘subscript𝜶subscript𝒞𝑘subscript𝐿𝒑𝒒superscript𝜶𝜶\langle{\mathcal{C}}_{k}|L_{\bm{p},\bm{q}}(\bm{a}^{\dagger},\bm{a})|{\mathcal{C}}_{k}\rangle\sim\frac{\textnormal{{\hbox{\bar{{n}}}}}^{|\bm{p}+\bm{q}|/2}}{|\mathcal{C}_{k}|}\sum_{\bm{\alpha}\in{\mathcal{C}}_{k}}L_{\bm{p},\bm{q}}(\bm{\alpha}^{\star},\bm{\alpha})~{}, (5)

where the one-norm |𝒑+𝒒|𝒑𝒒|\bm{p}+\bm{q}| is the degree of L𝒑,𝒒(𝜶,𝜶)subscript𝐿𝒑𝒒superscript𝜶𝜶L_{\bm{p},\bm{q}}(\bm{\alpha}^{\star},\bm{\alpha}). A ladder error can be detected whenever the above average is independent of k𝑘k [44].

Polytope QSCs

We have found numerous QSCs whose constellations form vertices of real [45] or complex [46, 47] polytopes [Uncaptioned image]. Polytope vertices are both sufficiently well-separated and uniform, providing protection against both types of noise. Code and polytope tables for the two cases are in Appxs. A and B, respectively.

We characterize ladder-error protection of polytope QSCs with three “distances”: dsubscript𝑑d_{\downarrow}, tsubscript𝑡t_{\downarrow}, and dsubscript𝑑d_{\updownarrow}. The first is the number of detectable losses (plus one), signifying that any pure-loss ladder error L𝒑=𝟎,𝒒subscript𝐿𝒑0𝒒L_{\bm{p}=\bm{0},\bm{q}} with |𝒒|<d𝒒subscript𝑑|\bm{q}|<d_{\downarrow} is detectable. Similarly, tsubscript𝑡t_{\downarrow} is the number of correctable losses (plus one), signifying that any ladder error with |𝒑|,|𝒒|<t𝒑𝒒subscript𝑡|\bm{p}|,|\bm{q}|<t_{\downarrow} is detectable. The degree distance dsubscript𝑑d_{\updownarrow} signifies that the code detects ladder errors with degree |𝒑+𝒒|<d𝒑𝒒subscript𝑑|\bm{p}+\bm{q}|<d_{\updownarrow}. These three parameters satisfy

(d+1)/2tddsubscript𝑑12subscript𝑡subscript𝑑subscript𝑑\lfloor(d_{\updownarrow}+1)/2\rfloor\leq t_{\downarrow}\leq d_{\updownarrow}\leq d_{\downarrow} (6)

and can vary quite significantly.

Our notation for an n𝑛n-mode polytope QSC with K𝐾K logical codewords is ((n,K,dE,d))𝑛𝐾subscript𝑑𝐸subscript𝑑(\!(n,K,d_{E},d_{\updownarrow})\!) or, more generally, ((n,K,dE,t,d,d))𝑛𝐾subscript𝑑𝐸subscript𝑡subscript𝑑subscript𝑑(\!(n,K,d_{E},\langle t_{\downarrow},d_{\updownarrow},d_{\downarrow}\rangle)\!). The four-component cat code is a ((1,2,2.0,2,2,2))122.0222(\!(1,2,2.0,\langle 2,2,2\rangle)\!) QSC, detecting d1=1subscript𝑑11d_{\downarrow}-1=1 loss error while sporting the relatively high resolution of 2.02.02.0. Since it can detect one gain simultaneously with one loss, this code also corrects t1=1subscript𝑡11t_{\downarrow}-1=1 loss error.

Each logical constellation of the four-component cat code is a line segment, and the code constellation forms the vertices of a square. More generally, logical constellations of the 2p2𝑝2p-component ((1,2,4sin2π2p,p,p,p))124superscript2𝜋2𝑝𝑝𝑝𝑝(\!(1,2,4\sin^{2}\frac{\pi}{2p},\langle p,p,p\rangle)\!) cat code are two p𝑝p-gons whose vertices interleave for maximal resolution. There is a tradeoff between loss protection and resolution, with the latter of order O(1/p2)𝑂1superscript𝑝2O(1/p^{2}) for a large number p1𝑝1p-1 of correctable losses. Utilizing higher dimensions, we pick other complex polytopes that maintain the same resolution while offering increased loss protection over the cat codes.

A simple code straddling the p=2,3𝑝23p=2,3 cat codes in terms of performance is the ((2,2,1.5,2,3,3))221.5233(\!(2,2,1.5,\langle 2,3,3\rangle)\!) simplex code,

𝒞0={12(ωμ,ω2μ)|μ𝟧}=𝒞1,subscript𝒞0conditional-set12superscript𝜔𝜇superscript𝜔2𝜇𝜇subscript5subscript𝒞1\mathcal{C}_{0}=\left\{\textstyle{\frac{1}{\sqrt{2}}}(\omega^{\mu},\omega^{2\mu})\,|\,\mu\in\mathsf{\mathbb{Z}_{5}}\right\}=-\mathcal{C}_{1}~{}, (7)

where ω=ei2π5𝜔superscript𝑒𝑖2𝜋5\omega=e^{i\frac{2\pi}{5}}. This code admits a lower resolution than the p=2𝑝2p=2 cat code, but detects one more loss in any of the two modes. Equivalently, it admits a higher resolution than the p=3𝑝3p=3 cat code’s resolution of unity, but corrects one fewer loss. Simplices exist in any dimension, yielding the infinite ((n,2,21/n,3))𝑛221𝑛3(\!(n,2,2-1/n,3)\!) QSC family that approaches the resolution of the p=2𝑝2p=2 cat code with increasing n𝑛n while detecting one more loss in any mode.

The Möbius-Kantor ((2,3,1.0,3,4,4))231.0344(\!(2,3,1.0,\langle 3,4,4\rangle)\!) code maintains the resolution of the p=3𝑝3p=3 cat code while adding one more logical state and detecting one more loss. Each of its three logical constellations form the 8 vertices of a Möbius-Kantor polygon (3{3}33333\{3\}3 in Coxeter notation; see Appx. B), and such polygons combine to form the 24 vertices of a 3{4}33433\{4\}3 polygon. This code corrects one more loss than the 𝟤𝖳2𝖳\mathsf{2T}-qutrit [Uncaptioned image] [36], a ((2,3,1.0,2,4,4))231.0244(\!(2,3,1.0,\langle 2,4,4\rangle)\!) QSC whose logical constellations each make up the 8 vertices of a complex octagon 2{4}42442\{4\}4. These two codes differ despite the fact that both code constellations map to the vertices of the same real 4D polytope via the mapping (x+iy,z+iw)(x,y,z,w)𝑥𝑖𝑦𝑧𝑖𝑤𝑥𝑦𝑧𝑤(x+iy,z+iw)\to(x,y,z,w), demonstrating subtleties in using real polytopes to define complex QSCs.

Logical constellations of the powerful ((3,2,1.0,4,5,9))321.0459(\!(3,2,1.0,\langle 4,5,9\rangle)\!) Hessian code consist of the 27 vertices of a Hessian polytope,

𝒞0={12(ημ,ην,0)perms.|μ,ν𝟥}=𝒞1,subscript𝒞0conditional-set12superscript𝜂𝜇superscript𝜂𝜈0perms.𝜇𝜈subscript3subscript𝒞1\mathcal{C}_{0}=\left\{\textstyle{\frac{1}{\sqrt{2}}}(\eta^{\mu},-\eta^{\nu},0)\cup\text{perms.}\,|\,\mu,\nu\in\mathsf{\mathbb{Z}_{3}}\right\}=-\mathcal{C}_{1}~{}, (8)

where η=ei2π3𝜂superscript𝑒𝑖2𝜋3\eta=e^{i\frac{2\pi}{3}}, and “perms.” is shorthand for the two cyclic permutations of the vector to the left for each μ,ν𝜇𝜈\mu,\nu. This code corrects as many losses as the p=4𝑝4p=4 cat code, but has the resolution of the p=3𝑝3p=3 cat code. Moreover, it can detect up to 8 losses, a feature available only to the p9𝑝9p\geq 9 cat codes.

There is a ((2,2,22,5,6,12))22225612(\!(2,2,2-\sqrt{2},\langle 5,6,12\rangle)\!) code that maintains the same resolution as the p=4𝑝4p=4 cat code, but corrects one more and detects 8 more losses. Its logical constellations each form the 24 vertices of a 4{3}44344\{3\}4 polygon, combining into a 48-vertex 2{6}42642\{6\}4 polygon.

An overachieving cousin of the above code is the ((4,2,22,6,8,12))42226812(\!(4,2,2-\sqrt{2},\langle 6,8,12\rangle)\!) Witting code, consisting of two Witting polytopes with 240 vertices each. This code corrects as many losses as a p=6𝑝6p=6 cat code, has the resolution of a p=4𝑝4p=4 cat code, and detects up to 11 losses. It is the first member of the infinite ((2r,2,22,8))superscript2𝑟2228(\!(2^{r},2,2-\sqrt{2},8)\!) family of codes that are based on orbits of the real Clifford group [Uncaptioned image] [48, 49, 50, 51].

A lower bound on dsubscript𝑑d_{\updownarrow} for Clifford, simplex, or other QSCs can be obtained whenever their logical constellations form designs [52]. A constellation 𝒞ksubscript𝒞𝑘\mathcal{C}_{k} is a complex spherical design [53, 54] of strength τ𝜏\tau if averages of monomials L𝒑,𝒒subscript𝐿𝒑𝒒L_{\bm{p},\bm{q}} of total degree |𝒑+𝒒|τ𝒑𝒒𝜏|\bm{p}+\bm{q}|\leq\tau over 𝒞ksubscript𝒞𝑘\mathcal{C}_{k} (1) are equal to those over the entire unit n𝑛n-sphere,

1|𝒞k|𝜶𝒞kL𝒑,𝒒(𝜶,𝜶)=Ωn𝑑𝜶L𝒑,𝒒(𝜶,𝜶).1subscript𝒞𝑘subscript𝜶subscript𝒞𝑘subscript𝐿𝒑𝒒superscript𝜶𝜶subscriptsubscriptΩ𝑛differential-d𝜶subscript𝐿𝒑𝒒superscript𝜶𝜶\frac{1}{|\mathcal{C}_{k}|}\sum_{\bm{\alpha}\in\mathcal{C}_{k}}L_{\bm{p},\bm{q}}(\bm{\alpha}^{\star},\bm{\alpha})=\int_{\Omega_{n}}d\bm{\alpha}L_{\bm{p},\bm{q}}(\bm{\alpha}^{\star},\bm{\alpha})~{}. (9)

Design strength is preserved under unitary rotations 𝑹𝑹\bm{R}, so logical constellations 𝒞k=𝑹k𝒞0subscript𝒞𝑘subscript𝑹𝑘subscript𝒞0\mathcal{C}_{k}=\bm{R}_{k}\mathcal{C}_{0} consisting of rotated versions of a complex spherical τ𝜏\tau-design 𝒞0subscript𝒞0\mathcal{C}_{0} yield a QSC whose degree distance is at least τ+1𝜏1\tau+1. In this way, construction of good QSCs can be accomplished by finding well-separated spherical designs 𝒞0subscript𝒞0\mathcal{C}_{0} of high strength coupled with a choice of rotations {𝑹k}k=0K1superscriptsubscriptsubscript𝑹𝑘𝑘0𝐾1\{\bm{R}_{k}\}_{k=0}^{K-1} (with 𝑹0subscript𝑹0\bm{R}_{0} the identity) that permits to control the resolution dEsubscript𝑑𝐸d_{E} of the code constellation k𝑹k𝒞0subscript𝑘subscript𝑹𝑘subscript𝒞0\bigcup_{k}\bm{R}_{k}\mathcal{C}_{0} while achieving high logical dimension K𝐾K.

Corroborating our parameter-based analysis, we numerically compare the performance of multi- and single-mode codes using the channel fidelity [55, 56, 57, 58, 59]. We observe that, for quKit encodings (for K>2𝐾2K>2), even simple multi-mode constellations, such as the simplex (7), are able to utilize the extra dimensions efficiently and outperform single-mode constellations over a range of loss rates (see Appx. D). The more non-trivial K=6𝐾6K=6 Möbius-Kantor 2{8}3absent283\subset 2\{8\}3 encoding (see Table B.1) consistently outperforms various combinations of cat codes for a wide range of energies and noise parameters.

CSS-based QSCs

Concatenations of CSS codes [Uncaptioned image] [60, 61, 62] with the two-component cat code [Uncaptioned image] [29], 𝒞0={(+1)}=𝒞1subscript𝒞01subscript𝒞1\mathcal{C}_{0}=\{(+1)\}=-\mathcal{C}_{1}, can also be interpreted as QSCs, albeit with a weight-based notion of ladder-error protection. Such codes are actively studied [11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24, 25], but have so far been interpreted in the framework of the outer qubit code and not in terms of underlying modal degrees of freedom. Our interpretation parallels a standard way to construct (classical) spherical codes by mapping binary codes to the (real) sphere [31, Sec. 2.5; 32, Sec. 1.2].

A [[n,k,(dX,dZ)]]delimited-[]𝑛𝑘subscript𝑑𝑋subscript𝑑𝑍[[n,k,(d_{X},d_{Z})]] qubit CSS code is constructed from two binary linear codes with distances dXsubscript𝑑𝑋d_{X} and dZsubscript𝑑𝑍d_{Z}, guaranteeing detection of Pauli X𝑋X-type and Z𝑍Z-type errors with weights less than the distances, respectively. Its codewords are equal superpositions of multi-qubit states labeled by binary strings. Concatenation is equivalent to mapping each binary string into a point on the n𝑛n-sphere via the coordinate-wise antipodal mapping 0+1010\to+1 and 11111\to-1. This yields an ((n,2k,dE=4dX/n,w=dZ))(\!(n,2^{k},d_{E}=4d_{X}/n,w_{\updownarrow}=d_{Z})\!) QSC that detects all errors L𝒑,𝒒subscript𝐿𝒑𝒒L_{\bm{p},\bm{q}} with Hamming weight Δ(𝒑+𝒒)<wΔ𝒑𝒒subscript𝑤\Delta(\bm{p}+\bm{q})<w_{\updownarrow} (see Appx. C). Asymptotically good qubit CSS codes thus yield QSCs whose distances dE,wsubscript𝑑𝐸subscript𝑤d_{E},w_{\updownarrow} are both separated from 0 as n𝑛n\to\infty.

X-type gates & stabilizers

Rotations on the n𝑛n-sphere provide groups of X𝑋X-type logical gates and stabilizers for QSCs. Elements of a logical group 𝖦𝖦\mathsf{G} permute logical constellations. Elements of a stabilizer subgroup 𝖧𝖦𝖧𝖦\mathsf{H}\subset\mathsf{G} permute points within each constellation, thereby leaving codewords invariant. Rotations are realized by passive linear-optical transformations using [35, Eq. (3.24)]. Rotation-based gates are noise-bias preserving [63] in that they do not convert rotations into losses.

For cat codes with 2p2𝑝2p components, 𝒞0={(ζ2j)|j𝗉}=ζ𝒞1subscript𝒞0conditional-setsuperscript𝜁2𝑗𝑗subscript𝗉𝜁subscript𝒞1\mathcal{C}_{0}=\{(\zeta^{2j})\,|\,j\in\mathsf{\mathbb{Z}_{p}}\}=\zeta\mathcal{C}_{1} with ζ=eiπp𝜁superscript𝑒i𝜋𝑝\zeta=e^{\mathrm{i}\frac{\pi}{p}}, the 1D rotation ζ𝜁\zeta permutes the two constellations, while powers of ζ2superscript𝜁2\zeta^{2} leave each constellation invariant. These rotations generate 𝖧=𝗉𝖦=𝟤𝗉𝖧subscript𝗉𝖦subscript2𝗉\mathsf{H}=\mathsf{\mathbb{Z}_{p}}\subset\mathsf{G}=\mathsf{\mathbb{Z}_{2p}} and are realized by transformations ζaasuperscript𝜁superscript𝑎𝑎\zeta^{a^{\dagger}a} and ζ2aasuperscript𝜁2superscript𝑎𝑎\zeta^{2a^{\dagger}a}.

Simplex constellations (7) can be permuted with the (1001)1001-\left(\begin{smallmatrix}1&0\\ 0&1\end{smallmatrix}\right) rotation and are invariant under powers of ω(100ω)𝜔100𝜔\omega\left(\begin{smallmatrix}1&0\\ 0&\omega\end{smallmatrix}\right), corresponding to the groups 𝟧𝟧×𝟤subscript5subscript5subscript2\mathsf{\mathbb{Z}_{5}}\subset\mathsf{\mathbb{Z}_{5}}\times\mathsf{\mathbb{Z}_{2}}, respectively. The latter group is generated by the two-mode transformations (1)a1a1+a2a2superscript1superscriptsubscript𝑎1subscript𝑎1superscriptsubscript𝑎2subscript𝑎2(-1)^{a_{1}^{\dagger}a_{1}+a_{2}^{\dagger}a_{2}} and ωa1a1+2a2a2superscript𝜔superscriptsubscript𝑎1subscript𝑎12superscriptsubscript𝑎2subscript𝑎2\omega^{a_{1}^{\dagger}a_{1}+2a_{2}^{\dagger}a_{2}}.

A stabilizer group for the Hessian code (8) is 𝖧𝖾𝟥=η,X,Zsubscript𝖧𝖾3𝜂𝑋𝑍\mathsf{He_{3}}=\langle\eta,X,Z\rangle, the 27-element qutrit Pauli/𝖧𝖾𝖧𝖾\mathsf{He}isenberg group consisting of powers of η𝜂\eta and the X,Z𝑋𝑍X,Z qutrit Pauli matrices. Appending by the logical-X𝑋X rotation I𝐼-I, where I𝐼I is the 3-by-3 identity, yields the logical group 𝖧𝖾𝟥×𝟤subscript𝖧𝖾3subscript2\mathsf{He_{3}\times\mathbb{Z}_{2}}. These groups are realized by phase-shifters and SWAP gates. Larger 𝖧𝖦𝖧𝖦\mathsf{H}\subset\mathsf{G} can be picked using the fact that all constellations form polytopes. The largest such groups are the 648-element and 1296-element symmetry groups of the corresponding Hessian and double-Hessian polytopes, respectively. These offer other ways to implement the logical-X𝑋X Pauli gate, but do not yield other gates.

Qudit QSCs offer larger logical-gate groups. The two groups are 𝟤𝟤𝖨subscript22𝖨\mathsf{\mathbb{Z}_{2}}\subset\mathsf{2I} for the 24-cell ((2,5,0.382,4,6,8))250.382468(\!(2,5,0.382,\langle 4,6,8\rangle)\!) real polytope code, with the former generated by the 5-by-5 matrix I𝐼-I, and the latter the binary icosahedral group 𝟤𝖨2𝖨\mathsf{2I}. Since the stabilizer group acts trivially, the logical group acts on the 5 codewords as a 5D permutation representation of the icosahedral group 𝖨=𝟤𝖨/𝟤𝖨2𝖨subscript2\mathsf{I}=\mathsf{2I/\mathbb{Z}_{2}}.

CSS-based QSCs inherit logical-X𝑋X stabilizers (gates) by mapping each X𝑋X-type stabilizer (logical Pauli) to a transversal linear-optical transformation via the component-wise mapping σx(1)aasubscript𝜎𝑥superscript1superscript𝑎𝑎\sigma_{x}\to(-1)^{a^{\dagger}a}. For example, the σx4superscriptsubscript𝜎𝑥tensor-productabsent4\sigma_{x}^{\otimes 4} stabilizer of the [[4,2,2]]delimited-[]422[[4,2,2]] code [Uncaptioned image] is mapped to the joint parity j=14(1)ajajsuperscriptsubscripttensor-product𝑗14superscript1superscriptsubscript𝑎𝑗subscript𝑎𝑗\bigotimes_{j=1}^{4}(-1)^{a_{j}^{\dagger}a_{j}}.

Z-type gates & stabilizers

The Z𝑍Z-type “stabilizer” for 2p2𝑝2p-component cat codes is F(a)=a2pn¯p𝐹𝑎superscript𝑎2𝑝superscript¯n𝑝F(a)=a^{2p}-\textnormal{$\bar{\textsc{n}}$}^{p}, which annihilates each point in the dilated code constellation n¯𝒞¯n𝒞\sqrt{\textnormal{$\bar{\textsc{n}}$}}\mathcal{C}. The corresponding polynomial F(α)𝐹𝛼F(\alpha) can be thought of as a potential on the sphere that is minimized only at the code-constellation points [64].

Polytope QSCs can require multiple polynomials to be stabilized. Simplex codes (7) are stabilized by F1=a12a24n¯3subscript𝐹1superscriptsubscript𝑎12superscriptsubscript𝑎24superscript¯n3F_{1}=a_{1}^{2}a_{2}^{4}-\textnormal{$\bar{\textsc{n}}$}^{3} and F2=a13a2n¯2subscript𝐹2superscriptsubscript𝑎13subscript𝑎2superscript¯n2F_{2}=a_{1}^{3}a_{2}-\textnormal{$\bar{\textsc{n}}$}^{2}. Hessian codewords (8) are stabilized by the F1=a1a2a3subscript𝐹1subscript𝑎1subscript𝑎2subscript𝑎3F_{1}=a_{1}a_{2}a_{3}, F2=a13+a23+a33subscript𝐹2superscriptsubscript𝑎13superscriptsubscript𝑎23superscriptsubscript𝑎33F_{2}=a_{1}^{3}+a_{2}^{3}+a_{3}^{3}, and F3=a16+a26+a36n¯3/4subscript𝐹3superscriptsubscript𝑎16superscriptsubscript𝑎26superscriptsubscript𝑎36superscript¯n34F_{3}=a_{1}^{6}+a_{2}^{6}+a_{3}^{6}-\textnormal{$\bar{\textsc{n}}$}^{3}/4. The degree of F1,2subscript𝐹12F_{1,2} is lower than the code’s degree distance (d=5subscript𝑑5d_{\updownarrow}=5) and detectable-loss distance (d=9subscript𝑑9d_{\downarrow}=9), unlike for the cat codes. This property makes this code similar to degenerate stabilizer codes [Uncaptioned image], i.e., codes whose check-operator weight is smaller than their distance.

Stabilizer polynomials commute with logical transformations U𝑹subscript𝑈𝑹U_{\bm{R}} for any 𝑹𝑹\bm{R} in the logical group and can be obtained by averaging ladder operators (4) over the symmetry group of the code constellation’s polytope.

Other polynomials act as logical gates on QSCs, evaluating to the same value for all points in 𝒞ksubscript𝒞𝑘\mathcal{C}_{k} in a way that depends on k𝑘k. For the cat codes, G=ap𝐺superscript𝑎𝑝G=a^{p} evaluates to ±n¯p/2plus-or-minussuperscript¯n𝑝2\pm\textnormal{$\bar{\textsc{n}}$}^{p/2} on the two codewords, respectively, yielding a logical-Z𝑍Z gate. The monomial G=a1a22𝐺subscript𝑎1superscriptsubscript𝑎22G=a_{1}a_{2}^{2} projects to a logical-Z𝑍Z gate within the simplex codespace. The smallest loss-only Z𝑍Z-gate of the Hessian code is G1=a13a26subscript𝐺1superscriptsubscript𝑎13superscriptsubscript𝑎26G_{1}=a_{1}^{3}a_{2}^{6} or its two cyclic permutations, and only a permutation-symmetric combination of all three operators commutes with the stabilizer group. A lower-degree monomial G2=a1a1a23subscript𝐺2subscriptsuperscript𝑎1subscript𝑎1superscriptsubscript𝑎23G_{2}=a^{\dagger}_{1}a_{1}a_{2}^{3} realizes another Z𝑍Z-gate with the help of gain operators. Combinations Gj+Gjsubscript𝐺𝑗superscriptsubscript𝐺𝑗G_{j}+G_{j}^{\dagger} generate logical Z𝑍Z-rotations within the F𝐹F-annihilated subspace [64], and have been realized for p=2𝑝2p=2 cat codes [4].

CSS-based QSCs inherit gates/stabilizers by mapping each Z𝑍Z-type gate/stabilizer to a monomial via the component-wise mapping σzasubscript𝜎𝑧𝑎\sigma_{z}\to a. For example, the [[4,2,2]]delimited-[]422[[4,2,2]] code’s σzσzIItensor-productsubscript𝜎𝑧subscript𝜎𝑧𝐼𝐼\sigma_{z}\otimes\sigma_{z}\otimes I\otimes I gate is mapped to a1a2subscript𝑎1subscript𝑎2a_{1}a_{2}. These codes also require stabilizers aj2n¯/nsuperscriptsubscript𝑎𝑗2¯n𝑛a_{j}^{2}-\textnormal{$\bar{\textsc{n}}$}/n on each mode j𝑗j in order to stabilize the inner cat-code constellation.

Correcting errors

Protection against rotation-based noise for 2p2𝑝2p-component cat codes is done passively using a Lindbladian whose jump operator is the Z𝑍Z-type stabilizer F𝐹F [64] and/or a Hamiltonian FFsuperscript𝐹𝐹F^{\dagger}F [65, 66]. Both techniques have been realized for p=2𝑝2p=2 [3, 5]. General QSCs admit the same type of passive protection but require several Fjsubscript𝐹𝑗F_{j}’s.

Microwave cavities coupled to superconducting circuits [67] provide a fertile ground for realizing such passive protection, and we outline how an existing superconducting circuit element called an “ATS” [68] can be tuned to realize the more complicated jump operators of several QSCs (see Appx. E). In particular, we show that a recent surface-cat concatenated-code proposal [17] can be readily modified with a Z𝑍Z-type surface-code stabilization scheme, thereby utilizing the full power of the code against Z𝑍Z-type noise in exclusively passive fashion.

Ladder errors (4) map the k𝑘kth codeword (1) into error states in span{|𝜶,𝜶𝒞k}spanket𝜶𝜶subscript𝒞𝑘\text{span}\{|\bm{\alpha}\rangle,\bm{\alpha}\in\mathcal{C}_{k}\}. The stabilizer group 𝖧𝖧\mathsf{H} splits up into several irreducible representations (irreps) acting on this span. Ladder-error protection is done by measuring syndromes associated with irreps and mapping back into the codespace. In order for correction to be possible, the stabilizer group has to be able to resolve all error spaces associated with a given error set.

The 444-component cat-code stabilizer is the parity (1)aasuperscript1superscript𝑎𝑎(-1)^{a^{\dagger}a}. Its eigenvalues correspond to the two irreps of 𝖧=𝖹𝟤𝖧subscript𝖹2\mathsf{H}=\mathsf{Z_{2}}, distinguishing between no error and a single loss a𝑎a. This technique [1] led to the first demonstration of break-even QEC using p=2𝑝2p=2 cat codes [2]. Similar multimode parities detect X𝑋X-errors for CSS-based QSCs.

For the simplex code (7), eigenvalues of the two-mode stabilizer ωa1a1+2a2a2superscript𝜔superscriptsubscript𝑎1subscript𝑎12superscriptsubscript𝑎2subscript𝑎2\omega^{a_{1}^{\dagger}a_{1}+2a_{2}^{\dagger}a_{2}} label the five irreps of 𝖹𝟧subscript𝖹5\mathsf{Z_{5}}. They allow for correction of {a1,a2,a1a2,a22}subscript𝑎1subscript𝑎2subscript𝑎1subscript𝑎2superscriptsubscript𝑎22\{a_{1},a_{2},a_{1}a_{2},a_{2}^{2}\}, falling short of correcting all two-mode losses due to a12superscriptsubscript𝑎12a_{1}^{2} not being simultaneously correctable with a2subscript𝑎2a_{2}.

For the Hessian code (8), the transformations realizing 𝖧𝖾𝟥subscript𝖧𝖾3\mathsf{He_{3}} can be measured to resolve the group’s 11 irreps. The general procedure for this and other non-Abelian codes resembles that of molecular codes [26, Sec. V.D].

Conclusion

We introduce a framework for constructing quantum analogs of the classical spherical codes, encapsulating several physically relevant quantum coding schemes for bosonic, spin, and molecular systems. We apply our framework to obtain multi-mode coherent-state codes based on polytopes, CSS codes, and classical codes. These QSCs outperform previous cat-code constructions [29, 30, 36] both in terms of code parameters and a numerical performance comparison. We show how passive protection of several instances of these QSCs can be realized in microwave cavities.

There are many other ways of constructing spherical codes, e.g., as group-orbit codes [Uncaptioned image] [69, 70, 71], as spherical embeddings of association schemes [32], through computer searches [72, 73], and many others [74, 31, 32, 75], as well as ways of constructing spherical designs [76, 77, 78]. As such, we anticipate that this work will pave the way for many novel, well-protected, and experimentally feasible logical qubits.

Data availability statement

Mathematica notebooks generated during the current study are available from the corresponding author on reasonable request. Further details and references about spherical codes described in this manuscript are available at the Error-correction Zoo website at http://errorcorrectionzoo.org and the corresponding repositories at http://github.com/errorcorrectionzoo.

Acknowledgements.
We thank Francesco Arzani, Ansgar Burchards, Jonathan Conrad, Aurélie Denys, Philippe Faist, Michael Gullans, Wenhao He, Liang Jiang, Greg Kuperberg, Gideon Lee, Anthony Leverrier, Pavel Panteleev, Shraddha Singh, and Guo (Jerry) Zheng for helpful discussions. VVA is especially grateful to Kyungjoo Noh for discussion about realizations of these codes. This work is supported in part by NSF QLCI grant OMA-2120757 and NSF grants CCF-2110113 (NSF-BSF) and CCF-2104489. JTI thanks the Joint Quantum Institute at the University of Maryland for support through a JQI fellowship. Our figures were drawn using Mathematica 13 following the prescription of Ref. [79]. Contributions to this work by NIST, an agency of the US government, are not subject to US copyright. Any mention of commercial products does not indicate endorsement by NIST. VVA thanks Ryhor Kandratsenia and Olga Albert for providing daycare support throughout this work.

Appendix A Real-polytope QSCs

logical constellation code constellation n𝑛n K𝐾K t,d,dsubscript𝑡subscript𝑑subscript𝑑\langle t_{\downarrow},d_{\updownarrow},d_{\downarrow}\rangle dEsubscript𝑑𝐸d_{E} related code
line segment 2K2𝐾2K-gon 111 K𝐾K 2,2,2222\langle 2,2,2\rangle 4sin2π2K4superscript2𝜋2𝐾4\sin^{2}\frac{\pi}{2K} two-component cat quK𝐾Kit
icosahedron 222 666 222 1.1061.1061.106
dodecahedron 222 101010 222 0.5090.5090.509
24-cell 222 121212 222 1.0001.0001.000 𝟤𝟤𝖳subscript22𝖳\mathsf{\mathbb{Z}_{2}}\subset\mathsf{2T} group-GKP
288-cell 222 242424 222 0.5860.5860.586 𝟤𝟤𝖮subscript22𝖮\mathsf{\mathbb{Z}_{2}}\subset\mathsf{2O} group-GKP
hyper-icosahedron 222 606060 222 0.3820.3820.382 𝟤𝟤𝖨subscript22𝖨\mathsf{\mathbb{Z}_{2}}\subset\mathsf{2I} group-GKP
hyper-dodecahedron 222 300300300 222 0.0730.0730.073
D𝐷D-orthoplex D/2𝐷2\left\lceil D/2\right\rceil D𝐷D 1,2,2122\langle 1,2,2\rangle 2.0002.0002.000 D=4𝐷4D=4: 𝟤𝖰subscript2𝖰\mathsf{\mathbb{Z}_{2}}\subset\mathsf{Q} group-GKP
D𝐷D-cube D/2𝐷2\left\lceil D/2\right\rceil 2D1superscript2𝐷12^{D-1} 222 4/D4𝐷4/D
p𝑝p-gon Kp𝐾𝑝Kp-gon 111 K𝐾K p,p,p𝑝𝑝𝑝\langle p,p,p\rangle 4sin2πKp4superscript2𝜋𝐾𝑝4\sin^{2}\frac{\pi}{Kp} p𝑝p-component cat quK𝐾Kit
tetrahedron dodecahedron 222 555 333 0.5090.5090.509
octahedron 5-octahedron 222 555 444 0.3820.3820.382
icosahedron 2-icosahedron 222 222 4,6,6466\langle 4,6,6\rangle 0.2110.2110.211
hyper-tetrahedron hyper-dodecahedron 222 120120120 333 0.0730.0730.073
hyper-octahedron 24-cell 222 333 2,4,4244\langle 2,4,4\rangle 1.0001.0001.000 𝖰𝟤𝖳𝖰2𝖳\mathsf{Q}\subset\mathsf{2T} group-GKP, 𝟤𝖳2𝖳\mathsf{2T}-qutrit
288-cell 222 666 2,4,4244\langle 2,4,4\rangle 0.5860.5860.586 𝖰𝟤𝖮𝖰2𝖮\mathsf{Q}\subset\mathsf{2O} group-GKP
hyper-icosahedron 222 151515 444 0.3820.3820.382 𝖰𝟤𝖨𝖰2𝖨\mathsf{Q}\subset\mathsf{2I} group-GKP
hyper-dodecahedron 222 757575 444 0.0730.0730.073
hyper-cube, -octahedron 24-cell 222 222 2,4,4244\langle 2,4,4\rangle 1.0001.0001.000
24-cell 288-cell 222 222 5,6,125612\langle 5,6,12\rangle 0.5860.5860.586 𝟤𝖳𝟤𝖮2𝖳2𝖮\mathsf{2T}\subset\mathsf{2O} group-GKP
hyper-icosahedron 222 555 4,6,8468\langle 4,6,8\rangle 0.3820.3820.382 𝟤𝖳𝟤𝖨2𝖳2𝖨\mathsf{2T}\subset\mathsf{2I} group-GKP
hyper-dodecahedron 222 252525 666 0.0730.0730.073
hyper-icosahedron hyper-dodecahedron 222 555 121212 0.0730.0730.073
D𝐷D-simplex D𝐷D-bisimplex D/2𝐷2\left\lceil D/2\right\rceil 222 2,3,3233\langle 2,3,3\rangle 22/D22𝐷2-2/D
(2r1)superscript2𝑟1(2^{r}-1)-simplex (2r1)superscript2𝑟1(2^{r}-1)-cube 2r1superscript2𝑟12^{r-1} 22rr1superscript2superscript2𝑟𝑟12^{2^{r}-r-1} 333 4/(2r1)4superscript2𝑟14/(2^{r}-1) shortened Hadamard
D𝐷D-demicube D𝐷D-cube D/2𝐷2\left\lceil D/2\right\rceil 222 min(4,D)4𝐷\min(4,D) 4/D4𝐷4/D single parity-check
2rsuperscript2𝑟2^{r}-orthoplex 2rsuperscript2𝑟2^{r}-cube 2r1superscript2𝑟12^{r-1} 22rr1superscript2superscript2𝑟𝑟12^{2^{r}-r-1} 444 22rsuperscript22𝑟2^{2-r} augmented Hadamard
Table A.1: QSCs whose logical and code constellations both make up the vertices of a real polytope; D2𝐷2D\geq 2 corresponds to spatial dimension, and the parameter r2𝑟2r\geq 2.

Logical constellations 𝒞ksubscript𝒞𝑘\mathcal{C}_{k} of a real polytope QSC form the vertices of a real polytope. The figure that results from the union of all logical polytopes is called a polytope compound, and its vertices form the code constellation 𝒞𝒞\mathcal{C}. Polytope QSCs can thus be constructed from established polytope compounds.

Regular real polytope compounds have been classified in three [80] and four [45, 81] dimensions. We collect all QSCs whose logical and code constellations each form a single regular real polytope in Table A.1. We leave to future work QSCs made up of polytope compounds whose code constellation forms vertices of multiple regular polytopes [80][45, Table VII][81, Sec. 10] as well as recently discovered variations of compounds with the same parameters [81, Sec. 10]. We include a few QSCs constructed from notable non-regular polytopes. All polytopes used in our constructions are listed in Table A.2.

The first column of the table lists the polytope whose vertices make up the logical constellations 𝒞ksubscript𝒞𝑘\mathcal{C}_{k}. All 𝒞ksubscript𝒞𝑘\mathcal{C}_{k} make up the same polytope for every code, with the exception being the “hyper-cube, -octahedron” code, in which 𝒞0subscript𝒞0\mathcal{C}_{0} (𝒞1subscript𝒞1\mathcal{C}_{1}) makes up the vertices of a hyper-cube (hyper-octahedron).

Since the n𝑛n-sphere is complex while the polytopes are real, we have to embed the polytopes into the sphere. For even dimension D𝐷D, the standard method of doing this is via the mapping

D(x1,x2,,xD)(x1+ix2,x3+ix4,,xD1+ixD)D/2.containssuperscript𝐷subscript𝑥1subscript𝑥2subscript𝑥𝐷subscript𝑥1isubscript𝑥2subscript𝑥3isubscript𝑥4subscript𝑥𝐷1isubscript𝑥𝐷superscript𝐷2\mathbb{R}^{D}\ni(x_{1},x_{2},\cdots,x_{D})\to(x_{1}+\mathrm{i}x_{2},x_{3}+\mathrm{i}x_{4},\cdots,x_{D-1}+\mathrm{i}x_{D})\in\mathbb{C}^{D/2}~{}. (A1)

Other mappings can be obtained by permuting the real coordinates. For odd D𝐷D, one has to embed the polytope into D+1𝐷1D+1 dimensions and then apply a mapping like the one above. Convenient coordinates exist for polytopes embedded in higher dimensions, e.g., vertices of a D𝐷D-simplex have coordinates (1,1,1,D)D+1111𝐷superscript𝐷1(1,1,\cdots 1,-D)\in\mathbb{R}^{D+1} and permutations thereof [32, Sec. 1.5]. Mappings into higher-dimensional spaces can also be used, e.g., the 2p2𝑝2p-component cat-code constellation can be mapped into 𝒞={ζj𝜶,j𝟤𝗉}𝒞superscript𝜁𝑗𝜶𝑗subscript2𝗉\mathcal{C}=\{\zeta^{j}\bm{\alpha},j\in\mathsf{\mathbb{Z}_{2p}}\} for any n𝑛n-dimensional unit vector 𝜶𝜶\bm{\alpha}. If one prefers to use real-valued vertices, then Dsuperscript𝐷\mathbb{R}^{D} can be directly embedded into Dsuperscript𝐷\mathbb{C}^{D}.

The parameters t,dsubscript𝑡subscript𝑑t_{\downarrow},d_{\downarrow} can depend on which of the above mappings one uses; we calculate them numerically by evaluating Eq. (5) for a given 𝒞𝒞\mathcal{C}. A mapping-independent lower bound on the degree distance dsubscript𝑑d_{\updownarrow} can be obtained from the strength of the design formed by the logical polytopes. Real polytope vertices can form (real) spherical designs [52], which are convertible into complex spherical designs via [53, Lemma 3.3]. The design strengths τ𝜏\tau of D𝐷D-dimensional polytope vertices are listed in Table A.2, column 5, yielding dτ+1subscript𝑑𝜏1d_{\updownarrow}\geq\tau+1 for a code consisting of such polytopes. This bound appears to be tight for real polytopes and holds as long as the polytope formed by 𝒞𝒞\mathcal{C} is the same dimension as those formed by each 𝒞ksubscript𝒞𝑘\mathcal{C}_{k}. Otherwise, the logical polytopes will not share a common sphere on which their vertices form designs. An exception to this restriction is for 𝒞ksubscript𝒞𝑘\mathcal{C}_{k} that are 1D line segments and is due to the fact that any pair of segments shares a common circle. The degree distance of a QSC consisting of segments is thus at least two.

polytope dim Schlafli/Coxeter vertices design dEsubscript𝑑𝐸d_{E} dEsubscript𝑑𝐸d_{E} (numerical) reference
line segment 111 {}\{\phantom{2}\} 222 111 444 4.0004.0004.000
triangle 222 {3}3\{3\} 333 222 333 3.0003.0003.000
square 222 {4}4\{4\} 444 333 222 2.0002.0002.000
pentagon 222 {5}5\{5\} 555 444 5(1φ)51𝜑\sqrt{5}(1-\varphi) 1.3821.3821.382
\vdots \vdots \vdots \vdots \vdots \vdots \vdots
p𝑝p-gon 222 {p}𝑝\{p\} p𝑝p p1𝑝1p-1 4sin2πp4superscript2𝜋𝑝4\sin^{2}\frac{\pi}{p}
tetrahedron 333 {3,3}33\{3,3\} 444 222 8/3838/3 2.6672.6672.667
octahedron 333 {3,4}34\{3,4\} 666 333 222 2.0002.0002.000
cube 333 {4,3}43\{4,3\} 888 333 4/3434/3 1.3331.3331.333
icosahedron 333 {3,5}35\{3,5\} 121212 555 4/(1+φ2)41superscript𝜑24/(1+\varphi^{2}) 1.1061.1061.106
dodecahedron 333 {5,3}53\{5,3\} 202020 555 225/322532-2\sqrt{5}/3 0.5090.5090.509
2-icosahedron 333 β{3,4}𝛽34\mathit{\beta\{3,4\}} 242424 555 2(1φ)2/(1+φ2)2superscript1𝜑21superscript𝜑22(1-\varphi)^{2}/(1+\varphi^{2}) 0.21110.21110.2111 [80]
5-octahedron 333 [5{3,4}]2{3,5}delimited-[]534235\mathit{[5\{3,4\}]2\{3,5\}} 303030 555 (1φ)2superscript1𝜑2(1-\varphi)^{2} 0.3820.3820.382 [80]
hyper-tetrahedron 444 {3,3,3}333\{3,3,3\} 555 222 5/2525/2 2.5002.5002.500 [82]
hyper-octahedron 444 {3,3,4}334\{3,3,4\} 888 333 222 2.0002.0002.000 [82]
hyper-cube 444 {4,3,3}433\{4,3,3\} 161616 333 111 1.0001.0001.000 [82]
24-cell 444 {3,4,3}343\{3,4,3\} 242424 555 111 1.0001.0001.000 [Uncaptioned image] [82]
288-cell 444 o3m4m3o 484848 777 22222-\sqrt{2} 0.5860.5860.586 [Uncaptioned image] [83]
hyper-icosahedron 444 {3,3,5}335\{3,3,5\} 120120120 111111 (1φ)2superscript1𝜑2(1-\varphi)^{2} 0.3820.3820.382 [Uncaptioned image] [82]
hyper-dodecahedron 444 {5,3,3}533\{5,3,3\} 600600600 111111 (735)/47354(7-3\sqrt{5})/4 0.0730.0730.073 [Uncaptioned image] [82]
D𝐷D-simplex D𝐷D {3D1}superscript3𝐷1\{3^{D-1}\} D+1𝐷1D+1 222 2+2/D22𝐷2+2/D [Uncaptioned image]
D𝐷D-bisimplex D𝐷D [2{3D1}]delimited-[]2superscript3𝐷1\mathit{[2\{3^{D-1}\}]} 2(D+1)2𝐷12(D+1) 222 22/D22𝐷2-2/D [83]
D𝐷D-orthoplex D𝐷D {3D2,4}superscript3𝐷24\{3^{D-2},4\} 2D2𝐷2D 333 222 2.0002.0002.000 [Uncaptioned image]
D𝐷D-demicube D𝐷D {31,D3,1}superscript31𝐷31\mathit{\{3^{1,D-3,1}\}} 2D1superscript2𝐷12^{D-1} min(3,D1)3𝐷1\min(3,D-1) 8/D8𝐷8/D [83]
D𝐷D-cube D𝐷D {4,3D2}4superscript3𝐷2\{4,3^{D-2}\} 2Dsuperscript2𝐷2^{D} 333 4/D4𝐷4/D
Table A.2: Polytope data used to construct QSCs in Table A.1. Non-italicised polytopes make up the convex regular polytopes in real dimension D𝐷D. φ=1+52𝜑152\varphi=\frac{1+\sqrt{5}}{2} is the golden ratio.

Points on the real 4D sphere are in one-to-one correspondence with quaternions, which in turn parameterize the group 𝖲𝖴(𝟤)𝖲𝖴2\mathsf{SU(2)} [84]. Vertices of the hyper-octahedron, 24-cell, (disphenoidal) 288-cell, and hyper-icosahedron correspond to quaternions forming the quaternion 𝖰𝖰\mathsf{Q}, binary tetrahedral 𝟤𝖳2𝖳\mathsf{2T}, binary octahedral 𝟤𝖮2𝖮\mathsf{2O}, and binary icosahedral 𝟤𝖨2𝖨\mathsf{2I} subgroups, respectively. Polytope QSCs consisting of such polytopes thus are related to 𝖲𝖴(𝟤)𝖲𝖴2\mathsf{SU(2)} group-GKP codes [Uncaptioned image] [26]. The 𝟤𝖳2𝖳\mathsf{2T}-qutrit code [36] is similarly related to the 𝖰𝟤𝖳𝖲𝖴(𝟤)𝖰2𝖳𝖲𝖴2\mathsf{Q}\subset\mathsf{2T}\subset\mathsf{SU(2)} group-GKP code, but the idea of using groups this way is limited to two modes because spheres in higher dimensions no longer correspond to groups.

An [n,k]𝑛𝑘[n,k] binary linear [Uncaptioned image] code C𝐶C can be converted into a QSC by taking logical constellations to be cosets of C𝐶C in 𝔽2nsuperscriptsubscript𝔽2𝑛\mathbb{F}_{2}^{n} under the antipodal mapping. The table lists QSCs arising this way from the Hadamard [Uncaptioned image] and single parity-check [Uncaptioned image] codes. These codes all have non-trivial dsubscript𝑑d_{\updownarrow} because the cosets correspond to known polytope compounds when embedded into the sphere [47, pg. 287].

Appendix B Complex-polytope QSCs

logical const-n code const-n n𝑛n K𝐾K t,d,dsubscript𝑡subscript𝑑subscript𝑑\langle t_{\downarrow},d_{\updownarrow},d_{\downarrow}\rangle dEsubscript𝑑𝐸d_{E} related code
Möbius-Kantor 2{6}32632\{6\}3 222 222 3,4,6346\langle 3,4,6\rangle 0.8450.8450.845
3{4}33433\{4\}3 222 333 3,4,4344\langle 3,4,4\rangle 1.0001.0001.000 𝖰𝟤𝖳𝖰2𝖳\mathsf{Q}\subset\mathsf{2T} group-GKP
2{8}32832\{8\}3 222 666 3,4,4344\langle 3,4,4\rangle 0.3670.3670.367
(2,4)24(2,4)-orthoplex 4{3}44344\{3\}4 222 333 2,4,4244\langle 2,4,4\rangle 1.0001.0001.000 𝖰𝟤𝖳𝖰2𝖳\mathsf{Q}\subset\mathsf{2T} group-GKP, 𝟤𝖳2𝖳\mathsf{2T}-qutrit
3{6}23623\{6\}2 [23{6}2]delimited-[]2362[2~{}3\{6\}2] 222 222 4,4,4444\langle 4,4,4\rangle 0.2110.2110.211
4{3}44344\{3\}4 2{6}42642\{6\}4 222 222 5,6,125612\langle 5,6,12\rangle 0.5860.5860.586 𝟤𝖳𝟤𝖮2𝖳2𝖮\mathsf{2T}\subset\mathsf{2O} group-GKP
3{4}33433\{4\}3 2{8}32832\{8\}3 222 222 3,6,123612\langle 3,6,12\rangle 0.3670.3670.367
2{6}42642\{6\}4 [22{6}4]delimited-[]2264[2~{}2\{6\}4] 222 222 4,8,8488\langle 4,8,8\rangle 0.2680.2680.268
3{5}33533\{5\}3 2{10}321032\{10\}3 222 222 9,12,3091230\langle 9,12,30\rangle 0.1320.1320.132
5{3}55355\{3\}5 2{6}52652\{6\}5 222 222 11,12,30111230\langle 11,12,30\rangle 0.0980.0980.098
3{4}53453\{4\}5 222 333 11,12,20111220\langle 11,12,20\rangle 0.0440.0440.044
(3,3)33(3,3)-orthoplex rectified Hessian 333 888 2,3,3233\langle 2,3,3\rangle 1.0001.0001.000
(3,6)36(3,6)-orthoplex rectified Hessian 333 444 2,4,6246\langle 2,4,6\rangle 1.0001.0001.000
Hessian double Hessian 333 222 4,5,9459\langle 4,5,9\rangle 1.0001.0001.000
Witting double Witting 444 222 6,8,126812\langle 6,8,12\rangle 0.5860.5860.586 Clifford group-orbit
(1,m)1𝑚(1,m)-cube (n,m)𝑛𝑚(n,m)-cube n𝑛n mn1superscript𝑚𝑛1m^{n-1} 1,2,m12𝑚\langle 1,2,m\rangle 4nsin2πm4𝑛superscript2𝜋𝑚\frac{4}{n}\sin^{2}\frac{\pi}{m}
(1,m)1𝑚(1,m)-orthoplex (n,m)𝑛𝑚(n,m)-orthoplex n𝑛n n𝑛n 1,2,m12𝑚\langle 1,2,m\rangle min(2,4sin2πm)24superscript2𝜋𝑚\min(2,4\sin^{2}\frac{\pi}{m})
Table B.1: QSCs whose logical and code constellations both make up the vertices of a non-real complex polytope; n1𝑛1n\geq 1 corresponds to complex dimension. d=msubscript𝑑𝑚d_{\downarrow}=m for the (n,m)𝑛𝑚(n,m)-cube/orthoplex codes are conjectured based on numerical results.

Complex polytopes are polytopes whose vertices are complex. As with real polytopes, there are a myriad polygons in the two complex dimensions, a handful of special polytopes in a few of the higher dimensions, and only two infinite families of non-real complex polytopes present in any dimension.

The two families are straightforward complex generalizations of the cube and orthoplex, respectively. A simple set of vertices of a real D𝐷D-dimensional cube consists of 2Dsuperscript2𝐷2^{D} vectors with coordinates ±1plus-or-minus1\pm 1. The vertices of the complex (n,m)𝑛𝑚(n,m)-cube (a.k.a. γnmsuperscriptsubscript𝛾𝑛𝑚\gamma_{n}^{m}) consist of mnsuperscript𝑚𝑛m^{n} complex vectors of dimension n𝑛n with m𝑚mth roots of unity at each coordinate. A similar generalization holds for the (n,m)𝑛𝑚(n,m)-orthoplex (a.k.a. βnmsuperscriptsubscript𝛽𝑛𝑚\beta_{n}^{m}), whose mn𝑚𝑛mn coordinates are n𝑛n-dimensional vectors whose single nonzero entry is an m𝑚mth root of unity.

A union of complex polytopes sharing a common center forms a complex polytope compound. Complex compounds yield complex QSCs whose code constellations are formed by the vertices of the compound and whose logical constellations are formed by the vertices of the participating polytopes. Complex compounds have not been as thoroughly studied as their real counterparts, and most of our codes come from the handful of constructions from Refs. [85, 86, 47]. In Table B.1, we collect the complex polytope QSCs that are the most interesting for a comparative study with the real polytope codes. All the polytopes used in our constructions are listed in Table B.2.

Complex polygons yield several interesting QSCs not available in the real case. We mentioned already in the main text that multiple complex polytopes can reduce to the same real polytope when mapped into the reals. As another example, compounds consisting of 5{3}55355\{3\}5 polygonal code constellations have exceptional loss detection capabilities, with dsubscript𝑑d_{\downarrow} as high as 30, but suffer from low resolution. There are many more polygons, and we leave a more extensive list of complex polytope QSCs to a follow-up work.

Complex polytopes also offer interesting many-mode alternatives to cat codes. The tensor product of n𝑛n single-mode 444-component cat codes is an ((n,2n,2/n,2,2,2))𝑛superscript2𝑛2𝑛222(\!(n,2^{n},2/n,\langle 2,2,2\rangle)\!) QSC whose code constellation can be thought of as an (n,4)𝑛4(n,4)-cube, constructed as a Kronecker product of n𝑛n (1,4)14(1,4)-cubes. The resolution of this code decreases as order O(1/n)𝑂1𝑛O(1/n), meaning that a constant energy per mode (usually picked to be n¯/n2¯n𝑛2\textnormal{$\bar{\textsc{n}}$}/n\approx 2 [17, 2]) is required in order to be able to resolve codewords without substantial intrinsic memory error. On the other hand, the (n,4)𝑛4(n,4)-orthoplex ((n,n,2.0,1,2,4))𝑛𝑛2.0124(\!(n,n,2.0,\langle 1,2,4\rangle)\!) QSC, whose logical constellations are (1,4)14(1,4)-orthoplexes, maintains constant resolution and has extra loss detection at the expense of a linear increase in the codespace dimension and no loss correction. It is an interesting open problem to find a QSC with K=O(n)𝐾𝑂𝑛K=O(n) that can correct one or more losses.

polytope dim Schlafli/Coxeter vertices design dEsubscript𝑑𝐸d_{E} dEsubscript𝑑𝐸d_{E} (numerical) reference
Möbius-Kantor 222 3{3}33333\{3\}3 888 333 222 2.0002.0002.000 [46, 85]
222 2{6}32632\{6\}3 161616 333 22/32232-2/\sqrt{3} 0.8450.8450.845 [46, 85]
222 3{4}33433\{4\}3 242424 555 111 1.0001.0001.000 [46, 85]
222 4{3}44344\{3\}4 242424 555 111 1.0001.0001.000 [46, 85]
222 3{6}23623\{6\}2 242424 333 (33)/2332(3-\sqrt{3})/2 0.6340.6340.634 [46, 85]
222 2{6}42642\{6\}4 484848 777 22222-\sqrt{2} 0.5860.5860.586 [46, 85]
222 2{8}32832\{8\}3 484848 555 222/322232-2\sqrt{2/3} 0.3670.3670.367 [46, 85]
222 [23{6}2]delimited-[]2362\mathit{[2~{}3\{6\}2]} 484848 333 2(1φ)2/(1+φ2)2superscript1𝜑21superscript𝜑22(1-\varphi)^{2}/(1+\varphi^{2}) 0.2110.2110.211
222 [22{6}4]delimited-[]2264\mathit{[2~{}2\{6\}4]} 969696 777 23232-\sqrt{3} 0.2680.2680.268
222 3{5}33533\{5\}3 120120120 111111 (1φ)2superscript1𝜑2(1-\varphi)^{2} 0.3820.3820.382 [46, 85]
222 5{3}55355\{3\}5 120120120 111111 (1φ)2superscript1𝜑2(1-\varphi)^{2} 0.3820.3820.382 [46, 85]
222 2{10}321032\{10\}3 240240240 111111 22(3+5)/3223532-\sqrt{2(3+\sqrt{5})/3} 0.1320.1320.132 [46, 85]
222 2{6}52652\{6\}5 240240240 111111 2φ52𝜑52-\sqrt{\varphi\sqrt{5}} 0.0980.0980.098 [46, 85]
222 3{4}53453\{4\}5 360360360 111111 4sin2(π/30)4superscript2𝜋304\sin^{2}(\pi/30) 0.0440.0440.044 [46, 85]
Hessian 333 3{3}3{3}3333333\{3\}3\{3\}3 272727 444 3/2323/2 1.5001.5001.500 [Uncaptioned image] [46]
double Hessian 333 2{4}3{3}3243332\{4\}3\{3\}3 545454 444 111 1.0001.0001.000 [86]
rectified Hessian 333 3{3}3{4}2333423\{3\}3\{4\}2 727272 555 111 1.0001.0001.000 [Uncaptioned image] [47]
Witting 444 3{3}3{3}3{3}333333333\{3\}3\{3\}3\{3\}3 240240240 777 111 1.0001.0001.000 [Uncaptioned image] [46]
double Witting 444 [2 3{3}3{3}3{3}3] 480480480 777 22222-\sqrt{2} 0.5860.5860.586
(n,m)𝑛𝑚(n,m)-cube n𝑛n m{4}2{3}2{3}2𝑚423232m\{4\}2\{3\}\cdots 2\{3\}2 mnsuperscript𝑚𝑛m^{n} min(3,m1)3𝑚1\min(3,m-1) 4nsin2πm4𝑛superscript2𝜋𝑚\frac{4}{n}\sin^{2}\frac{\pi}{m} [46]
(n,m)𝑛𝑚(n,m)-orthoplex n𝑛n 2{3}2{3}2{4}m232324𝑚2\{3\}2\{3\}\cdots 2\{4\}m nm𝑛𝑚nm min(3,m1)3𝑚1\min(3,m-1) min(2,4sin2πm)24superscript2𝜋𝑚\min(2,4\sin^{2}\frac{\pi}{m}) [46]
Table B.2: Non-real polytope data used to construct QSCs in Table B.1. Italicised polytopes are not regular. φ=1+52𝜑152\varphi=\frac{1+\sqrt{5}}{2} is the golden ratio.

Appendix C CSS-based QSCs

The antipodal mapping converts binary strings 𝒃=(b1,b2,,bn)𝒃subscript𝑏1subscript𝑏2subscript𝑏𝑛\bm{b}=(b_{1},b_{2},\cdots,b_{n}) labeling n𝑛n-qubit states into n𝑛n-mode coherent states normalized to an energy of unity,

𝜶𝒃=((1)b1,(1)b2,,(1)bn)/n.subscript𝜶𝒃superscript1subscript𝑏1superscript1subscript𝑏2superscript1subscript𝑏𝑛𝑛\bm{\alpha}_{\bm{b}}=\left((-1)^{b_{1}},(-1)^{b_{2}},\cdots,(-1)^{b_{n}}\right)/\sqrt{n}~{}. (C1)

Using [87, Thm. 7.3], there exists a basis of codewords for an [[n,k,(dX,dZ)]]delimited-[]𝑛𝑘subscript𝑑𝑋subscript𝑑𝑍[[n,k,(d_{X},d_{Z})]] CSS code that is labeled by length-k𝑘k binary strings bold-ℓ\bm{\ell} and that is expressed in terms of 𝖢𝖹superscriptsubscript𝖢𝖹perpendicular-to\mathsf{C_{Z}^{\perp}}, the dual of one of the underlying binary linear codes. Applying the antipodal mapping to the bold-ℓ\bm{\ell}th element of such a basis yields a codeword for the corresponding QSC,

|¯1|𝖢𝖹|𝒄𝖢𝖹|¯n𝜶+𝒄.similar-toket¯bold-ℓ1superscriptsubscript𝖢𝖹perpendicular-tosubscript𝒄superscriptsubscript𝖢𝖹perpendicular-toket¯nsubscript𝜶bold-ℓ𝒄|\overline{\bm{\ell}}\rangle\sim\frac{1}{\sqrt{|\mathsf{C_{Z}^{\perp}}|}}\sum_{\bm{c}\in\mathsf{C_{Z}^{\perp}}}|\sqrt{\textnormal{{\hbox{\bar{{n}}}}}}~{}\bm{\alpha}_{\bm{\ell}+\bm{c}}\rangle~{}. (C2)

Phase-flip errors

Using Eq. (5), the projection of a general ladder error acting a subset of modes SSSS\SS into the QSC codespace is equivalent to a Z𝑍Z-type error,

L𝒑,𝒒(SS)=jSSajpjajqj(¯nn)|𝒑+𝒒|/2jSSZjpj+qj,superscriptsubscript𝐿𝒑𝒒SSsubscriptproduct𝑗SSsuperscriptsubscript𝑎𝑗absentsubscript𝑝𝑗superscriptsubscript𝑎𝑗subscript𝑞𝑗superscript¯n𝑛𝒑𝒒2subscriptproduct𝑗SSsuperscriptsubscript𝑍𝑗subscript𝑝𝑗subscript𝑞𝑗L_{\bm{p},\bm{q}}^{(\SS)}=\prod_{j\in\SS}a_{j}^{\dagger p_{j}}a_{j}^{q_{j}}\quad\quad\to\quad\quad\left(\frac{\textnormal{{\hbox{\bar{{n}}}}}}{n}\right)^{|\bm{p}+\bm{q}|/2}\prod_{j\in\SS}Z_{j}^{p_{j}+q_{j}}~{}, (C3)

where we define Zj|n¯𝜶𝒃=(1)bj|n¯𝜶𝒃subscript𝑍𝑗ket¯nsubscript𝜶𝒃superscript1subscript𝑏𝑗ket¯nsubscript𝜶𝒃Z_{j}|\sqrt{\textnormal{$\bar{\textsc{n}}$}}\bm{\alpha}_{\bm{b}}\rangle=(-1)^{b_{j}}|\sqrt{\textnormal{$\bar{\textsc{n}}$}}\bm{\alpha}_{\bm{b}}\rangle. As long as the support size of the region SSSS\SS is less than dZsubscript𝑑𝑍d_{Z}, the distance of 𝖢𝖹subscript𝖢𝖹\mathsf{C_{Z}}, the properties of CSS codes can be used to show that the above error is detectable. This means that any ladder error with Hamming weight Δ(𝒑+𝒒)<dZΔ𝒑𝒒subscript𝑑𝑍\Delta(\bm{p}+\bm{q})<d_{Z} is detectable.

Bit-flip errors

The squared Euclidean distance between two code constellation elements 𝜶𝒃subscript𝜶𝒃\bm{\alpha}_{\bm{b}} and 𝜶𝒄subscript𝜶𝒄\bm{\alpha}_{\bm{c}} can be expressed in terms of the Hamming distance Δ(𝒃,𝒄)Δ𝒃𝒄\Delta(\bm{b},\bm{c}) between their corresponding binary strings,

𝜶𝒃𝜶𝒄2superscriptnormsubscript𝜶𝒃subscript𝜶𝒄2\displaystyle\left\|\bm{\alpha}_{\bm{b}}-\bm{\alpha}_{\bm{c}}\right\|^{2} =22𝜶𝒃𝜶𝒄absent22subscript𝜶𝒃subscript𝜶𝒄\displaystyle=2-2\bm{\alpha}_{\bm{b}}\cdot\bm{\alpha}_{\bm{c}} (C4a)
=22nj=1n(1)bj+cjabsent22𝑛superscriptsubscript𝑗1𝑛superscript1subscript𝑏𝑗subscript𝑐𝑗\displaystyle=2-\frac{2}{n}\sum_{j=1}^{n}(-1)^{b_{j}+c_{j}} (C4b)
=22nj=1n+[nΔ(𝒃,𝒄)][Δ(𝒃,𝒄)]absent22𝑛superscriptsubscript𝑗1𝑛delimited-[]𝑛Δ𝒃𝒄delimited-[]Δ𝒃𝒄\displaystyle=2-\frac{2}{n}\sum_{j=1}^{n}+[n-\Delta(\bm{b},\bm{c})]-[\Delta(\bm{b},\bm{c})] (C4c)
=4Δ(𝒃,𝒄)/n.absent4Δ𝒃𝒄𝑛\displaystyle=4\Delta(\bm{b},\bm{c})/n~{}. (C4d)

This quantity is bounded by 4dX/n4subscript𝑑𝑋𝑛4d_{X}/n, where dXsubscript𝑑𝑋d_{X} is the distance of the other underlying binary linear code 𝖢𝖷subscript𝖢𝖷\mathsf{C_{X}}.

Appendix D Performance of quKit QSCs

cat simplex
K𝐾K dEsubscript𝑑𝐸d_{E} n¯=n¯/n¯n¯n𝑛\textnormal{$\bar{\textsc{n}}$}=\textnormal{$\bar{\textsc{n}}$}/n Fmaxsubscript𝐹maxF_{\text{max}} dEsubscript𝑑𝐸d_{E} n¯¯n\bar{\textsc{n}} n¯/n¯n𝑛\textnormal{$\bar{\textsc{n}}$}/n Fmaxsubscript𝐹maxF_{\text{max}}
222 222 1.69201.69201.6920 0.98220.98220.9822 1.51.51.5 3.07763.07763.0776 1.53881.53881.5388 0.98410.98410.9841
333 111 2.67682.67682.6768 0.96030.96030.9603 0.88200.88200.8820 5.39195.39195.3919 2.69602.69602.6960 0.96040.96040.9604
444 0.58580.58580.5858 3.64033.64033.6403 0.93180.93180.9318 0.88200.88200.8820 5.34425.34425.3442 2.67212.67212.6721 0.95520.95520.9552
555 0.38200.38200.3820 4.59934.59934.5993 0.89920.89920.8992 0.69090.69090.6909 5.18685.18685.1868 2.59342.59342.5934 0.95320.95320.9532
666 0.26800.26800.2680 5.56085.56085.5608 0.86420.86420.8642 0.41730.41730.4173 5.54965.54965.5496 2.77482.77482.7748 0.94120.94120.9412
(a) Sweet spot data and other code parameters are listed where K𝐾K is logical dimension, n𝑛n is the number of modes, n¯¯n\bar{\textsc{n}} the total energy required and Fmaxsubscript𝐹maxF_{\text{max}} the fidelity achieved at the code’s sweet spot.
Refer to caption
(b) Channel fidelity Fsubscript𝐹F_{\mathcal{E}} is plotted at each code’s respective sweet spot energies.
Figure D.1: Comparing cat (111-mode) and simplex (222-mode) quKit codes for varying values of K𝐾K, it is observed that the simplex family provides more pronounced advantages in code parameters and performance with growing logical dimension. The sweet-spot energy was calculated at the loss rate γ=0.095𝛾0.095\gamma=0.095.

Here, we present results of a numerical comparison of several new intrinsically multi-mode polytope constellations to single-mode and multi-mode instantiations of cat-code (i.e., polygon-based) constellations. We observe that multi-mode QSCs efficiently utilize the extra dimensions to store more logical information, all while consuming a comparable (in most cases, lower) energy per mode.

Our performance metric, as guided by [55, 56, 57, 58], is the channel fidelity

FΨ|ρ|Ψsubscript𝐹quantum-operator-productΨsubscript𝜌ΨF_{\mathcal{E}}\equiv\langle\Psi|\rho_{\mathcal{E}}|\Psi\rangle (D1)

where, for a qubit state, |Ψ=(|0A0B+|1A1B)/2ketΨketsubscript0𝐴subscript0𝐵ketsubscript1𝐴subscript1𝐵2|\Psi\rangle=(|0_{A}0_{B}\rangle+|1_{A}1_{B}\rangle)/\sqrt{2} is the maximally entangled state between the source qubit A𝐴A and the ancilla qubit B𝐵B. The outgoing density matrix, ρAB(|ΨΨ|)subscript𝜌tensor-productsubscript𝐴subscript𝐵ketΨbraΨ\rho_{\mathcal{E}}\equiv\mathcal{E}_{A}\otimes\mathcal{I}_{B}(|\Psi\rangle\langle\Psi|), is obtained by the action of the combined encoding-noise-recovery channel \mathcal{E} on the source qubit and identity \mathcal{I} on the ancilla. The channel fidelity Fsubscript𝐹F_{\mathcal{E}} is an intrinsic property of the channel which measures how well the entanglement between the information qubit and an ancillary system in preserved upon application of the channel \mathcal{E}. For more motivation behind our choice of metric, we refer the interested reader to [59, Appx. A].

The channel \mathcal{E} is considered to be the composition of the encoding, noise and recovery channels. We assume that noise occurs only via the pure-loss channel, described by Kraus operators [88]

E(γ1γ)/2a^!(1γ)n^/2,subscript𝐸superscript𝛾1𝛾2superscript^𝑎superscript1𝛾^𝑛2E_{\ell}\equiv\left(\frac{\gamma}{1-\gamma}\right)^{\ell/2}\frac{\hat{a}^{\ell}}{\sqrt{\ell!}}(1-\gamma)^{\hat{n}/2}~{}, (D2)

where 00\ell\geq 0 quantifies the amount of photons lost, and where γ𝛾\gamma is the loss rate. For a selected encoding and this error channel, we optimize the recovery to obtain the maximum Fsubscript𝐹F_{\mathcal{E}}. This optimization problem can be formulated as a semidefinite program [57], which we solve using the Python library CVXPY [89, 90].

The above technique can be adapted to bosonic codes by setting a maximum Fock-space cutoff (in order to make the underlying space finite-dimensional) [59]. We avoid such truncation by working in the coherent-state basis. In such a basis, the action of the pure loss channel can be expressed using a different set of Kraus operators whose cardinality and matrix dimension are equal to the size of the code constellation [36, Appx. A]. That way, we are constrained more by the size of the code constellation than the number of modes.

A K𝐾K-dimensional code is constructed by replicating a “base” logical constellation K𝐾K times while maintaining good resolution dEsubscript𝑑𝐸d_{E}. Cat codes use n𝑛n-gons as the base constellations, while simplex codes employ 𝒞0subscript𝒞0\mathcal{C}_{0} from Eq. (7).

The k𝑘kth logical constellation of a 222-gon quK𝐾Kit code with 0k<K0𝑘𝐾0\leq k<K is generated by multiplying the base line segment {1,1}11\{1,-1\} with eiπk/Ksuperscript𝑒i𝜋𝑘𝐾e^{\mathrm{i}\pi k/K}. The k𝑘kth logical constellation 𝑹k𝒞0subscript𝑹𝑘subscript𝒞0\bm{R}_{k}\mathcal{C}_{0} for the quK𝐾Kit simplex codes with 0k<K{2,3,4}0𝑘𝐾2340\leq k<K\in\{2,3,4\} is obtained by letting {𝑹0,𝑹1,𝑹2,𝑹3}={I,I,Z,Z}subscript𝑹0subscript𝑹1subscript𝑹2subscript𝑹3𝐼𝐼𝑍𝑍\{\bm{R}_{0},\bm{R}_{1},\bm{R}_{2},\bm{R}_{3}\}=\{I,-I,Z,-Z\}, where I𝐼I is the two-dimensional identity and Z𝑍Z is the Pauli-Z𝑍Z matrix. The K=5𝐾5K=5 (K=6𝐾6K=6) simplex constellations are generated using the unitary rotations {ωkI|k5}conditional-setsuperscript𝜔𝑘𝐼𝑘subscript5\{\omega^{k}I\,|\,k\in\mathbb{Z}_{5}\} ({e2kπi/6I|k6}conditional-setsuperscript𝑒2𝑘𝜋𝑖6𝐼𝑘subscript6\{e^{2k\pi i/6}I\,|\,k\in\mathbb{Z}_{6}\}).

Sweet-spot comparison

Given a loss rate γ𝛾\gamma, one can tune the energy of a given code to obtain the sweet spot energy value — the n¯¯n\bar{\textsc{n}} that gives the highest fidelity Fmaxsubscript𝐹maxF_{\text{max}}. For cat codes, it has been observed [91, 92, 59] that this sweet spot value is finite, and that it does not drastically change with small changes in the loss rate γ𝛾\gamma. We observe similar behavior in all the QSCs we examine.

We evaluate the performance of each code at its respective sweet spot in order to compare the highest possible performance of each code under a given loss rate. Figure 1(a) lists the code parameters of the simplex and 222-gon based cat quKit codes. The advantage of using simplex codes over the cat becomes pronounced for larger memories. As we scan the table in the figure, we see that the fidelity Fmaxsubscript𝐹maxF_{\text{max}} of simplex codes decreases slower with growing dimension K𝐾K compared to that of cat codes, meaning that simplex codes utilize the available phase space more effectively when packing more quantum information. This is corroborated by the simplex quK𝐾Kits maintaining higher resolution dEsubscript𝑑𝐸d_{E} for large K𝐾K.

The energy required per mode (n¯/n¯n𝑛\textnormal{$\bar{\textsc{n}}$}/n) for optimal simplex performance also increases at a slower rate than that of cat codes. Notably, for K=6𝐾6K=6, even the total energy (n¯¯n\bar{\textsc{n}}) needed by simplex codes is lower than that of the corresponding cat code. This trend is consistent in code performance, quantified by the channel fidelity, as shown in Fig. 1(b).

logical const-n code const-n tsubscript𝑡t_{\downarrow} dEsubscript𝑑𝐸d_{E} n¯¯n\bar{\textsc{n}} n¯/n¯n𝑛\textnormal{$\bar{\textsc{n}}$}/n Fmaxsubscript𝐹maxF_{\text{max}}
222-gon 121212-gon 111 0.2680.2680.268 5.56085.56085.5608 5.56085.56085.5608 0.86420.86420.8642
333-gon 181818-gon 222 0.1210.1210.121 8.95848.95848.9584 8.95848.95848.9584 0.88820.88820.8882
3-gon3-gontensor-product3-gon3-gon3\text{-gon}\otimes 3\text{-gon} 9-gon6-gontensor-product9-gon6-gon9\text{-gon}\otimes 6\text{-gon} 222 0.2340.2340.234 9.18019.18019.1801 4.59014.59014.5901 0.95850.95850.9585
Möbius-Kantor 2{8}32832\{8\}3 222 0.3670.3670.367 5.79925.79925.7992 2.89962.89962.8996 0.99010.99010.9901
Table D.1: Sweet spot data and other code parameters for quKit codes with K=6𝐾6K=6 are listed where n𝑛n is the number of modes, n¯¯n\bar{\textsc{n}} the total energy required and Fmaxsubscript𝐹maxF_{\text{max}} the fidelity achieved at the code’s ‘sweet spot’.
Refer to caption
(a) Channel fidelity Fsubscript𝐹F_{\mathcal{E}} is plotted at fixed loss rate γ=0.095𝛾0.095\gamma=0.095.
Refer to caption
(b) Channel fidelity Fsubscript𝐹F_{\mathcal{E}} is plotted at fixed total energy n¯=8.9584¯n8.9584\textnormal{$\bar{\textsc{n}}$}=8.9584.
Figure D.2: As shown in (a), the Möbius-Kantor code demonstrates a universal improvement over polygon based codes and outperforms them over a range of energies and loss rates, as exemplified in (b) by choosing n¯¯n\bar{\textsc{n}} corresponding to the sweet spot energy (at γ=0.095𝛾0.095\gamma=0.095) of the 3-gon18-gon3-gon18-gon3\text{-gon}\subset 18\text{-gon} code.

Overall advantage of a qudit encoding

We also compare overall performance of a multi-mode QSC to various cat-like codes by sweeping both energy and loss rate. We fix K=6𝐾6K=6 and construct codes out of various logical constellations: the 222-gon, 333-gon ({1,e2πi/3,e4πi/3}1superscript𝑒2𝜋𝑖3superscript𝑒4𝜋𝑖3\{1,e^{2\pi i/3},e^{4\pi i/3}\}), 3-gon3-gontensor-product3-gon3-gon3\text{-gon}\otimes 3\text{-gon}({(e2πim1/3,e2πim2/3)| 0m1,m22}conditional-setsuperscript𝑒2𝜋𝑖subscript𝑚13superscript𝑒2𝜋𝑖subscript𝑚23formulae-sequence 0subscript𝑚1subscript𝑚22\{(e^{2\pi im_{1}/3},e^{2\pi im_{2}/3})\,|\,0\leq m_{1},m_{2}\leq 2\}) and the Möbius-Kantor polygon. The first two are single-mode cat codes, the third distributes logical information over two modes using tensor products of single-mode cat codes, while the last is an intrinsically two-mode code.

Results from a numerical comparison in Fig. 2(a) show a universal advantage across the swept energy-and-loss-rate parameter space. Similar trends are observed for various other γ𝛾\gamma values (not shown here). Notably, the Möbius-Kantor surpasses other codes even at their optimal values, as exemplified in Fig. 2(b), where we choose the n¯¯n\bar{\textsc{n}} corresponding to the 333-gon \subset 181818-gon code’s sweet spot.

We append sweet-spot data for this set of codes in Tab. D.1, which corroborates the simplex-cat-code data in Fig. 1(a). We observe that the Möbius-Kantor code provides robust protection against up to 2 losses, boasts higher resolution (dEsubscript𝑑𝐸d_{E}), requires lower energy per mode, and consistently outperforms all the mentioned polygon based codes in the sweet-spot comparison.

When encoding a greater number of logical dimensions, multimodal QSCs prove significantly more resource efficient and clearly outperform cat codes.

Appendix E Lindbladian stabilization

The Z𝑍Z-type (i.e., dephasing or rotation error) correction for cat codes is done autonomously by engineering Lindbladians with a desired “correcting” jump operator F=κ(a2pn¯p)𝐹𝜅superscript𝑎2𝑝superscript¯n𝑝F=\kappa(a^{2p}-\textnormal{$\bar{\textsc{n}}$}^{p}) and correction rate κ𝜅\kappa. Engineering such terms is possible in microwave cavities coupled to superconducting circuits [67]. The typical scheme proceeds by coupling the physical system to an ancillary or buffer mode b𝑏b via the Hamiltonian coupling Fb+H.c.𝐹superscript𝑏H.c.Fb^{\dagger}+\text{H.c.}, setting the ancillary mode to have a high loss rate, and then showing that the effective Lindbladian acting on the physical a𝑎a-mode system has jump operator F𝐹F (see, e.g., Ref. [64]).

Multi-mode coherent-state QSCs require more jump operators, and each jump operator can now consist of multiple monomials in the lowering operators ajsubscript𝑎𝑗a_{j}. However, due to the flexibility provided by a recently developed circuit element called an asymmetrically threaded SQUID, or ATS [68], the above scheme can be extended to realize the more complex jumps required for QSCs. We sketch out a general scheme below and apply it to a CSS and a polytope QSC.

The cost of our basic scheme — one ancillary mode per jump — is only an upper bound. While more advanced schemes are outside the scope of this work, we note that a single ATS can be used to simulataneously realize multiple jumps using as little as one ancillary mode [17, Appx. B.2].

E.0.1 General scheme

A desired jump operator is a sum of monomials of some maximum degree and a potential constant term that is the P𝑃Pth power of n¯¯n\bar{\textsc{n}},

F=monomial(a_j)constant¯nP.𝐹monomial(a_j)constantsuperscript¯n𝑃F=\sum\text{monomial({\hbox{a_{j}}})}-\text{constant}\cdot\textnormal{{\hbox{\bar{{n}}}}}^{P}~{}. (E1)

Leveraging previous schemes [68, supplement][17, Appx. B.2], we describe a slightly more general scheme to implement a dissipator with this jump operator, which generates time evolution according to the equation of motion (E4).

Let the harmonic component of the j𝑗jth physical mode have frequency ωjsubscript𝜔𝑗\omega_{j}, while the ancilla b𝑏b-mode evolves at ωbsubscript𝜔𝑏\omega_{b}. In the rotating frame w.r.t. these components, the multi-mode density matrix ρabsubscript𝜌𝑎𝑏\rho_{ab} describing a set of modes coupled via an ATS evolves according to

ρ˙ab=i[Hdrive+HATS,ρ]+κb𝒟[b](ρ).subscript˙𝜌𝑎𝑏𝑖subscript𝐻drivesubscript𝐻ATS𝜌subscript𝜅𝑏𝒟delimited-[]𝑏𝜌\dot{\rho}_{ab}=-i[H_{\text{drive}}+H_{\text{ATS}},\rho]+\kappa_{b}{\cal D}[b](\rho)\,. (E2)

We describe each term and its purpose:

  1. 1.

    The drive term, Hdrive=n¯Pb+H.c.subscript𝐻drivesuperscript¯n𝑃𝑏H.c.H_{\text{drive}}=-\textnormal{$\bar{\textsc{n}}$}^{P}b+\text{H.c.}, will yield the constant part of the jump operator F𝐹F (E1) once the effective equation of motion on the physical modes is derived. This term can be set to zero if no constant term is necessary.

  2. 2.

    All of the magic comes from the ATS term [17, Eq. (B11)]111Up to constant prefactors, this term is also obtained from the last two terms in [68, Eq. (S2)] by setting φΔ=φ+π/2subscript𝜑Δ𝜑𝜋2\varphi_{\Delta}=\varphi+\pi/2, ΔEJ=0Δsubscript𝐸𝐽0\Delta E_{J}=0, and φΣ=ϵ(t)+π/2subscript𝜑Σitalic-ϵ𝑡𝜋2\varphi_{\Sigma}=\epsilon(t)+\pi/2, and expanding to first order in ϵitalic-ϵ\epsilon. The quadratic term in that equation can absorbed into the bare oscillator Hamiltonian prior going into the rotating frame by expanding φ𝜑\varphi in a set of normal modes.,

    HATS=ϵ(t)sin(φ+ϕbbeiωbt+jϕjajeiωjt+H.c.)with pump tonesϵ(t)=pξpeiΩpt+H.c.,formulae-sequencesubscript𝐻ATSitalic-ϵ𝑡𝜑subscriptitalic-ϕ𝑏𝑏superscript𝑒𝑖subscript𝜔𝑏𝑡subscript𝑗subscriptitalic-ϕ𝑗subscript𝑎𝑗superscript𝑒𝑖subscript𝜔𝑗𝑡H.c.with pump tonesitalic-ϵ𝑡subscript𝑝subscript𝜉𝑝superscript𝑒𝑖subscriptΩ𝑝𝑡H.c.H_{\text{ATS}}=\epsilon(t)\sin\left(\varphi+\phi_{b}be^{-i\omega_{b}t}+\mathord{\sum}_{j}\phi_{j}a_{j}e^{-i\omega_{j}t}+\text{H.c.}\right)\quad\text{with pump tones}\quad\epsilon(t)=\mathord{\sum}_{p}\xi_{p}e^{i\Omega_{p}t}+\text{H.c.}\,, (E3)

    which depends on static real parameters {ϕb,ϕj,φ}subscriptitalic-ϕ𝑏subscriptitalic-ϕ𝑗𝜑\{\phi_{b},\phi_{j},\varphi\} and tunable real parameters {ωb,ωj,ξp,Ωp}subscript𝜔𝑏subscript𝜔𝑗subscript𝜉𝑝subscriptΩ𝑝\{\omega_{b},\omega_{j},\xi_{p},\Omega_{p}\}. The static flux φ{0,π/2}𝜑0𝜋2\varphi\in\{0,\pi/2\} [68, Eq. (S3)] allows us to interpolate between a sine and cosine ATS term. One pump tone, with amplitude ξpsubscript𝜉𝑝\xi_{p} and frequency ΩpsubscriptΩ𝑝\Omega_{p}, is necessary for each monomial in the jump operator (E1). Tuning the frequency allows us to select the specific desired monomial, while tuning the drive allows us to tune the monomial’s coefficient.

  3. 3.

    The dissipative part, κb𝒟[b]subscript𝜅𝑏𝒟delimited-[]𝑏\kappa_{b}{\cal D}[b] for sufficiently large κb>0subscript𝜅𝑏0\kappa_{b}>0, ensures that the ancilla is sufficiently lossy. The steady-state space of this evolution is spanned by any state of the ajsubscript𝑎𝑗a_{j} modes, tensored with the vacuum Fock state |0ket0|0\rangle on the b𝑏b mode. Assuming the Hamiltonian terms to be perturbations to this strong Lindbladian evolution, one can then derive an effective equation of motion within this steady-state space using either second-order perturbation theory or what is colloquially known as “adiabatic elimination” [94, 95, 96].

Expanding the ATS term yields an infinite series, with combinatorially many monomials consisting of products of drive-tone terms {ξpeiΩpt}subscript𝜉𝑝superscript𝑒𝑖subscriptΩ𝑝𝑡\{\xi_{p}e^{i\Omega_{p}t}\}, physical mode operators {ajeiωjt}subscript𝑎𝑗superscript𝑒𝑖subscript𝜔𝑗𝑡\{a_{j}e^{-i\omega_{j}t}\}, the ancillary mode term beiωbt𝑏superscript𝑒𝑖subscript𝜔𝑏𝑡be^{-i\omega_{b}t}, and the flux bias φ𝜑\varphi. The expansion is approximated by truncating to an order such that the highest-degree term is one higher than the degree of the highest-order monomial in the desired jump operator (E1). The phase term φ{0,π/2}𝜑0𝜋2\varphi\in\{0,\pi/2\} ensures that the expansion contains the monomial of correct (even or odd) degree.

The pump-tone frequencies {Ωp}subscriptΩ𝑝\{\Omega_{p}\} are then tuned to particular linear combinations of {ωj,ωb}subscript𝜔𝑗subscript𝜔𝑏\{\omega_{j},\omega_{b}\} so that any of the monomials that are also present in the desired jump operator become time-independent. That way, all other terms can be treated as higher-order “fast-rotating” corrections in what is known as the “rotating-wave approximation”. Combining with the drive term, the Hamiltonian terms in Eq. (E3) are then approximated by Fb+H.c.𝐹superscript𝑏H.c.Fb^{\dagger}+\text{H.c.}. Verifying that the many remaining terms in the expansion are all time-dependent can be done using the algebraic manipulation plugin sneg [97, 98] in Mathematica.

The desired equation of the density matrix ρ=trmode b(ρab)𝜌subscripttrmode 𝑏subscript𝜌𝑎𝑏\rho=\text{tr}_{\text{mode }b}(\rho_{ab}) on the physical modes upon adiabatically eliminating the ancilla is then

ρ˙=κ𝒟[F](ρ)+,˙𝜌𝜅𝒟delimited-[]𝐹𝜌\dot{\rho}=\kappa{\cal D}[F](\rho)+\cdots\,, (E4)

for a to-be-determined correction rate κ𝜅\kappa, and up to higher-order corrections “\cdots” stemming from corrections to the approximations.

E.0.2 CSS QSCs

Our Z𝑍Z-type stabilization for CSS-type concatenated encodings provides an autonomous alternative to the discrete measurement of Z𝑍Z-type error syndromes. Such jump operators can be readily “plugged in” to any concatenated cat-CSS code, including a recent concatenated surface-cat code proposal [17].

Our jump operator for the surface-cat code example consists of a product of lowering operators acting on sides one through four of each plaquette, F=a1a2a3a4n¯2𝐹subscript𝑎1subscript𝑎2subscript𝑎3subscript𝑎4superscript¯n2F=a_{1}a_{2}a_{3}a_{4}-\textnormal{$\bar{\textsc{n}}$}^{2}, of a square lattice. This is a special case of the general form (E1) with one degree-four monomial and constant term with P=2𝑃2P=2. The sole monomial requires only one pump tone, with amplitude ξ1ξsubscript𝜉1𝜉\xi_{1}\equiv\xi and frequency Ω1ΩsubscriptΩ1Ω\Omega_{1}\equiv\Omega, and zero flux bias, φ=0𝜑0\varphi=0. The ATS sine term is expanded to fifth order. The condition selecting the desired monomial is

Ω1=ω1+ω2+ω3+ω4ωb,yielding the monomiala1a2a3a4b+H.c..subscriptΩ1subscript𝜔1subscript𝜔2subscript𝜔3subscript𝜔4subscript𝜔𝑏yielding the monomialsubscript𝑎1subscript𝑎2subscript𝑎3subscript𝑎4superscript𝑏H.c.\Omega_{1}=\omega_{1}+\omega_{2}+\omega_{3}+\omega_{4}-\omega_{b}\,,\quad\quad\text{yielding the monomial}\quad\quad a_{1}a_{2}a_{3}a_{4}b^{\dagger}+\text{H.c.}\,. (E5)

This monomial is multiplied by a product of accompanying constants, ϕ1ϕ2ϕ3ϕ4ϕbξsubscriptitalic-ϕ1subscriptitalic-ϕ2subscriptitalic-ϕ3subscriptitalic-ϕ4subscriptitalic-ϕ𝑏𝜉\phi_{1}\phi_{2}\phi_{3}\phi_{4}\phi_{b}\xi, to yield an effective correction rate κ(ϕ1ϕ2ϕ3ϕ4ϕbξ)2/κbproportional-to𝜅superscriptsubscriptitalic-ϕ1subscriptitalic-ϕ2subscriptitalic-ϕ3subscriptitalic-ϕ4subscriptitalic-ϕ𝑏𝜉2subscript𝜅𝑏\kappa\propto(\phi_{1}\phi_{2}\phi_{3}\phi_{4}\phi_{b}\xi)^{2}/\kappa_{b}, after adiabatic elimination.

The above scheme is done for each plaquette of the surface-code architecture. The additional aj2n¯superscriptsubscript𝑎𝑗2¯na_{j}^{2}-\textnormal{$\bar{\textsc{n}}$} dissipators — required for restricting each mode j𝑗j to antipodal coherent states — are realized in said architecture using a single ATS [17, Appx. B.2]. Together, these provide autonomous protection against all Z𝑍Z-type errors, utilizing the full error-correcting power of the (outer) surface code for such noise. Extension to other QLDPC codes is straightforward, barring any issues with long-range physical connectivity.

E.0.3 Hessian QSC

The Hessian code requires three jump operators, two of which consist of three monomials. Each jump operator can be realized using the general scheme above, providing a non-trivial QSC example that should be realizable with state-of-the-art ATS technology.

  1. 1.

    The jump F=a1a2a3𝐹subscript𝑎1subscript𝑎2subscript𝑎3F=a_{1}a_{2}a_{3} has no constant term, so no drive term is necessary. Only one pump tone, with amplitude ξ1ξsubscript𝜉1𝜉\xi_{1}\equiv\xi and frequency Ω1ΩsubscriptΩ1Ω\Omega_{1}\equiv\Omega, is required, and φ=π/2𝜑𝜋2\varphi=\pi/2 to obtain a cosine ATS. The ATS term is expanded to second order. The conditions selecting the desired monomial are

    Ω=ω1+ω2+ω3ωbyielding the monomiala1a2a3b+H.c..Ωsubscript𝜔1subscript𝜔2subscript𝜔3subscript𝜔𝑏yielding the monomialsubscript𝑎1subscript𝑎2subscript𝑎3superscript𝑏H.c.\Omega=\omega_{1}+\omega_{2}+\omega_{3}-\omega_{b}\quad\quad\text{yielding the monomial}\quad\quad a_{1}a_{2}a_{3}b^{\dagger}+\text{H.c.}\,. (E6)

    This monomial corresponds to an effective correction rate κ(ϕ1ϕ2ϕ3ϕbξ)2/κbproportional-to𝜅superscriptsubscriptitalic-ϕ1subscriptitalic-ϕ2subscriptitalic-ϕ3subscriptitalic-ϕ𝑏𝜉2subscript𝜅𝑏\kappa\propto(\phi_{1}\phi_{2}\phi_{3}\phi_{b}\xi)^{2}/\kappa_{b} after adiabatic elimination.

  2. 2.

    The jump F=a13+a23+a33𝐹superscriptsubscript𝑎13superscriptsubscript𝑎23superscriptsubscript𝑎33F=a_{1}^{3}+a_{2}^{3}+a_{3}^{3} also has no constant term. The three terms require three drive tones, with parameters {ξp,Ωp}subscript𝜉𝑝subscriptΩ𝑝\{\xi_{p},\Omega_{p}\} for p{1,2,3}𝑝123p\in\{1,2,3\}. The phase φ=π/2𝜑𝜋2\varphi=\pi/2 so that the ATS term becomes a cosine. The ATS cosine term is expanded to second order. The conditions selecting the desired monomials are

    ξp=1/ϕj=p3Ωp=3ωj=pωb,yielding the monomialsaj=p3b+H.c..subscript𝜉𝑝absent1superscriptsubscriptitalic-ϕ𝑗𝑝3subscriptΩ𝑝absent3subscript𝜔𝑗𝑝subscript𝜔𝑏yielding the monomialssuperscriptsubscript𝑎𝑗𝑝3superscript𝑏H.c.\begin{aligned} \xi_{p}&=1/\phi_{j=p}^{3}\\ \Omega_{p}&=3\omega_{j=p}-\omega_{b}\end{aligned},\quad\quad\text{yielding the monomials}\quad\quad a_{j=p}^{3}b^{\dagger}+\text{H.c.}\,. (E7)

    Each of these monomials is multiplied by a product of respective accompanying constants, ϕj=p3ϕbξp=ϕbsuperscriptsubscriptitalic-ϕ𝑗𝑝3subscriptitalic-ϕ𝑏subscript𝜉𝑝subscriptitalic-ϕ𝑏\phi_{j=p}^{3}\phi_{b}\xi_{p}=\phi_{b}, where we have used the drive-tone amplitudes to cancel the non-tunable coupling strengths ϕjsubscriptitalic-ϕ𝑗\phi_{j}. This yields an effective correction rate κϕb2/κbproportional-to𝜅superscriptsubscriptitalic-ϕ𝑏2subscript𝜅𝑏\kappa\propto\phi_{b}^{2}/\kappa_{b}, after adiabatic elimination.

  3. 3.

    The jump F=a16+a26+a36n¯3/4𝐹superscriptsubscript𝑎16superscriptsubscript𝑎26superscriptsubscript𝑎36superscript¯n34F=a_{1}^{6}+a_{2}^{6}+a_{3}^{6}-\textnormal{$\bar{\textsc{n}}$}^{3}/4 has three monomials and a constant term with power P=3𝑃3P=3 and coefficient 1/4141/4. The three monomials require three drive tones, with parameters {ξp,Ωp}subscript𝜉𝑝subscriptΩ𝑝\{\xi_{p},\Omega_{p}\} for p{1,2,3}𝑝123p\in\{1,2,3\}. The phase φ=0𝜑0\varphi=0 so that the ATS term remains a sine. This term is then expanded to fourth order. The conditions selecting the desired monomials are

    ξp=1/ϕj=p6Ωp=6ωj=pωb,yielding the monomialsaj=p6b+H.c..subscript𝜉𝑝absent1superscriptsubscriptitalic-ϕ𝑗𝑝6subscriptΩ𝑝absent6subscript𝜔𝑗𝑝subscript𝜔𝑏yielding the monomialssuperscriptsubscript𝑎𝑗𝑝6superscript𝑏H.c.\begin{aligned} \xi_{p}&=1/\phi_{j=p}^{6}\\ \Omega_{p}&=6\omega_{j=p}-\omega_{b}\end{aligned},\quad\quad\text{yielding the monomials}\quad\quad a_{j=p}^{6}b^{\dagger}+\text{H.c.}\,. (E8)

    Each of these monomials is multiplied by ϕbsubscriptitalic-ϕ𝑏\phi_{b}. The fluxes gjsubscript𝑔𝑗g_{j} are required to be equal for all three j𝑗j in order to realize the jump. This yields an effective correction rate κϕb2/κbproportional-to𝜅superscriptsubscriptitalic-ϕ𝑏2subscript𝜅𝑏\kappa\propto\phi_{b}^{2}/\kappa_{b}, after adiabatic elimination.

References

  • Sun et al. [2014] L. Sun, A. Petrenko, Z. Leghtas, B. Vlastakis, G. Kirchmair, K. M. Sliwa, A. Narla, M. Hatridge, S. Shankar, J. Blumoff, L. Frunzio, M. Mirrahimi, M. H. Devoret, and R. J. Schoelkopf, Tracking photon jumps with repeated quantum non-demolition parity measurements., Nature 511, 444 (2014).
  • Ofek et al. [2016] N. Ofek, A. Petrenko, R. Heeres, P. Reinhold, Z. Leghtas, B. Vlastakis, Y. Liu, L. Frunzio, S. M. Girvin, L. Jiang, M. Mirrahimi, M. H. Devoret, and R. J. Schoelkopf, Extending the lifetime of a quantum bit with error correction in superconducting circuits, Nature 536, 441 (2016).
  • Leghtas et al. [2015] Z. Leghtas, S. Touzard, I. M. Pop, A. Kou, B. Vlastakis, A. Petrenko, K. M. Sliwa, A. Narla, S. Shankar, M. J. Hatridge, M. Reagor, L. Frunzio, R. J. Schoelkopf, M. Mirrahimi, and M. H. Devoret, Confining the state of light to a quantum manifold by engineered two-photon loss, Science 347 (2015).
  • Touzard et al. [2018] S. Touzard, A. Grimm, Z. Leghtas, S. Mundhada, P. Reinhold, C. Axline, M. Reagor, K. Chou, J. Blumoff, K. Sliwa, S. Shankar, L. Frunzio, R. J. Schoelkopf, M. Mirrahimi, and M. Devoret, Coherent Oscillations inside a Quantum Manifold Stabilized by Dissipation, Phys. Rev. X 8, 021005 (2018).
  • Grimm et al. [2020] A. Grimm, N. E. Frattini, S. Puri, S. O. Mundhada, S. Touzard, M. Mirrahimi, S. M. Girvin, S. Shankar, and M. H. Devoret, Stabilization and operation of a Kerr-cat qubit, Nature 584, 205 (2020).
  • Campagne-Ibarcq et al. [2020] P. Campagne-Ibarcq, A. Eickbusch, S. Touzard, E. Zalys-Geller, N. E. Frattini, V. V. Sivak, P. Reinhold, S. Puri, S. Shankar, R. J. Schoelkopf, L. Frunzio, M. Mirrahimi, and M. H. Devoret, Quantum error correction of a qubit encoded in grid states of an oscillator, Nature 584, 368 (2020).
  • Flühmann et al. [2019] C. Flühmann, T. L. Nguyen, M. Marinelli, V. Negnevitsky, K. Mehta, and J. Home, Encoding a qubit in a trapped-ion mechanical oscillator, Nature 566, 513 (2019).
  • de Neeve et al. [2022] B. de Neeve, T.-L. Nguyen, T. Behrle, and J. P. Home, Error correction of a logical grid state qubit by dissipative pumping, Nat. Phys. 18, 296 (2022).
  • Sivak et al. [2022] V. V. Sivak, A. Eickbusch, B. Royer, S. Singh, I. Tsioutsios, S. Ganjam, A. Miano, B. L. Brock, A. Z. Ding, L. Frunzio, S. M. Girvin, R. J. Schoelkopf, and M. H. Devoret, Real-time quantum error correction beyond break-even (2022).
  • Dahan et al. [2022] R. Dahan, G. Baranes, A. Gorlach, R. Ruimy, N. Rivera, and I. Kaminer, Creation of Optical Cat and GKP States Using Shaped Free Electrons (2022).
  • Cohen and Mirrahimi [2014] J. Cohen and M. Mirrahimi, Dissipation-induced continuous quantum error correction for superconducting circuits, Phys. Rev. A 90, 062344 (2014).
  • Fukui et al. [2017] K. Fukui, A. Tomita, and A. Okamoto, Analog Quantum Error Correction with Encoding a Qubit into an Oscillator, Phys. Rev. Lett. 119, 180507 (2017).
  • Fukui et al. [2018] K. Fukui, A. Tomita, A. Okamoto, and K. Fujii, High-Threshold Fault-Tolerant Quantum Computation with Analog Quantum Error Correction, Phys. Rev. X 8, 021054 (2018).
  • Guillaud and Mirrahimi [2019] J. Guillaud and M. Mirrahimi, Repetition Cat Qubits for Fault-Tolerant Quantum Computation, Phys. Rev. X 9, 041053 (2019).
  • Vuillot et al. [2019] C. Vuillot, H. Asasi, Y. Wang, L. P. Pryadko, and B. M. Terhal, Quantum error correction with the toric Gottesman-Kitaev-Preskill code, Phys. Rev. A 99, 032344 (2019).
  • Noh and Chamberland [2020] K. Noh and C. Chamberland, Fault-tolerant bosonic quantum error correction with the surface–Gottesman-Kitaev-Preskill code, Phys. Rev. A 101, 012316 (2020).
  • Chamberland et al. [2022] C. Chamberland, K. Noh, P. Arrangoiz-Arriola, E. T. Campbell, C. T. Hann, J. Iverson, H. Putterman, T. C. Bohdanowicz, S. T. Flammia, A. Keller, G. Refael, J. Preskill, L. Jiang, A. H. Safavi-Naeini, O. Painter, and F. G. Brandão, Building a Fault-Tolerant Quantum Computer Using Concatenated Cat Codes, PRX Quantum 3, 010329 (2022).
  • Larsen et al. [2021] M. V. Larsen, C. Chamberland, K. Noh, J. S. Neergaard-Nielsen, and U. L. Andersen, Fault-Tolerant Continuous-Variable Measurement-based Quantum Computation Architecture, PRX Quantum 2, 030325 (2021).
  • Guillaud and Mirrahimi [2021] J. Guillaud and M. Mirrahimi, Error Rates and Resource Overheads of Repetition Cat Qubits, Phys. Rev. A 103, 042413 (2021).
  • Noh et al. [2022] K. Noh, C. Chamberland, and F. G. Brandão, Low-Overhead Fault-Tolerant Quantum Error Correction with the Surface-GKP Code, PRX Quantum 3, 010315 (2022).
  • Zhang et al. [2022] J. Zhang, Y.-C. Wu, and G.-P. Guo, Concatenation of the Gottesman-Kitaev-Preskill code with the XZZX surface code (2022).
  • Régent et al. [2022] F.-M. L. Régent, C. Berdou, Z. Leghtas, J. Guillaud, and M. Mirrahimi, High-performance repetition cat code using fast noisy operations (2022).
  • Stafford and Menicucci [2022] M. P. Stafford and N. C. Menicucci, Biased Gottesman-Kitaev-Preskill repetition code (2022).
  • Lieu et al. [2022] S. Lieu, Y.-J. Liu, and A. V. Gorshkov, Candidate for a passively-protected quantum memory in two dimensions (2022).
  • Gouzien et al. [2023] E. Gouzien, D. Ruiz, F.-M. L. Regent, J. Guilland, and N. Sangouard, Computing 256-bit Elliptic Curve Logarithm in 9 Hours with 126133 Cat Qubits (2023).
  • Albert et al. [2020] V. V. Albert, J. P. Covey, and J. Preskill, Robust Encoding of a Qubit in a Molecule, Physical Review X 10, 031050 (2020).
  • Byron Bay Quantum Computing Workshop [2021] Byron Bay Quantum Computing Workshop, GKP retrospective & bosonic codes panel discussion with Gottesman, Devoret, Divincenzo, & Girvin (2021).
  • Gottesman [2022] D. Gottesman, Opportunities and Challenges in Fault-Tolerant Quantum Computation (2022).
  • Cochrane et al. [1999] P. T. Cochrane, G. J. Milburn, and W. J. Munro, Macroscopically distinct quantum-superposition states as a bosonic code for amplitude damping, Phys. Rev. A 59, 2631 (1999).
  • Leghtas et al. [2013] Z. Leghtas, G. Kirchmair, B. Vlastakis, R. J. Schoelkopf, M. H. Devoret, and M. Mirrahimi, Hardware-Efficient Autonomous Quantum Memory Protection, Phys. Rev. Lett. 111, 120501 (2013).
  • Conway and Sloane [1999] J. H. Conway and N. J. A. Sloane, Sphere Packings, Lattices and Groups, Grundlehren der mathematischen Wissenschaften, Vol. 290 (Springer New York, New York, NY, 1999).
  • Ericson and Zinoviev [2001] T. Ericson and V. Zinoviev, Codes on Euclidean Spheres (North Holland, 2001).
  • Pedernales et al. [2020] J. S. Pedernales, F. Cosco, and M. B. Plenio, Decoherence-Free Rotational Degrees of Freedom for Quantum Applications, Phys. Rev. Lett. 125, 090501 (2020).
  • Cerf et al. [2007] N. J. Cerf, G. Leuchs, and E. S. Polzik, Quantum Information with Continuous Variables of Atoms and Light (Imperial College Press, 2007).
  • Serafini [2017] A. Serafini, Quantum Continuous Variables: A Primer of Theoretical Methods (CRC Press, Boca Raton FL, 2017).
  • Denys and Leverrier [2022] A. Denys and A. Leverrier, The 2T-qutrit, a two-mode bosonic qutrit (2022).
  • Schlegel et al. [2022] D. S. Schlegel, F. Minganti, and V. Savona, Quantum error correction using squeezed Schrodinger cat states, Physical Review A 106, 022431 (2022).
  • Xu et al. [2022] Q. Xu, G. Zheng, Y.-X. Wang, P. Zoller, A. A. Clerk, and L. Jiang, Autonomous quantum error correction and fault-tolerant quantum computation with squeezed cat qubits (2022).
  • Hillmann and Quijandr\́mathrm{missing}ia [2022] T. Hillmann and F. Quijandr\́mathrm{i}a, Quantum error correction with dissipatively stabilized squeezed cat qubits (2022).
  • Albert et al. [2019] V. V. Albert, S. O. Mundhada, A. Grimm, S. Touzard, M. H. Devoret, and L. Jiang, Pair-cat codes: autonomous error-correction with low-order nonlinearity, Quantum Sci. Technol. 4, 035007 (2019).
  • Gross [2021] J. A. Gross, Encoding a qubit in a spin, Phys. Rev. Lett. 127, 010504 (2021).
  • Turchette et al. [2000] Q. A. Turchette, C. J. Myatt, B. E. King, C. A. Sackett, D. Kielpinski, W. M. Itano, C. Monroe, and D. J. Wineland, Decoherence and decay of motional quantum states of a trapped atom coupled to engineered reservoirs, Phys. Rev. A 62, 053807 (2000).
  • Grimsmo et al. [2020] A. L. Grimsmo, J. Combes, and B. Q. Baragiola, Quantum Computing with Rotation-Symmetric Bosonic Codes, Phys. Rev. X 10, 011058 (2020).
  • Knill and Laflamme [1997] E. Knill and R. Laflamme, Theory of quantum error-correcting codes, Phys. Rev. A 55, 900 (1997).
  • Coxeter [1973] H. S. M. Coxeter, Regular polytopes, 3rd ed. (Dover Publications, New York, 1973).
  • Shephard [1952] G. C. Shephard, Regular Complex Polytopes, P. Lond. Math. Soc. s3-2, 82 (1952).
  • Coxeter [1991] H. S. M. Coxeter, Regular complex polytopes, 2nd ed. (Cambridge University Press, Cambridge [England]; New York, 1991).
  • Sidelnikov [1997a] V. Sidelnikov, On a finite group of matrices and codes on the Euclidean sphere, Problems Inform. Transmission 33, 29 (1997a).
  • Sidelnikov [1997b] V. Sidelnikov, On a finite group of matrices generating orbit codes on Euclidean sphere, in IEEE Int. Symp. Inf. Theory - Proc. (IEEE, Ulm, Germany, 1997) p. 436.
  • Sidelnikov [1999] V. Sidelnikov, Spherical 7-Designs in 2n-Dimensional Euclidean Space, J. Algebraic Comb. 10, 279 (1999).
  • Nebe et al. [2001] G. Nebe, E. M. Rains, and N. J. A. Sloane, The Invariants of the Clifford Groups, Designs, Codes and Cryptography 24, 99 (2001).
  • Delsarte et al. [1977] P. Delsarte, J. M. Goethals, and J. J. Seidel, Spherical codes and designs, Geom. Dedicata 6, 363 (1977).
  • Roy and Suda [2014] A. Roy and S. Suda, Complex Spherical Designs and Codes, J. Combin. Designs 22, 105 (2014).
  • Mohammadpour and Waldron [2023] M. Mohammadpour and S. Waldron, Complex spherical designs from group orbits (2023).
  • Schumacher [1996] B. Schumacher, Sending entanglement through noisy quantum channels, Phys. Rev. A 54, 2614 (1996).
  • Reimpell and Werner [2005] M. Reimpell and R. F. Werner, Iterative Optimization of Quantum Error Correcting Codes, Phys. Rev. Lett. 94, 080501 (2005).
  • Fletcher et al. [2007] A. S. Fletcher, P. W. Shor, and M. Z. Win, Optimum quantum error recovery using semidefinite programming, Phys. Rev. A 75, 012338 (2007).
  • Fletcher et al. [2008] A. S. Fletcher, P. W. Shor, and M. Z. Win, Channel-Adapted Quantum Error Correction for the Amplitude Damping Channel, IEEE Trans. Inf. Theory 54, 5705 (2008).
  • Albert et al. [2018] V. V. Albert, K. Noh, K. Duivenvoorden, D. J. Young, R. T. Brierley, P. Reinhold, C. Vuillot, L. Li, C. Shen, S. M. Girvin, B. M. Terhal, and L. Jiang, Performance and structure of single-mode bosonic codes, Phys. Rev. A 97, 032346 (2018).
  • Calderbank and Shor [1995] A. R. Calderbank and P. W. Shor, Good Quantum Error-Correcting Codes Exist, Phys. Rev. A 54, 1098 (1995).
  • Steane [1996a] A. M. Steane, Error Correcting Codes in Quantum Theory, Phys. Rev. Lett. 77, 793 (1996a).
  • Steane [1996b] A. Steane, Multiple-particle interference and quantum error correction, Proc. R. Soc. A 452, 2551 (1996b).
  • Puri et al. [2020] S. Puri, L. St-Jean, J. A. Gross, A. Grimm, N. E. Frattini, P. S. Iyer, A. Krishna, S. Touzard, L. Jiang, A. Blais, S. T. Flammia, and S. M. Girvin, Bias-preserving gates with stabilized cat qubits, Sci. Adv. 610.1126/sciadv.aay5901 (2020).
  • Mirrahimi et al. [2014] M. Mirrahimi, Z. Leghtas, V. V. Albert, S. Touzard, R. J. Schoelkopf, L. Jiang, and M. H. Devoret, Dynamically protected cat-qubits: a new paradigm for universal quantum computation, New J. Phys. 16, 045014 (2014).
  • Goto [2016] H. Goto, Bifurcation-based adiabatic quantum computation with a nonlinear oscillator network: Toward quantum soft computing, Sci. Rep. 6, 21686 (2016).
  • Puri et al. [2017] S. Puri, S. Boutin, and A. Blais, Engineering the quantum states of light in a Kerr-nonlinear resonator by two-photon driving, npj Quantum Inf. 3, 18 (2017).
  • Girvin [2023] S. M. Girvin, Introduction to Quantum Error Correction and Fault Tolerance (2023).
  • Lescanne et al. [2020] R. Lescanne, M. Villiers, T. Peronnin, A. Sarlette, M. Delbecq, B. Huard, T. Kontos, M. Mirrahimi, and Z. Leghtas, Exponential suppression of bit-flips in a qubit encoded in an oscillator, Nat. Phys. 16, 509 (2020).
  • Slepian [1968] D. Slepian, Group Codes for the Gaussian Channel, Bell Syst. Tech. J. 47, 575 (1968).
  • Loeliger [1991] H.-A. Loeliger, Signal sets matched to groups, IEEE Trans. Inform. Theory 37, 1675 (1991).
  • Mittelholzer and Lahtonen [1996] T. Mittelholzer and J. Lahtonen, Group codes generated by finite reflection groups, IEEE Trans. Inform. Theory 42, 519 (1996).
  • Sloane et al. [2000] N. Sloane, R. Hardin, W. Smith, and others, Spherical codes (2000).
  • Ballinger et al. [2009] B. Ballinger, G. Blekherman, H. Cohn, N. Giansiracusa, E. Kelly, and A. Schürmann, Experimental Study of Energy-Minimizing Point Configurations on Spheres, Exp. Math. 18, 257 (2009).
  • Cameron et al. [1978] P. Cameron, J. Goethals, and J. Seidel, Strongly regular graphs having strongly regular subconstituents, Journal of Algebra 55, 257 (1978).
  • Waldron [2018] S. F. D. Waldron, An Introduction to Finite Tight Frames, Appl. Harm. Numer. Anal. (Springer New York, New York, NY, 2018).
  • Bajnok [1991] B. Bajnok, Construction of Designs on the 2-Sphere, Eur. J. Combin. 12, 377 (1991).
  • Reznick [1995] B. Reznick, Some constructions of spherical 5-designs, Linear Algebra Appl. 226-228, 163 (1995).
  • Xiang [2022] Z. Xiang, Explicit spherical designs, J. Algebraic Comb. 5, 347 (2022).
  • Coxeter and Shephard [1992] H. S. M. Coxeter and G. C. Shephard, Portraits of a Family of Complex Polytopes, Leonardo 25, 239 (1992).
  • Skilling [1976] J. Skilling, Uniform compounds of uniform polyhedra, Math. Proc. Camb. Phil. Soc. 79, 447 (1976).
  • McMullen [2018] P. McMullen, New Regular Compounds of 4-Polytopes, in New Trends in Intuitive Geometry, Vol. 27 (Springer Berlin Heidelberg, Berlin, Heidelberg, 2018) pp. 307–320.
  • Mamone et al. [2010] S. Mamone, G. Pileio, and M. H. Levitt, Orientational Sampling Schemes Based on Four Dimensional Polytopes, Symmetry 2, 1423 (2010).
  • Klitzing [2023] R. Klitzing, Polytopes & their Incidence Matrices (2023).
  • Hanson [2005] A. J. Hanson, Visualizing quaternions, in ACM SIGGRAPH 2005 Courses on - SIGGRAPH ’05 (ACM Press, Los Angeles, California, 2005) p. 1.
  • Coxeter et al. [1997] H. S. M. Coxeter, J. C. Fisher, and J. B. Wilker, Coordinates for the Regular Complex Polygons, J. London Math. Soc. 55, 527 (1997).
  • Coxeter [1997] H. S. M. Coxeter, Reciprocating the Regular Polytopes, J. London Math. Soc. 55, 549 (1997).
  • Bruss and Leuchs [2007] D. Bruss and G. Leuchs, eds., Lectures on quantum information (Wiley-VCH, Weinheim, 2007).
  • Nielsen and Chuang [2011] M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information: 10th Anniversary Edition, 10th ed. (Cambridge University Press, New York, NY, USA, 2011).
  • Agrawal et al. [2018] A. Agrawal, R. Verschueren, S. Diamond, and S. Boyd, A rewriting system for convex optimization problems, Journal of Control and Decision 5, 42 (2018).
  • Diamond and Boyd [2016] S. Diamond and S. Boyd, CVXPY: A Python-Embedded Modeling Language for Convex Optimization, Journal of Machine Learning Research 17, 1 (2016).
  • Michael et al. [2016] M. H. Michael, M. Silveri, R. T. Brierley, V. V. Albert, J. Salmilehto, L. Jiang, and S. M. Girvin, New Class of Quantum Error-Correcting Codes for a Bosonic Mode, Phys. Rev. X 6, 031006 (2016).
  • Li et al. [2017] L. Li, C.-l. Zou, V. V. Albert, S. Muralidharan, S. M. Girvin, and L. Jiang, Cat Codes with Optimal Decoherence Suppression for a Lossy Bosonic Channel, Phys. Rev. Lett. 119, 030502 (2017).
  • Note [1] Up to constant prefactors, this term is also obtained from the last two terms in [68, Eq. (S2)] by setting φΔ=φ+π/2subscript𝜑Δ𝜑𝜋2\varphi_{\Delta}=\varphi+\pi/2, ΔEJ=0Δsubscript𝐸𝐽0\Delta E_{J}=0, and φΣ=ϵ(t)+π/2subscript𝜑Σitalic-ϵ𝑡𝜋2\varphi_{\Sigma}=\epsilon(t)+\pi/2, and expanding to first order in ϵitalic-ϵ\epsilon. The quadratic term in that equation can absorbed into the bare oscillator Hamiltonian prior going into the rotating frame by expanding φ𝜑\varphi in a set of normal modes.
  • Reiter and Sω𝜔{\omega{}}rensen [2012] F. Reiter and A. S. Sω𝜔{\omega{}}rensen, Effective operator formalism for open quantum systems, Phys. Rev. A 85, 032111 (2012).
  • Azouit et al. [2016] R. Azouit, A. Sarlette, and P. Rouchon, Adiabatic elimination for open quantum systems with effective Lindblad master equations, in 2016 IEEE 55th Conf. Decis. Control (IEEE, 2016) pp. 4559–4565.
  • Azouit et al. [2017] R. Azouit, F. Chittaro, A. Sarlette, and P. Rouchon, Towards generic adiabatic elimination for bipartite open quantum systems, Quantum Sci. Technol. 2, 044011 (2017).
  • Zitko [2011] R. Zitko, SNEG – Mathematica package for symbolic calculations with second-quantization-operator expressions, Computer Physics Communications 182, 2259 (2011).
  • Zitko [2022] R. Zitko, rokzitko/sneg: April 2022 release (2022).