Quantum storage in quantum ferromagnets

Yingkai Ouyang1, 2, 3, ∗ 1Department of Physics and Astronomy, University of Sheffield, Sheffield, UK 2Singapore University of Technology and Design, 8 Somapah Road, Singapore 3Centre for Quantum Technologies, National University of Singapore, 3 Science Drive 2, Singapore We must protect inherently fragile quantum data to unlock the potential of quantum technologies. A perti- nent concern in schemes for quantum storage is their potential for near-term implementation. Since Heisenberg ferromagnets are readily available, we investigate their potential for robust quantum storage. We propose to use permutation-invariant quantum codes to store quantum data in Heisenberg ferromagnets, because the ground space of any Heisenberg ferromagnet must be symmetric under any permutation of the underlying . By exploiting an area law on the expected energy of Pauli errors, we show that increasing the effective dimension of Heisenberg ferromagnets can improve the storage lifetime. When the effective dimension of Heisenberg fer- romagnets is maximal, we also obtain an upper bound for the storage error. This result relies on perturbation theory, where we use Davis’ divided difference representation for Frechet´ derivatives along with the recursive structure of these divided differences. Our numerical bounds allow us to better understand how quantum mem- ory lifetimes can be enhanced in Heisenberg ferromagnets.

Introduction.— Decoherence quickly renders unprotected tions from satisfying constraint (iii). Consequently, determin- quantum data unreliable. To combat this, it becomes necessary ing whether such physical systems can satisfy constraint (iii) to encode quantum data into codes. is especially pertinent. In this paper, we study Heisenberg fer- The challenge in designing robust quantum memories arises romagnets as media for quantum storage because they com- from the difficulty of simultaneously (i) utilizing an easily ac- prise of non-commuting two-body interactions and therefore cessible physical system, (ii) having a quantum code that lies sidestep the no-go result of [12]. We also study to what extent within the ground space of the system’s Hamiltonian H, and Heisenberg ferromagnets satisfy constraint (iii). (iii) having an increased storage lifetime τ with an increasing The Heisenberg ferromagnet (HF) is a model of quantum number of qubits N in the physical system. Self-correcting magnetism, and is prevalent in many naturally occurring phys- quantum memories [1, 2] should satisfy (ii) and (iii), but are ical systems, and thereby satisfies constraint (i). For instance, challenging to implement in a multitude of desirable settings the HF is found in various cuprates [19, 20], in solid Helium-3 [3–14]. Indeed, constraint (i) does easily not hold, which frus- [21], and more generally in systems with interacting electrons trates the design of reliable quantum storage. For instance, [22]. Even in many physical systems that cannot be naturally quantum memories based on stabilizer codes which correct at interpreted as ferromagnets, effective HFs can nonetheless be least one error and also satisfy (ii) unfortunately reside in un- engineered, for instance by symmetrizing systems dominated physical systems with many-body interactions, and can only by dipole interactions using dynamic pulse sequences [23]. be approximately constructed [4, 15–17]. Of these constraints, Effective HFs have also been engineered in ultracold atomic it is most pertinent to satisfy (i), because physically unrealistic gases [24] and quantum dots [25]. Specifically, we study - quantum memories will be difficult to engineer. half HFs in the absence of an external magnetic field, with There are two reasons to store quantum data within the Hamiltonian of the form ground space and thereby satisfy constraint (ii). First, a grow- ing energy gap can suppress excitations from the ground space H = − ∑ J(1 − πi, j). (1) [18]. Second, storing quantum data within the ground space {i, j}∈E Here, 1 denotes an N- identity operator, π denotes a avoids unnecessary errors that can occur even in the complete i, j swap operator on the ith and jth qubits, J denotes the exchange absence of noise. Any state within the ground space is an constants, and E denotes the set of interactions. Such HFs eigenstate of the physical system, and for such states, they are have J > 0 and ground state energy set to zero. arXiv:1904.01458v3 [quant-ph] 3 Feb 2020 left unchanged by a unitary operation U that the system’s nat- τ By storing quantum data using permutation-invariant (PI) ural dynamics induces, after a storage time of τ elapses. By codes in HFs, we automatically satisfy constraint (ii). This avoiding the need to uncompute U , we would not suffer from τ is because symmetric states lie within the ground space of an imperfect reversal of U caused by our imprecise knowl- τ any HF, and quantum data in PI codes, by being invari- edge of τ. ant under any permutation of their underlying qubits, are Storage within the ground space, while satisfying constraint symmetric states. The possibility of quantum error correc- (ii), is not enough to result in self-correcting quantum memo- tion using a PI code was first affirmed by Ruskai, via con- ries and thereby satisfy constraint (iii). Moreover, many phys- struction of an explicit PI- 9,1,3 code that corrects one ically realistic systems satisfying constraint (i) comprising of error [26] 1. Ruskai’s resultJ [26]K was subsequently im- two-local terms are surprisingly incompatible with constraint proved to yield PI- 7,1,3 codes [27]. In [28], Ouyang (iii) [12]. However, this no-go result does not preclude phys- J K ical systems comprising of non-commuting two-body interac-

1 Here, N,K,D denotes a quantum code that encodes K logical qubits into N physicalJ qubitsK and with distance D. When one needs to correct t errors ∗ y.ouyang@sheffield.ac.uk on a quantum code, it suffices to have D > 2t. 2 N 2 constructs PI- (2t + 1) ,1,2t + 1 codes that correct t arbi- where r = 0,1, and |Dwi are Dicke states of weight w [33, trary errors, therebyJ generalizingK Ruskai’s nine-qubit code. 34]. Further generalizing these codes, Ouyang introduces PI- We quantify a HF’s dimension using properties of its un- 2 (2t + 1) (d − 1),log2 d,2t + 1 codes [29]. Approximate derlying graph of interactions [35]. This graph G has vertices quantumJ error correcting PI codesK were also investigated in labeled from 1 to N, and edges E that correspond to the in- qubit [28, 30, 31] and bosonic settings [32]. Research on PI teractions in the HF’s Hamiltonian H. Given a subset S of codes shows that quantum correction is possible within the {1,...,N}, let ∂E S denote its edge-boundary with respect to ground space of HFs, which is only suggestive that constraint the edge set E, which is the set of edges in E with exactly (iii) can be compatible with PI codes. This is because the cod- one vertex in S. When every subset S satisfies the inequality 1−1/δ ing parameters of PI codes alone, being independent of the |∂E S| ≥ cmin(|S|,N − |S|) , the graph and HF have di- parameters in HFs, are not enough to determine what happens mension δ with isoperimetric constant c. when physical noise applies to PI codes stored in HFs. To Given a set P of N-qubit Pauli errors that afflict at most better understand the extent in which HFs with PI codes can N/2 qubits, let satisfy constraint (iii), we study bounds on the storage error of PI codes under the action of two different noise models, where hPi = min hψ|PHP|ψi (3) both bounds depend on properties of the underlying HF. |ψi∈C Our first noise model applies to HFs of any geometry, and denote the minimum expected energy of P ∈ on the code introduces Pauli errors. These Pauli errors occur with proba- P . When is a GNU code, we derive a lower bound on hPi in bilities that are thermodynamically related to their expected C C terms of the edge boundary V(P), where V(P) denotes the set energies on the codespace of a specific family of PI codes of vertices on which P acts non-trivially. In particular, Theo- [28]. To derive an upper bound on the storage error, we find rem 1 below gives an area law on the minimum size of hPi. an area law on the expected energy of a Pauli error, which demonstrates that a quantum memory based in a HF can have Theorem 1. Let C be an N-qubit GNU code with parameters a macroscopic energy barrier for Pauli errors. This allows us g = n = 2t +1 and u = 2, where N = 2(2t +1)2 and t ≥ 1. Then to show that the storage error decreases with increasing dimen- with respect to the Hamiltonian H given by (1) with exchange sionality of the HF. constants J and set of interactions E, for every N-qubit Pauli Our second noise model introduces unitary errors proba- P in P, we have bilistically, where each unitary arises from a local perturba- tion of the underlying Hamiltonian. Such a noise model can hPi ≥ χJ|∂E (V(P))|, describe the effects of unwanted physical interactions, such as spurious local fields afflicting each particle independently. where χ = min{2µ,1 − 4µ} and We use perturbation theory to bound the storage error by us- ing Davis’ divided difference representation of these taking µ = (1 + 5t + 6t2)/(4 + 32t + 32t2). Frechet´ derivatives. Because we require complete knowledge of the Hamiltonian’s spectrum, we restrict our analysis to ex- The significance of Theorem 1 lies in the geometric inter- actly solvable mean-field HFs. In such HFs every pair of spins pretation it imparts to hPi. Namely, when the graph G has interacts with equal strength. Since such HFs have an infinite dimension δ and isoperimetric number c, we have the isoperi- effective dimension, analyzing them is indicative of the ulti- metric inequality mate limits of robust quantum storage in HFs. With respect to both noise models, we provide upper bounds hPi ≥ Jχc|P|1−1/δ , (4) for the storage error of quantum memories in HFs that are nu- merically tractable. In both cases, we find that quantum mem- where |P| = |V(P)| denotes the weight of P. For a HF on a ories in HFs are partially self-correcting in the sense that is an 1D spin-chain, hPi ≥ Jχ. For a HF on a square lattice, [36] optimal system size for fixed noise parameters that minimizes with our result implies that hPi ≥ Jχp|P|. Whenever δ > 1, our upper bounds on the storage error. the expected energy of P grows with its weight, and we have Energy of Pauli errors and their geometry.— We use GNU a macroscopic energy barrier [37]. This suggests that when codes [28] to elucidate the dependence of a HF’s dimension δ > 1, HFs can be good quantum memories. To see this, con- with respect to the storage error of quantum data. GNU codes sider a noise model that introduces Pauli errors P ∈ P with depend on three parameters g, n, and u, and encode a single probability proportional to e−βhPi and with effective inverse qubit into N = gnu qubits. Here g and n quantify the distance temperature β. The corresponding channel is T , which ran- of the GNU code with respect to bit-flip and phase-flip errors domly introduces Pauli errors, and has the form respectively, while u is a scaling parameter where u ≥ 1. When g = n = 2t +1, the GNU code corrects t errors. A GNU code has T (ρ) = ∑ (e−βhPi/Z )PρP, (5) logical codewords P∈P

s n j gnu |rLi = |D i, (2) ∑ n−1 g j 2 0≤ j≤n 2 Dicke states are uniform superpositions of computational basis states la- mod( j,2)=r beled by binary vectors of Hamming weight w.

2 10 2 10 2

10 4 10 3

t 6

u 10

r r o o r r r

8 r e e

10

e 4 e

g 10 t = 1, N = 18 g J = 10 GHz, t = 1, N = 7 a a r

10 r

o 3

10 o t

t = 2, N = 50 t J = 10 GHz, t = 1, N = 7 s s 3 12 t = 3, N = 98 J = 10 GHz, t = 2, N = 25 10 5 t = 4, N = 162 10 J = 104 GHz, t = 2, N = 25 10 14 t = 5, N = 242 J = 105 GHz, t = 2, N = 25 t = 6, N = 338 6 16 J = 10 GHz, t = 3, N = 49 10 10 6 2 4 6 8 0 50 100 150 200 dimension storage lifetime (ns) FIG. 1: When a HF stores an encoded qubit within a GNU code on 2 FIG. 2: When a mean-field HF stores an encoded qubit within an N = 2(2t + 1) physical qubits that corrects t errors, we use (7) to N-qubit PI that corrects t errors, we use Theorem 3 to obtain upper u obtain upper bounds on the storage error t with respect to the HF’s bounds for the corresponding storage error ε after a target storage t J = c = dimension δ and . Here β 13 and 1. lifetime of τ. The baseline lifetime and storage error for an unpro- tected qubit are 12ns and 0.00048 respectively. The shaded region indicates where ε is smaller than the baseline. −βhPi where Z = ∑P∈P e . The corresponding probability of obtaining an uncorrectable error, which is also the storage er- ror after perfectly performing quantum error correction, is then Now, let perturbations A1,...,Aα to the Hamiltonian H oc- cur with probabilities p1,..., pα respectively. These perturba- 1 −βhPi tions model the effect of qubits coupling to spurious classical ut = ∑ e . (6) Z P∈P fields. Each perturbation A j is a linear combination of oper- |P|≥t+1 ators that acts non-trivially on a single qubit, and induces a −ixτ unitary evolution Uj = g(H + A j), where g(x) = e denotes From the isoperimetric inequality (4) and the bound an exponential function. The corresponding noisy channel is |∂EV(P)| ≤ ∆|P|, where ∆ is the maximum vertex degree of Nτ , where for any initial state ρ, we have G, we obtain α † Nτ (ρ) = p jUjρU . (8) −1 ∑ j N/2 ! N/2 ! j=1 N 1−1/δ N w −βJχcw w −βJ∆w In what follows, we focus our attention solely N , which we ut ≤ ∑ 3 e ∑ 3 e . τ w=t+1 w w=0 w call a random coherent noise channel, which is parameterized (7) by its noise strength a = max j kA jk/N. For any perturbation A j, the Taylor series of the unitary We illustrate (7) in Figure 1 with ∆ = 4,c = 1,βJ = 13, and g(H + A j) gives ∞ vary the dimension δ and number of correctible errors t. Fig- [k] ure 1 shows that long-range interactions can help to suppress g(H + A j) = g(H) + ∑ Dg (H,A j)/k!, (9) k=1 ut because increasing δ decreases ut . From Figure 1, when where δ < 4, the optimal code corrects between 1 error to 4 errors. k [k] d This shows that for a system with sufficient few dimensions, Dg (H,A j) = g(H + ξA j)| = (10) dξ k ξ 0 our scheme exhibits a partial self-correction property, where are Frechet´ derivatives of g(H) in the matrix direction A increasing the system size cannot indefinitely reduce the stor- j [38, 39]. Now the correctible component of g(H + A ) with age error. j respect to a code that corrects t errors comprises of Frechet´ Random coherent noise and storage error.— A good quan- derivatives of order at most t, because these Frechet´ derivatives tum memory preserves entanglement. Given a quantum code are polynomials in A of order no more than t. Therefore, we with logical codewords |0 i,...,|(M − 1) i, consider the j C L √ L study only the high order Frechet´ derivatives. These Frechet´ | i = M−1 | ji ⊗ | j i/ M. storage er- entangled state ΨC ∑ j=0 L The derivatives allow us to bound the storage error. ror of C with respect to a noisy channel N is Lemma 2. Given a quantum code C that corrects t errors, let ∞ [k] 1 R j = ∑k=t+1 Dg (H,A j) and define kRkC = max j{kR j|ψik : ε(N ,C ) = min |ΨC ihΨC | − R(N (|ΨC ihΨC |)) 1 , 2 R 2 |ψi ∈ C }. Then ε ≤ kRkC + kRkC . where R = I ⊗ R, N = I ⊗ N , I is an identity chan- From the integral representation of R j [40], we exploit the fact that g(H + A j) is unitary for Hermitian H and A to get nel, the minimization is over all recovery maps R, and k · k1 denotes the trace norm. To keep notation simple, when the {kD[t+1](H,A )| ik | i ∈ } code and noise model are already specified, we use the max g j ψ : ψ C C N kRkC ≤ max , (11) shorthand ε = ε(N ,C ). j (t + 1)!

3 which depends only on a single Frechet´ derivative instead of involves maximization. To avoid maximizing, we overesti- infinitely many. Given (11), one can clearly bound kRkC in mate the size of ht+1 by overcounting the contributions from terms of just kHk and aN. However, such a bound increases g(0,λn1 ,...,λnt+1 ) when there are repeated arguments in g. with increasing kHk, and exhibits a behavior contrary to the First we use (14) for divided differences even when there are numerical evidence in Figure 1. Increasing the number of r repeated arguments. Second, for divided arguments with r long-range interactions increases both kHk and dimensional- repeated entries, we count the contributions from leaves that ity, but since increasing dimensionality should decrease the terminate with all possible divided differences with repeated storage error, this suggests that the storage error should instead identical arguments. decrease with increasing kHk. We solve this conundrum by If the contribution to the divided differences is domi- using Davis’ representation [41] of Frechet´ derivatives, which nated by leaves with no repeating indices, the total con- t t+1 reveals the intricate dependence of Frechet´ derivatives on the tribution of such leaves to ht+1 is at most S 2 /δ0, −1 −1 spectral decomposition H = ∑ j≥0 λ jΠ j. Here, λ j strictly in- where S = δ0 + ··· + δt+1. The contribution to ht+1 crease with j and Π j are eigenprojectors. Namely, we can by leaves that terminate with r repeated arguments is [k] r−1 t+1−r write Dg (H,A j)/k! as (τ /(r − 1)!)S /δ0. From this we get ht+1 ≤ θ where

g( ,..., )( A )...( A ) . t+1 t+1 t+1 r−1 ∑ λn0 λnk Πnk j Πn1 j Πn0 (12) S n/2 + 1 r−1 τ t+2−r n0,...,nk θ = + ∑ S . (15) δ0 δ0 r=2 (r − 1)! Here, g(λ ,...,λ ) are divided differences that arise nat- n0 nk This leads us to our final result, which follows directly from urally from the theory of Lagrange interpolation. To un- Lemma 2, (13) and (15). ravel (12), we leverage on the remarkable properties of di- vided differences. First, divided differences are invariant un- Theorem 3. Let H be a mean-field HF where every pair of der any permutation of their arguments. Hence, we can al- qubits interacts with exchange constant J, and C be any PI ways arrange the arguments of a divided difference in non- code that corrects t errors. Let Nτ be the random coher- decreasing order. Second, divided differences generalize ent noise channel as given in (8). Then when we have per- 2 scalar derivatives, because the divided difference of a vec- fect quantum error correction, ε(Nτ ,C ) ≤ Θ + Θ , where tor with k identical arguments is proportional to the (k − Θ = (aN)t+1(θ + τt+1/(t + 1)!), and θ is given in (15). 1)th derivative of the underlying function. For the exponen- k−1 tial function, we have |g(y1,...,yk)| = τ /(k − 1)! when From Theorem 3, we can see that the quantum memory is y1 = ··· = yk. For instance, |g(2,2)| ≤ τ. Third, a di- partially self-correcting, because the bound on the storage er- vided difference when not evaluated on identical arguments ror contains a term (Nτ)t /t! which diverges for large N, since can be recursively defined; whenever yi and y j are distinct, N is quadratic in t. Hence for every noise parameter a and g(y) = (g(y[not i]) − g(y[not j]))/(yi − y j), where y[not i] de- exchange constant J, our scheme for a quantum memory has notes a vector obtained from y by deleting its ith component. an optimal system size. Figure 2 illustrates only results for From (11) and (12), we find that optimal system sizes. Recently, a superconducting qubit was stored between 12ns  τt+1  to 20ns with a fidelity of 0.9995 [42]. Using our noise model, kRk ≤ (aN)t+1 h + . (13) C t+1 (t + 1)! these experimental parameters can be recast into a baseline storage error of 5(10−4) with a memory lifetime of 12ns and Here τt+1/(t + 1)! arises from the divided difference with all a noise strength of a = 0.04MHz. Given this noise model, we arguments equal to zero. The term ht+1 is the sum of all use Theorem 3 to obtain upper bounds on the storage error ε of an encoded qubit within a PI code in a mean-field HF, and we |g(0,λn1 ,...,λnt+1 )| where n1 + ··· + nt+1 > 0. Evaluating a bound on ht+1 requires knowing the eigenval- depict these numerical results in Figure 2. Here, the number of 2 ues of H. Since finding the eigenvalues of H for arbitrary qubits for t = 1 is seven [27], and when t ≥ 2, N = (2t + 1) . E is in general a difficult problem [35], we study an exactly From Figure 2, if one uses a seven-qubit PI code with J = 3 solvable HF where every pair of spins interacts equally with 10 GHz, the qubit’s storage lifetime can be improved to over Ji, j = J. We call such a HF a mean-field HF, and its ground 100ns. In addition, if one uses a 25 qubit PI code that corrects state energy is λ0 = 0 and its higher energy eigenvalues are 2 errors [28], the qubit’s storage lifetime can be enhanced to 4 6 λ1 = JN,λ2 = 2J(N − 1), λ3 = 3J(N − 2), and in general, over 120ns when J = 10 GHZ. Similarly, if J = 10 GHZ, the λ j = J j(N +1− j) [35]. We proceed to bound δ j, which is the qubit’s storage lifetime can be enhanced to over 150ns using a minimum energy needed to transition away from λ j. For in- 49 qubit PI code that corrects 3 errors. From this, we can see stance, δ0 = λ1 −λ0 = JN, δ1 = λ2 −λ1 = J(N −2), and δ2 = how increasing J and the number of qubits in HFs can enhance the storage lifetime. λ3 −λ2 = J(N −4). In general, δbN/2c = 2+(N −2bN/2c) and δ j = J(N − 2 j) for all j = 0,...,bN/2c − 1. Importantly, δ j is Discussions.— Here, we study quantum storage in a phys- non-increasing in j and is maximal when j = 0. Exploiting the ically abundant physical system, the HF. Our scheme, being recursive structure of divided differences, one can show that based on a physical model that is simple to realize, will be easier to implement than those built upon many-body interac- t+1 −1 −1 tions. Because Pauli errors on PI codes can exhibit a macro- |g(0,λn1 ,...,λnt )| ≤ 2 δ0 (δn1 ...δnt+1 ) . (14) scopic energy barrier, we see evidence that a quantum memory When there are repeated arguments for the divided differ- based in a HF can become increasingly robust with increasing ence, the bound is more complicated than (14), because it dimensionality of the HF. In addition, we study the impact of

4 increasing the coupling strengths within HFs on the targeted to implementation. storage error. For this, we analyze an infinite-dimensional HF, namely the mean-field HF, and find numerically tractable up- per bounds on the storage error. Our derivation of the bounds I. ACKNOWLEDGMENTS relies on a novel approach based on the connection between matrix perturbation theory and divided differences. We find Acknowledgements.— YO thanks Joseph F. Fitzsimons for that increasing the size of the coupling strengths can be bene- coming up with the idea for studying Heisenberg ferromagnets ficial in extending the storage lifetime of a HF-based quantum for quantum storage, and also for helping to prove Theorem 1. memory when used in concert with a PI code. YO also thanks Jon Tyson for bringing to his attention the di- Since our analysis technique extends to any physical system vided difference representation of matrix Frechet´ derivatives. with a completely understood spectral structure, it applies also YO acknowledges support from the Singapore National Re- to other code-inspired Hamiltonians, and lays the foundations search Foundation under NRF Award NRF-NRFF2013-01, the for analyzing using quantum memories using our new method- U.S. Air Force Office of Scientific Research under AOARD ology. While PI codes can be robustly prepared in supercon- grant FA2386-18-1-4003. YO acknowledges support from the ducting charge qubits, by inducing Rabi-like oscillations be- EPSRC (Grant No. EP/M024261/1) and the QCDA project tween Dicke states of adjacent weights [43], it remains to see (Grant No. EP/R043825/1) which has received funding from how the initialization Hamiltonian can be integrated with the the QuantERA ERANET Cofund in Quantum Technologies HF. Furthermore, constructing explicit protocols for the de- implemented within the European Unions Horizon 2020 Pro- coding of PI codes can bring quantum memories in HFs closer gramme.

[1] B. M. Terhal, “Quantum error correction for quantum memo- vol. 326, no. 10, pp. 2566–2633, 2011. ries,” Rev. Mod. Phys., vol. 87, pp. 307–346, Apr 2015. [15] S. P. Jordan and E. Farhi, “Perturbative gadgets at arbitrary or- [2] B. J. Brown, D. Loss, J. K. Pachos, C. N. Self, and J. R. ders,” Phys. Rev. A, vol. 77, p. 062329, Jun 2008. Wootton, “Quantum memories at finite temperature,” Reviews [16] S. A. Ocko and B. Yoshida, “Nonperturbative gadget for topo- of Modern Physics, vol. 88, no. 4, p. 045005, 2016. logical quantum codes,” Phys. Rev. Lett., vol. 107, p. 250502, [3] E. Dennis, A. Kitaev, A. Landahl, and J. Preskill, “Topological Dec 2011. quantum memory,” Journal of Mathematical Physics, vol. 43, [17] C. G. Brell, S. T. Flammia, S. D. Bartlett, and A. C. Doherty, no. 9, pp. 4452–4505, 2002. “Toric codes and quantum doubles from two-body Hamiltoni- [4] A. Kitaev, “Anyons in an exactly solved model and beyond,” ans,” New Journal of Physics, vol. 13, no. 5, p. 053039, 2011. Annals of Physics, vol. 321, no. 1, pp. 2 – 111, 2006. January [18] M. Marvian and D. A. Lidar, “Error suppression for hamilto- Special Issue. nian in markovian environments,” Physical [5] A. Kay and R. Colbeck, “Quantum Self-Correcting Stabilizer Review A, vol. 95, no. 3, p. 032302, 2017. Codes, arXiv:0810.3557,” [19] N. Motoyama, H. Eisaki, and S. Uchida, “Magnetic susceptibil- [6] S. Bravyi and B. Terhal, “A no-go theorem for a two- ity of ideal spin 1 /2 heisenberg antiferromagnetic chain sys- dimensional self-correcting quantum memory based on stabi- tems, sr2cuo3 and srcuo2,” Phys. Rev. Lett., vol. 76, pp. 3212– lizer codes,” New Journal of Physics, vol. 11, no. 4, p. 43029, 3215, Apr 1996. 2009. [20] C. Chung, J. Marston, and R. H. McKenzie, “Large-N solu- [7] S. Chesi, B. Rothlisberger,¨ and D. Loss, “Self-correcting quan- tions of the Heisenberg and Hubbard-Heisenberg models on tum memory in a thermal environment,” Phys. Rev. A, vol. 82, the anisotropic triangular lattice: application to Cs2CuCl4 p. 22305, Aug. 2010. and to the layered organic superconductors κ-(BEDT-TTF) 2X [8] S. Chesi, D. Loss, S. Bravyi, and B. M. Terhal, “Thermody- (BEDT-TTF≡ bis (ethylene-dithio) tetrathiafulvalene); X≡ an- namic stability criteria for a quantum memory based on stabi- ion,” Journal of Physics: Condensed Matter, vol. 13, no. 22, lizer and subsystem codes,” New Journal of Physics, vol. 12, p. 5159, 2001. no. 2, p. 025013, 2010. [21] D. Thouless, “Exchange in solid 3He and the Heisenberg Hamil- [9] F. Pastawski, L. Clemente, and J. I. Cirac, “Quantum memo- tonian,” Proceedings of the Physical Society, vol. 86, no. 5, ries based on engineered dissipation,” Phys. Rev. A - Atomic, p. 893, 1965. Molecular, and Optical Physics, vol. 83, pp. 1–14, Oct. 2011. [22] S. Blundell, Magnetism in Condensed Matter. Great Clarendon 1010.2901. Street, Oxford OX2 6DP: Oxford master series in condensed [10] S. Bravyi and J. Haah, “Quantum Self-Correction in the 3D Cu- matter physics, first ed., 2003. bic Code Model,” Phys. Rev. Lett., vol. 111, p. 200501, Nov. [23] G. S. Uhrig, “Keeping a quantum bit alive by optimized π-pulse 2013. sequences,” Phys. Rev. Lett., vol. 98, p. 100504, Mar 2007. [11] C. G. Brell, “A proposal for self-correcting stabilizer quantum [24] L.-M. Duan, E. Demler, and M. D. Lukin, “Controlling spin ex- memories in 3 dimensions (or slightly less),” New Journal of change interactions of ultracold atoms in optical lattices,” Phys- Physics, vol. 18, no. 1, p. 013050, 2016. ical review letters, vol. 91, no. 9, p. 090402, 2003. [12] I. Marvian and D. A. Lidar, “Quantum error suppression with [25] H. Tamura, K. Shiraishi, and H. Takayanagi, “Tunable exchange commuting hamiltonians: Two local is too local,” Phys. Rev. interaction in quantum dot devices,” Japanese journal of applied Lett., vol. 113, p. 260504, Dec 2014. physics, vol. 43, no. 5B, p. L691, 2004. [13] M. B. Hastings, “Topological Order at Nonzero Temperature,” [26] M. B. Ruskai, “Pauli Exchange Errors in Quantum Computa- Phys. Rev. Lett., vol. 107, p. 210501, Nov. 2011. tion,” Phys. Rev. Lett., vol. 85, pp. 194–197, July 2000. [14] B. Yoshida, “Feasibility of self-correcting quantum memory [27] H. Pollatsek and M. B. Ruskai, “Permutationally invariant codes and thermal stability of topological order,” Annals of Physics, for quantum error correction,” Linear Algebra and its Applica-

5 tions, vol. 392, no. 0, pp. 255–288, 2004. Since ρ is permutation-invariant, both πi, jρ and ρπi, j are equal [28] Y. Ouyang, “Permutation-invariant quantum codes,” Phys. Rev. to ρ, and hence A, vol. 90, no. 6, p. 062317, 2014. [29] Y. Ouyang, “Permutation-invariant qudit codes from polynomi- hPi = J ∑ Ti, j, (A1) als,” Linear Algebra and its Applications, vol. 532, pp. 43 – 59, {i, j}∈E 2017. [30] Y. Ouyang and J. Fitzsimons, “Permutation-invariant codes en- where coding more than one qubit,” Phys. Rev. A, vol. 93, p. 042340, Apr 2016. Ti, j = 1 − Tr(ρ(πi, jPπi, j)P). (A2) [31] F. G. S. L. Brandao,˜ E. Crosson, M. B. S¸ahinoglu,˘ and J. Bowen, “Quantum error correcting codes in eigenstates of translation- If swapping the ith and jth qubits leaves P invariant, we have invariant spin chains,” Phys. Rev. Lett., vol. 123, p. 110502, Sep ⊗N (πi, jPπi, j)P = I , and the corresponding term Ti, j is zero. 2019. Now denote Xi, Yi and Zi as Pauli operators that apply a single [32] Y. Ouyang and R. Chao, “Permutation-invariant constant- X, Y and Z Pauli on the ith qubit and leave the remaining qubits excitation quantum codes for amplitude damping,” IEEE Trans- actions on Information Theory, 2019. untouched. [33]G.T oth´ and O. Guhne,¨ “Entanglement and Permutational Sym- Conversely, if swapping the ith and jth spins changes P, the metry,” Phys. Rev. Lett., vol. 102, p. 170503, May 2009. operator (πi, jPπi, j)P is either Xi, j, Yi, j, or Zi, j. Because ρ is [34] M. Bergmann and O. Guhne,¨ “Entanglement criteria for Dicke permutation-invariant, it suffices to evaluate states,” Journal of Physics A: Mathematical and Theoretical, vol. 46, no. 38, p. 385304, 2013. Tr(ρX1,2), Tr(ρY1,2), Tr(ρZ1,2). (A3) [35] Y.Ouyang, “Computing spectral bounds of the heisenberg ferro- magnet from geometric considerations,” Journal of Mathemati- Since ρ is supported on a PI-code that corrects at least a single cal Physics, vol. 60, no. 7, p. 071901, 2019. error, the Knill-Laflamme conditions [44] imply that for all [36] B. Bollobas´ and I. Leader, “Edge-isoperimetric inequalities in Q ∈ Q = {X1,2,Y1,2,Z1,2}, we have h0L|Q|1Li = h1L|Q|0Li = the grid,” Combinatorica, vol. 11, no. 4, pp. 299–314, 1991. 0, and h0L|Q|0Li = h1L|Q|1Li [28]. Hence for all Q ∈ Q and [37] E. Campbell, “A theory of single-shot error correction for ad- ρ supported on our PI-codespace, versarial noise,” Quantum Science and Technology, 2019. [38] R. Bhatia, Matrix analysis, vol. 169. Springer Science & Busi- Tr(ρQ) = h0L|ρ|0Lih0L|Q|0Li + h1L|ρ|1Lih1L|Q|1Li ness Media, 2013. h0 |Q|0 i + h1 |Q|1 i [39] E. Deadman and S. D. Relton, “Taylor’s theorem for matrix = h0 |Q|0 i = L L L L functions with applications to condition number estimation,” L L 2 Linear Algebra and its Applications, vol. 504, pp. 354 – 371, n = 2−n hDm |Q|Dm i. 2016. ∑ ` g` g` [40] Y. Ouyang, D. R. White, and E. Campbell, “Compila- 0≤`≤n tion by stochastic hamiltonian sparsification,” arXiv preprint N  N−2 arXiv:1910.06255, 2019. Using the Vandermonde binomial identity g` = g` + [41] C. Davis, “Explicit functional calculus,” Linear Algebra and its 2N−2 + N−2, it follows that Applications, vol. 6, pp. 193 – 199, 1973. g`−1 g`−2 [42] R. Barends, J. Kelly, A. Megrant, A. Veitia, D. Sank, E. Jef-  1 − 4µ, Q = Z frey, T. C. White, J. Mutus, A. G. Fowler, B. Campbell, Tr(ρQ) = , (A4) Y. Chen, Z. Chen, B. Chiaro, A. Dunsworth, C. Neill, 2µ, Q ∈ {X,Y} P. OaEUR(tm)Malley,ˆ P. Roushan, A. Vainsencher, J. Wenner, A. N. Korotkov, A. N. Cleland, and J. M. Martinis, “Supercon- where for our GNU code with parameters g,n and u, we have ducting quantum circuits at the surface code threshold for fault  N−2 tolerance,” Nature, vol. 508, pp. 500 EP –, Apr 2014. n g`− µ := 2−n 1 . [43] C. Wu, Y. Wang, C. Guo, Y. Ouyang, G. Wang, and X.-L. Feng, ∑ ` N  “Initializing a permutation-invariant quantum error-correction 0≤`≤n g` code,” Phys. Rev. A, vol. 99, p. 012335, Jan 2019. [44] E. Knill and R. Laflamme, “Theory of quantum error-correcting In particular, for all u > 1, we have 0 < µ < 1/2, which im- codes,” Phys. Rev. A, vol. 55, pp. 900–911, Feb. 1997. plies that either Ti, j = 0 or Ti, j is strictly positive. The lower [45] R. S. Varga, Gersgorinˇ and his circles. Springer-Verlag, first ed., bound for hPi then depends crucially on the number of edges 2004. for which Ti, j is strictly positive. This implies that

hPi ≥ J|∂E (V(P))|min{1 − 4µ,2µ}. (A5)

Appendix A: Proof of Theorem 1 Substituting u = 2,g = n = 2t + 1 then gives the result.

We now show that the larger the weight of P, the larger the Appendix B: Proof of Lemma 2 expected energy of PρP, where ρ is supported on our chosen PI codespace. By definition of hPi, we have † Consider N (ρ) = ∑ j p jUjρUj for any density matrix ρ, where p j are probabilities and Uj are unitary matrices. Let hPi = − ∑ J(Tr(PρPπi, j) − Tr(ρπi, j)). C be a quantum code with logical codewords |rLi for r = {i, j}∈E 0,...,d − 1. Let Uj = G j + B j, where G j and B j denote the

6 correctible and uncorrectable parts of Uj respectively with re- where spect to C . It suffices to show that † mg,g = |rLihrL| − ΠgR(U|rLihrL|U )Πg, p d−1 q  † j † † mg,b = ΠgR(U|rLihrL|U )Πb, ε(N ,C ) ≤ ∑ ∑ hrL|B B|rLi + hrL|B B|rLi . d † j r=0 mb,g = ΠbR(U|rLihrL|U )Πg, (B1) † mb,b = ΠbR(U|rLihrL|U )Πb. (C2) We now proceed with the proof. From the linearity of the Let U = G + B where G denotes the correctible part of U and channels and , it is clear that N R B denotes the uncorrectable part of U with respect to the code . Second, since corrects errors on the correctible subspace 1 C R ε(N ,C ) =min |ΨC ihΨC | − R(N (|ΨC ihΨC |)) and its Kraus operators do not induce interactions between the R 2 1 correctible and uncorrectable subspaces, it follows that 1 1 d−1 =min (|rihs| ⊗ |rLihsL| † ∑ ΠgR(U|rLihrL|U )Πg R 2 d r,s=0 † † ! = ΠgR(G|rLihrL|G )Πg + ΠgR(B|rLihrL|G )Πg  † − |rihs| ⊗ p U |r ihs |U . † † ∑ jR j L L j + ΠgR(G|rLihrL|B )Πg + ΠgR(B|rLihrL|B )Πg j 1 † = |rLihrL|hrL|G G|rLi. (C3) Using the triangle inequality for the trace norm and the mono- † tonicity of the trace distance under the partial trace, we get Since hrL|U U|rLi =1, it follows from the definition of G and † † B that hrL|G G|rLi + hrL|B B|rLi = 1. Hence d−1 p j ε( , ) ≤ M , (B2)  † †  N C ∑ ∑ r,Uj 1 hrL|B B|rLi|rLihrL| ΠgR(U|rLihrL|B )Πb j 2d r=0 Mr,U = † † . ΠbR(B|rLihrL|U )Πg ΠbR(B|rLihrL|B )Πb (C4) where

† We now require a block-matrix generalization of the Gersgorin˘ Mr,U = |rLihrL| − R(U|rLihrL|U ). circle theorem [45] given by the following lemma.

It remains to obtain an upper bound on kMr,U k1. Lemma 5. Let M be a Hermitian matrix with the block de- composition Lemma 4.  AC q M = , † † † kMr,U k1 ≤ 2 hrL|B B|rLi + 2hrL|B B|rLi. (B3) C B where A and B are Hermitian matrices of the same size. Then We defer the proof of Lemma 4, which essentially involves decomposing Mr,U into a block matrix on the correctible and kMk1 ≤ kAk1 + kBk1 + 2kCk1. (C5) uncorrectable subspaces of the quantum code C . Thus, it read- ily follows that Proof. Since M is Hermitian, there is a unitary matrix V such that kMk1 = Tr(VM). Such a unitary matrix V admits a block p d−1 q  j † † decomposition ε(N ,C ) ≤ hrL|B B|rLi + hrL|B B|rLi . ∑ d ∑ j r=0 V V  V = 11 12 . (C6) Since probabilities sum to one, we can use Holder’s¨ inequality V21 V22 and find that the maximization over r gives the upper bound Then we can see that required in the lemma. † kMk1 = Tr(V11A +V12C +V21C +V22B). (C7)

Appendix C: Proof of Lemma 4 Since the trace of the product of two matrices is an inner prod- uct, we can use Holder’s¨ inequality to obtain an upper bound on kMk . Such a bound requires knowledge on the operator Let Πg and Πb denote projectors onto the correctible and 1 uncorrectable subspaces of the quantum code C respectively. norm (the maximum singular value) of the submatrices of V. For this, we use the parallelogram law. Note that It follows that Πg|rLi = |rLi and Πb|rLi = 0. Hence, by writing Mr,U into a size 2 block matrix induced by the projectors Πg     x V11x +V12y and Πb we get V = y V21x +V22y       mg,g mg,b x V11x −V12y Mr,U = , (C1) V = . (C8) mb,g mb,b −y V21x −V22y

7 Whenever (x;y) is a vector of unit norm, it follows that the Hence vectors on the right side of the above equation are also vectors of unit norms. Now let us apply the parallelogram law, which q † † gives us kMr,U k1 ≤ 2 hrL|B B|rLi + 2hrL|B B|rLi, (C18)

2 2 2 2 2kV11xk + 2kV12yk = kV11x +V12yk + kV11x −V12yk . (C9) and the result follows.

Since V is unitary and (x;y) is a vector of unit norm, we al- ways have

2 2 2kV11xk + 2kV12yk ≤ 2. (C10) Appendix D: Proof of Eq. (14)

Whenever y = 0, we must have kxk = 1, and in this scenario In Figure 3, we depict g( , , , ) as a binary tree we find that kV xk2 ≤ 1. This implies that kV k ≤ 1. Re- λ0 λ2 λ2 λ3 11 11 with root labeled by g( , , , ). Children of the root peating this argument, we find that the operator norm of all λ0 λ2 λ2 λ3 are obtained by removing the largest and the smallest argu- submatrices of V is at most 1. Hence the result follows from ments from g( , , , ), which are and respectively. applying Holder’s¨ inequality on (C7). λ0 λ2 λ2 λ3 λ0 λ3 g(λ0,λ2,λ2,λ3) Using Lemma 5, we get 3 0 † † kMr,U k1 ≤ hrL|B B|rLi + 2kΠgR(U|rLihrL|B )Πbk1 g(λ2,λ2,λ3) g(λ0,λ2,λ2) † 3 2 + kΠbR(B|rLihrL|B )Πbk1. (C11) 2 0

Now recall that for matrices S and T, kSTk1 ≤ kSkkTk1 and g(λ2, λ3) g(λ2,λ2) g(λ2,λ2) g(λ0, λ2) kSTk1 ≤ kTkkSk1. Using this fact repeatedly, we find that 3 2 2 0 † † kΠgR(U|rLihrL|B )Πbk1 ≤ kR(U|rLihrL|B )k1, (C12) g(λ3) g(λ2) g(λ2) g(λ0) kΠ (B|r ihr |B†)Π k ≤ k (B|r ihr |B†)k . (C13) bR L L b 1 R L L 1 FIG. 3: A tree illustrates the recursive structure of the divided differ- ence of g(λ ,λ ,λ ,λ ), where g is the exponential function. Furthermore, because R is a trace-preserving map, we have 0 2 2 3

† † kΠgR(U|rLihrL|B )Πbk1 ≤ k|rLihrL|B k1, (C14) † † −1 kΠbR(B|rLihrL|B )Πbk1 ≤ kB|rLihrL|B k1. (C15) Hence, the root’s children inherit a factor of (λ3 − λ0) from √ the root. This procedure iterates until no distinct arguments † Using that fact that for a matrix S, we have kSk1 = Tr S S, it in the divided difference remain. We can see that the left- follows that most vertex in Figure 3 is a leaf labeled by g(λ3) that con- tributes at most (λ − λ )−1(λ − λ )−2 to the bound. The q 3 0 3 2 † † g( , ) ( − k|rLihrL|B k1 = Tr B|rLihrL|B blue vertex labeled by λ2 λ2 contributes at most λ3 )−1( − )−1 to the bound. Using the triangle inequal- q λ0 λ3 λ2 τ † = hrL|B B|rLi, (C16) ity and the monotonicity of δi, we find that g(λ0,λ2,λ2,λ3) ≤ 3 −1 −1 −1 2 −1 −1 −1 max{2 δ0 δ2 δ3 ,2 δ0 δ3 τ}. Here, the factor of δ0 −1 and arises as a coarse upper bound to (λ3 − λ0) , while δ2 and q δ3 correspond to the smallest energy gaps obtainable when de- † † † kB|rLihrL|B k1 = Tr B|rLihrL|B B|rLihrL|B scending the tree from the root’s children onwards. General- izing this argument for distinct non-zero λn ,...,λn where = hr |B†B|r i. (C17) 1 t+1 L L n1 > ··· > nt+1 gives the result.

8