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

Polynomial-Time Preparation of Low-Temperature Gibbs States for 2D Toric Code

Zhiyan Ding [email protected]. Department of Mathematics, University of California, Berkeley Bowen Li [email protected]. Department of Mathematics, City University of Hong Kong Lin Lin [email protected]. Department of Mathematics, University of California, Berkeley Applied Mathematics and Computational Research Division, Lawrence Berkeley National Laboratory Ruizhe Zhang [email protected]. Simons Institute for the Theory of Computing
Abstract

We propose a polynomial-time algorithm for preparing the Gibbs state of the two-dimensional toric code Hamiltonian at any temperature, starting from any initial condition, significantly improving upon prior estimates that suggested exponential scaling with inverse temperature. Our approach combines the Lindblad dynamics using a local Davies generator with simple global jump operators to enable efficient transitions between logical sectors. We also prove that the Lindblad dynamics with a digitally implemented low temperature local Davies generator is able to efficiently drive the quantum state towards the ground state manifold. Despite this progress, we explain why protecting quantum information in the 2D toric code with passive dynamics remains challenging.

1 Introduction

The ability (or the lack of it) to efficiently prepare Gibbs states has far-reaching implications in quantum information theory, condensed matter physics, quantum chemistry, statistical mechanics, and optimization. Given a quantum Hamiltonian H2N×2NH\in\mathbb{C}^{2^{N}\times 2^{N}}, we would like to prepare the associated thermal state σβeβH\sigma_{\beta}\propto e^{-\beta H}, where β\beta is the inverse temperature. A number of quantum algorithms have been designed to efficiently prepare high-temperature Gibbs states with a small β=𝒪(poly(N1)\beta=\mathcal{O}(\mathrm{poly}(N^{-1})[44, 18, 52, 26, 1]. However, as the temperature lowers (i.e., β\beta becomes large), the complexity of these algorithms can scale exponentially in the number of qubits NN, rendering them impractical for low-temperature regimes where the Gibbs state has a significant overlap with the ground state of HH.

Recent advancements have rekindled interest in designing quantum Gibbs samplers based on Lindblad dynamics [41, 47, 14, 16, 17, 55, 21]. These algorithms rely on a specific form of open quantum system dynamics to drive the system toward its thermal equilibrium, an idea pioneered by Davies in the 1970s [19, 20]. The efficiency of such algorithms largely depends on the mixing time of the underlying dynamics, which can vary significantly across different systems and different forms of Lindbladians.

The computational complexity of preparing quantum Gibbs states, computing partition functions, and the potential for establishing quantum advantage in these tasks is a topic of ongoing debate in the literature. On one hand, at high enough temperatures, there exists polynomial-time classical algorithms to sample from Gibbs states and to estimate partition functions [7, 40, 56, 10]. On the other hand, in the low-temperature regime, preparing classical Gibbs states is already \NP-hard in the worst case [4, 48], and we do not expect efficient quantum algorithms in these cases. The development of these new Gibbs samplers has also contributed to advancements in our complexity-theoretic understanding [45, 8, 46]. [45] proved that simulating the Lindbladian proposed in [17] to time T=\poly(N)T=\poly(N) at β=Ω(log(N))\beta=\Omega(\log(N)) for a kk-local Hamiltonian is \BQP-complete. [8, 46] constructed a family of kk-local Hamiltonians such that quantum Gibbs sampling at constant temperatures (lower than the classically simulatable threshold) can be efficiently achieved with the block-encoding framework [16], and the task is classically intractable assuming no collapse of the polynomial hierarchy.

None of these constructions imply efficient preparation of Gibbs states at low temperatures. Indeed, when the temperature is sufficiently low, the Gibbs state can exhibit a high overlap with the ground state, and cooling to these temperatures is expected to be \QMA-hard in the worst case. However, it is important to recognize that \QMA-hardness does not preclude the possibility of developing efficient Gibbs samplers for specific Hamiltonians. Consequently, understanding and controlling the mixing time for specific systems (or specific classes of systems) is a fundamental open question in this field and can provide valuable insights into the practical performance of these Gibbs samplers [2, 50, 49, 22, 11, 23, 51, 36, 31, 5, 30, 6, 45].

In this work, we propose a novel Gibbs sampler with nonlocal jump operators, and analyze its convergence rate for preparing low-temperature Gibbs states of the 2D toric code [33], a paradigmatic model in quantum information theory, quantum error correction, and condensed matter physics. In the context of thermalization, when the toric code is exposed to thermal noise modeled by a specific form of Lindbladians called the Davies generator, the seminal work of Alicki et al. [2] showed that the inverse spectral gap (which leads to an upper bound of the mixing time) grows exponentially with inverse temperature β\beta. Using a different argument based on energy barriers, Temme et al. [50, 49] confirmed that the thermalization time (i.e., mixing time of the Lindblad dynamics) of the 2D toric code with local noise should indeed scale exponentially with β\beta. However, it is unknown whether the exp(β)\exp(\beta) factor in the mixing time is unavoidable for all Lindblad dynamics on this problem. This leads to the central question of this work:

Can we design a quantum algorithm, based on Lindblad dynamics, that efficiently prepares the low-temperature Gibbs state of the 2D toric code with a polynomial runtime in both the inverse temperature β\beta and the number of qubits NN?

Notations.

For a finite-dimensional Hilbert space \mathcal{H}, we denote by ()\mathcal{B}(\mathcal{H}) the space of bounded operators with identity element 𝟏{\bf 1}. We let 𝒟():={ρ();ρ0,Tr(ρ)=1}\mathcal{D}(\mathcal{H}):=\{\rho\in\mathcal{B}(\mathcal{H})\,;\ \rho\geq 0\,,\ \operatorname{Tr}(\rho)=1\} be the set of quantum states. Denoting by XX^{\dagger} the adjoint operator of X()X\in\mathcal{B}(\mathcal{H}), we recall the Hilbert–Schmidt (HS) inner product on ()\mathcal{B}(\mathcal{H}): X,Y:=Tr(XY)\langle X,Y\rangle:=\operatorname{Tr}(X^{\dagger}Y). With some abuse of notation, the adjoint of a superoperator Φ:()()\Phi:\mathcal{B}(\mathcal{H})\to\mathcal{B}(\mathcal{H}) for ,\langle\cdot,\cdot\rangle is also denoted by Φ\Phi^{\dagger}. Moreover, {X,Y}=XY+YX\{X,Y\}=XY+YX and [X,Y]=XYYX[X,Y]=XY-YX for X,Y()X,Y\in\mathcal{B}(\mathcal{H}) denote the anti-commutator and commutator, respectively. We will use the standard asymptotic notations: 𝒪\mathcal{O}, Ω\Omega, and Θ\Theta. Precisely, we write f=Ω(g)f=\Omega(g) if g=𝒪(f)g=\mathcal{O}(f), and f=Θ(g)f=\Theta(g) if f=𝒪(g)f=\mathcal{O}(g) and g=𝒪(f)g=\mathcal{O}(f).

1.1 Contribution

We derive a perhaps counterintuitive result regarding the 2D toric code: by employing a Lindbladian composed of a local Davies generator supplemented with simple global jump operators that facilitate transitions between logical sectors to overcome the energy barrier, it is possible to efficiently prepare the Gibbs state of the 2D toric code at any temperature. Furthermore, ignoring the information in the logical space, the local Davies generator alone suffices to efficiently drive the density matrix towards the ground state manifold in the zero temperature limit. This approach circumvents the previously established exponential dependence of the mixing time on β\beta, achieving a polynomial scaling with system size (the number of qubits NN).

To be specific, given β>0\beta>0, we construct the following Davies generator:

β=j=1Nσjx+σjy+σjz:=local full+𝖷1+𝖹1+𝖷2+𝖹2:=global,\mathcal{L}_{\beta}=\underbrace{\sum^{N}_{j=1}\mathcal{L}_{\sigma^{x}_{j}}+\mathcal{L}_{\sigma^{y}_{j}}+\mathcal{L}_{\sigma^{z}_{j}}}_{:=\mathcal{L}_{\text{\rm local full}}}+\underbrace{\mathcal{L}_{\mathsf{X}_{1}}+\mathcal{L}_{\mathsf{Z}_{1}}+\mathcal{L}_{\mathsf{X}_{2}}+\mathcal{L}_{\mathsf{Z}_{2}}}_{:=\mathcal{L}_{\rm global}}\,, (1.1)

where []\mathcal{L}_{[\cdot]} is the standard Davies generator defined via Eq. 2.4 based on 2D toric code Hamiltonian (4.1), and 𝖷1,𝖹1,𝖷2,𝖹2\mathsf{X}_{1},\mathsf{Z}_{1},\mathsf{X}_{2},\mathsf{Z}_{2} are global logic operators for 2D toric code, see Eq. 4.4 in Section 4. Then, we prove the following main results:

Theorem 1 (Fast mixing of 2D toric code).

The spectral gap of the Gibbs sampler β\mathcal{L}_{\beta} in (1.1) has the following lower bound

Gap(β)=max{e𝒪(β),Ω(N3)}.{\rm Gap}(-\mathcal{L}_{\beta})=\max\left\{e^{-\mathcal{O}(\beta)},\Omega(N^{-3})\right\}\,.

Our proof is valid for all temperatures, which includes two important regimes: 1. finite temperature regime β=𝒪(1)\beta=\mathcal{O}(1); 2. low temperature regime β1\beta\gg 1. In the first case of β=𝒪(1)\beta=\mathcal{O}(1), the local Davies generator local full\mathcal{L}_{\text{\rm local full}} is sufficient to ensure the spectral gap of β\mathcal{L}_{\beta} at least exp(Θ(β))\exp(-\Theta(\beta)) and independent of NN. This is consistent with the result in [2, Theorem 2]. In the second case of β1\beta\gg 1, the Davies generator with global jumps global\mathcal{L}_{\rm global} significantly increases the spectral gap of β\mathcal{L}_{\beta}, achieving a lower bound of order poly(N1)\mathrm{poly}\left(N^{-1}\right) and independent of β\beta.

The choice of the global jump operators in (1.1) is natural. As β\beta\to\infty, the thermal state converges toward the ground state of the Hamiltonian. Since the 2D toric code has four degenerate ground states that are not locally connected, the local Davies generator local full\mathcal{L}_{\text{\rm local full}} cannot efficiently transit between these ground states, resulting in a slow mixing process. To overcome this difficulty, we introduce the global jump operators global\mathcal{L}_{\rm global} that enable transitions between different ground states at low temperatures, which ensures fast mixing even at zero temperature. This phenomenon is not unique to the 2D toric code. For example, in our paper, we also consider a simpler 1D Ising model in Appendix A, construct a similar Davies generator with global jump operators, and demonstrate a fast mixing result similar to 1 to illustrate the proof concept.

Our proof implies a more detailed characterization of the mixing process. The algebra of observables for 2D toric code can be expressed as a tensor product between a logical observable space and a syndrome space. The spectral gap of the Lindbladian in the logical space and the syndrome space can be analyzed independently. We find that for the standard Davies generator with only local jump operators, the exponentially vanishing spectral gap on β\beta is only due to the action on the logical space, where the 2D toric code has four linearly independent ground states that are not locally connected. On the other hand, we show that the spectral gap of the local Davies generator, when restricted to the syndrome space (by tracing out the logical subspace), has a lower bound that decays polynomially with the size of the system and remains independent of β\beta in the low-temperature regime. This is summarized in the following proposition:

Proposition 2.

Let local full\mathcal{L}_{\text{\rm local full}} be defined as in Eq. (1.1). Then a part of local full\mathcal{L}_{\text{\rm local full}} admits the syndrome subspace as an invariant subspace and exhibits a “large” spectral gap.

Specifically, there is a subset of Paulis {pi}{σjx/y/z}\{p_{i}\}\subset\{\sigma^{x/y/z}_{j}\}, which defines local=ipi\mathcal{L}_{\rm local}=\sum_{i}\mathcal{L}_{p_{i}}, a decomposition =2N=logicsyndrome\mathcal{H}=\mathbb{C}^{2^{N}}=\mathcal{H}_{\rm logic}\otimes\mathcal{H}_{\rm syndrome} with logic4\mathcal{H}_{\rm logic}\cong\mathbb{C}^{4} and syndrome2N2\mathcal{H}_{\rm syndrome}\cong\mathbb{C}^{2^{N-2}}, and a corresponding decomposition ()(logic)(syndrome)\mathcal{B}(\mathcal{H})\cong\mathcal{B}\left(\mathcal{H}_{\rm logic}\right)\otimes\mathcal{B}\left(\mathcal{H}_{\rm syndrome}\right), such that

local((logic)(syndrome))=(logic)local((syndrome)).\mathcal{L}_{\rm local}\left(\mathcal{B}\left(\mathcal{H}_{\rm logic}\right)\otimes\mathcal{B}\left(\mathcal{H}_{\rm syndrome}\right)\right)=\mathcal{B}\left(\mathcal{H}_{\rm logic}\right)\otimes\mathcal{L}_{\rm local}\left(\mathcal{B}\left(\mathcal{H}_{\rm syndrome}\right)\right)\,.

The spectral gap of local-\mathcal{L}_{\rm local} restricted to the syndrome space (syndrome)\mathcal{B}\left(\mathcal{H}_{\rm syndrome}\right) is lower bounded by

max{exp(𝒪(β)),Ω(N3)}.\max\left\{\exp(-\mathcal{O}(\beta)),\Omega(N^{-3})\right\}\,.

The above proposition implies that, even in the absence of the global jump operator, the system thermalizes quickly within the syndrome space. This is a key step towards proving 1; see Section 3 for further details. Here the operator local\mathcal{L}_{\text{local}} in 2 is carefully designed so that analyzing syndrome mixing reduces to studying the spectral gap of two distinct Glauber dynamics with quasi-1D classial Ising Hamiltonians (precisely, on the “snake” and “comb”, see Fig. 3) at low temperature. We then introduce a new iterative method to establish lower bounds of their spectral gaps. To the best of our knowledge, this result concerning low-temperature thermalization for such quasi-1D classical Ising models is also novel in the literature.

1.2 Implications

Mixing time for preparing low-temperature Gibbs state.

In this paper, we mainly focus on estimating the spectral gap of the Lindblad generator. It is well known that a lower bound on the spectral gap provides an upper bound on the mixing time of the Lindblad dynamics [51]. Specifically, for a detailed balanced Lindblad generator \mathcal{L} with a spectral gap lower bounded by α\alpha and the unique fixed point σβ\sigma_{\beta}, the following holds:

etρσβtr2χ2(etρ,σβ)χ2(ρ,σβ)e2αt,\|e^{t\mathcal{L}^{\dagger}}\rho-\sigma_{\beta}\|^{2}_{\mathrm{tr}}\leq\chi^{2}(e^{t\mathcal{L}^{\dagger}}\rho,\sigma_{\beta})\leq\chi^{2}(\rho,\sigma_{\beta})e^{-2\alpha t},

where χ2(ρ,σβ)=Tr[(ρσβ)σβ1/2(ρσβ)σβ1/2]\chi^{2}(\rho,\sigma_{\beta})=\mathrm{Tr}[(\rho-\sigma_{\beta})\sigma_{\beta}^{-1/2}(\rho-\sigma_{\beta})\sigma_{\beta}^{-1/2}] is the χ2\chi^{2}-divergence. Notice that

maxρχ2(ρ,σβ)(λmin(σβ))12NeβH,\max_{\rho}\chi^{2}(\rho,\sigma_{\beta})\leq(\lambda_{\min}(\sigma_{\beta}))^{-1}\leq 2^{N}e^{\beta\|H\|}\,,

with λmin()\lambda_{\min}(\cdot) denoting the minimal eigenvalue. For 2D toric code Hamiltonian, we have the operator norm H=𝒪(N)\|H\|=\mathcal{O}(N). Thus, it readily gives an upper bound of the mixing time tmix(ϵ):={t0;etρσβtrϵ,quantum states ρ}t_{\mathrm{mix}}(\epsilon):=\{t\geq 0\,;\ \|e^{t\mathcal{L^{{\dagger}}}}\rho-\sigma_{\beta}\|_{\rm tr}\leq\epsilon,\ \forall\,\text{quantum states $\rho$}\} for a fixed ϵ\epsilon:

tmix(ϵ)=𝒪(1α(N+β)).t_{\rm mix}(\epsilon)=\mathcal{O}\left(\frac{1}{\alpha}\left(N+\beta\right)\right)\,.

By substituting the spectral gap estimate from 1, the mixing time of β\mathcal{L}_{\beta} in Eq. 1.1 scales as 𝒪(min{βpoly(N),exp(cβ)(N+β)})\mathcal{O}\left(\min\left\{\beta\mathrm{poly}(N),\exp(c\beta)(N+\beta)\right\}\right) for some universal constant cc, which is a significant improvement over the 𝒪(exp(cβ)(N+β))\mathcal{O}\left(\exp(c\beta)(N+\beta)\right) bound given in [2] when β1\beta\gg 1. It is worth mentioning that in both cases, the mixing time scales polynomially with the number of qubits, NN, which is referred to as fast mixing in the literature. In our case, the NN dependence arises from both the spectral gap and the prefactor maxρχ2(ρ,σβ)\max_{\rho}\chi^{2}(\rho,\sigma_{\beta}). While the latter dependence can sometimes be improved to 𝒪(log(N))\mathcal{O}(\log(N)) by considering the relative entropy distance [5, 6, 30], achieving the rapid mixing regime. This would also require translating the spectral gap lower bound into a constant or Ω(log(N)1)\Omega(\log(N)^{-1}) lower bound for the modified logarithmic Sobolev inequality (MLSI) constant. This presents an interesting direction for future exploration.

Thermal state versus ground state preparation:

The 2D toric code is a stabilizer Hamiltonian, and its ground state can be efficiently prepared by measuring all stabilizers. The ground state can also be prepared using carefully designed dissipative state engineering approaches (see e.g. [53]). However, these approaches do not easily generalize to algorithms for thermal state preparation. The fast mixing result at low temperatures in this work suggests that thermal state preparation methods can also serve as a generic tool for approximate ground state preparation by selecting a sufficiently low temperature. Specifically, as discussed in 2, the local Davies generator within the syndrome space has a spectral gap that decays only polynomially with system size as β\beta approaches infinity. This implies that at fixed sufficiently low temperature (β1\beta\gg 1), once a quasi-particle pair (an elementary excited state of the 2D toric code, see Section 4.1) appears in the system, the local Davies generator can eliminate it in 𝒪(poly(N))\mathcal{O}(\text{poly}(N)) time. Consequently, if the goal is to efficiently prepare some ground state while discarding logical information, it is sufficient to use local Davies generators, and the time required is much shorter than that for equilibrating all logical sectors using local Davies generators. This result may be of independent interest, as nature itself often prepares ground states by cooling.

Fast thermalization from a ground state:

A natural question regarding the thermalization of the toric code Hamiltonian is: if one starts from the ground state, doesn’t it take exponential time in β\beta to create even one quasi-particle pair in the first place? While this is correct, and may seem paradoxical, it does not contradict our main result that the thermalization time scales polynomially in β\beta. The reason is that although creating one quasi-particle pair from the ground state takes an exponential amount of time in β\beta to create one quasi-particle pair from the ground state, the fraction of the excited state in the thermal state is exponentially small in β\beta. In the case of 2D toric code, the slow thermalization of local Davies generators at low temperatures mainly stems from the need to achieve an equal population across all ground states. In other words, the challenge of thermalization lies mainly in transitioning between orthogonal ground states, as discussed in Section 1.1. In our work, we prove that spectral gap of the global jump operators restricted to the logical space is, independent of β\beta at low temperatures. This allows the system to equilibrate rapidly, even starting from a ground state.

Fast thermal state preparation versus quantum memory:

A good self correcting quantum memory (SCQM) should be able to store a quantum state in contact with a cool thermal bath for a duration that increases exponentially with the size of the system. If the thermalization time of a quantum system only scales polynomially with system size, it cannot be considered a viable candidate for SCQM. Ref. [2] demonstrated that, at any constant temperature, the 2D toric code has a spectral gap that remains independent of system size, and thus is not a good SCQM, and to date valid candidates for SCQM are only known in 4D or higher [3]. Our refined estimate indicates that local Davies generator at low temperature (or even at zero temperature) can be efficient in annihilating quasi-particles in the syndrome space.

A natural question is: does it make the 2D toric code a candidate for passively protected quantum memories (sometimes known as autonomous quantum error correction, autonomous quantum memory protection) [13, 37, 38]? Specifically, consider a Lindbladian =e+r\mathcal{L}=\mathcal{L}_{e}+\mathcal{L}_{r} where e\mathcal{L}_{e} is a Davies generator modeling thermal noise at some finite temperature β1\beta^{-1}, and r\mathcal{L}_{r} is a Davies generator at near-zero temperature (implemented either digitally or using analogue devices), and the number of terms (each of up to unit strength) in r\mathcal{L}_{r} can scale polynomially in NN. Starting from a pure ground state ρ0\rho_{0} carrying well-defined logical information, we run the dynamics for some fixed time tt, and then apply a single round of decoding map d\mathcal{E}_{d} to obtain a final density matrix ρf\rho_{f}. Can we ensure that the trace distance between ρ0\rho_{0} and ρf\rho_{f} decreases super-polynomially in NN? Unfortunately, the answer is very likely negative for the 2D toric code. The reason is that once e\mathcal{L}_{e} creates a syndrome in the form of a quasi-particle excitation, this quasi-particle can be diffused by either e\mathcal{L}_{e} or r\mathcal{L}_{r} at zero energy cost. As a result, in the worst case, a quasi-particle may diffuse across the torus in polynomial time before it is annihilated, which changes the logical information. In this sense, the key difference between protecting logical information through passive dynamics and using quantum error correction is that the latter actively guides the diffusion and annihilation of quasi-particles in a way that preserves the logical information.

Notably, this problem does not arise in 2D Ising models for which local excitations cannot diffuse without incurring an energy cost. This is a key aspect in recent designs of passively protected quantum memories [38]. To our knowledge, the theoretical analysis of such models remains an open question.

1.3 Related works

The thermalization of stabilizer Hamiltonians using a local Davies generator has been explored in several prior works [2, 50, 49, 22, 11, 23]. In [2], the authors demonstrated that the local Davies generator achieves fast thermalization for the 2D toric code. In particular, the spectral gap of the Davies generator, when considering all local Pauli coupling operators, is lower bounded by exp(Θ(β))\exp(-\Theta(\beta)). The bound is valid for all temperatures and results in a mixing time (defined via the trace distance) scaling as tmix=Nexp(Θ(β))t_{\rm mix}=N\exp(\Theta(\beta)). Along the same direction, [50, 49] considered general stabilizer codes and establish a lower bound on the spectral gap (or Poincaré constant) using the generalized energy barrier ϵ¯\overline{\epsilon} [49, Definition 13]. Specifically, for any given β>0\beta>0, the spectral gap can be lower bounded by CNexp(βϵ¯)C_{N}\exp\left(-\beta\overline{\epsilon}\right), where CNC_{N} is a technical constant that typically scales as 1/N1/N. Although these works address thermalization across all temperatures, these lower bounds on the spectral gap are insufficient for efficiently preparing low-temperature thermal states, as the gap decays exponentially in β\beta. Specifically, when using a local Davies generator to transition between ground states along a local Pauli path, the energy must first increase, requiring the dynamics to overcome the energy barrier to fully thermalize. Furthermore, as analyzed in [49, 34], an exponentially small spectral gap of order exp(Θ(β))\exp(-\Theta(\beta)) in a local Davies generator appears inevitable when such energy barriers are present.

In our work, to overcome the bottleneck posed by the energy barrier, we modify the local Davies generator by incorporating appropriate global coupling operators that can directly connect the degenerate ground states, thereby avoiding the issues associated with the generalized energy barrier defined by local Pauli paths. By refining the analysis in [2], we demonstrate that the resulting dynamics exhibits a spectral gap that decays polynomially with the size of the system but remains independent of β\beta, ensuring fast mixing even at low temperatures. We note that the polynomially decaying spectral gap primarily arises from the mixing rate within the syndrome space, which remains unaffected by the introduction of the new global jump operators acting on the logical space. Consequently, our refined mixing time estimate rigorously demonstrates that the local Davies generator at low temperatures (and even at zero temperature) can efficiently annihilate quasi-particles residing in the syndrome space. A similar phenomenon has been numerically observed in the study of quantum memory [22, 11, 23].

There is extensive literature on the mixing properties of local Davies generators for general local commuting Hamiltonians [31, 5, 30, 6]. However, most studies consider a fixed finite temperature β=𝒪(1)\beta=\mathcal{O}(1) [31, 5, 6, 30], particularly in the context of 1D local commuting Hamiltonians, where the mixing time implicitly depends on β\beta, or a high temperature β1\beta\ll 1 [31, 30]. Extending these general approaches to low-temperature thermal state preparation and explicitly calculating the temperature dependence remains an interesting and challenging problem.

There is also a long line of works studying the fast mixing and rapid mixing of different types of classical Markov chains for spin systems. However, many classical results suffer from low-temperature bottlenecks: the dynamics can mix in polynomial time only when the inverse temperature β\beta is below some threshold βc\beta_{c}. For example, it is well-known that the Glauber dynamics for the ferromagnetic Ising model with NN spins on a dd-dimensional lattice has mixing time Θ(NlogN)\Theta(N\log N) when β<βc(d)\beta<\beta_{c}(d) [42, 39], and eΘ(N11/d)e^{\Theta(N^{1-1/d})} when β>βc(d)\beta>\beta_{c}(d) [43], where NN is the total number of spins and the constant in the Θ\Theta notation depends on β\beta. Some classical results managed to overcome this bottleneck by carefully designing the initial distribution [25], or studying other Markov chains or other graphical models [29, 24, 27, 28, 9, 15]. However, we note that these classical techniques cannot be applied directly to obtain our results.

1.4 Organization

In the following part of the paper, we start with a brief introduction to the Davies generator and properties of its spectral gap in Section 2 . A technical overview of the proof of 1 is provided in Section 3. The detailed introduction to the 2D toric code and its proof can be found in Section 4. Additionally, in Appendix A, we discuss a simpler case of the 1D ferromagnetic Ising chain for completeness.

Acknowledgment

This material is partially supported by the U.S. Department of Energy, Office of Science, National Quantum Information Science Research Centers, Quantum Systems Accelerator (Z.D.), by the Challenge Institute for Quantum Computation (CIQC) funded by National Science Foundation (NSF) through grant number OMA-2016245 (L.L.), and by DOE Grant No. DE-SC0024124 (R.Z.). L.L. is a Simons Investigator in Mathematics. We thank Garnet Chan, Li Gao, Yunchao Liu, John Preskill for helpful discussions and feedbacks.

2 Preliminaries

Let HH be a quantum many-body Hamiltonian on 2N\mathcal{H}\cong\mathbb{C}^{2^{N}} and σβ=eβH/𝒵β\sigma_{\beta}=e^{-\beta H}/\mathcal{Z}_{\beta} be the associated thermal state, where β\beta is the inverse temperature and 𝒵β=Tr(eβH)\mathcal{Z}_{\beta}=\operatorname{Tr}(e^{-\beta H}) is the partition function. In this section, we shall recall the canonical form of the Davies generator with σβ\sigma_{\beta} being the invariant state and some basic facts for its spectral gap analysis.

A superoperator Φ:()()\Phi:\mathcal{B}(\mathcal{H})\to\mathcal{B}(\mathcal{H}) is a quantum channel if it is completely positive and trace preserving (CPTP). Lindblad dynamics is a C0C_{0}-semigroup of quantum channels with the generator defined by (ρ):=limt0+t1(𝒫t(ρ)ρ)\mathcal{L}^{\dagger}(\rho):=\lim_{t\to 0^{+}}t^{-1}(\mathcal{P}^{\dagger}_{t}(\rho)-\rho) for ρ𝒟()\rho\in\mathcal{D}(\mathcal{H}). Here and in what follows, we adopt the convention that the adjoint operators (i.e., those with {\dagger}) are the maps in Schrödinger picture acting on quantum states. Both generators \mathcal{L} and \mathcal{L}^{\dagger} are usually referred to as Lindbladian. Davies generator is a special class of Lindbladians derived from the weak coupling limit of open quantum dynamics with a large thermal bath.

We first introduce the Bohr frequencies of HH by

BH:={ω=λiλj:λi,λjSpec(H)},B_{H}:=\{\omega=\lambda_{i}-\lambda_{j}\,:\ \forall\leavevmode\nobreak\ \lambda_{i},\lambda_{j}\in{\rm Spec}(H)\}\,,

where Spec(H){\rm Spec}(H) is the spectral set of HH. Let {Sa}a𝒜\{S_{a}\}_{a\in\mathcal{A}} be a set of coupling operators with 𝒜\mathcal{A} being a finite index set that satisfies

{Sa}a𝒜={Sa}a𝒜.\{S_{a}\}_{a\in\mathcal{A}}=\{S_{a}^{\dagger}\}_{a\in\mathcal{A}}\,. (2.1)

The jump operators {Sa(ω)}a,ω\{S_{a}(\omega)\}_{a,\omega} for the Davies generator are given by the Fourier components of the Heisenberg evolution of SaS_{a}:

eiHtSaeiHt=ωBHeiωtλiλj=ωPiSaPj:=ωBHeiωtSa(ω).e^{iHt}S_{a}e^{-iHt}=\sum_{\omega\in B_{H}}e^{i\omega t}\sum_{\lambda_{i}-\lambda_{j}=\omega}P_{i}S_{a}P_{j}:=\sum_{\omega\in B_{H}}e^{i\omega t}S_{a}(\omega)\,. (2.2)

where Pi/jP_{i/j} is the projection into the eigenspace λi/j\lambda_{i/j}. By (2.1), we have Sa(ω)=Sa(ω)S_{a}(\omega)^{\dagger}=S_{a}(-\omega) for any ωBH\omega\in B_{H}.

We then introduce the Davies generator in the Heisenberg picture:

β(X):=a𝒜Sa(X),X(),\mathcal{L}_{\beta}(X):=\sum_{a\in\mathcal{A}}\mathcal{L}_{S_{a}}(X)\,,\quad X\in\mathcal{B}(\mathcal{H})\,, (2.3)

with

Sa(X):=ωBHγa(ω)(Sa(ω)XSa(ω)12{Sa(ω)Sa(ω),X}),\mathcal{L}_{S_{a}}(X):=\sum_{\omega\in B_{H}}\gamma_{a}(\omega)\left(S_{a}(\omega)^{\dagger}\,X\,S_{a}(\omega)-\frac{1}{2}\left\{S_{a}(\omega)^{\dagger}S_{a}(\omega),X\right\}\right)\,, (2.4)

where the transition rate function γa(ω)>0\gamma_{a}(\omega)>0 is given by the Fourier transform of the bath auto-correlation function satisfying the KMS condition [32]:

γa(ω)=eβωγa(ω).\gamma_{a}(-\omega)=e^{\beta\omega}\gamma_{a}(\omega)\,.

In this work, we always choose the transition rate function γa(ω)\gamma_{a}(\omega) as the Glauber form:

γa(ω)=2eβω+1,\gamma_{a}(\omega)=\frac{2}{e^{\beta\omega}+1}\,, (2.5)

For later use, we define ga(ω):=eβω/2γa(ω)g_{a}(\omega):=e^{\beta\omega/2}\gamma_{a}(\omega) and find ga(ω)=ga(ω)g_{a}(\omega)=g_{a}(-\omega). Then, letting

La(ω)=ga(ω)Sa(ω),L_{a}(\omega)=\sqrt{g_{a}(\omega)}S_{a}(\omega)\,, (2.6)

we reformulate the Davies generator (2.3)–(2.4) as follows:

β(X)\displaystyle\mathcal{L}_{\beta}(X) =a𝒜ωBHeβω/2(La(ω)XLa(ω)12{La(ω)La(ω),X})\displaystyle=\sum_{a\in\mathcal{A}}\sum_{\omega\in B_{H}}e^{-\beta\omega/2}\Big{(}L_{a}(\omega)^{\dagger}XL_{a}(\omega)-\frac{1}{2}\left\{L_{a}(\omega)^{\dagger}L_{a}(\omega),X\right\}\Big{)} (2.7)
=12a𝒜ωBHeβω/2La(ω)[X,La(ω)]+eβω/2[La(ω),X]La(ω)\displaystyle=\frac{1}{2}\sum_{a\in\mathcal{A}}\sum_{\omega\in B_{H}}e^{-\beta\omega/2}L_{a}(\omega)^{\dagger}[X,L_{a}(\omega)]+e^{\beta\omega/2}[L_{a}(\omega),X]L_{a}(\omega)^{\dagger}
=12a𝒜ωBHγa(ω)Sa(ω)[X,Sa(ω)]+γa(ω)[Sa(ω),X]Sa(ω).\displaystyle=\frac{1}{2}\sum_{a\in\mathcal{A}}\sum_{\omega\in B_{H}}\gamma_{a}(\omega)S_{a}(\omega)^{\dagger}[X,S_{a}(\omega)]+\gamma_{a}(-\omega)[S_{a}(\omega),X]S_{a}(\omega)^{\dagger}\,.

where the second step follows from

La(ω)=eβω/4γa(ω)Sa(ω)=eβω/4eβωγa(ω)Sa(ω)=La(ω).\displaystyle L_{a}(\omega)^{\dagger}=e^{\beta\omega/4}\sqrt{\gamma_{a}(\omega)}S_{a}(-\omega)=e^{\beta\omega/4}\sqrt{e^{-\beta\omega}\gamma_{a}(-\omega)}S_{a}(-\omega)=L_{a}(-\omega)\,.

We now define the GNS inner product associated with the Gibbs state σβ\sigma_{\beta}:

Y,Xσβ=Tr(YXσβ).\langle Y,X\rangle_{\sigma_{\beta}}=\operatorname{Tr}(Y^{\dagger}X\sigma_{\beta})\,. (2.8)

It is known [32, 21] that the Davies generator satisfies the GNS detailed balance:

Y,β(X)σβ=β(Y),Xσβ,\langle Y,\mathcal{L}_{\beta}(X)\rangle_{\sigma_{\beta}}=\langle\mathcal{L}_{\beta}(Y),X\rangle_{\sigma_{\beta}}\,,

and thus the associated Lindblad dynamics etβe^{t\mathcal{L}_{\beta}^{\dagger}} admits σβ\sigma_{\beta} as an invariant state, i.e.,

β(σβ)=0.\mathcal{L}_{\beta}^{\dagger}(\sigma_{\beta})=0\,.

It follows that β\mathcal{L}_{\beta} is similar to a self-adjoint operator for the HS inner product, called the master Hamiltonian, and has only real spectrum. To be precise, we introduce the transform φX:=Xσβ1/2\varphi_{X}:=X{\sigma_{\beta}}^{1/2}, which gives Y,Xσβ=φY,φX\langle Y,X\rangle_{\sigma_{\beta}}=\langle\varphi_{Y},\varphi_{X}\rangle. Then, the master Hamiltonian ~β\widetilde{\mathcal{L}}_{\beta} is given by the similar transform of β\mathcal{L}_{\beta} via φX\varphi_{X}:

~β:=φβφ1,\widetilde{\mathcal{L}}_{\beta}:=\varphi\circ\mathcal{L}_{\beta}\circ\varphi^{-1}\,, (2.9)

satisfying Y,β(X)σβ=φY,~βφX\langle Y,\mathcal{L}_{\beta}(X)\rangle_{\sigma_{\beta}}=\langle\varphi_{Y},\widetilde{\mathcal{L}}_{\beta}\varphi_{X}\rangle. It is easy to see that ~β-\widetilde{\mathcal{L}}_{\beta} is positive semi-definite and σβ\sqrt{\sigma_{\beta}} is the zero-energy ground state of ~β-\widetilde{\mathcal{L}}_{\beta}. Thus, the spectral gap of β\mathcal{L}_{\beta} is the same as the ground state spectral gap of the Hamiltonian ~β-\widetilde{\mathcal{L}}_{\beta}.

We say that β\mathcal{L}_{\beta} is primitive if σβ\sigma_{\beta} is the unique invariant state; equivalently, the kernel ker(β)\ker(\mathcal{L}_{\beta}) is of one dimension, spanned by 𝟏{\bf 1}. In this case, we have

limtetβ(ρ)=σβ,ρ𝒟(),\displaystyle\lim_{t\to\infty}e^{t\mathcal{L}_{\beta}^{\dagger}}(\rho)=\sigma_{\beta}\,,\quad\forall\rho\in\mathcal{D}(\mathcal{H})\,,

and the spectral gap Gap(β){\rm Gap}(\mathcal{L}_{\beta}) of the primitive Davies generator can be characterized by the variational form:

Gap(β)=infX0,Tr(σβX)=0X,β(X)σβX,Xσβ.{\rm Gap}(\mathcal{L}_{\beta})=\inf_{X\neq 0\,,\operatorname{Tr}(\sigma_{\beta}X)=0}\frac{\langle X,\mathcal{L}_{\beta}(X)\rangle_{\sigma_{\beta}}}{\langle X,X\rangle_{\sigma_{\beta}}}\,.

Moreover, thanks to the detailed balance, the operator norm of β\mathcal{L}_{\beta} can be computed by

βσβσβ=supX0X,β(X)σβX,Xσβ.\left\lVert\mathcal{L}_{\beta}\right\rVert_{\sigma_{\beta}\to\sigma_{\beta}}=\sup_{X\neq 0}\frac{\langle X,\mathcal{L}_{\beta}(X)\rangle_{\sigma_{\beta}}}{\langle X,X\rangle_{\sigma_{\beta}}}\,.

By [54, 57], a sufficient and necessary condition for the primitivity of β\mathcal{L}_{\beta} is the \mathbb{C}-algebra generated by all the jump operators {Sa(ω)}a,w\{S_{a}(\omega)\}_{a,w} is the whole algebra ()\mathcal{B}(\mathcal{H}). By this condition, one can check that for the choice of {Sa}a𝒜={σix,σiy,σiz}i=1N\{S_{a}\}_{a\in\mathcal{A}}=\{\sigma_{i}^{x},\sigma_{i}^{y},\sigma_{i}^{z}\}_{i=1}^{N}, the associated Davies generator is always primitive. Without loss of generality, when discussing the Gibbs samplers in Appendix A and Section 4, we always let {σix,σiy,σiz}i=1N\{\sigma_{i}^{x},\sigma_{i}^{y},\sigma_{i}^{z}\}_{i=1}^{N} be a subset of {Sa}a𝒜\{S_{a}\}_{a\in\mathcal{A}} to guarantee the primitivity.

The following lemmas are collected from [2, Lemmas 1 and 2] with proof omitted, which shall be used repeatedly in our subsequent spectral gap analysis.

Lemma 3.

Let A,B()A,B\in\mathcal{B}(\mathcal{H}) be positive operators on a Hilbert space \mathcal{H}, i.e., A,B0A,B\geq 0. The spectral gaps of AA and BB are their smallest non-zero positive eigenvalues, denoted by Gap(A){\rm Gap}(A) and Gap(B){\rm Gap}(B), respectively. We have:

  • If AA has a non-trivial kernel and A2gAA^{2}\geq gA for some real g>0g>0, then

    Gap(A)g.{\rm Gap}(A)\geq g\,. (2.10)
  • If ker(A+B)\ker(A+B) is non-trivial such that ker(A+B)=ker(B)\ker(A+B)=\ker(B), then

    Gap(A+B)Gap(B).\mathrm{Gap}(A+B)\geq\mathrm{Gap}(B)\,. (2.11)
  • If AA and BB are commuting and ker(A+B)\ker(A+B) is non-trivial, then

    Gap(A+B)min{Gap(A),Gap(B)}.\mathrm{Gap}(A+B)\geq\min\{\mathrm{Gap}(A),\mathrm{Gap}(B)\}\,. (2.12)
  • If AA has gap lower bound gAg_{A} and φ,BφgB\langle\varphi,B\varphi\rangle\geq g_{B} for all normalized φker(A)\varphi\in\ker(A), then

    A+BgAgBgA+B,A+B\succeq\frac{g_{A}g_{B}}{g_{A}+\left\lVert B\right\rVert}\,, (2.13)

    where B\left\lVert B\right\rVert denotes the operator norm of BB.

Remark 4.

The second statement in Lemma 3 means that for any primitive Davies generator with σβ\sigma_{\beta} being invariant, adding any other Davies generator keeping the invariant state can only increase the spectral gap.

3 Technical overview

In this section, we provide a technical overview of the proof of 1. In the analysis, there are three main steps:

  1. (a)

    Decompose \mathcal{H} into logic and syndrome subspaces and the observable algebra ()\mathcal{B}(\mathcal{H}) correspondingly.

  2. (b)

    Demonstrate efficient transition between logic subspaces.

  3. (c)

    Demonstrate fast mixing inside the syndrome subspace.

The decomposition in the first step leverages the special structure of the stabilizer Hamiltonian, following the approach outlined in previous work by [2]. Specifically, for the 2D toric code, we decompose =logicsyndrome\mathcal{H}=\mathcal{H}_{\rm logic}\otimes\mathcal{H}_{\rm syndrome} according to the eigendecomposition of HtoricH^{\rm toric}. Since HtoricH^{\rm toric} has a four-dimensional ground state space, it encodes two logical qubits, making dim(logic)=4\dim(\mathcal{H}_{\rm logic})=4. The syndrome subspace syndrome\mathcal{H}_{\rm syndrome} is then spanned by the electric and magnetic excited states, which are characterized by the bond configurations of the local observables in HtoricH^{\rm toric}. That is, it can be further decomposed into the electric and magnetic excited subspaces syndrome=bmbe\mathcal{H}_{\rm syndrome}=\mathcal{H}^{\rm m}_{\rm b}\otimes\mathcal{H}^{\rm e}_{\rm b}. According to the decomposition of \mathcal{H}, we can naturally decompose the observable algebra ()=𝒬1𝒬2𝒜mfull𝒜efull\mathcal{B}(\mathcal{H})=\mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e}, where 𝒬1𝒬2\mathcal{Q}_{1}\otimes\mathcal{Q}_{2} is generated by the logical operators, i.e., the global operators 𝖷1,𝖹1,𝖷2,𝖹2\mathsf{X}_{1},\mathsf{Z}_{1},\mathsf{X}_{2},\mathsf{Z}_{2} appearing in (1.1). The syndrome algebras 𝒜mfull\mathcal{A}^{\rm full}_{\rm m} and 𝒜efull\mathcal{A}^{\rm full}_{\rm e} are spanned by linear transformations acting on the syndrome subspaces bm\mathcal{H}^{\rm m}_{\rm b} and be\mathcal{H}^{\rm e}_{\rm b}, respectively. Here, the \otimes symbol represents multiplication between commuting matrices, and every element in 𝒬1,𝒬2,𝒜mfull\mathcal{Q}_{1},\mathcal{Q}_{2},\mathcal{A}^{\rm full}_{\rm m}, and 𝒜efull\mathcal{A}^{\rm full}_{\rm e} is understood as a matrix defined over the entire Hilbert space \mathcal{H}. We put the detailed discussion of the above decomposition in Section 4.1.

After decomposing ()\mathcal{B}(\mathcal{H}), we can further decompose the Davies generator (1.1):

β=local+global:=gapped+rest,\mathcal{L}_{\beta}=\underbrace{\mathcal{L}_{\rm local}+\mathcal{L}_{\rm global}}_{:=\mathcal{L}^{\rm gapped}}+\mathcal{L}^{\rm rest}\,,

with gapped\mathcal{L}^{\rm gapped} defined in (4.13) and rest=local fulllocal\mathcal{L}^{\rm rest}=\mathcal{L}_{\text{\rm local full}}-\mathcal{L}_{\rm local}. Here we can show that gapped\mathcal{L}^{\rm gapped} is ergodic, i.e., Ker(gapped)=Span{𝟏}\mathrm{Ker}(\mathcal{L}^{\rm gapped})={\rm Span}\{{\bf 1}\}. By Lemma 3 (item 2), we can lower bound Gap(β)\mathrm{Gap}(\mathcal{L}_{\beta}) by Gap(gapped)\mathrm{Gap}(\mathcal{L}^{\rm gapped}). More importantly, the generator gapped\mathcal{L}^{\rm gapped} is block diagonal with respect to the following decomposition:

()=B1{𝟏,𝖷1,𝖸1,𝖹1},B2{𝟏,𝖷2,𝖸2,𝖹2}B1,B2,\mathcal{B}(\mathcal{H})=\bigoplus_{B_{1}\in\{{\bf 1},\mathsf{X}_{1},\mathsf{Y}_{1},\mathsf{Z}_{1}\},B_{2}\in\{{\bf 1},\mathsf{X}_{2},\mathsf{Y}_{2},\mathsf{Z}_{2}\}}\mathcal{B}_{B_{1},B_{2}}\,,

where B1,B2=B1B2𝒜mfull𝒜efull\mathcal{B}_{B_{1},B_{2}}=B_{1}\otimes B_{2}\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e} and the operators 𝖷i,𝖸i,𝖹i\mathsf{X}_{i},\mathsf{Y}_{i},\mathsf{Z}_{i} are given in (4.4). Because Ker(gapped)=Span{𝟏}\mathrm{Ker}(\mathcal{L}^{\rm gapped})={\rm Span}\{{\bf 1}\} and gapped\mathcal{L}^{\rm gapped} is block diagonal, we have

Gap(β)Gap(gapped)min{Gap(gapped|𝟏,𝟏),λmin(gapped|B1,B2)|B1𝟏orB2𝟏}.\mathrm{Gap}\left(-\mathcal{L}_{\beta}\right)\geq\mathrm{Gap}\left(-\mathcal{L}^{\rm gapped}\right)\\ \geq\min\left\{\mathrm{Gap}\left(-\mathcal{L}^{\rm gapped}\middle|_{\mathcal{B}_{{\bf 1},{\bf 1}}}\right),\lambda_{\min}\left(-\mathcal{L}^{\rm gapped}|_{\mathcal{B}_{B_{1},B_{2}}}\right)\middle|B_{1}\neq{\bf 1}\ \text{or}\ B_{2}\neq{\bf 1}\right\}\,. (3.1)

Thus, for a lower bound estimate of Gap(gapped)\mathrm{Gap}(-\mathcal{L}^{\rm gapped}), it suffices to consider Gap(gapped|𝟏,𝟏)\mathrm{Gap}(-\mathcal{L}^{\rm gapped}|_{\mathcal{B}_{{\bf 1},{\bf 1}}}) and λmin(gapped)|B1,B2\lambda_{\min}(-\mathcal{L}^{\rm gapped})|_{\mathcal{B}_{B_{1},B_{2}}} with B1𝟏orB2𝟏B_{1}\neq{\bf 1}\ \text{or}\ B_{2}\neq{\bf 1}. The latter term characterizes the transition rate between different logical subspaces. If β\mathcal{L}_{\beta} contains only local coupling operators, the system requires a long evolution time to overcome the energy barrier and transit between different logical subspaces, which leads to a slow mixing time scaling as exp(Θ(β))\exp(\Theta(\beta)) according to [2]. In our work, an important observation is that the global logical operators 𝖷1,𝖹1,𝖷2,𝖹2\mathsf{X}_{1},\mathsf{Z}_{1},\mathsf{X}_{2},\mathsf{Z}_{2} can directly flip the logical qubits, enabling transitions between different logical subspaces without the need to overcome the energy barrier. The efficient transition in logic density space implies a fast decaying of exp(tgapped)\exp(t\mathcal{L}^{\rm gapped}) in the logic subspace {B1,B2}B1𝟏orB2𝟏\{\mathcal{B}_{B_{1},B_{2}}\}_{B_{1}\neq{\bf 1}\ \text{or}\ B_{2}\neq{\bf 1}} and a lower bound of {gapped|B1,B2}B1𝟏orB2𝟏\{-\mathcal{L}^{\rm gapped}|_{\mathcal{B}_{B_{1},B_{2}}}\}_{B_{1}\neq{\bf 1}\ \text{or}\ B_{2}\neq{\bf 1}}. Specifically, in 8 in Section 4.2, we show

gapped|B1,B2Gap(local|𝟏,𝟏).-\mathcal{L}^{\rm gapped}|_{\mathcal{B}_{B_{1},B_{2}}}\succeq\mathrm{Gap}\left(-\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{{\bf 1},{\bf 1}}}\right)\,. (3.2)

According to the above analysis, the remaining thing to lower bound the spectral gap of β\mathcal{L}_{\beta} is to study the spectral gap of local\mathcal{L}_{\rm local} on 𝟏,𝟏\mathcal{B}_{{\bf 1},{\bf 1}}. Roughly speaking, this requires us to prove 2. Similar task is done in [2]. However, we emphasize that the lower bound and proof technique in [2] is not suitable for our purpose. In [2, Proposition 2], while the spectral gap is independent of the system size, it decays as exp(Θ(β))\exp(-\Theta(\beta)), indicating slow mixing in the syndrome subspace at low temperatures. A main contribution of this work is to show that this lower bound is not tight when β1\beta\gg 1, and the system actually mixes fast in the syndrome subspace at low temperatures. To this end, we develop a new iteration argument and a decomposition trick to prove that Gap(local|𝟏,𝟏)min{exp(Θ(β)),poly(1/N)}\mathrm{Gap}\left(\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{{\bf 1},{\bf 1}}}\right)\geq\min\left\{\exp(-\Theta(\beta)),\mathrm{poly}(1/N)\right\}, which provides a much sharper lower bound of the spectral gap in the lower temperature regime. This result is summarized in 9 in Section 4.2.1, which also provides a proof of 2. Plugging this into (3.1) and (3.2), we can conclude 1.

4 2D toric code

Let N=2L2N=2L^{2} spins be on the edges of the toroidal lattice, modeled by the Hilbert space 2N\mathcal{H}\cong\mathbb{C}^{2^{N}}. The Hamiltonian is given by

Htoric=sXspZp,H^{\rm toric}=-\sum_{s}\textbf{X}_{s}-\sum_{p}\textbf{Z}_{p}\,, (4.1)

where indices ss and pp denote a star and plaquette that consist of four sites around a node of the lattice and the center of a cell, respectively, as drawn in Fig. 1, and the associated observables Xs\textbf{X}_{s} and Zp\textbf{Z}_{p} are given by

Xs=isσix,Zp=ipσiz,\textbf{X}_{s}=\prod_{i\in s}\sigma^{x}_{i}\,,\quad\textbf{Z}_{p}=\prod_{i\in p}\sigma^{z}_{i}\,,

which commute with each other: [Xs,Zp]=[Xs,Xs]=[Zp,Zp]0[\textbf{X}_{s},\textbf{Z}_{p}]=[\textbf{X}_{s},\textbf{X}_{s^{\prime}}]=[\textbf{Z}_{p},\textbf{Z}_{p^{\prime}}]\equiv 0 for any stars s,ss,s^{\prime} and plaquettes p,pp,p^{\prime}. In addition, due to the periodic boundary condition, it holds that

sXs=1,pZp=1,\prod_{s}\textbf{X}_{s}=1\,,\quad\prod_{p}\textbf{Z}_{p}=1\,, (4.2)

This section is devoted to the spectral gap analysis of the Davies generator (Gibbs sampler) in (1.1) and the proof of 1, building on the discussion in Section 3. We will first discuss the ground states and the decomposition of the observable algebra of HtoricH^{\rm toric} in Section 4.1. Then, in Section 4.2, we first address the step (b) outlined in Section 3, reducing the problem to step (c): the study of the spectral gap of the local Davies generator on the syndrome space. This will be explored in detail in Section 4.2.1.

Refer to caption
Figure 1: 2D toric code. Each blue plaquette contains one 4-local operator Zp=iplaquette pσiz\textbf{Z}_{p}=\prod_{i\,\in\,\text{plaquette }p}\sigma^{z}_{i}. Each red star contains one 4-local operator Xs=istar sσix\textbf{X}_{s}=\prod_{i\,\in\,\text{star }s}\sigma^{x}_{i}. The dashed line on the right/bottom is the same as the solid line on the left/up to indicate the periodic boundary condition.

4.1 Ground states and observable algebra

In this section, we introduce the ground state space of HtoricH^{\rm toric} and the associated observable algebra ()\mathcal{B}(\mathcal{H}), following [2, 3, 12], which is important for the step (a) of the road map in Section 3.

Analogous to the 1D ferromagnetic Ising chain (see Appendix A), the 2D toric code Hamiltonian HtoricH^{\text{toric}} is frustration-free, i.e., its ground state is simultaneously an eigenvector with eigenvalue 1 for all local terms 𝐗s\mathbf{X}_{s} and 𝐙p\mathbf{Z}_{p}. Due to the toric structure with periodic boundary conditions, there are two topologically protected degrees of freedom (i.e., two logical qubits), resulting in a four-dimensional ground state space.

We now construct the ground state space of HtoricH^{\rm toric} explicitly. Given a vector |ϕ\ket{\phi} satisfying Zp|ϕ=|ϕ\textbf{Z}_{p}\ket{\phi}=\ket{\phi} for all pp, e.g., |0N\ket{0^{N}} or |1N\ket{1^{N}} (note that there are many others), we can construct a ground state as follows:

|ψ=Xstar|ϕ,Xstar:=sstar(I+Xs)=α{0,1}NsstarXsαs,\ket{\psi}=\textbf{X}_{\rm star}\ket{\phi}\,,\quad\textbf{X}_{\rm star}:=\prod_{s\in\rm star}\left(I+\textbf{X}_{s}\right)=\sum_{\alpha\in\{0,1\}^{N}}\prod_{s\in\rm star}\textbf{X}_{s}^{\alpha_{s}}\,, (4.3)

which, as one can easily check, satisfies Zp|ψ=|ψ\textbf{Z}_{p}\ket{\psi}=\ket{\psi}, Xs|ψ=|ψ\textbf{X}_{s}\ket{\psi}=\ket{\psi} for all p,sp,s. To find all the ground states, we first define some global observables for two logical qubits (see Fig. 2):

𝖷1:=jorange dotsσjx,𝖹1:=jgreen squaresσjz,𝖸1:=i𝖹1𝖷1,\displaystyle\mathsf{X}_{1}:=\prod_{j\in\text{orange dots}}\sigma^{x}_{j}\,,\quad\mathsf{Z}_{1}:=\prod_{j\in\text{green squares}}\sigma^{z}_{j}\,,\quad\mathsf{Y}_{1}:=i\mathsf{Z}_{1}\mathsf{X}_{1}\,, (4.4)
𝖷2:=jblue dotsσjx,𝖹2:=jred squaresσjz,𝖸2:=i𝖹2𝖷2.\displaystyle\mathsf{X}_{2}:=\prod_{j\in\text{blue dots}}\sigma^{x}_{j}\,,\quad\mathsf{Z}_{2}:=\prod_{j\in\text{red squares}}\sigma^{z}_{j}\,,\quad\mathsf{Y}_{2}:=i\mathsf{Z}_{2}\mathsf{X}_{2}\,.

Here, {𝖷1,𝖸1,𝖹1}\{\mathsf{X}_{1},\mathsf{Y}_{1},\mathsf{Z}_{1}\} and {𝖷2,𝖸2,𝖹2}\{\mathsf{X}_{2},\mathsf{Y}_{2},\mathsf{Z}_{2}\} generates two commuting Pauli algebras. Moreover, the operators 𝖷1\mathsf{X}_{1} and 𝖷2\mathsf{X}_{2} give one-to-one correspondences between four topological equivalent classes [3, Fig. 3.2] (see also [12, Section 3.2]). Specifically, we consider four orthogonal vectors that satisfy Zp|ϕ=|ϕ\textbf{Z}_{p}\ket{\phi}=\ket{\phi}:

|ϕo=|0N,|ϕ|=𝖷1|0N,|ϕ=𝖷2|0N,|ϕ+=𝖷1𝖷2|0N.\ket{\phi_{o}}=\ket{0^{N}}\,,\quad\ket{\phi_{|}}=\mathsf{X}_{1}\ket{0^{N}}\,,\quad\ket{\phi_{-}}=\mathsf{X}_{2}\ket{0^{N}}\,,\quad\ket{\phi_{+}}=\mathsf{X}_{1}\mathsf{X}_{2}\ket{0^{N}}\,.

They generate the four orthogonal ground states via (4.3):

|ψg=Xstar|ϕgwithg=o,|,,+,\displaystyle\ket{\psi_{g}}=\textbf{X}_{\rm star}\ket{\phi_{g}}\quad\text{with}\quad g=o,\ |,\ -,\ +\,, (4.5)

which span the whole ground state space of HtoricH^{\rm toric}. In this form, |ψg\ket{\psi_{g}} with g=o,|,,+g=o,\ |,\ -,\ + corresponds to the logical qubits |00\ket{00}, |10\ket{10}, |01\ket{01}, and |11\ket{11}, respectively.

Refer to caption
Figure 2: Four global logic operators in 2D toric code. Orange: 𝖷1=jorange dotsσjx\mathsf{X}_{1}=\prod_{j\in\text{orange dots}}\sigma^{x}_{j}. Green: 𝖹1=jgreen squaresσjz\mathsf{Z}_{1}=\prod_{j\in\text{green squares}}\sigma^{z}_{j}. Blue: 𝖷2=jblue dotsσjx\mathsf{X}_{2}=\prod_{j\in\text{blue dots}}\sigma^{x}_{j}. Red: 𝖹2=jred squaresσjz\mathsf{Z}_{2}=\prod_{j\in\text{red squares}}\sigma^{z}_{j}. Squares represent the qubits on 𝖹1,𝖹2\mathsf{Z}_{1},\mathsf{Z}_{2} logic operators, while dots represent the qubits on 𝖷1,𝖷2\mathsf{X}_{1},\mathsf{X}_{2} logic operators.

Next, to formulate the observable algebra, we first consider how the bit flip will influence the energy, namely, the excitation of HtoricH^{\rm toric}. For this, we delicately construct a snake path (as a subset of N=2L2N=2L^{2} sites) passing through the centers of all the cells and a comb path passing through all the nodes of the toroidal lattice (see Fig. 3) such that

  • {σjx}jsnake\{\sigma_{j}^{x}\}_{j\in{\rm snake}} generates all the XX-type excitations (also called “magnetic” excitations): for any two plaquettes (p1,p2p_{1},p_{2}), we can find a path ll on the snake connecting them (blue path in Fig. 3). Then we define

    Wlm=jlσjx,W_{l}^{m}=\prod_{j\in l}\sigma^{x}_{j}\,, (4.6)

    and then have the excited state |ψp1,p2=Wlm|ψ\ket{\psi_{p_{1},p_{2}}}=W_{l}^{m}\ket{\psi} satisfying

    Zp1/p2|ψp1,p2=|ψp1,p2,Htoric|ψp1,p2=Eground+2,\textbf{Z}_{p_{1}/p_{2}}\ket{\psi_{p_{1},p_{2}}}=-\ket{\psi_{p_{1},p_{2}}},\quad H^{\rm toric}\ket{\psi_{p_{1},p_{2}}}=E_{\rm ground}+2\,,

    where |ψ\ket{\psi} is a ground state of HtoricH^{\rm toric}.

  • {σjz}jcomb\{\sigma_{j}^{z}\}_{j\in{\rm comb}} generates all the ZZ-type excitations (also called “electric” excitations): for any two stars (s1,s2s_{1},s_{2}), we can find a path ll on the comb connecting them (red path in Fig. 3). We then define

    Wle=jlσjz.W_{l}^{e}=\prod_{j\in l}\sigma^{z}_{j}\,. (4.7)

    For a ground state |ψ\ket{\psi}, we define the excited state |ψs1,s2=Wle|ψ\ket{\psi_{s_{1},s_{2}}}=W_{l}^{e}\ket{\psi} satisfying

    Xs1/s2|ψs1,s2=|ψs1,s2,Htoric|ψs1,s2=Eground+2.\textbf{X}_{s_{1}/s_{2}}\ket{\psi_{s_{1},s_{2}}}=-\ket{\psi_{s_{1},s_{2}}},\quad H^{\rm toric}\ket{\psi_{s_{1},s_{2}}}=E_{\rm ground}+2\,.

    thanks to Xs1/s2Wle=WleXs1/s2\textbf{X}_{s_{1}/s_{2}}W_{l}^{e}=-W_{l}^{e}\textbf{X}_{s_{1}/s_{2}}.

  • The snake and comb form a partition of all but two spins. In addition, the snake does not intersect with 𝖹1,𝖹2\mathsf{Z}_{1},\mathsf{Z}_{2}, and the comb does not intersect with 𝖷1,𝖷2\mathsf{X}_{1},\mathsf{X}_{2}. This ensures that the excitation operators Wlm/eW_{l}^{m/e} constructed above commute with global observables: for a ground state |ψ\ket{\psi},

    𝖮Wlm/e|ψ=Wlm/e𝖮|ψ,\mathsf{O}W_{l}^{m/e}\ket{\psi}=W_{l}^{m/e}\mathsf{O}\ket{\psi}\,,

    where 𝖮=𝖷1\mathsf{O}=\mathsf{X}_{1}, 𝖷1\mathsf{X}_{1}, 𝖹1\mathsf{Z}_{1}, and 𝖹2\mathsf{Z}_{2}.

These elementary excited states |ψp1,p2,|ψs1,s2\ket{\psi_{p_{1},p_{2}}},\ket{\psi_{s_{1},s_{2}}} are often called quasi-particles pairs (or quasi-particles for short). All excited states of HtoricH^{\rm toric} can be expressed using these quasi-particles.

Following the above discussions, we decompose the space \mathcal{H} according to the magnetic excitations observed by Zp\textbf{Z}_{p} and the electric excitations observed by Xs\textbf{X}_{s}:

=22b2N=2L2withb=bmbe,\mathcal{H}=\mathbb{C}^{2}\otimes\mathbb{C}^{2}\otimes\mathcal{H}_{\rm b}\cong\mathbb{C}^{2^{N=2L^{2}}}\quad\text{with}\quad\mathcal{H}_{\rm b}=\mathcal{H}_{\rm b}^{\rm m}\otimes\mathcal{H}_{\rm b}^{\rm e}\,, (4.8)

where bm/e\mathcal{H}_{\rm b}^{\rm m/e} is the space of electric/magnetic excited states spanned by

{lWlm/e|ψg:Wlm/egiven in (4.6) and (4.7),g=o,|,,+}.\Big{\{}\prod_{l}W_{l}^{\rm m/e}\ket{\psi_{g}}\,:\ W_{l}^{\rm m/e}\ \text{given in \eqref{eqn:magnetic_operator} and \eqref{eqn:electric_operator}},g=o,|,-,+\Big{\}}\,. (4.9)

Moreover, a basis vector in b\mathcal{H}_{\rm b} can be identified as

|m|e=|m1,mL2|e1,eL2{+1,1}2L2,\ket{m}\ket{e}=\ket{m_{1},\ldots m_{L^{2}}}\ket{e_{1},\ldots e_{L^{2}}}\in\left\{+1,-1\right\}^{2L^{2}}\,, (4.10)

such that #{mj=1},#{ej=1}2\#\left\{m_{j}=-1\right\},\#\left\{e_{j}=-1\right\}\in 2\mathbb{Z}, due to the periodic boundary condition, with mj=±1m_{j}=\pm 1 (resp., ej=±1e_{j}=\pm 1) denoting the observation under Zp\textbf{Z}_{p} (resp., Xs\textbf{X}_{s}). Here and in what follows, we sort {Zp}p\{\textbf{Z}_{p}\}_{p} and {Xs}s\{\textbf{X}_{s}\}_{s} along the snake and comb such that they can be indexed by j{1,,L2}j\in\{1,\cdots,L^{2}\} (see Fig. 3). We emphasize that dim(b)=22L22\mathrm{dim}(\mathcal{H}_{b})=2^{2L^{2}-2} follows from the parity constraint (4.2), while each |m|e\ket{m}\ket{e} is a 2L22L^{2}-dimensional vector.

Now, we are ready to decompose the observable algebra ()\mathcal{B}(\mathcal{H}). Let 𝒬1\mathcal{Q}_{1} and 𝒬2\mathcal{Q}_{2} be the observable algebras over two logical qubits, generated by 𝖷1\mathsf{X}_{1} and 𝖹1\mathsf{Z}_{1} and by 𝖷2\mathsf{X}_{2} and 𝖹2\mathsf{Z}_{2}, respectively. We denote by 𝒜m/efull\mathcal{A}_{\rm m/e}^{\rm full} the linear operator spaces on bm/e\mathcal{H}_{\rm b}^{\rm m/e}. Then we have the following decomposition of the observable algebra for the 2D toric code. The proof is straightforward and given in Section 4.3 for completeness.

Lemma 5.

The algebra of observables for 2D toric code can be decomposed into

()𝒬1𝒬2𝒜mfull𝒜efull,\mathcal{B}(\mathcal{H})\cong\mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\otimes\mathcal{A}_{\rm m}^{\rm full}\otimes\mathcal{A}_{\rm e}^{\rm full}\,, (4.11)

associated with the decomposition (4.8), where the algebras 𝒜mfull\mathcal{A}^{\rm full}_{\rm m} and 𝒜efull\mathcal{A}^{\rm full}_{\rm e} are generated by {𝐙p}p{σjx}jsnake\{\mathbf{Z}_{p}\}_{p}\cup\{\sigma^{x}_{j}\}_{j\in\text{\rm snake}} and {𝐗s}s{σjz}jcomb\{\mathbf{X}_{s}\}_{s}\cup\{\sigma^{z}_{j}\}_{j\in\text{\rm comb}}. In addition, the four subalgebras 𝒬1\mathcal{Q}_{1}, 𝒬2\mathcal{Q}_{2}, 𝒜mfull\mathcal{A}_{\rm m}^{\rm full}, and 𝒜efull\mathcal{A}_{\rm e}^{\rm full} are commutative with each other.

To give further interpretation on the decomposition (4.11), we define four subspaces g\mathcal{H}_{g} with g=o,|,,+g=o,\ |,\ -,\ + spanned by the ground state |ψg\ket{\psi_{g}} in (4.5) and its excited states Wlm/e|ψgW_{l}^{m/e}\ket{\psi_{g}} via (4.6) and (4.7). Then it is easy to see dim(g)=2N2\dim(\mathcal{H}_{g})=2^{N-2} with N=2L2N=2L^{2}. The algebra 𝒬1𝒬2\mathcal{Q}_{1}\otimes\mathcal{Q}_{2} gives all the linear transformation between subspaces g\mathcal{H}_{g} with g=o,|,,+g=o,\ |,\ -,\ +, while on each subspace g\mathcal{H}_{g}, the linear maps are characterized by 𝒜mfull𝒜efull\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e}. In particular, recalling the basis (4.10), 𝒜mfull\mathcal{A}^{\rm full}_{\rm m} (resp., 𝒜efull\mathcal{A}^{\rm full}_{\rm e}) transfers the bond |m\ket{m} (resp., |e\ket{e}), for example,

𝒜mfull(|m|e)=(𝒜mfull|m)|e.\mathcal{A}^{\rm full}_{\rm m}(\ket{m}\ket{e})=\left(\mathcal{A}^{\rm full}_{\rm m}\ket{m}\right)\ket{e}\,.
Refer to caption
Figure 3: Orange dot Snake and green square comb for 2D toric code. The blue magnetic path operators WlmW_{l}^{m} (4.6) and red electric path operators WleW_{l}^{e} (4.7) act along the snake and the comb, respectively. We index {Zp}p\{\textbf{Z}_{p}\}_{p} as {Zj}j=1L2\{\textbf{Z}_{j}\}_{j=1}^{L^{2}} along the snake from left to right and up to down, and index the spins from 22 to L2L^{2} on the snake in the same way so that each spin jj corresponds a pair of plaquettes j1j-1 and jj. One can similarly index the stars and spins on comb from left to right and up to down.

4.2 Spectral gap of the Gibbs sampler

In this section, we will focus on steps (b) and (c) of the roadmap in Section 3. For this, we decompose the generator (1.1) as follows:

β=local full+global=gapped+rest,\mathcal{L}_{\beta}=\mathcal{L}_{\text{\rm local full}}+\mathcal{L}_{\rm global}=\mathcal{L}^{\rm gapped}+\mathcal{L}^{\rm rest}\,, (4.12)

where

gapped:=jsnakeσjx+jcombσjz:=local+𝖷1+𝖷2+𝖹1+𝖹2:=global.\displaystyle\mathcal{L}^{\rm gapped}:=\underbrace{\sum_{j\in\rm snake}\mathcal{L}_{\sigma^{x}_{j}}+\sum_{j\in\rm comb}\mathcal{L}_{\sigma^{z}_{j}}}_{:=\mathcal{L}_{\rm local}}+\underbrace{\mathcal{L}_{\mathsf{X}_{1}}+\mathcal{L}_{\mathsf{X}_{2}}+\mathcal{L}_{\mathsf{Z}_{1}}+\mathcal{L}_{\mathsf{Z}_{2}}}_{:=\mathcal{L}_{\rm global}}\,. (4.13)

Here rest:=local fulllocal\mathcal{L}^{\rm rest}:=\mathcal{L}_{\text{\rm local full}}-\mathcal{L}_{\rm local} is a local Lindbladian with other Pauli couplings not included in local\mathcal{L}_{\rm local}. From Lemma 3 (item 2), it suffices to limit our discussion to the part gapped\mathcal{L}^{\rm gapped} and show that it is primitive with the desired spectral gap lower bound max{exp(𝒪(β)),Ω(N3)}\max\left\{\exp(-\mathcal{O}(\beta)),\Omega(N^{-3})\right\}.

Before we proceed, we prepare the explicit formulations of the Lindbladians involved in gapped\mathcal{L}^{\rm gapped} for subsequent analysis. For σjx\sigma_{j}^{x} with jsnakej\in{\rm snake}, we have

eitHtoricσjxeitHtoric=eit(Zp+Zp)σjxeit(Zp+Zp),e^{itH^{\rm toric}}\sigma_{j}^{x}e^{-itH^{\rm toric}}=e^{-it\left(\textbf{Z}_{p^{\prime}}+\textbf{Z}_{p}\right)}\sigma_{j}^{x}e^{it\left(\textbf{Z}_{p^{\prime}}+\textbf{Z}_{p}\right)}\,,

where pp and pp^{\prime} are the two plaquettes with j=ppj=p\cap p^{\prime} being the intersection site. The bond observable (Zp+Zp)-(\textbf{Z}_{p^{\prime}}+\textbf{Z}_{p}) has eigenvalues 2,0,2-2,0,2 with the associated eigenprojections denoted by PjP^{-}_{j}, Pj0P_{j}^{0}, Pj+P^{+}_{j}, respectively, which can be represented as follows:

Pj0=12(IZpZp),Pj±=14(IZp)(IZp).P_{j}^{0}=\frac{1}{2}(I-\textbf{Z}_{p^{\prime}}\textbf{Z}_{p})\,,\quad P_{j}^{\pm}=\frac{1}{4}(I\mp\textbf{Z}_{p^{\prime}})(I\mp\textbf{Z}_{p})\,. (4.14)

We then compute the Fourier components of σjx\sigma_{j}^{x} (jsnakej\in{\rm snake}):

eitHtoricσjxeitHtoric=e4itPjσjxPj+aj+e4itPj+σjxPjaj+Pj0σjxPj0aj0,\displaystyle e^{itH^{\rm toric}}\sigma_{j}^{x}e^{-itH^{\rm toric}}=e^{-4it}\underbrace{P_{j}^{-}\sigma_{j}^{x}P_{j}^{+}}_{a_{j}}+e^{4it}\underbrace{P_{j}^{+}\sigma_{j}^{x}P_{j}^{-}}_{a_{j}^{\dagger}}+\underbrace{P_{j}^{0}\sigma_{j}^{x}P_{j}^{0}}_{a_{j}^{0}}\,, (4.15)

due to Pj±σjxPj0=Pj+σjxPj+=PjσjxPj=0P_{j}^{\pm}\sigma_{j}^{x}P_{j}^{0}=P^{+}_{j}\sigma_{j}^{x}P^{+}_{j}=P^{-}_{j}\sigma_{j}^{x}P^{-}_{j}=0, where aja_{j}, aja_{j}^{\dagger}, aj0a_{j}^{0} correspond to the Bohr frequencies 4-4, 44, 0 of the Hamiltonian (4.1), respectively. Then, we can write σjx\mathcal{L}_{\sigma_{j}^{x}} by Eqs. 2.3 and 2.7 with Glauber-type transition (2.5):

σjx(A)=12[aj0,[aj0,A]]+12{h+aj[A,aj]+h[aj,A]aj}+12{h+[aj,A]aj+haj[A,aj]},\mathcal{L}_{\sigma_{j}^{x}}(A)=-\frac{1}{2}[a_{j}^{0},[a_{j}^{0},A]]+\frac{1}{2}\left\{h_{+}a_{j}^{\dagger}[A,a_{j}]+h_{-}[a_{j},A]a_{j}^{\dagger}\right\}\\ +\frac{1}{2}\left\{h_{+}[a_{j}^{\dagger},A]a_{j}+h_{-}a_{j}[A,a_{j}^{\dagger}]\right\}\,, (4.16)

where the constants h±h_{\pm} are given by

h+=γ(4)=2e4β+1,h=γ(4)=2e4β+1.h_{+}=\gamma(-4)=\frac{2}{e^{-4\beta}+1}\,,\quad h_{-}=\gamma(4)=\frac{2}{e^{4\beta}+1}\,. (4.17)

The computation for σjz\mathcal{L}_{\sigma_{j}^{z}} with jcombj\in{\rm comb} is quite similar to σjx\mathcal{L}_{\sigma_{j}^{x}} with jsnakej\in{\rm snake}, so we only sketch it below. For σjz\sigma_{j}^{z} with jcombj\in{\rm comb}, we have

eitHtoricσjxeitHtoric\displaystyle e^{itH^{\rm toric}}\sigma_{j}^{x}e^{-itH^{\rm toric}} =eit(Xs+Xs)σjzeit(Xs+Xs)\displaystyle=e^{-it\left(\textbf{X}_{s^{\prime}}+\textbf{X}_{s}\right)}\sigma_{j}^{z}e^{it\left(\textbf{X}_{s^{\prime}}+\textbf{X}_{s}\right)}
=e4itQjσjzQj+bj+e4itQj+σjzQjbj+Qj0σjzQj0bj0,\displaystyle=e^{-4it}\underbrace{Q_{j}^{-}\sigma_{j}^{z}Q_{j}^{+}}_{b_{j}}+e^{4it}\underbrace{Q_{j}^{+}\sigma_{j}^{z}Q_{j}^{-}}_{b_{j}^{\dagger}}+\underbrace{Q_{j}^{0}\sigma_{j}^{z}Q_{j}^{0}}_{b_{j}^{0}}\,,

with ss and ss^{\prime} being two starts uniquely determined by j=ssj=s\cap s^{\prime}. Here, the projections QjQ^{-}_{j}, Qj0Q_{j}^{0}, Qj+Q^{+}_{j} are the eigenprojections of the bond observable (Xs+Xs)-(\textbf{X}_{s^{\prime}}+\textbf{X}_{s}) for eigenvalues 2,0,2-2,0,2, which can be similarly formulated as (4.14) by replacing Zp\textbf{Z}_{p} by Xs\textbf{X}_{s}. With the Glauber-type transition rate (2.5), the generator σjx\mathcal{L}_{\sigma_{j}^{x}} is given by

σjz(A)=12[bj0,[bj0,A]]+12{h+bj[A,bj]+h[bj,A]bj}+12{h+[bj,A]bj+hbj[A,bj]},\mathcal{L}_{\sigma_{j}^{z}}(A)=-\frac{1}{2}[b_{j}^{0},[b_{j}^{0},A]]+\frac{1}{2}\left\{h_{+}b_{j}^{\dagger}[A,b_{j}]+h_{-}[b_{j},A]b_{j}^{\dagger}\right\}\\ +\frac{1}{2}\left\{h_{+}[b_{j}^{\dagger},A]b_{j}+h_{-}b_{j}[A,b_{j}^{\dagger}]\right\}\,, (4.18)

where the constants h±h_{\pm} is the same as (4.17). According to our construction, we have

σjzσβσβ,σjxσβσβ=Θ(1),uniformly in β.\big{\|}\mathcal{L}_{\sigma_{j}^{z}}\big{\|}_{\sigma_{\beta}\to\sigma_{\beta}},\ \,\big{\|}\mathcal{L}_{\sigma_{j}^{x}}\big{\|}_{\sigma_{\beta}\to\sigma_{\beta}}=\Theta(1)\,,\quad\text{uniformly in $\beta$}\,.

We proceed to compute the Lindbladian with global couplings. For 𝖮=𝖷1,𝖷2,𝖹1,𝖹2\mathsf{O}=\mathsf{X}_{1},\mathsf{X}_{2},\mathsf{Z}_{1},\mathsf{Z}_{2}, thanks to eitHtoric𝖮eitHtoric=𝖮e^{itH^{\rm toric}}\mathsf{O}e^{-itH^{\rm toric}}=\mathsf{O} by [Zp,𝖮]=[Xs,𝖮]=0[\textbf{Z}_{p},\mathsf{O}]=[\textbf{X}_{s},\mathsf{O}]=0, we readily have

𝖮(A)=12[𝖮,[𝖮,A]],\mathcal{L}_{\mathsf{O}}(A)=-\frac{1}{2}[\mathsf{O},[\mathsf{O},A]]\,, (4.19)

since 𝖮\mathsf{O} has only the component with Bohr frequency zero. It is clear that

𝖮σβσβ=Θ(1),uniformly in β.\left\lVert\mathcal{L}_{\mathsf{O}}\right\rVert_{\sigma_{\beta}\to\sigma_{\beta}}=\Theta(1)\,,\quad\text{uniformly in $\beta$}\,. (4.20)

Now, for the step (b) of the roadmap in Section 3, we first give the following lemma. Its proof is postponed to Section 4.3 for ease of reading.

Lemma 6.

The generators global\mathcal{L}_{\rm global}, jsnakeσjx\sum_{j\in\rm snake}\mathcal{L}_{\sigma^{x}_{j}}, and jcombσjz\sum_{j\in\rm comb}\mathcal{L}_{\sigma^{z}_{j}} defined in (4.13) only nontrivially act on a subalgebra of ()\mathcal{B}(\mathcal{H}). Specifically, we have

global(𝒬1𝒬2𝒜mfull𝒜efull)=global(𝒬1𝒬2)𝒜mfull𝒜efull,\mathcal{L}_{\rm global}\left(\mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e}\right)=\mathcal{L}_{\rm global}\left(\mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\right)\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e}\,, (4.21)
jsnakeσjx(𝒬1𝒬2𝒜mfull𝒜efull)=𝒬1𝒬2jsnakeσjx(𝒜mfull)𝒜efull,\sum_{j\in\rm snake}\mathcal{L}_{\sigma^{x}_{j}}\left(\mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e}\right)=\mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\otimes\sum_{j\in\rm snake}\mathcal{L}_{\sigma^{x}_{j}}\left(\mathcal{A}^{\rm full}_{\rm m}\right)\otimes\mathcal{A}^{\rm full}_{\rm e}\,, (4.22)

and

jcombσjz(𝒬1𝒬2𝒜mfull𝒜efull)=𝒬1𝒬2𝒜mfulljcombσjz(𝒜efull).\sum_{j\in\rm comb}\mathcal{L}_{\sigma^{z}_{j}}\left(\mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e}\right)=\mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes\sum_{j\in\rm comb}\mathcal{L}_{\sigma^{z}_{j}}\left(\mathcal{A}^{\rm full}_{\rm e}\right)\,. (4.23)

In particular, it holds that

Ker(jsnakeσjx)=\displaystyle\mathrm{Ker}\Big{(}\sum_{j\in\rm snake}\mathcal{L}_{\sigma^{x}_{j}}\Big{)}= 𝒬1𝒬2𝟏𝒜efull,\displaystyle\leavevmode\nobreak\ \mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\otimes{\rm{\bf 1}}\otimes\mathcal{A}^{\rm full}_{\rm e}\,, (4.24)
Ker(jcombσjz)=\displaystyle\mathrm{Ker}\Big{(}\sum_{j\in\rm comb}\mathcal{L}_{\sigma^{z}_{j}}\Big{)}= 𝒬1𝒬2𝒜mfull𝟏,\displaystyle\leavevmode\nobreak\ \mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes{\rm{\bf 1}}\,,
Ker(global)=\displaystyle\mathrm{Ker}\left(\mathcal{L}_{\rm global}\right)= 𝟏𝟏𝒜mfull𝒜efull.\displaystyle\leavevmode\nobreak\ {\rm{\bf 1}}\otimes{\rm{\bf 1}}\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e}\,.

Note from (4.19) that for any B1,B2{I,𝖷i,𝖸i,𝖹i}B_{1},B_{2}\in\{I,\mathsf{X}_{i},\mathsf{Y}_{i},\mathsf{Z}_{i}\}, the operator B1B2B_{1}B_{2} is an eigenvector of global\mathcal{L}_{\rm global}: global(B1B2)=cB1B2\mathcal{L}_{\rm global}\left(B_{1}B_{2}\right)=c\cdot B_{1}B_{2} for some constant cc. Combining this with Lemma 6, we obtain the following properties of gapped\mathcal{L}^{\rm gapped}.

Corollary 7.

It holds that

  • gapped\mathcal{L}^{\rm gapped} is block diagonal for the following orthogonal decomposition (4.25) in both HS inner product and GNS inner product:

    𝒬1𝒬2𝒜mfull𝒜efull=Bi{I,𝖷i,𝖸i,𝖹i},i=1,2B1,B2,\mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e}=\bigoplus_{B_{i}\in\{I,\mathsf{X}_{i},\mathsf{Y}_{i},\mathsf{Z}_{i}\},\leavevmode\nobreak\ i=1,2}\mathcal{B}_{B_{1},B_{2}}\,, (4.25)

    with

    B1,B2=B1B2𝒜mfull𝒜efull.\mathcal{B}_{B_{1},B_{2}}=B_{1}\otimes B_{2}\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e}\,.
  • gapped\mathcal{L}^{\rm gapped} is primitive: Ker(gapped)=Span{𝟏}\mathrm{Ker}(\mathcal{L}^{\rm gapped})={\rm Span}\{{\bf 1}\}.

By (2.11) in Lemma 3 with the primitivity of gapped\mathcal{L}^{\rm gapped}, we have

Gap(β)=Gap(restgapped)Gap(gapped).\mathrm{Gap}(-\mathcal{L}_{\beta})=\mathrm{Gap}(-\mathcal{L}^{\rm rest}-\mathcal{L}^{\rm gapped})\geq\mathrm{Gap}(-\mathcal{L}^{\rm gapped})\,.

Further, from the block diagonal form of gapped\mathcal{L}^{\rm gapped} for (4.25), we obtain (3.1):

Gap(β)Gap(gapped)min{Gap(gapped|𝟏,𝟏),λmin(gapped|B1,B2)forB1𝟏orB2𝟏}.\mathrm{Gap}\left(-\mathcal{L}_{\beta}\right)\geq\mathrm{Gap}\left(-\mathcal{L}^{\rm gapped}\right)\\ \geq\min\left\{\mathrm{Gap}\left(-\mathcal{L}^{\rm gapped}\middle|_{\mathcal{B}_{{\bf 1},{\bf 1}}}\right),\lambda_{\min}\left(-\mathcal{L}^{\rm gapped}|_{\mathcal{B}_{B_{1},B_{2}}}\right)\leavevmode\nobreak\ \text{for}\leavevmode\nobreak\ B_{1}\neq{\bf 1}\ \text{or}\ B_{2}\neq{\bf 1}\right\}\,. (4.26)

For the first term in (4.26), we note that global|𝟏,𝟏=0\mathcal{L}_{\rm global}|_{\mathcal{B}_{{\bf 1},{\bf 1}}}=0 from (4.24), and then have

Gap(gapped|𝟏,𝟏)=Gap(local|𝟏,𝟏).\mathrm{Gap}\left(-\mathcal{L}^{\rm gapped}\middle|_{\mathcal{B}_{{\bf 1},{\bf 1}}}\right)=\mathrm{Gap}\left(-\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{{\bf 1},{\bf 1}}}\right)\,. (4.27)

We next consider the lower bound estimation of λmin(gapped|B1,B2)\lambda_{\min}\left(-\mathcal{L}^{\rm gapped}|{\mathcal{B}_{B_{1},B_{2}}}\right). For any Bi{I,𝖷i,𝖸i,𝖹i}B_{i}\in\{I,\mathsf{X}_{i},\mathsf{Y}_{i},\mathsf{Z}_{i}\} with i=1,2i=1,2 such that B1𝟏B_{1}\neq\mathbf{1} or B2𝟏B_{2}\neq\mathbf{1}, the operators 𝖷1,𝖷2,𝖹1,𝖹2\mathsf{X}_{1},\mathsf{X}_{2},\mathsf{Z}_{1},\mathsf{Z}_{2} either commute or anti-commute with B1B2B_{1}B_{2}, and there always exists one, say 𝖮\mathsf{O}, among them that anti-commutes with B1B2B_{1}B_{2}. Then, by (4.19), we have 𝖮(A)=2A\mathcal{L}_{\mathsf{O}}(A)=-2A, which implies

B1B2,global(B1B2)σβB1B2,𝖮(B1B2)σβ=2.\left\langle B_{1}B_{2},-\mathcal{L}_{\rm global}(B_{1}B_{2})\right\rangle_{\sigma_{\beta}}\geq\left\langle B_{1}B_{2},-\mathcal{L}_{\mathsf{O}}(B_{1}B_{2})\right\rangle_{\sigma_{\beta}}=2\,. (4.28)

Furthermore, according to Eqs. 4.22, 4.23 and 4.24 in Lemma 6, we find

Ker(local|B1,B2)=B1B2.\mathrm{Ker}\left(\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{B_{1},B_{2}}}\right)=B_{1}B_{2}\,.

Finally, by Lemma 3 (item 4), there holds

gapped|B1,B2\displaystyle-\mathcal{L}^{\rm gapped}\Big{|}_{\mathcal{B}_{B_{1},B_{2}}} Gap(local|B1,B2)B1B2,global(B1B2)σβGap(local|B1,B2)+globalσβσβ\displaystyle\succeq\frac{\mathrm{Gap}\left(\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{B_{1},B_{2}}}\right)\left\langle B_{1}B_{2},-\mathcal{L}_{\rm global}(B_{1}B_{2})\right\rangle_{\sigma_{\beta}}}{\mathrm{Gap}\left(\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{B_{1},B_{2}}}\right)+\left\lVert\mathcal{L}_{\rm global}\right\rVert_{\sigma_{\beta}\rightarrow\sigma_{\beta}}} (4.29)
=Θ(Gap(local|B1,B2)Gap(local|B1,B2)+1),\displaystyle=\Theta\left(\frac{\mathrm{Gap}\left(\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{B_{1},B_{2}}}\right)}{\mathrm{Gap}\left(\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{B_{1},B_{2}}}\right)+1}\right)\,,

where we have used (4.20). Plugging this into (4.26), we obtain the following proposition, concluding step (b) of the proof.

Proposition 8.

Let β\mathcal{L}_{\beta}, gapped\mathcal{L}^{\rm gapped}, global\mathcal{L}_{\rm global} and local\mathcal{L}_{\rm local} be defined in (4.12)-(4.13). Suppose Gap(local|𝟏,𝟏)=𝒪(1)\mathrm{Gap}\left(\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{{\bf 1},{\bf 1}}}\right)=\mathcal{O}(1). Then, for any Bi{I,𝖷i,𝖸i,𝖹i}B_{i}\in\{I,\mathsf{X}_{i},\mathsf{Y}_{i},\mathsf{Z}_{i}\} with i=1,2i=1,2 such that B1𝟏B_{1}\neq\mathbf{1} or B2𝟏B_{2}\neq\mathbf{1}, we have

gapped|B1,B2Ω(Gap(local|𝟏,𝟏)),-\mathcal{L}^{\rm gapped}|_{\mathcal{B}_{B_{1},B_{2}}}\succeq\Omega\left(\mathrm{Gap}\left(\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{{\bf 1},{\bf 1}}}\right)\right)\,, (4.30)

Moreover, it holds that

Gap(β)=Ω(Gap(local|𝟏,𝟏)).\mathrm{Gap}\left(\mathcal{L}_{\beta}\right)=\Omega\left(\mathrm{Gap}\left(-\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{{\bf 1},{\bf 1}}}\right)\right)\,. (4.31)
Proof.

Since the action of local\mathcal{L}_{\rm local} on B1,B2\mathcal{B}_{B_{1},B_{2}} is independent of B1,B2B_{1},B_{2} by (4.22) and (4.23), we obtain Gap(local|B1,B2)=Gap(local|𝟏,𝟏)\mathrm{Gap}(\mathcal{L}_{\rm local}|_{\mathcal{B}_{B_{1},B_{2}}})=\mathrm{Gap}(\mathcal{L}_{\rm local}|_{\mathcal{B}_{{\bf 1},{\bf 1}}}), which, along with (4.29), implies (4.30). Then, the estimate (4.31) directly follows from (4.26) and (4.27). ∎

Thanks to 8 above, to finish the proof of 1, it suffices to study the spectral gap of local\mathcal{L}_{\rm local} on the syndrome space 𝟏,𝟏\mathcal{B}_{{\bf 1},{\bf 1}}, i.e., step (c) outlined in Section 3. This is the goal of the next section.

4.2.1 Analysis of the quasi-1D structure

In this section, we will analyze the spectral gap of the local Davies generator local\mathcal{L}_{\rm local} in the syndrome space. The main result is stated as follows.

Proposition 9.

Let local\mathcal{L}_{\rm local} be defined in (4.13) and 𝟏,𝟏\mathcal{B}_{{\bf 1},{\bf 1}} be defined in 7, we have

Gap(local|𝟏,𝟏)=max{e𝒪(β),Ω(N3)}.\mathrm{Gap}\left(-\mathcal{L}_{\rm local}\middle|_{\mathcal{B}_{{\bf 1},{\bf 1}}}\right)=\max\left\{e^{-\mathcal{O}(\beta)},\Omega(N^{-3})\right\}\,.

Then, 1 is a corollary of 8 and 9. To prove the above proposition, we consider

1:=jsnakeσjx,2:=jcombσjz.\mathcal{L}_{1}:=\sum_{j\in\rm snake}\mathcal{L}_{\sigma^{x}_{j}},\quad\mathcal{L}_{2}:=\sum_{j\in\rm comb}\mathcal{L}_{\sigma^{z}_{j}}\,. (4.32)

From Lemma 6, it is straightforward to see that 1\mathcal{L}_{1} and 2\mathcal{L}_{2} commutes and only act nontrivally on 𝒜mfull\mathcal{A}^{\rm full}_{\rm m} and 𝒜efull\mathcal{A}^{\rm full}_{\rm e}, respectively. Thus, we can analyze the spectral gap of 1\mathcal{L}_{1} and 2\mathcal{L}_{2} separately. 9 directly follows from the following two lemmas.

Lemma 10.

Let 𝒜mfull\mathcal{A}^{\rm full}_{\rm m} be defined as in Lemma 5, we have

Gap(1|𝒜mfull)=max{e𝒪(β),Ω(N3)}.\mathrm{Gap}\left(-\mathcal{L}_{1}\middle|_{\mathcal{A}^{\rm full}_{\rm m}}\right)=\max\left\{e^{-\mathcal{O}(\beta)},\Omega(N^{-3})\right\}\,. (4.33)
Lemma 11.

Let 𝒜efull\mathcal{A}^{\rm full}_{\rm e} be defined as in Lemma 5, we have

Gap(2|𝒜efull)=max{e𝒪(β),Ω(N2.5)}.\mathrm{Gap}\left(-\mathcal{L}_{2}\middle|_{\mathcal{A}^{\rm full}_{\rm e}}\right)=\max\left\{e^{-\mathcal{O}(\beta)},\Omega(N^{-2.5})\right\}\,.

The exp(𝒪(β))\exp(-\mathcal{O}(\beta)) spectral gap of Gap(1|𝒜mfull)\mathrm{Gap}(-\mathcal{L}_{1}|_{\mathcal{A}^{\rm full}_{\rm m}}) and Gap(2|𝒜efull)\mathrm{Gap}(-\mathcal{L}_{2}|_{\mathcal{A}^{\rm full}_{\rm e}}) has been proved in [2, Section 7]. In what follows, we focus on the gap estimate of order poly(N1){\rm poly}(N^{-1}). We first prove Lemma 10.

Proof of Lemma 10.

Recall that 𝒜mfull=(bm)\mathcal{A}^{\rm full}_{\rm m}=\mathcal{B}\left(\mathcal{H}^{\rm m}_{\rm b}\right) is spanned by the basis matrices |mm|\ket{m}\bra{m^{\prime}}, where m,m{1,1}L2m,m^{\prime}\in\{-1,1\}^{L^{2}} and #{mi=1},#{mi=1}2\#\left\{m_{i}=-1\right\},\#\left\{m^{\prime}_{i}=-1\right\}\in 2\mathbb{Z}. We note that this set of basis matrices {|mm|}\left\{\ket{m}\bra{m^{\prime}}\right\} also forms an orthogonal basis for 𝒜mfull\mathcal{A}^{\rm full}_{\rm m} with respect to the GNS inner product defined as in (2.8) by the reduced Gibbs state Tr𝒜efull(σβ)exp(βpZp)\operatorname{Tr}_{\mathcal{A}^{\rm full}_{\rm e}}(\sigma_{\beta})\propto\exp(-\beta\sum_{p}\textbf{Z}_{p}) 111Note from Lemma 5 that σβ𝒜mfull𝒜efull\sigma_{\beta}\in\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e}..

We next decompose the space 𝒜mfull\mathcal{A}^{\rm full}_{\rm m} such that 1\mathcal{L}_{1} presents a block diagonal form. Note that 𝒜mfull\mathcal{A}^{\rm full}_{\rm m} is spanned by |mm|\ket{m^{\prime}}\bra{m} with |m,|mbm\ket{m},\ket{m^{\prime}}\in\mathcal{H}^{\rm m}_{\rm b} having an even number of - signs. For any Λ{1,2,,L2}\Lambda\subset\{1,2,\ldots,L^{2}\} with even L2|Λ|L^{2}-|\Lambda|, we introduce the subspace 𝒜mfull(Λ)\mathcal{A}^{\rm full}_{\rm m}(\Lambda) of 𝒜mfull\mathcal{A}^{\rm full}_{\rm m} spanned by |mm|\ket{m^{\prime}}\bra{m}, where m=mm=m^{\prime} on Λ\Lambda and m=mm=-m^{\prime} on the the complement Λc:={1,2,,L2}\Λ\Lambda^{c}:=\{1,2,\ldots,L^{2}\}\backslash\Lambda. It is straightforward to check that the subspaces 𝒜mfull(Λ)\mathcal{A}^{\rm full}_{\rm m}(\Lambda) are orthogonal with respect to both the GNS and HS inner products, and thereby

𝒜mfull=Λ{1,,L2}:L2Λeven𝒜mfull(Λ).\displaystyle\mathcal{A}^{\rm full}_{\rm m}=\bigoplus_{\begin{subarray}{c}\Lambda\subset\{1,\dots,L^{2}\}:\\ L^{2}-\Lambda\leavevmode\nobreak\ \text{even}\end{subarray}}\mathcal{A}^{\rm full}_{\rm m}(\Lambda)\,. (4.34)

In addition, by writing |mm|=(|mm|)Λ(|mm|)Λc𝒜mfull(Λ)\ket{m^{\prime}}\bra{m}=(\ket{m}\bra{m})_{\Lambda}\otimes(\ket{-m}\bra{m})_{\Lambda^{c}}\in\mathcal{A}^{\rm full}_{\rm m}(\Lambda), we can further decompose

𝒜mfull(Λ)=𝒜ab(Λ)(Λ),\displaystyle\mathcal{A}^{\rm full}_{\rm m}(\Lambda)=\mathcal{A}^{\rm ab}(\Lambda)\otimes\mathcal{F}(\Lambda)\,,

where 𝒜ab(Λ)\mathcal{A}^{\rm ab}(\Lambda) is the Abelian algebra generated by the projections on bonds restricted to Λ\Lambda and (Λ)\mathcal{F}(\Lambda) is the space spanned by flips of bonds restricted to Λc\Lambda^{c}. We observe that any partition of {1,2,,L2}\{1,2,\ldots,L^{2}\} into ΛΛc\Lambda\cup\Lambda^{c} induces a partition of the spins jj on the snake, equivalently, pairs of neighboring plaquettes/bonds {Zj1,Zj}\{\textbf{Z}_{j-1},\textbf{Z}_{j}\} into three sets: Γflip\Gamma_{\rm flip} for both bonds in Λc\Lambda^{c}, Γab\Gamma_{\rm ab} for both bonds in Λ\Lambda, and Γint\Gamma_{\rm int} for one bond in Λ\Lambda and the other in Λc\Lambda^{c}.

The following lemma extends [2, Lemma 5], which was originally stated for the 1D ferromagnetic Ising model. We provide a self-contained proof in Section 4.3 with explicit computations of the local matrix representations of the master Hamiltonian of 1\mathcal{L}_{1}.

Lemma 12.

1\mathcal{L}_{1}, defined in (4.32), on 𝒜mfull\mathcal{A}^{\rm full}_{\rm m} is block diagonal for the decomposition (4.34).

Now, we are ready to estimate the gap of 1\mathcal{L}_{1} on each 𝒜mfull(Λ)\mathcal{A}^{\rm full}_{\rm m}(\Lambda) by the following three cases.

  • If Γint\Gamma_{\rm int}\neq\emptyset, by Eq. 4.54 in the proof of Lemma 12, under an orthonormal basis with respect to the GNS inner product, the master Hamiltonian of σjx-\mathcal{L}_{\sigma^{x}_{j}} on 𝒜mfull(Λ)\mathcal{A}^{\rm full}_{\rm m}(\Lambda) takes the form:

    Ij2[h+120000h++120000h+120000h++12]IL2j12.I_{j-2}\otimes\left[\begin{matrix}\frac{h_{-}+1}{2}&0&0&0\\ 0&\frac{h_{+}+1}{2}&0&0\\ 0&0&\frac{h_{-}+1}{2}&0\\ 0&0&0&\frac{h_{+}+1}{2}\\ \end{matrix}\right]\otimes I_{L^{2}-j}\succeq\frac{1}{2}\,. (4.35)

    This implies

    A,1|𝒜mfull(Λ)AσβjΓintA,σjx|𝒜mfull(Λ)Aσβ|Γint|12A,Aσβ.\displaystyle-\left\langle A,\mathcal{L}_{1}|_{\mathcal{A}^{\rm full}_{\rm m}(\Lambda)}A\right\rangle_{\sigma_{\beta}}\geq\sum_{j\in\Gamma_{\rm int}}-\left\langle A,\mathcal{L}_{\sigma^{x}_{j}}|_{\mathcal{A}^{\rm full}_{\rm m}(\Lambda)}A\right\rangle_{\sigma_{\beta}}\geq\left|\Gamma_{\rm int}\right|\cdot\frac{1}{2}\langle A,A\rangle_{\sigma_{\beta}}\,.

    for any A𝒜mfull(Λ)A\in\mathcal{A}^{\rm full}_{\rm m}(\Lambda), where the first step is by σjx0-\mathcal{L}_{\sigma_{j}^{x}}\succeq 0. It follows that

    1|𝒜mfull(Λ)|Γint|2.-\mathcal{L}_{1}|_{\mathcal{A}^{\rm full}_{\rm m}(\Lambda)}\succeq\frac{\left|\Gamma_{\rm int}\right|}{2}\,. (4.36)
  • If Γflip={1,2,,L2}\Gamma_{\rm flip}=\{1,2,\ldots,L^{2}\} (i.e., Λ=\Lambda=\emptyset), by (4.53), the master Hamiltonian of σjx-\mathcal{L}_{\sigma^{x}_{j}} on 𝒜mfull()\mathcal{A}^{\rm full}_{\rm m}(\emptyset) is

    Ij2[h++h200001100110000h++h2]IL2j.I_{j-2}\otimes\left[\begin{matrix}-\frac{h_{+}+h_{-}}{2}&0&0&0\\ 0&-1&1&0\\ 0&1&-1&0\\ 0&0&0&-\frac{h_{+}+h_{-}}{2}\\ \end{matrix}\right]\otimes I_{L^{2}-j}\,. (4.37)

    This case has already been discussed in [2] (see after Proposition 2), from which we have

    1|𝒜mfull()12.-\mathcal{L}_{1}|_{\mathcal{A}^{\rm full}_{\rm m}(\emptyset)}\succeq\frac{1}{2}\,. (4.38)
  • If Γab={1,2,,L2}\Gamma_{\rm ab}=\{1,2,\ldots,L^{2}\} (i.e., Λ=[L2]\Lambda=[L^{2}]), [2, Proposition 2] has given a spectral gap lower bound that exponentially decays in β\beta. However, this lower bound is far from sharp at low temperature. Next, we shall derive a lower bound of gap polynomially decaying in NN but independent of β\beta, which is summarized in the following lemma.

    Lemma 13.

    Given the notation above, we have, when β=Ω(lnN)\beta=\Omega(\ln N),

    Gap(1|𝒜mfull([L2]))=Ω(N3).\mathrm{Gap}\left(-\mathcal{L}_{1}\middle|_{\mathcal{A}^{\rm full}_{\rm m}\left([L^{2}]\right)}\right)=\Omega(N^{-3})\,.

Once Lemma 13 is proved, we can combine it with Lemma 12, as well as (4.36) and (4.38), to obtain the desired (4.33):

Gap(1|𝒜mfull)min{12,Gap(1|𝒜mfull([L2]))}=Ω(N3).\mathrm{Gap}\left(-\mathcal{L}_{1}|_{\mathcal{A}_{\rm m}^{\rm full}}\right)\geq\min\left\{\frac{1}{2},\mathrm{Gap}\left(-\mathcal{L}_{1}\middle|_{\mathcal{A}^{\rm full}_{\rm m}\left([L^{2}]\right)}\right)\right\}=\Omega(N^{-3})\,.

We next prove Lemma 13 to conclude the proof of Lemma 10. From (4.52), the matrix representation of the master Hamiltonian of 1-\mathcal{L}_{1} can be written as:

KL2β=KL2+(KL2βKL2),K_{L^{2}}^{\beta}=K_{L^{2}}+(K_{L^{2}}^{\beta}-K_{L^{2}})\,,

where KL2K_{L^{2}} is the matrix representation at zero temperature (i.e., β\beta\to\infty):

KL2=i=0L22IikTIL22i:=KT+i=0L22IikDIL22i:=KD,K_{L^{2}}=\underbrace{\sum^{L^{2}-2}_{i=0}I_{i}\otimes k_{T}\otimes I_{L^{2}-2-i}}_{:=K_{T}}+\underbrace{\sum^{L^{2}-2}_{i=0}I_{i}\otimes k_{D}\otimes I_{L^{2}-2-i}}_{:=K_{D}}\,, (4.39)

with

kT=[0000011001100000],kD=[0000000000000002].k_{T}=\left[\begin{matrix}0&0&0&0\\ 0&1&-1&0\\ 0&-1&1&0\\ 0&0&0&0\\ \end{matrix}\right],\quad k_{D}=\left[\begin{matrix}0&0&0&0\\ 0&0&0&0\\ 0&0&0&0\\ 0&0&0&2\\ \end{matrix}\right]\,. (4.40)

and KL2βKL2K_{L^{2}}^{\beta}-K_{L^{2}} is given by

KL2βKL2=i=0N2Ii[2η2η2+1002ηη2+1000000002ηη2+1002η2η2+1]IL22i,K_{L^{2}}^{\beta}-K_{L^{2}}=\sum^{N-2}_{i=0}I_{i}\otimes\left[\begin{matrix}\frac{2\eta^{2}}{\eta^{2}+1}&0&0&-\frac{2\eta}{\eta^{2}+1}\\ 0&0&0&0\\ 0&0&0&0\\ -\frac{2\eta}{\eta^{2}+1}&0&0&-\frac{2\eta^{2}}{\eta^{2}+1}\\ \end{matrix}\right]\otimes I_{L^{2}-2-i}\,,

where η=exp(2β)\eta=\exp(-2\beta). The operator norm of the self-adjoint operator KL2βKL2K_{L^{2}}^{\beta}-K_{L^{2}} can be directly estimated as

KL2βKL2=Θ(L2e2β)=𝒪(L8),when β5lnL .\left\lVert K_{L^{2}}^{\beta}-K_{L^{2}}\right\rVert=\Theta(L^{2}e^{-2\beta})=\mathcal{O}(L^{-8})\,,\quad\text{when $\beta\geq 5\ln L$\,.}

Therefore, to obtain the desired gap estimate (A.9), we only need to consider the gap at zero temperature and prove Gap(KL2)Ω(L6)\mathrm{Gap}(K_{L^{2}})\geq\Omega(L^{-6}).

Refer to caption
Figure 4: Action of 1\mathcal{L}_{1} (β\beta\to\infty) on Γab\Gamma_{\rm ab}. kTk_{T} and kDk_{D} are defined in (4.40). Each black dot represents a bond corresponding to the plaquette observable Zj\textbf{Z}_{j} ordered by the snake.

For this, we consider the following configuration space

𝒮L2=Span(𝒜L2)2L2,with 𝒜L2={+,}L2,\mathcal{S}_{L^{2}}=\mathrm{Span}\left(\mathcal{A}_{L^{2}}\right)\cong\mathbb{C}^{2^{L^{2}}},\quad\text{with $\mathcal{A}_{L^{2}}=\left\{+,-\right\}^{\otimes{L^{2}}}$}\,, (4.41)

with the space decomposition 𝒮L2=k𝒮L2,k\mathcal{S}_{L^{2}}=\bigoplus_{k}\mathcal{S}_{{L^{2}},k}, where 𝒮L2,k=Span(𝒜L2,k)\mathcal{S}_{{L^{2}},k}=\mathrm{Span}\left(\mathcal{A}_{{L^{2}},k}\right) with

𝒜L2,k:={a{+,}L2|a has k ``" signs}.\mathcal{A}_{{L^{2}},k}:=\left\{a\in\left\{+,-\right\}^{\otimes{L^{2}}}\leavevmode\nobreak\ \middle|\leavevmode\nobreak\ \text{$a$ has $k$ $``-"$ signs}\right\}\,.

It is clear from the construction that the action of 1-\mathcal{L}_{1} on 𝒜mfull(Λ)\mathcal{A}^{\rm full}_{\rm m}(\Lambda) with Λ=Γab={1,2,}\Lambda=\Gamma_{\rm ab}=\{1,2,\ldots\} is the same as the action of KL2βK^{\beta}_{L^{2}} on keven𝒮L2,k\bigoplus_{k\,\text{even}}\mathcal{S}_{{L^{2}},k} under the identification |mm||m\ket{m}\bra{m}\rightarrow\ket{m}. One can also readily check that

KL2:𝒮L2,k𝒮L2,k,K_{L^{2}}:\mathcal{S}_{L^{2},k}\to\mathcal{S}_{L^{2},k}\,, (4.42)

that is, KL2K_{L^{2}} is block diagonal for the decomposition 𝒮L2=k𝒮L2,k\mathcal{S}_{L^{2}}=\bigoplus_{k}\mathcal{S}_{L^{2},k}.

When k=0k=0, we have KL2|𝒜L2,0=0K_{L^{2}}|_{\mathcal{A}_{L^{2},0}}=0 and dim(𝒜L2,0)=1\mathrm{dim}\left(\mathcal{A}_{L^{2},0}\right)=1. Thus,

Gap(1|𝒜mfull([L2]))mink=2,3,4,λmin(KL2|𝒮L2,k).\mathrm{Gap}\left(-\mathcal{L}_{1}|_{\mathcal{A}^{\rm full}_{\rm m}([L^{2}])}\right)\geq\min_{k=2,3,4,\cdots}\lambda_{\min}\left(K_{L^{2}}|_{\mathcal{S}_{L^{2},k}}\right)\,.

In principle, we only need to consider the admissible configuration, that is, the subspaces 𝒮L2,k\mathcal{S}_{L^{2},k} with even kk. However, it is simpler to prove a result for all kk’s via iterative reduction. We emphasize that such a relaxation does not produce any additional dependence on the system size N=2L2N=2L^{2}, and hence will not give a worse spectral gap lower bound.

We now consider the lower bound of KL2K_{L^{2}} on each 𝒮L2,k\mathcal{S}_{L^{2},k} for k>0k>0. Define

λL2,k:=λmin(KL2|𝒮L2,k).\lambda_{L^{2},k}:=\lambda_{\min}\left(K_{L^{2}}|_{\mathcal{S}_{L^{2},k}}\right)\,.

We first note that when k=L2k=L^{2}, it also holds that dim(𝒮L2,L2)=1\dim\left(\mathcal{S}_{L^{2},L^{2}}\right)=1, and

KL2|𝒮L2,L2=KD|𝒮L2,L22(L21)1.K_{L^{2}}|_{\mathcal{S}_{L^{2},L^{2}}}=K_{D}|_{\mathcal{S}_{L^{2},L^{2}}}\succeq 2(L^{2}-1)\geq 1\,.

When k3k\geq 3 and L23L^{2}\geq 3, we use the following iteration to reduce the estimation of λL2,k\lambda_{L^{2},k} to λL2,2\lambda_{L^{2},2}. By the representation (4.39) of KL2K_{L^{2}}, we find

KL2=KL21I1+IL22(kT+kD).K_{L^{2}}=K_{L^{2}-1}\otimes I_{1}+I_{L^{2}-2}\otimes(k_{T}+k_{D})\,.

In addition, there holds

𝒮L2,k=Span(𝒜L21,k|+)Span(𝒜L21,k1|).\mathcal{S}_{L^{2},k}=\mathrm{Span}\left(\mathcal{A}_{L^{2}-1,k}\otimes\ket{+}\right)\oplus\mathrm{Span}\left(\mathcal{A}_{L^{2}-1,k-1}\otimes\ket{-}\right)\,.

Thus, for any given a=a+|++a|𝒮L2,ka=a_{+}\otimes\ket{+}+a_{-}\otimes\ket{-}\in\mathcal{S}_{L^{2},k} with |a+|2+|a|2=1|a_{+}|^{2}+|a_{-}|^{2}=1, we can derive

aKL2a\displaystyle a^{*}K_{L^{2}}a a(KL21I1)a\displaystyle\geq a^{*}(K_{L^{2}-1}\otimes I_{1})a (4.43)
=(a+)KL21a++(a)KL21a\displaystyle=(a_{+})^{*}K_{L^{2}-1}a_{+}+(a_{-})^{*}K_{L^{2}-1}a_{-}
min{λL21,k,λL21,k1},\displaystyle\geq\min\left\{\lambda_{L^{2}-1,k},\lambda_{L^{2}-1,k-1}\right\}\,,

where the first inequality follows from kT+kD0k_{T}+k_{D}\succeq 0, the second inequality follows from (4.42) and a+𝒮L21,ka_{+}\in\mathcal{S}_{L^{2}-1,k}, a𝒮L21,k1a_{-}\in\mathcal{S}_{L^{2}-1,k-1}.

Therefore, it suffices to estimate λL2,2\lambda_{L^{2},2} to finish. To do so, we find, by (4.39),

KL2|𝒜L2,2=(KT)|𝒜L2,2+(KD)|𝒜L2,2,K_{L^{2}}|_{\mathcal{A}_{L^{2},2}}=(K_{T})|_{\mathcal{A}_{L^{2},2}}+(K_{D})|_{\mathcal{A}_{L^{2},2}}\,,

and KT(φ)=0K_{T}(\varphi)=0 for φ𝒜L2,2\varphi\in\mathcal{A}_{L^{2},2} if and only if IikTIL22i(φ)=0I_{i}\otimes k_{T}\otimes I_{L^{2}-2-i}(\varphi)=0 for all ii, which implies

Ker(KT|𝒜L2,2)=Span{a𝒜L2,2a}.\mathrm{Ker}\left(K_{T}|_{\mathcal{A}_{L^{2},2}}\right)=\mathrm{Span}\bigg{\{}\sum_{a\in\mathcal{A}_{L^{2},2}}a\bigg{\}}\,.

Then, it follows that

(2L2(L21)a𝒜L2,2a)KD(2L2(L21)a𝒜L2,2a)=4L2(L21)(a𝒜L2,2a)(a𝒜L2,2a)=4L2,\bigg{(}\sqrt{\frac{2}{L^{2}(L^{2}-1)}}\sum_{a\in\mathcal{A}_{L^{2},2}}a\bigg{)}^{*}K_{D}\bigg{(}\sqrt{\frac{2}{L^{2}(L^{2}-1)}}\sum_{a\in\mathcal{A}_{L^{2},2}}a\bigg{)}\\ =\frac{4}{L^{2}(L^{2}-1)}\bigg{(}\sum_{a\in\mathcal{A}^{--}_{L^{2},2}}a\bigg{)}^{*}\bigg{(}\sum_{a\in\mathcal{A}^{--}_{L^{2},2}}a\bigg{)}=\frac{4}{L^{2}}\,,

where 𝒜L2,2:={a{+,}L2|a has 2 consecutive ``" signs}\mathcal{A}_{{L^{2}},2}^{--}:=\left\{a\in\left\{+,-\right\}^{\otimes{L^{2}}}\leavevmode\nobreak\ \middle|\leavevmode\nobreak\ \text{$a$ has $2$ consecutive $``-"$ signs}\right\}. By Lemma 3 (item 4) and KD|𝒜L2,2=𝒪(1)\|K_{D}|_{\mathcal{A}_{L^{2},2}}\|=\mathcal{O}(1), we have

KT|𝒜L2,2+KD|𝒜L2,2(4/L2)Gap(KT|𝒜L2,2)Gap(KT|𝒜L2,2)+𝒪(1).K_{T}|_{\mathcal{A}_{L^{2},2}}+K_{D}|_{\mathcal{A}_{L^{2},2}}\succeq\frac{(4/L^{2})\mathrm{Gap}\left(K_{T}|_{\mathcal{A}_{L^{2},2}}\right)}{\mathrm{Gap}\left(K_{T}|_{\mathcal{A}_{L^{2},2}}\right)+\mathcal{O}(1)}\,. (4.44)

We next estimate the spectral gap of (KT)|𝒜L2,2(K_{T})|_{\mathcal{A}_{L^{2},2}}. We note from (4.39)-(4.40) that KTK_{T} can be identified as a ferromagnetic Heisenberg-1/21/2 model on L2L^{2} particles:

KT=12j=1L21(σjxσj+1x+σjyσj+1y+σjzσj+1zI).K_{T}=-\frac{1}{2}\sum_{j=1}^{L^{2}-1}\left(\sigma_{j}^{x}\sigma_{j+1}^{x}+\sigma_{j}^{y}\sigma_{j+1}^{y}+\sigma_{j}^{z}\sigma_{j+1}^{z}-I\right)\,.

From [35, Proposition 2] for the Heisenberg model, we have

Gap((KT)|𝒜L2,2)Gap(KT)=Ω(L4).\mathrm{Gap}\left((K_{T})|_{\mathcal{A}_{L^{2},2}}\right)\geq\mathrm{Gap}\left(K_{T}\right)=\Omega(L^{-4})\,.

Plugging this back into (4.44), we have

λL2,2=Ω(L6).\lambda_{L^{2},2}=\Omega\left(L^{-6}\right)\,. (4.45)

Using Eq. 4.43, it follows that min3kL2λL2,k=Ω(L6)\min_{3\leq k\leq L^{2}}\lambda_{L^{2},k}=\Omega\left(L^{-6}\right) and Gap(1|𝒜mfull([L2]))=Ω(L6)\mathrm{Gap}\left(-\mathcal{L}_{1}|_{\mathcal{A}^{\rm full}_{\rm m}([L^{2}])}\right)=\Omega\left(L^{-6}\right). To summarize, there holds

Gap(1)min{12,Gap(1|𝒜mfull([L2]))}=Ω(L6)=Ω(N3),\mathrm{Gap}\left(-\mathcal{L}_{1}\right)\geq\min\left\{\frac{1}{2},\mathrm{Gap}\left(-\mathcal{L}_{1}|_{\mathcal{A}^{\rm full}_{\rm m}([L^{2}])}\right)\right\}=\Omega\left(L^{-6}\right)=\Omega\left(N^{-3}\right)\,, (4.46)

for β5lnL\beta\geq 5\ln L. This concludes the proof of Lemma 13, and consequently Lemma 10. ∎

We next prove Lemma 11, whose basic ideas are similar to that of Lemma 10 but require some more technical arguments due to the comb structure. To be specific, recall the formula (4.18), the local master Hamiltonian representation of σjz\mathcal{L}_{\sigma_{j}^{z}} is the same as that of σjx\mathcal{L}_{\sigma_{j}^{x}}. However, here 1\mathcal{L}_{1} acts on a 1D straight line (snake), where each observable Zp\textbf{Z}_{p} connects only two neighboring qubits along the chain, while 2\mathcal{L}_{2} acts on a comb-like 1D structure, where some observables Xs\textbf{X}_{s} can connect three neighboring qubits on the comb. This difference prevents us from directly applying the proof of Lemma 10.

Proof of Lemma 11.

The starting point of the proof of Lemma 11 follows from that of Lemma 10. We recall that 𝒜efull\mathcal{A}^{\rm full}_{\rm e} is spanned by the orthogonal basis |ee|\ket{e^{\prime}}\bra{e} for the GNS inner product induced by the reduced Gibbs state Tr𝒜mfull(σβ)exp(βsXs)\operatorname{Tr}_{\mathcal{A}^{\rm full}_{\rm m}}(\sigma_{\beta})\propto\exp(-\beta\sum_{s}\textbf{X}_{s}), where e,e{+,}L2e,e^{\prime}\in\{+,-\}^{L^{2}} and #{ei=1},#{ei=1}2\#\left\{e_{i}=-1\right\},\#\left\{e^{\prime}_{i}=-1\right\}\in 2\mathbb{Z}. Moreover, we decompose 𝒜efull=Λ𝒜efull(Λ)\mathcal{A}^{\rm full}_{\rm e}=\bigoplus_{\Lambda}\mathcal{A}^{\rm full}_{\rm e}(\Lambda), where Λ{1,2,,L2}\Lambda\subset\{1,2,\ldots,L^{2}\} with even L2|Λ|L^{2}-|\Lambda| and the subspace 𝒜efull(Λ)\mathcal{A}^{\rm full}_{\rm e}(\Lambda) is spanned by |ee|\ket{e^{\prime}}\bra{e}, where e=ee=e^{\prime} on Λ\Lambda and e=ee=-e^{\prime} on {1,2,,L2}\Λ\{1,2,\ldots,L^{2}\}\backslash\Lambda. We also partition pairs of neighboring bonds into three sets: Γflip\Gamma_{\rm flip} for both bonds in Λc\Lambda^{c}, Γab\Gamma_{\rm ab} for both bonds in Λ\Lambda, and Γint\Gamma_{\rm int} for one bond in Λ\Lambda and the other in Λc\Lambda^{c}.

Then, a similar lemma as Lemma 12 holds for 2\mathcal{L}_{2}, since its proof only needs the properties of local Lindbladians σjz\mathcal{L}_{\sigma_{j}^{z}} that are the same as those of σjx\mathcal{L}_{\sigma_{j}^{x}}.

Lemma 14.

2\mathcal{L}_{2}, defined in (4.32), on 𝒜efull\mathcal{A}^{\rm full}_{\rm e} is block diagonal for the decomposition:

𝒜efull=Λ{1,,L2}:L2Λeven𝒜efull(Λ).\mathcal{A}^{\rm full}_{\rm e}=\bigoplus_{\begin{subarray}{c}\Lambda\subset\{1,\dots,L^{2}\}:\\ L^{2}-\Lambda\leavevmode\nobreak\ \text{even}\end{subarray}}\mathcal{A}^{\rm full}_{\rm e}(\Lambda)\,.

Next, we consider three cases: 1. Γint\Gamma_{\rm int}\neq\emptyset; 2. Γflip={1,,L2}:=[L2]\Gamma_{\rm flip}=\{1,\ldots,L^{2}\}:=[L^{2}]; 3. Γab=[L2]\Gamma_{\rm ab}=[L^{2}]. Noting that again, the arguments of (4.36) and (4.38) only uses local Lindbladians σjx\mathcal{L}_{\sigma_{j}^{x}}, we have similar estimates for 2\mathcal{L}_{2}:

2|𝒜efull(Λ),Γint12and2|𝒜efull()12.-\mathcal{L}_{2}|_{\mathcal{A}_{\rm e}^{\rm full}(\Lambda),\,\Gamma_{\rm int}\neq\emptyset}\succeq\frac{1}{2}\quad\text{and}\quad-\mathcal{L}_{2}|_{\mathcal{A}_{\rm e}^{\rm full}(\emptyset)}\succeq\frac{1}{2}\,.

We now consider the third case Γab=[L2]\Gamma_{\rm ab}=[L^{2}] and prove the following result, which finishes the proof of Lemma 11.

Lemma 15.

Given the notation above, we have, when β=Ω(lnN)\beta=\Omega(\ln N),

Gap(2|𝒜mfull([L2]))=Ω(N2.5).\mathrm{Gap}\left(-\mathcal{L}_{2}\middle|_{\mathcal{A}^{\rm full}_{\rm m}\left([L^{2}]\right)}\right)=\Omega(N^{-2.5})\,.

Following the notation in the proof of Lemma 13, we still consider the subspace 𝒮L2\mathcal{S}_{L^{2}} in (4.41) and denote by KL2βK^{\beta}_{L^{2}} the matrix representation of the master Hamiltonian of 2-\mathcal{L}_{2}. Moreover, we similarly have

KL2βKL2=Θ(L2e2β)=𝒪(L8),for β5lnL,\left\lVert K_{L^{2}}^{\beta}-K_{L^{2}}\right\rVert=\Theta(L^{2}e^{-2\beta})=\mathcal{O}(L^{-8})\,,\quad\text{for $\beta\geq 5\ln L$}\,,

where KL2:=KL2K_{L^{2}}:=K^{\infty}_{L^{2}}. One can also see that each local term in KL2K_{L^{2}} has the same form as the one in (4.40), but the tensor structure is different222Since the qubits and star observables on the comb cannot be ordered along a line, some local term in KL2K_{L^{2}} is of the form IaIbII\otimes a\otimes I\otimes b\otimes I with a,ba,b being non-identity 2×22\times 2 matrix. Since some observable Xs\textbf{X}_{s} is altered by three σjz\mathcal{L}_{\sigma_{j}^{z}}, there are some sites where we find three local terms in KL2K_{L^{2}} nontrivially acting on it.; see Fig. 5. We then decompose 𝒮L2\mathcal{S}_{L^{2}} according to number of ``"``-" signs in the entry of basis:

𝒮L2=k𝒮L2,k,𝒮L2,k=Span(𝒜L2,k),\mathcal{S}_{L^{2}}=\bigoplus_{k}\mathcal{S}_{L^{2},k}\,,\quad\mathcal{S}_{{L^{2}},k}=\mathrm{Span}\left(\mathcal{A}_{{L^{2}},k}\right)\,, (4.47)

with

𝒜L2,k:={a{+,}L2|a has k ``" signs}.\mathcal{A}_{{L^{2}},k}:=\left\{a\in\left\{+,-\right\}^{\otimes{L^{2}}}\middle|\text{$a$ has $k$ $``-"$ signs}\right\}\,.

Then, KL2K_{L^{2}} is block diagonal for (4.47) and

Gap(KL2)mink=2,3,4,λmin(KL2|𝒮L2,k).\mathrm{Gap}\left(K_{L^{2}}\right)\geq\min_{k=2,3,4,\cdots}\lambda_{\min}\left(K_{L^{2}}|_{\mathcal{S}_{{L^{2}},k}}\right)\,.

We first consider the subspace 𝒮L2,2\mathcal{S}_{{L^{2}},2}, whose basis vectors contain only two ``"``-" signs. Our strategy for lower bounding λmin(KL2|𝒮L2,2)\lambda_{\min}\left(K_{L^{2}}|_{\mathcal{S}_{{L^{2}},2}}\right) is to reduce this problem to a straight line case as in Fig. 4. For this, we introduce a set of lines that covering all vertices in the comb. More specifically, we may regard the comb as a connected graph (more precisely, a tree) with L2L^{2} vertices of degree at most 3, each corresponding to a star Xs\textbf{X}_{s} that interacts the comb. Let 𝐃comb{\bf D}_{\rm comb} be the set of degree-one vertices (i.e., end-points) in the comb, and lu,vl_{u,v} be the shortest path between the vertices uu and vv in the comb. Define

𝐏comb:={lu,v:uv𝐃comb}.\displaystyle{\bf P}_{\rm comb}:=\{l_{u,v}:\forall\leavevmode\nobreak\ u\neq v\in{\bf D}_{\rm comb}\}\,. (4.48)

Then, we know that 𝐏comb{\bf P}_{\rm comb} contains L(L1)/2=𝒪(L2)L(L-1)/2=\mathcal{O}(L^{2}) (simple) paths, and the maximum length of the paths is comb=3L2\ell_{\rm comb}=3L-2. Two examples of these paths are given in Fig. 6.

For an arbitrary unit vector α𝒮L2,2\alpha\in\mathcal{S}_{{L^{2}},2}:

α=a𝒜L2,2pa|a,\displaystyle\alpha=\sum_{a\in\mathcal{A}_{L^{2},2}}p_{a}\ket{a}\,,

by 16, there exists a path l~\tilde{l} in 𝐏comb{\bf P_{\rm comb}} such that for

αl~:=a𝒜L2,k:#{sl~:as=``"}>1pa|a,\displaystyle\alpha_{\tilde{l}}:=\sum_{a\in\mathcal{A}_{L^{2},k}:\#\{s\in\tilde{l}:a_{s}=``-"\}>1}p_{a}\ket{a}\,,

it holds that

αl~22=Ω(1/L2).\displaystyle\|\alpha_{\tilde{l}}\|_{2}^{2}=\Omega(1/L^{2})\,. (4.49)

Let (KL2)l~(K_{L^{2}})_{\tilde{l}} be the restriction of KL2K_{L^{2}} to the path l~\tilde{l}. More specifically,

(KL2)l~:=e=(u,v)l~(kT+kD)u,v,\displaystyle(K_{L^{2}})_{\tilde{l}}:=\sum_{e=(u,v)\in\tilde{l}}(k_{T}+k_{D})_{u,v}\,,

where (kT+kD)u,v(k_{T}+k_{D})_{u,v} is a local term that applies kT+kDk_{T}+k_{D} to the “qubits” at uu and vv. Then, we have

αKL2α\displaystyle\alpha^{*}K_{L^{2}}\alpha\geq α(KL2)l~α\displaystyle\leavevmode\nobreak\ \alpha^{*}(K_{L^{2}})_{\tilde{l}}\alpha (4.50)
\displaystyle\geq αl~(KL2)l~αl~\displaystyle\leavevmode\nobreak\ \alpha^{*}_{\tilde{l}}(K_{L^{2}})_{\tilde{l}}\alpha_{\tilde{l}}
\displaystyle\geq αl~22λcomb,2\displaystyle\leavevmode\nobreak\ \|\alpha_{\tilde{l}}\|_{2}^{2}\cdot\lambda_{\ell_{\rm comb},2}
=\displaystyle= Ω(L2)Ω(L3)=Ω(L5),\displaystyle\leavevmode\nobreak\ \Omega(L^{-2})\cdot\Omega(L^{-3})=\Omega(L^{-5})\,,

where the first step follows from KL2(KL2)l~K_{L^{2}}-(K_{L^{2}})_{\tilde{l}} is positive semi-definite, the second step follows from ααl~,(KL2)l~αl~=0\langle\alpha-\alpha_{\tilde{l}},(K_{L^{2}})_{\tilde{l}}\alpha_{\tilde{l}}\rangle=0, the third step follows from (KL2)l~(K_{L^{2}})_{\tilde{l}} is equivalent to a 1D chain of length |l~|comb=O(L)|\tilde{l}|\leq\ell_{\rm comb}=O(L) and al~𝒮|l~|,2a_{\tilde{l}}\in\mathcal{S}_{|\tilde{l}|,2}, and the last step follows from (4.49) and (4.45).

Thus, we conclude

KL2|𝒮L2,2=Ω(L5)=Ω(N2.5).K_{L^{2}}|_{\mathcal{S}_{{L^{2}},2}}=\Omega(L^{-5})=\Omega(N^{-2.5})\,.

We proceed to estimate λmin(KL2|𝒮L2,k)\lambda_{\min}\big{(}K_{L^{2}}|_{\mathcal{S}_{{L^{2}},k}}\big{)} for k3k\geq 3. For an arbitrary unit vector α𝒮L2,k\alpha\in\mathcal{S}_{{L^{2}},k}:

α=a𝒜L2,kpa|a,\displaystyle\alpha=\sum_{a\in\mathcal{A}_{L^{2},k}}p_{a}\ket{a}\,,

by 16, there exists a path l~\tilde{l} in 𝐏comb{\bf P_{\rm comb}} such that for

αl~,q:=a𝒜L2,k:#{sl~:as=``"}=qpa|a2qk,\displaystyle\alpha_{\tilde{l},q}:=\sum_{\begin{subarray}{c}a\in\mathcal{A}_{L^{2},k}:\\ \#\{s\in\tilde{l}:a_{s}=``-"\}=q\end{subarray}}p_{a}\ket{a}\leavevmode\nobreak\ \leavevmode\nobreak\ \leavevmode\nobreak\ \forall 2\leq q\leq k\,,

we have

q=2kαl~,q22=Ω(1/L2).\displaystyle\sum_{q=2}^{k}\|\alpha_{\tilde{l},q}\|_{2}^{2}=\Omega(1/L^{2})\,. (4.51)

Let (KL2)l~(K_{L^{2}})_{\tilde{l}} be the restriction of KL2K_{L^{2}} to the path l~\tilde{l}. Then, we find, similar to (4.50),

αKL2α\displaystyle\alpha^{*}K_{L^{2}}\alpha\geq α(KL2)l~α\displaystyle\leavevmode\nobreak\ \alpha^{*}(K_{L^{2}})_{\tilde{l}}\alpha
\displaystyle\geq q=2kαl~,q(KL2)l~αl~,q\displaystyle\leavevmode\nobreak\ \sum^{k}_{q=2}\alpha^{*}_{\tilde{l},q}(K_{L^{2}})_{\tilde{l}}\alpha_{\tilde{l},q}
\displaystyle\geq q=2kαl~,q22λcomb,q\displaystyle\leavevmode\nobreak\ \sum_{q=2}^{k}\|\alpha_{\tilde{l},q}\|_{2}^{2}\cdot\lambda_{\ell_{\rm comb},q}
\displaystyle\geq q=2kαl~,q22λcomb,2\displaystyle\leavevmode\nobreak\ \sum_{q=2}^{k}\|\alpha_{\tilde{l},q}\|_{2}^{2}\cdot\lambda_{\ell_{\rm comb},2}
=\displaystyle= Ω(L2)Ω(L3)=Ω(L5),\displaystyle\leavevmode\nobreak\ \Omega(L^{-2})\cdot\Omega(L^{-3})=\Omega(L^{-5})\,,

where the first step follows from KL2(KL2)l~K_{L^{2}}-(K_{L^{2}})_{\tilde{l}} is positive semi-definite, the second step follows from αl~,q1(KL2)l~αl~,q2=0\alpha_{\tilde{l},q_{1}}^{*}(K_{L^{2}})_{\tilde{l}}\alpha_{\tilde{l},q_{2}}=0 for q1q2q_{1}\neq q_{2}, the third step follows from (KL2)l~(K_{L^{2}})_{\tilde{l}} is equivalent to a 1D chain of length |l~|comb=O(L)|\tilde{l}|\leq\ell_{\rm comb}=O(L), and αl~,q\alpha_{\tilde{l},q} is in the subspace 𝒮|l~|,q\mathcal{S}_{|\tilde{l}|,q}, the fourth step follows from the recursion relation (4.43) that λcomb,qλcomb,2\lambda_{\ell_{\rm comb},q}\geq\lambda_{\ell_{\rm comb},2}, an the last step follow from (4.45) and (4.51). Thus, we have

KL2|𝒮L2,k=Ω(L5)=Ω(N2.5).\displaystyle K_{L^{2}}|_{\mathcal{S}_{{L^{2}},k}}=\Omega(L^{-5})=\Omega(N^{-2.5})\,.

Combining them all together, we now can conclude

Gap(2|𝒜efull([L2]))min2kL2λmin(KL2|𝒮L2,k)𝒪(N4)=Ω(N2.5).\displaystyle\mathrm{Gap}\left(-\mathcal{L}_{2}|_{\mathcal{A}_{\rm e}^{\rm full}([L^{2}])}\right)\geq\min_{2\leq k\leq L^{2}}\lambda_{\min}\left(K_{L^{2}}|_{\mathcal{S}_{{L^{2}},k}}\right)-\mathcal{O}(N^{-4})=\Omega(N^{-2.5})\,.

This concludes the proof of Lemma 15, and thus the proof of Lemma 11. ∎

Fact 16.

Let 2kL212\leq k\leq L^{2}-1. For any unit vector α𝒮L2,k\alpha\in\mathcal{S}_{L^{2},k} of the form:

α=a𝒜L2,kpa|a,\displaystyle\alpha=\sum_{a\in\mathcal{A}_{L^{2},k}}p_{a}\ket{a}\,,

there exists a path l~𝐏comb\tilde{l}\in{\bf P_{\rm comb}} defined as (4.48) such that

a𝒜L2,k:#{sl~:as=``"}>1|pa|2=Ω(1/L).\displaystyle\sum_{\begin{subarray}{c}a\in\mathcal{A}_{L^{2},k}:\\ \#\{s\in\tilde{l}:a_{s}=``-"\}>1\end{subarray}}|p_{a}|^{2}=\Omega(1/L)\,.
Proof.

We first observe that for any a{+,}L2a\in\{+,-\}^{L^{2}} with kk ``"``-", by the construction of 𝐏comb{\bf P}_{\rm comb}, there exists at least one path l𝐏combl\in{\bf P}_{\rm comb} that contains at last two ``"``-" on it. Then, we have

l𝐏comba𝒜L2,k:#{sl:as=``"}>1|pa|2=\displaystyle\sum_{l\in{\bf P}_{\rm comb}}\sum_{\begin{subarray}{c}a\in\mathcal{A}_{L^{2},k}:\\ \#\{s\in l:a_{s}=``-"\}>1\end{subarray}}|p_{a}|^{2}= a𝒜L2,k|pa|2#{l𝐏comb:#{sl:as=``"}>1}\displaystyle\leavevmode\nobreak\ \sum_{a\in\mathcal{A}_{L^{2},k}}|p_{a}|^{2}\cdot\#\{l\in{\bf P}_{\rm comb}:\#\{s\in l:a_{s}=``-"\}>1\}
\displaystyle\geq a𝒜L2,k|pa|21\displaystyle\leavevmode\nobreak\ \sum_{a\in\mathcal{A}_{L^{2},k}}|p_{a}|^{2}\cdot 1
=\displaystyle= 1,\displaystyle\leavevmode\nobreak\ 1\,,

where the first step follows from exchanging the summations, the second step follows from our observation, and the last step follows from α\alpha is a unit vector. Since 𝐏comb{\bf P}_{\rm comb} contains L2L^{2} paths, by the averaging argument, there must exists an l~𝐏comb\tilde{l}\in{\bf P}_{\rm comb} such that

a𝒜L2,k:#{sl:as=``"}>1|pa|21/L,\displaystyle\sum_{\begin{subarray}{c}a\in\mathcal{A}_{L^{2},k}:\\ \#\{s\in l:a_{s}=``-"\}>1\end{subarray}}|p_{a}|^{2}\geq 1/L\,,

which proves the proposition. ∎

Refer to caption
Figure 5: The underlying graph of jcombσjz\sum_{j\in\rm comb}\mathcal{L}_{\sigma^{z}_{j}} acting on 𝒜efull\mathcal{A}^{\rm full}_{\rm e}. Here, each black dot represents a qubit corresponding to a bond observation of Xsj\textbf{X}_{s_{j}}. The transition matrix kT+kDk_{T}+k_{D} induced by σjz\mathcal{L}_{\sigma^{z}_{j}} acts on each neighboring qubits.
Refer to caption
Figure 6: Examples of lines on the graph.

4.3 Proof of Lemmas

We collect proofs of some technical lemmas for the spectral gap analysis.

Proof of Lemma 5.

The decomposition (4.11) is straightforward by the construction. Here we only prove the fact that the algebra 𝒜mfull=(bm)\mathcal{A}^{\rm full}_{\rm m}=\mathcal{B}(\mathcal{H}_{\rm b}^{\rm m}) can be generated by {𝐙p}p{σjx}jsnake\{\mathbf{Z}_{p}\}_{p}\cup\{\sigma^{x}_{j}\}_{j\in\text{\rm snake}}. The claim for 𝒜efull\mathcal{A}^{\rm full}_{\rm e} can be similarly proved. Indeed, each basis (admissible bond) |m\ket{m} in bm\mathcal{H}_{\rm b}^{\rm m} can be identified as a Pauli string σj1xσjkx\sigma^{x}_{j_{1}}\cdots\sigma^{x}_{j_{k}} with jij_{i} on the snake. Each operator on bm\mathcal{H}_{\rm b}^{\rm m} can be written as the linear combination of |mm|\ket{m^{\prime}}\bra{m}, where |m\ket{m} and |m\ket{m^{\prime}} are admissible bonds. Meanwhile, each |mm|\ket{m^{\prime}}\bra{m} with |m|m\ket{m^{\prime}}\neq\ket{m} is a composition of flips of neighboring states. Thus, it suffices to consider the case |mm|\ket{m^{\prime}}\bra{m} that maps the state |m=σj1xσjkx|+1L2\ket{m}=\sigma^{x}_{j_{1}}\cdots\sigma^{x}_{j_{k}}\ket{+1^{L^{2}}}333More rigorously, |m=σj1xσjkx|+1L2\ket{m}=\sigma^{x}_{j_{1}}\cdots\sigma^{x}_{j_{k}}\ket{+1^{L^{2}}} is defined as the bond configuration for {Zp}p\{\textbf{Z}_{p}\}_{p} associated with the spin configuration σj1xσjkx|ψo\sigma^{x}_{j_{1}}\cdots\sigma^{x}_{j_{k}}\ket{\psi_{o}}, where |ψo\ket{\psi_{o}} is a ground state (4.5). to a neighboring one |m=σj2xσjkx|+1L2\ket{m^{\prime}}=\sigma^{x}_{j_{2}}\cdots\sigma^{x}_{j_{k}}\ket{+1^{L^{2}}}, which can be easily constructed using {Zp}p\{\textbf{Z}_{p}\}_{p} and {σjx}jsnake\{\sigma^{x}_{j}\}_{j\in{\rm snake}}. Here we give the representation for |mm|\ket{m^{\prime}}\bra{m}:

|mm|=Imj1Zj12|m0,|m|m(jj11,j1L2I+mjZj2)σj1x=|mm|+|mm|.\ket{m^{\prime}}\bra{m}=\underbrace{\frac{I-m_{j_{1}}\textbf{Z}_{j_{1}}}{2}}_{\ket{m}\rightarrow 0,\ \ket{m^{\prime}}\rightarrow\ket{m^{\prime}}}\cdot\underbrace{\Big{(}\prod_{j\neq j_{1}-1,j_{1}}^{L^{2}}\frac{I+m_{j}\textbf{Z}_{j}}{2}\Big{)}\sigma^{x}_{j_{1}}}_{=\ket{m^{\prime}}\bra{m}+\ket{m}\bra{m^{\prime}}}\,.

The case of |mm|\ket{m}\bra{m} can be similarly done. ∎

Proof of Lemma 6.

The formula (4.22) follows from the representation of 𝖮\mathcal{L}_{\mathsf{O}} with 𝖮=𝖷1,𝖷2,𝖹1,𝖹2\mathsf{O}=\mathsf{X}_{1},\mathsf{X}_{2},\mathsf{Z}_{1},\mathsf{Z}_{2} and the fact from (5) that these global jumps commute with the algebras 𝒜m/efull\mathcal{A}_{\rm m/e}^{\rm full}. For the formula (4.22), it suffices to note that {σjx}jsnake\{\sigma_{j}^{x}\}_{j\in{\rm snake}} and the projections (4.14) belong to 𝒜mfull\mathcal{A}_{\rm m}^{\rm full} and commute with 𝒬1𝒬2𝒜efull\mathcal{Q}_{1}\otimes\mathcal{Q}_{2}\otimes\mathcal{A}_{\rm e}^{\rm full}. The formula (4.23) follows from the same reason by the computation (4.18). The first two statements of (4.24) can be proved by a very similar argument as [2, Lemma 6]. To show Ker(global)=𝟏𝟏𝒜mfull𝒜efull\mathrm{Ker}\left(\mathcal{L}_{\rm global}\right)={\bf 1}\otimes{\bf 1}\otimes\mathcal{A}^{\rm full}_{\rm m}\otimes\mathcal{A}^{\rm full}_{\rm e}, we only need to note that operators 𝖷1,𝖷2,𝖹1,𝖹2\mathsf{X}_{1},\mathsf{X}_{2},\mathsf{Z}_{1},\mathsf{Z}_{2} span the whole algebra 𝒬1𝒬2\mathcal{Q}_{1}\otimes\mathcal{Q}_{2}. ∎

Proof of Lemma 12.

We first recall that we order the plaquette observables {Zp}p\{\textbf{Z}_{p}\}_{p} as {Zj}j=1L2\{\textbf{Z}_{j}\}_{j=1}^{L^{2}} and index {σjx}j=2L2\{\sigma_{j}^{x}\}_{j=2}^{L^{2}} along the snake. We note from (4.14) that the projections, as elements in 𝒜mfull\mathcal{A}^{\rm full}_{\rm m}, can be represented as

Pj0=(|++|+|++|)j1,j,Pj±=(||)j1,j,P_{j}^{0}=\left(\ket{+-}\bra{+-}+\ket{-+}\bra{-+}\right)_{j-1,j},\ P_{j}^{\pm}=\left(\ket{\mp}\bra{\mp}\right)_{j-1,j}\,,

where the subscript means that the operator only nontrivially acts on bonds j1j-1 and jj on the snake associated with observables {Zp}p\{\textbf{Z}_{p}\}_{p}. This enables us to compute the jumps aj,aj,aj0𝒜mfulla_{j},a_{j}^{\dagger},a_{j}^{0}\in\mathcal{A}^{\rm full}_{\rm m} defined in (4.15) for the local Lindbladian 1\mathcal{L}_{1} in (4.32):

aj=(|++|)j1,j,aj=(|++|)j1,j,,\displaystyle a_{j}=\left(\ket{++}\bra{--}\right)_{j-1,j}\,,\quad a_{j}^{\dagger}=\left(\ket{--}\bra{++}\right)_{j-1,j}\,,\,,

and

aj0=(|++|+|++|)j1,j.a_{j}^{0}=\left(\ket{+-}\bra{-+}+\ket{-+}\bra{+-}\right)_{j-1,j}\,.

Without loss of generality, we consider the Lindbladian σjx\mathcal{L}_{\sigma^{x}_{j}} in (4.16) for a fixed jΓflip,Γab,Γintj\in\Gamma_{\rm flip},\Gamma_{\rm ab},\Gamma_{\rm int} on 𝒜mfull(Λ)\mathcal{A}^{\rm full}_{\rm m}(\Lambda) for each Λ\Lambda. Due to the locality of the jump operators, σjx\mathcal{L}_{\sigma^{x}_{j}} only changes the pair of bonds associated with jj. For simplicity, we shall omit the subscripts j1,jj-1,j.

  • For jΓabj\in\Gamma_{\rm ab}, we consider the local basis

    A1=|++++|,A2=|++|,A3=|++|,A4=||,\displaystyle A_{1}=\ket{++}\bra{++}\,,\quad A_{2}=\ket{+-}\bra{+-}\,,\quad A_{3}=\ket{-+}\bra{-+},\quad A_{4}=\ket{--}\bra{--}\,,

    that are orthogonal in both HS and GNS inner products for the reduced Gibbs state σ~β=Tr𝒜efull(σβ)exp(βpZp)\widetilde{\sigma}_{\beta}=\operatorname{Tr}_{\mathcal{A}^{\rm full}_{\rm e}}(\sigma_{\beta})\propto\exp(-\beta\sum_{p}\textbf{Z}_{p}). Moreover, we compute

    eβpZpA1=η1A1,eβpZpA2=A2,\displaystyle e^{-\beta\sum_{p}\textbf{Z}_{p}}A_{1}=\eta^{-1}A_{1}\,,\quad e^{-\beta\sum_{p}\textbf{Z}_{p}}A_{2}=A_{2}\,,
    eβpZpA3=A3,eβpZpA4=ηA4,\displaystyle e^{-\beta\sum_{p}\textbf{Z}_{p}}A_{3}=A_{3}\,,\quad e^{-\beta\sum_{p}\textbf{Z}_{p}}A_{4}=\eta A_{4}\,,

    where η=e2β\eta=e^{-2\beta}, which implies that

    A1σ~β=𝒵β1/2η1/2,A2σ~β=A3σ~β=𝒵β1/2,A4σ~β=𝒵β1/2η1/2,\displaystyle\|A_{1}\|_{\widetilde{\sigma}_{\beta}}=\mathcal{Z}_{\beta}^{-1/2}\eta^{-1/2}\,,\quad\|A_{2}\|_{\widetilde{\sigma}_{\beta}}=\|A_{3}\|_{\widetilde{\sigma}_{\beta}}=\mathcal{Z}_{\beta}^{-1/2}\,,\quad\|A_{4}\|_{\widetilde{\sigma}_{\beta}}=\mathcal{Z}_{\beta}^{-1/2}\eta^{1/2}\,,

    with 𝒵β\mathcal{Z}_{\beta} being the partition function of σβ\sigma_{\beta}. Then, we find, by using (4.16),

    σjx(A1)\displaystyle\mathcal{L}_{\sigma^{x}_{j}}(A_{1}) =h+||h|++++|=h+A4hA1\displaystyle=h_{+}\ket{--}\bra{--}-h_{-}\ket{++}\bra{++}=h_{+}A_{4}-h_{-}A_{1}

    due to

    [A1,aj]=[|++++|,|++|]=|++|,\displaystyle[A_{1},a_{j}]=[\ket{++}\bra{++},\ket{++}\bra{--}]=\ket{++}\bra{--}\,,
    [A1,aj]=[|++++|,|++|]=|++|.\displaystyle[A_{1},a_{j}^{\dagger}]=[\ket{++}\bra{++},\ket{--}\bra{++}]=-\ket{--}\bra{++}\,.

    Similarly, we have σjx(A4)=h+A4+hA1\mathcal{L}_{\sigma^{x}_{j}}(A_{4})=-h_{+}A_{4}+h_{-}A_{1}, by

    [A4,aj]=[||,|++|]=|++|,\displaystyle[A_{4},a_{j}]=[\ket{--}\bra{--},\ket{++}\bra{--}]=-\ket{++}\bra{--}\,,
    [A4,aj]=[||,|++|]=|++|.\displaystyle[A_{4},a_{j}^{\dagger}]=[\ket{--}\bra{--},\ket{--}\bra{++}]=\ket{--}\bra{++}\,.

    In the same way, we can also compute

    σjx(A2)=A3A2,σjx(A3)=A2A3.\mathcal{L}_{\sigma^{x}_{j}}(A_{2})=A_{3}-A_{2}\,,\quad\mathcal{L}_{\sigma^{x}_{j}}(A_{3})=A_{2}-A_{3}\,.

    This allows us to compute the local matrix representation of the master Hamiltonian φ1φ1\varphi\circ\mathcal{L}_{1}\circ\varphi^{-1} via Ai,1Ajσ~βAiσ~βAjσ~β\frac{\langle A_{i},\mathcal{L}_{1}A_{j}\rangle_{\widetilde{\sigma}_{\beta}}}{\left\lVert A_{i}\right\rVert_{\widetilde{\sigma}_{\beta}}\left\lVert A_{j}\right\rVert_{\widetilde{\sigma}_{\beta}}}:

    [h=2η2η2+100h/η=2ηη2+101100110ηh+=2ηη2+100h+=2η2+1].\left[\begin{matrix}-h_{-}=-\frac{2\eta^{2}}{\eta^{2}+1}&0&0&h_{-}/\eta=\frac{2\eta}{\eta^{2}+1}\\ 0&-1&1&0\\ 0&1&-1&0\\ \eta h_{+}=\frac{2\eta}{\eta^{2}+1}&0&0&-h_{+}=-\frac{2}{\eta^{2}+1}\\ \end{matrix}\right]\,. (4.52)
  • For jΓflipj\in\Gamma_{\rm flip}, we let

    A1=|++|,A2=|++|,A3=|++|,A4=|++|.\displaystyle A_{1}=\ket{++}\bra{--}\,,\quad A_{2}=\ket{+-}\bra{-+}\,,\quad A_{3}=\ket{-+}\bra{+-}\,,\quad A_{4}=\ket{--}\bra{++}\,.

    For A1,A4A_{1},A_{4}, noting that

    [A1,aj]=[|++|,|++|]=0,\displaystyle[A_{1},a_{j}]=[\ket{++}\bra{--},\ket{++}\bra{--}]=0\,,
    [A1,aj]=[|++|,|++|]=|++++|||,\displaystyle[A_{1},a_{j}^{\dagger}]=[\ket{++}\bra{--},\ket{--}\bra{++}]=\ket{++}\bra{++}-\ket{--}\bra{--}\,,

    and

    [A4,aj]=[|++|,|++|]=|||++++|,\displaystyle[A_{4},a_{j}]=[\ket{--}\bra{++},\ket{++}\bra{--}]=\ket{--}\bra{--}-\ket{++}\bra{++}\,,
    [A4,aj]=[|++|,|++|]=0,\displaystyle[A_{4},a_{j}^{\dagger}]=[\ket{--}\bra{++},\ket{--}\bra{++}]=0\,,

    we have

    σjx(A1)\displaystyle\mathcal{L}_{\sigma^{x}_{j}}(A_{1}) =12{h+[aj,A1]aj+haj[A1,aj]}=12(h++h)A1,\displaystyle=\frac{1}{2}\left\{h_{+}[a_{j}^{\dagger},A_{1}]a_{j}+h_{-}a_{j}[A_{1},a_{j}^{\dagger}]\right\}=-\frac{1}{2}(h_{+}+h_{-})A_{1}\,,

    and

    σjx(A4)=12{h+aj[A4,aj]+h[aj,A4]aj}=12(h++h)A4.\displaystyle\mathcal{L}_{\sigma^{x}_{j}}(A_{4})=\frac{1}{2}\left\{h_{+}a_{j}^{\dagger}[A_{4},a_{j}]+h_{-}[a_{j},A_{4}]a_{j}^{\dagger}\right\}=-\frac{1}{2}(h_{+}+h_{-})A_{4}\,.

    Similarly, a direct computation also gives

    σjx(A2)=A3A2,σjx(A3)=A2A3.\mathcal{L}_{\sigma^{x}_{j}}(A_{2})=A_{3}-A_{2}\,,\quad\mathcal{L}_{\sigma^{x}_{j}}(A_{3})=A_{2}-A_{3}\,.

    The local matrix representation of φ1φ1\varphi\circ\mathcal{L}_{1}\circ\varphi^{-1} via Ai,1Ajσ~βAiσ~βAjσ~β\frac{\langle A_{i},\mathcal{L}_{1}A_{j}\rangle_{\widetilde{\sigma}_{\beta}}}{\left\lVert A_{i}\right\rVert_{\widetilde{\sigma}_{\beta}}\left\lVert A_{j}\right\rVert_{\widetilde{\sigma}_{\beta}}} is given by

    [h++h200001100110000h++h2].\left[\begin{matrix}-\frac{h_{+}+h_{-}}{2}&0&0&0\\ 0&-1&1&0\\ 0&1&-1&0\\ 0&0&0&-\frac{h_{+}+h_{-}}{2}\\ \end{matrix}\right]\,. (4.53)
  • For jΓintj\in\Gamma_{\rm int}, we let444Without loss of generality, we place jΛj\in\Lambda in the second position. The other case of j1Λj-1\in\Lambda is symmetric.

    A1=|+++|,A2=|+|,A3=|+++|,A4=|+|.\displaystyle A_{1}=\ket{++}\bra{-+}\,,\quad A_{2}=\ket{+-}\bra{--}\,,\quad A_{3}=\ket{-+}\bra{++}\,,\quad A_{4}=\ket{--}\bra{+-}\,.

    and find that they are eigenvectors of σjx\mathcal{L}_{\sigma^{x}_{j}}:

    σjx(A1)=h+12A1,σjx(A2)=h++12A2,\mathcal{L}_{\sigma^{x}_{j}}(A_{1})=-\frac{h_{-}+1}{2}A_{1}\,,\quad\mathcal{L}_{\sigma^{x}_{j}}\left(A_{2}\right)=-\frac{h_{+}+1}{2}A_{2}\,,

    and

    σjx(A3)=h+12A3,σjx(A4)=h++12A4.\mathcal{L}_{\sigma^{x}_{j}}(A_{3})=-\frac{h_{-}+1}{2}A_{3},\quad\mathcal{L}_{\sigma^{x}_{j}}(A_{4})=-\frac{h_{+}+1}{2}A_{4}\,.

    In this case, the local matrix representation of φ1φ1\varphi\circ\mathcal{L}_{1}\circ\varphi^{-1} via Ai,1Ajσ~βAiσ~βAjσ~β\frac{\langle A_{i},\mathcal{L}_{1}A_{j}\rangle_{\widetilde{\sigma}_{\beta}}}{\left\lVert A_{i}\right\rVert_{\widetilde{\sigma}_{\beta}}\left\lVert A_{j}\right\rVert_{\widetilde{\sigma}_{\beta}}} is

    [h+120000h++120000h+120000h++12].\left[\begin{matrix}-\frac{h_{-}+1}{2}&0&0&0\\ 0&-\frac{h_{+}+1}{2}&0&0\\ 0&0&-\frac{h_{-}+1}{2}&0\\ 0&0&0&-\frac{h_{+}+1}{2}\\ \end{matrix}\right]\,. (4.54)

The above calculation concludes the proof that 1\mathcal{L}_{1} is block diagonal for the decomposition of 𝒜mfull\mathcal{A}^{\rm full}_{\rm m} in (4.34), namely, 1:(+)(Λ)(+)(Λ)\mathcal{L}_{1}:\mathcal{B}(\mathcal{H}_{+})(\Lambda)\to\mathcal{B}(\mathcal{H}_{+})(\Lambda) for any Λ{1,2,,L2}\Lambda\subset\{1,2,\dots,L^{2}\} such that L2|Λ|L^{2}-|\Lambda| is even. ∎

References

  • ACL [23] Dong An, Andrew M Childs, and Lin Lin. Quantum algorithm for linear non-unitary dynamics with near-optimal dependence on all parameters. arXiv preprint arXiv:2312.03916, 2023.
  • AFH [09] R Alicki, M Fannes, and M Horodecki. On thermalization in kitaev’s 2d model. Journal of Physics A: Mathematical and Theoretical, 42(6):065303, January 2009.
  • AHHH [10] R. Alicki, M. Horodecki, P. Horodecki, and R. Horodecki. On thermal stability of topological qubit in kitaev’s 4d model. Open Systems and Information Dynamics, 17(01):1–20, 2010.
  • Bar [82] Francisco Barahona. On the computational complexity of ising spin glass models. Journal of Physics A: Mathematical and General, 15(10):3241, 1982.
  • BCG+ [23] Ivan Bardet, Ángela Capel, Li Gao, Angelo Lucia, David Pérez-García, and Cambyse Rouzé. Rapid thermalization of spin chain commuting Hamiltonians. Phys. Rev. Lett., 130(6):060401, 2023.
  • BCG+ [24] Ivan Bardet, Ángela Capel, Li Gao, Angelo Lucia, David Pérez-García, and Cambyse Rouzé. Entropy decay for davies semigroups of a one dimensional quantum lattice. Communications in Mathematical Physics, 405(2):42, 2024.
  • BCGW [21] Sergey Bravyi, Anirban Chowdhury, David Gosset, and Pawel Wocjan. On the complexity of quantum partition functions. arXiv preprint arXiv:2110.15466, 2021.
  • BCL [24] Thiago Bergamaschi, Chi-Fang Chen, and Yunchao Liu. Quantum computational advantage with constant-temperature gibbs sampling. arXiv preprint arXiv:2404.14639, 2024.
  • BCP+ [21] Antonio Blanca, Pietro Caputo, Daniel Parisi, Alistair Sinclair, and Eric Vigoda. Entropy decay in the swendsen–wang dynamics on Zd. In Proceedings of the 53rd Annual ACM SIGACT Symposium on Theory of Computing, pages 1551–1564, 2021.
  • BLMT [24] Ainesh Bakshi, Allen Liu, Ankur Moitra, and Ewin Tang. High-Temperature Gibbs States are Unentangled and Efficiently Preparable. arXiv preprint arXiv:2403.16850, 2024.
  • BLP+ [16] Benjamin J. Brown, Daniel Loss, Jiannis K. Pachos, Chris N. Self, and James R. Wootton. Quantum memories at finite temperature. Rev. Mod. Phys., 88:045005, Nov 2016.
  • Bom [13] Héctor Bombín. An introduction to topological quantum codes. arXiv preprint arXiv:1311.0277, 2013.
  • BW [00] Jeff P Barnes and Warren S Warren. Automatic quantum error correction. Phys. Rev. Lett., 85(4):856, 2000.
  • CB [21] Chi-Fang Chen and Fernando G.S.L. Brandão. Fast Thermalization from the Eigenstate Thermalization Hypothesis. arXiv preprint arXiv:2112.07646, 2021.
  • CGG+ [21] Zongchen Chen, Andreas Galanis, Leslie A Goldberg, Will Perkins, James Stewart, and Eric Vigoda. Fast algorithms at low temperatures via markov chains. Random Structures & Algorithms, 58(2):294–321, 2021.
  • CKBG [23] Chi-Fang Chen, Michael J Kastoryano, Fernando GSL Brandão, and András Gilyén. Quantum thermal state preparation. arXiv preprint arXiv:2303.18224, 2023.
  • CKG [23] Chi-Fang Chen, Michael J Kastoryano, and András Gilyén. An efficient and exact noncommutative quantum gibbs sampler. arXiv preprint arXiv:2311.09207, 2023.
  • CS [17] Anirban Narayan Chowdhury and Rolando D. Somma. Quantum algorithms for Gibbs sampling and hitting-time estimation. Quantum Inf. Comput., 17(1-2):41–64, 2017.
  • Dav [70] EB Davies. Quantum stochastic processes ii. Commun. Math. Phys., 19(2):83–105, 1970.
  • Dav [74] E. Brian Davies. Markovian master equations. Commun. Math. Phys., 39:91–110, 1974.
  • DLL [24] Zhiyan Ding, Bowen Li, and Lin Lin. Efficient quantum gibbs samplers with Kubo–Martin–Schwinger detailed balance condition. arXiv/2404.05998, 2024.
  • FHGW [14] C. Daniel Freeman, C. M. Herdman, D. J. Gorman, and K. B. Whaley. Relaxation dynamics of the toric code in contact with a thermal reservoir: Finite-size scaling in a low-temperature regime. Phys. Rev. B, 90:134302, Oct 2014.
  • Fre [18] Christian Daniel Freeman. The Toric Code at Finite Temperature. PhD thesis, UC Berkeley, 2018.
  • GJ [17] Heng Guo and Mark Jerrum. Random cluster dynamics for the ising model is rapidly mixing. In Proceedings of the Twenty-Eighth Annual ACM-SIAM Symposium on Discrete Algorithms, pages 1818–1827. SIAM, 2017.
  • GS [22] Reza Gheissari and Alistair Sinclair. Low-temperature ising dynamics with random initializations. In Proceedings of the 54th Annual ACM SIGACT Symposium on Theory of Computing, pages 1445–1458, 2022.
  • GSLW [19] András Gilyén, Yuan Su, Guang Hao Low, and Nathan Wiebe. Quantum singular value transformation and beyond: exponential improvements for quantum matrix arithmetics. In Proceedings of the 51st Annual ACM SIGACT Symposium on Theory of Computing, pages 193–204, 2019.
  • GŠV [19] Andreas Galanis, Daniel Štefankovič, and Eric Vigoda. Swendsen-wang algorithm on the mean-field potts model. Random Structures & Algorithms, 54(1):82–147, 2019.
  • HPR [19] Tyler Helmuth, Will Perkins, and Guus Regts. Algorithmic pirogov-sinai theory. In Proceedings of the 51st Annual ACM SIGACT Symposium on Theory of Computing, pages 1009–1020, 2019.
  • JS [93] Mark Jerrum and Alistair Sinclair. Polynomial-time approximation algorithms for the ising model. SIAM Journal on computing, 22(5):1087–1116, 1993.
  • KACR [24] Jan Kochanowski, Alvaro M. Alhambra, Angela Capel, and Cambyse Rouzé. Rapid thermalization of dissipative many-body dynamics of commuting hamiltonians. arXiv/2404.16780, 2024.
  • KB [16] Michael J. Kastoryano and Fernando G.S.L. Brandão. Quantum Gibbs samplers: The commuting case. Commun. Math. Phys., 344:915–957, 2016.
  • KFGV [77] Andrzej Kossakowski, Alberto Frigerio, Vittorio Gorini, and Maurizio Verri. Quantum detailed balance and KMS condition. Commun. Math. Phys., 57(2):97–110, 1977.
  • Kit [03] A Yu Kitaev. Fault-tolerant quantum computation by anyons. Annals of physics, 303(1):2–30, 2003.
  • KKCG [24] Michael J. Kastoryano, Lasse B. Kristensen, Chi-Fang Chen, and András Gilyén. A little bit of self-correction. arXiv/2408.14970, 2024.
  • KN [97] Tohru Koma and Bruno Nachtergaele. The spectral gap of the ferromagnetic XXZ-chain. Letters in Mathematical Physics, 40:1–16, 1997.
  • KT [13] Michael J Kastoryano and Kristan Temme. Quantum logarithmic Sobolev inequalities and rapid mixing. J. Math. Phys., 54(5):1–34, 2013.
  • LKV+ [13] Zaki Leghtas, Gerhard Kirchmair, Brian Vlastakis, Robert J Schoelkopf, Michel H Devoret, and Mazyar Mirrahimi. Hardware-efficient autonomous quantum memory protection. Phys. Rev. Lett., 111(12):120501, 2013.
  • LLG [24] Simon Lieu, Yu-Jie Liu, and Alexey V. Gorshkov. Candidate for a passively protected quantum memory in two dimensions. Phys. Rev. Lett., 133:030601, Jul 2024.
  • LS [16] Eyal Lubetzky and Allan Sly. Information percolation and cutoff for the stochastic ising model. Journal of the American Mathematical Society, 29(3):729–774, 2016.
  • MH [21] Ryan L Mann and Tyler Helmuth. Efficient algorithms for approximating quantum partition functions. Journal of Mathematical Physics, 62(2), 2021.
  • ML [20] Evgeny Mozgunov and Daniel Lidar. Completely positive master equation for arbitrary driving and small level spacing. Quantum, 4(1):1–62, 2020.
  • MO [94] Fabio Martinelli and Enzo Olivieri. Approach to equilibrium of glauber dynamics in the one phase region: I. the attractive case. Communications in Mathematical Physics, 161(3):447–486, 1994.
  • Pis [96] Agoston Pisztora. Surface order large deviations for ising, potts and percolation models. Probability Theory and Related Fields, 104:427–466, 1996.
  • PW [09] David Poulin and Pawel Wocjan. Preparing ground states of quantum many-body systems on a quantum computer. Phys. Rev. Lett., 102(13):130503, 2009.
  • RFA [24] Cambyse Rouzé, Daniel Stilck Franca, and Álvaro M. Alhambra. Efficient thermalization and universal quantum computing with quantum Gibbs samplers. arXiv preprint arXiv:2403.12691, 2024.
  • RW [24] Joel Rajakumar and James D Watson. Gibbs sampling gives quantum advantage at constant temperatures with o(1)o(1)-local hamiltonians. arXiv preprint arXiv:2408.01516, 2024.
  • RWW [23] Patrick Rall, Chunhao Wang, and Pawel Wocjan. Thermal state preparation via rounding promises. Quantum, 7:1132, 2023.
  • Sly [10] Allan Sly. Computational transition at the uniqueness threshold. In 2010 IEEE 51st Annual Symposium on Foundations of Computer Science, pages 287–296. IEEE, 2010.
  • Tem [17] Kristan Temme. Thermalization time bounds for pauli stabilizer hamiltonians. Communications in Mathematical Physics, 350(2):603–637, 2017.
  • TK [15] Kristan Temme and Michael J. Kastoryano. How fast do stabilizer hamiltonians thermalize? arXiv/1505.07811, 2015.
  • TKR+ [10] Kristan Temme, Michael James Kastoryano, Mary Beth Ruskai, Michael Marc Wolf, and Frank Verstraete. The χ2\chi^{2}-divergence and mixing times of quantum Markov processes. J. Math. Phys., 51(12), 2010.
  • VAGGdW [17] Joran Van Apeldoorn, András Gilyén, Sander Gribling, and Ronald de Wolf. Quantum SDP-solvers: Better upper and lower bounds. In FOCS 2017, pages 403–414. IEEE, 2017.
  • VWC [09] Frank Verstraete, Michael M. Wolf, and I. Cirac. Quantum computation and quantum-state engineering driven by dissipation. Nat. Phys., 5(9):633–636, 2009.
  • Wol [12] Michael M. Wolf. Quantum channels & operations: guided tour. Lecture Notes. URL http://www-m5. ma. tum. de/foswiki/pub M, 2012.
  • WT [23] Pawel Wocjan and Kristan Temme. Szegedy walk unitaries for quantum maps. Commun. Math. Phys., 402(3):3201–3231, 2023.
  • YL [23] Chao Yin and Andrew Lucas. Polynomial-time classical sampling of high-temperature quantum gibbs states. arXiv preprint arXiv:2305.18514, 2023.
  • ZB [23] Yikang Zhang and Thomas Barthel. Criteria for Davies irreducibility of Markovian quantum dynamics. J. Phys. A: Math. Theor, 2023.

Appendix A Low temperature Gibbs state preparation for 1D ferromagnetic Ising model

This section studies the low-temperature thermal state preparation for a 1D ferromagnetic Ising chain with periodic boundary condition. Although the Hamiltonian of this model is classical (a diagonal matrix in the computational basis), the jump operators are quantum and involve a significant amount of non-diagonal elements. The analysis of this model also shows the generality of core techniques for the 2D toric code in Section 4.

We consider NN spins on a ring structure, modeled by the Hilbert space 2N\mathcal{H}\cong\mathbb{C}^{2^{N}}. The Hamiltonian is

HIsing=Jj=1Nσjzσj+1z,J>0,H^{\rm Ising}=-J\sum_{j=1}^{N}\sigma^{z}_{j}\sigma^{z}_{j+1}\,,\quad J>0\,, (A.1)

with σN+1z=σ1z\sigma_{N+1}^{z}=\sigma_{1}^{z} (see Fig. 7, top), where {Zj:=σjzσj+1z}j=1N\{\textbf{Z}_{j}:=\sigma^{z}_{j}\sigma^{z}_{j+1}\}_{j=1}^{N} are bond observables satisfying

j=1NZj=I.\prod_{j=1}^{N}\textbf{Z}_{j}=I\,. (A.2)

We propose the following Gibbs smaller for the above Ising model:

β=j=1N(σjx+σjy+σjz)+𝖷,\mathcal{L}_{\beta}=\sum_{j=1}^{N}\left(\mathcal{L}_{\sigma^{x}_{j}}+\mathcal{L}_{\sigma^{y}_{j}}+\mathcal{L}_{\sigma^{z}_{j}}\right)+\mathcal{L}_{\mathsf{X}}\,, (A.3)

which is the sum of local Lindbladians with Pauli couplings plus a global one with coupling operator 𝖷:=j=1Nσjx\mathsf{X}:=\prod_{j=1}^{N}\sigma_{j}^{x}. Here σjx/y/z\mathcal{L}_{\sigma_{j}^{x/y/z}} and 𝖷\mathcal{L}_{\mathsf{X}} are defined via (2.4). Their explicit formulas can be similarly derived as in (4.16), (4.18), and (4.19). The main result of this section is the following spectral gap lower bound of β\mathcal{L}_{\beta} in Eq. A.3.

Theorem 17.

For the Davies generator (A.3), we have

Gap(β)max{Θ(e4βJ),Θ(N3)}.\mathrm{Gap}(\mathcal{L}_{\beta})\geq\max\left\{\Theta(e^{-4\beta J}),\Theta(N^{-3})\right\}\,.
Refer to caption
Figure 7: 1D Ising model. Top: The 1D ferromagnetic Ising Hamiltonian HIsingH^{\rm Ising}. The dased dot and line denote the periodic boundary condition. Middle: The logic operator 𝖷=Πi=1Nσix\mathsf{X}=\Pi_{i=1}^{N}\sigma^{x}_{i}. Bottom: The logic operator 𝖹=σ1z\mathsf{Z}=\sigma^{z}_{1}.

The proof is similar to that of 1, based on the structures of the ground states of the Ising model and the associated observable algebra ()\mathcal{B}(\mathcal{H}). We know that HIsingH^{\rm Ising} has a two-fold degenerate ground state space spanned by |0N\ket{0^{N}} and |1N\ket{1^{N}}, and it is frustration-free, namely, any ground state |φ\ket{\varphi} of HIsingH^{\rm Ising} satisfies σjzσj+1z|φ=|φ\sigma^{z}_{j}\sigma^{z}_{j+1}\ket{\varphi}=\ket{\varphi} for each jj. This two two-fold degeneracy encodes a single logical qubit in the sense that with some abuse of notation, |0N\ket{0^{N}} and |1N\ket{1^{N}} can be identified as the logical qubit |0\ket{0} and as |1\ket{1}, respectively:

|0|0N,|1|1N.\ket{0}\simeq\ket{0^{N}}\,,\quad\ket{1}\simeq\ket{1^{N}}\,.

The excited states are given by acting Pauli string Xj1XjnX_{j_{1}}\cdots X_{j_{n}} on the ground states |0N\ket{0^{N}} and |1N\ket{1^{N}}. We define observables (see Fig. 7)

𝖷=i=1Nσix,𝖹=σ1z.\displaystyle\mathsf{X}=\prod_{i=1}^{N}\sigma^{x}_{i}\,,\quad\mathsf{Z}=\sigma^{z}_{1}\,.

The observable 𝖷\mathsf{X} flips the logical qubit 𝖷|0/1=|1/0\mathsf{X}\ket{0/1}=\ket{1/0}. The measurement outcome by 𝖹\mathsf{Z} indicates which state we are observing:

0|𝖹|0=1,1|𝖹|1=1.\bra{0}\mathsf{Z}\ket{0}=1\,,\quad\bra{1}\mathsf{Z}\ket{1}=-1\,.

We note that 𝖷\mathsf{X} and 𝖹\mathsf{Z} commute with local terms {σjzσj+1z}\{\sigma^{z}_{j}\sigma^{z}_{j+1}\}, and thus all the eigenspaces of HIsingH^{\rm Ising} are the invariant subspaces of 𝖷\mathsf{X} and 𝖹\mathsf{Z}.

Let 𝒬(2)\mathcal{Q}\cong\mathcal{B}(\mathbb{C}^{2}) be the algebra on the logical qubit generated by 𝖷\mathsf{X} and 𝖹\mathsf{Z}. Then we have 𝒬=Span{I,𝖷,𝖸,𝖹}\mathcal{Q}={\rm Span}\{I,\mathsf{X},\mathsf{Y},\mathsf{Z}\}, where 𝖸=i𝖹𝖷\mathsf{Y}=i\mathsf{Z}\mathsf{X}. To describe the excited states and observable algebra ()\mathcal{B}(\mathcal{H}), we first define an admissible bond on a ring of NN sites as a configuration containing an even number of 1-1:

|b=|b1,b2,,bN{+1,1}Nsuch that#{bi=1}2.\ket{b}=\ket{b_{1},b_{2},\ldots,b_{N}}\in\left\{+1,-1\right\}^{N}\ \text{such that}\quad\#\left\{b_{i}=-1\right\}\in 2\mathbb{Z}\,.

The space spanned by the admissible bond, denoted by +\mathcal{H}_{+}, is of dimension 2N12^{N-1}. An important observation is that a natural orthonormal tensor basis of 2N\mathbb{C}^{2^{N}} consisting of |ϵ1ϵN\ket{\epsilon_{1}\ldots\epsilon_{N}} with ϵi=0/1\epsilon_{i}=0/1 can be uniquely written as |ϵ1|b2+\ket{\epsilon_{1}}\ket{b}\in\mathbb{C}^{2}\otimes\mathcal{H}_{+}, where the first qubit |ϵ1\ket{\epsilon_{1}} is regarded as a logical qubit and bjb_{j} is determined by the bond observable Zj=σjzσj+1z\textbf{Z}_{j}=\sigma^{z}_{j}\sigma^{z}_{j+1}:

Zj|ϵ1ϵN=bj|ϵ1ϵN.\textbf{Z}_{j}\ket{\epsilon_{1}\ldots\epsilon_{N}}=b_{j}\ket{\epsilon_{1}\ldots\epsilon_{N}}\,.

We define the full bond algebra 𝒜bfull\mathcal{A}^{\rm full}_{b} by all the linear transformations on +\mathcal{H}_{+}. Then, as a consequence of the above decomposition of |ϵ1ϵN\ket{\epsilon_{1}\ldots\epsilon_{N}}, the observable algebra can be decomposed as follows [2, Lemma 4].

Lemma 18.

The algebra of observables on a ring can be decomposed into

()=𝒬𝒜bfull.\mathcal{B}(\mathcal{H})=\mathcal{Q}\otimes\mathcal{A}^{\rm full}_{b}\,.

In particular, we have the orthogonal decomposition in GNS inner product

()=(I𝒜bfull)(𝖷𝒜bfull)(𝖸𝒜bfull)(𝖹𝒜bfull).\mathcal{B}(\mathcal{H})=(I\otimes\mathcal{A}^{\rm full}_{b})\oplus(\mathsf{X}\otimes\mathcal{A}^{\rm full}_{b})\oplus(\mathsf{Y}\otimes\mathcal{A}^{\rm full}_{b})\oplus(\mathsf{Z}\otimes\mathcal{A}^{\rm full}_{b})\,. (A.4)

We are now ready to sketch the proof of 17. For simplicity, we use bj=±b_{j}=\pm for the bond configuration. Similarly to Lemma 5, the algebra 𝒜bfull\mathcal{A}_{b}^{\rm full} is generated by the observables {Zj}j=1N\{\textbf{Z}_{j}\}^{N}_{j=1} and {σjx}j=2N\{\sigma^{x}_{j}\}_{j=2}^{N}, which commute with 𝖷\mathsf{X}. Then, by a direct computation, we have 𝖷(I𝒜bfull)=0\mathcal{L}_{\mathsf{X}}\left(I\otimes\mathcal{A}^{\rm full}_{\rm b}\right)=0. Moreover, there holds

𝖷(𝖹/𝖸𝒜bfull)=2(𝖹/𝖸)𝒜bfull,(𝖹/𝖸),𝖷(𝖹/𝖸)σβ=2.\mathcal{L}_{\mathsf{X}}\left(\mathsf{Z}/\mathsf{Y}\otimes\mathcal{A}^{\rm full}_{\rm b}\right)=-2(\mathsf{Z}/\mathsf{Y})\otimes\mathcal{A}^{\rm full}_{\rm b},\quad-\langle(\mathsf{Z}/\mathsf{Y}),\mathcal{L}_{\mathsf{X}}(\mathsf{Z}/\mathsf{Y})\rangle_{\sigma_{\beta}}=2\,. (A.5)

Similarly, we derive 𝖹(I𝒜bfull)=0\mathcal{L}_{\mathsf{Z}}\left(I\otimes\mathcal{A}^{\rm full}_{\rm b}\right)=0, and

𝖹(𝖷/𝖸𝒜bfull)=2(𝖷/𝖸)𝒜bfull,(𝖷/𝖸),𝖹(𝖷/𝖸)σβ=2.\mathcal{L}_{\mathsf{Z}}\left(\mathsf{X}/\mathsf{Y}\otimes\mathcal{A}^{\rm full}_{\rm b}\right)=-2\left(\mathsf{X}/\mathsf{Y}\right)\otimes\mathcal{A}^{\rm full}_{\rm b}\,,\quad-\langle(\mathsf{X}/\mathsf{Y}),\mathcal{L}_{\mathsf{Z}}(\mathsf{X}/\mathsf{Y})\rangle_{\sigma_{\beta}}=2\,. (A.6)

Further, for the action of local Lindbladian σjx/σjy/σjz\mathcal{L}_{\sigma^{x}_{j}/\sigma^{y}_{j}/\sigma^{z}_{j}} on the decomposition (A.4), we have the following lemma, in analog with Lemma 6.

Lemma 19.

The Lindbladian with Pauli coupling σjx\mathcal{L}_{\sigma^{x}_{j}}, σjy\mathcal{L}_{\sigma^{y}_{j}}, and σjz\mathcal{L}_{\sigma^{z}_{j}} (j1j\geq 1) are block diagonal with respect to the decomposition (A.4):

σjx/σjy/σjz(I/𝖷/𝖸/𝖹𝒜bfull)I/𝖷/𝖸/𝖹𝒜bfull.\displaystyle\mathcal{L}_{\sigma^{x}_{j}/\sigma^{y}_{j}/\sigma^{z}_{j}}\left(I/\mathsf{X}/\mathsf{Y}/\mathsf{Z}\otimes\mathcal{A}_{\rm b}^{\rm full}\right)\subset I/\mathsf{X}/\mathsf{Y}/\mathsf{Z}\otimes\mathcal{A}_{\rm b}^{\rm full}\,.

In particular, it holds that for j2j\geq 2, σjx(𝒬𝒜bfull)=𝒬σjx(𝒜bfull)\mathcal{L}_{\sigma^{x}_{j}}(\mathcal{Q}\otimes\mathcal{A}_{\rm b}^{\rm full})=\mathcal{Q}\otimes\mathcal{L}_{\sigma^{x}_{j}}(\mathcal{A}_{\rm b}^{\rm full}), that is,

σjx(QA)=Qσjx(A)for Q𝒬 and A𝒜bfull.\mathcal{L}_{\sigma^{x}_{j}}(QA)=Q\mathcal{L}_{\sigma^{x}_{j}}(A)\quad\text{for $Q\in\mathcal{Q}$ and $A\in\mathcal{A}^{\rm full}_{\rm b}$}\,.

We now define the local Lindbladian:

~=j=2Nσjx\widetilde{\mathcal{L}}=\sum_{j=2}^{N}\mathcal{L}_{\sigma^{x}_{j}} (A.7)

which is primitive when restricted on (+)=𝒜bfull\mathcal{B}(\mathcal{H}_{+})=\mathcal{A}^{\rm full}_{\rm b} [2, Lemma 6]. Thanks to the properties (A.5)–(A.6) of 𝖷\mathcal{L}_{\mathsf{X}} and 𝖹\mathcal{L}_{\mathsf{Z}}, the Lindbladian ~+𝖷+𝖹\widetilde{\mathcal{L}}+\mathcal{L}_{\mathsf{X}}+\mathcal{L}_{\mathsf{Z}} (as a part of β\mathcal{L}_{\beta} in (A.3)) is primitive on ()\mathcal{B}(\mathcal{H}). Then, Lemma 3 (item 2) readily gives

Gap(β)Gap(~+𝖷+𝖹).\mathrm{Gap}\left(\mathcal{L}_{\beta}\right)\geq\mathrm{Gap}\left(\widetilde{\mathcal{L}}+\mathcal{L}_{\mathsf{X}}+\mathcal{L}_{\mathsf{Z}}\right)\,. (A.8)

Thus, it suffices to consider the spectral gap of the latter one. For this, we note from (A.5)–(A.6) that 𝖷/𝖹\mathcal{L}_{\mathsf{X}/\mathsf{Z}} is also block diagonal for the decomposition (A.4). Then, by Lemma 19, we only need to estimate the gap of ~+𝖷+𝖹\widetilde{\mathcal{L}}+\mathcal{L}_{\mathsf{X}}+\mathcal{L}_{\mathsf{Z}} on each invariant subspace I/𝖷/𝖸/𝖹𝒜bfullI/\mathsf{X}/\mathsf{Y}/\mathsf{Z}\otimes\mathcal{A}_{\rm b}^{\rm full}:

  • On I𝒜bfullI\otimes\mathcal{A}^{\rm full}_{\rm b}. Noting 𝖷|I𝒜bfull=𝖹|I𝒜bfull=0\mathcal{L}_{\mathsf{X}}|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}=\mathcal{L}_{\mathsf{Z}}|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}=0, we have

    Gap((~+𝖷+𝖹)|I𝒜bfull)=Gap(~|I𝒜bfull).\mathrm{Gap}\left(\left(\widetilde{\mathcal{L}}+\mathcal{L}_{\mathsf{X}}+\mathcal{L}_{\mathsf{Z}}\right)\Big{|}_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}\right)=\mathrm{Gap}\left(\widetilde{\mathcal{L}}|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}\right)\,.
  • On 𝖷/𝖸/𝖹𝒜bfull\mathsf{X}/\mathsf{Y}/\mathsf{Z}\otimes\mathcal{A}^{\rm full}_{\rm b}. Noting that any Davies generator is negative, we have

    (~+𝖷+𝖹)|𝖹𝒜bfull(~+𝖷)|𝖹𝒜bfull.-\left(\widetilde{\mathcal{L}}+\mathcal{L}_{\mathsf{X}}+\mathcal{L}_{\mathsf{Z}}\right)\Big{|}_{\mathsf{Z}\otimes\mathcal{A}^{\rm full}_{\rm b}}\succeq-\left(\widetilde{\mathcal{L}}+\mathcal{L}_{\mathsf{X}}\right)\Big{|}_{\mathsf{Z}\otimes\mathcal{A}^{\rm full}_{\rm b}}\,.

    Moreover, ker(~|𝖹𝒜bfull)=Span(𝖹I)\ker\big{(}\widetilde{\mathcal{L}}|_{\mathsf{Z}\otimes\mathcal{A}^{\rm full}_{\rm b}}\big{)}={\rm Span}(\mathsf{Z}\otimes I). It can be seen as follows. For any 𝖹A𝖹𝒜bfull\mathsf{Z}A\in\mathsf{Z}\otimes\mathcal{A}^{\rm full}_{\rm b} in the kernel, by the second part of Lemma 19, ~(𝖹A)=𝖹~(A)=0\widetilde{\mathcal{L}}(\mathsf{Z}A)=\mathsf{Z}\widetilde{\mathcal{L}}(A)=0. Thus, Aker(~|𝒜bfull)A\in\ker\big{(}\widetilde{\mathcal{L}}|_{\mathcal{A}^{\rm full}_{\rm b}}\big{)}. Since ~|𝒜bfull\widetilde{\mathcal{L}}|_{\mathcal{A}^{\rm full}_{\rm b}} is primitive, it must be the case that A=IA=I. Then, by Lemma 3 (item 4), there holds

    (~+𝖷)|𝖹𝒜bfullGap(~|I𝒜bfull)𝖹,𝖷(𝖹)σβGap(~|I𝒜bfull)+𝖷=Θ(Gap(~|I𝒜bfull)Gap(~|I𝒜bfull)+1).-\left(\widetilde{\mathcal{L}}+\mathcal{L}_{\mathsf{X}}\right)\Big{|}_{\mathsf{Z}\otimes\mathcal{A}^{\rm full}_{\rm b}}\succeq\frac{-\mathrm{Gap}\left(\widetilde{\mathcal{L}}|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}\right)\left\langle\mathsf{Z},\mathcal{L}_{\mathsf{X}}(\mathsf{Z})\right\rangle_{\sigma_{\beta}}}{\mathrm{Gap}\left(\widetilde{\mathcal{L}}|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}\right)+\left\lVert\mathcal{L}_{\mathsf{X}}\right\rVert}=\Theta\left(\frac{\mathrm{Gap}\left(\widetilde{\mathcal{L}}|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}\right)}{\mathrm{Gap}\left(\widetilde{\mathcal{L}}|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}\right)+1}\right)\,.

    The same estimates hold for (~+𝖷)|𝖸𝒜bfull-\left(\widetilde{\mathcal{L}}+\mathcal{L}_{\mathsf{X}}\right)\Big{|}_{\mathsf{Y}\otimes\mathcal{A}^{\rm full}_{\rm b}} and (~+𝖹)|𝖷𝒜bfull-\left(\widetilde{\mathcal{L}}+\mathcal{L}_{\mathsf{Z}}\right)\Big{|}_{\mathsf{X}\otimes\mathcal{A}^{\rm full}_{\rm b}}.

Combining the above arguments with (A.8), we find

Gap(β)\displaystyle\mathrm{Gap}\left(\mathcal{L}_{\beta}\right) min{Gap((~+𝖷+𝖹)|I𝒜bfull),λmin((~+𝖷+𝖹)|𝖷/𝖸/𝖹𝒜bfull)}\displaystyle\geq\min\left\{\mathrm{Gap}\left(-\left(\widetilde{\mathcal{L}}+\mathcal{L}_{\mathsf{X}}+\mathcal{L}_{\mathsf{Z}}\right)\middle|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}\right),\lambda_{\min}\left(-\left(\widetilde{\mathcal{L}}+\mathcal{L}_{\mathsf{X}}+\mathcal{L}_{\mathsf{Z}}\right)|_{\mathsf{X}/\mathsf{Y}/\mathsf{Z}\otimes\mathcal{A}^{\rm full}_{\rm b}}\right)\right\}
=Θ(Gap(~|I𝒜bfull)),\displaystyle=\Theta\left(\mathrm{Gap}\left(\widetilde{\mathcal{L}}|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}\right)\right)\,,

as Gap(~|I𝒜bfull)0\mathrm{Gap}\left(\widetilde{\mathcal{L}}|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}\right)\to 0. It has been proved in [2, Proposition 1] that for any NN,

Gap(~|I𝒜bfull)e4βJe4βJ+1.\mathrm{Gap}\left(\widetilde{\mathcal{L}}|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}\right)\geq\frac{e^{-4\beta J}}{e^{-4\beta J}+1}\,.

Therefore, to prove 17, it suffices to focus on the gap of ~\widetilde{\mathcal{L}} on 𝒜bfull\mathcal{A}^{\rm full}_{\rm b} and show

Gap(~|I𝒜bfull)=Ω(N3),whenβΩ(logN),\mathrm{Gap}\left(\widetilde{\mathcal{L}}|_{I\otimes\mathcal{A}^{\rm full}_{\rm b}}\right)=\Omega\left(N^{-3}\right)\,,\quad\text{when}\ \beta\geq\Omega(\log N)\,, (A.9)

which can be done as Lemma 13. We have completed the proof of 17.