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

Commuting Local Hamiltonian Problem on 2D beyond qubits

Sandy Irani [email protected] Computer Science Department, University of California, Irvine, CA, USA Jiaqing Jiang [email protected] Computing and Mathematical Sciences, California Institute of Technology, Pasadena, CA, USA
Abstract

We study the complexity of local Hamiltonians in which the terms pairwise commute. Commuting local Hamiltonians (CLHs) provide a way to study the role of non-commutativity in the complexity of quantum systems and touch on many fundamental aspects of quantum computing and many-body systems, such as the quantum PCP conjecture and the area law. Despite intense research activity since Bravyi and Vyalyi introduced the CLH problem two decades ago [BV03], its complexity remains largely unresolved; it is only known to lie in NP for a few special cases. Much of the recent research has focused on the physically motivated 2D case, where particles are located on vertices of a 2D grid and each term acts non-trivially only on the particles on a single square (or plaquette) in the lattice. In particular, Schuch [Sch11] showed that the CLH problem on 2D with qubits is in NP. Aharonov, Kenneth, and Vigdorovich [AKV18] then gave a constructive version of this result, showing an explicit algorithm to construct a ground state. Resolving the complexity of the 2D CLH problem with higher dimensional particles has been elusive. We prove two results for the CLH problem in 2D:

  • We give a non-constructive proof that the CLH problem in 2D with qutrits is in NP. As far as we know, this is the first result for the commuting local Hamiltonian problem on 2D beyond qubits. Our key lemma works for general qudits and might give new insights for tackling the general case.

  • We consider the factorized case, also studied in [BV03], where each term is a tensor product of single-particle Hermitian operators. We show that a factorized CLH in 2D, even on particles of arbitrary finite dimension, is equivalent to a direct sum of qubit stabilizer Hamiltonians. This implies that the factorized 2D CLH problem is in NP. This class of CLHs contains the Toric code as an example.

1 Introduction

Understanding the properties of ground states of local Hamiltonians is a central problem in condensed matter physics. Kitaev [Kit03a] famously formulated this problem as a decision problem, amenable to analysis from the perspective of computational complexity, by defining the local Hamiltonian problem (LHP), which asks whether the ground energy of a local Hamiltonian is below one threshold or greater than another. The LHP can be interpreted as the quantum generalization of the Boolean Satisfiability problem (SAT), where for SAT, all the terms {hi}i\{h_{i}\}_{i} are diagonal in the computational basis. Kitaev showed that the LHP is QMA-complete [KKR06], a quantum analog of the Cook–Levin theorem showing that SAT is NP-complete. Formally, a kk-local Hamiltonian H=ihiH=\sum_{i}h_{i} is a Hermitian operator on nn qudits, where each term hih_{i} only acts on kk qudits. Given two parameters a,ba,b with ba1poly(n)b-a\geq\frac{1}{poly(n)}, the LHP is to determine whether the ground energy of HH, namely the minimum eigenvalue, is smaller than aa or greater than bb.

It is widely believed that QMANP\mbox{\bf QMA}\neq\mbox{\bf NP}, which would imply that the LHP is strictly harder than the SAT. A natural question then to ask is what properties make quantum SAT (LHP) harder than classical SAT? Alternatively, what additional constraints can make LHP easier than QMA-complete? An intermediate model that sits in between classical and quantum Hamiltonians is the Commuting Local Hamiltonian problem (CLHP), in which the terms of the local Hamiltonian pairwise commute. Commuting local Hamiltonians form an important sub-class of Hamiltonians of physical interest and were first studied from the complexity perspective by Bravyi and Vyalyi  [BV03]. Compared to the general LHP, the idea that CLHs should be more classical in nature stems from the intuition [FS97] in quantum physics, which suggests that it is the non-commutativity that makes the quantum world different from classical. Moreover, since all the terms of a CLH can be simultaneously diagonalized by a single basis, every eigenstate of the full Hamiltonian can be specified up to degeneracies by the corresponding eigenvalue for each term. The fact that the eigenstates have a classical specification suggests that the structure of the eigenbasis is more classical in nature. However, this fact is not enough to prove CLHP is in NP, since the eigenstates themselves can potentially be highly entangled, as is true for the famous example of Kitaev’s toric code [Kit03b]. Thus, a quantum witness for CLHP may still be beyond a classical description.

Commuting Hamiltonians provide a lens to study many fundamental aspects of quantum computing and many-body systems. For example, the stabilizer framework [Got97] is the basis for most error-correcting codes, and stabilizer codes can be seen as the ground states of commuting Hamiltonians. Commuting Hamiltonians are commonly used as a test ground for attacking difficult problems such as the quantum-PCP conjecture [AE11a, Has12b, Has13], NLTS conjecture [ABN22], Gibbs states preparation [KB16] and fast thermalization [BCG+23, CRF20]. In particular, commuting Hamiltonians could potentially provide insight into the area law and its connection to the efficient expressibility of ground states. The Area Law states that for ground states of gapped Hamiltonians on a finite-dimensional lattice, the entanglement entropy between two regions of ground states scales with the boundary between the regions as opposed to their volume. The area law is known to hold for 1D Hamiltonians and is known for 2D only under the assumption that the Hamiltonian is uniformly gapped [AAG22]. It is widely hoped that proving the area law will lead to insight into whether such ground states can be efficiently expressed or constructed by quantum circuits. As a subclass of LHP, it is well known that the area law holds for CLHs, and yet, even in 2D, we do not know whether the ground states of CLHs can be efficiently represented or constructed.

1.1 Previous Results

Despite two decades of study, the complexity of CLHP still remains open, with a few special cases known to be in NP ([BV03, AE13, AE11b, Sch11, AKV18, Has12a]). Bravyi and Vyalyi initiated this line of work, showing that qudit 22-local CLHP is in NP  [BV03]. Their proof uses a decomposition lemma based on the theory of finite-dimensional CC^{*}-algebra representations. All subsequent work on the CLHP has made use of this framework.

Aharonov and Eldar [AE11b] extended the results to the 33-local case for qubits and qutrits. Specifically, they proved that 3-local qubit-CLHP is in NP. They also proved that 3-local qutrit CLHP is in NP on the Nearly Euclidean interaction graphs. All the above results are proved by showing that there is a trivial ground state, i.e. the ground state can be prepared by a constant depth quantum circuit. This constant depth circuit is the NP witness and the energy of this trivial ground state can be checked in classical polynomial time by a light-cone argument. However, for 4-local CLHP, even if the interaction graph is a 2D square lattice, there are systems like the Toric code which has no trivial ground states.

In this work, we mainly focus on 4-local CLHP on a 2D square lattice with qudits (abbreviated as qudit-CLHP-2D). Specifically, consider a 2D square lattice as in Fig. 1(a) with a qudit qq on each vertex and on each plaquette pp, there is a Hermitian term acting on the qudits on its four vertices. With some abuse of notations, we also use pp to denote the Hermitian term on the plaquette pp. The qudit-CLHP-2D is given an nn-qudit Hamiltonian H=ppH=\sum_{p}p where {p}p\{p\}_{p} are commuting, two parameters a,ba,b where ba1/poly(n)b-a\geq 1/poly(n), decide whether the minimum eigenvalue of HH is smaller than aa or greater than bb. There is an alternate 2D setting in which qudits are placed on the edges and there are Hermitian terms on “plaquettes” and “stars”. The two settings, i.e. qudits on vertices or qudits on edges, are equivalent when the underlying graph is a 2D square lattice. ( See Appendix C.)

q2q_{2}q3q_{3}q1q_{1}q4q_{4}pp
(a)
qqp1p_{1}^{\prime}p2p_{2}p1p_{1}p2p_{2}^{\prime}
(b)
q2q_{2}q3q_{3}q1q_{1}q4q_{4}pbp_{b}pap_{a}pcp_{c}
(c)
Figure 1: (a) Definition of the qudit-CLHP-2D. (b) The four terms which involve qq are p1,p2,p1,p2p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime}. (c) An example of 1D structure in Schuch’s [Sch11] proof. In the figure the symbol \circ represents the plaquette term acts non-trivially on the qudit. Specifically here among the 9 plaquettes in (c), only terms pa,pb,pcp_{a},p_{b},p_{c} acts non-trivially on at least one of q1,q2,q3,q4q_{1},q_{2},q_{3},q_{4}.

Along this line, Schuch [Sch11] first proved that the qubit-CLHP-2D is in NP. Schuch’s proof provides a witness showing that a low-energy state exists without giving an explicit description of the state. Indeed his proof leaves open the question of whether an explicit description of the state even exists. Aharonov, Kenneth, and Vigdorovich [AKV18] later gave a constructive proof, showing that after some transformation, the qubit-CLHP-2D is equivalent to the Toric code permitting boundaries. We say a proof is constructive if The NP prover shows that the ground energy is below aa by providing a circuit for preparing the ground state. A boundary roughly means a qubit qq where one of the four terms involving this qubit, i.e. p1,p2,p1,p2p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime} in Fig. 1(b), acts trivially on qq. A Hamiltonian is equivalent to the Toric code permitting boundaries if after choosing an appropriate basis for each qubit, terms {p}p\{p\}_{p} in the Hamiltonian act similarly as XX or ZZ on the non-boundary qubits. Both proofs  [Sch11, AKV18] for the qubit-CLHP-2D heavily depend on the restrictions to the qubits, since the induced algebra of a term pp on a qubit qq can only have a very limited structure. The limited structure for the qubits case does not hold for higher dimensional particles, not even for qutrits.

Hastings [Has12a] proved a subclass of the qudit-CLHP-2D is in NP with some restrictive assumptions. Roughly speaking, he grouped all qudits on the same vertical line as a supersite, then viewed the 2D lattice as a 1D line, which reduced the qudit-CLHP-2D back to the “qudit” 2-local case [BV03]. However, the main problem is that those supersites are not really qudit of constant dimension. In fact, the dimension of the supersite is sub-exponential in the number of qudits. Thus Hastings further assumes that several technical conditions hold, like certain operators in the proof can be efficiently represented by matrix product operators of bounded dimension.

Besides the qudit-CLHP-2D, Bravyi and Vyalyi [BV03] also addressed the special case of the commuting Hamiltonian problem where all the terms are factorized (CHP-factorized). That is H=ihiH=\sum_{i}h_{i}, and each term hih_{i} is a tensor product of single-qudit Hermitian operators. For example, the Toric code is an instance of CHP-factorized, since each term is X4X^{\otimes 4} or Z4Z^{\otimes 4}. In general, in CHP-factorized each hih_{i} is not necessarily local, and there are no constraints on the underlying interaction graph. Bravyi and Vyalyi give a non-constructive proof to show that the qubit-CHP-factorized is in NP. It is an open question whether their proof can be generalized to higher dimensional particles (qudit-CHP-factorized) or whether their proof for qubits can be made constructive.

All of the above results prove that subclasses of CLHP are in NP. On the other hand, Gosset, Mehta, and Vidick [GMV17] give a result that indicates CLHP might be harder than NP, or even as hard as the general LHP. They show that the ground space connectivity problem of commuting local Hamiltonian is QCMA-complete, which is as hard as the ground space connectivity problem for the general local Hamiltonian.

1.2 Main results and proof overview

1.2.1 Main results

In this work, we give two new results on the qudit-CLHP-2D. The family of Hamiltonians considered in both results contains the Toric code as a special case. Theorem 1 extends the results for the qubit-CLHP-2D [Sch11, AKV18] to qutrits.

Theorem 1

The qutrit-CLHP-2D is in NP.

As far as we know, Theorem 1 is the first result for CLHP on 2D lattice beyond qubits. As noted, the results for the qubit-CLHP-2D [Sch11, AKV18] heavily depend on the limited dimension of qubits — the induced algebras on qubits have a very limited structure — which does not hold for qutrits. Our key idea to circumvent this problem introduces a technique to decrease the dimension of qudits. Specifically, denote the qudit-CLHP-2D instance as H=ppH=\sum_{p}p and the Hilbert space of a qudit qq as q\mathcal{H}^{q}. Under certain conditions, we observe that it suffices to consider a new instance of qudit-CLHP by (1) Restricting q\mathcal{H}^{q} to a subspace of smaller dimension and (2) Constructing a new Hamiltonian by projecting all pp to the smaller subspace, and then rounding all non-1 eigenvalues of the projected terms to 0. Moreover, the new instance preserves the correct answer in that “no” instances are converted to “no” instances, and “yes” instances are converted to “yes” instances. Thus, we show that our “decrease dimension and rounding” method (Lemma 23) can be interpreted as a non-constructive self-reduction for the qudit-CLHP. Here self-reduction means we reduce the original qudit-CLHP to a new qudit-CLHP where some qudits have strictly smaller dimensions. We emphasize that the key lemma, Lemma 23, works for the qudit-CLHP rather than only for qutrits, even without 2D geometry. This lemma might be of independent interest and bring new insights to tackle the general CLHP. We will explain this in more detail in the proof overview.

We next consider the special case of qudit-CLHP-2D where all the terms are factorized (qudit-CLHP-2D-factorized). Our second result, namely Theorem 2, is a constructive proof showing that qudit-CLHP-2D-factorized is in NP. Although we do not give the details here, our proof could be considerably simplified to provide a non-constructive version based on the ideas of [BV03] and [Sch11]. Here we give a stronger constructive proof that characterizes the structure of the ground space.

Theorem 2 (Informal version of Theorem 46)

The Hamiltonian in the qudit-CLHP-2D-factorized is equivalent to a direct sum of qubit stabilizer Hamiltonian. In particular, a factorized 2D commuting local Hamiltonian always has a ground state which is equivalent to qubit stabilizer state. This implies qudit-CLHP-2D-factorized is in NP.

We first briefly explain terminologies in Theorem 2. We say a commuting Hamiltonian H=ihiH=\sum_{i}h_{i} on space :=qq\mathcal{H}_{*}:=\otimes_{q}\mathcal{H}^{q}_{*} is equivalent to a qubit stabilizer Hamiltonian, if (1) For each q\mathcal{H}^{q}_{*}, by choosing an appropriate basis, q\mathcal{H}^{q}_{*} is factorized as a tensor of Hilbert spaces of dimension 22. Thus each q\mathcal{H}^{q}_{*} can be interpreted as several qubits. We allow dim(q)=1dim(\mathcal{H}^{q}_{*})=1 which corresponds to 0 qubit. (2) Each term hih_{i} acts as a Pauli operator up to phases, with respect to the basis of those “qubits”. We say a subspace :=qq\mathcal{H}_{*}\subseteq\mathcal{H}:=\otimes_{q}\mathcal{H}_{q} is simple, if \mathcal{H}_{*} is a tensor product of subspaces of each qudit, i.e. =qq\mathcal{H}_{*}=\otimes_{q}\mathcal{H}^{q}_{*}. We say a commuting Hamiltonian HH on nn-qudit space =qq\mathcal{H}=\otimes_{q}\mathcal{H}^{q} is equivalent to a direct sum of qubit stabilizer Hamiltonian, if the \mathcal{H} is a direct sum of simple subspaces {}\{\mathcal{H}_{*}\}_{*}, such that i,hi\forall i,h_{i} keeps each \mathcal{H}_{*} invariant, and HH is equivalent to qubit stabilizer Hamiltonian on \mathcal{H}_{*}.

Although one might conjecture that factorized commuting Hamiltonians (CHP-factorized) are equivalent to a direct sum of stabilizer Hamiltonians even when not restricted to 2D, this is still an open question even for the qubit case [BV03]. To illustrate the difficulty, consider two factorized terms acting on two qudits qq and qq^{\prime}: h:=hqhqh:=h_{q}\otimes h_{q^{\prime}} and h^:=h^qh^q\hat{h}:=\hat{h}_{q}\otimes\hat{h}_{q^{\prime}}. If hqh^q0h_{q}\hat{h}_{q}\neq 0 and hqh^q0h_{q^{\prime}}\hat{h}_{q^{\prime}}\neq 0, then it must be the case that the factors in each individual qudit must commute or anti-commute which gives rise to a stabilizer-like structure. By contrast, if hqh^q=0h_{q}\hat{h}_{q}=0, then [h,h^]=0[h,\hat{h}]=0 for any choice of hq,h^qh_{q^{\prime}},\hat{h}_{q^{\prime}}. This means that hqh_{q^{\prime}} and h^q\hat{h}_{q^{\prime}} can have an arbitrary relationship to each other and the two commuting terms h,h^h,\hat{h} may look very different from stabilizers. This second possibility suggests, alternatively, that factorized CLHs in 2D could have a different topological order from stabilizer Hamiltonians. More precisely, one may conjecture that there is a factorized Hamiltonian, such that no ground state can be prepared by applying a constant depth quantum circuit to a stabilizer state. We show a negative answer to this conjecture, by proving that the qudit-CLHP-2D-factorized is equivalent to a direct sum of qubit stabilizer Hamiltonian.

1.2.2 Overview for Theorem 1.

We start by reducing the more general CLH problem to a slightly restricted case where the commuting terms are projections and the energy lower bound aa is equal to 0. We will argue that if the more restricted version is in NP then the more general CLHP is also in NP. Consider an instance of the more general problem with Hermitian terms {hi}i\{h_{i}\}_{i} and bounds aa and bb, where ba1/poly(n)b-a\geq 1/\mbox{poly}(n). The NP prover can provide a vector describing the energy eigenvalue λi\lambda_{i} for each individual term hih_{i} such that iλia\sum_{i}\lambda_{i}\leq a. Then the verifier can replace each term hih_{i} with h^i=IΠi,λi\hat{h}_{i}=I-\Pi_{i,\lambda_{i}}, where Πi,λi\Pi_{i,\lambda_{i}} is the projection onto the eigenspace of hih_{i} corresponding to eigenvalue λi\lambda_{i}. The new instance has a state |ψ\left|\psi\right\rangle where h^i|ψ=0\hat{h}_{i}\left|\psi\right\rangle=0 for all ii if and only if hi|ψ=λi|ψh_{i}\left|\psi\right\rangle=\lambda_{i}\left|\psi\right\rangle for all ii. Since the h^i\hat{h}_{i} are all commuting projectors, the eigenvalues of the Hamiltonian are non-negative integers. Thus, the verifier can set the new bb to be equal to 11, resulting in a promise gap of 11. Therefore, when describing the proof of Theorem 1, we shall assume the terms are projections and that the question is whether there is a |ψ\left|\psi\right\rangle where h^i|ψ=0\hat{h}_{i}\left|\psi\right\rangle=0 for all ii (i.e. a frustration-free ground state). Note that this reduction does not work for the factorized case because even if the hih_{i} is factorized, the resulting h^i\hat{h}_{i} is not necessarily factorized.

We start by introducing the framework of induced algebras. The basic ideas are sketched here and specified in more detail in Section 3. Denote ()\mathcal{L}(\mathcal{H}) as the set of all operators on a Hilbert space \mathcal{H}. A CC^{*}-algebra is any algebra 𝒜()\mathcal{A}\subseteq\mathcal{L}(\mathcal{H}) which is also closed under the \dagger operations and includes the identity. Consider a Hermitian term pp acting c\mathcal{H}\otimes\mathcal{H}^{c}. The operator pp can be decomposed as

p=i,jhij|ij|c.p=\sum_{i,j}h_{ij}^{\mathcal{H}}\otimes|i\rangle\langle j|^{\mathcal{H}^{c}}.

The induced algebra of pp on space \mathcal{H}, denoted as 𝒜p\mathcal{A}^{\mathcal{H}}_{p}, is the CC^{*}-algebra generated by {I}{hij}ij\{I_{\mathcal{H}}\}\cup\{h_{ij}^{\mathcal{H}}\}_{ij}. Note that the particular decomposition of pp is not critical other than the fact that the |ij||i\rangle\langle j| terms acting on c\mathcal{H}^{c} are linearly independent. The key technique introduced by [BV03] is the Structure Lemma which decouples two commuting terms in their overlapping space: consider two terms p,p^p,\hat{p} sharing only one qudit qq. If [p,p^]=0[p,\hat{p}]=0, then the induced algebras, 𝒜pq\mathcal{A}^{\mathcal{H}^{q}}_{p} and 𝒜p^q\mathcal{A}^{\mathcal{H}^{q}}_{\hat{p}}, must commute, meaning that every operator in 𝒜pq\mathcal{A}^{\mathcal{H}^{q}}_{p} commutes with every operator in 𝒜p^q\mathcal{A}^{\mathcal{H}^{q}}_{\hat{p}}. Furthermore, p,pp,p^{\prime} can be decoupled in q\mathcal{H}^{q}, in that is there exists a direct sum decomposition q=iiq\mathcal{H}^{q}=\bigoplus_{i}\mathcal{H}^{q}_{i}, where

  • each iq\mathcal{H}^{q}_{i} has a factorized structure: iq=iq1iq2\mathcal{H}^{q}_{i}=\mathcal{H}^{q1}_{i}\otimes\mathcal{H}^{q2}_{i}

  • and pp only acts non-trivially on iq1\mathcal{H}^{q1}_{i} and p^\hat{p} only acts non-trivially on iq2\mathcal{H}^{q2}_{i}.

Both proofs showing that the qubit-CLHP-2D is in NP [Sch11, AKV18] depend heavily on the properties of qubits. In particular, if qq is a qubit, then there are only two ways to have a direct sum decomposition of q\mathcal{H}^{q}, namely as the direct sum of two 11-dimensional spaces or as a single 22-dimensional space. Note that we must also consider the case in which an induced algebra is trivial, meaning that 𝒜pq={cI}c\mathcal{A}^{\mathcal{H}^{q}}_{p}=\{cI_{\mathcal{H}}\}_{c} which implies that the operator pp acts trivially on qubit qq. Using the Structure Lemma the following statement is true:

Fact 3

Any two commuting non-trivial induced algebras on a qubit must be diagonalizable in the same basis.

Note that Fact 3 is not true for qutrits. One may understand this statement intuitively by only basic linear algebra. In particular, if h1,h2,hh_{1},h_{2},h are Hermitian operators on a qubit, where hh is not proportional to the identity and both h1h_{1} and h2h_{2} commute with hh, then all three operators can be diagonalized in the same basis. This observation is not true for qutrits, as here exist operators h,h1,h2h,h_{1},h_{2} on a qutrit with hh nontrivial, such that hh commutes with h1h_{1} and h2h_{2}, but the three operators cannot be diagonalized simultaneously.

Since the structure of our proof follows the same outline as Shuch’s proof, we briefly explain how [Sch11] used Fact 3 to prove qubit-CLHP-2D is in NP. Consider an arbitrary qudit-CLHP-2D instance H=ppH=\sum_{p}p. As argued above, we can assume {p}p\{p\}_{p} are commuting projections, and the goal is to determine whether λ(H)=0\lambda(H)=0 or λ(H)1\lambda(H)\geq 1. Note that in this setting, proving λ(H)=0\lambda(H)=0 is equivalent to proving tr(p(Ip))>0tr(\prod_{p}(I-p))>0.

Now consider a qubit qq in a 2D lattice (as shown in Fig. 1(b)). We name the terms acting on qq as p1,p2,p1,p2p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime}. We define the set of removable qubits (defined implicitly in [Sch11]) as follows:

  • qRq\in R: If the induced algebras of p1,p2p_{1},p_{2} on qq can be diagonalized in the same basis; or this condition holds for p1,p2p_{1}^{\prime},p_{2}^{\prime}.

The set RR is called removable since they can be effectively traced out. Specifically, suppose qRq\in R and assume p1,p2p_{1},p_{2} can be diagonalized in basis {|ϕ1,|ϕ2}\{\left|\phi_{1}\right\rangle,\left|\phi_{2}\right\rangle\}. Then

tr[p(Ip)]=i=1,2tr[|ϕiϕi|(1p1)|ϕiϕi|(1p2)|ϕiϕi|pp1,p2(1p)]tr\left[\prod_{p}(I-p)\right]=\sum_{i=1,2}tr\left[|\phi_{i}\rangle\!\langle\phi_{i}|(1-p_{1})|\phi_{i}\rangle\!\langle\phi_{i}|(1-p_{2})|\phi_{i}\rangle\!\langle\phi_{i}|\prod_{p\neq p_{1},p_{2}}(1-p)\right]

Each quantity in the sum on the right is the trace of a product of two positive semi-definite Hermitian operators, and therefore, each quantity in the sum is non-negative. Thus, tr(p(Ip))>0tr(\prod_{p}(I-p))>0 if and only if one of the two quantities in the sum is positive. The prover will then provide a |ϕ1\left|\phi_{1}\right\rangle or |ϕ2\left|\phi_{2}\right\rangle for qubit qq for which the trace is positive. Note that the proof is non-constructive because the ground state may not lie entirely within either the space spanned by the space spanned by |ϕ1\left|\phi_{1}\right\rangle or the space spanned by |ϕ2\left|\phi_{2}\right\rangle. Suppose that the witness for qubit qq is |ϕ1\left|\phi_{1}\right\rangle. The verifier must verify that

tr[|ϕ1ϕ1|(1p1)|ϕ1ϕ1|(1p2)|ϕ1ϕ1|pp1,p2(1p)]>0.tr\left[|\phi_{1}\rangle\!\langle\phi_{1}|(1-p_{1})|\phi_{1}\rangle\!\langle\phi_{1}|(1-p_{2})|\phi_{1}\rangle\!\langle\phi_{1}|\prod_{p\neq p_{1},p_{2}}(1-p)\right]>0.

The other two terms that might act non-trivially on qubit qq are p1p_{1}^{\prime} and p2p_{2}^{\prime}. By Fact 3 we know either one of the induced algebras of p1,p2p_{1}^{\prime},p_{2}^{\prime} on qq acts trivially on qq, or they can be diagonalized in the same basis. By considering each case separately, it can be argued that the qubit qq can be traced out of all the terms. This process can be applied simultaneously for all the removable qubits. The remaining Hamiltonian will only contain terms that operate non-trivially on qubits that are not removable.

Using Fact 3, we know that if qq is not removable (qRq\not\in R), then one of p1,p2p_{1},p_{2} and one of p1,p2p_{1}^{\prime},p_{2}^{\prime} act trivially on qq. Now consider a graph where each vertex represents a plaquette term pp and two terms are connected if they operate non-trivially on a common qubit. Schuch argues that if none of the qubits are removable then this graph cannot have a vertex with a degree larger than two, namely the graph is a set of disjoint chains and cycles. The trace of each chain or cycle can be computed in classical polynomial time by representing each term as a tensor and contracting the tensors along the chain.

When moving to qutrits, we need to address the fact that the Structural Lemma allows for more complex decompositions of the Hilbert space of a qutrit, and in particular, Fact 3 no longer holds. Our key observation to tackle the problem, is to introduce a new way to decrease the dimension of the particle — Decrease Dimension and Rounding (Lemma 23, Sec. 4.1), which can be applied to qudits that have a property which we call semi-separable.

We first describe the stronger condition of a separable qudit, which was introduced in [AE11b]. Consider a CLHP H=ihiH=\sum_{i}h_{i}. A qudit is separable if there exists a non-trivial decomposition q=jjq\mathcal{H}^{q}=\bigoplus_{j}\mathcal{H}_{j}^{q} such that all the terms hih_{i} keep all subspaces jq\mathcal{H}^{q}_{j} invariant. Note that if there is a separable qudit and a solution (i.e. a frustration-free ground state) exists, then a ground state must lie entirely within one of the jq\mathcal{H}_{j}^{q}. Thus a prover can provide the projector Πjq\Pi_{j}^{q} onto the subspace of qudit qq which contains the solution. The verifier can replace each term hih_{i} with ΠjqhiΠjq\Pi_{j}^{q}\cdot h_{i}\cdot\Pi_{j}^{q} and the dimension of the problem has been reduced.

We now extend this notion and define a qudit to be semi-separable if we allow at most one term to not keep the decomposition invariant. Our key observation is that, for CLHP, even for a semi-separable qudit qq, an NP prover can similarly decrease the dimension of the qudit qq, in a non-constructive way. Specifically, the NP prover will choose a subspace jq\mathcal{H}_{j}^{q} and restrict all the terms in this subspace. Since there is one term (call it h0h_{0}) that is inconsistent with the decomposition, such restriction can not be done naturally. Instead, we project h0h_{0} on the subspace jq\mathcal{H}_{j}^{q} and then round all the not-1-eigenvalue to 0. By doing this we claim that we again get a new CLHP instance with a smaller dimension in qudit qq. More importantly, we prove that the original CLHP has a frustration-free ground state iff there exists a jj such that the new CLHP H|jH|_{j} also has a frustration-free ground state. The reduction is non-constructive because the ground state in the new instance may not be the same as the ground state in the original instance. The “semi-separable” technique is powerful especially for CLHP-2D where H=ppH=\sum_{p}p, since on 2D lattice as in Fig. 1(b), for any qudit qq, if we consider the decomposition of q\mathcal{H}^{q} induced by the induced algebra of p1p_{1} on qq, there are at most 2 terms, i.e. p1,p2p_{1}^{\prime},p_{2}^{\prime}, which do not keep the decomposition invariant. This observation is also true for CLHP embedded on a planar graph.

By the above argument, to prove the qutrit-CLHP-2D is in NP, w.l.o.g we can assume that there are no semi-separable qutrits. We further prove that the condition — the qutrit-CLHP-2D without semi-separable qutrits — leads to strong restrictions on the form of the Hamiltonian. In particular, we show that for the qutrit-CLHP-2D without semi-separable qutrits, if we consider again the graph of plaquette terms where two terms are connected by an edge if they act non-trivially on a common qutrit, then this graph must also consist of disjoint chains or cycles. The trace of the 0-energy space can be computed as before in classical polynomial time by contracting 1D chains of tensors. The NP witness will be the indexes of subspaces chosen when removing all semi-separable qudits, and the subspaces for the removable qutrits.

1.2.3 Overview of Theorem 2

Recall that Theorem 2 is a constructive proof for the qubit-CLHP-2D-factorized, where factorized means each term is a tensor product of single-qudit Hermitian operators. As we mentioned in the main results section, finding a constructive proof that CHP-factorized is in NP is made difficult because if the product of two terms hh and h^\hat{h} is equal to 0, then their terms on individual qudits can have an arbitrary relationship. Namely, it is possible that hqh^q±h^qhqh^{q}\hat{h}^{q}\neq\pm\hat{h}^{q}h^{q}. On the other hand, [BV03] showed that if all the terms are commuting obey the condition that hqh^q=±h^qhqh^{q}\hat{h}^{q}=\pm\hat{h}^{q}h^{q} for each qudit, then the Hamiltonian will be related to a qubit stabilizer Hamiltonian.

The key part to proving the Theorem 2, is to remove the possibility that hqh^q±h^qhqh^{q}\hat{h}^{q}\neq\pm\hat{h}^{q}h^{q} for some qq, without changing the ground space, which will imply a constructive proof by showing a correspondence with stabilizer Hamiltonians. In general, this removal is hard to achieve. Even for the qubit-CHP-factorized, it is still an open question whether there exists a constructive proof. However surprisingly, we can give a constructive proof for qudit-CHP-factorized, when the underlying interaction graph is 2D, i.e. qudit-CLHP-2D-factorized. Specifically, by using proof of contradiction, firstly we prove that qudit-CLHP-2D-factorized, if there are no separable qudits, then all terms must commute in a regular way. With a slight clarification, we show that the qudit-CLHP-2D-factorized without separable qudits is equivalent to qubit stabilizer Hamiltonian. For the more general case where there are separable qudits, we then notice that there exists a partition of the nn-qudit space into simple subspaces, such that the restricted Hamiltonian on each subspace has no separable qudits. This partition is achieved by the following: when there is a separable qudit qq w.r.t decomposition q=jjq\mathcal{H}^{q}=\oplus_{j}\mathcal{H}^{q}_{j}, we partition the whole space according to this decomposition. We recursively perform this partition until for each subspace, the restricted Hamiltonian has no separable qudits.

1.3 Structure of the manuscript

The manuscript is structured as follows. In Sec. 2 we give notations and definitions which are used throughout this manuscript. In Sec. 3 we review the necessary definitions and techniques of CC^{*}-algebra and the Structure Lemma required for proving that the qutrit-CLHP-2D is in NP. In Sec. 4 and Sec. 5, we give proofs for the qutrit-CLHP-2D and the qudit-CLHP-2D-factorized respectively.

The manuscript is written in a way that several proofs can be read separately. In summary, for readers interested in Lemma 23 (The Decrease Dimension and Rounding Lemma), the suggested order is Sec. 2 and Sec. 4.1. For readers interested in the qutrit-CLHP-2D, the suggested order is Sec 2, Sec. 3, Sec. 4. For readers interested in the qudit-CLHP-2D-factorized, the suggested order is Sec 2, Sec. 5, and then Sec. 3 if necessary.

1.4 Conclusion and future work

In this manuscript, we give two new results of the qudit-CLHP-2D. First, we proved that qutrit-CLHP-2D is in NP, by introducing a non-constructive way of self-reduction for the qudit-CLHP, when there are semi-separable qudits. This self-reduction (proven in Lemma 23) works for qudit and might be of independent interest. Second, we prove that qudit-CLHP-2D-factorized is in NP, by showing that the Hamiltonian is equivalent to a direct sum of qubit stabilizer Hamiltonian.

One direct question is whether our proof for qutrit-CLHP-2D can be made to be constructive. That is whether one can prepare the ground state by polynomial-size quantum circuits. Aharonov, Kenneth and Vigdorovich [AKV18] proved that the qubit-CLHP-2D is equivalent to the Toric code permitting boundary. It is natural to ask whether qutrit-CLHP-2D, or general qudit-CLHP-2D, can have different ground space properties from the stabilizer Hamiltonians. Another question is whether our constructive proof for the qudit-CLHP-2D-factorized can be modified to prepare the ground states of the qubit-CHP (without 2D geometry). Recall that the qubit-CHP is in NP by a non-constructive method [BV03].

A further question is to extend the frontier of the complexity of CLHP. In particular, we conjecture that Lemma 23 can be used in more general settings. Recall that Lemma 23 works for qudit rather than only for qutrits, and it implies w.l.o.g we can assume that there are no semi-separable qudits. It is interesting to see whether one can prove that the qudit-CLHP-2DNP\in\mbox{\bf NP} by combining Lemma 23 and other methods like [AKV18]. Another promising setting is considering 3-local qutrit-CLHP, without any geometry constraints. Recall that [AE11b] proved for 3-local qubit-CLHP is in NP, by showing that after removing all separable qubits, the resulting Hamiltonian can be viewed as a 2-local qudit-CLHP. It is possible that for 3-local qutrit-CLHP, if we remove all semi-separable qutrits, the Hamiltonian is again of a 2-local structure, which will imply 3-local qutrit-CLHP is in NP.

Most known results are trying to show that CLHP is in NP. On the other side, it will be very interesting to provide any evidence that CLHP might be harder than NP.

2 Preliminaries

2.1 Notation

Given a Hermitian operator HH, we use λ(H)\lambda(H) to denote its ground energy, i.e. minimum eigenvalue. For two operators, hh and hh^{\prime}, we use [h,h][h,h^{\prime}] to denote its commutator hhhhhh^{\prime}-h^{\prime}h. In particular, [h,h]=0[h,h^{\prime}]=0 means h,hh,h^{\prime} are commuting. Two sets of operators, SS and S^\hat{S}, commute if [h,h^]=0,hS,h^S^.[h,\hat{h}]=0,\forall h\in S,\hat{h}\in\hat{S}. For a set of Hermitian operators {hi}i\{h_{i}\}_{i}, we use ker{hi}i\ker\{h_{i}\}_{i} to denote its common 0-eigenspace, i.e. ker{hi}i:={|ψ|hi|ψ=0,i}\ker\{h_{i}\}_{i}:=\{\left|\psi\right\rangle\,|\,h_{i}\left|\psi\right\rangle=0,\forall i\}. We say ker{hi}i\ker\{h_{i}\}_{i} is non-trivial iff ker{hi}i{0}.\ker\{h_{i}\}_{i}\neq\{0\}.111This 0 means the zero vector. With some abuse of notations, we use 0 both for real number zero, and zero vector. For ease of illustration, we also denote a Hermitian H=ihiH=\sum_{i}h_{i} as a set {hi}i\{h_{i}\}_{i}. We say a Hermitian operator Π\Pi is a projection if Π2=Π\Pi^{2}=\Pi. When {hi}i\{h_{i}\}_{i} are commuting projections, we have λ(H)=0\lambda(H)=0 iff ker{hi}i\ker{\{h_{i}\}_{i}} is non-trivial.

Let \mathcal{H} be a finite-dimensional Hilbert space. We use ()\mathcal{L}(\mathcal{H}) to denote the set of linear operators on \mathcal{H}. For Hermitian hh, we use h0h\succeq 0 to denote hh is positive semidefinite, that is all of its eigenvalues are non-negative. We use II to denote the identity matrix. Let hh be a Hermitian operator on Hilbert space 𝒳=𝒵\mathcal{X}=\mathcal{H}\otimes\mathcal{Z}, we say hh keeps the decomposition =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} invariant if hh keeps the subspace i𝒵\mathcal{H}_{i}\otimes\mathcal{Z} invariant, i\forall i. We say the decomposition =i=1mi\mathcal{H}=\bigoplus_{i=1}^{m}\mathcal{H}_{i} is non-trivial if m2m\geq 2.

In the following, we use qq to denote a qudit, and q\mathcal{H}^{q} to denote the Hilbert space of the qudit qq. Consider an operator hh acting on nn qudits. We say that hh acts trivially on a qudit qq if hh acts as identity on q\mathcal{H}^{q}. When hh acts non-trivially on only kk of the nn qudits, we will interchangeably view hh as an operator on kk qudits or a global operator on nn qudits. The meaning will be clear in the context. We use trS()tr_{S}() for tracing out the qudits in SS, and use tr()tr() for tracing out all the qudits. We use ScS^{c} to denote the set of qudits outside SS.

2.2 Formal Problem Definitions

Commuting kk-Local Hamiltonian We say a Hermitian operator HH on nn qudits is a commuting kk-local Hamiltonian, if H=i=1mhiH=\sum_{i=1}^{m}h_{i} for m=poly(n)m=poly(n), where each hih_{i} only acts non-trivially on kk qudits, and hihj=hjhi,i,jh_{i}h_{j}=h_{j}h_{i},\forall i,j. We allow different qudits to have different dimensions. In particular, for qutrit kk-local commuting local Hamiltonian, we allow the dimension of each qudit to be either 1,21,2 or 33.

2D and Factorized Variants Consider a 2D square lattice as in Fig. 1(a), on each vertex there is a qudit qq, and on each plaquette pp there is a Hermitian term acting on the qudits on its four vertices. With some abuse of notations, we also use pp to denote the Hermitian term on the plaquette pp. We say a commuting (4-local) Hamiltonian is on 2D if there is an underlying 2D square lattice and plaquettee terms defined as above such that H=ppH=\sum_{p}p and all {p}p\{p\}_{p} are pairwise commuting.

We further say a commuting (4-local) Hamiltonian on 2D is factorized, if each pp is factorized on its vertices, that is p=pq1pq2pq3pq4p=p^{q_{1}}\otimes p^{q_{2}}\otimes p^{q_{3}}\otimes p^{q_{4}} for Hermitian terms pqip^{q_{i}} acting on qudit qiq_{i}, as shown in Fig. 1(a). We call pqip^{q_{i}} factors. For the Toric code, p{X4,Z4}p\in\{X^{\otimes 4},Z^{\otimes 4}\}.

Commuting kk-Local Hamiltonian problem Given a family of commuting kk-local Hamiltonian H=ihiH=\sum_{i}h_{i} on nn qudits and parameters a,ba,b\in\mathbb{R} with ba1/poly(n)b-a\geq 1/poly(n). The commuting kk-local Hamiltonian problem w.r.t (H,a,b)(H,a,b) is a promise problem that decides whether λ(H)a\lambda(H)\leq a or λ(H)b\lambda(H)\geq b. We denote this problem as qudit-CLHP(H,a,b)(H,a,b), abbreviated as qudit-CLHP when H,a,bH,a,b are clear in the context. Similarly, we abbreviate the 2D and 2D-factorized variants as qudit-CLHP-2D and qudit-CLHP-2D-factorized respectively.

We define a special case of the qudit-CLHP called qudit-CLHP-projection where each term hih_{i} is a projection, b=1b=1, and a=0a=0. Note that since {hi}i\{h_{i}\}_{i} are commuting projections, we know λ(H)\lambda(H) must be a non-negative integer. Thus in the No instance we use λ(H)1\lambda(H)\geq 1 rather than λ(H)1/poly(n)\lambda(H)\geq 1/poly(n). We define qudit-CLHP-2D-projection similarly.

2.3 More Definitions

Consider a commuting kk-local Hamiltonian H=ihiH=\sum_{i}h_{i} on nn qudits with Hermitian terms {hi}i\{h_{i}\}_{i}. Although hih_{i} acts non-trivially only on kk qudits, in this section we view it as an operator on nn qudits. We will name the nn qudits as q1,q2,,qnq_{1},q_{2},...,q_{n}. When we refer to an arbitrary qudit, we name it as qq.

Definition 4 (Separable qudit)

A qudit qq is separable w.r.t Hermitian terms {hi}i\{h_{i}\}_{i} if there exists a non-trivial decomposition of its Hilbert space q=j=1mjq\mathcal{H}^{q}=\bigoplus_{j=1}^{m}\mathcal{H}^{q}_{j} s.t all hih_{i} keep the decomposition invariant. Here non-trivial means m2m\geq 2. We use Πj\Pi_{j} to denote the projection onto jq\mathcal{H}^{q}_{j}.

The definition of separable is first introduced by [AE11b]. Roughly speaking, it says all the Hermitian terms {hi}i\{h_{i}\}_{i} are block-diagonalized in the same way. We introduce the notion of semi-separable qudit, which will play a key role in the proof of the qutrit-CLHP-2D.

Definition 5 (Semi-separable qudit)

A qudit qq is semi-separable w.r.t Hermitian terms {hi}i\{h_{i}\}_{i} if there exists a non-trivial decomposition of its Hilbert space q=j=1mjq\mathcal{H}^{q}=\bigoplus_{j=1}^{m}\mathcal{H}^{q}_{j} s.t all but one hih_{i} keeps the decomposition invariant. Here non-trivial means m2m\geq 2. We use Πj\Pi_{j} as the projection onto jq\mathcal{H}^{q}_{j}. By convention when referring to a specific qudit, we will denote the term which does not keep the decomposition invariant as h0h_{0}.

Semi-separable qudit is a relaxation of separable qudits, in the sense that we allow one term to be not block-diagonalized w.r.t the decomposition q=j=1mjq\mathcal{H}^{q}=\bigoplus_{j=1}^{m}\mathcal{H}^{q}_{j}. Note that by the definition of semi-separable, hih_{i} is Hermitian and we have [hi,Πj]=0,i0,j[h_{i},\Pi_{j}]=0,\forall i\neq 0,\forall j. We will repeatedly use this fact. It is also important to keep in mind that [h0,Πj][h_{0},\Pi_{j}] might not be equal to 0, since we allow h0h_{0} not keeping jq\mathcal{H}^{q}_{j} invariant.

3 Review of CC^{*}-algebras and the Structure Lemma

This section is a review of CC^{*}-algebra and the Structure Lemma [BV03], which is the key tool to analyze the structures in the commuting local Hamiltonians. A more detailed proof on those techniques can be seen in Sec. 7.3 of  [GHL+15]. The following notations and lemmas are rephrased from [AKV18] and Sec. 7.3 of  [GHL+15].

3.1 Basics of CC^{*}-algebras

Definition 6 (CC^{*}-algebra)

Let \mathcal{H} be a finite dimensional Hilbert space, a CC^{*}-algebra is any algebra 𝒜()\mathcal{A}\subseteq\mathcal{L}(\mathcal{H}) which is closed under the \dagger operations and includes the identity. We say that two CC^{*}-algebras, 𝒜\mathcal{A} and 𝒜\mathcal{A}^{\prime}, commute if [a,a]=0,a𝒜,a𝒜.[a,a^{\prime}]=0,\forall a\in\mathcal{A},a^{\prime}\in\mathcal{A}^{\prime}.

Definition 7 (Trivial operator and algebra)

Let \mathcal{H} be a finite-dimensional Hilbert space. We say an operator h()h\in\mathcal{L}(\mathcal{H}) acting trivially on \mathcal{H} if h=cIh=cI_{\mathcal{H}} for some constant cc. We say a CC^{*}-algebra on 𝒜()\mathcal{A}\subseteq\mathcal{L}(\mathcal{H}) is trivial if every operators in 𝒜\mathcal{A} is trivial, i.e. 𝒜={cI}c\mathcal{A}=\{cI_{\mathcal{H}}\}_{c}. If =12\mathcal{H}=\mathcal{H}_{1}\otimes\mathcal{H}_{2}, we say hh acts trivially on 1\mathcal{H}_{1} if h=cI1h2h=cI_{\mathcal{H}_{1}}\otimes h_{2} for h2(2)h_{2}\in\mathcal{L}(\mathcal{H}_{2}).

Definition 8 (Center of CC^{*}-algebra)

The center of a CC^{*}-algebra 𝒜\mathcal{A} is defined as the set of operators in 𝒜\mathcal{A} which commutes with 𝒜\mathcal{A}, that is

𝒵(𝒜):={a𝒜|[a,a]=0,a𝒜}.\displaystyle\mathcal{Z}(\mathcal{A}):=\{a\in\mathcal{A}|[a,a^{\prime}]=0,\forall a^{\prime}\in\mathcal{A}\}. (1)

Then we introduce the induced algebra, which connects a Hermitian operator and a CC^{*}-algebra.

Definition 9 (Induced algebra)

Let hh be a Hermitian operator acting on Hilbert space \mathcal{H}\otimes\mathcal{H}^{\prime}. Consider the decomposition

h=i,jhij|ij|\displaystyle h=\sum_{i,j}h_{ij}^{\mathcal{H}}\otimes|i\rangle\langle j|^{\mathcal{H}^{\prime}} (2)

where {|i}i\{\left|i\right\rangle\}_{i} is an orthogonal basis of \mathcal{H}^{\prime}. The induced algebra of hh on \mathcal{H}, denoted as 𝒜h\mathcal{A}_{h}^{\mathcal{H}}, is defined as the CC^{*}-algebra generated by {hij}ij{I}\{h_{ij}^{\mathcal{H}}\}_{ij}\cup\{I_{\mathcal{H}^{\prime}}\}. We abbreviate 𝒜h\mathcal{A}^{\mathcal{H}}_{h} as 𝒜h\mathcal{A}_{h} when \mathcal{H} is clear in the context. We abbreviate 𝒜hq\mathcal{A}^{\mathcal{H}_{q}}_{h} as 𝒜hq\mathcal{A}^{q}_{h} for qudit qq.

The induced algebra is independent of the chosen decomposition for Hermitian hh.

Lemma 10 (Claim B.3 of [AKV18])

In Definition 9 consider two decompositions

h=ijhijgij=ijh^ijg^ij.\displaystyle h=\sum_{ij}h_{ij}^{\mathcal{H}}\otimes g_{ij}^{\mathcal{H}^{\prime}}=\sum_{ij}\hat{h}_{ij}^{\mathcal{H}}\otimes\hat{g}_{ij}^{\mathcal{H}^{\prime}}. (3)

where the sets {gij}ij,{g^ij}ij\{g_{ij}^{\mathcal{H}^{\prime}}\}_{ij},\{\hat{g}_{ij}^{\mathcal{H}^{\prime}}\}_{ij} are linearly independent respectively. Then the CC^{*}-algebra generated by {hij}ij\{h_{ij}^{\mathcal{H}}\}_{ij} is the same as the one generated by {h^ij}ij\{\hat{h}_{ij}^{\mathcal{H}^{\prime}}\}_{ij}.

By Lemma 10 we know the induced algebra of hh on \mathcal{H}, i.e. 𝒜h\mathcal{A}^{\mathcal{H}}_{h} in Definition 9, is independent of the decomposition we choose, thus 𝒜h\mathcal{A}^{\mathcal{H}}_{h} is well-defined. Note that if there is a decomposition =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} such that 𝒜h\mathcal{A}_{h}^{\mathcal{H}} keeps i\mathcal{H}_{i} invariant, i\forall i, then it follows that hh keeps i\mathcal{H}_{i} invariant, i\forall i.

3.2 The Structure Lemma

The Structure Lemma [T+03] says that every finite-dimensional CC^{*}-algebra is a direct sum of algebras of all operators on a Hilbert space. See Sec. 7.3 of  [GHL+15] for an accessible proof of the Structure Lemma. The following statement is taken from [AKV18].

Lemma 11 (The Structure Lemma: classification of finite dimensional CC^{*}-algebras )

Let 𝒜L()\mathcal{A}\subseteq L(\mathcal{H}) be a CC^{*}-algebra where \mathcal{H} is finite dimensional. There exists a direct sum decomposition: =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} and a tensor product structure i=i1i2\mathcal{H}_{i}=\mathcal{H}_{i}^{1}\otimes\mathcal{H}_{i}^{2} such that

𝒜=i(i1)(i2)\mathcal{A}=\bigoplus_{i}\mathcal{L}(\mathcal{H}_{i}^{1})\otimes\mathcal{I}(\mathcal{H}_{i}^{2})

Furthermore, the center of 𝒜\mathcal{A} is spanned by {Πi}i\{\Pi_{i}\}_{i}, where Πi\Pi_{i} is the projection onto the subspace i\mathcal{H}_{i}.

Given a CC^{*}-algebra 𝒜\mathcal{A}, we denote the decomposition =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} in Lemma 11 as the decomposition induced by 𝒜\mathcal{A}. Note that here we do not argue whether the decomposition in Lemma 11 is unique or not. However, for clarity when we mention the decomposition induced by 𝒜\mathcal{A}, we always refer to the same canonical decomposition. For example, we can set the canonical decomposition to be the one obtained by the proof in Sec. 7.3 of  [GHL+15]. In the following we give some definitions of decompositions, and a further remark on Lemma 11.

Definition 12 (Trivial, Better decomposition)

Consider the decomposition of a finite-dimensional space =i=1mi\mathcal{H}=\bigoplus_{i=1}^{m}\mathcal{H}_{i}. We say the decomposition is trivial if m=1m=1. We say one decomposition is better222Here we measure “better” only in terms of mm. We do not require any relationship between the subspaces of the better decomposition =i=1mi\mathcal{H}=\bigoplus_{i=1}^{m}\mathcal{H}_{i} and the worse =i=1mi\mathcal{H}=\bigoplus_{i=1}^{m^{\prime}}\mathcal{H}^{\prime}_{i} for m>mm>m^{\prime}. Note that even for two commuting algebras 𝒜,𝒜^L()\mathcal{A},\hat{\mathcal{A}}\subseteq L(\mathcal{H}), the two decompositions of \mathcal{H} induced by 𝒜,𝒜^\mathcal{A},\hat{\mathcal{A}} might not be finer than each other. That’s why we use “better” rather than ”finer” here. We define in this way just to ease notations and make our proof more precise. than another if it has a bigger mm.

Lemma 11 implies all operators in 𝒜\mathcal{A} keep the decomposition =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} invariant. It is worth noting that the decomposition induced by 𝒜\mathcal{A} might not be the best decomposition that 𝒜\mathcal{A} keeps invariant. In particular, consider the CC^{*}-algebra 𝒜\mathcal{A} generated by II, i.e. {cI}c\{cI\}_{c\in\mathbb{C}}. The decomposition of \mathcal{H} induced by 𝒜\mathcal{A} is trivial, i.e. =1\mathcal{H}=\mathcal{H}_{1}, but 𝒜\mathcal{A} keeps any decomposition of \mathcal{H} invariant. Using Lemma 11, we can analyze how two induced algebras can commute with each other.

Corollary 13 (The Structure Lemma)

Let 𝒜h\mathcal{A}_{h} be a CC^{*}-algebra acting on a finite dimensional \mathcal{H}. Let =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i}, i=i1i2\mathcal{H}_{i}=\mathcal{H}_{i}^{1}\otimes\mathcal{H}_{i}^{2}, is the decomposition induced by 𝒜h\mathcal{A}_{h} by Lemma 11. Consider another CC^{*}-algebra 𝒜h\mathcal{A}_{h^{\prime}} on \mathcal{H} which commutes with 𝒜h\mathcal{A}_{h}, we have

𝒜h=i(i1)(i2)\displaystyle\mathcal{A}_{h}=\bigoplus_{i}\mathcal{L}(\mathcal{H}_{i}^{1})\otimes\mathcal{I}(\mathcal{H}_{i}^{2})
𝒜hi(i1)(i2)\displaystyle\mathcal{A}_{h^{\prime}}\subseteq\bigoplus_{i}\mathcal{I}(\mathcal{H}_{i}^{1})\otimes\mathcal{L}(\mathcal{H}_{i}^{2})

Especially, all operators in 𝒜h,𝒜h\mathcal{A}_{h},\mathcal{A}_{h^{\prime}} keep the decomposition =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} invariant.

Proof.

Firstly by Lemma 11 we can get the decomposition of \mathcal{H} induced by 𝒜h\mathcal{A}_{h}. Further let Πi\Pi_{i} be the projection onto i\mathcal{H}_{i}, by Lemma 11 we know Πi𝒵(𝒜h)𝒜h\Pi_{i}\in\mathcal{Z}(\mathcal{A}_{h})\subseteq\mathcal{A}_{h}. Since 𝒜h\mathcal{A}_{h^{\prime}} commutes with 𝒜h\mathcal{A}_{h}, thus 𝒜h\mathcal{A}_{h^{\prime}} commutes with Πi\Pi_{i}, thus 𝒜h\mathcal{A}_{h^{\prime}} keeps i\mathcal{H}_{i} invariant. Since only cIcI can commute with all operators in a Hilbert space, i.e. (Hi1)\mathcal{L}(H_{i}^{1}), thus we finish the proof. ∎

In the following, we give a sufficient condition that implies the decomposition of space induced by the CC^{*}-algebra is non-trivial.

Lemma 14 (Non-trivial decomposition)

Let 𝒜\mathcal{A} be a CC^{*}-algebra on a finite dimensional \mathcal{H}. Denote the decomposition induced by 𝒜\mathcal{A} in Lemma 11 be =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i}. Consider another CC^{*}-algebra 𝒜\mathcal{A}^{\prime} on \mathcal{H} which commutes with 𝒜\mathcal{A}. If h0𝒜,h0𝒜\exists h\neq 0\in\mathcal{A},h^{\prime}\neq 0\in\mathcal{A}^{\prime} such that hh=0hh^{\prime}=0. Then the decomposition =iHi\mathcal{H}=\bigoplus_{i}H_{i} is non-trivial.

Proof.

With contradiction suppose the decomposition is trivial, i.e.

=1=1112\displaystyle\mathcal{H}=\mathcal{H}_{1}=\mathcal{H}^{1}_{1}\otimes\mathcal{H}^{2}_{1}

By Corollary 13 we have

h𝒜=(11)12,\displaystyle h\in\mathcal{A}=\mathcal{L}(\mathcal{H}^{1}_{1})\otimes\mathcal{I}_{\mathcal{H}^{2}_{1}},
h𝒜11(12).\displaystyle h^{\prime}\in\mathcal{A}^{\prime}\subseteq\mathcal{I}_{\mathcal{H}^{1}_{1}}\otimes\mathcal{L}(\mathcal{H}^{2}_{1}).

Since h0,h0h\neq 0,h^{\prime}\neq 0, we have hh0hh^{\prime}\neq 0 which leads to a contradiction. ∎

3.3 Partitions Inducted by Commuting Operators

The following definitions will be used throughout Sec. 5.

Definition 15

Let h,hh,h^{\prime} be two Hermitian terms acting on 𝒳𝒵\mathcal{X}\otimes\mathcal{H}\otimes\mathcal{Z} where dim()=ddim(\mathcal{H})=d. Suppose that hh acts trivially on 𝒵\mathcal{Z}, hh^{\prime} acts trivially on 𝒳\mathcal{X}, [h,h]=0[h,h^{\prime}]=0, and at least one of h,hh,h^{\prime} acts non-trivially on \mathcal{H}. Let the decomposition =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} be the better one induced by 𝒜h\mathcal{A}_{h}^{\mathcal{H}} or 𝒜h\mathcal{A}_{h^{\prime}}^{\mathcal{H}}. We say that h,hh,h^{\prime} commute in (d1,,dm)(d_{1},...,d_{m})-way on \mathcal{H} if dim(i)=didim(\mathcal{H}_{i})=d_{i}.

Note that by Corollary 13, h,hh,h^{\prime} have a tensor-product structure on i\mathcal{H}_{i}. Since the dimension of any Hilbert space must be an integer, two terms on a qutrit qq of dimension 33 can only commute in the following ways.

Lemma 16

Let h,hh,h^{\prime} be two Hermitian terms acting on 𝒳𝒵\mathcal{X}\otimes\mathcal{H}\otimes\mathcal{Z} where dim()=3dim(\mathcal{H})=3. If hh acts trivially on 𝒵\mathcal{Z}, hh^{\prime} acts trivially on 𝒳\mathcal{X}, [h,h]=0[h,h^{\prime}]=0, and at least one of h,hh,h^{\prime} acts non-trivially on \mathcal{H}. Let the decomposition =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} be the better one induced by 𝒜h\mathcal{A}_{h}^{\mathcal{H}} or 𝒜h\mathcal{A}_{h^{\prime}}^{\mathcal{H}}. then h,hh,h^{\prime} must commute on \mathcal{H} via one of the following ways

  • (1,1,1)(1,1,1)-way: =123\mathcal{H}=\mathcal{H}_{1}\bigoplus\mathcal{H}_{2}\bigoplus\mathcal{H}_{3}, where dim(i)=1,idim(\mathcal{H}_{i})=1,\forall i.

  • (1,2)(1,2)-way: =12\mathcal{H}=\mathcal{H}_{1}\bigoplus\mathcal{H}_{2}, dim(1)=1,dim(2)=2dim(\mathcal{H}_{1})=1,dim(\mathcal{H}_{2})=2.

  • (3)(3)-way: =1\mathcal{H}=\mathcal{H}_{1}, dim(1)=3dim(\mathcal{H}_{1})=3. One of h,hh,h^{\prime} acts trivially on \mathcal{H}, and for another the induced algebra on \mathcal{H} is the full algebra ()\mathcal{L}(\mathcal{H}).

Proof.

By corollary 13, we get a decomposition =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} where i=i1i2\mathcal{H}_{i}=\mathcal{H}^{1}_{i}\otimes\mathcal{H}^{2}_{i}. Since the dimension of a subspace must be an integer we get the above 3 possible ways. Further for the (3)(3)-way, by assumption the decomposition induced by both 𝒜h,𝒜h\mathcal{A}_{h}^{\mathcal{H}},\mathcal{A}_{h^{\prime}}^{\mathcal{H}} are =1\mathcal{H}=\mathcal{H}_{1}. Since dim(1)=dim()=3dim(\mathcal{H}_{1})=dim(\mathcal{H})=3 is a prime, which means it can only be a tensor product of a one-dimensional Hilbert space and a three-dimensional Hilbert space. Thus for both 𝒜h,𝒜h\mathcal{A}_{h}^{\mathcal{H}},\mathcal{A}_{h^{\prime}}^{\mathcal{H}}, they equal to either {cI}c\{cI\}_{c} or ()\mathcal{L}(\mathcal{H}). Since at least one of h,hh,h^{\prime} acts non-trivially on \mathcal{H}, we know one of the induced algebra are ()\mathcal{L}(\mathcal{H}), w.l.o.g suppose 𝒜h=()\mathcal{A}_{h}^{\mathcal{H}}=\mathcal{L}(\mathcal{H}). Again by corollary 13, 𝒜h\mathcal{A}_{h^{\prime}}^{\mathcal{H}} should be {cI}c\{cI\}_{c}, thus hh^{\prime} acts trivially on \mathcal{H}. ∎

Similar arguments for qubits are widely used in the proof of the qubit-CLHP-2D is in NP [Sch11][AKV18]. We summarize it as below:

Lemma 17

If we change dim()dim(\mathcal{H}) to be 22 in the statement of Lemma 16, then h,hh,h^{\prime} must commute on \mathcal{H} via one of the following ways

  • (1,1)(1,1)-way if =12\mathcal{H}=\mathcal{H}_{1}\bigoplus\mathcal{H}_{2} where dim(1)=dim(2)=1dim(\mathcal{H}_{1})=dim(\mathcal{H}_{2})=1.

  • (2)(2)-way if =1\mathcal{H}=\mathcal{H}_{1} where dim(1)=2dim(\mathcal{H}_{1})=2. One of h,hh,h^{\prime} acts trivially on \mathcal{H}, and for another the induced algebra on \mathcal{H} is the full algebra ()\mathcal{L}(\mathcal{H}).

Note that Lemma 16 and Lemma 17 only involve 2 commuting terms h,hh,h^{\prime}, and their overlapping space is only \mathcal{H}. Those techniques do not directly apply to 2D Hamiltonians, where some of the terms overlap on 2 qudits.

4 Qutrit Commuting Local Hamiltonian on 2D

In this section, we will prove that the qutrit-CLHP-2D is in NP. This proof is non-constructive. Note that if the qutrit-CLHP-2D-projection is in NP, then the qutrit-CLHP-2D is in NP. The proof of this statement is in Appendix. B. Thus in this section, w.l.o.g we assume that all the terms pp are projections and prove that the qutrit-CLHP-2D-projection is in NP.

The proof sketch is as follows. In Sec. 4.1 we prove that we can further assume that there are no semi-separable qudits. In Sec. 4.2 we prove that for the qutrit-CLHP-2D without semi-separable qudits, there are strong restrictions on the form of Hamiltonian. Finally, in Sec. 4.3 we prove that with such restrictions, we can use Schuch’s method [Sch11] again.

4.1 Self-reduction for CLHP with semi-separable qudits

Lemmas in this section work for qudit-CLHP-projection333W.l.o.g we can assume that all terms are projections by Lemma 48., that is we do not assume that each particle is a qutrit, or the Hamiltonian is on 2D. Recall that qudit-CLHP-projection is: Consider a qudit-CLHP H=ihiH=\sum_{i}h_{i} where {hi}i\{h_{i}\}_{i} are kk-local projections for some constant kk, [hi,hj]=0[h_{i},h_{j}]=0 for iji\neq j. The question is to determine whether λ(H)=0\lambda(H)=0 or λ(H)1\lambda(H)\geq 1. Note that λ(H)=0\lambda(H)=0 iff all the commuting projections {hi}i\{h_{i}\}_{i} have a common 0-eigenvector, i.e. ker{hi}iker\{h_{i}\}_{i} is non-trivial. When λ(H)=0\lambda(H)=0, the common 0-eigenvectors of {hi}i\{h_{i}\}_{i} are the ground states of HH. We also denote the Hamiltonian HH as {hi}i\{h_{i}\}_{i}.

The key lemma in this section, Lemma 23, is to prove that when there is a semi-separable qudit, the prover can perform a non-constructive self-reduction for the qudit-CLHP-projection. Here self-reduction means reducing the qudit-CLHP-projection to another qudit-CLHP-projection, where the Hilbert space of the qudit has a smaller dimension. Before going into the formal proofs, in the following, we intuitively explain how Lemma 23 works. Specifically, temporarily assume that λ(H)=0\lambda(H)=0 thus we are in the Yes instance and try to prove λ(H)=0\lambda(H)=0. We begin with the example when there is a separable qudit, then generalize this idea to the case of semi-separable qudit, and after that we give the formal proofs.

When there is a separable qudit qq, the prover can easily perform a constructive self-reduction. Suppose qq is a separable qudit, by definition, there exists a non-trivial decomposition q=jjq\mathcal{H}^{q}=\bigoplus_{j}\mathcal{H}_{j}^{q} such that all the terms {hi}i\{h_{i}\}_{i} keep the decomposition invariant. Then there must be a subspace j0\mathcal{H}_{j_{0}} which contains a common 0-eigenstate of {hi}i\{h_{i}\}_{i}. Denote the projector onto j0q\mathcal{H}_{j_{0}}^{q} as Πj0\Pi_{j_{0}}. The prover can give the decomposition q=jjq\mathcal{H}^{q}=\bigoplus_{j}\mathcal{H}_{j}^{q} and the index j0j_{0}. The verifier checks that qq is a separable qudit, then restricts the space of qq from q\mathcal{H}^{q} to j0q\mathcal{H}_{j_{0}}^{q}, and restricts all terms {hi}i\{h_{i}\}_{i} to {hi<j0>:=Πj0hiΠj0}i\{h_{i}^{<j_{0}>}:=\Pi_{j_{0}}h_{i}\Pi_{j_{0}}\}_{i}, and ask the prover to prove that {hi<j0>}i\{h_{i}^{<j_{0}>}\}_{i} has a common 0-eigenstate. By definition of separable, the decomposition is non-trivial, thus the new instance {hi<j0>}i\{h_{i}^{<j_{0}>}\}_{i} is strictly simpler in the sense that we strictly decrease the dimension of the qudit qq. Note that this method is constructive — the common 0-eigenstate of the new instance {hi<j0>}i\{h_{i}^{<j_{0}>}\}_{i} is also the common 0-eigenstate of the original instance {hi}i\{h_{i}\}_{i}.

Our key observation is, for qudit-CLHP-projection, even for semi-separable qudit, the NP prover is able to perform a similar self-reduction, via a non-constructive way. If we follow the intuition of the separable qudit case, one might try restricting qjq\mathcal{H}^{q}\rightarrow\mathcal{H}^{q}_{j}, and transform every term to be ΠjhiΠj\Pi_{j}h_{i}\Pi_{j}. The problem is that since h0h_{0} does not keep jq\mathcal{H}^{q}_{j} invariant, and does not commute with Πj\Pi_{j}, the Πjh0Πj\Pi_{j}h_{0}\Pi_{j} is no longer a projection. One may also doubt whether the resulting Hamiltonian is commuting. A more serious problem is that, unlike the case for separable qudit, since h0h_{0} does not keep jq\mathcal{H}^{q}_{j} invariant, it is not clear how to connect the ground states of the original Hamiltonian to the ground states of the new Hamiltonian. We circumvent the problems by slightly changing the construction — rounding Πjh0Πj\Pi_{j}h_{0}\Pi_{j} to its 11-eigenspace.

Definition 18 (Reduced Hamiltonian)

Consider a semi-separable qudit qq w.r.t commuting projections {hi}i\{h_{i}\}_{i} and a non-trivial decomposition q=j=1mjq\mathcal{H}^{q}=\bigoplus_{j=1}^{m}\mathcal{H}^{q}_{j}, where Πj\Pi_{j} is the projection onto jq\mathcal{H}^{q}_{j}. For any jj, we define its jj-th reduced Hamiltonian to be {hi(j)}i\{h_{i}^{(j)}\}_{i}, or written as H(j):=ihi(j)H^{(j)}:=\sum_{i}h_{i}^{(j)}, where

  • hi(j)=ΠjhiΠjh_{i}^{(j)}=\Pi_{j}h_{i}\Pi_{j}, for i1i\geq 1.

  • h0(j)h_{0}^{(j)} is the projection onto the 11-eigenspace of Πjh0Πj\Pi_{j}h_{0}\Pi_{j}. Assign h0(j)h_{0}^{(j)} to be 0 when the 11-eigenspace is empty. It is equivalent to interpret h0(j)h^{(j)}_{0} is obtained by rounding all the strictly-smaller-than-1-eigenvalues of Πjh0Πj\Pi_{j}h_{0}\Pi_{j} to 0.

  • We restrict the space of qq from q\mathcal{H}^{q} to jq\mathcal{H}^{q}_{j}. Note that all terms hi(j)h_{i}^{(j)} including h0(j)h_{0}^{(j)} keeps jq\mathcal{H}^{q}_{j} invariant, thus this restriction of space is well-defined. In summary, the original Hamiltonian HH acts on q(qqq)\mathcal{H}^{q}\otimes\left(\otimes_{q^{\prime}\neq q}\mathcal{H}^{q^{\prime}}\right), the jj-th reduced Hamiltonian H(j)H^{(j)} acts on jq(qqq)\mathcal{H}^{q}_{j}\otimes\left(\otimes_{q^{\prime}\neq q}\mathcal{H}^{q^{\prime}}\right).

Note that the construction of reduced Hamiltonian is consistent with our previous intuition for the separable qudit — If qq is separable and h0h_{0} also keeps jq\mathcal{H}^{q}_{j} invariant, then Πjh0Πj\Pi_{j}h_{0}\Pi_{j} is a projection thus h0(j)=Πjh0Πjh_{0}^{(j)}=\Pi_{j}h_{0}\Pi_{j}. It is worth noting that the reduced Hamiltonian keeps the “geometry” of the original Hamiltonian:

Lemma 19

In Def. 18, if hih_{i} acts trivially on qudit qq^{\prime} w.r.t space q\mathcal{H}^{q^{\prime}}, then hi(j)h_{i}^{(j)} acts trivially on qudit qq^{\prime} w.r.t space q\mathcal{H}^{q^{\prime}} if qqq^{\prime}\neq q or jq\mathcal{H}^{q^{\prime}}_{j} if q=qq^{\prime}=q . Especially, If H=ihiH=\sum_{i}h_{i} is kk-local (or on 2D), so does the jj-th reduced Hamiltonian H(j)=ihi(j)H^{(j)}=\sum_{i}h_{i}^{(j)}.

Proof.

We prove that if h0h_{0} acts trivially on qudit qq^{\prime}, then so does h0(j)h_{0}^{(j)}, the proof for i0i\neq 0 is similar. By assumption, h0h_{0} acts trivially on qq^{\prime}, thus h0=Iqhh_{0}=I_{\mathcal{H}^{q^{\prime}}}\otimes h for some projection hh. Recall that qq is the semi-separable qudit in Def. 18. If q=qq^{\prime}=q, then

Πjh0Πj\displaystyle\Pi_{j}h_{0}\Pi_{j} =Πjh\displaystyle=\Pi_{j}\otimes h (4)
=Ijqh.\displaystyle=I_{\mathcal{H}^{q^{\prime}}_{j}}\otimes h. (5)

is a projection and acts trivially on jq\mathcal{H}_{j}^{q^{\prime}}. If qqq^{\prime}\neq q, Πjh0Πj=IqΠjhΠj\Pi_{j}h_{0}\Pi_{j}=I_{\mathcal{H}^{q^{\prime}}}\otimes\Pi_{j}h\Pi_{j}, the 1-eigenspace of Πjh0Πj\Pi_{j}h_{0}\Pi_{j} is also of form IqI_{q^{\prime}}\otimes... thus acts trivially on qq^{\prime}. ∎

Besides, the terms in the reduced Hamiltonian are commuting projections.

Lemma 20

If in Def. 18, {hi}i\{h_{i}\}_{i} are commuting projections, then for any jj, the jj-th reduced Hamiltonian {hi(j)}i\{h^{(j)}_{i}\}_{i} are commuting projections.

Proof.

Notice that, by the definition of semi-separable, we have [hi,Πj]=0,i0[h_{i},\Pi_{j}]=0,\forall i\neq 0. It is also important to keep in mind that [h0,Πj][h_{0},\Pi_{j}] might not equal 0.

Firstly we can check that all terms {hi(j)}i\{h^{(j)}_{i}\}_{i} are projections. Notice that h0(j)h_{0}^{(j)} is a projection by definition. For i0i\neq 0, since hih_{i} is a projection, and [hi,Πj]=0[h_{i},\Pi_{j}]=0, we know hi(j):=ΠjhiΠjh^{(j)}_{i}:=\Pi_{j}h_{i}\Pi_{j} is a projection444The most direct way to understand this is imagining Πj,hi\Pi_{j},h_{i} are diagonal 0,10,1 matrix, since they are commuting they can be diagonalized simultaneously.. In summary, all the terms are projections.

Then we prove that all terms are commuting. Notice that for any i0i\neq 0, for any ii^{\prime}, where ii^{\prime} can be 0, we have

(ΠjhiΠj)(ΠjhiΠj)\displaystyle(\Pi_{j}h_{i}\Pi_{j})(\Pi_{j}h_{i^{\prime}}\Pi_{j}) =(ΠjΠj)(ΠjhihiΠj)\displaystyle=(\Pi_{j}\Pi_{j})(\Pi_{j}h_{i}h_{i^{\prime}}\Pi_{j}) (6)
=(ΠjΠj)(ΠjhihiΠjΠj)\displaystyle=(\Pi_{j}\Pi_{j})(\Pi_{j}h_{i^{\prime}}h_{i}\Pi_{j}\Pi_{j}) (7)
=(ΠjΠj)(ΠjhiΠjhiΠj)\displaystyle=(\Pi_{j}\Pi_{j})(\Pi_{j}h_{i^{\prime}}\Pi_{j}h_{i}\Pi_{j}) (8)
=(ΠjhiΠj)(ΠjhiΠj)\displaystyle=(\Pi_{j}h_{i^{\prime}}\Pi_{j})(\Pi_{j}h_{i}\Pi_{j}) (9)

where Eq. (6) is from [hi,Πj]=0[h_{i},\Pi_{j}]=0, Eq. (7) is from [hi,hi]=0[h_{i},h_{i^{\prime}}]=0 and Πj2=Πj\Pi_{j}^{2}=\Pi_{j}, Eq. (8) is from [hi,Πj]=0[h_{i},\Pi_{j}]=0. Note that we never assume that hih_{i^{\prime}} commutes with Πj\Pi_{j}.

From the Eq. (9), we know [hi(j),hi(j)]=0[h_{i}^{(j)},h_{i^{\prime}}^{(j)}]=0 if i0,i0i\neq 0,i^{\prime}\neq 0. Besides, we know for i0i\neq 0, [hi(j),Πjh0Πj]=0[h_{i}^{(j)},\Pi_{j}h_{0}\Pi_{j}]=0. Thus hi(j)h_{i}^{(j)} keeps the 11-eigenspace555We emphasize it is the space spanned by all 11-eigenvector, rather than one of the eigenvector. of Πjh0Πj\Pi_{j}h_{0}\Pi_{j} invariant. Since hi(j)h_{i}^{(j)} is Hermitian, this implies hi(j)h_{i}^{(j)} commutes with the projection onto this 1-eigenspace of Πjh0Πj\Pi_{j}h_{0}\Pi_{j}, thus [hi(j),h0(j)]=0[h_{i}^{(j)},h_{0}^{(j)}]=0. In summary, all terms are commuting. ∎

In summary, we have

Corollary 21 (of Lemma 19 and Lemma 20)

If {hi}i\{h_{i}\}_{i} are kk-local (or on 2D) qudit commuting projections, then so does the jj-th reduced Hamiltonian {hi(j)}i\{h_{i}^{(j)}\}_{i}.

In addition, we give a cute lemma – Lemma 22. In the lemma description, the right side of Eq. (10) is just rounding all non-zero coefficients (1λ)(1-\lambda) to 11. This Lemma is simple itself but captures the key idea of “rounding” used in Lemma 23. It will explain why we can round all non-11 eigenvalue of Πjh0Πj\Pi_{j}h_{0}\Pi_{j} to 0, and only use the 11-eigenspace.

Lemma 22

Let f(j,λ)f(j,\lambda) be a non-negative function. Then

jλ1(1λ)f(j,λ)>0 iff jλ<1f(j,λ)>0.\displaystyle\sum_{j}\sum_{\lambda\leq 1}(1-\lambda)f(j,\lambda)>0\text{ iff }\sum_{j}\sum_{\lambda<1}f(j,\lambda)>0. (10)
Proof.

Since f(j,λ)f(j,\lambda) is non-negative. It suffices to notice that both the left and the right inequalities are equivalent to j,λ<1\exists j,\exists\lambda<1 s.t. f(j,λ)>0f(j,\lambda)>0. ∎

Now we are prepared to state our key lemma, which connects the original Hamiltonian to the reduced Hamiltonians. Inspired by Schuch’s idea [Sch11], we will decompose a non-negative term into summation over many non-negative terms. However, our method here uses very different decomposition rules, and has key differences from his, which will be discussed in more detail after the proof.

Lemma 23 (Decrease dimension and rounding)

Consider an instance of qudit-CLHP-projection on nn qudits, where the Hamiltonian is denoted as H=ihiH=\sum_{i}h_{i} for commuting projections {hi}i\{h_{i}\}_{i}. Suppose there is a semi-separable qudit qq w.r.t. {hi}i\{h_{i}\}_{i} and non-trivial decomposition q=j=1mjq\mathcal{H}^{q}=\bigoplus_{j=1}^{m}\mathcal{H}^{q}_{j}. For every jj, define the jj-th reduced Hamiltonian H(j)=ihi(j)H^{(j)}=\sum_{i}h_{i}^{(j)} as in Definition 18. Then

λ(H)=0 iff j s.t λ(H(j))=0.\displaystyle\lambda(H)=0\text{\quad iff \quad}\exists j\text{ s.t }\lambda(H^{(j)})=0. (11)
Proof.

Denote the nn-qudit space as =qq\mathcal{H}=\otimes_{q}\mathcal{H}^{q}. Define j=jqqqq\mathcal{H}_{j}=\mathcal{H}^{q}_{j}\otimes_{q^{\prime}\neq q}\mathcal{H}^{q^{\prime}}. For clarity, in this proof, we use II for the identity on ()\mathcal{L}(\mathcal{H}). When using tr(h)tr(h) we always view hh as an operator on \mathcal{H}, and we project out all the qudits, i.e. tr(h):=ii|h|itr(h):=\sum_{i}\langle i|h|i\rangle where {|i}i\{\left|i\right\rangle\}_{i} is the computational basis for \mathcal{H}. Especially, we view Πj\Pi_{j} as an operator in \mathcal{H}, while view IjI_{\mathcal{H}_{j}} as an operator in j\mathcal{H}_{j}.

Note that {hi}i\{h_{i}\}_{i} are commuting projections, proving λ(H)=0\lambda(H)=0 is equivalent to show that

tr[i(Ihi)]>0.\displaystyle tr\left[\prod_{i}(I-h_{i})\right]>0. (12)

Since {hi}i\{h_{i}\}_{i} are commuting, the relative order in the above formula is unimportant. Recall that Πj\Pi_{j} is the projection onto jq\mathcal{H}_{j}^{q}. By assumption i0\forall i\neq 0, hih_{i} keeps jq\mathcal{H}_{j}^{q} invariant, thus

tr[i(Ihi)]\displaystyle tr\left[\prod_{i}(I-h_{i})\right] =tr[(Ih0)i0(jΠj(Ihi)Πj)]\displaystyle=tr\left[\left(I-h_{0}\right)\prod_{i\neq 0}\left(\sum_{j}\Pi_{j}(I-h_{i})\Pi_{j}\right)\right] (13)
=tr[(Ih0)ji0(Πj(Ihi)Πj)]\displaystyle=tr\left[\left(I-h_{0}\right)\sum_{j}\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\right] (14)
=jtr[(Ih0)i0(Πj(Ihi)Πj)]\displaystyle=\sum_{j}tr\left[\left(I-h_{0}\right)\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\right] (15)

Eq (13) is from jΠj=I\sum_{j}\Pi_{j}=I, and for i0i\neq 0, hih_{i} is Hermitian and keeps jq\mathcal{H}^{q}_{j} invariant thus jΠjhiΠj=hi\sum_{j}\Pi_{j}h_{i}\Pi_{j}=h_{i}. Eq (14) is from {Πj}j\{\Pi_{j}\}_{j} are orthogonal from each other. Besides, since Πj2=Πj\Pi_{j}^{2}=\Pi_{j} and tr(MΠj)=tr(ΠjM)tr(M\Pi_{j})=tr(\Pi_{j}M) for arbitrary MM, we have

tr[i(Ihi)]\displaystyle tr\left[\prod_{i}(I-h_{i})\right] =jtr[(Ih0)Πji0(Πj(Ihi)Πj)Πj]\displaystyle=\sum_{j}tr\left[\left(I-h_{0}\right)\Pi_{j}\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\Pi_{j}\right] (16)
=jtr[Πj(Ih0)Πji0(Πj(Ihi)Πj)]\displaystyle=\sum_{j}tr\left[\Pi_{j}\left(I-h_{0}\right)\Pi_{j}\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\right] (17)
=jtr[Πj(IΠjh0Πj)Πji0(Πj(Ihi)Πj)]\displaystyle=\sum_{j}tr\left[\Pi_{j}\left(I-\Pi_{j}h_{0}\Pi_{j}\right)\Pi_{j}\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\right] (18)

Eq. (18) if from Πj2=Πj\Pi_{j}^{2}=\Pi_{j}. Note that Πjh0Πj\Pi_{j}h_{0}\Pi_{j} is an nn-qudit Hermitian operator, thus it can be diagonalized by a unitary matrix. Consider its eigenvalue decomposition, denote the eigenvalues and projections onto the corresponding eigenspace as λ\lambda, Πj,λ\Pi_{j,\lambda}. That is

Πjh0Πj=λλΠj,λ\displaystyle\Pi_{j}h_{0}\Pi_{j}=\sum_{\lambda}\lambda\Pi_{j,\lambda} (19)

Note that it might be possible that Πj,λ\Pi_{j,\lambda} acts non-trivially on some qqq^{\prime}\neq q as long as h0h_{0} acts non-trivially on qq^{\prime}. Besides, by definition of Πj,λ\Pi_{j,\lambda} we have

λΠj,λ=I.\displaystyle\sum_{\lambda}\Pi_{j,\lambda}=I. (20)

Since h0h_{0} is a projection, we have Πjh0Πj0\Pi_{j}h_{0}\Pi_{j}\succeq 0 and λ[0,1]\lambda\in[0,1]. Use Eqs (19),(20), we have Eq. (18) becomes

tr[i(Ihi)]\displaystyle tr\left[\prod_{i}(I-h_{i})\right] =jtr[(Πj(λ1(1λ)Πj,λ)Πj)i0(Πj(Ihi)Πj)]\displaystyle=\sum_{j}tr\left[\left(\Pi_{j}\left(\sum_{\lambda\leq 1}(1-\lambda)\Pi_{j,\lambda}\right)\Pi_{j}\right)\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\right] (21)
=jλ1(1λ)tr[(ΠjΠj,λΠj)i0(Πj(Ihi)Πj)]\displaystyle=\sum_{j}\sum_{\lambda\leq 1}\left(1-\lambda\right)tr\left[\left(\Pi_{j}\Pi_{j,\lambda}\Pi_{j}\right)\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\right] (22)

Define

f(j,λ)\displaystyle f(j,\lambda) :=tr[(ΠjΠj,λΠj)i0(Πj(Ihi)Πj)]\displaystyle:=tr\left[\left(\Pi_{j}\Pi_{j,\lambda}\Pi_{j}\right)\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\right] (23)

Since {Πj(Ihi)Πj}i0\{\Pi_{j}(I-h_{i})\Pi_{j}\}_{i\neq 0} are commuting projections, we know i0(Πj(Ihi)Πj)0\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\succeq 0 and is Hermitian. Note that ΠjΠj,λΠj0\Pi_{j}\Pi_{j,\lambda}\Pi_{j}\succeq 0 and is Hermitian. Since the trace of the product of two positive semi-definite Hermitian matrices is non-negative666Let A,BA,B to be arbitrary two Hermitian matrices where A0,B0A\succeq 0,B\succeq 0. Since A,BA,B are Hermitian, consider the eigenvalue decompositions A=iai|ϕiϕi|,ai0A=\sum_{i}a_{i}|\phi_{i}\rangle\langle\phi_{i}|,a_{i}\geq 0, B=jbj|ψjψj|,bj0B=\sum_{j}b_{j}|\psi_{j}\rangle\langle\psi_{j}|,b_{j}\geq 0. Then tr(AB)=i,jaibj|ϕi|ψj|20tr(AB)=\sum_{i,j}a_{i}b_{j}|\langle\phi_{i}|\psi_{j}\rangle|^{2}\geq 0. , we have f(j,λ)f(j,\lambda) is non-negative,

f(j,λ)0,j,λ.\displaystyle f(j,\lambda)\geq 0,\forall j,\lambda. (24)

By Lemma 22, tr[i(Ihi)]>0tr\left[\prod_{i}(I-h_{i})\right]>0 is equivalent to rounding all the non-zero coefficients in Eq. (22) to 1, that is equivalent as showing that

jλ<1tr[(ΠjΠj,λΠj)i0(Πj(Ihi)Πj)]>0\displaystyle\sum_{j}\sum_{\lambda<1}tr\left[\left(\Pi_{j}\Pi_{j,\lambda}\Pi_{j}\right)\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\right]>0 (25)
\displaystyle\Leftrightarrow jtr[(Πj(λ<1Πj,λ)Πj)i0(Πj(Ihi)Πj)]>0\displaystyle\sum_{j}tr\left[\left(\Pi_{j}\left(\sum_{\lambda<1}\Pi_{j,\lambda}\right)\Pi_{j}\right)\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\right]>0 (26)
\displaystyle\Leftrightarrow jtr[(Πj(Ih0(j))Πj)i0(Πj(Ihi)Πj)]>0\displaystyle\sum_{j}tr\left[\left(\Pi_{j}\left(I-h_{0}^{(j)}\right)\Pi_{j}\right)\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\right]>0 (27)
\displaystyle\Leftrightarrow jtr[(Πjh0(j))i0(Πj(Ihi)Πj)]>0.\displaystyle\sum_{j}tr\left[\left(\Pi_{j}-h_{0}^{(j)}\right)\prod_{i\neq 0}\left(\Pi_{j}(I-h_{i})\Pi_{j}\right)\right]>0. (28)
\displaystyle\Leftrightarrow jtr[i(Πjhi(j))]>0\displaystyle\sum_{j}tr\left[\prod_{i}\left(\Pi_{j}-h^{(j)}_{i}\right)\right]>0 (29)

Eq. (27) is from Eq. (20) and the definition of h0(j)h_{0}^{(j)}. Eq. (28) is from Πjh0(j)Πj=h0(j).\Pi_{j}h_{0}^{(j)}\Pi_{j}=h_{0}^{(j)}. Note that Eqs. (25-29) and Eq. (24) imply that

tr[i(Πjhi(j))]\displaystyle tr\left[\prod_{i}\left(\Pi_{j}-h^{(j)}_{i}\right)\right] =λ<1f(j,λ)\displaystyle=\sum_{\lambda<1}f(j,\lambda) (30)
0.\displaystyle\geq 0. (31)

Thus we further have

(29)\displaystyle(\ref{eq:24})\Leftrightarrow j s.t tr[i(Πjhi(j))]>0\displaystyle\exists j\text{ s.t }tr\left[\prod_{i}\left(\Pi_{j}-h^{(j)}_{i}\right)\right]>0 (32)
\displaystyle\Leftrightarrow j s.t tr(j)[i(Ijhi(j))]>0\displaystyle\exists j\text{ s.t }tr^{(j)}\left[\prod_{i}\left(I_{\mathcal{H}_{j}}-h^{(j)}_{i}\right)\right]>0 (33)

where in Eq. (33), the notation tr(j)tr^{(j)} means now we restrict the space of qudit qq from qjq\mathcal{H}^{q}\rightarrow\mathcal{H}^{q}_{j}, and the trace is over j\mathcal{H}_{j}. Note that Eq. (33) is well defined since i(Ijhi(j))\prod_{i}\left(I_{\mathcal{H}_{j}}-h^{(j)}_{i}\right) keeps j\mathcal{H}_{j} invariant.

By Lemma 20 we know {hi(j)}\{h_{i}^{(j)}\} are commuting projections on j=jq(qqq)\mathcal{H}_{j}=\mathcal{H}^{q}_{j}\otimes\left(\otimes_{q^{\prime}\neq q}\mathcal{H}^{q}\right). Eq. (33) is equivalent to say j\exists j such that the jj-th reduced Hamiltonian {hi(j)}i\{h^{(j)}_{i}\}_{i} has a common 0-eigenvector, where the space of qq is jq\mathcal{H}_{j}^{q} and the nn-qudit space is j\mathcal{H}_{j}. ∎

Corollary 24

If qudit-CLHP-projection without semi-separable qudit is in NP, then qudit-CLHP-projection is in NP.

Proof.

Lemma 23 says when there is a semi-separable qudit, an NP prover can efficiently reduce the qudit-CLHP-projection to a new qudit-CLHP-projection by strictly decreasing the dimension of qq. Since dim(q)dim(q) is a constant, an NP prover can reach to a qudit-CLHP-projection without semi-separable qudit by repeatedly performing Lemma 23 in poly(n)poly(n) time. ∎

To help better understand Lemma 23, let us discuss the differences between Lemma 23 and the method used in Schuch’s paper [Sch11]. We both decompose some of the terms and get a summation of non-negative quantities, but we do such decomposition and projection in different ways. The key difference is that in our method, we guarantee all the quantities still correspond to a commuting local Hamiltonian problem. Our method is more like self-reduction, and we can perform such self-reduction sequentially until this are no semi-separable qudits. On the contrary, in Schuch’s method, the two projections he used for each qudit777Schuch wrote his proof in terms of the qubit, but the first decomposition part also works for qudit., are not commuting with each other, and the quantity does not correspond to commuting local Hamiltonian anymore, thus this decomposition technique can only be performed once rather than sequentially.

4.2 Restrictions on the qutrit-CLHP-2D without semi-separable qudit

From now on we will consider the 2D geometry and start our proof for the qutrit-CLHP-2D-projection is in NP. Recall that we allow qudits in the qutrit-CLHP-2D-projection to have different dimensions, i.e. either 1,2 or 3. By Corollary 24, we can assume that there are no semi-separable qudits. This “no semi-separable qudits condition”, combined with the 2D geometry, will lead to strong restrictions on the form of the Hamiltonian, i.e. Lemma 26.

We define some notations. As shown in Fig. 2, when considering the qutrit-CLHP-2D-projection, for a qudit qq, we use p1,p2p_{1},p_{2} to denote the two plaquette projections in the diagonal direction, p1,p2p^{\prime}_{1},p_{2}^{\prime} to denote the two plaquettes projections in the anti-diagonal direction. In the whole Sec. 4.2, the relative positions of q,p1,p2,p1,p2q,p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime} will always obey Fig. 2. We give the following definitions.

q4q_{4}q5q_{5}qqq1q_{1}q2q_{2}q3q_{3}p1p_{1}^{\prime}p2p_{2}p4p_{4}^{\prime}p1p_{1}p2p_{2}^{\prime}p3p_{3}
Figure 2: Notations for the qutrit-CLHP-2D-projection.
Definition 25

Consider a qutrit-CLHP-2D-projection instance, for any qutrit qq of dimension 33, use the notations in Fig. 2. For A,B{(1,1,1),(1,2),(3)}A,B\in\{(1,1,1),(1,2),(3)\}, we say that p1,p2,p1,p2p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime} acting on qq via A×BA\times B-way, if p1,p2p_{1},p_{2} commute in AA-way888See Definition 15. More specifically, denote the set of qudits that p1,p2p_{1},p_{2} acting non-trivially on as S1,S2S_{1},S_{2}, then in Definition 15 :=q,𝒳:=qS1/qq,𝒵:=qS2/qq\mathcal{H}:=\mathcal{H}^{q},\mathcal{X}:=\otimes_{q^{\prime}\in S_{1}/q}\mathcal{H}^{q^{\prime}},\mathcal{Z}:=\otimes_{q^{\prime}\in S_{2}/q}\mathcal{H}^{q^{\prime}}. on q\mathcal{H}^{q}, and p1,p2p_{1}^{\prime},p_{2}^{\prime} commute in BB-way on q\mathcal{H}^{q}; or verse visa, i.e. p1,p2p_{1},p_{2} commute in BB-way on q\mathcal{H}^{q}, and p1,p2p_{1}^{\prime},p_{2}^{\prime} commute in AA-way on q\mathcal{H}^{q}.

Note that p1,p2p_{1},p_{2} only overlap on one qudit – qq – thus the above sentence “p1,p2p_{1},p_{2} commute in A-way on q\mathcal{H}^{q}” is well-defined. Same for p1,p2p^{\prime}_{1},p^{\prime}_{2}. On the other hand, p1,p1p_{1},p_{1}^{\prime} overlap on two qudits, thus we cannot say p1,p1p_{1},p_{1}^{\prime} commute in some way on q\mathcal{H}^{q}. Another clarification is, that it is possible that some of the terms are identity, eg. p1=p2=Ip_{1}=p_{2}=I, then the situation does not belong to Definition 25. Those cases will be considered separately and solved easily in the related proofs.

Recall that by Corollary 24, we can assume that there are no semi-separable qudits. This will imply certain ways of commuting cannot exist for the qutrit-CLHP-2D-projection.

Lemma 26 (Legal ways of commuting)

Consider a qutrit-CLHP-2D-projection Hamiltonian {p}p\{p\}_{p}, if there is no semi-separable qudit, then there is no qutrit qq with dim(q)=3dim(\mathcal{H}^{q})=3 such that terms p1,p2,p1,p2p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime} acting on qq via (1,2)×(3)(1,2)\times(3)-way or (1,2)×(1,2)(1,2)\times(1,2)-way.

Proof.

To ease notations, in this proof we use \mathcal{H} to denote the Hilbert space of qq, instead of using q\mathcal{H}^{q}. With contradiction, assume that there is a qudit qq with dim(q)=3dim(\mathcal{H}^{q})=3 such that

  • p1,p2p_{1},p_{2} commute via (1,2)(1,2)-way on q\mathcal{H}^{q}. The decomposition w.r.t p1,p2p_{1},p_{2} is =12\mathcal{H}=\mathcal{H}_{1}\bigoplus\mathcal{H}_{2}, where dim(1)=1dim(\mathcal{H}_{1})=1, 1=span(|ψ)\mathcal{H}_{1}=span(\left|\psi\right\rangle)999One may wonder how could 1\mathcal{H}_{1} written as 1112\mathcal{H}_{1}^{1}\otimes\mathcal{H}_{1}^{2} in the Structure Lemma. Conceptually one can interpret 11=span{|ψ}\mathcal{H}_{1}^{1}=span\{\left|\psi\right\rangle\}, and 12\mathcal{H}_{1}^{2} as a one-dimensional space as scalars {c}c\{c\}_{c}. for some |ψq\left|\psi\right\rangle\in\mathcal{H}^{q}, and dim(2)=2dim(\mathcal{H}_{2})=2, 2=2122\mathcal{H}_{2}=\mathcal{H}_{2}^{1}\otimes\mathcal{H}_{2}^{2}. Since 22 is a prime and 2=2×12=2\times 1, by definition the decomposition =12\mathcal{H}=\mathcal{H}_{1}\bigoplus\mathcal{H}_{2} is induced by p1p_{1} or p2p_{2}, and [p1,p2]=0[p_{1},p_{2}]=0, by Corollary 13, one can check that one of p1,p2p_{1},p_{2} must acts trivially on 2\mathcal{H}_{2}. W.l.o.g assume that dim(21)=2,dim(22)=1dim(\mathcal{H}_{2}^{1})=2,dim(\mathcal{H}_{2}^{2})=1, and p2p_{2} acts trivially on 2\mathcal{H}_{2}.

  • For p1,p2p_{1}^{\prime},p_{2}^{\prime}, consider the following cases:

    • (a)

      p1,p2p_{1}^{\prime},p_{2}^{\prime} commute via (3)(3)-way. In this case one of p1,p2p_{1}^{\prime},p_{2}^{\prime} must act trivially on qq. W.l.o.g assume that p1p_{1}^{\prime} is the term.

    • (b)

      p1,p2p_{1}^{\prime},p_{2}^{\prime} commute via (1,2)(1,2)-way. Similarly as above notaions for p1,p2p_{1},p_{2} we have =12\mathcal{H}=\mathcal{H}^{\prime}_{1}\bigoplus\mathcal{H}^{\prime}_{2}, and define |ψ\left|\psi^{\prime}\right\rangle, i\mathcal{H}_{i}^{\prime} similarly. And similarly assume that p2p_{2}^{\prime} acts trivially on 2\mathcal{H}^{\prime}_{2}.

\square Case (a): We will prove qq is semi-separable, which leads to a contradiction. Consider the decomposition from (1,2)(1,2)-way for p1,p2p_{1},p_{2}, that is =12\mathcal{H}=\mathcal{H}_{1}\bigoplus\mathcal{H}_{2}. Since p1p_{1}^{\prime} acts trivially on qq, it keeps those subspaces invariant as well. In summary, all terms but p2p_{2}^{\prime} keep the decomposition invariant. Since the decomposition =12\mathcal{H}=\mathcal{H}_{1}\bigoplus\mathcal{H}_{2} is non-trivial, by definition qq is semi-separable.

\square Case (b): Consider term p2p_{2}, by definition, p2p_{2} is Hermitian, keeps 1,2\mathcal{H}_{1},\mathcal{H}_{2} invariant and acts trivially on 2\mathcal{H}_{2}. We can write

p2=|ψψ|A+(Iq|ψψ|)B.\displaystyle p_{2}=\left|\psi\right\rangle\left\langle\psi\right|\otimes A+(I_{q}-\left|\psi\right\rangle\left\langle\psi\right|)\otimes B. (34)

Here IqI_{q} is the identity on q\mathcal{H}^{q}. And A,BA,B (might be 0) act non-trivially at most on the remaining three qudits q1,q4,q5q_{1},q_{4},q_{5}, as in Fig. 2. Rewriting h:=AB,h^=Bh:=A-B,\hat{h}=B, we have

p2=|ψψ|h+Iqh^.\displaystyle p_{2}=\left|\psi\right\rangle\left\langle\psi\right|\otimes h+I_{q}\otimes\hat{h}. (35)

Since p2p_{2} is Hermitian, we know h,h^h,\hat{h} are Hermitian. Similarly, we can write

p2=|ψψ|h+Iqh^.\displaystyle p_{2}^{\prime}=\left|\psi^{\prime}\right\rangle\left\langle\psi^{\prime}\right|\otimes h^{\prime}+I_{q}\otimes\hat{h}^{\prime}. (36)

where h,h^h^{\prime},\hat{h}^{\prime} are Hermitian and act non-trivially at most on q1,q2,q3q_{1},q_{2},q_{3}. If h=0h=0 then p2p_{2} acts trivially on qq. Using similar argument as case (a) we conclude qq is semi-separable. The case for h=0h^{\prime}=0 is similar. So w.l.o.g, we assume that

h0,h0,\displaystyle h\neq 0,h^{\prime}\neq 0, (37)

In the following, we are going to prove that either qq or q1q_{1} is semi-separable, which will lead to a contradiction. W.l.o.g assume that both |ψ,|ψ\left|\psi\right\rangle,\left|\psi^{\prime}\right\rangle are unit vectors.

  • If |ψ=eiθ|ψ\left|\psi\right\rangle=e^{i\theta}\left|\psi^{\prime}\right\rangle for some θ\theta: then by definition, p1,p2,p1,p2p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime} keep 1,2\mathcal{H}_{1},\mathcal{H}_{2} invariant. Thus qq is separable.

  • If |ψ|ψ\left|\psi\right\rangle\perp\left|\psi^{\prime}\right\rangle: then one can verify that both p2,p2p_{2},p_{2}^{\prime} keeps 1,2\mathcal{H}_{1},\mathcal{H}_{2} invariant, since p1p_{1} also keeps 1,2\mathcal{H}_{1},\mathcal{H}_{2} invariant. By definition we know qq is semi-separable.

  • If |ψ|ψ|0,1|\langle\psi|\psi^{\prime}\rangle|\neq 0,\neq 1: notice that

    p2p2=ψ|ψ|ψψ|hh+|ψψ|hh^+|ψψ|h^h+Ih^h^,\displaystyle p_{2}p_{2}^{\prime}=\langle\psi|\psi^{\prime}\rangle\left|\psi\right\rangle\left\langle\psi^{\prime}\right|\otimes hh^{\prime}+\left|\psi\right\rangle\left\langle\psi\right|\otimes h\hat{h}^{\prime}+\left|\psi^{\prime}\right\rangle\left\langle\psi^{\prime}\right|\otimes\hat{h}h^{\prime}+I\otimes\hat{h}\hat{h}^{\prime}, (38)
    p2p2=ψ|ψ|ψψ|hh+|ψψ|hh^+|ψψ|h^h+Ih^h^.\displaystyle p_{2}^{\prime}p_{2}=\langle\psi^{\prime}|\psi\rangle\left|\psi^{\prime}\right\rangle\left\langle\psi\right|\otimes h^{\prime}h+\left|\psi^{\prime}\right\rangle\left\langle\psi^{\prime}\right|\otimes h^{\prime}\hat{h}+\left|\psi\right\rangle\left\langle\psi\right|\otimes\hat{h}^{\prime}h+I\otimes\hat{h}^{\prime}\hat{h}. (39)

    Since dim(q)=3dim(q)=3, there exists |ζ0q\left|\zeta\right\rangle\neq 0\in\mathcal{H}^{q} s.t |ζ|ψ,|ζ|ψ\left|\zeta\right\rangle\perp\left|\psi\right\rangle,\left|\zeta\right\rangle\perp\left|\psi^{\prime}\right\rangle. By p2p2=p2p2p_{2}p_{2}^{\prime}=p_{2}^{\prime}p_{2} we have ζ|p2p2|ζ=ζ|p2p2|ζ\langle\zeta|p_{2}p_{2}^{\prime}|\zeta\rangle=\langle\zeta|p_{2}^{\prime}p_{2}|\zeta\rangle, thus

    h^h^=h^h^.\displaystyle\hat{h}\hat{h}^{\prime}=\hat{h}^{\prime}\hat{h}. (40)
  • Since |ψ|ψ|1|\langle\psi|\psi^{\prime}\rangle|\neq 1, there exists |ϕ\left|\phi\right\rangle such that |ϕ|ψ,|ϕ⟂̸|ψ\left|\phi\right\rangle\perp\left|\psi\right\rangle,\left|\phi\right\rangle\not\perp\left|\psi^{\prime}\right\rangle. By Eq. (40) and ϕ|p2p2|ϕ=ϕ|p2p2|ϕ\langle\phi|p_{2}p_{2}^{\prime}|\phi\rangle=\langle\phi|p_{2}^{\prime}p_{2}|\phi\rangle we have

    h^h=hh^.\displaystyle\hat{h}h^{\prime}=h^{\prime}\hat{h}. (41)
  • Similarly to the above point, we have

    hh^=h^h.\displaystyle h\hat{h}^{\prime}=\hat{h}^{\prime}h. (42)
  • Finally by Eqs. (40) (41) (42) and p2p2=p2p2p_{2}p_{2}^{\prime}=p_{2}^{\prime}p_{2} we have

    ψ|ψ|ψψ|hh=ψ|ψ|ψψ|hh\displaystyle\langle\psi|\psi^{\prime}\rangle\left|\psi\right\rangle\left\langle\psi^{\prime}\right|\otimes hh^{\prime}=\langle\psi^{\prime}|\psi\rangle\left|\psi^{\prime}\right\rangle\left\langle\psi\right|\otimes h^{\prime}h (43)

    Left multiplying ϕ|\left\langle\phi\right| to both sides of Eq. (43), and use the fact that ψ|ψ0,ϕ|ψ=0,ϕ|ψ0\langle\psi^{\prime}|\psi\rangle\neq 0,\langle\phi|\psi\rangle=0,\langle\phi|\psi^{\prime}\rangle\neq 0, we conclude that

    hh=0.\displaystyle h^{\prime}h=0. (44)

    Similarly, we would get

    hh=0.\displaystyle hh^{\prime}=0. (45)

In summary, we showed that

  • (i)

    h,h^,h,h^h,\hat{h},h^{\prime},\hat{h}^{\prime} are Hermitian.

  • (ii)

    {h,h^}\{h,\hat{h}\} commute with {h,h^}\{h^{\prime},\hat{h}^{\prime}\},.

  • (iii)

    hh=hh=0hh^{\prime}=h^{\prime}h=0.

To ease the notations, we abbreviate the induced algebra of Hermitian term pp on qq, i.e. 𝒜pq\mathcal{A}_{p}^{\mathcal{H}^{q}}, as 𝒜pq\mathcal{A}_{p}^{q}. Intuitively the above (i)(ii)(iii)(i)(ii)(iii) say that 𝒜p2q1\mathcal{A}_{p_{2}}^{q_{1}},𝒜p2q1\mathcal{A}_{p^{\prime}_{2}}^{q_{1}} should commute with each other, and some elements are orthogonal to each other. We will show that this implies q1q_{1} is semi-separable. In the following we write down a careful proof, especially because h,h^h,\hat{h} are operators which might act non-trivially on three qudits q1,q4,q5q_{1},q_{4},q_{5} instead of only on q1q_{1}.

Let {|ij|q4,q5}ij\{|i\rangle\langle j|^{q_{4},q_{5}}\}_{ij} be the computational basis for q4,q5q_{4},q_{5}, {|kl|q2,q3}kl\{|k\rangle\langle l|^{q_{2},q_{3}}\}_{kl} for q2,q3q_{2},q_{3}. Consider the decomposition of h,h^h,\hat{h}, we rewrite p2p_{2} as

p2=|ψψ|ijhijq1|ij|q4,q5+Iqijh^ijq1|ij|q4,q5\displaystyle p_{2}=\left|\psi\right\rangle\left\langle\psi\right|\otimes\sum_{ij}h^{q_{1}}_{ij}\otimes|i\rangle\langle j|^{q_{4},q_{5}}+I_{q}\otimes\sum_{ij}\hat{h}^{q_{1}}_{ij}\otimes|i\rangle\langle j|^{q_{4},q_{5}}

where {|ij|q4,q5}ij\{|i\rangle\langle j|^{q_{4},q_{5}}\}_{ij} are linearly independent. Since {|ψψ|,Iq}\{\left|\psi\right\rangle\left\langle\psi\right|,I_{q}\} are linearly independent, we know {|ψψ||ij|q4,q5}ij{I|ij|q4,q5}ij\{\left|\psi\right\rangle\left\langle\psi\right|\otimes|i\rangle\langle j|^{q_{4},q_{5}}\}_{ij}\cup\{I\otimes|i\rangle\langle j|^{q_{4},q_{5}}\}_{ij} are linearly independent. By lemma 10, 𝒜p2q1\mathcal{A}_{p_{2}}^{q_{1}} is generated by {hijq1}ij{h^ijq1}ij\{h^{q_{1}}_{ij}\}_{ij}\cup\{\hat{h}^{q_{1}}_{ij}\}_{ij}. Define similar notations for p2p_{2}^{\prime}, we will have 𝒜p2q1\mathcal{A}_{p_{2}^{\prime}}^{q_{1}} is generated by {hklq1}kl{h^klq1}kl\{h^{\prime q_{1}}_{kl}\}_{kl}\cup\{\hat{h}^{\prime q_{1}}_{kl}\}_{kl}.

Note that {q4,q5}\{q_{4},q_{5}\} and {q2,q3}\{q_{2},q_{3}\} are disjoint, the fact that [h,h]=0[h,h^{\prime}]=0 implies the two sets {hijq1}ij\{h_{ij}^{q_{1}}\}_{ij}, {hklq1}kl\{h_{kl}^{{}^{\prime}q_{1}}\}_{kl} commute. Similarly, (ii) implies the generators of 𝒜p2q1\mathcal{A}_{p_{2}}^{q_{1}} and 𝒜p2q1\mathcal{A}_{p_{2}^{\prime}}^{q_{1}} commute, thus 𝒜p2q1\mathcal{A}_{p_{2}}^{q_{1}} and 𝒜p2q1\mathcal{A}_{p_{2}^{\prime}}^{q_{1}} commute. If we name the 44 terms on q1q_{1} as p2,p2,p3,p4p_{2}^{\prime},p_{2},p_{3},p_{4}^{\prime} as in Fig.2. Let ~\tilde{\mathcal{H}} be the Hilbert space of q1q_{1}, consider the decomposition ~=i~i\tilde{\mathcal{H}}=\bigoplus_{i}\tilde{\mathcal{H}}_{i} induced by the induced algebra of p2p_{2} on q1q_{1}, i.e. 𝒜p2q1\mathcal{A}_{p_{2}}^{q_{1}}. By Corollary 13 we know all 𝒜pq1\mathcal{A}^{q_{1}}_{p}, pp4p\neq p_{4}^{\prime} keeps the decomposition invariant. Thus all terms expect that p4p_{4}^{\prime} keeps the decomposition ~=i~i\tilde{\mathcal{H}}=\bigoplus_{i}\tilde{\mathcal{H}}_{i} invariant.

Furthermore, (iii)(iii) implies hijq1hklq1=0h^{q_{1}}_{ij}h^{\prime q_{1}}_{kl}=0, i,j,k,l\forall i,j,k,l. By Eq. (37) we assume that h0,h0h\neq 0,h^{\prime}\neq 0, thus hijq10,hklq10\exists h^{q_{1}}_{ij}\neq 0,h^{\prime q_{1}}_{kl}\neq 0. Consider Lemma 14, let 𝒜:=𝒜p2q1\mathcal{A}:=\mathcal{A}_{p_{2}}^{q_{1}}, 𝒜:=𝒜p2q1\mathcal{A}^{\prime}:=\mathcal{A}_{p^{\prime}_{2}}^{q_{1}}, we know the previous decomposition ~=i~i\tilde{\mathcal{H}}=\bigoplus_{i}\tilde{\mathcal{H}}_{i} induced by 𝒜=𝒜p2q1\mathcal{A}=\mathcal{A}_{p_{2}}^{q_{1}} is non-trivial.

Combining the implications of (i)(ii)(iii)(i)(ii)(iii), we conclude q1q_{1} is semi-separable, which leads to a contradiction. ∎

4.3 Schuch’s method and its extensions

Schuch [Sch11] proved that the qubit-CLHP-2D-projection is in NP. In this section, we illustrate that his idea can be generalized to prove that a subclass of qudit-CLHP-2D-projection is in NP, see Theorem 31. In the next section, i.e. Sec. 4.4, we will show that the qutrit-CLHP-2D-projection without semi-separable qudits falls into this subclass.

The proof for Theorem 31 is similar to [Sch11]. The main difference is that we generalize the definitions of removable qudits from Lemma 29 to Lemma 28. This generalization brings some subtlety so we write the proof in detail even though the proof proceeds in a similar way as [Sch11].

Definition 27 (Removable qudit)

For qudit-CLHP-2D-projection, consider a qudit qq and terms p1,p2,p1,p2p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime} acting on it, as shown in Fig. 2. Denote q\mathcal{H}^{q} as \mathcal{H}. Suppose there exists a decomposition =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} such that both p1,p2p_{1},p_{2} keep the decomposition invariant. Let PiP_{i} be the projection onto i\mathcal{H}_{i}. Similarly define the decomposition =jj\mathcal{H}=\bigoplus_{j}\mathcal{H}^{\prime}_{j} and PjP^{\prime}_{j} for for p1,p2p_{1}^{\prime},p_{2}^{\prime}. We say a qudit qq is removable if one of the following holds:

  • (i)

    i,j\forall i,j, at most one of p1,p2p_{1},p_{2} acts non-trivially on i\mathcal{H}_{i}, at most one of p1,p2p^{\prime}_{1},p^{\prime}_{2} acts non-trivially on j\mathcal{H}_{j}^{\prime}, and rank(PiPj)1rank(P_{i}P^{{}^{\prime}}_{j})\leq 1101010The rank is w.r.t viewing Pi,PjP_{i},P_{j}^{\prime} as local operators in (q)\mathcal{L}(\mathcal{H}^{q})..

  • (ii)

    p1,p2p_{1}^{\prime},p_{2}^{\prime} act trivially on \mathcal{H}. Besides, \mathcal{H} has a tensor-product structure as =1^2^\mathcal{H}=\hat{\mathcal{H}_{1}}\otimes\hat{\mathcal{H}_{2}}, such that p1(1^)2^p_{1}\in\mathcal{L}(\hat{\mathcal{H}_{1}})\otimes\mathcal{I}_{\hat{\mathcal{H}_{2}}}, p21^(2^)p_{2}\in\mathcal{I}_{\hat{\mathcal{H}_{1}}}\otimes\mathcal{L}(\hat{\mathcal{H}_{2}}). Or similar conditions hold when exchanging the name of p1,p2p_{1},p_{2} with p1,p2p_{1}^{\prime},p_{2}^{\prime}.

We give some examples of removable qudits.

Lemma 28

For qudit-CLHP-2D-projection, for any qudit qq, if p1,p2p_{1},p_{2} commute in (1,1,,1)(1,1,...,1)-way on q\mathcal{H}^{q}, p1,p2p_{1}^{\prime},p_{2}^{\prime} commute in (d1,,dt)(d_{1}^{\prime},...,d_{t}^{\prime})-way on q\mathcal{H}^{q} where did_{i}^{\prime} is a prime, i\forall i. Then qq is removable.

Proof.

Denote :=q\mathcal{H}:=\mathcal{H}^{q}. Let =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} be the decomposition w.r.t to p1,p2p_{1},p_{2} and (1,1,,1)(1,1,...,1)-way. Let =ii\mathcal{H}^{\prime}=\bigoplus_{i}\mathcal{H}^{\prime}_{i} be the decomposition w.r.t to p1,p2p^{\prime}_{1},p^{\prime}_{2} and (d1,,dt)(d_{1}^{\prime},...,d_{t}^{\prime})-way. PiP_{i},PjP_{j}^{\prime} are the projections onto i,j\mathcal{H}_{i},\mathcal{H}_{j}^{\prime}. By Definition 15, the way of commuting is obtained by the Structure Lemma, by Corollary 13, we know j\mathcal{H}_{j}^{\prime} has a tensor-product structure as j=j1j2\mathcal{H}_{j}^{\prime}=\mathcal{H}_{j}^{\prime 1}\otimes\mathcal{H}_{j}^{\prime 2}, where

p1\displaystyle p_{1}^{\prime} =i(j1)Ij2\displaystyle=\bigoplus_{i}\mathcal{L}(\mathcal{H}^{{}^{\prime}1}_{j})\otimes{I_{\mathcal{H}^{{}^{\prime}2}_{j}}} (46)
p2\displaystyle p_{2}^{\prime} iIj1(j2)\displaystyle\subseteq\bigoplus_{i}{I_{\mathcal{H}^{{}^{\prime}1}_{j}}}\otimes\mathcal{L}(\mathcal{H}^{{}^{\prime}2}_{j}) (47)

Since djd_{j}^{\prime} is a prime, we know that at most one of p1,p2p_{1}^{\prime},p_{2}^{\prime} acts non-trivially on j\mathcal{H}_{j}^{\prime}. Similarly, since 11 is a prime, at most one of p1,p2p_{1},p_{2} acts non-trivially on i\mathcal{H}_{i}. Furthermore, note that rank(Pi)=1rank(P_{i})=1, i\forall i. Thus rank(PiPj)min{rank(Pi),rank(Pj)}=1rank(P_{i}P_{j}^{\prime})\leq\min\{rank(P_{i}),rank(P_{j}^{\prime})\}=1. ∎

Lemma 29

For the qubit-CLHP-2D-projection, for any qubit qq, if one of {p1,p2}\{p_{1},p_{2}\} or {p1,p2}\{p_{1}^{\prime},p_{2}^{\prime}\} commute in (1,1)(1,1)-way on q\mathcal{H}^{q}, then qq is removable.

Proof.

Denote :=q\mathcal{H}:=\mathcal{H}^{q}. W.l.o.g assume p1,p2p_{1},p_{2} commute in (1,1)(1,1)-way w.r.t =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i}. If both p1,p2p_{1}^{\prime},p_{2}^{\prime} acts trivially on qq, then p1,p2p_{1}^{\prime},p_{2}^{\prime} also keep =ii\mathcal{H}=\bigoplus_{i}\mathcal{H}_{i} invariant. One can check that Definition. 27 (i) holds. If at least one of p1,p2p^{\prime}_{1},p_{2}^{\prime} acts non-trivially on \mathcal{H}, the by Lemma 17 p1,p2p^{\prime}_{1},p_{2}^{\prime} must commute on \mathcal{H} via (1,1)(1,1) or (2)(2)-way. By Lemma 28, qq is removable. ∎

We name those qudits as removable since we will “remove” them in the proof of Theorem 31. Before proving Theorem 31, we summarize [Sch11]’s result as below. Although written in terms of qubit, [Sch11]’s proof directly works for the following lemma for qudits.

Lemma 30 (One-dimensional structure [Sch11])

Consider a qudit-CLHP-2D-projection Hamiltonian H=ppH=\sum_{p}p on nn qudits. Let SS be the set of qudits where qS\forall q\in S, there exist p{p1,p2}p\in\{p_{1},p_{2}\} and p{p1,p2}p^{\prime}\in\{p_{1}^{\prime},p_{2}^{\prime}\}, such that both the induced algebra of p,pp,p^{\prime} on q\mathcal{H}^{q} are the full algebra (q)\mathcal{L}(\mathcal{H}^{q}). Then for any quantum channels {𝒩pq:(q)}p,q\{\mathcal{N}_{p}^{q}:\mathcal{L}(\mathcal{H}^{q})\rightarrow\mathbb{C}\}_{p,q}, the product p(qSC𝒩pq)[Ip]\prod_{p}\left(\otimes_{q\in S^{C}}\mathcal{N}_{p}^{q}\right)[I-p] has a one-dimensional structure thus tr(p(qSC𝒩pq)[Ip])tr\left(\prod_{p}\left(\otimes_{q\in S^{C}}\mathcal{N}_{p}^{q}\right)[I-p]\right) can be computed in classical polynomial time.

Here (qSC𝒩pq)[Ip]\left(\otimes_{q\in S^{C}}\mathcal{N}_{p}^{q}\right)[I-p] means applying qSC𝒩pq\otimes_{q\in S^{C}}\mathcal{N}_{p}^{q} on (Ip)(I-p), which can be interpreted as tracing out qudits in ScS^{c}.

Theorem 31

Consider a qudit-CLHP-2D-projection instance, if for each qudit qq, either (a) qq is removable, or (b) there exists p{p1,p2}p\in\{p_{1},p_{2}\} and p{p1,p2}p^{\prime}\in\{p_{1}^{\prime},p_{2}^{\prime}\} such that both the induced algebra of p,pp,p^{\prime} on q\mathcal{H}^{q} are the full algebra (q)\mathcal{L}(\mathcal{H}^{q}), then the corresponding qudit-CLHP-2D-projection instance is in NP.

Proof.

The proof follows the idea in  [Sch11]. The main difference is that we generalize the definitions of removable qudits from Lemma 29 to Lemma 28. This generalization brings some subtlety so we write the proof in detail even though the proof proceeds in a similar way as Schuch’s.

Imagine the 2D grid as a chess board and color the plaquettes as black and white. Denote the set of the black plaquettes as 𝒫B\mathcal{P}_{B}, the white plaquettes as 𝒫W\mathcal{P}_{W}. Use same notations as Definition 27, for any removable qudit qq, w.l.o.g assume that p1,p2p_{1},p_{2} are black, p1,p2p_{1}^{\prime},p_{2}^{\prime} are white. If qq satisfies Definition 27 (i), one can notice that

tr((Ip1)(Ip2)(Ip1)(Ip2))\displaystyle tr\left((I-p_{1})(I-p_{2})(I-p^{\prime}_{1})(I-p^{\prime}_{2})\right) (48)
=tr([iPi(Ip1)(Ip2)Pi][jPj(Ip1)(Ip2)Pj])\displaystyle=tr\left(\left[\sum_{i}P_{i}(I-p_{1})(I-p_{2})P_{i}\right]\left[\sum_{j}P^{{}^{\prime}}_{j}(I-p^{\prime}_{1})(I-p^{\prime}_{2})P^{{}^{\prime}}_{j}\right]\right) (49)
=i,jtr([Pi(Ip1)(Ip2)Pi][Pj(Ip1)(Ip2)Pj])\displaystyle=\sum_{i,j}tr\left([P_{i}(I-p_{1})(I-p_{2})P_{i}][P^{{}^{\prime}}_{j}(I-p^{\prime}_{1})(I-p^{\prime}_{2})P^{{}^{\prime}}_{j}]\right) (50)

Note that by definition of removable qudit (i), at most one of p1,p2p_{1},p_{2} acts non-trivially on i\mathcal{H}_{i}, w.o.l.g assume that it is p1p_{1}. This means Pi(Ip2)PiP_{i}(I-p_{2})P_{i} is PiP_{i} tensor some operator. Formally,

Pi(Ip2)Pi\displaystyle P_{i}(I-p_{2})P_{i} =trq(Pi(Ip2)Pi)/trq(Pi)Pi\displaystyle=tr_{q}(P_{i}(I-p_{2})P_{i})/tr_{q}(P_{i})\otimes P_{i} (51)
=trq(Pi(Ip2)Pi)/trq(Pi)IqPi\displaystyle=tr_{q}(P_{i}(I-p_{2})P_{i})/tr_{q}(P_{i})\otimes I_{q}\cdot P_{i} (52)

Similarly, we assume that p1p^{\prime}_{1} acts non-trivially on j\mathcal{H}^{\prime}_{j}. We have

tr(Pi(Ip1)(Ip2)PiPj(Ip1)(Ip2)Pj)\displaystyle tr\left(P_{i}(I-p_{1})(I-p_{2})P_{i}P^{{}^{\prime}}_{j}(I-p^{\prime}_{1})(I-p^{\prime}_{2})P^{{}^{\prime}}_{j}\right) (53)
=tr(Pi(Ip1)Pi(Ip2)PiPj(Ip1)Pj(Ip2)Pj)\displaystyle=tr\left(P_{i}(I-p_{1})P_{i}(I-p_{2})P_{i}P^{{}^{\prime}}_{j}(I-p^{\prime}_{1})P_{j}(I-p^{\prime}_{2})P^{{}^{\prime}}_{j}\right) (54)
=tr(Pi(Ip1)trq(Pi(Ip2)Pi)/trq(Pi)IqPiPj(Ip1)trq(Pj(Ip2)Pj)/trq(Pj)IqPj)\displaystyle=tr\left(P_{i}(I-p_{1})\cdot tr_{q}(P_{i}(I-p_{2})P_{i})/tr_{q}(P_{i})\otimes I_{q}\cdot P_{i}\cdot P^{{}^{\prime}}_{j}(I-p^{\prime}_{1})\cdot tr_{q}(P^{\prime}_{j}(I-p^{\prime}_{2})P^{\prime}_{j})/tr_{q}(P^{\prime}_{j})\otimes I_{q}\cdot P^{{}^{\prime}}_{j}\right) (55)

Note that Tr(MN)=Tr(NM)Tr(MN)=Tr(NM) for any M,NM,N. Further, by definition of (i), there exist un-normalized vectors |αq,|βq\left|\alpha\right\rangle_{q},\left|\beta\right\rangle_{q} on qq such that PiPj=|αqβ|qP_{i}P_{j}^{\prime}=\left|\alpha\right\rangle_{q}\left\langle\beta\right|_{q}. Then:

=tr(|βqα|q(Ip1)trq(Pi(Ip2)Pi)/trq(Pi)Iq|αqβ|q(Ip1)trq(Pj(Ip2)Pj)/trq(Pj)Iq)\displaystyle=tr\left(\left|\beta\right\rangle_{q}\left\langle\alpha\right|_{q}(I-p_{1})\cdot tr_{q}(P_{i}(I-p_{2})P_{i})/tr_{q}(P_{i})\otimes I_{q}\cdot\left|\alpha\right\rangle_{q}\left\langle\beta\right|_{q}(I-p^{\prime}_{1})\cdot tr_{q}(P^{\prime}_{j}(I-p^{\prime}_{2})P^{\prime}_{j})/tr_{q}(P^{\prime}_{j})\otimes I_{q}\right) (56)
=tr(α|q(Ip1)|αqtrq(Pi(Ip2)Pi)/trq(Pi)β|q(Ip1)|βqtrq(Pj(Ip2)Pj)/trq(Pj))\displaystyle=tr\left(\left\langle\alpha\right|_{q}(I-p_{1})\left|\alpha\right\rangle_{q}\cdot tr_{q}(P_{i}(I-p_{2})P_{i})/tr_{q}(P_{i})\cdot\left\langle\beta\right|_{q}(I-p^{\prime}_{1})\left|\beta\right\rangle_{q}\cdot tr_{q}(P^{\prime}_{j}(I-p^{\prime}_{2})P^{\prime}_{j})/tr_{q}(P^{\prime}_{j})\right) (57)

Recall that {p1,p2,p1,p2}\{p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime}\} are commuting projections. In summary, the above calculations show two things:

  • In Eq. (50), each quantity, i.e. Eq. (53), is the trace of the product of two positive semi-definite Hermitian matrices, thus each quantity is non-negative. Proving tr((Ip1)(Ip2)(Ip1)(Ip2))>0tr((I-p_{1})(I-p_{2})(I-p_{1}^{\prime})(I-p_{2}^{\prime}))>0 is equivalent to show that one of the quantitys is >0>0.

  • In Eq. (57), we somehow “project out qudit qq” for all terms. Then calculating Eq. (53) is equivalent to computing the trace of a term without qudit qq, i.e. Eq. (57). This is why qq is named removable.

  • Also note that in Eqs. (53)-(57), we do not change the relative order of (Ipi)(I-p_{i}) or (Ipi)(I-p_{i}^{\prime}). This is important when considering multiple removable qudits at the same time.

If qq satisfies Definition 27 (ii), we can also “tracing out qq”. Denote dqd_{q} as the dimension of qq. Interpret qq as two qudits q1,q2q_{1},q_{2} w.r.t 1^,2^\hat{\mathcal{H}_{1}},\hat{\mathcal{H}_{2}}, we have

tr((Ip1)(Ip2)(Ip1)(Ip2))\displaystyle tr\left((I-p_{1})(I-p_{2})(I-p^{\prime}_{1})(I-p^{\prime}_{2})\right) (58)
=tr(trq((Ip1))/dq2trq(Ip2)/dq1trq((Ip1))/dqtrq((Ip2))/dq)\displaystyle=tr\left(tr_{q}((I-p_{1}))/{d_{q_{2}}}\cdot tr_{q}(I-p_{2})/{d_{q_{1}}}\cdot tr_{q}((I-p^{\prime}_{1}))/{d_{q}}\cdot tr_{q}((I-p^{\prime}_{2}))/{d_{q}}\right) (59)

Eqs. (48)-(59) illustrate how to project out a single removable qudit. Similarly, when calculating tr(p(Ip))tr(\prod_{p}(I-p)) we can project out all the removable qudits. Specifically, we first perform Eq. (48)-Eq. (50) simultaneously for all removable qudits, and then perform Eq. (53)-(59) to project out all removable qudits. It is worth noting that we should be careful about the relative orders of each (Ip)(I-p) — To perform the calculations in Eq. (48)-Eq. (50), one requires that for each qq, terms p1,p2p_{1},p_{2} are put in the left, and p1,p2p_{1}^{\prime},p_{2}^{\prime} in the right.111111If we begin with tr(Pi1(Ip1)Pi1Pj1(Ip1)Pj1Pi2(Ip2)Pi2Pj2(Ip1)Pj2)tr\left(P_{i_{1}}(I-p_{1})P_{i_{1}}P^{\prime}_{j_{1}}(I-p_{1}^{\prime})P^{\prime}_{j_{1}}P_{i_{2}}(I-p_{2})P_{i_{2}}P^{\prime}_{j_{2}}(I-p_{1}^{\prime})P^{\prime}_{j_{2}}\right) the above method doesn’t work since this quantity might be negative. To perform such decomposition for all removable qudits simultaneously, it suffices to put all the black terms on the left and all the white terms on the right. Denote the set of removable qudits as RR, we have

tr(p(Ip))\displaystyle tr\left(\prod_{p}(I-p)\right) =tr(p𝒫B(Ip)p𝒫W(Ip))\displaystyle=tr\left(\prod_{p\in\mathcal{P}_{B}}(I-p)\prod_{p^{\prime}\in\mathcal{P}_{W}}(I-p^{\prime})\right) (60)
=iq,jq;qRtr(qRPiqq(p𝒫B(Ip))qRPiqqqRPjqq(p𝒫W(Ip))qRPjqq)\displaystyle=\sum_{i_{q},j_{q};q\in R}tr(\prod_{q\in R}P^{q}_{i_{q}}\left(\prod_{p\in\mathcal{P}_{B}}(I-p)\right)\prod_{q\in R}P^{q}_{i_{q}}\prod_{q\in R}P^{{}^{\prime}q}_{j_{q}}\left(\prod_{p\in\mathcal{P}_{W}}(I-p)\right)\prod_{q\in R}P^{{}^{\prime}q}_{j_{q}}) (61)

Then for each removable qudit qq, we perform similar operations as Eq. (53)-(59) to project out qq .

Finally for the quantity in Eq. (61), we project out all removable qudits for every (Ip)(I-p). All the remaining qudits originally correspond to type (b) in the Theorem description. By Lemma 30 we know the quantity in Eq. (61) can be computed in polynomial time.

In summary, proving tr(p(Ip))>0tr(\prod_{p}(I-p))>0 is equivalent to proving that one of the quantity in Eq. (61) is >0>0, where the quantity is tracing a product which has a one-dimensional structure, thus can be computed in classical polynomial time. Thus we conclude the qudit-CLHP-2D-projection which satisfies Theorem 31’s conditions is in NP. ∎

4.4 Qutrit-CLHP-2D is in NP

Finally, to prove that the qutrit-CLHP-2D is in NP, it suffices to notice that the qutrit-CLHP-2D-projection without semi-separable qudit satisfying conditions in Theorem 31.

Lemma 32

For any qutrit-CLHP-2D-projection instance without semi-separable qudit, every qudit satisfies one of the two conditions in Theorem 31.

Proof.

Consider any qudit qq. If dim(q)=1dim(q)=1, it is removable due to Definition 27 (i). W.l.o.g. assume dim(q){2,3}dim(q)\in\{2,3\}. If p1,p2p_{1}^{\prime},p_{2}^{\prime} act trivially on q\mathcal{H}^{q}, since qq is not semi-separable, we know p1,p2p_{1},p_{2} must commute via (3)(3)-way or (2)(2)-way. Furthermore, by the Structure Lemma, i.e. Corollary 13, we know there is a tensor product structure of q=q,1q,2\mathcal{H}^{q}=\mathcal{H}^{q,1}\otimes\mathcal{H}^{q,2} such that p1(q,1)q,2p_{1}\in\mathcal{L}(\mathcal{H}^{q,1})\otimes\mathcal{I}_{\mathcal{H}^{q,2}}, p2q,1(q,2)p_{2}\in\mathcal{I}_{\mathcal{H}^{q,1}}\otimes\mathcal{L}(\mathcal{H}^{q,2}) , thus qq is removable due to Definition 27 (ii). A similar argument works when p1,p2p_{1},p_{2} act trivially on q\mathcal{H}^{q}. Thus w.l.o.g, we assume that at least one of p1,p2p_{1},p_{2}, and one of p1,p2p_{1}^{\prime},p_{2}^{\prime} act non-trivially on qq. Then

  • When dim(q)=2dim(q)=2, by Lemma 17 we know either one of {p1,p2}\{p_{1},p_{2}\} or {p1,p2}\{p_{1}^{\prime},p_{2}^{\prime}\} commute in (1,1)(1,1)-way, or both of them commute in (2)(2)-way. For the first case, qq is removable due to Lemma 29. For the second case, qq satisfies Theorem 31 condition (b).

  • When dim(q)=3dim(q)=3, since there is no semi-separable qudit, for any qudit qq, by Lemma 26 and Lemma 16, we know either one of {p1,p2}\{p_{1},p_{2}\} or {p1,p2}\{p_{1}^{\prime},p_{2}^{\prime}\} commute in (1,1,1)(1,1,1)-way, then qq is removable by Lemma 28. Or both of them commute in (3)(3)-way, then qq satisfies Theorem 31 condition (b).

Combined with Corollary 49, Corollary 24, Lemma 32 and Theorem 31, we finally conclude that

Corollary 33

The qutrit-CLHP-2D problem is in NP.

5 Factorized commuting local Hamiltonian on 2D

In this section, we give a constructive proof to show that qudit-CLHP-2D-factorized is in NP, by proving that qudit-CLHP-2D-factorized is equivalent to a direct sum of qubit stabilizer Hamiltonian (see Theorem 46). Note that in this section we do not assume that {p}p\{p\}_{p} are projections. The reason for this is that if we start with an arbitrary qudit-CLHP-2D-factorized Hamiltonian, such as the Toric code, and replace each term with a projection that preserve the kernel (as in Lemma 48), then the new Hamiltonian is not guaranteed to be a factorized projection. For example, if we take a Toric code term such as h=XXXXh=XXXX and replace hh with (IIIIXXXX)/2(IIII-XXXX)/2, the resulting term is no longer factorized. By contrast, for the qutrit-CLHP-2D, where we do not require that the terms be factorized, the assumption that the terms are projections does not weaken the results due to Corollary 49.

The structure of this section is as follows. In Sec. 5.1 we give notations and definitions. In Sec. 5.2 we prove that if there are no separable qudits, then the Hamiltonian is equivalent to a direct sum of qubit stabilizer Hamiltonian. Finally in Sec. 5.3 we prove Theorem 46.

5.1 Notations, definitions and lemmas

Notation. Let hh be a Hermitian operator on Hilbert space \mathcal{H}. Let \mathcal{H}^{\prime} be a subspace of \mathcal{H} and suppose hh keeps \mathcal{H}^{\prime} invariant. We define ker(h):={|ψ|h|ψ=0}\ker_{\mathcal{H}}(h):=\{\left|\psi\right\rangle\in\mathcal{H}\,|\,h\left|\psi\right\rangle=0\} and ker(h):={|ψ|h|ψ=0}\ker_{\mathcal{H}^{\prime}}(h):=\{\left|\psi\right\rangle\in\mathcal{H}^{\prime}\,|\,h\left|\psi\right\rangle=0\}. For two orthogonal subspaces V,WV,W\subseteq\mathcal{H}, their direct sum are defined as VW={v+w|vV,wW}V\bigoplus W=\{v+w\,|\,v\in V,w\in W\}. For two Hermitian operators h,h()h,h^{\prime}\in\mathcal{L}(\mathcal{H}), we write hh=±hhhh^{\prime}=\pm h^{\prime}h if either hh=hhhh^{\prime}=h^{\prime}h or hh=hhhh^{\prime}=-h^{\prime}h. We say an nn-qudit Hermitian term hh is factorized if h=qhqh=\otimes_{q}h^{q} where hqh^{q} is Hermitian, q\forall q. When a factorized Hermitian h=0h=0, we always rewrite hh to be tensor of zeros, i.e. hq=0,qh^{q}=0,\forall q.

We say a Hilbert space \mathcal{H} is equivalent to mm-qubit space, denoted as (2)m\mathcal{H}\sim(\mathbb{C}^{2})^{\otimes m}, if there exists tensor-product structure =1m\mathcal{H}=\mathcal{H}_{1}\otimes...\otimes\mathcal{H}_{m} where dim(i)=2dim(\mathcal{H}_{i})=2. When considering Hilbert space (2)n,\mathcal{H}\sim(\mathbb{C}^{2})^{\otimes n}, we define Pauli groups as the group of operators generated by {I,X,Y,Z}n\{I,X,Y,Z\}^{\otimes n}, with respect the basis which makes \mathcal{H} factorized as (2)n(\mathbb{C}^{2})^{\otimes n}. We denote elements in the Pauli group as Pauli operators. First, we formally define what it means for a Hamiltonian to be equivalent to a qubit stabilizer Hamiltonian or a direct sum of qubit stabilizer Hamiltonians.

Definition 34 (Equivalence to qubit stabilizer Hamiltonian)

Consider a commuting (but not necessarily local) Hamiltonian H=ihiH=\sum_{i}h_{i} acting on space :=qq\mathcal{H}_{*}:=\otimes_{q}\mathcal{H}^{q}_{*}. We say HH is equivalent to a qubit stabilizer Hamiltonian on \mathcal{H}_{*}, if (1) For all qq, q(2)mq\mathcal{H}^{q}_{*}\sim(\mathbb{C}^{2})^{\otimes m^{q}} for some integer mqm^{q}. (2) Each hih_{i} acts as a Pauli operator up to phases on \mathcal{H}_{*}, with respect to the basis which makes (2)(qmq)\mathcal{H}_{*}\sim(\mathbb{C}^{2})^{\otimes(\sum_{q}m^{q})}.

In the above definition, we allow mq=0m_{q}=0, where dim(q)=1dim(\mathcal{H}^{q}_{*})=1, and all hih_{i} acts as II up to phases on q\mathcal{H}^{q}_{*}.

Definition 35 (Simple subspace)

Consider an n-qudit space =qq\mathcal{H}=\otimes_{q}\mathcal{H}^{q}. We say a subspace \mathcal{H}_{*} of \mathcal{H} is simple, if \mathcal{H}_{*} is a tensor product of subspace of each qudit, i.e. =qq\mathcal{H}_{*}=\otimes_{q}\mathcal{H}^{q}_{*} where q\mathcal{H}^{q}_{*} is a subspace of q\mathcal{H}^{q}.

Definition 36 (Equivalence to direct sum of qubit stabilizer Hamiltonian)

Given a commuting Hamiltonian H=ihiH=\sum_{i}h_{i} acting on space =qq\mathcal{H}=\otimes_{q}\mathcal{H}^{q}. We say HH is equivalent to a direct sum of qubit stabilizer Hamiltonians, if there exists a set of simple subspace {:=qq}P\{\mathcal{H}_{*}:=\otimes_{q}\mathcal{H}^{q}_{*}\}_{*\in P} such that (1) {}P\{\mathcal{H}_{*}\}_{*\in P} are pairwise orthogonal, and =P\mathcal{H}=\bigoplus_{*\in P}\mathcal{H}_{*}; (2) P\forall*\in P, HH keeps \mathcal{H}_{*} invariant, {hi}i\{h_{i}\}_{i} keeps \mathcal{H}_{*} invariant, and HH is equivalent to qubit stabilizer Hamiltonian when restricted to .\mathcal{H}_{*}.

Remark 37 (Terminology of “Equivalent to qubit stabilizer state” used in Theorem 2)

Use notations in Definition 36, if an nn-qudit Hamiltonian H=ihiH=\sum_{i}h_{i} is equivalent to a direct sum of qubit stabilizer Hamiltonians, there exists a simple subspace =qq\mathcal{H}_{*}=\otimes_{q}\mathcal{H}^{q}_{*} which contains a ground state of HH, denoted as |ψ\left|\psi_{*}\right\rangle. Since HH is equivalent to qubit stabilizer Hamiltonian on \mathcal{H}_{*}, we know |ψ\left|\psi_{*}\right\rangle can be chosen to be a qubit stabilizer state w.r.t to the basis which makes q(2)mq\mathcal{H}^{q}_{*}\sim(\mathbb{C}^{2})^{\otimes m_{q}} in Definition 34. In this sense we say HH has a ground state which is equivalent to qubit stabilizer state. The notion of “equivalent to qubit stabilizer state” is only used for intuitively explaining how we prove that qudit-CLHP-2D-factorized is in NP in Theorem 2. To avoid ambiguity, we will not use this notion further in the following context.

We now give the definitions and lemmas for commuting in a singular/regular way. For technical reasons, our definition is slightly different from [BV03]121212[BV03] defines the case where hh^=0h\hat{h}=0 and q,hqh^q=±h^qhq\forall q,h^{q}\hat{h}^{q}=\pm\hat{h}^{q}h^{q} as commuting in a singular way. We define this case as commuting in a regular way..

Definition 38

[Singular/regular way] Consider two factorized Hermitian terms h,h^h,\hat{h} acting on qudits, with [h,h^]=0[h,\hat{h}]=0.

  • We say h,h^h,\hat{h} commute in a regular way, if q\forall q, hqh^q=±h^qhqh^{q}\hat{h}^{q}=\pm\hat{h}^{q}h^{q}.

  • We say h,h^h,\hat{h} commute in a singular way if there \exists qudit qq, hqh^q±h^qhqh^{q}\hat{h}^{q}\neq\pm\hat{h}^{q}h^{q}.

Recall that when h=0h=0, we always rewrite hh to be tensor of zeros. Thus if h,h^h,\hat{h} commute in a singular way, then h0,h^0h\neq 0,\hat{h}\neq 0. We say a set of factorized Hermitian terms {hi}i\{h_{i}\}_{i} commute in a regular way if i,j\forall i,j, the hi,hjh_{i},h_{j} commute in a regular way. In the following, we introduce a lemma which states how factorized terms can commute with each other.

Lemma 39 (Rephrase of Lemma 9 in [BV03])

Consider two factorized Hermitian terms h,h^h,\hat{h} acting on nn qudits, with [h,h^]=0[h,\hat{h}]=0. If they only share one qudit qq, then [hq,h^q]=0[h^{q},\hat{h}^{q}]=0. If they share two qudits q1,q2q_{1},q_{2}, then one of the following holds: (1) hh^0h\hat{h}\neq 0. In this case hq1h^q1=±h^q1hq1h^{q_{1}}\hat{h}^{q_{1}}=\pm\hat{h}^{q_{1}}h^{q_{1}} and hq2h^q2=±h^q2hq2h^{q_{2}}\hat{h}^{q_{2}}=\pm\hat{h}^{q_{2}}h^{q_{2}}. (2) hh^=0h\hat{h}=0. In this case one of hq1h^q1h^{q_{1}}\hat{h}^{q_{1}}, or hq2h^q2h^{q_{2}}\hat{h}^{q_{2}} equals to 0.

Corollary 40

Consider two factorized Hermitian terms h,h^h,\hat{h} acting on nn qudits, with [h,h^]=0[h,\hat{h}]=0. If h,h^h,\hat{h} share two qudits q1,q2q_{1},q_{2}, and commute in a singular way, then one of hq1h^q1h^{q_{1}}\hat{h}^{q_{1}}, or hq2h^q2h^{q_{2}}\hat{h}^{q_{2}} equals to 0. For the other one qudit, denoted as qq, we have hqh^q±h^qhqh^{q}\hat{h}^{q}\neq\pm\hat{h}^{q}h^{q}.

We also prove some useful lemmas.

Lemma 41

Consider two Hermitian terms h,h^h,\hat{h} acting on a Hilbert space \mathcal{H}. If hh^=αh^hh\hat{h}=\alpha\hat{h}h for some α\alpha\in\mathbb{R}, then hh keeps ker(h^)\ker_{\mathcal{H}}(\hat{h}) invariant.

Proof.

It suffices to notice that |ψker(h^)\forall\left|\psi\right\rangle\in\ker_{\mathcal{H}}(\hat{h}), we have h^h(|ψ)=αhh^|ψ=0\hat{h}h(\left|\psi\right\rangle)=\alpha h\hat{h}\left|\psi\right\rangle=0. This implies that h|ψker(h^)h\left|\psi\right\rangle\in\ker_{\mathcal{H}}(\hat{h}). ∎

Lemma 42

Consider a qudit qq and Hermitian terms h^q0\hat{h}^{q}\neq 0 and {hiq}i\{h_{i}^{q}\}_{i} acting on q\mathcal{H}^{q}. Suppose that i\forall i, h^qhiq=±hiqh^q\hat{h}^{q}h_{i}^{q}=\pm h_{i}^{q}\hat{h}^{q}, and furthermore there exists i0i_{0} such that hi0q0h_{i_{0}}^{q}\neq 0, and h^qhi0q=0.\hat{h}^{q}h_{i_{0}}^{q}=0. Then qq is separable with respect to {h^q}{hiq}i\{\hat{h}^{q}\}\cup\{h_{i}^{q}\}_{i}.

Proof.

Define 1:=ker(h^)\mathcal{H}_{1}:=\ker(\hat{h}). Since both h^,hi0\hat{h},h_{i_{0}} are non-zero and h^hi0=0\hat{h}h_{i_{0}}=0, we know 1{0}\mathcal{H}_{1}\neq\{0\} and 1q.\mathcal{H}_{1}\neq\mathcal{H}^{q}. This implies the decomposition q=11\mathcal{H}^{q}=\mathcal{H}_{1}\bigoplus\mathcal{H}_{1}^{\perp} is non-trivial. By lemma 41, we know that all operators in {h^}{hi}i\{\hat{h}\}\cup\{h_{i}\}_{i} keep 1\mathcal{H}_{1} invariant. Since they are Hermitian, they also keep 1\mathcal{H}_{1}^{\perp} invariant. Thus qq is separable. ∎

5.2 A weaker version without separable qudits

In this section we prove in Theorem 45 that if there are no separable qudits, then qudit-CLHP-2D-factorized is equivalent to qubit stabilizer Hamiltonian. In particular, we prove that in this situation, all the terms must commute in a regular way.

Lemma 43

Consider qudit-CLHP-2D-factorized Hamiltonian H=ppH=\sum_{p}p acting on nn qudits. If there are no separable qudits, then all the terms {p}p\{p\}_{p} commute in a regular way. Moreover, for any qudit qq and any term p,p^{p1,p2,p1,p2}p,\hat{p}\in\{p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime}\} acting on qq, as shown in Fig. 3 (a), we have (1) (pq)2=cpqIq(p^{q})^{2}=c_{pq}I_{q} for some constant cpqc_{pq}. (2) pqp^q=±p^qpqp^{q}\hat{p}^{q}=\pm\hat{p}^{q}p^{q}. (3) The CC^{*}-algebra generated by {pq}p{p1,p2,p1,p2}\{p^{q}\}_{p\in\{p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime}\}} is the whole algebra (q)\mathcal{L}(\mathcal{H}^{q}).

Proof.

We first define some notations to help the illustration. To match the notations in Definition 34, we denote the Hilbert space of qudit qq as q\mathcal{H}_{*}^{q}, the nn-qudit space as :=qq\mathcal{H}_{*}:=\otimes_{q}\mathcal{H}_{*}^{q}. As shown in Fig. 3 (a), Consider a qudit qq, we denote the terms acting on qq as p1,p2,p1,p2p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime}. Recall that all the terms are factorized, thus we can use p1qp_{1}^{q} to represent the factor of p1p_{1} on qq. Use similar notations for other terms. For any two terms h,h^{p1,p2,p1,p2}h,\hat{h}\in\{p_{1},p_{2},p^{\prime}_{1},p^{\prime}_{2}\}, we use symbols ,0¯,×-,\underline{0},\times to represent the relationship between hq,h^qh^{q},\hat{h}^{q}.

  • -”: If hqh^q=±h^qhqh^{q}\hat{h}^{q}=\pm\hat{h}^{q}h^{q}, and h,h^h,\hat{h} commute in a regular way.

  • 0¯\underline{0}”: If hqh^q=0h^{q}\hat{h}^{q}=0, and h,h^h,\hat{h} commute in a singular way.

  • ×\times”: If hqh^q±h^qhqh^{q}\hat{h}^{q}\neq\pm\hat{h}^{q}h^{q}, and h,h^h,\hat{h} commute in a singular way.

Using the above symbols, we will draw a graph to represent the relationship of {p1,p2,p1,p2}\{p_{1},p_{2},p^{\prime}_{1},p^{\prime}_{2}\} on qudit qq. For example, the graph in Fig. 3 (f) means: The relationship between p1qp_{1}^{q} and p2qp_{2}^{q} is “-”. That is p1qp2q=±p2qp1qp_{1}^{q}p_{2}^{q}=\pm p_{2}^{q}p_{1}^{q}, and p1,p2p_{1},p_{2} commute in a regular way. Similar for p1qp^{{}^{\prime}q}_{1} and p2qp^{{}^{\prime}q}_{2}. The relationship between p1qp^{q}_{1} and p1qp^{{}^{\prime}q}_{1} is “0¯\underline{0}”. That is p1qp1q=0p_{1}^{q}p_{1}^{{}^{\prime}q}=0, p1,p1p_{1},p_{1}^{\prime} commute in a singular way. Similar for p2qp^{{}^{\prime}q}_{2} and p2qp^{q}_{2}. The relationship for p1q,p2qp_{1}^{q},p_{2}^{{}^{\prime}q} is “×\times”. That is p1qp2q±p2qp1qp_{1}^{q}p_{2}^{{}^{\prime}q}\neq\pm p_{2}^{{}^{\prime}q}p_{1}^{q}, and p1,p2p_{1},p_{2}^{\prime} commute in a singular way. Similar for p1qp_{1}^{{}^{\prime}q} and p2qp_{2}^{q}. We use #q0¯\#_{q}\underline{0} to represent the number of 0¯\underline{0} in the graph for qq, and similar for symbol “×\times”. For example, in Fig. 3 (f), we have #q0¯=#q×=2.\#_{q}\underline{0}=\#_{q}\times=2.

For each qq, we draw such a graph and assign ,0¯,×-,\underline{0},\times to each pair h,h^{p1q,p2q,p1q,p2q}h,\hat{h}\in\{p_{1}^{q},p_{2}^{q},p_{1}^{{}^{\prime}q},p_{2}^{{}^{\prime}q}\}. For qq in the boundary131313Here we mean the physical boundary of the lattice, not the (virtual) boundary defined in [AKV18]. of the lattice, some terms might be missing. We draw the graph for the existing terms similarly. We use #0¯\#\underline{0} to represent the total number of 0¯\underline{0} when considering all qudits, that is #0¯=q#q0¯\#\underline{0}=\sum_{q}\#_{q}\underline{0}. Similarly for #×\#\times. Note that by Corollary 40, we have #0¯=#×\#\underline{0}=\#\times. Now we are prepared to prove Lemma 43. We first prove:

Claim 1: There is no qudit qq such that two of the terms acting on qq, i.e. two of p1,p2,p1,p2p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime} in Fig. 3(a), can commute in a singular way. With contraction suppose there are such qudits, then #0¯=#×1\#\underline{0}=\#\times\geq 1. Then there must be a qudit qq such that of #q0¯#q×1\#_{q}\underline{0}\geq\#_{q}\times\geq 1. We will prove that qq is separable thus leading to a contradiction. Here we suppose qq is in not on the boundary, thus there are four terms p1,p2,p1,p2p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime} acting on qq. The case where qq is on the boundary and some terms are missing can be analyzed similarly.

Note that we always have p1qp2q=p2qp1qp_{1}^{q}p_{2}^{q}=p_{2}^{q}p_{1}^{q} since p1,p2p_{1},p_{2} only shares one qudit. Similar for p1qp_{1}^{{}^{\prime}q} and p2qp_{2}^{{}^{\prime}q}. One can check that up to rotating the lattice, the relationships of terms on qq must be related to one of the graphs in Fig. 3 (b-f). Please read the captions of Fig. 3 for a precise description of the classification. Note that if two terms h,h^h,\hat{h} on qq satisfies 0¯\underline{0}, by definition of commuting in a singular way, we have p0,p^0p\neq 0,\hat{p}\neq 0, thus pq0,p^q0p^{q}\neq 0,\hat{p}^{q}\neq 0.

For Fig. 3 (b-e), let h^q:=p1q\hat{h}^{q}:=p_{1}^{q}, {hiq}i:={p2q,p1q,p2q}\{h_{i}^{q}\}_{i}:=\{p_{2}^{q},p_{1}^{{}^{\prime}q},p_{2}^{{}^{\prime}q}\}. By Lemma 42, we know qq is separable. For Fig. 3 (f), let 𝒜\mathcal{A} to be the CC^{*}-algebra generated by p1q,p2qp_{1}^{q},p_{2}^{{}^{\prime}q}, 𝒜\mathcal{A}^{\prime} to be the CC^{*}-algebra generated by p1q,p2qp_{1}^{{}^{\prime}q},p_{2}^{q}. By Lemma 14 we know qq is separable.

qqp1p_{1}^{\prime}p1p_{1}p2p_{2}p2p_{2}^{\prime}
(a)
0
(b)
0×\times
(c)
0×\times
(d)
00×\times×\times
(e)
00×\times×\timesp1p_{1}^{\prime}p1p_{1}p2p_{2}p2p_{2}^{\prime}
(f)
Figure 3: Relationships of factors on qq. The classification is organized in the increasing order of #q×\#_{q}\times. (b) Case where #q×=0,#q0¯=1\#_{q}\times=0,\#_{q}\underline{0}=1. The cases where #q×=0,#q0¯1\#_{q}\times=0,\#_{q}\underline{0}\geq 1 can be handled in the same way. (c)(d) Case when #q×=1,#q0¯=1.\#_{q}\times=1,\#_{q}\underline{0}=1. Cases when #q×=1,#q0¯1\#_{q}\times=1,\#_{q}\underline{0}\geq 1 can be handled in the same way. (e)(f) Cases when #q×=2,#q0¯=2.\#_{q}\times=2,\#_{q}\underline{0}=2.

Claim 2: The conditions (1)(2)(3) in Lemma 43 hold. Consider arbitrary qudit qq, let 𝒜\mathcal{A} be the CC^{*}-algebra generated by {pq}p{p1,p2,p1,p2}\{p^{q}\}_{p\in\{p_{1},p_{2},p_{1}^{\prime},p_{2}^{\prime}\}}. Since qq is not separable, by lemma 11, we know the the center of 𝒜\mathcal{A}, i.e.𝒵(𝒜)\mathcal{Z}(\mathcal{A}), must be trivial. Besides, lemma 11 also implies q=12\mathcal{H}^{q}=\mathcal{H}_{1}\otimes\mathcal{H}_{2}, 𝒜=(1)2\mathcal{A}=\mathcal{L}(\mathcal{H}_{1})\otimes\mathcal{I}_{\mathcal{H}_{2}}. Since qq is not separable, we must further have 2\mathcal{H}_{2} is of dimension 11, thus 𝒜=(q)\mathcal{A}=\mathcal{L}(\mathcal{H}^{q}) (condition (3)). Otherwise qq is again separable by considering q=i1|ϕiϕi|\mathcal{H}^{q}=\bigoplus_{i}\mathcal{H}_{1}\otimes\left|\phi_{i}\right\rangle\left\langle\phi_{i}\right| where {ϕi}\{\phi_{i}\} is a basis of 2\mathcal{H}_{2}.

Since Claim 1 is true, all terms are commuting in a regular way. This means all factors are either anti-commuting or commuting (condition (2)), thus \forall term pp, (pq)2(p^{q})^{2} commute with every factor, i.e. (pq)2𝒵(𝒜)(p^{q})^{2}\in\mathcal{Z}(\mathcal{A}). Since we already argued that 𝒵(𝒜)\mathcal{Z}(\mathcal{A}) must be trivial, we have (pq)2=cpqI(p^{q})^{2}=c_{pq}I for some constant cpqc_{pq} (condition (1)). ∎

[BV03] showed that for qudit-factorized-CHP141414Recall that CHP represents commuting Hamiltonian problem where the Hamiltonian might not be local., if all terms commute in a regular way, then one can transform the Hamiltonian into a qubit stabilizer Hamiltonian. Especially, they proved the following lemma.

Lemma 44 (Lemma 12 of [BV03])

Let q\mathcal{H}^{q}_{*} be a Hilbert space, G1,,Gr(q)G_{1},...,G_{r}\in\mathcal{L}(\mathcal{H}^{q}_{*}) be Hermitian operators such that

Ga2=I,GaGb=±GbGa,a,b{1,,r}.G_{a}^{2}=I,G_{a}G_{b}=\pm G_{b}G_{a},\forall a,b\in\{1,...,r\}.

and such that the algebra generated by G1,,GrG_{1},...,G_{r} coincides with (q)\mathcal{L}(\mathcal{H}^{q}_{*}). Then there exists an integer nn, a tensor product structure q=(2)n\mathcal{H}^{q}_{*}=(\mathbb{C}^{2})^{\otimes n} and a unitary operator Uq(q)U^{q}\in\mathcal{L}(\mathcal{H}^{q}_{*}) such that UqGaUqU^{q}G_{a}U^{q\dagger} is a tensor of Pauli operators and the identity (up to sign) for all a{1,,r}a\in\{1,...,r\}. Here nn may be equal to 0.

Finally, we are prepared to prove the main theorem in this section.

Theorem 45

Consider a qudit-CLHP-2D-factorized Hamiltonian H=ppH=\sum_{p}p on nn-qudit space. If there are no separable qudits, then HH is equivalent to qubit stabilizer Hamiltonian on \mathcal{H}.

Proof.

To match the notations in Definition 34, we denote the Hilbert space of qudit qq as q\mathcal{H}_{*}^{q}, the nn-qudit space as :=qq\mathcal{H}_{*}:=\otimes_{q}\mathcal{H}_{*}^{q}. Consider a qudit-CLHP-2D-factorized H=ppH=\sum_{p}p. W.o.l.g assume that all pp are non-zero. Note that since pp is Hermitian, if p2=0p^{2}=0, then p=0p=0. Consider arbitrary qudit qq, use notations in Lemma 43 we know cpq0,p,qc_{pq}\neq 0,\forall p,q. For every pp, define p~\tilde{p} be a normalized version of pp, that is p~=qp~q\tilde{p}=\otimes_{q}\tilde{p}^{q} where p~q=pq/cpq\tilde{p}^{q}=p^{q}/c_{pq}. For any qudit qq, view ,p~q,...,\tilde{p}^{q},... as ,Ga,...,G_{a},.... By Lemma 43, one can check that {p~q}p\{\tilde{p}^{q}\}_{p} satisfies the condition of Lemma 44. Thus by choosing appropriate basis of each qudit, q\mathcal{H}^{q}_{*} will have a factorized structure as (2)mq({\mathbb{C}^{2}})^{\otimes m^{q}} for integer mqm^{q}, and {p~q}\{\tilde{p}^{q}\} are tensor of Pauli operators up to sign. Thus pqp^{q} acts as a Pauli operator on this basis, up to phases. Thus HH is equivalent to qubit stabilizer Hamiltonian HH by definition.

5.3 The full version

Finally, we remove the constraints of no separable qudits in Theorem 45 and prove the following.

Theorem 46

Any qudit-CLHP-2D-factorized Hamiltonian H=ppH=\sum_{p}p is equivalent to a direct sum of qubit stabilizer Hamiltonian.

Proof.

Denote the space that HH acting on as =qq\mathcal{H}=\bigotimes_{q}\mathcal{H}^{q}. Recall that if a qudit qq is separable with respect to decomposition q=iiq\mathcal{H}^{q}=\bigoplus_{i}\mathcal{H}_{i}^{q}, for any chosen index ii we can restrict all terms on this subspace, and get a new instance of qudit-CLHP-2D-factorized. If we repetitively perform this restriction whenever this is a separable qudit, after polytime we will reach the case with no separable qudits.

To prove theorem 46, it suffices to imagine the restricting process as a decision tree. Specifically, we write down a root node and define the space of the root node as \mathcal{H}, and repeat the following process: Transverse all the leaf nodes. Denote the leaf node considered currently as *, and its space as \mathcal{H}_{*}. If HH restricting on \mathcal{H}_{*} has separable qudits, choose an arbitrary such separable qudit. Denote this qudit as qq and the corresponding decomposition as q=iiq\mathcal{H}^{q}=\bigoplus_{i}\mathcal{H}^{q}_{i}. For every ii, we build a child node wiw_{i} to *, and define the space of wiw_{i} as restricting q\mathcal{H}^{q} to iq\mathcal{H}_{i}^{q} in \mathcal{H}_{*}. We repeat this process until for every leaf node, there are no separable qudits. In the final tree, every leaf node * corresponds to a simple subspace \mathcal{H}_{*}. By the definition of the tree, we know {}\{\mathcal{H}_{*}\}_{*} are orthogonal to each other and =\mathcal{H}=\bigoplus_{*}\mathcal{H}_{*}. By Theorem 45, \mathcal{H} is equivalent to qubit stabilizer Hamiltonian on \mathcal{H}_{*}. Thus we prove that HH is equivalent to a direct sum of qubit stabilizer Hamiltonian. ∎

Corollary 47

Qudit-CLHP-2D-factorized is in NP.

Proof.

Consider a qudit-CLHP-2D-factorized problem with Hamiltonian H=ppH=\sum_{p}p and parameters a,ba,b. By Theorem 46 we know there exists a *, where =qq\mathcal{H}_{*}=\bigotimes_{q}\mathcal{H}^{q}_{*} is a simple subspace, such that there is a ground state lies in \mathcal{H}_{*}, and HH is equivalent to qubit stabilizer Hamiltonian on \mathcal{H}_{*}. The NP prover is supposed to provide the subspace {q}q\{\mathcal{H}^{q}_{*}\}_{q}, and provide the qudit unitary UqU^{q} in Theorem 44 for each qq. Using that information, the verifier firstly checks that all terms {p}p\{p\}_{p} keep the subspaces {q}q\{\mathcal{H}^{q}_{*}\}_{q} invariant. Then the verifier uses polynomial time to transform q\mathcal{H}^{q}_{*} to be tensor of qubit space, i.e.(2)mq(\mathbb{C}^{2})^{\otimes m_{q}}, and transform HH on \mathcal{H}_{*} to be a summation of Pauli operators up to phases, denoted as H|=hahhH|_{*}=\sum_{h}a_{h}h. Here hh is a Pauli operator.

Then the verifier is going to verify λ(H|)a\lambda(H|_{*})\leq a. Since {ahh}h\{a_{h}h\}_{h} are commuting, there is a ground state which is the common eigenstate of every hh, denote the corresponding eigenvalue as λh\lambda_{h}. The prover is supposed to provide such {λh}h\{\lambda_{h}\}_{h}. The verifier verifies that hahλha\sum_{h}a_{h}\lambda_{h}\leq a, and verifies there is a state which is the common 1-eigenstate of commuting Pauli operators {h/λh}\{h/\lambda_{h}\}. The common 1-eigenstate verification can be done in polynomial time by standard stabilizer formalism. Note that although we describe the prover in an interactive way, he can in fact send all the witnesses at the same time. ∎

Appendix A Acknowledgement

Thanks Thomas Vidick and Daniel Ranard for helpful discussion.

Appendix B Relationship between general case and projection case

Lemma 48

If qudit-CLHP-2D-projection is in NP, then qudit-CLHP-2D is in NP.

Proof.

Consider a qudit-CLHP-2D(H,a,b)(H,a,b), where H=ppH=\sum_{p}p, a,ba,b\in\mathbb{R} and ba1/poly(n)b-a\geq 1/poly(n). Denote the ground energy λ:=λ(H)\lambda:=\lambda(H). Since {p}p\{p\}_{p} are commuting with each other, there exists a ground state |ψ\left|\psi\right\rangle and {λp}p\{\lambda_{p}\in\mathbb{R}\}_{p} such that p|ψ=λp|ψp\left|\psi\right\rangle=\lambda_{p}\left|\psi\right\rangle,p\forall p and pλp=λ\sum_{p}\lambda_{p}=\lambda. Let Πp\Pi_{p} be the projection onto the λp\lambda_{p}-eigenspace of pp. Let p^=IΠp\hat{p}=I-\Pi_{p}, H^=pp^\hat{H}=\sum_{p}\hat{p}. Since {p}p\{p\}_{p} are commuting, we know that {p^}p\{\hat{p}\}_{p} are also commuting.

The NP prover is supposed to list such {λp}p\{\lambda_{p}\}_{p}, then the verifier can check pλp<a\sum_{p}\lambda_{p}<a and compute p^\hat{p} and H^\hat{H}. Then the prover is supposed to prove qudit-CLHP-2D-projection(H^)(\hat{H}) is a Yes instance — that is proving there exists |ψ\left|\psi\right\rangle such that p^|ψ=0,p^\hat{p}\left|\psi\right\rangle=0,\forall\hat{p}. Thus if qudit-CLHP-2D-projection is in NP, then qudit-CLHP-2D is in NP. ∎

Corollary 49

If the qutrit-CLHP-2D-projection is in NP, then the qutrit-CLHP-2D is in NP.

Appendix C Qudits on the vertexs or on the edges

In this manuscript, we consider commuting local Hamiltonian on a 2D square lattice, where qudits are on the vertices and Hermitian terms are on the plaquettes. There is another setting that put qudits on the edges, and Hermitian terms on ”plaquettes” and ”stars”. Specifically, as shown in Fig. 4 for each plaquette pp, there is a Hermitian term BpB_{p} acting on the qudits on its edges, i.e. q1,q2,q3,q4q_{1},q_{2},q_{3},q_{4}. For each vertex vv, consider the star consisting of vv and edges adjacent to vv, there is a Hermitian term AvA_{v} acting on qudits on its edges, i.e. q3,q4,q5,q6q_{3},q_{4},q_{5},q_{6}. The Hamiltonian is H=pBp+vAvH=\sum_{p}B_{p}+\sum_{v}A_{v}. We abbreviate this setting as “qudits on 2D edges” and the setting in Sec. 2 as “qudits on 2D vertices”. In the following we will show that the two settings are equivalent.

q1q_{1}q2q_{2}q3q_{3}q4q_{4}q5q_{5}q6q_{6}ppvv
Figure 4: Qudits on edges to qudits on vertices

(1) “qudits on 2D edges” \Rightarrow “qudits on 2D vertices”. Begin from “qudits on 2D edges”, as shown in Fig. 5, the qudits on the edges can in fact be viewed as qudits on the vertices of another 2D square lattice defined by the dash lines. The terms BpB_{p} and AsA_{s} will correspond to plaquette terms in the dashed lattice. Thus our techniques directly apply to the setting for qudits on the edges.

Figure 5: Qudits on 2D edges to qudits on 2D vertices

(1) “qudits on 2D vertices” \Rightarrow “qudits on 2D edges”. Begin from “qudits on 2D vertices”, as shown in Fig. 6, the qudits on the vertices can in fact be viewed as qudits on the edges of another 2D square lattice defined by the dash lines. The plaquette terms will correspond to plaquette and star terms in the dashed lattice.

Figure 6: Qudits on 2D edges to qudits on 2D vertices

References

  • [AAG22] Anurag Anshu, Itai Arad, and David Gosset. An area law for 2d frustration-free spin systems. In Proceedings of the 54th Annual ACM SIGACT Symposium on Theory of Computing. ACM, jun 2022.
  • [ABN22] Anurag Anshu, Nikolas Breuckmann, and Chinmay Nirkhe. Nlts hamiltonians from good quantum codes. arXiv preprint arXiv:2206.13228, 2022.
  • [AE11a] Dorit Aharonov and Lior Eldar. On the complexity of commuting local hamiltonians, and tight conditions for topological order in such systems. In 2011 IEEE 52nd Annual Symposium on Foundations of Computer Science, pages 334–343, 2011.
  • [AE11b] Dorit Aharonov and Lior Eldar. On the complexity of commuting local hamiltonians, and tight conditions for topological order in such systems. In 2011 IEEE 52nd Annual Symposium on Foundations of Computer Science, pages 334–343. IEEE, 2011.
  • [AE13] Dorit Aharonov and Lior Eldar. The commuting local hamiltonian on locally-expanding graphs is in np. arXiv preprint arXiv:1311.7378, 2013.
  • [AKV18] Dorit Aharonov, Oded Kenneth, and Itamar Vigdorovich. On the complexity of two dimensional commuting local hamiltonians. arXiv preprint arXiv:1803.02213, 2018.
  • [BCG+23] Ivan Bardet, Ángela Capel, Li Gao, Angelo Lucia, David Pérez-García, and Cambyse Rouzé. Rapid thermalization of spin chain commuting hamiltonians. Physical Review Letters, 130(6):060401, 2023.
  • [BV03] Sergey Bravyi and Mikhail Vyalyi. Commutative version of the k-local hamiltonian problem and common eigenspace problem. arXiv preprint quant-ph/0308021, 2003.
  • [CRF20] Ángela Capel, Cambyse Rouzé, and Daniel Stilck França. The modified logarithmic sobolev inequality for quantum spin systems: classical and commuting nearest neighbour interactions. arXiv preprint arXiv:2009.11817, 2020.
  • [FS97] Gerald B Folland and Alladi Sitaram. The uncertainty principle: a mathematical survey. Journal of Fourier analysis and applications, 3:207–238, 1997.
  • [GHL+15] Sevag Gharibian, Yichen Huang, Zeph Landau, Seung Woo Shin, et al. Quantum hamiltonian complexity. Foundations and Trends® in Theoretical Computer Science, 10(3):159–282, 2015.
  • [GMV17] David Gosset, Jenish C Mehta, and Thomas Vidick. Qcma hardness of ground space connectivity for commuting hamiltonians. Quantum, 1:16, 2017.
  • [Got97] Daniel Gottesman. Stabilizer codes and quantum error correction. California Institute of Technology, 1997.
  • [Has12a] Matthew B Hastings. Matrix product operators and central elements: Classical description of a quantum state. Geometry & Topology Monographs, 18(115-160):276, 2012.
  • [Has12b] Matthew B Hastings. Trivial low energy states for commuting hamiltonians, and the quantum pcp conjecture. arXiv preprint arXiv:1201.3387, 2012.
  • [Has13] Matthew B. Hastings. Trivial low energy states for commuting hamiltonians, and the quantum pcp conjecture. Quantum Info. Comput., 13(5–6):393–429, may 2013.
  • [KB16] Michael J Kastoryano and Fernando GSL Brandao. Quantum gibbs samplers: The commuting case. Communications in Mathematical Physics, 344:915–957, 2016.
  • [Kit03a] A Yu Kitaev. Fault-tolerant quantum computation by anyons. Annals of physics, 303(1):2–30, 2003.
  • [Kit03b] A.Yu. Kitaev. Fault-tolerant quantum computation by anyons. Annals of Physics, 303(1):2–30, jan 2003.
  • [KKR06] Julia Kempe, Alexei Kitaev, and Oded Regev. The complexity of the local hamiltonian problem. Siam journal on computing, 35(5):1070–1097, 2006.
  • [Sch11] Norbert Schuch. Complexity of commuting hamiltonians on a square lattice of qubits. arXiv preprint arXiv:1105.2843, 2011.
  • [T+03] Masamichi Takesaki et al. Theory of operator algebras II, volume 125. Springer, 2003.