Abstract 1 Introduction 2 Preliminaries 3 Our Algorithms 4 Analysis 5 Discussion References

Improved Algorithms for Quantum MaxCut via Partially Entangled Matchings

Anuj Apte ORCID University of Chicago, IL, USA Eunou Lee Korea Institute for Advanced Study, Seoul, South Korea Kunal Marwaha ORCID University of Chicago, IL, USA Ojas Parekh ORCID Sandia National Laboratories, Albuquerque, NM, USA James Sud ORCID University of Chicago, IL, USA
Abstract

We introduce a 0.611-approximation algorithm for Quantum MaxCut and a 1+540.809-approximation algorithm for the EPR Hamiltonian of [12]. A novel ingredient in both of these algorithms is to partially entangle pairs of qubits associated to edges in a matching, while preserving the direction of their single-qubit Bloch vectors. This allows us to interpolate between product states and matching-based states with a tunable parameter.

Keywords and phrases:
Quantum computing, Quantum MaxCut, Maximum matching
Funding:
Eunou Lee: E.L is supported by a KIAS Individual Grant CG093801 at Korea Institute for Advanced Study.
Kunal Marwaha: K.M acknowledges support from AFOSR (FA9550-21-1-0008).
Ojas Parekh: O.P. acknowledges that this material is based upon work supported by the U.S. Department of Energy, Office of Science, Accelerated Research in Quantum Computing, Fundamental Algorithmic Research toward Quantum Utility (FAR-Qu).
James Sud: J.S acknowledges that this work is funded in part by the STAQ project under award NSF Phy-232580; in part by the US Department of Energy Office of Advanced Scientific Computing Research, Accelerated Research for Quantum Computing Program.
Copyright and License:
[Uncaptioned image] © Anuj Apte, Eunou Lee, Kunal Marwaha, Ojas Parekh, and James Sud; licensed under Creative Commons License CC-BY 4.0
2012 ACM Subject Classification:
Theory of computation Approximation algorithms analysis
Related Version:
Full Version: https://arxiv.org/pdf/2504.15276 [2]
Acknowledgements:
This work is supported by a collaboration between the US DOE and other Agencies. The work of A.A is supported by the Data Science Institute at the University of Chicago. This article has been authored by an employee of National Technology & Engineering Solutions of Sandia, LLC under Contract No. DE-NA0003525 with the U.S. Department of Energy (DOE). The employee owns all right, title and interest in and to the article and is solely responsible for its contents. The United States Government retains and the publisher, by accepting the article for publication, acknowledges that the United States Government retains a non-exclusive, paid-up, irrevocable, world-wide license to publish or reproduce the published form of this article or allow others to do so, for United States Government purposes. The DOE will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan https://www.energy.gov/downloads/doe-public-access-plan.
Funding:
K.M and J.S. acknowledge that this material is based upon work supported by the National Science Foundation Graduate Research Fellowship under Grant No. 2140001.
Editors:
Anne Benoit, Haim Kaplan, Sebastian Wild, and Grzegorz Herman

1 Introduction

Given a graph G(V,E,w) with positive edge weights w:E+ and a 2-local Hamiltonian term h, define the n-qubit Hamiltonian

HG=def(i,j)E(G)wijhij,

where hij is the local term h applied on qubits (i,j). For two particular local terms h, we are interested in the problem of computing the maximum energy of HG, which we denote as λmax(HG), for any G. This is not an easy task in general: deciding if λmax(HG) is above some threshold with inverse polynomial accuracy is known to be 𝖰𝖬𝖠-hard [18].

In the first problem, we choose the local term

hijQMC=def12(IiIjXiXjYiYjZiZj)=2|ψijψ|ij,

where |ψ=12(|01|10) is the singlet state. This problem has recently been studied under the name Quantum MaxCut (QMC) [7]. In the statistical mechanics literature, Hamiltonians HG defined by hQMC are instances of the zero-field quantum Heisenberg XXX1/2 model. The decision version of QMC is 𝖰𝖬𝖠-hard [18].

In the second problem, we choose the local term

hijEPR=def12(IiIj+XiXjYiYj+ZiZj)=2|ϕ+ijϕ+|ij,

where |ϕ+=12(|00+|11). This problem, named EPR by [12], is thought to be easier than QMC, since Hamiltonians HG defined by hEPR are stoquastic (sign-problem free) [18, 5]. In fact, it is not yet clear if EPR can be solved in polynomial time.

In lieu of exactly computing λmax(HG), we may try to approximate this value. In both problems, the local term h is positive semidefinite, so λmax(HG)0. We judge an approximation by its approximation ratio. Suppose we can find efficiently computable functions ,u such that for all graphs G,

0(G)λmax(HG)u(G).

Then, the approximation ratio α is at least

αminG(G)λmax(HG)minG(G)u(G).

In most works, the upper bound u is determined by solving a semidefinite programming (SDP) relaxation of the maximization problem [3, 7, 16, 19, 20, 9]. The source of such SDP relaxations are generally hierarchies of SDPs that provide increasingly better upper bounds u at the cost of solving larger-sized SDPs. The quantum moment-SOS hierarchy, typically based on Pauli operators, is widely employed for quantum local Hamiltonian problems. This hierarchy is an instance of the NPA hierarchy [15] and is also known as the quantum Lasserre hierarchy [16]. The second level of the quantum moment-SOS hierarchy is necessary for good approximations, since critical monogamy-of-entanglement properties begin to emerge at that level [16, 17].

The lower bound is usually determined by an algorithm that prepares a state with this energy. Several approximation algorithms have been proposed for QMC and EPR, using product states [7, 16, 17, 12], matchings [1, 16, 17, 14, 10, 8], and short variational circuits [1, 13, 12, 11, 8]. During the preparation of this manuscript, the best known approximation ratios for these problems were improved to α0.603 for QMC [8] and α1+540.809 for EPR [11].

We provide a new algorithm for both problems. For each algorithm, we start from a good product state. We then choose a matching in G and partially rotate matched qubits towards an entangled state. Our novel contribution is to perform this rotation such that individual single-qubit Bloch vectors are preserved, up to a rescaling of magnitude. This allows us to evaluate the energy of our entangled state in terms of the energy of the product state. For EPR, we use a fractional matching and provide a circuit-based algorithm. For QMC, we explicitly describe a tensor-product of single and two-qubit states, which interpolates between the product state and an integer matching-based state with a single tunable parameter. For QMC, we additionally introduce the technique of choosing a maximum-weight matching with respect to rescaled edge weights. One interpretation of this approach is that we widen the search space of our algorithm to include matchings other than the maximum-weight matching of G. Combining these techniques allows us to achieve state-of-the-art approximation ratios on each problem:

Theorem 1.

For EPR, αφ20.809, where φ=def1+521.618 is the golden ratio.

Theorem 2.

For QMC, α0.611.

 Remark 3.

Theorem 1 was shown simultaneously and independently by [11] using a different algorithm. The algorithm of [11] also partially entangles edges while preserving single-qubit marginals. However, we believe both our algorithm and analysis are simpler.

 Remark 4.

Theorem 2 requires an improved upper bound in addition to new algorithmic techniques. The technique of finding a matching with respect to rescaled edge weights was simultaneously and independently used by [8]; however, they use this as part of an improved approximation algorithm for QMC on triangle-free graphs rather than general graphs. Our QMC approximation does rely on a strengthening by [8] of a class of upper bounds first used in the work [14]. Without this strengthened bound, we still achieve an approximation ratio α0.610.

Our approximation algorithm finds a quantum state ρG such that Tr(HGρG)(G). For the EPR Hamiltonian, the analysis of our algorithm is optimal: there exist graphs G (such as the single-edge graph) where

(G)u(G)ψG|HG|ψGλmax(G)=1+54.

For QMC, however, it is possible that a better analysis of this algorithm (particularly of u(G)) would give a larger approximation ratio.

We stress the novelty of this work is new algorithmic techniques for QMC and EPR. We use these techniques to immediately improve the approximation ratios for each algorithm, however we believe that these techniques, potentially combined with previous methods, may lead to further improvements in approximation ratios for both algorithms. In the remainder of this work, we introduce necessary notation, describe each algorithm, and prove Theorems 1 and 2. We defer some technical lemmas to the appendix, which can be found in the full version of this paper [2].

2 Preliminaries

2.1 Graph theory

Let G(V,E,w) denote a graph with vertex set V, edge set E, and positive edge weights w:E+. In this work, we always take V=[n]=def{1,2,,n}. We let WG=def(i,j)Ewij. When G is inferred by context, we simply write W. Define N(v) as the set of neighbors of a vertex vV. For convenience, we index edges by either e or (i,j), or simply ij in subscripts.

A fractional matching of a graph G(V,E,w) is a function m:E(G)[0,1] that assigns a value to each edge in G, such that the sum over values jN(i)mij is at most 1 for any vertex i. If mij{0,1} for all edges (i,j), we say that m is an integral matching. For an integral matching m, we say an edge (i,j) is in the matching if mij=1. The weight Wt(m) of a matching m is defined as Wt(m)=def(i,j)E(G)wijmij. We define FMG (or MG) to be the maximum total weight of any fractional (or integral) matching of G. For convenience, we sometimes let MG also denote the set of edges in the maximum-weight integral matching of G. When the graph is clear from context we drop the subscript G. Optimal integral and fractional matchings can be computed in polynomial time, for example with linear programming [6].

2.2 Quantum computation

We refer to σ=(X,Y,Z) as the canonical Pauli matrices. As such, we can define the Bloch vector B(ρ) of any 1-qubit density matrix ρ as the vector (bx,by,bz) in the unit sphere S2 such that

ρ=12(I+B(ρ)σ)=def12(I+bxX+byY+bzZ).

Pure states correspond to unit Bloch vectors. Let |ψ=12(|01|10) be the singlet state, and |ϕ+=12(|00+|11) be the EPR state.

2.3 Previous algorithms

In our work, we use two previously known algorithms as subroutines:

  1. 1.

    The algorithm PROD takes a graph G(V,E,w) as an input, and outputs a good product state ρPROD. For EPR, we define ρPROD=def|0n0|n. For QMC, we define PROD to be the output of the [7] rounding algorithm applied to the some fixed constant level k of the quantum moment-SOS hierarchy, as in [17] and [14]. This is the only part of our algorithm for QMC that relies on an SDP. As described in [17, Sec 2.2], the output from the SDP can be interpreted as a pseudo-density matrix. It is called a pseudo-density matrix because it satisfies only some of the constraints of a valid density matrix. Furthermore, the SDP at any constant can be solved in polynomial time.

  2. 2.

    The algorithm MATCH takes a graph G(V,E,w) as an input, and outputs a product of 2-qubit states ρMATCH. It was first formally proposed in [14]. The algorithm first finds a maximum-weight integral matching MG. For EPR (and for QMC), it outputs the tensor product of |ϕ+ijϕ+|ij (and |ψijψ|ij for QMC) for every pair of vertices (i,j) in the matching, and the maximally mixed state for every vertex i not in the matching. This obtains energy 2 on matched edges and 1/2 on unmatched edges, for a total of (3MG+WG)/2.

3 Our Algorithms

To motivate our algorithms, consider the following simple graph G:

This graph is bipartite, so QMC and EPR are equivalent under local rotations [12]. We thus focus on EPR for simplicity. It is easy to compute that λmax(HG)=3. The optimal product state ρPROD is ρ03, where ρ0=def|00|, and achieves energy 2. The algorithm MATCH computes the maximum matching MG={(a,b)} and returns the state ρMATCH as described in Section 2. This state gains energy 2 on edge (i,j) and 1/2 on edge (b,c), achieving total energy 5/2. Thus, the algorithm of [14], which returns the better of ρPROD and ρMATCH, achieves 5/60.833 of the optimal energy.

Our algorithms do better by interpolating between ρPROD and ρMATCH, rather than taking the better of the two. For our example G, the unitary

U=eiθ(XaYa2XbYb2),

takes ρPROD to ρ, a tensor product of a two-qubit state ρab and a single-qubit state ρc. The parameter θ sets the entanglement for ρab: when θ=0, ρab is a product state; when θ=π/4, ρab is fully rotated into the EPR state. As such, we view this approach as smoothly interpolating between ρPROD and ρMATCH in superposition. It can be easily verified that the energy obtained by ρ on edge (a,b) is Tr[habEPRρ]=1+2cosθsinθ, and the single qubit marginals of a and b are given by

ρa=defTrbc[ρ]=cos2θρ0+2sin2θI=ρb.

In particular, note that the single qubit marginals are simply rescaled and shifted by the identity. This fact aids in the analysis: we can still evaluate the energy of (b,c) in terms of ρPROD, even though the edge is not in our matching:

Tr[hbcEPRρ] =Tr[hbcEPR(ρbρc)]
=Tr[hbcEPR(ρbρ0)]
=Tr[hbcEPR((cos2θρ0+2sin2θ(I/2))ρ0)]
=Tr[hbcEPR(cos2θρPROD+sin2θIρ0)]
=cos2θ+sin2θ=cos2θ.

Thus the total energy is given by

Tr[(habEPR+hbcEPR)ρ]=1+2cosθsinθ+cos2θ.

The parameter θ can be optimized over; in this example, taking θ.554 yields

Tr[(habEPR+hbcEPR)ρ]2.618,

outperforming both ρPROD and ρMATCH.

In this example, the analysis simplifies nicely because the initial product state is the symmetric state ρ0n, and because the graph is bipartite. For EPR on general graphs, we provide a circuit-based algorithm that prepares an entangled state based on a fractional matching in the graph. This state still works by partially entangling edges according to the matching while preserving single-qubit marginals, but it does not have a simple interpretation in terms of ρMATCH. We leave open the possibility that a (1+54)-approximation algorithm can be achieved with a tensor product of one and two-qubit states that simply interpolates between ρPROD and ρMATCH. For QMC, our algorithm does indeed interpolate between ρPROD and ρMATCH. However the algorithm and analysis become slightly more complicated. Crucial to our analysis is the following lemma:

Lemma 5 (Energy obtained by QMC algorithm on matched and unmatched edges).

Given two 1-qubit pure states ρi,ρj and a real parameter θ[0,π/2], there exists a 2-qubit pure state ρij such that

Tr[hQMCρij] =(1+sinθ)(1B(ρi)B(ρj))2,
B(Trj[ρij]) =cosθB(ρi),
B(Tri[ρij]) =cosθB(ρj).

The last two constraints in Lemma 5 indicate that the Bloch vectors of the single-qubit marginals of ρij are rescaled Bloch vectors of ρi and ρj, respectively. This is analogous to the argument in our example, where the marginals are rescaled and shifted by the identity. This again allows us to compute the energy of unmatched edges in terms of ρPROD. The parameter θ again sets the entanglement in ρij: when θ=0, ρij is a product state; when θ=π/2, ρij is maximally entangled. The proof of Lemma 5 is given in the appendix of [2].

3.1 EPR

Using the intuition from our example, we introduce the following algorithm for EPR:

Algorithm 1 (Fractional matching algorithm for EPR).

Given a weighted graph G(V,E,w):

  1. 1.

    Define the 1-qubit Hamiltonian

    P=def12(XY),

    and the angle

    θ=defln(φ)20.240, (1)

    where φ=1+52 is the golden ratio.

  2. 2.

    Find a fractional matching (muv)(u,v)E of maximum weight (e.g., via linear programming).

  3. 3.

    Output the state

    |χ=def(u,v)EeiγuvPuPv|0n, (2)

    where

    γuv =12cos1exp[θmuv][0,π/2]. (3)

The unitary eiγPuPv rotates the state |00ij towards the Bell state |ϕ+; this circuit was proposed in the approximation algorithm of [12].

3.2 QMC

Our algorithm for QMC relies on the following subroutine, which we call PMATCH since it places partially entangled 2-qubit states on the edges of a matching, while MATCH places maximally entangled states:

Algorithm 2 (PMATCH).

Given a weighted graph G(V,E,w) and a real parameter θ[0,π/2]:

  1. 1.

    Prepare the state ρPROD=i[n]ρi by running the PROD algorithm on QMC from Section 2 for some fixed, constant SDP level .

  2. 2.

    For each edge (i,j)E, compute tij=defB(ρi)B(ρj) from ρPROD.

  3. 3.

    Find a maximum-weight integral matching M~E on the reweighted graph G(V,E,w~), where

    w~ij=wij(sinθ(1tij(1+sinθ))2)+, (4)

    and E={(i,j)E|w~ij>0}. Here, we use the notation ()+=defmax(0,) from [14].

  4. 4.

    Let ρPMATCH(θ) be the state starting from ρPROD, but for each edge (i,j) in M~, replace ρiρj with the state ρij described in Lemma 5, parametrized by ρi,ρj,θ.

  5. 5.

    Output the state ρPMATCH(θ).

Our algorithm for QMC chooses the better of the algorithms PMATCH and MATCH:

Algorithm 3 (Combined algorithm for QMC).

Given a weighted graph G(V,E,w):

  1. 1.

    Prepare the state ρMATCH by running the MATCH algorithm from Section 2.

  2. 2.

    Prepare the state ρPMATCH(θ) by running PMATCH with parameter θ=1.286.

  3. 3.

    Output the state out of ρMATCH and ρPMATCH(θ) obtaining larger energy on the QMC Hamiltonian HG.

4 Analysis

We upper-bound the maximum energy λmax(HG) of QMC and EPR Hamiltonians using a quantifiable monogamy of entanglement:

Lemma 6 (Monogamy of Entanglement, e.g. [14, Lemma 1]).

For both EPR and QMC, and for all graphs G(V,E,w), we have

λmax(HG)WG+FMG.

For QMC, our analysis improves with better upper bounds on λmax(HG). We use an inequality of the form WG+MG/d; this first appears in [14, Lemma 4] with the constant d=4/5. By a detailed analysis and numerical verification on graphs with up to 13 vertices, this constant was recently improved to 14/15 in [8]:

Lemma 7 (Strengthened monogamy of entanglement, [8, Lemma 3.10]).

For QMC, and for all graphs G(V,E,w), we have

λmax(HG)WG+MGd,

where d=1415. Additionally, the bound holds for the optimal value of the 13-th level of the quantum moment SoS hierarchy.

4.1 EPR

We now prove Theorem 1. Given the state in Equation 2

|χ=def(i,j)EeiγijPiPj|0n, (5)

[12, Lemma 9] showed that the energy on the local term hijEPR is at least

χ|hijEPR|χ12(1+AijBij+sin2γij(Aij+Bij)), (6)

where

Aij =defkN(i){j}cos2γik,
Bij =defkN(j){i}cos2γjk.

In Algorithm 1, the output state |χ is in the form of Equation 5. Using the angles γij specified by Equation 3, we have

Aij =exp[θkN(i){j}mik]exp[θ(1mij)], (7)
Bij =exp[θkN(j){i}mjk]exp[θ(1mij)], (8)
sin2γij =1cos22γij=1exp[2θmij]. (9)

The inequalities in the first two lines follow because m is a matching. For example, we have mij+kN(i){j}mik1. Using Equation 6, the energy of |χ on hijQMC is at least

T(θ,mij)=def12(1+exp[2θ(1mij)]+21exp[2θmij]exp[θ(1mij)]). (10)

We combine Equation 10 with Lemma 6 to bound the approximation ratio on any graph G(V,E,w):

χ|HG|χλmax(HG)(i,j)EwijT(θ,mij)(i,j)Ewij(1+mij).

Each term in the numerator and denominator is positive, so the approximation ratio is at least the approximation ratio of the worst edge

min(i,j)ET(θ,mij)1+mijminx[0,1]T(θ,x)1+x. (11)

Recall from Equation 1 that we take θ=12lnφ. We empirically identified θ as the angle that maximizes the approximation ratio in this analysis. Substituting this value into T(θ,x) in Equation 10, the RHS of Equation 11 yields the minimization problem

minx[0,1][12(1+x)(1+φ(1x)+21φxφ12(1x))],

whose value is found to be φ/2 by the following lemma:

Lemma 8.

Given φ=def1+52, we have that

minx[0,1][12(1+x)(1+φ(1x)+21φxφ12(1x))]=1+54=φ2. (12)

The proof of Lemma 8 is given in the appendix of [2].

4.2 QMC

We now analyze Algorithm 3 and prove the 0.611-approximation in Theorem 2. We first introduce some helpful notation. Fix a graph G(V,E,w). Let ρ~ be the pseudo-density matrix outputted by solving the SDP of the level- quantum moment-SOS hierarchy (described in Section 2.3). We define

gij=defTr[hijQMCρ~],sij=def13Tr[(XiXj+YiYj+ZiZj)ρ~].

The first value is the energy obtained by ρ~ on hijQMC; the second value is the expected value of ρ~ with respect to the traceless components of the term hijQMC. It is straightforward to show that gij=12(13sij). The output of the SDP gives an upper bound to the optimal energy:

λmax(HG)(i,j)Ewijgij=(i,j)Ewij13sij2. (13)

Fix a positive integer k and suppose Lemma 7 is true for d=2k2k+1. Then, consider a reweighting G~(V,E,u) of the graph G where uij0. Let xE be a vector with

xij=defd(gij1)+=d(1+3sij2)+.

Here, we again use the notation ()+=defmax(,0) from [14], which ensures

(1+3sij2)+[0,1],(i,j)E.

It is shown (e.g. in [8, Lemma 3.4]) that x is in the integral matching polytope of G~. This means that x is a convex combination of integral matchings {m}1k. Since the weight of any integral matching m on G~ is at most MG~, the weight of x on G~ is also at most MG~. Thus,

(i,j)Euijd(1+3sij2)+MG~. (14)

Equation 14 in this context was first used in [16] and also appears in [14, Lemma 4] and [10, Lemma C.2].

We now find a lower bound for the energy obtained by the matching state ρPMATCH(θ) in PMATCH. We consider the energy on a case-by-case basis in the following lemma:

Lemma 9.

Consider the state ρPMATCH and matching M~ from Algorithm 2. For x{0,1,2}, let S~x be the subset of edges not in M~ (e.g. S~xEM~) where each edge in S~x has x matched endpoints in M~. Then the energy of ρPMATCH on an edge (i,j) is exactly

Tr[hijQMCρPMATCH]={12(1+sinθ)(1tij),(i,j)M~,12(1tij),(i,j)S~0,12(1tijcosθ),(i,j)S~1,12(1tijcos2θ),(i,j)S~2.

where, tij is defined in Algorithm 2.

Proof.

The case when (i,j)M~ follows directly from how we choose ρPMATCH in Algorithm 2, Step 4. For all other edges, recall that tij is the inner product of the Bloch vectors of qubits i and j with respect to the product state ρPROD. The reduced density matrix of ρPMATCH on qubits i and j is a product state σiσj. By [12, Lemma 10], the energy of σiσj on hijQMC is

12(1B(σi)B(σj)).

We handle each remaining case separately:

  1. 1.

    (i,j)S~0: Here, the product state is exactly ρiρj, and so B(σi)B(σj)=tij by definition.111Although M~ is a maximum matching, its edge set E is a subset of E, and so S~0 may be non-empty.

  2. 2.

    (i,j)S~1: Without loss of generality, i belongs to a matched edge in M~ and j is not in a matched edge. Lemma 5 implies that B(σi) is rescaled to B(σi)cosθ, and so B(σi)B(σj)=tijcosθ.

  3. 3.

    (i,j)S~2: i and j belong to two different matched edges. Therefore, it has both Bloch vectors rescaled, i.e. B(σi)B(σj)=(B(ρi)B(ρj))cos2θ=tijcos2θ.

Lemma 9 allows us to compute the energy obtained by the state ρPMATCH(θ):

PMATCH(θ)=def(i,j)EwijTr[hijQMCρPMATCH]
=(i,j)M~wij2(1+sinθ)(1tij)+(i,j)S~0wij2(1tij)
+(i,j)S~1wij2(1tijcosθ)+(i,j)S~2wij2(1tijcos2θ)
=(i,j)Ewij2(1tijcos2θ)+(i,j)M~wij2((1+sinθ)(1tij)(1tijcos2θ))
+(i,j)S~0wijtij2(cos2θ1)+(i,j)S~1wijtij2(cos2θcosθ)
=(i,j)Ewij2(1tijcos2θ)+(i,j)M~wij2(sinθ(1tij(1+sinθ)))
(i,j)S~0wijtij2sin2θ+(i,j)S~1wijtij2(cos2θcosθ). (15)

Note that M~ is a maximum-weight integral matching with respect to the rescaled weights defined in Equation 4. Thus, we can invoke Equation 14 to replace the sum over M~ in Equation 15 with a sum over E. Since the rescaled weights are positive for all edges in M~, we have

PMATCH(θ) (i,j)Ewij2(1tijcos2θ+d(sinθ(1tij(1+sinθ)))(1+3sij2)+)
(i,j)S~0wijtij2sin2θ+(i,j)S~1wijtij2(cos2θcosθ). (16)

Note that tij is a random variable because PROD is a randomized algorithm. The analysis of PROD in [7] is based on [4], which shows that 𝔼[tij]=F(sij). A definition of F in the context of QMC appears in [17, Equation 3]. For our purposes, it is enough to note that F is an odd function because it has the form

F(s)=csG(s2), (17)

where c>0 is a constant and G0 is a hypergeometric function. Since Section 4.2 is linear in tij, we use linearity of expectation to conclude

𝔼[PMATCH(θ)]
(i,j)Ewij2(1F(sij)cos2θ+dsinθ(1F(sij)(1+sinθ))(1+3sij2)+)
(i,j)S~0wij2F(sij)sin2θ+(i,j)S~1wij2((cos2θcosθ)F(sij)). (18)

The following lemma lets us analyze the performance of Algorithm 3, which takes the better of two algorithms, MATCH and PMATCH.

Lemma 10 (Reducing worst-case bounds to a single edge).

Suppose we have a collection of k approximation algorithms {A}[k] and an upper bound on the maximum energy λmax(HG)eEwebe, where beB=(0,bmax] for all eE, and bmax is a constant. Furthermore, suppose that the energy that algorithm A earns on edge e is a function of be, which we denote a(be), such that A earns eEwea(be). Then, by running each of A[k] and taking the output with the maximum energy, we can obtain an approximation ratio

αmaxμminbμa(b)b,

where the maximum is taken over valid probability distributions

{μ|[k]μ=1, 0μ1[k]},

and the minimum is taken over bB.

Proof.

Given a specific graph G, outputting the best of A is at least as good as outputting the result of A with probability μ, for all [k] and for any possible μ. Now, given a fixed distribution μ, the approximation ratio is at least

α μeEwea(be)eEwebe
=μeE(webeeEwebe)a(be)be
mineEμa(be)be
minbBμa(b)b.

The third line follows because

{webeeEwebe}eE

is a distribution over E. Since the statement is true for all possible distributions μ, we take the maximum over μ to finish the lemma.

We apply Lemma 10 with the SDP upper bound in Equation 13 and our two algorithms MATCH and PMATCH as described in Algorithm 3. For PMATCH, we consider the minimum of three different functions, depending on if an edge is in S~0, in S~1, or otherwise. For MATCH, we derived in Section 2.3 that the energy obtained by MATCH on G is 3M+W2. Since M is a maximum-weight matching with respect to G, we may invoke Equation 14 to express

M(i,j)Ewijd(1+3sij2)+,

implying that MATCH achieves energy at least

3M+W2(i,j)Ewij2(1+3d(1+3sij2)+).

We have lower-bounded the energy that each algorithm earns on an edge (i,j) as a function of sij. Note that the SDP upper bound is trivial (i.e. at most 0) when sij13, so we may search over s[1,1/3). Altogether, Lemma 10 implies that Algorithm 3 obtains approximation ratio at least

α maxμ[0,1],θ[0,π/2]min{mins[1,1/3)μEPM(θ)+(1μ)EM,
mins[1,1/3)μEPM(θ)+(1μ)EM,
mins[1,1/3)μEPM′′(θ)+(1μ)EM}, (19)
EPM(θ) =def1F(s)cos2θ+dsinθ(1F(s)(1+sinθ))(1+3s2)+13s, (20)
EPM(θ) =def1F(s)cosθ+dsinθ(1F(s)(1+sinθ))(1+3s2)+13s, (21)
EPM′′(θ) =def1F(s)+dsinθ(1F(s)(1+sinθ))(1+3s2)+13s, (22)
EM =def1+3d(1+3s2)+13s. (23)

In the above, μ is demoted to a scalar (which fully describes a probability distribution over two variables). Equation 23 is the approximation ratio obtained by MATCH on an edge with s=sij. The expected approximation ratio obtained by PMATCH is Equation 22 on an edge in S~0, Equation 21 on an edge in S~1, and is Equation 20 on all other edges.

We further simplify Equation 19. Recall that F is an odd function (Equation 17). So, for all θ[0,π/2] we have EPM′′(θ)EPM(θ)EPM(θ) when s>0 and EPM(θ)EPM(θ)EPM′′(θ) when s0. So we may rewrite Equation 19 as

α maxμ[0,1],θ[0,π/2]min{mins[1,0)μEPM(θ)+(1μ)EM,mins[0,1/3)μEPM′′(θ)+(1μ)EM}. (24)

One may then search over the three free parameters μ, θ, and s using an enumeration of the feasible ranges. For a more efficient approach, it is also possible to solve a linear program where the number of variables and constraints is based on an enumeration of the feasible ranges for s and θ. We obtain α>0.611 for d=14/15 at θ=1.286 and μ=0.861. By Lemma 7, our algorithm thus achieves a 0.611-approximation in expectation if =13. If the factor d in Lemma 7 is improved to 1, we would obtain α>0.614 at θ=1.288 and μ=0.881.

 Remark 11.

We may simplify Equation 19 by assuming s[1,0). This is because if s[0,1/3) then EM1 and EPM′′(θ)1 for all θ[0,π/2]. This follows from F being odd, and in addition |F(s)||s| for all s[1,1].

5 Discussion

In this work, we introduce two novel techniques for constructing algorithms for the Quantum MaxCut and EPR Hamiltonians on arbitrary graphs. The main techniques are a) partially entangling edges in a matching towards an entangled state while maintaining the direction of the marginal Bloch vectors as b) choosing matching with respect to reweighted graphs. We show that these techniques immediately lead to state-of-the-art approximation ratios for both problems. We believe that these techniques may be combined with other methods to obtain further improvements, and encourage further studies in this direction.

References

  • [1] Anurag Anshu, David Gosset, and Karen Morenz. Beyond product state approximations for a quantum analogue of Max Cut. LIPIcs, Volume 158, TQC 2020, 158:7:1–7:15, 2020. doi:10.4230/LIPIcs.TQC.2020.7.
  • [2] Anuj Apte, Eunou Lee, Kunal Marwaha, Ojas Parekh, and James Sud. Improved algorithms for quantum maxcut via partially entangled matchings, 2025. arXiv:2504.15276.
  • [3] Fernando G. S. L. Brandão and Aram W. Harrow. Product-state Approximations to Quantum Ground States, December 2014. arXiv:1310.0017.
  • [4] Jop Briet, Fernando Mario de Oliveira Filho, and Frank Vallentin. The positive semidefinite Grothendieck problem with rank constraint, 2010. doi:10.1007/978-3-642-14165-2_4.
  • [5] Toby Cubitt and Ashley Montanaro. Complexity classification of local Hamiltonian problems, March 2016. arXiv:1311.3161.
  • [6] Jack Edmonds. Paths, Trees, and Flowers. Canadian Journal of Mathematics, 17:449–467, January 1965. doi:10.4153/CJM-1965-045-4.
  • [7] Sevag Gharibian and Ojas Parekh. Almost optimal classical approximation algorithms for a quantum generalization of Max-Cut. LIPIcs, Volume 145, APPROX/RANDOM 2019, 145:31:1–31:17, 2019. doi:10.4230/LIPIcs.APPROX-RANDOM.2019.31.
  • [8] Sander Gribling, Lennart Sinjorgo, and Renata Sotirov. Improved approximation ratios for the Quantum Max-Cut problem on general, triangle-free and bipartite graphs, April 2025. arXiv:2504.11120.
  • [9] Felix Huber, Kevin Thompson, Ojas Parekh, and Sevag Gharibian. Second order cone relaxations for quantum Max Cut, November 2024. arXiv:2411.04120.
  • [10] Zackary Jorquera, Alexandra Kolla, Steven Kordonowy, Juspreet Singh Sandhu, and Stuart Wayland. Monogamy of Entanglement Bounds and Improved Approximation Algorithms for Qudit Hamiltonians, November 2024. arXiv:2410.15544.
  • [11] Nathan Ju and Ansh Nagda. Improved approximation algorithms for the EPR Hamiltonian, April 2025. doi:10.48550/arXiv.2504.10712.
  • [12] Robbie King. An Improved Approximation Algorithm for Quantum Max-Cut. Quantum, 7:1180, November 2023. doi:10.22331/q-2023-11-09-1180.
  • [13] Eunou Lee. Optimizing quantum circuit parameters via SDP, September 2022. arXiv:2209.00789.
  • [14] Eunou Lee and Ojas Parekh. An improved Quantum Max Cut approximation via matching, February 2024. arXiv:2401.03616.
  • [15] Miguel Navascues, Stefano Pironio, and Antonio Acin. A convergent hierarchy of semidefinite programs characterizing the set of quantum correlations. New Journal of Physics, 10(7):073013, July 2008. doi:10.1088/1367-2630/10/7/073013.
  • [16] Ojas Parekh and Kevin Thompson. Application of the Level-$2$ Quantum Lasserre Hierarchy in Quantum Approximation Algorithms. LIPIcs, Volume 198, ICALP 2021, 198:102:1–102:20, 2021. doi:10.4230/LIPIcs.ICALP.2021.102.
  • [17] Ojas Parekh and Kevin Thompson. An Optimal Product-State Approximation for 2-Local Quantum Hamiltonians with Positive Terms, June 2022. doi:10.48550/arXiv.2206.08342.
  • [18] Stephen Piddock and Ashley Montanaro. The complexity of antiferromagnetic interactions and 2D lattices, December 2015. arXiv:1506.04014.
  • [19] Jun Takahashi, Chaithanya Rayudu, Cunlu Zhou, Robbie King, Kevin Thompson, and Ojas Parekh. An SU(2)-symmetric Semidefinite Programming Hierarchy for Quantum Max Cut, August 2023. arXiv:2307.15688.
  • [20] Adam Bene Watts, Anirban Chowdhury, Aidan Epperly, J. William Helton, and Igor Klep. Relaxations and Exact Solutions to Quantum Max Cut via the Algebraic Structure of Swap Operators. Quantum, 8:1352, May 2024. doi:10.22331/q-2024-05-22-1352.