<<

arXiv:2105.01193v1 [quant-ph] 3 May 2021 ttso hlo lwdph unu icis[ output circuits the quantum (low-depth) over shallow based energy of states proposed the minimizing been have variationally preparation on algorithms range heuristic state wide vein, ground this a for In across platforms. implemented hardware be of quantum can few in- and consume increased that resources been algorithms has devising there in future, terest near the in available true the in present state. structure ground entanglement not the do approximations such capture Hartree-Fock methods or of mean-field classical state the Indeed, as the store efficiently. over even systems improvement such cannot some expect which to machines natural ap- classical Hamiltoni- is of local it task of Here the energy ans. consider state we ground is paper the systems, proximating this quantum comput- In of as unknown. properties still such advantage temperature tasks, quantum low simulation the ing important of scien- other extent for for quan- The useful and be science chemistry. can materials tum physics, they in that applications anticipated tific is it and neatos nteSplmna aeil ediscuss to we results Material, our two-local Supplemental of with extensions qubits the of In system a interactions. for results our simplicity, For the state we Hamiltonians. estimating local in of energy circuits state ground quantum shallow of formance a as stands challenge. algorithms pressing quantum variational and are quantum circuits shallow limitations by offered advantage some the Characterizing and problems [ an- known specific rigorously for been have alyzed algorithms variational though n h yaiso unu aybd ytm [ systems many-body quantum of dynamics the ing 2 obgn let begin, To oiae ysalqatmcmuesta a be may that computers quantum small by Motivated nti ae,w eiergru onso h per- the on bounds rigorous derive we paper, this In unu optr r aal fecetycomput- efficiently of capable are computers Quantum eateto obntrc n piiainadInstitut and Optimization and Combinatorics of Department mrvdapoiainagrtm o one-ereloc bounded-degree for algorithms approximation Improved 5 – 8 ,n eea ramn fterecc exists. efficacy their of treatment general no ], 1 rdc state product prxmto ai civdb ie rdc tt.Tea The circuit quantum state. shallow product given of a family by a achieved describe optimi ratio we approximation algorithms Here existing Most states. graphs. bounded-degree on tt iha nryta slwrthan lower is that energy an with state urneso nw loihsfrbuddocrec cla re bounded-occurrence we for state, algorithms product random known initial of an guarantees to applied When graph. eetn u eut to results our extend We ehv a Ω( = Var have we eateto EECS of Department ekly S n iosIsiuefrteTer fComputi of Theory the for Institute Simons and USA Berkeley, ecnie h ako prxmtn h rudsaeenerg state ground the approximating of task the consider We nrgAnshu Anurag G ( = 4 ,E V, etrfrTertclPyis ascuet nttt o Institute Massachusetts Physics, Theoretical for Center | v eagah n osdra consider and graph, a be ) i 3 k ihma energy mean with lclHamiltonians. -local eateto hmsr,Uiest fTrno aaaand Canada Toronto, of University Chemistry, of Department n 1 n h nryipoeeti rprinlt h ubrof number the to proportional is improvement energy the and ) ai Gosset David , & k hleg nttt o unu optto,Uiest o University Computation, Quantum for Institute Challenge lclHmloin n nage nta states. initial entangled and Hamiltonians -local e 2 0 2 ae .Mrn Korol Morenz J. Karen , = – e 4 0 h .Al- ]. v ya mutpootoa oVar to proportional amount an by | H 1 | ], v i n aineVr= Var variance and o unu optn,Uiest fWtro,Canada Waterloo, of University Computing, Quantum for e ag spsil.W ilas eitrse nefficient in interested states be prepare also that will algorithms We quantum possible. as large uhta h prxmto ratio approximation the that such prxmto ratios. approximation hsc,Hmloin ftefr q ( Eq. form the of Hamiltonians physics, Hamiltonian hr h oli ocmuea estimate an compute to is goal the where hspoieapyial oiae xeso fteclas- the of sical extension and motivated physically science a computer provide thus in considered cost-functions table l,oemycnie nIigHmloinfrwhich for Hamiltonian Ising an consider h may one ple, [ computers with polynomially n inverse scaling error additive an within ia piiainadi ihu oso eeaiy since generality, of loss without λ is and optimization sical with fetmtn h ags eigenvalue largest the estimating of h omtn em uha esneginteractions Heisenberg as non- such involve terms may commuting Hamiltonians graph quantum gener- the two-local More ally, of problem. Max-Cut optimization the classical well-studied finding eigenvalue to maximum equivalent its is compu- computing the basis—and in diagonal tational is, classical—that is Hamiltonian ovnett nta prxmt h ags eigenvalue largest λ the approximate instead to convenient ie once ya de easm ihu loss without assume We that edge. generality an of by connected tices rbe fapoiaigtegon nryo small- or energy ground eigenvalue the est approximating of problem 1 / ij min ij max eie eciiglclitrcin nonee in encountered interactions local describing Besides sblee ob nrcal o unu rclassical or quantum for intractable be to believed is 4( ( = htatnnrval nyo qubits on only nontrivially act that I ( ( H prxmto algorithm approximation n − H I = ) X ;ti ovninmthsteoeue nclas- in used one the matches convention this ); sclcntan aifcinproblems. satisfaction constraint ssical = + eteeeg vrtesto product of set the over energy the ze i oe n eeaieteperformance the generalize and cover ftolclqatmHamiltonians quantum two-local of y X htcnb sdt mrv the improve to used be can that s Z | grtmtksa nu an input as takes lgorithm V − 9 j i 3 .Hr ecnie h prxmto task approximation the consider we Here ]. − Z λ | g ekly aiona USA. California, Berkeley, ng, n ed Soleimanifar Mehdi and , max j Y uisadnaetniho interactions nearest-neighbor and qubits λ ) h i min / v Y ehooy USA Technology, f ,where 2, | ( j ( k − H − ( h H H H ij Z − fteHmloin twl be will It Hamiltonian. the of ) ≤ k i .I h os ae h problem the case, worst the In ). 2 = e Z /n 0 j ) { wt Pauli (with ) 2 natpclcase, typical a In . i,j Z | X lHamiltonians al v .W r neetdi the in interested are We 1. }∈ i stePuioeao.This operator. Pauli the is n upt a outputs and , E etn [ setting de nthe in edges h California, f ij λ r max n ≡ -qubit 1 ,Y X, 4 e/λ ( a noeno- encode can ) 10 H | .Frexam- For ]. ψ { fE.( Eq. of ) e max ,j i, i and ≤ ihgood with } ( λ H Z max tver- at sas is ) h oper- 1 G ij ( to ) H (1) a , = ) 2 ators); the resulting optimization problem can be viewed v1 v2 ... vn where each vi is a single- as a quantum analogue of Max-Cut [11]. Quantum ap- |qubiti ⊗ state. | i⊗ ⊗ | i | i proximation algorithms aim to estimate the largest eigen- Theorem 1. Given a product state v , we can efficiently value of such Hamiltonians and have been studied in sev- | i eral previous works. This includes the Heisenberg inter- compute a depth-(d + 1) quantum circuit U such that the state ψ = U v satisfies actions mentioned above [11, 12] and more general set- | i | i tings in which the interaction terms h are restricted to ij Var (H)2 be positive semidefinite [13–15], or traceless [16, 17]. ψ H ψ v H v +Ω v . (2) h | | i ≥ h | | i d2 E Despite considerable interest, the ultimate limits of ef-  | |  ficient algorithms for quantum approximation algorithms are poorly understood. Approximation ratios approach- This result applies broadly to quantum optimization ing 1 are only known to be achievable for certain special problems, but does not provide any improvement when families of graphs, including lattices or bounded-degree specialized to the classical setting. To see this, note that planar graphs using tensor product of O(1)-qubit states condition (i) is not satisfied in the purely classical case [18] or high degree graphs using tensor products of single- where v is a computational basis state and H is diagonal | i qubit states [13, 18, 19]. In certain cases, one may ascer- in the computational basis. Indeed, we have Varv(H)=0 tain limitations on efficient achievable approximation ra- whenever v is an eigenstate of H. On the other hand, tios from the classical Probabilistically Checkable Proof condition (i)| i is fairly mild in the quantum setting, which (PCP) theorem [20–22], though stronger and more gen- can be seen from the following expression for the vari- eral limitations may follow from the quantum PCP con- ance: jecture if some version of it can be proven [23]. Var (H)= ( v h h v v h v v h v ) . A quantum typically outputs v h | ij kl| i − h | ij | i · h | kl| i i,j k,l = an estimate of the form v H v where v is a quantum { }∩{X }6 ∅ h | | i | i state computed by the algorithm. A central challenge is Since G is d-regular, the number of terms in the sum to understand the structure of quantum states v that | i is O(d E ). So condition (i) is satisfied if the sum is achieve high approximation ratios in the general case. proportional| | to the number of terms appearing in it. Most existing algorithms are based on tensor products of one- or few-qubit states, while Ref. [12] also considers Simple examples demonstrate that neither of the two states prepared by shallow quantum circuits. In this work conditions alone is enough to even guarantee the exis- tence of a state with approximation ratio better than v we describe conditions under which the performance of | i such algorithms can be improved. We restrict our at- for large regular graphs. Condition (ii) alone is not suf- tention to local Hamiltonians on bounded-degree graphs ficient because it is possible for a product state to have and consider an improvement strategy based on shallow maximal energy λmax(H) (i.e., this occurs for all classi- quantum circuits. cal Hamiltonians). To see that condition (i) is not suf- ficient, one can consider the Max-Cut Hamiltonian on Improvement of product states To this end, suppose (say) an even cycle graph, and let v be an equal su- we are given an n-qubit state v and a Hamiltonian | i | i perposition of two eigenstates of H, one with maximal Eq. (1) defined on a graph G = (V, E) with maximum de- energy E and one with energy E Θ( E ). The re- gree d 2. It will be convenient to assume (without loss | | | |− | | 1/2 sulting state has approximation ratio 1 O( E − ) and of generality)≥ that G is d-regular—we can ensure this by −p | | variance Varv(H) = Ω( E ). Thus condition (i) is satis- possibly adding some local terms hij which are equal to fied, but the approximation| | ratio cannot be improved by zero. We imagine that v may be the output of some ap- an additive constant. proximation algorithm| suchi as the ones described above. In the special case where v achieves the largest en- Our aim is to efficiently compute a state with energy | i larger than v H v . Moreover, we would like to increase ergy of any product state, we are able to strengthen the h | | i bound Eq. (2). We say that the product state v is lo- this energy by an amount proportional to E in order to | i guarantee that the approximation ratio is larger| | by some cally optimal for H if for any single-qubit Pauli Q, we additive constant. We show that this is possible if the have following two conditions hold: d iφQ iφQ v e− He v =0, dφ h | | i φ=0 (i) The variance of the energy, defined by or equivalently v [Q,H] v = 0. As we show in the Sup- 2 2 h | | i Varv(H)= v H v v H v , plemental Material, the bound in Eq. (2) can be improved 2 h | | i − h | | i Varv(H) to v H v + Ω( d E ) for locally optimal states. satisfies Varv(H)=Ω( E )[24]. h | | i | | | | Generally, however, the improvement stated in Eq. (2) (ii) The state v is a product state. That is, v = is optimal in the sense that there exists a Hamiltonian | i | i 3

iθ P P † e kℓ k ℓ . Thus H and a product state v with Varv(H) = Θ( E ) for v Vij hij Vij v where Vij = k,ℓ Nij which | i | | h | | i { }∈ Q 2 ψ hij ψ = v hij v + Varv(H) h | | i h | | i λ (H) v H v O . (3) m max − h | | i≤ d2 E ∞ i  | |  v θkℓPkPℓ,hij v . (7) m!h | − m| i m=2  k,ℓ Nij  For example, Eq. (3) is satisfied by the Hamiltonian with X { X}∈ hij = Zi + Zj on any d-regular graph and the product n Here, [A, B]m is the m-nested commutator state v = (cos(θ) 0 + sin(θ) 1 )⊗ , for any θ (0,π/2). [A, [A, . . . [A, B]]]. Using the fact that vk Pk vk = 0 for In this| simplei case,| thei right-hand| i side can be∈ computed h | | i 2 2 all k, the m = 1 term simplifies to Var(H) 4 sin (θ) exactly and is equal to 2 2 . d E · sin (2θ) | | iθ v [P P ,h ] v = iθ v [P P ,h ] v . To establish Theorem 1, we consider a variational fam- − kℓh | k ℓ ij | i − ij h | i j ij | i k,ℓ Nij ily of states obtained from v = i V vi by applying a { X}∈ quantum circuit composed| ofi nearest⊗ ∈ neighbor| i commut- (8) ing gates on the interaction graph G. In particular, let At this stage, we make the choice P1, P2,...,Pn be any collection of single-qubit operators such that P 1 and θij = θ sign ( i v [PiPj ,hij ] v ) , (9) k ik≤ · − h | | i where the parameter θ will be determined later. Substi- vi Pi vi = 0 for all i V. h | | i ∈ tuting in Eq. (8) gives Following [12], we define the circuit iθkℓ v [PkPℓ,hij ] v = θ vi, vj [PiPj ,hij ] vi, vj . iθij PiPj i P{i,j}∈E θij PiPj − h | | i |h | | i| V (θ~)= e = e . (4) k,ℓ Nij { X}∈ i,j E (10) { O}∈ For m> 1, we have ~ Here, θ is an array of real parameters θij i,j E. Since by assumption, the interaction graph {G is}d{ -regular,}∈ the v θkℓPkPℓ,hij v quantum circuit V (θ~) can be implemented with circuit h | − m| i  k,ℓ Nij  depth d + 1. It is not hard to see that this variational { X}∈ m θ v [Pk1 Pℓ1 , [..., [Pk Pℓ ,hij ]]] v . family includes as a special case the level-1 Quantum Ap- ≤ |h | m m | i| k1,ℓ1 , k2,ℓ2 ,... proximate Optimization Algorithm (QAOA) for 2-local { } { } km,ℓXm Nij classical Hamiltonians [3]. For a given choice of oper- { }∈ ators Pi i V , the following theorem lower bounds the The only nonzero terms are those in which the expres- { } ∈ improvement in the energy after applying the the quan- sion vs Ps vs does not appear. To upper bound the tum circuit V (θ~) to v . h | | i | i number of nonzero terms, we count the number of tu- ples ( k ,ℓ , k ,ℓ ,... k ,ℓ ) such that no vertex Theorem 2. Let v be a product state and ψ = V (θ~) v 1 1 2 2 m m in V { i, j }appears{ } exactly{ once.} An upper bound is be the state prepared| i by the quantum circuit| Eq.i (4). De-| i provided\{ in} the following. fine the positive real parameter α by Claim 1. Let m 2. The number of ordered tuples of E ≥ m α = i,j E vi, vj [PiPj ,hij ] vi, vj , (5) edges ( k ,ℓ , k ,ℓ ),..., k ,ℓ ) N × in which { }∈ |h | | i| 1 1 2 2 m m ij no vertex{ in }V { i, j appears{ exactly} ∈ once is at most where the expectation is with respect to the uniform dis- (2m√d)m. \{ } tribution over the edges. There is an efficient classical ~ algorithm to select parameters θ satisfying Proof. First, we count all such tuples ( k ,ℓ , k ,ℓ ,... k ,ℓ ) in which the edge ψ H ψ v H v +Ω E α2/d . (6) { 1 1} { 2 2} { m m} h | | i ≥ h | | i | | i, j does not appear. Each one can be generated by { }  choosing a tuple of vertices (v1, v2,...vm) incident to i, j and then specifying a neighbor, either i or j, for { } Proof. Write N for the set of edges k,ℓ E incident each of them. An upper bound is obtained by counting ij { } ∈ to a given edge i, j E. The latter edge is included as the number of tuples (v1, v2,...vm) such that each vp well, i.e., i, j { N}. ∈ Consider the energy of a term occurs at least twice and then multiplying by 2m. Any { } ∈ ij tuple (v1, v2,...vm) of this form can be generated as ψ h ψ = v V (θ~)†h V (θ~) v . follows. First, for each i = 1, 2,...,m we choose a h | ij | i h | ij | i color c(i) 1, 2 ...,m/2 . We set v = v ′ whenever ∈ { } k k The gates in V (θ~) which are associated with edges that c(k) = c(k′). We then assign a neighbor of i or j to are not incident with i, j can be cancelled, leaving each color 1, 2,...,m/2 . Since vertices i and j each { } { } 4 have at most d neighbors, we see that the number of Pi = (Xi + Yi)/√2 for all i. By a direct calculation we tuples (v1, v2,...vm) such that each vp occurs at least see that twice is at most (m/2)m (2d)m/2. The number of · m 2 tuples of edges ( k ,ℓ , k ,ℓ ),... k ,ℓ N × α1 = Im ( 11 hij 00 ) { 1 1} { 2 2 { m m}} ∈ ij E | h | | i | in which no vertex in V i, j appears exactly once, | | i,j E \{ } { X}∈ and the edge i, j does not occur, is then at most 2 m m { m/} 2 α = Re ( 11 h 00 ) (12) 2 (m/2) (2d) . 2 E | h | ij| i | · · | | i,j E In order to account for the appearance of the edge { X}∈ i, j , we fix the number of places u where the edge ap- and therefore pears{ } and then count as before for the m u places. This − α1 + α2 number is 0n HQ H 0n = 11 h 00 2 E . h | 2 | i |h | ij | | ≤ | | 2 m i,j E   { X}∈ m √ m u √ m ((m u) 2d) − (2m d) . 1 n n u − ≤ This means max α , α E − 0 HQ H 0 which u=0   1 2 2 X together with Eq.{ (11) implies} ≥ | that| h when| t = 2,| wei canef- ficiently find a series of operators Pi such that the param- 1 eter α satisfies α (2 E )− Varv(H). By plugging this 2 Finally, using Eq. (9) and the fact that h , P 1, ≥ | | Varv (H) k ij k k ik≤ in Eq. (6), we obtain ψ H ψ v H v + Ω( ). we can upper bound h | | i ≥ h | | i d E Thus if t = 2 we obtain a better lower bound than| | the m m one claimed in Theorem 1. Otherwise, if t = 1, then a θ v [Pk1 Pℓ1 , [..., [Pkm Pℓm ,hij ]]] v (2θ) . |h | | i| ≤ simple calculation (reproduced in the Supplemental Ma- Thus, the sum of all m> 1 terms in Eq. (7) has magni- terial) shows that one can efficiently compute a product 2 Var(H) tude at most state with energy at least v H v + Ω( 2 ). In gen- h | | i d E eral, the choice between t = 1 and t = 2 can| be| efficiently ∞ 1 m ∞ m+2 4m√d θm 4e√dθ 32e2dθ2 determined. Thus we obtain Theorem 1. In the Supple- m! ≤ ≤ m=2 m=0 mental Material, we show that if v is locally optimal for X   X   H, then 0n HQ H 0n = 0 and |t i= 2, so we obtain the assuming θ 1 (where we used the bound mm/m! h | 1 | i ≤ 8e√d ≤ better bound described above. em). Combining with Eqs. (7,10) and summing over all Let us briefly illustrate how these results can be ap- i, j E, we get { } ∈ plied to the quantum Max-Cut Hamiltonian considered in Refs. [11, 12]. The Hamiltonian is built from local terms h = w Π , where 0 w 1 and Π = ψ H ψ v H v + E θα 32e2dθ2 . ij ij ij ≤ ij ≤ ij h | | i ≥ h | | i | | − (I XiXj YiYj ZiZj)/4 is the projector onto the − − − We may then choose θ = O(α/d) to get the desired lower antisymmetric state of two qubits. This Hamiltonian has the special feature that any product state v is locally bound. | i optimal, and moreover, we have v⊥, v⊥ h v , v = |h i j | ij | i j i| Let us now see how Theorem 1 is obtained as a conse- vi, vj hij vi, vj . Therefore h | | i quence of Theorem 2. The lower bound (6) applies to any 2 1 2 Var (H)= v h v E − v H v choice of operators Pi i V . We will choose these oper- v h | ij | i ≥ | | h | | i { } ∈ i,j E ators in a way that gives the variance bound Eq. (2). In { X}∈ the following, for convenience and without loss of gener- using Cauchy-Schwarz. We may then efficiently compute ality, we shall work in a local basis in which our initial a state ψ such that product state is v 0n . Our starting point is the ob- | i | i ≡ | i servation that the variance of a 2-local Hamiltonian can v H v 4 ψ H ψ v H v +Ω h | | i . (13) be expressed in this basis as h | | i ≥ h | | i d E 3  | |  n n n n Varv(H)= 0 HQ1H 0 + 0 HQ2H 0 , We see that if the initial state has approximation ratio h | | i h | | i v H v / E = r then the state ψ improves this to r + where Qt is the projector onto computational basis states hΩ(|r4/d| i)[|25|]. | i with Hamming weight t 1, 2 . This implies that ∈{ } The preceding example demonstrates the power of n n 0 HQtH 0 Varv(H)/2 (11) Theorem 1 and shows that for the quantum Max-Cut h | | i≥ problem, the approximation ratio of any product state for some t 1, 2 . Suppose t = 2 and let X ,Y , and can be improved by applying a shallow quantum circuit. ∈ { } i i Zi be the Pauli operators. We define α1 to be the RHS For more general two-local Hamiltonians, we can guaran- of Eq. (5) with Pi = Xi for all i, and similarly α2 with tee an improvement in the approximation ratio whenever 5 the condition Varv(H) = Ω( E ) holds, which we expect one may consider the trivial algorithm in which each for most (but not all) product| | states. Below we dis- variable is chosen independently and uniformly at ran- cuss two natural extensions of our results. First, we ask dom. Remarkably, efficient algorithms which improve whether one can improve approximation ratios attained over the approximation ratio achieved by this simple by more general families of quantum states. Along these strategy are not likely to exist in the general case [28]. lines, we provide an extension of Theorem 1 to the more On the other hand, for structured cases such as bounded- general case where v is any state prepared by a quan- degree graphs, improvement is possible. In particular, on tum circuit of depth| Di = O(1). Next, we show how one degree-d graphs, one can efficiently find an assignment 1 can improve the approximation ratio achieved by a ran- satisfying a µ + Ω( d ) fraction of constraints [29]. Here dom product state v . Using Theorem 2, we show that µ is the expected fraction of constraints satisfied by a the approximation ratio| i can be improved by Ω(1/d) for uniformly random assignment. It has been shown that any Hamiltonian with nontrivial two-local interactions, when a degree-d graph is triangle-free, there are efficient and by Ω(1/√d) if the interaction graph is triangle-free. “local” algorithms that find a binary string satisfying a µ + Ω( 1 ) fraction of constraints by starting with a uni- Improvement of bounded-depth states Recall that for √d any n-qubit quantum circuit and any qubit j [n], we formly random assignment [30, 31] or quantum superpo- may define the lightcone (j) [n] which consists∈ of all sition [32] and then locally updating each bit/qubit as a output qubits that are causallyL ⊆ connected to j. Define function of the state of its neighbors. the maximum lightcone size ℓ = maxj [n] (j). We have Below we show that this optimal dependence on d can ∈ L ℓ 2D for any depth D circuit composed of two-qubit be recovered and generalized to the local Hamiltonian ≤ gates. setting by applying our algorithm in Theorem 2 to a ran- n domly chosen product state. For randomly chosen v , Theorem 3. Let v = W 0 where W is a quantum | i circuit with maximum| i lightcone| i size ℓ. There is an effi- the parameter α in Theorem 2 can be related to the 2- cient classical algorithm that computes a quantum circuit norm of the quadratic terms in the Pauli expansion of U such that ψ = U v satisfies the Hamiltonian. More precisely, for an n-qubit operator | i | i O = f ij σi σj where σ = I and σ , σ , σ i0,y>0 For constant depth circuits we have ℓ = O(1) and we X X get the same asymptotic energy improvement as we es- Theorem 4. There is an efficient randomized algorithm tablished previously in Theorem 1 for product states. which computes a depth-d +1 circuit U such that ψ = However, in this case the circuit U that we construct U v satisfies | i is not constant-depth. In the Supplemental Material, we | i show that the improvement stated above can also be ob- 2 E E quad(H) tained for states v that are the unique ground states of v ψ H ψ v v H v +Ω . | i h | | i≥ h | | i d E a gapped local Hamiltonian F . In that case, ℓ is replaced  | |  by the locality of the Hamiltonian F . Thus, Theorem 3 If the graph is triangle-free then the right-hand side can extends to a broad class of tensor network states (such as quad(H) be replaced with Ev v H v +Ω . PEPS of low bond dimension) that have a gapped parent h | | i √d Hamiltonian.   The proof of Theorem 4 is provided in the Supple- The theorem provides limitations on the energy that mental Material. We also show that for triangle-free can be achieved by any state v produced by a bounded- graphs one can efficiently compute product states match- depth circuit. In particular,| sincei ψ H ψ λ (H), h | | i ≤ max ing the approximation ratios quoted above using a lo- we find that cal classical algorithm similar to the ones described in Var(H)2 Refs. [30, 31]. Thus, low depth quantum circuits are v H v λmax(H) Ω . not necessary to achieve the asymptotic Ω(1/√d) scal- h | | i≤ − ℓ10d2 E  | |  ing in this case. Nevertheless, one may take the output This shows that the approximation ratio achievable by product state of such algorithms and improve it further constant-depth states v with Var(H) = Ω( E ) is using the shallow quantum circuit from Theorem 1. | i | | bounded away from 1. An interesting direction for future Discussion For local Hamiltonian problems on work is to explore whether one can use this fact to exhibit bounded-degree graphs, we showed that the approxima- new local Hamiltonian systems with the almost-linear tion ratio achieved by a product state can be improved NLTS (No Low-energy Trivial States) property [26, 27]. by a shallow quantum circuit, assuming a mild condition Improvement of random assignments Given an in- on its variance. Our quantum algorithm generalizes the stance of a (classical) constraint satisfaction problem, level-1 Quantum Approximate Optimization Algorithm 6

(QAOA) and extends its applicability beyond classical [11] Sevag Gharibian and Ojas Parekh, “Almost optimal clas- cost functions. By applying our algorithm to randomly sical approximation algorithms for a quantum generaliza- chosen product states we generalized known algorithms tion of Max-Cut,” in Approximation, randomization, and for bounded-occurrence classical constraint satisfaction combinatorial optimization. Algorithms and techniques, LIPIcs. Leibniz Int. Proc. Inform., Vol. 145 (Schloss problems. Our results quantify the improvement that Dagstuhl. Leibniz-Zent. Inform., Wadern, 2019) pp. Art. shallow quantum circuits can provide over methods based No. 31, 17. on product states. [12] Anurag Anshu, David Gosset, and Karen Morenz, “Be- Acknowledgments AA acknowledges support from yond product state approximations for a quantum ana- the NSF QLCI program through grant number OMA- logue of max cut,” in 15th Conference on the Theory of Quantum Computation, Communication and Cryptogra- 2016245. DG acknowledges the support of the Natural phy (TQC 2020) (Schloss Dagstuhl-Leibniz-Zentrum f¨ur Sciences and Engineering Research Council of Canada Informatik, 2020). through grant number RGPIN-2019-04198, the Cana- [13] Sevag Gharibian and Julia Kempe, “Approximation al- dian Institute for Advanced Research, and IBM Re- gorithms for qma-complete problems,” SIAM Journal on search. KJMK acknowledges support from NSERC Computing 41, 1028–1050 (2012). Vanier Canada Graduate Scholarship. MS was supported [14] Sean Hallgren, Eunou Lee, and Ojas Parekh, “An by NSF grant CCF-1729369, a Samsung Advanced Insti- approximation algorithm for the max-2-local hamilto- nian problem,” in Approximation, Randomization, and tute of Technology Global Research Cluster and grant Combinatorial Optimization. Algorithms and Techniques number FXQi-RFP-1811A from the Foundational Ques- (APPROX/RANDOM 2020) (Schloss Dagstuhl-Leibniz- tions Institute and Fetzer Franklin Fund, a donor advised Zentrum f¨ur Informatik, 2020). fund of Silicon Valley Community Foundation. [15] Ojas Parekh and Kevin Thompson, “Beating random as- signment for approximating quantum 2-local hamiltonian problems,” (2020), arXiv:2012.12347 [quant-ph]. [16] Aram W Harrow and Ashley Montanaro, “Extremal eigenvalues of local hamiltonians,” Quantum 1, 6 (2017). [1] Seth Lloyd, “Universal quantum simulators,” Science , [17] Sergey Bravyi, David Gosset, Robert K¨onig, and Kristan 1073–1078 (1996). Temme, “Approximation algorithms for quantum many- [2] Alberto Peruzzo, Jarrod McClean, Peter Shadbolt, Man- body problems,” Journal of Mathematical Physics 60, Hong Yung, Xiao-Qi Zhou, Peter J Love, Al´an Aspuru- 032203 (2019). Guzik, and Jeremy L O’brien, “A variational eigenvalue [18] Nikhil Bansal, Sergey Bravyi, and Barbara M Terhal, solver on a photonic quantum processor,” Nature com- “Classical approximation schemes for the ground-state munications 5, 4213 (2014). energy of quantum and classical ising spin hamiltoni- [3] Edward Farhi, Jeffrey Goldstone, and Sam Gutmann, “A ans on planar graphs,” arXiv preprint arXiv:0705.1115 quantum approximate optimization algorithm,” arXiv (2007). preprint arXiv:1411.4028 (2014). [19] Fernando GSL Brandao and Aram W Harrow, “Product- [4] Abhinav Kandala, Antonio Mezzacapo, Kristan Temme, state approximations to quantum states,” Communica- Maika Takita, Markus Brink, Jerry M Chow, and tions in Mathematical Physics 342, 47–80 (2016). Jay M Gambetta, “Hardware-efficient variational quan- [20] , , , tum eigensolver for small molecules and quantum mag- , and , “Proof verifica- nets,” Nature 549, 242–246 (2017). tion and the hardness of approximation problems,” [5] Jarrod R McClean, Sergio Boixo, Vadim N Smelyanskiy, J. ACM 45, 501–555 (1998). Ryan Babbush, and Hartmut Neven, “Barren plateaus [21] Sanjeev Arora and , “Probabilistic check- in quantum neural network training landscapes,” Nature ing of proofs: a new characterization of NP,” communications 9, 1–6 (2018). J. ACM 45, 70–122 (1998). [6] Edward Farhi, David Gamarnik, and Sam Gut- [22] , “The PCP theorem by gap amplification,” mann, “The quantum approximate optimization algo- J. ACM 54, Art. 12, 44 (2007). rithm needs to see the whole graph: A typical case,” [23] Dorit Aharonov, Itai Arad, and Thomas Vidick, “Guest (2020), arXiv:2004.09002 [quant-ph]. column: the quantum pcp conjecture,” Acm sigact news [7] Sergey Bravyi, Alexander Kliesch, Robert Koenig, 44, 47–79 (2013). and Eugene Tang, “Obstacles to variational quan- [24] I.e. there is a universal constant c, such that asymptoti- tum optimization from symmetry protection,” cally Varv(H) ≥ c · |E|. Phys. Rev. Lett. 125, 260505 (2020). [25] A better bound can be obtained by directly computing [8] Sergey Bravyi, David Gosset, and Ramis Movassagh, the parameter α for a randomized choice of operators {Pi}. In that case, Eα ≥ Ω(hv|H|vi/|E|) which results in “Classical algorithms for quantum mean values,” Nature 2 Physics 17, 337–341 (2021). an improvement of Ω hv|H|vi /(d|E|). [9] In particular, a decision version of this problem is com- [26] Anurag Anshu and Chinmay Nirkhe, “Circuit lower plete for the complexity class QMA which is a quantum bounds for low-energy states of quantum code hamilto- generalization of NP. nians,” (2021), arXiv 2011.02044. [10] Vijay V Vazirani, Approximation algorithms (Springer [27] Michael H. Freedman and Matthew B. Hastings, “Quan- Science & Business Media, 2013). tum systems on non-k-hyperfinite complexes: A gener- alization of classical statistical mechanics on expander graphs,” Quantum Info. Comput. 14, 144–180 (2014). 7

[28] Johan Haastad, “Some optimal inapproximability re- [31] Matthew B Hastings, “Classical and quantum bounded sults,” J. ACM 48, 798–859 (2001). depth approximation algorithms,” arXiv preprint [29] Johan Haastad, “On bounded occurrence constraint sat- arXiv:1905.07047 (2019). isfaction,” Inform. Process. Lett. 74, 1–6 (2000). [32] Edward Farhi, Jeffrey Goldstone, and Sam Gutmann, [30] Boaz Barak, Ankur Moitra, Ryan O’Donnell, Prasad “A quantum approximate optimization algorithm applied Raghavendra, Oded Regev, David Steurer, Luca Tre- to a bounded occurrence constraint problem,” (2015), visan, Aravindan Vijayaraghavan, David Witmer, and arXiv:1412.6062 [quant-ph]. John Wright, “Beating the random assignment on con- [33] Ryan O’Donnell, Analysis of boolean functions (Cam- straint satisfaction problems of bounded degree,” (2015), bridge University Press, 2014). arXiv:1505.03424 [cs.CC]. 8

Supplemental Materials

IMPROVEMENT OF PRODUCT STATES

In this section we provide the full details of the proof of Theorem 1. It will be convenient to work in a local basis defined by v , such that v = 0n and | i | i | i Var (H)= 0n H2 0n ( 0n H 0n )2. v h | | i− h | | i For ease of notation we write Var(H) = Varv(H). Recall the quantity α defined in Eq. (5): E α = i,j E vi, vj [PiPj ,hij ] vi, vj , (S1) { }∈ |h | | i| We will use the following proposition.

Proposition 1. Let Q2 be the projector onto computational basis states with Hamming weight 2. We can efficiently choose operators Pi i V such that { } ∈ 1 α 0n HQ H 0n . (S2) ≥ E · h | 2 | i | |

Proof. Let α1 be Eq. (5) with Pi = Xi for all i, and let α2 be Eq. (5) with Pi = (Xi + Yi)/√2 for all i. Direct calculation shows that

2 n 2 n α = Im 11 0 − h 0 1 E h |ij ⊗ h | ij | i | | i,j E { X}∈  2 n 2 n α = Re 11 0 − h 0 (S3) 2 E h |ij ⊗ h | ij | i | | i,j E { X}∈ 

We can express 0n HQ H 0n as h | 2 | i n n n 2 n 2 0 HQ H 0 . = 11 0 − h 0 h | 2 | i |h |ij ⊗ h | ij | i| i,j E { X}∈ n 2 n 2 n 2 n 2 = Im 11 0 − h 0 + Re 11 0 − h 0 h |ij ⊗ h | ij | i h |ij ⊗ h | ij | i i,j E { X}∈   n 2 n n 2 n Im 11 0 − h 0 + Re 11 0 − h 0 ≤ h |ij ⊗ h | ij | i h |ij ⊗ h | ij | i i,j E { X}∈   α + α = E 1 2 , | | · 2 where we used the fact that h 1 in going from the second to the third line above. Now the last line implies that k ij k≤ either α1 or α2 achieves the bound from Eq. (S2). Moreover, the choice of α1 or α2 can be efficiently determined.

Proof of Theorem 1. Let Qt be the projector onto computational basis states with Hamming weight t 1, 2 . Since H is two-local we have ∈{ }

Var(H)= 0n HQ H 0n + 0n HQ H 0n . h | 1 | i h | 2 | i Therefore 0n HQ H 0n Var(H)/2 for some t 1, 2 . If t = 2 then we may use Proposition 1 which gives h | t | i≥ ∈{ } 1 max α , α Var(H). { 1 2}≥ 2 E | | Combining this with Theorem 2, we arrive at Var(H)2 ψ H ψ 0n H 0n +Ω h | | i ≥ h | | i d E  | |  9 which is better than the desired lower bound. Next suppose 0n HQ H 0n Var(H)/2. Define h | 1 | i≥ n L = ( 1)aj P − j j=1 X where each P is a single-qubit Pauli operator acting nontrivially only on qubit j, and a 0, 1 is chosen so that j j ∈{ } i 0n ( 1)aj [P ,H] 0n = 0n [P ,H] 0n . h | − j | i |h | j | i| iθL n Define θ = e− 0 where θ is a real parameter that we will fix later. Then | i | i n θ H θ = 0n H 0n + θ 0n [P ,H] 0n + Err, h | | i h | | i |h | j | i| j=1 X where

imθm Err = 0n [L,H] 0n | | m! h | m| i m 2 X≥ θm4m E ≤ | | m! m 2 X≥ 16θ2 E e4θ. ≤ | | In the second line we used the fact that

[L,h ] = [( 1)ai P + ( 1)aj P ,h ] ij m − i − j ij m can be expanded as a sum of 2m terms each of norm at most 2m. Now define

1 n β = 0n [P ,H] 0n . (S4) E |h | j | i| j=1 | | X and note that since [P ,H] 2d for all j we have k j k≤ n 1 β 2d 4. ≤ E ≤ j=1 | | X Then

θ H θ 0n H 0n + E θβ 16θ2e4θ . h | | i ≥ h | | i | | −  Choosing θ = β/32 gives

β2 β2 θ H θ 0n H 0n + E eβ/8 h | | i ≥ h | | i | | 32 − 64   β2 β2 0n H 0n + E e1/2 ≥ h | | i | | 32 − 64   0n H 0n +0.001 E β2. (S5) ≥ h | | i · | |

Now let β1 be given by Eq. (S4) with Pi = Xi for all i, and let β2 be given by Eq. (S4) with Pi = Yi for all i. Then 10

n β + β 1 1 2 = 0n [X ,H] 0n + 0n [Y ,H] 0n 2 2 E |h | j | i| |h | j | i| j=1 | | X 1 n 0n [X ,H] 0n 2 + 0n [Y ,H] 0n 2 ≥ 4d E |h | j | i| |h | j | i| j=1 | | X 1 n = 2Im( eˆ H 0n ) 2 + 2Re ( eˆ H 0n ) 2 4d E | h j | | i | | h j| | i | j=1 | | X 1 = 0n HQ H 0n d E h | 1 | i | | 1 Var(H). ≥ 2d E | |

Therefore either β1 or β2 is larger than the RHS above. Plugging this into Eq. (S5) we arrive at

Var(H)2 θ H θ 0n H 0n +Ω . h | | i ≥ h | | i d2 E  | | 

Finally, let us discuss a special case in which the bound from Theorem 1 can be improved. We say that a product state v is locally optimal for H if, for any single-qubit Pauli Q we have | i

d iφQ iφQ v e− He v =0, dφ h | | i φ=0 or equivalently

v [Q,H] v =0. (S6) h | | i As in the above, for simplicity we shall work in a local basis defined by v, so that v = 0n . | i | i Claim 2. Suppose 0n is locally optimal for H. Then for any string z 0, 1 n with Hamming weight z = 1 we have | i ∈ { } | |

z H 0n =0. (S7) h | | i

n 1 Proof. Without loss of generality consider the case where z = 10 − . Then

2Im( z H 0n ) = 0n [X ,H] 0n = 0 and 2Re( z H 0n ) = 0n [Y ,H] 0n =0, | h | | i | |h | 1 | i| | h | | i | |h | 1 | i| where we used Eq. (S6).

Claim 3. Suppose 0n is locally optimal for H. We may efficiently choose P and θ such that | i { i} { ij } Var(H)2 ψ H ψ 0n H 0n +Ω . (S8) h | | i ≥ h | | i d E  | |  Proof. Since H is two-local we have

Var(H)= 0n HQ H 0n + 0n HQ H 0n = 0n HQ H 0n , h | 1 | i h | 2 | i h | 2 | i where in the last equality we used claim 2. The claim then follows directly by combining Proposition 1 and Theorem 2. 11

IMPROVEMENT OF RANDOM STATES

We prove the first part of Theorem 4 regarding general degree-d graphs, which is implied by the following lemma.

Lemma 1. Let v = v v ... v where each v is a Haar random single-qubit state. Then there is an efficient | i | 1i⊗| 2i ⊗| ni i randomized process with random coins r, that constructs the matrices Pi (depending on both r and v ) such that the resulting state ψ satisfies | i | r,vi quad(H)2 E ψ H ψ E v H v +Ω . r,vh r,v| | r,vi≥ vh | | i d E  | | 

Proof. Pick v = i vi , where each vi is chosen uniformly at random from Haar measure on qubits. Also choose n | i ⊗ | i | i π uniformly random real numbers µi i.i.d in the interval [0, 2 ]. The latter choice is made using the coins r. Given v , r iµi iµi | i define Pi = e vi vi⊥ + e− vi⊥ vi (we drop the labels v , r from Pi for convenience). Observe that vi Pi vi =0 and P 1, as| required.ih | Then| αih |(as given in Eq (5) )| cani be evaluated to be h | | i k ik≤ v,r

E i(µi+µj ) i(µi+µj ) αv,r = i,j E e vi⊥, vj⊥ hij vi, vj e− vi, vj hij vi⊥, vj⊥ { }∈ h | | i− h | | i

E  i(µi+µj )  =2 i,j E Im e vi⊥, vj⊥ hij vi, vj . (S9) { }∈ h | | i  

iκi,j Let v⊥, v⊥ h v , v = e v⊥, v⊥ h v , v be the polar decomposition. Then h i j | ij | i j i |h i j | ij | i j i|

i(µi+µj ) Im e v⊥, v⊥ h v , v = sin (µ + µ + κ ) v⊥, v⊥ h v , v . | h i j | ij | i j i | | i j i,j | · |h i j | ij | i j i|   Note that

π π 4 2 2 2 E sin (µ + µ + κ ) = sin (µ + µ + κ ) dµ dµ , r| i j i,j | π2 | i j i,j | i j ≥ 5 Z0 Z0 for all κi,j . Then Eq S9 ensures that

4 E E E i(µi+µj ) E rαv,r =2 i,j E r Im e vi⊥, vj⊥ hij vi, vj i,j E vi⊥, vj⊥ hij vi, vj . { }∈ | h | | i |≥ 5 · { }∈ |h | | i|   Then we can evaluate

E 4 E v,rαv,r i,j E vi⊥, vj⊥ hij vi, vj dvidvj ≥ 5 · { }∈ | h | | i | Z  4 E 2 i,j E vi⊥, vj⊥ hij vi, vj dvidvj ≥ 5 · { }∈ |h | | i| Z 4 E = i,j E tr vi⊥, vj⊥ vi⊥, vj⊥ hij vi, vj vi, vj hij dvidvj 5 · { }∈ | ih | | ih | Z 4 E  = i,j E 11 Ui† Vj† hij (Ui Vj ) 00 00 Ui† Vj† hij (Ui Vj ) 11 dUidVj 5 · { }∈ h | ⊗ ⊗ | ih | ⊗ ⊗ | i Z     4 E = i,j E 1100 Ui†1 Vj†1 Ui†2 Vj†2 hi1,j1 hi2,j2 (Ui1 Vj1 Ui2 Vj2) 0011 dUidVj , 5 · { }∈ h | ⊗ ⊗ ⊗ ⊗ ⊗ ⊗ ⊗ | i Z   (S10) where in the second last equality we fixed a basis 0 , 1 for each qubit and introduced random unitaries Ui, Vj that specify v = U 0 , v = V 0 . Using the well known{| i | propertiesi} of Haar integral, we have | ii i| i | j i j | i

U † V † U † V † h h (U V U V ) dU dV i1 ⊗ j1 ⊗ i2 ⊗ j2 i1,j1 ⊗ i2,j2 i1 ⊗ j1 ⊗ i2 ⊗ j2 i j Z = aI I + bS  I + cI S + dS S . (S11) i1,i2 ⊗ j1,j2 i1,i2 ⊗ j1,j2 i1,i2 ⊗ j1,j2 i1,i2 ⊗ j1,j2 12

Above, I is the identity operator, S is the swap operator and the subscripts represent the qubits on which the operator acts. Coefficients a,b,c,d can be evaluated using the following system of equations, obtained from Eq. S11 by tracing each of the four operators.

2 (tri,j hij ) = 16a +8b +8c +4d

trj (trihij trihij )=8a +4b + 16c +8d

tri (trj hij trj hij )=8a + 16b +4c +8d 2 tri,j hij =4a +8b +8c + 16d. One can solve for d to obtain  2 2 (tr h ) tri,j h tr (tr h tr h )+tr (tr h tr h ) d = i,j ij + ij j i ij i ij i j ij j ij . 36 9  − 18 i,j i j In order to obtain a simpler lower bound and see that d is positive, we expand hij = x,y fx,yσx σy in the two qubit Pauli basis. Then ⊗ P 2 2 i,j 2 i,j 2 (tri,j hij ) = 16 f0,0 , tri,j hij =4 fx,y , x,y    X 

2 2 i,j i,j trj (trihij trihij )=8 f0,y , tri (trj hij trj hij )=8 fy,0 . y y X   X   Hence,

2 2 (tr h ) tri,j h tr (tr h tr h )+tr (tr h tr h ) i,j ij + ij j i ij i ij i j ij j ij 36 9  − 18 4 2 2 2 2 = f i,j + f i,j f i,j f i,j 9 0,0 x,y − 0,y − y,0 x,y y y !   X  X   X   4 2 = f i,j . 9 x,y x>0,y>0 ! X  Conjugating Eq. S11 with 1100 ( ) 0011 , it can be seen that only the term corresponding to d survives and evaluates to 1. Thus, Eq. S10 gives h | · | i

E α v,r v,r ≥ 2 tr h2 4E (tri,j hij ) i,j ij trj (trihij trihij )+tri (trj hij trj hij ) i,j E + 5 { }∈ 36 9  − 18 !

16E i,j 2 16 quad(H) = i,j E fx,y = . 45 { }∈ 45 E x>0,y>0 ! | | X  Thus, using the convexity of square function,

162 quad(H) 2 1 quad(H) 2 E α2 . v,r v,r ≥ 452 E ≥ 8 E  | |   | |  This completes the proof by employing Theorem 2.

Triangle-free graphs

In this section we establish the second part of Theorem 4, which concerns triangle-free graphs. The proof is based on the following exact expression. It will be convenient in what follows to work in a local basis in which the product state of interest is v = 0n . | i | i 13

Lemma 2 (Improvement for triangle-free Hamiltonians). Suppose G is a triangle-free, degree-d graph. Suppose 2 we are given single-qubit Hermitian operators Pi i [n] satisfying Pi = I and 0 Pi 0 =0 for all i [n], and consider i P θ P P { } ∈ h | | i ∈ the state ψ = e {r,s}∈E rs r s 0n as a function of the real parameters θ . Define | i | i { rs} α = 00 [h , P P ] 00 kl |h | kl k l | i| We can efficiently choose θ θ for each edge i, j E so that ij ∈ {± } { } ∈ 1 1 2d 2 1 d αkl d 1 ψ h ψ = Tr(h )+ Tr(h Z Z )cos − (2θ)+ Tr(h (Z + Z ))cos (2θ). + sin(2θ)cos − (2θ) (S12) h | kl| i 4 kl 4 kl k l 4 kl k l 2 for all edges k,l E { } ∈ Proof. We have

n n ψ h ψ = 0 V † h (θ)V 0 (S13) h | kl| i h | kl kl kl| i iθklPk Pl iθklPk Pl where hkl(θkl)= e− hkle and

iθksPk Ps iθrlPr Pl Vkl = e e (S14) k,s E k,l r,l E k,l { }∈Y\{ } { }∈Y\{ } = (cos(θ)+ i sin(θks)PkPs) (cos(θ)+ i sin(θrl)PrPl) . (S15) k,s E k,l r,l E k,l { }∈Y\{ } { }∈Y\{ } Plugging Eq. (S15) into Eq. (S13) and using 0 P 0 = 0 and the fact that G is triangle-free gives h | i| i 2d 2 A B A + B ψ h ψ = cos2(θ) − −| |−| | sin2(θ) | | | | h | kl| i A N(k) l B N(l) k ⊆ X\{ } ⊆ X\{ }   0n P P P P h (θ ) P P P P 0n · h | k s r l kl kl k s r l | i s A r B ! s A r B ! Y∈ Y∈ Y∈ Y∈ In the above we also used our choice θij = θ for all edges i, j E. Observe that the matrix element appearing in the above depends only on the parity| (even/odd)| of A and{ B}. ∈ In particular, | | | | ψ h ψ = F + F + F + F h | kl| i EE EO OE OO where the even/even term is

2 d 1 d 1 j j F = 00 h (θ ) 00 − cos2(θ) − − sin2(θ) (S16) EE h | kl kl | i  j  j=0,2,...   X   1 d 1 2  = 00 h (θ ) 00 1+cos − (2θ) , (S17) h | kl kl | i4 and by similar calculations one arrives at 

1 2d 2 F = 10 h (θ ) 10 1 cos − (2θ) (S18) EO h | kl kl | i4 − 1 2d 2  F = 01 h (θ ) 01 1 cos − (2θ) (S19) OE h | kl kl | i4 − 1 d 1 2 F = 11 h (θ ) 11 1 cos − (2θ) (S20) OO h | kl kl | i4 −  d 1 Now for ease of presentation in the following we write c = cos − (2θ) and axy = xy hkl(θkl) xy , for x, y 0, 1 . Then expanding the above expression gives h | | i ∈{ }

ψ h ψ = F + F + F + F (S21) h | kl| i EE EO OE OO a + a 1 = a + (1 c) 01 10 a + (1 c)2 (a + a a a ) . (S22) 00 − 2 − 00 4 − 11 00 − 01 − 10   14

Now let

b = xy h xy . (S23) xy h | kl| i So that

a = cos2(θ)b + sin2(θ)b + i cos(θ) sin(θ ) 00 [h , P P ] 00 . 00 00 11 kl h | kl k l | i

We now fix the sign of θkl so that 1 a = cos2(θ)b + sin2(θ)b + sin(2θ)α . (S24) 00 00 11 2 kl With this choice we have 1 a = cos2(θ)b + sin2(θ)b sin(2θ)α . (S25) 11 11 00 − 2 kl 2 2 To compute the third term in the above equation we used the fact that Pk = Pl = I and 0 Pk 0 = 0 Pl 0 = 0 which implies h | | i h | | i

11 [P P ,h ] 11 = 00 P P [P P ,h ]P P 00 = 00 [P P ,h ] 00 . h | k l kl | i h | k l k l kl k l| i −h | k l kl | i Similarly, by a direct calculation we get

a01 + a10 = b10 + b01. (S26)

Plugging Eqs. (S24, S25, S26) into Eq. (S22) we get b + b 1 ψ h ψ = b + (a b )c + (1 c) 01 10 b + (1 c)2 (b + b b b ) h | kl| i 00 00 − 00 − 2 − 00 4 − 00 11 − 10 − 01   αkl d 1 2 d 1 = b + sin(2θ)cos − (2θ) + sin (θ)cos − (2θ) (b b ) 00 2 11 − 00 d 1 b01 + b10 1 d 1 2 + (1 cos − (2θ)) b + (1 cos − (2θ)) (b + b b b ) (S27) − 2 − 00 4 − 00 11 − 10 − 01   Rearranging the above expression we arrive at

αkl d 1 1 1 2d 2 1 d ψ h ψ = sin(2θ)cos − (2θ)+ b + cos − (2θ)+ cos (2θ) h | kl| i 2 00 4 4 2   1 1 2d 2 1 d 1 1 2d 2 + b + cos − (2θ) cos (2θ) + (b + b ) cos − (2θ) (S28) 11 4 4 − 2 01 10 4 − 4     1 By noting that b = Tr(h ), ( 1)x+yb = Tr(h Z Z ), and b b = Tr(h (Z + Z )), we arrive x,y xy kl x,y − xy kl k l 00 − 11 2 kl k l at Eq. (S12). P P

Using the expression in (S12), we prove the bound for triangle-free graphs from Theorem 4:

Proof. As shown above, the exact formula for the energy of ψ = V (θ~) 0n on a triangle-free graph is | i | i 1 1 2d 2 ψ h ψ = Tr(h )+ Tr(h Z Z )cos − (2θ) h | kl| i 4 kl 4 kl k l 1 d αkl d 1 + Tr(h (Z + Z ))cos (2θ). + sin(2θ)cos − (2θ). (S29) 4 kl k l 2

Here αkl depends on the choices of Pk, Pl. We either choose Pi = Xi for all i, or Pi = (X + Y )i/√2 for all i, each with probability 1/2. Then

E(α )=2 Re( 00 h 11 ) +2 Im( 00 h 11 ) 2 00 h 11 . kl | h | kl| i | | h | kl| i |≥ |h | kl| i| 15

Substituting in Eq. (S29) gives

1 1 2d 2 E ( ψ h ψ ) Tr(h )+ Tr(h Z Z )cos − (2θ) h | kl| i ≥ 4 kl 4 kl k l 1 d d 1 + Tr(h (Z + Z ))cos (2θ). + 00 h 11 sin(2θ)cos − (2θ). (S30) 4 kl k l |h | kl| i| Now instead of using the starting state 0n , suppose we start from a random computational basis state s = X(s) 0n . Running through the above argument| in thei rotated basis defined by s we see that for a suitable random| i choice| ofi P we have | i { i}

1 00 hkl 11 + 01 hkl 10 d 1 E ( ψ h ψ ) (Tr(h )) + |h | | i| |h | | i| sin(2θ)cos − (2θ) h | kl| i ≥ 4 kl 2 1 1 d 1 (Tr(h )) + Tr(h X X ) sin(2θ)cos − (2θ). (S31) ≥ 4 kl 4| kl k l | Here we used the fact that

Es (Tr(X(s)hklX(s)ZkZl)) = Es (Tr(X(s)hklX(s)(Zk + Zl))) = 0. Summing Eq. (S31) over all edges k,l E gives { } ∈ 1 1 d 1 E ( ψ H ψ ) (Tr(H)) + sin(2θ)cos − (2θ) Tr (h X X ) h | | i ≥ 4 4 | kl k l | k,l E { X}∈ Since there is nothing special about the X-basis we can again use our freedom to randomize the local basis of each qubit to get

1 1 d 1 E ( ψ H ψ ) Tr(H)+ sin(2θ)cos − (2θ) Tr (h Q R ) h | | i ≥ 4 36 | kl k l | k,l E Q,R X,Y,Z { X}∈ ∈{X } 1 1 d 1 2 Tr(H)+ sin(2θ)cos − (2θ) Tr (h Q R ) /4 ≥ 4 36 | kl k l | k,l E Q,R X,Y,Z { X}∈ ∈{X } 1 d 1 quad(H) = Tr(H) + sin(2θ)cos − (2θ) , (S32) 4 36 where in the second-to-last line we used the fact that Tr (hklQkRl) 4 which follows from hkl 1. Finally, we can | |≤ d k1 k≤ find the maximum value of the second term with respect to θ by noting that sin(2θ)cos − (2θ) reaches a maximum when θ = arcsin( 1 ). Using this fact we get √d

1 quad(H) E( ψ H ψ ) Tr(H)+Ω . (S33) h | | i ≥ 4 √  d 

IMPROVEMENT OF BOUNDED-DEPTH STATES

We prove Theorem 3. Given the d-regular graph G = (V, E), we consider the state v = W 0 n, where W has a maximum lightcone of size ℓ. The aim is to increase the energy of v with respect to H.| Thei light| i cones of the edges have sizes at most 2ℓ. Define | i n F = W 1 1 W †. (S34) | ih |j j=1 X The locality of F is ℓ. Let A = i[H, F ] and define ψ = eiAθ v (thus U = eiAθ in the statement of Theorem 3). We can write | i | i

A = i[he, F ] := Ae, (S35) e E e E X∈ X∈ 16 where n Ae = i [he, W 1 1 j W †]= i [he, W 1 1 j W †]. (S36) | ih | † | ih | j=1 j:supp(he) supp(W 1 1 j W )=φ X ∩ X | ih | 6

Any j satisfying supp(he) supp(W 1 1 j W †) = φ is in the light cone of he. Thus there are 2ℓ such j’s. For any ∩ | ih | 6 2 ≤ such j, W 1 1 W † has locality ℓ. Thus, A is supported on 2ℓ qubits. Further, | ih |j e ≤ Ae 2ℓ max [he, W 1 1 jW †] 2ℓ, † k k≤ · j:supp(he) supp(W 1 1 j W ) k | ih | k≤ ⊂ | ih | where we used

[h , W 1 1 W †] = [h , W 1 1 W † I/2] 2 h W 1 1 W † I/2 1. k e | ih |j k k e | ih |j − k≤ k ekk | ih |j − k≤ We have

∞ ( iθ)m ψ H ψ = v H v i v [A, H] v θ + − v [A, H] v , (S37) h | | i h | | i− h | | i m! h | m| i m=2 X Now, using the identities F v = 0 and F I v v , we find | i ≥ − | ih | i v [A, H] v θ = v [[H, F ],H] v θ =2 v HFH v θ 2θ v H(I v v )H v =2θVar(H). (S38) − h | | i h | | i h | | i ≥ h | − | ih | | i Thus, let us focus on the terms with m 2. We upper bound ≥

v [A, H]m v [A, he]m E max [A, he]m . (S39) h | | i≤ k k ≤ | | e k k e E X∈ Now, consider for each e,

[A, he]m = [Aem , [Aem−1 ... [Ae1 ,he]]], (S40) e1,...e Xm where we used Eq. S35. Most terms are zero and we will bound the number of non-zero terms. We will use the following simple fact.

2 Fact 1. Let S V . The number of e such that the support of Ae overlaps with S is at most S ℓ d. For each such e and any operator⊂ O on S , the support of [O , A ] is at most S +2ℓ2. | | S S e | |

Proof. Since e is such that the support of A overlaps with S, there exist a j satisfying supp(h ) supp(W 1 1 W †) = e e ∩ | ih |j 6 φ for which the support of [h , W 1 1 W †] overlaps with S. Thus, either supp(h ) S = φ or j belongs to the light e | ih |j e ∩ 6 cone of S. Since the support of he overlaps with the light cone of j in the latter case, we have that the support of he overlaps with the light cone of the light cone of S (in both the cases). To upper bound the number of possible e, we hence we count the size of the light cone of the light cone of S ( S ℓ2) and then count the number of edges intersecting with this light cone ( d S ℓ2). For any such e, the support≤ of | [O| , A ] is contained in the union of S and ≤ | | S e the support of Ae. This completes the proof.

Using Fact 1, let us estimate the number of (e1,e2,...em) that contribute to Eq. S40. Setting S to be the set 2 of two vertices of e, we find that the number of e1 is at most 2dℓ . Arguing inductively, suppose we have fixed 2 e1,e2,...ek 1. The support size of [Aek−1 , [Aem−1 ... [Ae1 ,he]]] is at most 2 + 2(k 1)ℓ (by Fact 1). Thus, the number of e− contributing to Eq. S40 is at most (2+2(k 1)ℓ2)ℓ2d. Hence, the total number− of (e ,...e ) is at most k − 1 m 2 2 2 2 2 2 2 m 2m 2 2 m 4 m (2dℓ ) (2+2ℓ )ℓ d (2+4ℓ )ℓ d . . . (2 + 2(m 1)ℓ )ℓ d 2 (m 1)! ℓ − (2dℓ ) (m 1)!(4dℓ ) . · · − ≤ · − · · ≤ − Thus,

4 m [A, he]m (m 1)!(4dℓ ) max [Aem , [Aem−1 ... [Ae1 ,he]]] k k∞ ≤ − e1,...em k k 4 m m (m 1)!(4dℓ ) 2 he max Ae1 Ae2 ... Aem ≤ − · k k · e1,...em k k·k k ·k k (a) (m 1)!(4dℓ4)m 2m (2ℓ)m = (m 1)! (16dℓ5)m, (S41) ≤ − · · − · 17 where (a) uses A 2ℓ. Combining with Eq. S39, this ensures that k ek≤

∞ (iθ)m ∞ θm v [[H, A]] v E (m 1)! (16dℓ5)m m! h | m| i ≤ | | m! − · m=2 m=2 X X ∞ E (16dℓ5θ)m ≤ | | · m=2 X 2 E (16dℓ5θ)2, (S42) ≤ | | · 1 where the last inequality assumes θ 5 (our choice below will satisfy this). Thus, using Eq. S37 and Eq. S38, ≤ 32dℓ θ H θ v H v +2θVar(H) 2 E (16dℓ5θ)2 h | | i ≥ h | | i − | | · Var(H) = v H v +2θ E 2(16dℓ5)2θ . (S43) h | | i | | E −  | |  Var(H) 1 Setting θ = 210d2ℓ10 E 32dℓ5 , we conclude that | | ≤ Var(H)2 ψ H ψ v H v + . (S44) h | | i ≥ h | | i 210d2ℓ10 E | | We highlight that the above proof can be applied with minor modifications to the more general case in which F is a ℓ-local Hamiltonian with the unique ground state v and constant spectral gap γ = Ω(1). In this case, we set the ground energy of F at 0, leading to the relations F v| i= 0 and F γ (I v v ). Thus, the first order contribution in (S38) is replaced by | i  − | ih |

i v [A, H] v θ 2γθVar(H). − h | | i ≥ The higher order contributions are upper bounded in a manner similar to above.

IMPROVEMENT FOR GENERAL k-LOCAL HAMILTONIANS

Let G = (V, E) be a hypergraph with hyperedges of size at most k and n = V qubits on its vertices. We denote the number of hyperedges that contain i V by deg(i) and assume deg(i) | d| for all i V . Consider a k-local ∈ ≤ ∈ Hamiltonian H = R E hR where each local term hR acts non-trivially on a subset R E of qubits with R k ∈ ∈ n | | ≤ and hR 1. Here without loss of generality we assume the input product states is v = 0 . We use a similar argument|| || ≤as in theP proof of Theorem 1 to relate the improvement after applying an extension| i | of thei quantum circuit V (θ) in Eq. (4) to the variance Var(H). To this end, we write

k Var(H)= 0n HQ H 0n h | t | i t=1 X where Qt are the projector onto the computational basis states with Hamming weight t. Note that operators Qt with Hamming weight > k do not contribute. It holds that there exists a t such that 1 0n HQ H 0n Var(H). h | t | i≥ k

Depending on t, our choice of circuit V (θ) is a generalization of what we had before in Theorem 1. The set S contains all the collection of t different vertices j1, j2,...,jt which fully reside in the support of at least one local { } iθL term hR of the Hamiltonian H. That is, there exists an R such that j1, j2,...,jt supp(hR). Define V (θ)= e− where { }⊆

a L = ( 1) j1,...,jt P P ...P . − j1 j2 jt j1,j2,...,jt S { X }∈ 18

Here each Pj is a single-qubit Pauli operator with the property 0 Pj 0 = 0 that acts nontrivially only on qubit j and a 0, 1 is chosen so that h | | i j1,...,jt ∈{ } n a n n n i 0 ( 1) j1,...,jt [P P ...P ,H] 0 = 0 [P P ...P ,H] 0 . h | − j1 j2 jt | i |h | j1 j2 jt | i| iθL n Let θ = e− 0 . Then | i | i θ H θ = 0n H 0n + θ 0n [P P ...P ,H] 0n + Err, h | | i h | | i |h | j1 i2 jt | i| j1,...,jt S { X}∈ where the higher order terms Err can be bounded as

imθm Err = 0n [L,H] 0n | | m! h | m| i m 2 X≥ m m θ k E 2kd ≤ | | m! t 1 m 2    X≥ − 2 k 2 2kd k θ 2kd θ E e (t−1) . (S45) ≤ t 1 | |   − 

In the second line, we used the fact that [L,H]m = R[L,hR]m and each term [L,hR] can be expanded as a sum m k m of at most kd t 1 non-zero terms each of normP at most 2 . This is because the operators Pj1 , Pj2 ,...,Pjt − commute with each other for different choices of j1, j2,...,jt and only those that overlap with the support of hR  { } k may contribute. The number of such operators (i.e. S supp(hR) ) can be bounded by kd t 1 as follows: There are | ∩ | − at most k vertices in supp(hR) and each vertex is in the support of d other terms in the Hamiltonian. Given a vertex ≤ k j supp(hR) and an overlapping Hamiltonian term hR′ such that j supp(hR′ ), there are t 1 choices of vertices ∈ ∈ − k j , j ,...,j supp(h ′ ) that contain j. Hence, from the definition of set S follows that S supp(h ) kd 1 2 t R  R t 1 (one{ can obtain}⊆ tighter bounds using 0 P 0 = 0). Now define | ∩ |≤ − h | j | i  1 β = 0n [P P ...P ,H] 0n . (S46) E |h | j1 j2 jt | i| j1,...,jt S | | { X}∈

It holds that β 2 k . To see this, note that 0n [P P ...P ,H] 0n 0n [P P ...P ,h ] 0n . Us- ≤ t |h | j1 j2 jt | i| ≤ R |h | j1 j2 jt R | i| ing the assumption 0 Pj 0 = 0, it follows that the only choices of vertices j1,...,jt that may contribute in 0n [P P ...P ,hh ]|0n | iare those which are fully contained in supp(h ). TheP{ number of} such vertices is bounded |h | j1 j2 jt R | i| R by k . Using 0n [P P ...P ,h ] 0n 2, we arrive at the claimed bound β 2 k . We have t |h | j1 j2 jt R | i| ≤ ≤ t

 2  n n k 2 2kd k θ θ H θ = 0 H 0 + E θβ 2kd θ e (t−1) . h | | i h | | i | | − t 1   −  !

β Choosing θ = O 2 2 k 2 gives k d ( 1)  t− 

2 n n E β θ H θ 0 H 0 +Ω | | 2 . (S47) h | | i≥ | | i  2 2 k  k d t 1 −    Now let β1 be given by Eq. (S46) with Pj1 Pj2 ...Pjt = Xj1 Xj2 Xjt for all j1,...,jt S. Define eˆ = X X X 0n and the operator ⊗ ⊗···⊗ { } ∈ | j1,...,jt i j1 ⊗ j2 ⊗···⊗ jt | i i π 0 e− 2t p = i π . e 2t 0   19

Let β be given by Eq. (S46) with P P ...P = p p p for all j ,...,j S. Then, one can see that 2 j1 j2 jt j1 ⊗ j2 ⊗···⊗ jt { 1 t} ∈ β + β 1 1 2 = Im( eˆ H 0n ) + Re( eˆ H 0n ) 2 E | h j1,...,jt | | i | | h j1,...,jt | | i | j1,...,jt S | | { X}∈ 1 Im( eˆ H 0n ) 2 Re( eˆ H 0n ) 2 d | h j1 ,...,jt | | i | + | h j1,...,jt | | i | ≥ E · d2 d2 j1,...,jt S ! | | { X}∈ 1 = 0n HQ H 0n d E |h | t | i| | | 1 Var(H). ≥ kd E | | 1 This implies either β1 or β2 is larger than kd E Var(H). By plugging this into Eq. (S47) we arrive at | |

2 n n Var(H) θ H θ 0 H 0 +Ω 2 . h | | i ≥ h | | i  4 4 k  k d t 1 E − | |    This bound is minimized by allowing t = k/2+1 which results in the following overall lower bound on the improve- ment to the energy of the input state 0n⌈: ⌉ | i Var(H)2 θ H θ 0n H 0n +Ω . (S48) h | | i ≥ h | | i 2O(k)d4 E  | |  We note that the Ω(1/d4) dependence of (S48) on the degree is quadratically worse than the bound we obtained for 2-local Hamiltonians in Theorem 1; It would be interesting to recover the Ω(1/d2) scaling in this case.

LOCAL CLASSICAL ALGORITHMS

Theorem 5. Consider a two-local Hamiltonian H = i,j E hij where G = (V, E) is a d-regular triangle-free graph. There is an efficient randomized algorithm that computes{ a}∈ product state v = n v satisfying P | i ⊗i=1| ii 1 quad(H) E v H v Tr(H)+Ω . (S49) vh | | i≥ 4 √  d  Note that in Eq. (S49) the first term on the right hand side is equal to the expected energy of H with respect to a random state ρ = I/2n.

Proof of Theorem 5. It will be convenient to work in a local Pauli basis X, Y or Z chosen at random and independently for each qubit. We write hij in this randomly chosen product basis. Let us define wij = Tr(hij )/4 and

1 1 1 ux = Tr(h X X ) uy = Tr(h Y Y ) uz = Tr(h Z Z ). ij 4 ij i j ij 4 ij i j ij 4 ij i j Due to the random choice of basis we have 1 E[(ua )2]= quad(h ) a x,y,z . (S50) ij 9 ij ∈{ }

n Following [S30, S31], we start with a random i.i.d assignment of pure product v = i=1 vi states to the vertices. We then select a subset of vertices A uniformly at random. For any vertex i |Ai, let⊗N(i|) =i i, j E : j / A be the neighboring edges that contain exactly one vertex in A (i.e. vertex i).∈ The remaining edges{{ } that ∈ are not∈ in} i AN(i) either connect two vertices that are not in A or connect two vertices in A. We denote the former by M and ∪ ∈ the latter by M ′. 20

1 x y z The initial random pure state at each vertex ρi can be represented by ρi = 2 (I + ri Xi + ri Yi + ri Zi), where x y z R3 x 2 y 2 z 2 (ri , ri , ri ) is the Bloch vector with norm ri + ri + ri = 1. For a vertex i A, the total energy of the edges N(i)∈ is given by | | | | | | ∈

tr[h ρ ρ ]= w + (ux rxrx + uy ryry + uz rzrz)+ D (~r , ~r ) (S51) ij i ⊗ j ij ij i j ij i j ij i j ij i j j: i,j N(i) j: i,j N(i) j: i,j N(i) j: i,j N(i) { X}∈ { X}∈ { X}∈ { X}∈ where

ab a b a a a a Dij (~ri, ~rj )= cij ri rj + (dij ri + eij rj ). a=b a x,y,z X6 ∈{X } ab a a for some coefficients cij , dij ,eij . Using Cauchy–Schwarz inequality, we see that the first two terms in Eq. (S51) can be maximized by applying a local unitary on each vertex i A which rotates the state ρi to a stateρ ˜i with the Bloch vector ∈

1/2 a a a x x 2 y y 2 z z 2 − Ri = ( uij rj ) ( uij rj ) + ( uij rj ) + ( uij rj ) (S52) j: i,j N(i)  j: i,j N(i) j: i,j N(i) j: i,j N(i)  { X}∈ { X}∈ { X}∈ { X}∈ for a x,y,z . When the denominator of Eq. (S52) is zero, the vector R~ is chosen uniformly at random. Hence, we get ∈{ }

1/2 tr[h ρ˜ ρ ]= w + ( ux rx)2 + ( uy ry)2 + ( uz rz)2 ij i ⊗ j ij ij j ij j ij j j: i,j N(i) i,j N(i)  j: i,j N(i) j: i,j N(i) j: i,j N(i)  { X}∈ { X}∈ { X}∈ { X}∈ { X}∈ (S53)

+ Dij (R~ i, ~rj ), j: i,j N(i) { X}∈ A property of this construction is that E[Ra] = 0 for a x,y,z and R , R are independent of each other for i ∈ { } i j i, j M ′. This follows from the triangle-freeness, the definition of the set N(i), and the initial uniform i.i.d. distribution{ } ∈ of the state of vertices. Moreover, we have

E[rarb ]= E[Rarb ]=0 whenever a = b . j k j k 6 Using these observations, the expected value of the total energy after the local improvements is

E tr[h ρ˜ ρ˜ ] + E tr[h ρ ρ ] + E tr[h ρ˜ ρ ]  ij i ⊗ j   ij i ⊗ j   ij i ⊗ j  i,j M ′ i,j M i A j: i,j N(i) { X}∈ { X}∈ X∈ { X}∈      1/2 = w + E ( ux rx)2 + ( uy ry)2 + ( uz rz)2 ij  ij j ij j ij j  i,j E i A  j: i,j N(i) j: i,j N(i) j: i,j N(i)  { X}∈ X∈ { X}∈ { X}∈ { X}∈   w + E uz rz (S54) ≥ ij  | ij j | i,j E i A j: i,j N(i) { X}∈ X∈ { X}∈   The first term i,j E wij corresponds to the expected energy when the product states are chosen uniformly at random. The second{ }∈ term is a lower bound on the improvement achieved by the local updates which we now show is at least Ω( E /√Pd). | | For a fixed choice of the set A V , we define the random variable ξ = uz rz. Using the “second ⊆ i j: i,j N(i) ij j moment method,” for t [0, 1], we get { }∈ ∈ P E[ξ2]2 Pr ξ t E[ξ2] (1 t2)2 (S55) | i|≥ ≥ − E[ξ4] h p i 21

One way to sample uniform pure states over Bloch sphere is to uniformly draw φ [0, 2π], rz [ 1, 1] and set ∼ j ∼ − rx = 1 (rz)2 cos φ and ry = 1 (rz)2 sin φ. Given this we have E[rz ] = 0, E[(rz )2]=1/3, E[(rz )3] = 0, and j − j j − j j j j E z 4 E 4 E 2 2 [(rj )q]=1/5. Hence, using Corollaryq 9.6 of [S33], we have [ξ ] 9 [ξ ] . Plugging this in (S55) implies that for a fixed choice of the set A, the expectation with respect to the random≤ · distribution of the initial product states for an arbitrary choice of t [0, 1] is ∈

1 E uz rz t(1 t2)2 E[ξ2]1/2 | ij j | ≥ 9 · − · j: i,j N(i) { X}∈   1/2 1 = t(1 t2)2 (uz )2E (rz)2 9 · − ·  ij j  j: i,j N(i) { X}∈    1/2  1 t(1 t2)2 (uz )2 . (S56) ≥ 9√3 · − ·  ij  j: i,j N(i) { X}∈   Finally, we calculate the expectation with respect to the set A V . Note that the set N(i) is also a random variable determined by the set A. Conditioned on the event that the vertex⊆ i A and using Theorem 9.24 of [S33], we have ∈

1 1 Pr (uz )2 (uz )2 .  ij ≥ 2 ij  ≥ 4e2 j: i,j N(i) j: i,j E { X}∈ {X}∈   Thus, we get

1/2 1/2 1 E (uz )2 = (uz )2   ij   8√2e2  ij  i A j: i,j N(i) i V j: i,j E X∈ { X}∈  X∈ {X}∈     1 1  = (uz )2. (S57) 4√2e2 · √ ij d i,j E { X}∈ Finally, taking the expectation over the random choice of local basis and using Eq. (S50) we get

1/2 z 2 1 1 E (u ) quad(hij )   ij   ≥ 36√2e2 · √ i A j: i,j N(i) d i,j E X∈ { X}∈  { X}∈     1 1 quad(H) (S58) ≥ 36√2e2 · √d We arrive at (S49) by plugging this into (S56) and using (S54).