Correlation Decay up to Uniqueness in Spin Systems

Liang Li
Shandong University
li.liang@sdu.edu.cn
This work was done when this author visited Microsoft Research Asia.
   Pinyan Lu
SHUFE
lu.pinyan@mail.shufe.edu.cn
   Yitong Yin
Nanjing University
yinyt@nju.edu.cn
Abstract

We give a complete characterization of the two-state anti-ferromagnetic spin systems which are of strong spatial mixing on general graphs. We show that a two-state anti-ferromagnetic spin system is of strong spatial mixing on all graphs of maximum degree at most Δ if and only if the system has a unique Gibbs measure on infinite regular trees of degree up to Δ, where Δ can be either bounded or unbounded. As a consequence, there exists an FPTAS for the partition function of a two-state anti-ferromagnetic spin system on graphs of maximum degree at most Δ when the uniqueness condition is satisfied on infinite regular trees of degree up to Δ. In particular, an FPTAS exists for arbitrary graphs if the uniqueness is satisfied on all infinite regular trees. This covers as special cases all previous algorithmic results for two-state anti-ferromagnetic systems on general-structure graphs.

Combining with the FPRAS for two-state ferromagnetic spin systems of Jerrum-Sinclair and Goldberg-Jerrum-Paterson, and the very recent hardness results of Sly-Sun and independently of Galanis-S̆tefankovic̆-Vigoda, this gives a complete classification, except at the phase transition boundary, of the approximability of all two-state spin systems, on either degree-bounded families of graphs or family of all graphs.

1 Introduction

Spin systems are well studied in the areas of Statistical Physics, Applied Probability and Computer Science as a general framework to capture the essence of how local interactions and constrains affect the macroscopic properties of particle systems. A system is usually described by a graph, with each vertex in one of a fixed number of states called spins, and edges specifying the neighborhood relation of the system.

Let G(V,E) be a graph and q be the number of spin states. A configuration of the system is one of the q|V| possible assignments σ:V[q]. Each configuration σ has an energy H(σ) as a sum over all edges and vertices, such that the contribution of an edge (u,v)E is determined by a symmetric function of the spin states σ(u) and σ(v), and the contribution of a vertex vV is determined by a function of its spin state σ(v). The weight of a configuration σ is w(σ)=exp(H(σ)T), where T is the temperature. We focus on the two-state spin systems. Up to normalization, a two-state spin system is fully captured by three parameters (β,γ,λ), where β and γ determine the symmetric function for edge contribution and λ, also known as the external field, determines the function for vertex contribution. The Gibbs measure is a natural probability distribution over all configurations such that the probability of a configuration σ is w(σ)Z, where the normalizing factor Z=w(σ) is called the partition function. The partition function encodes rich information regarding the macroscopic behavior of the spin system. However, for almost all nontrivial settings it is #P-hard to compute the precise value of partition functions.

One of the most important properties of spin systems is the correlation decay, which says that the correlation between the marginal Gibbs distributions of two vertices decays rapidly with respect to the distance between the two vertices. This property is also called weak spatial mixing [26]. Of greater algorithmic significance is the strong spatial mixing, which says that the correlations decay in the presence of arbitrary fixed spins at other vertices. For two-state spin systems, the strong spatial mixing may imply efficient approximation algorithms for the partition function. It is then an important question to characterize the systems which exhibit strong spatial mixing on arbitrary instances of graphs by the parameters of the systems.

A two-state spin system is ferromagnetic if adjacent vertices favor agreement of spins, and is anti-ferromagnetic if otherwise. For all two-state ferromagnetic spin systems, the partition functions can be efficiently approximated [12, 10]. In the anti-ferromagnetic region, the correlation decay plays a central role in the approximability of partition functions. It is believed (see [19]) that for such models the approximability of the partition function is characterized by the uniqueness of Gibbs measure on infinite regular trees, which is equivalent to the weak spatial mixing on the infinite regular trees. This condition is called the uniqueness condition.

In a seminal work [26], Weitz shows that for the hardcore model strong spatial mixing is characterized by the uniqueness condition, and in a recent work of Sinclair, Srivastava, and Thurley [21] the same characterization is proved for the anti-ferromagnetic Ising model. Both models are important special two-state anti-ferromagnetic spin systems. On the hardness side, it is proved for the hardcore model by Sly [23] and Galanis et al. [6], and very recently for the general two-state anti-ferromagnetic spin systems by Sly and Sun [24] and independently by Galanis, S̆tefankovic̆, and Vigoda [7] that violating the uniqueness condition implies the inapproximability of partition functions. Two questions remain open for our complete understanding of the correlation decay and computation in two-state spin systems: the characterizations of the strong spatial mixing and approximability of general two-state spin systems.

1.1 Our results

We characterize the two-state anti-ferromagnetic spin systems which exhibit strong spatial mixing on general degree-bounded graphs or arbitrary graphs by the uniqueness of Gibbs measure on infinite regular trees.

Theorem 1.

For any finite Δ2 or Δ=, a two-state anti-ferromagnetic spin system is of strong spatial mixing on graphs of maximum degree at most Δ if and only if the system exhibits uniqueness on infinite d-regular trees for all dΔ.

Due to Weitz’s self-avoiding walk tree construction [26], the strong spatial mixing of a two-state spin system on degree-bounded graphs immediately implies an FPTAS for the partition function. Indeed, we show an even stronger notion of correlation decay introduced in a previous work [15], namely the computationally efficient correlation decay, which gives FPTAS not only for the degree-bounded graphs but also for arbitrary graphs with unbounded degrees, when the corresponding uniqueness condition is satisfied.

Theorem 2.

For any finite Δ3 or Δ=, there exists an FPTAS for the partition function of the two-state anti-ferromagnetic spin system on graphs of maximum degree at most Δ if for all dΔ the system parameters lie in the interior of the uniqueness region of infinite d-regular tree.

In the above two theorems, the Δ= case represents the graphs of unbounded degree.

Due to a very recent hardness results of Sly and Sun [24] for general two-state spin systems and an independent result of Galanis, S̆tefankovic̆ and Vigoda [7] for a less general setting, violating the uniqueness condition implies inapproximability of the partition function.

Theorem 3 (Due to [24] and [7]).

For any finite Δ3 or Δ=, unless NP=RP there does not exist an FPRAS for the partition function of the two-state anti-ferromagnetic spin system on graphs of maximum degree at most Δ if for some dΔ the system parameters lie in the interior of the non-uniqueness region of infinite d-regular tree.

The original theorem in [24] holds for d-regular graphs with fixed d, which immediately implies the hardness for degree-bounded graphs or arbitrary graphs. And the hardness condition in both [24] and [7] requires the non-uniqueness on a d-regular tree with d3. But the uniqueness on the infinite 2-regular tree (i.e. the infinite path) always holds for any two-state anti-ferromagnetic spin system, thus the condition in Theorem 3 suffices.

For graphs of maximum degree 2 or less, the partition function can be computed exactly in polynomial time. Thus Theorem 2 and Theorem 3 together with the FPRAS for two-state ferromagnetic spin system [12, 10] give an almost complete (except at the phase transition boundary) classification of the approximability of the partition function of all two-state spin systems, on either degree-bounded families of graphs or family of all graphs.

1.1.1 Regularity and monotonicity

In Statistical Physics, the correlation decay is usually studied on regular graphs or even structurally symmetric graphs (e.g. Bethe lattice). Although for hardness results considering only regular graphs will strengthen the lower bounds, from algorithmic perspective it is more general to consider spin systems on general graphs. The approximation algorithms in [26] for the hardcore model and [21] for the anti-ferromagnetic Ising model are both for general graphs with bounded maximum degrees. As discussed in [21, 24], up to translation of parameters the hardcore and Ising models together are complete for general two-state spin systems on d-regular graphs with fixed d. However, this does not cover the most general case, namely the general two-state spin system on general graphs, of either bounded or unbounded degree. A fundamental reason for this gap is the non-monotonicity of general spin systems.

The well studied hardcore and anti-ferromagnetic Ising models (along with all two-state spin systems with β,γ1) are both monotone spin systems, in the sense that the uniqueness on infinite d-regular tree implies the uniqueness on all infinite regular trees with smaller degrees. This monotonicity does not necessarily hold in general two-state spin systems.

In [26], Weitz established the following implication in the hardcore model:

Claim (Theorem 2.3 in [26]).

Strong spatial mixing on a d-regular tree implies the strong spatial mixing on all graphs of maximum degree at most d.

In [26], Weitz also remarked without proof that this implication holds for all two-state spin systems (indeed this is rigorously proved for anti-ferromagnetic Ising model in [21]). With this to be true, devising approximation algorithms for two-state spin systems on degree-bounded graphs is reduced to verifying the strong spatial mixing on d-regular trees. This has been accepted as a fact about the two-state spin systems and has become a building block for approximation algorithms for such systems (see Theorem 2.8 in [21] and the discussions in [22, 23]). It was also raised as a conjecture in [22] whether the claim holds for general multi-state spin systems.

As a byproduct of our analysis (see Section 5), we find that this well-believed implication between the strong spatial mixing on d-regular tree and on graphs of maximum degree at most d holds only for monotone spin systems (including the hardcore and anti-ferromagnetic Ising models). For general two-state spin systems the worst case for uniqueness as well as strong spatial mixing among all degree-bounded graphs is indeed a regular tree, but may no longer be the one of the highest degree. A bright side of this complication is that higher degrees may yield much faster correlation decay, making possible the FPTAS for graphs with unbounded degrees.

These new phenomena suggest that the general two-state spin systems have much richer structure than the well-studied monotone spin systems such as the hardcore and anti-ferromagnetic Ising models. The former approach via the strong spatial mixing on d-regular tree which succeeds in monotone spin systems on general graphs and general spin systems on regular graphs, meets a barrier when dealing with general spin systems on general graphs. We give a unified approach to the correlation decay in general two-state spin systems, through the strong spatial mixing on arbitrary trees instead of d-regular trees. This approach was initiated in our previous work [15] dealing with graphs of unbounded degrees. In this paper, we devise a unified potential-based analysis which adapts to both the irregularity of the arbitrary tree and the non-monotonicity of general two-state spin system and give tight correlation decay results for all two-state anti-ferromagnetic spin systems on degree-bounded families of graphs and family of all graphs.

1.1.2 Implications of the main results

By solving the uniqueness condition we can restate our main results in various threshold forms. Theorem 2 covers as special cases all previous algorithmic results for two-state anti-ferromagnetic spin systems on general-structure graphs as well as clears up previously uncovered cases.

In terms of interactions.

We can fix the external field λ and discuss the tractable region of (β,γ). Since the roles of β and γ are symmetric, we can further fix one of them and discuss the tractable range of the other. This formulation was used in [10, 15].

Our main result can be restated as follows: for any Δ, there is a critical threshold γc(β,λ,Δ) for the uniqueness on infinite regular trees up to degree Δ such that if γc(β,λ,Δ)<γ<1β there is an FPTAS for graphs of maximum degree at most Δ; and in particular, γc=γc(β,λ,)>1 is an absolute constant such that if γ(γc,1β) there is an FPTAS for arbitrary graphs.

This covers as special cases all algorithmic results in [10] regarding the anti-ferromagnetic systems and all results in [15], extends the tractable regions in these previous works, and considers the degree-bounded graphs as well.

In terms of external field.

Motivated by the studies of hardcore and anti-ferromagnetic Ising models, we can fix (β,γ) and discuss the tractable range of external field.

Due to the symmetric role of β and γ, we may assume that βγ. Our main results can be restated in specific settings as follows:

  • Hard constraints (when β=0): For any Δ, λc(γ,Δ)=min1<d<Δγd+1dd(d1)d+1 is a critical threshold for the uniqueness on infinite regular trees up to degree Δ such that if λ<λc(γ,Δ) there exists an FPTAS for graphs of maximum degree at most Δ.

    For γ1, the critical threshold equals λc(γ,Δ)=γΔ(Δ1)Δ1(Δ2)Δ. There is no external field λ>0 satisfying uniqueness on infinite regular trees of unbounded degrees. This is consistent with the hardness result for the hardcore model without degree bound [4, 23]. One particularly interesting case is when γ=1, in which case the model is the hardcore model with fugacity λ, and λc(1,Δ)=(Δ1)Δ1(Δ2)Δ is the critical threshold. This covers the result of [26].

    For γ>1, in addition to the results for degree-bounded graphs, there exists an absolute positive constant λc(γ)=mind>1γd+1dd(d1)d+1 which lower bounds λc(γ,Δ) for all Δ, such that if λ<λc(γ) there exists an FPTAS for arbitrary graphs.

  • Soft constraints (when β>0): If βγ>Δ2Δ the uniqueness condition holds for any external field λ>0 and there always exists an FPTAS for graphs of maximum degree at most Δ.

    Now suppose βγΔ2Δ. Let Δ¯ be the smallest d satisfying βγd1d+1. For every integer Δ¯d<Δ, there exist two thresholds λ1(β,γ,d) and λ2(β,γ,d) for the uniqueness on the infinite (d+1)-regular tree, such that if λ<λ1(β,γ,d) or λ>λ2(β,γ,d) for all integers Δ¯d<Δ, there is an FPTAS for graphs of maximum degree at most Δ.

    Furthermore, if γ1, the above uniqueness condition can be simplified as that λ<λc or λ>λ¯c, where λc=λc(β,γ,Δ)minΔ¯d<Δλ1(β,γ,d) and λ¯c=λ¯c(β,γ,Δ)maxΔ¯d<Δλ2(d). In particular, for β=γ, the anti-ferromagnetic Ising model, we have λ¯c=1/λc, and the uniqueness holds when |logλ|>logλ¯c. This covers the result of anti-ferro-Ising model in [21].

    For unbounded maximum degree Δ, if γ1, there is no λ>0 satisfying the uniqueness on infinite regular trees of all degrees, which is consistent with the hardness results in [10]; and if γ>1, the uniqueness for all infinite regular trees holds when λ<λ1(β,γ,d) or λ>λ2(β,γ,d) for all dΔ¯, under which condition there exists an FPTAS for arbitrary graphs.

1.2 Related works

The approximation of partition functions of spin systems has been extensively studied[12, 13, 9, 3, 11, 25, 4, 5, 17]. In a seminal work [12], Jerrum and Sinclair devised a fully polynomial-time randomized approximation scheme (FPRAS) for the ferromagnetic Ising model. Later in [10], the FPRAS was extended to all two-state ferromagnetic spin systems by translating the parameters to the ferromagnetic Ising model. Also in [10], Goldberg, Jerrum, and Paterson gave an FPRAS and inapproximability results for two-state anti-ferromagnetic spin systems on arbitrary graphs. A gadget based on random regular bipartite graphs was proposed by Dyer, Frieze, and Jerrum in [4] and was also analyzed by Mossel, Weitz, and Wormald in [19] to study the inapproximability on degree-bounded graphs. It is widely believed that the transition of approximability of anti-ferromagnetic spin systems is captured by the phase transition of uniqueness on infinite trees. This was raised openly as a conjecture in [19]. The conjecture was proved by Sly in [23] for the hardcore model. This result was improved by Galanis et al. in [6] to a wide range of parameters. Very recently, Sly and Sun [24] proved the hardness of all two-state anti-ferromagnetic spin systems of non-uniqueness on infinite regular trees. A same result for anti-ferromagnetic Ising model without external field was independently proved by Galanis, S̆tefankovic̆, and Vigoda in [7].

The correlation decay technique developed independently by Weitz [26] and Bandyopadhyay and Gamarnik [1] is a powerful tool for devising deterministic fully polynomial-time approximation schemes (FPTAS) for partition functions (other important examples include [8, 2]). In [26], Weitz introduced the concept of strong spatial mixing and used it to devise FPTAS for the hardcore model up to the uniqueness threshold. The other most important two-state anti-ferromagnetic spin system, the anti-ferromagnetic Ising model, was studied recently by Sinclair, Srivastava, and Thurley in [21], where a more powerful message-decay method was introduced to analyze the strong spatial mixing and give FPTAS up to uniqueness threshold. A powerful technique was developed by Restrepo et al. in [20] which makes use of the specific structure of graphs for strong spatial mixing. A broader tractable region than the region of uniqueness is achieved on grid lattice by exploiting the structure of the graph. In a previous work [15], we gave an FPTAS for two-state anti-ferromagnetic spin systems without external field on arbitrary graphs with unbounded degrees, up to a continuous relaxation of the uniqueness threshold. The approach used in the current paper was initiated in [15], however the analysis in [15] cannot separate the uniqueness up to certain degree, thus fails in dealing with degree-bounded families of graphs.

2 Definitions and preliminaries

A two-state spin system is described by a graph G=(V,E). A configuration of the system is one of the 2|V| possible assignments σ:V{0,1} of states to vertices. We also use two colors blue and green to denote these two states. The weight of a configuration can be described as a product of contributions of individual edges and vertices. Let A=[A0,0A0,1A1,0A1,1] and b=(b0,b1). The weight of a configuration σ:V{0,1} is given by

w(σ)=(u,v)EAσ(u),σ(v)vVbσ(v).

The Gibbs measure is a probability distribution over all configurations defined by ρ(σ)=w(σ)Z(G). The normalization factor Z(G)=σw(σ) is called the partition function.

We can normalize the contributions of a {𝚋𝚕𝚞𝚎,𝚐𝚛𝚎𝚎𝚗} edge and of a green vertex to be 1. So that A=[β11γ] for some β,γ0, and b=(b0,b1)=(λ,1) for some λ>0. Since the roles of blue and green are symmetric, we can assume that βγ without loss of generality. The three parameters (β,γ,λ) with 0βγ and λ>0 completely specify a two-state spin system. A two-state spin system with β=γ is an Ising model and a two-state spin system with β=0,γ=1 or symmetrically β=1,γ=0 is a hardcore model.

A two-state spin system is called anti-ferromagnetic if adjacent vertices favor disagreeing spins, i.e. if βγ<1. Without loss of generality, we focus on the cases that βγ.

Definition 4.

(β,γ,λ) is anti-ferromagnetic if 0βγ, γ>0, βγ<1, and λ>0.

By the symmetry of β and γ and the triviality of the case β=γ=0, this definition is complete for all nontrivial two-state anti-ferromagnetic systems

2.1 Correlation decay

The Gibbs measure defines a marginal distribution of state for each vertex. Let pv denote the probability of vertex v to be colored blue. Let σΛ{0,1}Λ be a configuration of ΛV. We call vertices vΛ fixed vertices, and vΛ free vertices. We use pvσΛ to denote the marginal probability of v to be colored blue conditioning on the configuration of Λ being fixed as σΛ.

Definition 5.

A spin system on a family of graphs is said to be of strong spatial mixing if for any graph G=(V,E) in the family, any vV,ΛV and σΛ,τΛ{0,1}Λ,

|pvσΛpvτΛ|exp(Ω(dist(v,S))),

where SΛ is the subset on which σΛ and τΛ differ, and dist(v,S) is the shortest distance from v to any vertex in S.

The weak spatial mixing can be defined by measuring the decay with respect to dist(v,Λ) instead of dist(v,S). The spatial mixing property is also called correlation decay in Statistical Physics.

Let T be a tree rooted by v. Define RTσΛ=pvσΛ/(1pvσΛ) to be the ratio between the probabilities that the root v is blue and green, when imposing the condition σΛ (with the convention that RTσΛ= when pvσΛ=1). Suppose that v has d children and Ti is the subtree rooted by the i-th child. Due to the independence of subtrees, we have an easy recursion for calculating RTσΛ:

RTσΛ =λi=1dβRTiσΛ+1RTiσΛ+γ. (1)

Let G(V,E) be a graph. Similarly define that RG,vσΛ=pvσΛ/(1pvσΛ). In contrast to the case of tree, there is no easy recursion for calculating RG,vσΛ for a general graph G because of the dependencies caused by cycles. In [26], a construction called the self-avoiding walk (SAW) tree was introduced which reduces the computing of marginal distribution in a general graph to that in a tree. Specifically, given a graph G(V,E) and a vertex vV. The SAW tree TSAW(G,v) is a tree rooted at v with a new vertex set VSAW (which effectively enumerates all paths originating from v in G and may include fixed leaves). Moreover, any vertex sets SΛV are mapped to respective SSAWΛSAWVSAW and any configuration σΛ{0,1}Λ is mapped to a σΛSAW{0,1}ΛSAW. We abuse the notation and write S=SSAW and σΛ=σΛSAW if no ambiguity is caused. Given a graph G(V,E), vV and SV, let distG(v,S) be the shortest distance in G from v to any vertex in S.

Theorem 6 (Theorem 3.1 of Weitz [26]).

Let G(V,E) be a graph, vV, σΛ{0,1}Λ a configuration on ΛV, and SV. Let T=TSAW(G,v). It holds that the maximum degree of T equals the maximum degree of G, distG(v,S)=distT(v,S), and RG,vσΛ=RTσΛ. Moreover, any neighborhood of v in T can be constructed in time proportional to the size of the neighborhood.

2.2 The uniqueness condition

We consider the uniqueness of Gibbs measure on the infinite (d+1)-regular trees 𝕋^d, in which the recursion is given by fd(x)λ(βx+1x+γ)d due to the symmetric structure of 𝕋^d.

Let x^d be the positive fixed point of fd(x), that is, x^d=f(x^d). It is known [14, 18] that the two-state anti-ferromagnetic spin system on 𝕋^d undergoes a phase transition at |fd(x^d)|=1 with uniqueness when |fd(x^d)|1. This motivates the following definition.

Definition 7.

Let (β,γ,λ) be anti-ferromagnetic. Let x^d be the positive fixed point of function fd(x)=λ(βx+1x+γ)d. We say that (β,γ,λ) is up-to-Δ unique, if for all integers 1d<Δ,

|fd(x^d)|=λd(1βγ)(βx^d+1)d1(x^d+γ)d+1=d(1βγ)x^d(βx^d+1)(x^d+γ)<1.

In particular, (β,γ,λ) is universally unique if it is up-to- unique.

Being up-to-Δ unique is equivalent to that the system is of weak spatial mixing on infinite regular trees up to degree Δ. The uniqueness condition can be described in various threshold forms, which are given in Appendix A.

The uniqueness is defined by requiring that |fd(x^d)|<1. The following lemma states that |fd(x^d)| is bounded by an absolute constant as long as the uniqueness condition holds.

Lemma 8.

Let (β,γ,λ) be anti-ferromagnetic. If (β,γ,λ) is up-to-Δ unique then there exists an absolute constant c<1 which depends only on β, γ, λ and Δ, such that |fd(x^d)|c for all 1d<Δ.

Proof.

The lemma holds trivially for finite Δ. It then remains to show that in case of universal uniqueness, |fd(x^d)| cannot be arbitrarily close to 1 as d grows to infinity. If (β,γ,λ) is universally unique, due to Lemma 21.2, we must have γ>1. For anti-ferromagnetic (β,γ,λ), β1γ, thus the fixed point x^d=λ(βx^d+1x^d+γ)dλγd, therefore |fd(x^d)|=d(1βγ)x^d(βx^d+1)(x^d+γ)dλγd. The lemma follows.

3 The strong spatial mixing on general graphs

In this section we prove Theorem 1. The necessity of the uniqueness condition is trivial since strong spatial mixing on general graphs implies weak spatial mixing on regular trees. It then remains to prove the following theorem.

Theorem 9.

Let (β,γ,λ) be anti-ferromagnetic. For any finite Δ2 or Δ=, if (β,γ,λ) is up-to-Δ unique, then the two-state spin system of parameters (β,γ,λ) is of strong spatial mixing on graphs of maximum degree at most Δ.

Our approach is to prove the strong spatial mixing on arbitrary trees of maximum degree at most Δ, which by the self-avoiding walk tree construction implies the theorem. Note that unlike in [26] and [21], we analyze the decay on arbitrary tree instead of regular tree. This is because for general two-state anti-ferromagnetic spin systems the worst case for strong spatial mixing among all graphs of maximum degree at most d may no longer be the d-regular tree. We will explain this in detail in Section 5.

Given any graph G(V,E) of maximum degree at most Δ, any configuration σΛ{0,1}Λ on ΛV and any SΛ, fixing an arbitrary vertex vV, by Theorem 6, a self-avoiding walk tree T=TSAW(G,v) can be constructed such that the maximum degree of T is bounded by Δ, distG(v,S)=distT(v,S) and RG,vσΛ=RTσΛ. Recall that RG,vσΛ=pvσΛ/(1pvσΛ) thus pvσΛ=RG,vσΛ/(1+RG,vσΛ). For any σΛ and τΛ,

|pvσΛpvτΛ|=|RG,vσΛ1+RG,vσΛRG,vτΛ1+RG,vτΛ||RG,vσΛRG,vτΛ|=|RTσΛRTτΛ|.

This motivates the following definition.

Definition 10.

Let T be a tree rooted by vertex v, τΛ{0,1}Λ be a configuration on ΛV and SΛ be a vertex set. Define Rv and δv as that RvRTτΛRv+δv for all σΛ{0,1}Λ which differ from τΛ only on S.

It is then sufficient to prove Theorem 9 by constructing such Rv and δv for T=TSAW(G,v) and showing that δvexp(Ω(dist(v,S))).

Let T be a tree rooted by v, who has d children v1,,vd, and Ti be the subtree rooted by vi. It holds that

RTσΛ =f(RT1σΛ,,RTdσΛ)λi=1dβRTiσΛ+1RTiσΛ+γ.

The lower and upper bounds Rv and Rv+δv can be recursively constructed as follows. The base cases are: (1) vS, in which case Rv=0 and δv=; and (2) vΛS, i.e. v is fixed to be the same color in all σΛ, in which case δv=0 and Rv= (or Rv=0) if v is fixed to be blue (or green). For vΛ, since f(R1,,Rd) is monotonically decreasing for anti-ferromagnetic (β,γ,λ),

Rv =f(Rv1+δv1,,Rvd+δvd),
Rv+δv =f(Rv1,,Rvd),

where Rvi and Rvi+δvi are the corresponding lower and upper bounds for RTiσΛ, 1id. In particular, when d=0 the empty product equals 1 by convention, thus Rv=Rv+δv=λ, which is consistent with the case that v is a free vertex having no children.

By the monotonicity of f(R1,,Rd), it is easy to check that the Rv and Rv+δv constructed above satisfy the requirement of Definition 10. Our goal is to show that δv decays exponentially in depth of recursion when the uniqueness holds. A straightforward approach is trying to prove that δ contracts at a constant rate in each recursion step. But this does not have to be true to guarantee the exponential decay. Indeed there are cases that the error does not decay in single steps but decay in a long run. To overcome this, we use a potential Φ to amortize the contraction and show that δΦ contracts at a constant rate. We choose the potential function to be

Φ(R)=1R(βR+1)(R+γ).

We are analyzing the decay on an arbitrary tree with irregular degrees. In order to adapt this irregularity, the potential function cannot have d as an input, but only caries the information regarding the distribution at the current vertex, yet it has to be able to provide correct compensation to the step-wise decay at any state of R and for all spin systems satisfying sufficient uniqueness. A heuristic procedure which leads us to this good potential function is discussed in Appendix C.

Let φ(R) be a monotone function satisfying that φ(R)=Φ(R). We define that

ϵvφ(Rv+δv)φ(Rv),

and accordingly, ϵviφ(Rvi+δvi)φ(Rvi), 1id.

We define a function α(d;x1,,xd) as follows:

α(d;x1,,xd) (1βγ)(λi=1dβxi+1xi+γ)12(βλi=1dβxi+1xi+γ+1)12(λi=1dβxi+1xi+γ+γ)12i=1dxi12(βxi+1)12(xi+γ)12.

The following lemma is obtained from applying the Mean Value Theorem. Similar routines were previously used in [15, 20]. The proof of the lemma is in Appendix B.

Lemma 11.

The followings hold for ϵv.

  1. 1.

    (relation to δv) ϵv=δvΦ(R~) for some R~[Rv,Rv+δv].

  2. 2.

    (absolute bound) Assuming that γ>1 or the maximum degree of T is bounded by a constant, if vΛ then Rv+δv=O(1) and if viS for all 1id then ϵv=O(1).

  3. 3.

    (stepwise contraction) There exist Ri~[Rvi,Rvi+δvi], 1id, such that

    ϵvα(d;R~1,,R~d)max1id{ϵvi}.

To prove the strong spatial mixing, we first relate δv to ϵv by Item 1 of Lemma 11, and then apply induction on the depth in T, with Item 2 of Lemma 11 as basis and Item 3 of Lemma 11 as induction step. We then need to bound the contraction rate α(d;x1,,xd). Note that z(βz+1)(z+γ)(1+βγ)21 for z[0,), thus it holds unconditionally for all xi[0,), 1id, that

α(d;x1,,xd) d, (2)
α(d;x1,,xd) dλγd+1. (3)

With the uniqueness, the following much tighter contraction bound can be proved.

Lemma 12.

Let (β,γ,λ) be anti-ferromagnetic. If (β,γ,λ) is up-to-Δ unique, then there exists a constant α<1 such that for any integer 1d<Δ and any x1,,xd[0,+), 1id, it holds that α(d;x1,,xd)α.

This lemma is the technical core of our analysis. It crucially relies on the choice of potential function. Before delving into the formal proof of Lemma 12, we note that Theorem 9 can be implied by this lemma.

Proof of Theorem 9.

Let T=TSAW(G,v) for a G whose maximum degree is at most Δ. Then the maximum degree of T is at most Δ, thus the root v has at most Δ children in T, and every other vertex in T has less than Δ children. We recursively construct Ru, δu and ϵu for every subtree in T.

Let t=dist(v,S). By repeatedly applying Item 3 of Lemma 11, without loss of generality, we have a path u1u2ut2 in T with u1=v such that ϵujα(dj;x1,,xdj)ϵuj+1 for j=1,2,,t3, where dj is the number of children of uj and xi[0,), 1idj.

Note that d1Δ, and dj<Δ for all other j. Assume that (β,γ,λ) is up-to-Δ unique. If Δ is bounded, then by Lemma 12 there exists a constant α<1, such that ϵujαϵuj+1 for 2jt3, and ϵvd1ϵu2Δϵ2 due to (2); and if Δ=, then by Lemma 12, ϵujαϵuj+1 for all 1jt3. In both cases we have ϵv=O(αtϵut2).

Due to Item 1 of Lemma 11, δv=ϵvΦ(R~)=O(1Φ(R~)αtϵut2) for some R~[Rv,Rv+δv]. We then bound Φ(R~) and ϵut2. Due to Item 2 of Lemma 21, the fact that (β,γ,λ) is up-to-Δ unique implies that either Δ is bounded or γ>1. Note that v must be free or the theorem is trivial to prove, and none of ut2’s children is in S because dist(v,S)=t. Thus by Item 2 of Lemma 11, ϵut2=O(1) and R~Rv+δv=O(1), which implies that Φ(R~)=1R~(βR~+1)(R~+γ)=Ω(1).

In conclusion, if (β,γ,λ) is up-to-Δ unique, there exists a constant α<1, such that δv=O(αt). As discussed in the beginning of this section, this proves Theorem 9.

Proof of Lemma 12.

The rest of this section is devoted to the proof of Lemma 12. Given that (β,γ,λ) is up-to-Δ unique, there exists an absolute constant α<1 such that α(d;x1,,xd)α for any 1d<Δ and x1,,xd0.

We define the symmetric version of α(d;x1,,xd):

αd(x) α(d;x,,xd)=d(1βγ)(xλ(βx+1x+γ)d)12(βx+1)12(x+γ)12(βλ(βx+1x+γ)d+1)12(λ(βx+1x+γ)d+γ)12.

The following lemma shows that the symmetric case dominates the maximum of α(d;x1,,xd) by using the inequalities of Cauchy-Schwarz and arithmetic and geometric means.

Lemma 13.

Let (β,γ,λ) be anti-ferromagnetic. Then for any integer d and any x1,,xd[0,+), there exists an x¯[0,+) such that α(d;x1,,xd)α(d,x¯).

Proof.

Let zi=βxi+1xi+γ. Then zi(β,1γ] and xi=1γziziβ. Express α(d;x1,,xd) in terms of zi:

α(d;x1,,xd) =(λi=1dzi)12(βλi=1dzi+1)12(λi=1dzi+γ)12i=1d(zi1γ)12(ziβ)12.

Due to Cauchy-Schwarz inequality,

i=1d(zi1γ)12(ziβ)12d(1di=1d(zi1γ)(ziβ))12=d(1+βγ1di=1d(ziγ+βzi1))12.

Due to the inequality of arithmetic and geometric means,

d(1+βγ1di=1d(ziγ+βzi1))12 d(1+βγγ(i=1dzi)1dβ(i=1dzi)1d)12.

Let z¯=(i=1dzi)1d. Then combining the above calculations,

α(d;x1,,xd) (λz¯d)12d(1+βγγz¯βz¯1)12(βλz¯d+1)12(λz¯d+γ)12=dλz¯d(z¯1γ)(z¯β)(βλz¯d+1)(λz¯d+γ).

Let x¯ be such that βx¯+1x¯+γ=z¯. Then x¯[0,+) and by substituting βx¯+1x¯+γ for z¯, we have

α(d;x1,,xd) d(1βγ)(x¯λ(βx¯+1x¯+γ)d)12(βx¯+1)12(x¯+γ)12(βλ(βx¯+1x¯+γ)d+1)12(λ(βx¯+1x¯+γ)d+γ)12=αd(x¯).

Lemma 14.

Let (β,γ,λ) be anti-ferromagnetic. If (β,γ,λ) is up-to-Δ unique, then there exists a constant α<1 such that for any integer 1d<Δ, it holds that αd(x)α for all x0.

Proof.

Fix d to be any positive integer. We characterize the value of x at which αd(x) achieves its maximum. We denote that fd(x)=λ(βx+1x+γ)d. Taking derivative of αd(x) with respect to x, we get that

αd(x)=d(1βγ)G(x)2G(x),

where G(x)=xfd(x)(βfd(x)+1)(fd(x)+γ)(βx+1)(x+γ), whose derivative is

G(x) =fd(x)d(1βγ)x(βfd(x)+1)(fd(x)+γ)(βx+1)2(x+γ)2(γβx2d(1βγ)xγβfd(x)2(βfd(x)+1)(fd(x)+γ)).

As x ranges over [0,), the function γβx2d(1βγ)x is strictly decreasing in x and ranges from + to , and the function γβfd(x)2(βfd(x)+1)(fd(x)+γ) is strictly increasing in x and has a bounded range. Thus, the equation

γβx2d(1βγ)x =γβfd(x)2(βfd(x)+1)(fd(x)+γ). (4)

has unique solution in (0,), denoted by xd. Moreover, it holds that

G(x){>0if 0x<xd,=0if x=xd,<0if x>xd. (5)

The same also holds for αd(x). Thus, for any fixed d, αd(x) achieves its maximum when x=xd.

Therefore, for all x0,

αd(x)αd(xd) =d(1βγ)(xdfd(xd)(βxd+1)(xd+γ)(βfd(xd)+1)(fd(xd)+γ))12
=(d(1βγ)(γβxd2)(βxd+1)(xd+γ)fd(xd)(γβfd(xd)2))12 (6)
α~d(xd).

Equation (6) is obtained by substituting (βfd(xd)+1)(fd(xd)+γ) according to (4).

Let x^d be the positive fixed point of fd(x), that is, x^d=fd(x^d). We then claim that if (β,γ,λ) is up-to-Δ unique, then α~d(xd)α~d(x^d) for any integral 1d<Δ, To see that this claim is sufficient to imply the lemma, note that after substituting x^d=fd(x^d), we have α~d(x^d)=d(1βγ)x^d(βx^d+1)(x^d+γ)=|fd(x^d)|. And due to Lemma 8, if (β,γ,λ) is up-to-Δ unique then there exists a constant c<1 such that |fd(x^d)|<c for all integer 1d<Δ.

We then prove the claim. Assume that (β,γ,λ) is up-to-Δ unique and 1d<Δ. It is then sufficient to show that α~d(x) is decreasing if x^dxd and is increasing if x^d>xd.

Case 1: x^dxd. Due to (5), G(x^d)0. Note that

G(x^d)=d(1βγ)(γβx^d2)x^d2(βx^d+1)3(x^d+γ)3(1d(1βγ)x^d1(βx^d+1)(x^d+γ)).

Due to the uniqueness, |fd(x^d)|=d(1βγ)x^d(βx^d+1)(x^d+γ)<1, thus 1d(1βγ)x^d1(βx^d+1)(x^d+γ)>0. Combining with that G(x)0, we have γβx^d20. Since the function fd(x) is monotonically decreasing and x^d is its fixed point, γβfd(xd)2γβfd(x^d)2=γβx^d20. Since xd satisfies (4), γβxd2 and γβfd(xd)2 must be simultaneously positive or negative, thus it also holds that γβxd20. Then both (γβx2)(βx+1)(x+γ) and fd(x)(γβfd(x)2) are positive and monotonically decreasing in x[x^d,xd]. Therefore, α~d(xd)α~d(x^d).

Case 2: x^d>xd. By the same argument as above, it holds that γβfd(x^d)2=γβx^d2<0, γβfd(xd)2<0, and γβxd2<0. Thus both (γβx2)(βx+1)(x+γ) and fd(x)(γβfd(x)2) are negative and monotonically decreasing in x[xd,x^d], hence their product is positive and increasing in x[xd,x^d]. Therefore, α~d(xd)α~d(x^d).

Lemma 12 is proved by combining Lemma 13 and 14. This completes the proof of Theorem 9.

Strong spatial mixing on regular trees.

As a byproduct of our analysis, we prove a strong spatial mixing theorem for regular trees. When the graphs G itself is a regular tree. All vertices (except the root) has the same arity. And all ds (excerpt the one of the root) that appear in the proof are the same and equal that arity. Then the condition that the uniqueness holds on all infinite regular trees of degree up to Δ can be replaced by the uniqueness on infinite Δ-regular tree.

Theorem 15.

For two-state anti-ferromagnetic spin systems, on any infinite Δ-regular tree the uniqueness implies the strong spatial mixing.

The same result can be obtained by combining the same theorem for the hardcore model [26] and anti-ferromagnetic Ising model [21] and translating the parameters of general two-state anti-ferromagnetic spin systems to these models (as discussed in[24, 21]). However, unlike the hardcore and the anti-ferromagnetic Ising models, for general two-state anti-ferromagnetic spin systems Theorem 15 itself is not sufficient to imply the strong spatial mixing on graphs of maximum degree at most Δ. This is discussed in Section 5.

4 Algorithmic implications

In this section we prove Theorem 2. That is, if an anti-ferromagnetic (β,γ,λ) is up-to-Δ unique then there exists an FPTAS for the partition function Z(G) for any graph G of maximum degree at most Δ, and in particular the universal uniqueness implies an FPTAS for arbitrary graph G.

It is well-known that Z(G) can be computed from pvσΛ by the following standard procedure. Let v1,,vn enumerate the vertices in G. For 0in, let σi be the configuration fixing the first i vertices v1,,vi as follows: σi(vj)=σi1(vj) for 1ji1 and σi(vi) is fixed so that piPr[σi(vi)σi1]1/3. In particular, σn{0,1}V is a configuration of V. It holds for the Gibbs measure of σn that ρ(σn)=p1p2pn as well as that ρ(σn)=w(σn)Z(G), thus Z(G)=w(σn)p1p2pn, where the weight w(σn)=(u,v)EAσn(u),σn(v)vVbσn(v) can be computed precisely for any particular σn in time polynomial in n. Note that pi equals to either pviσi1 or 1pviσi1. Therefore, if pvσΛ can be approximated within an additive error ϵ in time polynomial in n and 1ϵ, then the configurations σi can be efficiently constructed such that all pi are bounded away from 0, thus the product p1p2pn can be approximated within a factor of (1±nϵ) in time polynomial in n and 1ϵ, which implies an FPTAS for Z(G).

Bounded degree graphs.

Let G be a graph whose maximum degree is at most Δ and v be any vertex. A self-avoiding walk tree T=TSAW(G,v) can be constructed so that RG,vσΛ=RTσΛ. We can use the recursive procedure described in Section 3 to compute the upper and lower bounds of RTσΛ, with the setting that for all the vertices more than t steps away from the root v, the trivial bounds 0RTσΛ is used. Then the proof of Theorem 9 shows that the recursive procedure returns R0 and R1 such that R0RTσΛR1, and R1R0=O(αt) for some constant α<1 assuming that (β,γ,λ) is up-to-Δ unique. Note that RTσΛ=RG,vσΛ=pvσΛ1pvσΛ. Let p0=R0R0+1 and p1=R1R1+1. Then p0pvσΛp1 and

p1p0=R1R1+1R0R0+1R1R0=O(αt). (7)

The recursive procedure runs in time O(Δt) since it only needs to construct the first t levels of the self-avoiding walk tree. If Δ is bounded, by setting t=logαϵ, this gives an algorithm which approximates pvσΛ within an additive error O(ϵ) in time polynomial in n and 1ϵ, which implies an FPTAS for Z(G).

Arbitrary graphs.

Let G be an arbitrary graph and v be any vertex. Let T=TSAW(G,v). We use the method of Computationally Efficient Correlation Decay introduced in [15] to deal with the unbounded degrees. Intuitively, using this method we observe correlation decay in a refined metric instead of graph distance such that in this new metric a neighborhood of moderate size is sufficient to guarantee desirable correlation decay.

Similarly, we use the recursive procedure described in Section 3 to compute the upper and lower bounds of RTσΛ, but this time the termination condition relies on a new depth defined as follows.

Definition 16.

Let T be a rooted tree and M>1 be a constant. For any vertex v in T, define the M-based depth of v, denoted M(v), as such: M(v)=0 if v is the root, and M(v)=M(u)+logM(d+1) if v is one of the d children of u.

Let M>1 to be fixed. Denote by B() the set of all vertices with M-based depth < along with their children and grandchildren in T. It can be verified by induction that |B()|n2M. The recursion is applied to estimate the RTσΛ when the current vB() until v is no longer in B() in which case the trivial bounds 0RTσΛ is used.

Let ϵv be defined as in Section 3. Repeatedly applying Item 3 of Lemma 11, without loss of generality, we have a path u1u2uk in T from the root u1=v to a uk with M(uk) and M(uk1)<, such that ϵujα(dj;x1,,xdj)ϵuj+1 for j=1,2,,k, where dj is the number of children of uj and xi[0,), 1idj. The key to overcome the explosion caused by unbounded degrees is to observe that the contraction α(d;x1,,xd) decreases dramatically as the degree d grows.

Lemma 17.

Let (β,γ,λ) be anti-ferromagnetic. If (β,γ,λ) satisfies the universal uniqueness, then there exist constants α<1 and M>1 such that for any integer d1, and any xi[0,), 1id, it holds that α(d;x1,,xd)αlogM(d+1).

Proof.

Assume the universal uniqueness of (β,γ,λ). Due to Lemma 12, there exists a constant α<1 such that α(d;x1,,xd)α. By Item 2 of Lemma 21, the universal uniqueness implies that γ>1, thus there exists a constant M>1 such that dλγd+1αlogM(d+1) for all dM. Due to (3), it holds that α(d;x1,,xd)dλγd+1αlogM(d+1) for all dM. Note that α(d;x1,,xd)α means that α(d;x1,,xd)αlogM(d+1) for d<M. Therefore, α(d;x1,,xd)αlogM(d+1) for all d. ∎

By Lemma 17, there exist constants α<1 and M>1 such that

ϵv ϵukj=1kαlogM(dj+1)ϵukαj=1klogM(dj+1)=ϵukαM(uk)ϵukα.

With the notation used in Section 3, S is the complement of B(). Note that all uk’s children are in B() thus none of them are in S, and by Item 2 of Lemma 21 the universal uniqueness implies that γ>1. Thus by Item 2 of Lemma 11 it holds that ϵuk=O(1). Therefore, ϵvϵukα=O(α).

Let δv=R1R0, where R0 and R1 are the bounds returned by the recursive procedure such that R0RTσΛR1. By the same analysis as in Section 3, δv=ϵvΦ(R~)=O(ϵv)=O(α). Then by (7), the marginal probability pvσΛ is approximated within an additive error of O(α). The running time of the recursion is O(nB())=O(n3M). By setting =logαϵ, we have an algorithm which approximates pvσΛ within an additive error of O(ϵ) in time polynomial in n and 1ϵ, which implies an FPTAS for Z(G) for arbitrary graph G.

Heterogeneous spin systems.

Our analysis in last and this sections actually holds for heterogeneous spin systems which allow that each vertex v has a distinct constant external field λv>0.

Theorem 18.

For a two-state anti-ferromagnetic heterogeneous spin system with parameters β, γ, and external field λv at each vertex v, for any finite Δ2 or Δ=, if for all v the (β,γ,λv) is up-to-Δ unique then the spin system is of strong spatial mixing and has FPTAS on graphs of maximum degree at most Δ.

5 Non-monotonicity of general two-state spin system

In this section, we prove the following theorem.

Theorem 19.

There exist two-state anti-ferromagnetic spin systems which exhibit strong spatial mixing on infinite d-regular tree but does not exhibit strong spatial mixing on all graphs of maximum degree at most d.

In a seminal work [26], Weitz proved that for the hardcore model the strong spatial mixing on an infinite d-regular tree implies the strong spatial mixing on graphs of maximum degree at most d (Theorem 2.3 in [26]). He further remarked that the same implication holds for all two-state spin systems. That is,

for any two-state spin system, strong spatial mixing on an infinite d-regular tree implies the strong spatial mixing on graphs of maximum degree at most d.

This claim played important roles in current understanding of correlation decay in two-state spin systems as well as devising FPTAS for such systems. An algorithmic form of this claim was cited in [21] as a theorem for all two-state spin systems (Theorem 2.8 in [21]) and was proved for the anti-ferromagnetic Ising model. It was raised as a conjecture in [22] whether this claim holds for multi-state spin systems.

Here we clarify that this claim holds only for the two-state spin systems under limited settings but does not hold for all general two-state spin systems. This disproves the conjecture in [22] and shows that the common belief that the d-regular tree represents the worst case for strong spatial mixing among all graphs of maximum degree at most d cannot be generalized to general two-state spin systems. We first describe a region that the claim is true.

Lemma 20.

For 0β,γ1, the strong spatial mixing on infinite d-regular tree implies the strong spatial mixing on trees of maximum degree at most d.

Proof.

Given a rooted tree of maximum degree at most d, for each vertex of k children with k<d1, we can attach d1k dummy children with fixed (distributions of) spin states. This is the method used in [26] and [21]: for the hardcore model the dummy children are fixed to be unoccupied and for the anti-ferromagnetic Ising model the dummy children are of uniform distributions over spin states. In both cases, the dummy children have no effect on their parent. In general, we fix the distribution to be (p0,p1) at each dummy child satisfying

p0+p1 =1,
βp0+p1 =p0+γp1.

When 0β,γ1, this system has solutions in 0p0,p11. With the ratio Ri at the i-th child, 1ik, and Ri=p0/p1 for the dummy children k<id1, the ratio at the parent is given by the recursion

λi=1d1βRi+1Ri+γ=λi=1kβRi+1Ri+γ,

which is identical to the original quantity.

Due to the self-avoiding walk tree construction (Theorem 6), it holds that for 0β,γ1, strong spatial mixing on infinite d-regular tree implies the strong spatial mixing and the FPTAS on graphs of maximum degree at most d.

For 0β,γ1, the spin system shows the following monotone property: the uniqueness on infinite d-regular tree implies the uniqueness on all infinite regular trees of smaller degree. This can be verified by the following reasoning: due to Theorem 15, on infinite d-regular tree the uniqueness implies the strong spatial mixing, which for 0β,γ1, implies the strong spatial mixing (including the uniqueness) on all infinite regular trees of smaller degree.

There exist two-state anti-ferromagnetic spin systems which are non-monotone. We can choose anti-ferromagnetic (β,γ,λ) satisfying that γ>1 and λ>λc𝗎𝗉𝗉𝖾𝗋(β,γ), where λc𝗎𝗉𝗉𝖾𝗋(β,γ) is the upper threshold for universal uniqueness given in Item 8 of Lemma 21. Due to Item 3 of Lemma 21, for γ>1 the uniqueness holds on d-regular trees for all sufficiently large d. On the other hand, due to Item 8 of Lemma 21, (β,γ,λ) cannot be universally unique when λ>λc𝗎𝗉𝗉𝖾𝗋(β,γ), which means that there exists a finite d such that the system is non-unique on d-regular tree.

For such non-monotone systems, due to Theorem 15, the uniqueness implies the strong spatial mixing on d-regular tree for sufficiently large d, but the strong spatial mixing does not hold on a regular tree of smaller degree (because of the non-uniqueness on the smaller tree). Therefore the implication between the strong spatial mixing on d-regular tree and on graphs of maximum degree at most d does not hold for general two-state spin systems.

Acknowledgment.

We would like to thank Alistair Sinclair and Piyush Srivastava for the helpful comments to an early version of this paper.

Acknowledgment for the 2021 revision.

We would like to thank Xiaoyu Chen, Weiming Feng and Xinyuan Zhang for locating an error in Lemma 21 in the original (2013) version of the paper and suggesting its fix.

References

  • [1] A. Bandyopadhyay and D. Gamarnik. Counting without sampling: Asymptotics of the log-partition function for certain statistical physics models. Random Struct. Algorithms, 33:452–479, December 2008.
  • [2] M. Bayati, D. Gamarnik, D. Katz, C. Nair, and P. Tetali. Simple deterministic approximation algorithms for counting matchings. In Proceedings of the thirty-ninth annual ACM Symposium on Theory of Computing, STOC ’07, pages 122–127, New York, NY, USA, 2007. ACM.
  • [3] M. Dyer, M. Jerrum, and E. Vigoda. Rapidly mixing markov chains for dismantleable constraint graphs. In Proceedings of a DIMACS/DIMATIA Workshop on Graphs, Morphisms and Statistical Physics, 2001.
  • [4] M. E. Dyer, A. M. Frieze, and M. Jerrum. On counting independent sets in sparse graphs. SIAM Jounal on Computing, 31(5):1527–1541, 2002.
  • [5] M. E. Dyer and C. S. Greenhill. On markov chains for independent sets. Journal of Algorithms, 35(1):17–49, 2000.
  • [6] A. Galanis, Q. Ge, D. Štefankovič, E. Vigoda, and L. Yang. Improved inapproximability results for counting independent sets in the hard-core model. In The 15th International Workshop on Randomization and Computation, RANDOM ’11, 2011.
  • [7] A. Galanis, D. Stefankovic, and E. Vigoda. Inapproximability of the partition function for the antiferromagnetic ising and hard-core models. Arxiv preprint arXiv:1203.2226, 2012.
  • [8] D. Gamarnik and D. Katz. Correlation decay and deterministic fptas for counting list-colorings of a graph. In Proceedings of the eighteenth annual ACM-SIAM symposium on Discrete algorithms, SODA ’07, pages 1245–1254, Philadelphia, PA, USA, 2007. Society for Industrial and Applied Mathematics.
  • [9] L. A. Goldberg and M. Jerrum. A polynomial-time algorithm for estimating the partition function of the ferromagnetic ising model on a regular matroid. In L. Aceto, M. Henzinger, and J. Sgall, editors, Proceedings of the 38th International Colloquium on Automata, Languages and Programming, volume 6755 of ICALP ’11, pages 521–532. Springer, 2011.
  • [10] L. A. Goldberg, M. Jerrum, and M. Paterson. The computational complexity of two-state spin systems. Random Struct. Algorithms, 23(2):133–154, 2003.
  • [11] M. Jerrum. A very simple algorithm for estimating the number of k-colorings of a low-degree graph. Random Struct. Algorithms, 7(2):157–166, 1995.
  • [12] M. Jerrum and A. Sinclair. Polynomial-time approximation algorithms for the ising model. SIAM Journal on Computing, 22(5):1087–1116, 1993.
  • [13] M. Jerrum, A. Sinclair, and E. Vigoda. A polynomial-time approximation algorithm for the permanent of a matrix with nonnegative entries. Journal of the ACM, 51:671–697, July 2004.
  • [14] F. Kelly. Stochastic models of computer communication systems. Journal of the Royal Statistical Society. Series B (Methodological), 47(3):379–395, 1985.
  • [15] L. Li, P. Lu, and Y. Yin. Approximate counting via correlation decay in spin systems. In Proceedings of the Twenty-Third Annual ACM-SIAM Symposium on Discrete Algorithms, SODA ’12, pages 922–940, Kyoto, Japan, 2012. SIAM.
  • [16] L. Li, P. Lu, and Y. Yin. Correlation Decay up to Uniqueness in Spin Systems. In Proceedings of the Twenty-Fourth Annual ACM-SIAM Symposium on Discrete Algorithms, SODA ’13, pages 67-84, New Orleans, Louisiana, USA, 2013. SIAM.
  • [17] M. Luby and E. Vigoda. Approximately counting up to four (extended abstract). In Proceedings of the twenty-ninth annual ACM symposium on Theory of computing, STOC ’97, pages 682–687, New York, NY, USA, 1997. ACM.
  • [18] F. Martinelli, A. Sinclair, and D. Weitz. Fast mixing for independent sets, colorings, and other models on trees. Random Struct. Algorithms, 31(2):134–172, 2007.
  • [19] E. Mossel, D. Weitz, and N. Wormald. On the hardness of sampling independent sets beyond the tree threshold. Probability Theory and Related Fields, 143:401–439, 2009.
  • [20] R. Restrepo, J. Shin, P. Tetali, E. Vigoda, and L. Yang. Improved mixing condition on the grid for counting and sampling independent sets. In Proceedings of the 52nd Annual IEEE Symposium on Foundations of Computer Science, FOCS ’11, 2011.
  • [21] A. Sinclair, P. Srivastava, and M. Thurley. Approximation algorithms for two-state anti-ferromagnetic spin systems on bounded degree graphs. In Proceedings of the Twenty-Third Annual ACM-SIAM Symposium on Discrete Algorithms, pages 941–953. SIAM, 2012.
  • [22] A. Sly. Uniqueness thresholds on trees versus graphs. The Annals of Applied Probability, 18(5):1897–1909, 2008.
  • [23] A. Sly. Computational transition at the uniqueness threshold. In Proceedings of the 2010 IEEE 51st Annual Symposium on Foundations of Computer Science, FOCS ’10, pages 287–296, Washington, DC, USA, 2010. IEEE Computer Society.
  • [24] A. Sly and N. Sun. The computational hardness of counting in two-spin models on d-regular graphs. Arxiv preprint arXiv:1203.2602, 2012.
  • [25] E. Vigoda. Improved bounds for sampling coloring. In Proceedings of the 40th Annual Symposium on Foundations of Computer Science, FOCS ’99, pages 51–, Washington, DC, USA, 1999. IEEE Computer Society.
  • [26] D. Weitz. Counting independent sets up to the tree threshold. In Proceedings of the thirty-eighth annual ACM symposium on Theory of computing, STOC ’06, pages 140–149, New York, NY, USA, 2006. ACM.

Appendix A The Uniqueness Thresholds

The following lemma translates the uniqueness condition into its various threshold forms. The lemma (except for item 2) is not used in the proofs of the main results but is used in the interpretation of the main results and comparisons to the previous results which are mostly stated in threshold forms.

Lemma 21.

Let (β,γ,λ) be anti-ferromagnetic.

  1. 1.

    (β,γ,λ) is up-to-2 unique.

  2. 2.

    If γ1, then the uniqueness does not hold on infinite d-regular tree for all sufficiently large d.

  3. 3.

    If γ>1, then the uniqueness holds on infinite d-regular tree for all sufficiently large d.

  4. 4.

    For any Δ (including Δ=), there exists a critical threshold γc=γc(β,λ,Δ) such that (β,γ,λ) is up-to-Δ unique if and only if γ(γc,1β). In particular, γc(β,λ,)>1 and γc(β,λ,)=γc(β,λ,Δ) for some finite Δ.

  5. 5.

    If β=0, for any Δ (including Δ=), there exists a critical threshold λc=λc(γ,Δ)=min1<d<Δγd+1dd(d1)d+1 such that (β,γ,λ) is up-to-Δ unique if and only if λ(0,λc).

  6. 6.

    If βγ>Δ2Δ, then for any external field λ, (β,γ,λ) is up-to-Δ unique.

  7. 7.

    If β>0, for any Δ (including Δ=) that βγΔ2Δ, the regime for the λ satisfying up-to-Δ uniqueness is as follows.

    Let Δ¯1+βγ1βγ. For any integer Δ¯d<Δ, let x1(d)x2(d) be the two positive roots of equation d(1βγ)x(βx+1)(x+γ)=1, more specifically,

    x1(d)=1βγ+d(1βγ)(1βγ+d(1βγ))24βγ2β,
    x2(d)=1βγ+d(1βγ)+(1βγ+d(1βγ))24βγ2β.

    Let λi(d)xi(d)(xi(d)+γβxi(d)+1)d, where i=1,2, be defined for integers Δ¯d<Δ.

    Then (β,γ,λ) is up-to-Δ unique if and only if λ belongs to the following regime

    Δ¯d<Δ((0,λ1(d))(λ2(d),)). (8)

    And if furthermore γ1, then (β,γ,λ) is up-to-Δ unique if and only if λ(0,λc)(λ¯c,), where

    λc =λc(β,γ,Δ)minΔ¯d<Δλ1(d),
    λ¯c =λ¯c(β,γ,Δ)maxΔ¯d<Δλ2(d)=λ2(Δ1).

    In particular, when β=γ, it holds that λcλ¯c=1, and thus (β,β,λ) is up-to-Δ unique if and only if |logλ|>logλ¯c.

  8. 8.

    (β,γ,λ) can be universally unique only when γ>1. If γ>1, there exist finite positive constants λc𝗅𝗈𝗐𝖾𝗋=λc𝗅𝗈𝗐𝖾𝗋(β,γ) and λc𝗎𝗉𝗉𝖾𝗋=λc𝗎𝗉𝗉𝖾𝗋(β,γ) where λc𝗅𝗈𝗐𝖾𝗋λc𝗎𝗉𝗉𝖾𝗋, such that (β,γ,λ) is universally unique if λ<λc𝗅𝗈𝗐𝖾𝗋 and (β,γ,λ) is universally unique only if λ<λc𝗎𝗉𝗉𝖾𝗋.

Remark (An error in the SODA’13 version).

In the earlier version of the paper, there was an error in Item 7 of Lemma 21 (which was Lemma 3.1 in the SODA’13 conference version [16]). In that version, the uniqueness regime (8) was unconditionally and wrongly simplified to the bi-interval simple form (0,λc)(λ¯c,), which actually holds more restrictively when both β,γ1. However, when γ>1, there exist such βγ<1 that the uniqueness regime (8) may become more complicated than consisting of at most two intervals. For example, when β=1122, γ=100, and Δ=23, the regime for the λ such that (β,γ,λ) is up-to-Δ unique, consists of three intervals (0,102.664)(104.339,106.967)(109.444,).

The current version corrects the error and changes the statements of Item 7 and Item 8 of Lemma 21. We further remark that the error does not affect the main result of the paper, namely the implication from the up-to-Δ uniqueness to strong spatial mixing and FPTAS, because Lemma 21 is not used in the proofs of the main results (except for Item 2), but is only used for the interpretation of the uniqueness regime in various threshold forms and comparison with other known results.

Proof of Lemma 21.

Let fd(x)=λ(βxd+1xd+γ)d and x^d=fd(x^d) be the positive fixed point of fd(x). Then

|fd(x^d)|=d(1βγ)x^d(βx^d+1)(x^d+γ).

Let (β,γ,λ) be anti-ferromagnetic. That is, 0βγ, γ>0, and βγ<1, thus β<1.

  1. 1.

    It is easy to verify that for βγ<1, (βx+1)(x+γ)(1βγ)x>0 for any positive x>0. Therefore, when d=1, we have |f1(x^1)|=(1βγ)x^1(βx^1+1)(x^1+γ)<1, which means that (β,γ,λ) is up-to-2 unique.

  2. 2.

    For all sufficiently large d, it holds that

    λβdexp(d(1βγ)d3) <d(1βγ)3β, and
    λexp(γdd(1βγ)3) >γd(1βγ)3.

    By contradiction, suppose that |fd(x^d)|=d(1βγ)x^d(βx^d+1)(x^d+γ)1. Then,

    1d(1βγ)x^d(βx^d+1)(x^d+γ)=d(1βγ)βx^d+γx^d+(1+βγ)d(1βγ)βx^d+γx^d+2.
    • Case.1:

      x^dγ. Then γx^d1. Thus,

      1d(1βγ)βx^d+γx^d+2d(1βγ)βx^d+3,

      which implies that x^dd(1βγ)3β. However, it holds that

      x^d =λ(βx^d+1x^d+γ)dλ(βx^d+1x^d)dλ(β+βd(1βγ)3)d
      λβdexp(d(1βγ)d3)<d(1βγ)3β,

      a contradiction.

    • Case.2:

      x^d<γ. Then βx^dβγ<1. Thus,

      1d(1βγ)βx^d+γx^d+2d(1βγ)γx^d+3,

      which implies that x^dγd(1βγ)3. However, it holds that

      x^d =λ(βx^d+1x^d+γ)dλ(x^d+1)dλ(1+γd(1βγ)3)d
      λexp(γdd(1βγ)3)>γd(1βγ)3,

      a contradiction.

  3. 3.

    Let γ>1. The fixed point x^d=λ(βx^d+1x^d+γ)dλγd, and

    |fd(x^d)|=d(1βγ)x^d(βx^d+1)(x^d+γ)dλγd,

    which is strictly less than 1 for all sufficiently large d. Thus the uniqueness holds on infinite d-regular tree for all sufficiently large d.

  4. 4.

    We first show that there exists a critical threshold γc=γc(β,λ,Δ) such that (β,γ,λ) is up-to-Δ unique if and only if γ(γc,1β). It is sufficient to show that if an anti-ferromagnetic (β,γ,λ) is up-to-Δ unique then (β,γ,λ) is up-to-Δ unique for any γ>γ and βγ<1.

    Recall that x^d is the positive fixed point of fd(x)=λ(βx+1x+γ)d. Also let x^d denote the positive solution to x=λ(βx+1x+γ)d.

    We first show that x^d<x^d. By contradiction, assume that x^dx^d. Since for anti-ferromagnetic (β,γ,λ), fd(x) is monotonically decreasing, we have

    x^d=λ(βx^d+1x^d+γ)dλ(βx^d+1x^d+γ)d>λ(βx^d+1x^d+γ)d=x^d,

    a contradiction.

    Since

    λ(β+(1βγ)x^d+γ)d=x^d<x^d=λ(β+(1βγ)x^d+γ)d,

    we have

    (1βγ)x^d+γ<(1βγ)x^d+γ.

    For x^d<x^d, it also holds that

    x^dβx^d+1=1β+1x^d<1β+1x^d=x^dβx^d+1.

    Multiplying the above two inequalities together, we have

    d(1βγ)x^d(βx^d+1)(x^d+γ)<d(1βγ)x^d(βx^d+1)(x^d+γ).

    Note that these are the absolute derivatives at the respective fixed points when the parameters are (β,γ,λ) and (β,γ,λ). Therefore if (β,γ,λ) is up-to-Δ unique then (β,γ,λ) is up-to-Δ unique.

    Due to Part 2 of the lemma, if γ1, |fd(x^d)|>1 for all sufficiently large d, thus γc(β,λ,)>1. And due to Part 3 of the lemma, for any γγc(β,λ,)>1, |fd(x^d)| is arbitrarily close to 0 as d grows to infinity, thus γc(β,λ,)=γc(β,λ,Δ) for a finite Δ.

  5. 5.

    When β=0, |fd(x^d)|=d(1βγ)x^d(βx^d+1)(x^d+γ)=dx^dx^d+γ, the uniqueness condition |fd(x^d)|<1 is equivalent to that x^d<γd1 (here we assume d>1 since due to Part 1 of the lemma, for d=1 the uniqueness always holds). Recall that x^d=λ(1x^d+γ)d. Then |fd(x^d)|<1 if and only if

    λ=x^d(x^d+γ)d<γd+1dd(d1)d+1.

    Let λc=λc(γ,Δ)=min1<d<Δγd+1dd(d1)d+1. It holds that (β,γ,λ) is up-to-Δ unique if and only if λ(0,λc).

  6. 6.

    We note that |fd(x^d)|=d(1βγ)x^d(βx^d+1)(x^d+γ) is not monotone in x^d. It achieves its maximum value at x^d=βγ. Therefore, if for any 1d<Δ, d(1βγ)x(βx+1)(x+γ)<1 for x=βγ, then (β,γ,λ) is up-to-Δ unique for any λ. This condition holds when βγ>d1d+1 for all 1d<Δ, i.e. when βγ>Δ2Δ.

  7. 7.

    Let θ(d)1βγ+d(1βγ). It holds that θ(d)2βγ for all dΔ¯=1+βγ1βγ. Thus, x1(d),x2(d) are well-defined and can be expressed as:

    x1(d)=θ(d)θ(d)24βγ2β and x2(d)=θ(d)+θ(d)24βγ2β.

    Recall that λi(d)xi(d)(xi(d)+γβxi(d)+1)d for i=1,2.

    The following monotonicity and unboundedness of λ2(d) is easy to verify.

    Claim 22.

    λ2(d) is monotonically increasing in d and goes to infinity as d grows.

    Proof.

    Observe that for 0βγ, βγ<1 and dΔ¯, it holds that ββγd1d+1<1 and x+γβx+1 is increasing in x. Moreover,

    x2(d)+γβx2(d)+1 = 1β(d1)(1βγ)+θ(d)24βγ(d+1)(1βγ)+θ(d)24βγ
    d+1d1(d1)(1βγ)(d+1)(1βγ)=1.

    It is easy to verify that x2(d) is increasing in d and unbounded as d grows. Therefore, we can conclude that λ2(d)=x2(d)(x2(d)+γβx2(d)+1)d is increasing in d and is unbounded as d grows.

    We proceed to analyze the uniqueness regime. For 1d<Δ¯, it holds that βγ>d1d+1, in which case it always holds that |fd(x^d)|<1 due to Part 6 of this lemma.

    Recall that βγΔ2Δ. It remains to analyze the |fd(x^d)| for those Δ¯d<Δ. For such d’s, we have βγd1d+1, and since x1(d) and x2(d) are the roots of equation d(1βγ)x(βx+1)(x+γ)=1, it holds that |fd(x^d)|=d(1βγ)x^d(βx^d+1)(x^d+γ)<1 if and only if x^d<x1(d) or x^d>x2(d). Note that x(x+γβx+1)d is monotonically increasing in x. Thus:

    x^d<x1(d) λ=x^d(x^d+γβx^d+1)d<x1(d)(x1(d)+γβx1(d)+1)d=λ1(d),
    x^d>x2(d) λ=x^d(x^d+γβx^d+1)d>x2(d)(x2(d)+γβx2(d)+1)d=λ2(d),

    which means for any Δ¯d<Δ, |fd(x^d)|<1 if and only if λ(0,λ1(d))(λ2(d),).

    Therefore, (β,γ,λ) is up-to-Δ unique if and only if λ belongs to the regime

    Δ¯d<Δ((0,λ1(d))(λ2(d),)),

    which is exactly the uniqueness regime (8). Recall that

    λc=λc(β,γ,Δ)minΔ¯d<Δλ1(d) and λ¯c=λ¯c(β,γ,Δ)maxΔ¯d<Δλ2(d)=λ2(Δ1),

    where the last equation is due to the monotonicity of λ2(d).

    We then show that for γ1, the uniqueness regime (8) becomes (0,λc)(λ¯c,). We need the following claims, whose proofs are postponed to the end of this section.

    Claim 23.

    For Δ¯d0<d1, it always holds that

    (0,mind0d<d1λ1(d))(maxd0d<d1λ2(d),)d0d<d1((0,λ1(d))(λ2(d),));

    furthermore, if λ1(d+1)λ2(d) for all integers d0d<d11, then

    (0,mind0d<d1λ1(d))(maxd0d<d1λ2(d),)=d0d<d1((0,λ1(d))(λ2(d),)).
    Claim 24.

    The followings hold:

    • if γ1, then λ1(d+1)λ2(d) for all integers dΔ¯;

    • if γ>1, there is a finite d0=d0(β,γ)Δ¯ such that λ1(d+1)λ2(d) for all dd0.

    Now assume that γ1. By Claim 24, λ1(d+1)λ2(d) for all integers Δ¯d<Δ1. Then due to Claim 23, the uniqueness regime (8) becomes

    Δ¯d<Δ((0,λ1(d))(λ2(d),))=(0,λc)(λ¯c,).

    In particular, consider β=γ, the anti-ferromagnetic Ising model. It is easy to verify that x1(d)x2(d)=γβ=1, and thus

    λ1(d)λ2(d)=x1(d)x2(d)(x1(d)+γβx1(d)+1x2(d)+γβx2(d)+1)d=1.

    Therefore, λ1(d)=1/λ2(d) is monotonically decreasing, due to the monotonicity of λ2(d) guaranteed by Claim 22. As a result, λc=λ1(Δ1), λ¯c=λ2(Δ1), and λcλ¯c=λ1(Δ1)λ2(Δ1)=1. Then (β,β,λ) is up-to-Δ unique if and only if |logλ|>logλ¯c.

  8. 8.

    The case of β=0 is taken cared of in Part 5 of this lemma. In such case, due to Part 5 of the lemma, (0,γ,λ) is up-to-Δ unique if and only if λ(0,λc(γ,Δ)) where λc(γ,Δ)=min1<d<Δγd+1dd(d1)d+1. In particular, λc(γ,)=mind>1γd+1dd(d1)d+1 becomes 0 if γ1 and becomes a positive and finite λc(γ)>0 if γ>1. Therefore, (0,γ,λ) is universally unique only if γ>1, and when γ>1, (0,γ,λ) is universally unique if and only if λ(0,λc(γ)). We can let λc𝗅𝗈𝗐𝖾𝗋(γ)=λc𝗎𝗉𝗉𝖾𝗋(γ)=λc(γ) and the lemma follows.

    For the case of β>0, due to Part 7 of the lemma, for all ΔΔ¯1+βγ1βγ, the tuple (β,γ,λ) is up-to-Δ unique if and only if λ belongs to the following regime:

    Δ¯d<Δ((0,λ1(d))(λ2(d),)).

    When γ1, this regime becomes (0,λc(β,γ,Δ))(λ¯(β,γ,Δ),), where λc(β,γ,Δ)=minΔ¯d<Δλ1(d) and λ¯c(β,γ,Δ)=maxΔ¯d<Δλ2(d)=λ2(Δ1).

    Claim 25.

    The followings hold:

    • if γ1, then λ1(d) approaches 0 as d grows to infinity;

    • if γ>1, then λ1(d) is unbounded as d grows to infinity.

    Proof.

    For integers dΔ¯, let θ(d)1βγ+d(1βγ), and let

    λ^1(d)2γd+1θ(d)(d+1d1)d. (9)

    When γ1, it is easy to verify that λ^1(d) approaches 0 as d grows to infinity.

    Moreover, λ1(d) is upper bounded by λ^1(d). Specifically, for dΔ¯,

    x1(d) =θ(d)θ(d)24βγ2β=2γθ(d)+θ(d)24βγ2γθ(d).

    Since x+γβx+1 is increasing in x when βγ<1, we have

    x1(d)+γβx1(d)+12γθ(d)+γ2βγθ(d)+1=γd+1d1.

    Thus, for all dΔ¯,

    λ1(d) =x1(d)(x1(d)+γβx1(d)+1)d2γd+1θ(d)(d+1d1)d=λ^1(d). (10)

    Therefore, when γ1, λ1(d) approaches 0 as d grows to infinity.

    On the other hand, when γ>1, for dΔ¯,

    λ1(d)=x1(d)(x1(d)+γβx1(d)+1)dx1(d)γd=2γd+1θ(d)+θ(d)24βγ,

    which is obviously unbounded as d grows.

    When λ1, due to above claim, λ1(d) approaches 0 as d grows to infinity, which means λc(β,γ,)=0; and by Claim 22, λ2(d) goes to infinity as d grows to infinity, which means λ¯c(β,γ,)=. Therefore, (β,γ,λ) can be universally unique only when γ>1.

    Now assume γ>1. By Claim 25, λ1(d) is unbounded as d grows. Thus, λ1(d) achieves its minimum value at a finite d=d(β,γ)Δ¯. And it is obvious that λ1(d) is always finite and positive for any finite dΔ¯. Therefore, λc(β,γ,)=λc(β,γ,d(β,γ)) is finite and positive. Let λc𝗅𝗈𝗐𝖾𝗋(β,γ)=λc(β,γ,)>0 denote this finite positive threshold.

    On the other hand, by Claim 24, for γ>1, there is a finite d0=d0(β,γ)Δ¯ such that λ1(d+1)λ2(d) for all dd0. Let λc𝗎𝗉𝗉𝖾𝗋(β,γ)=mindd0λ1(d). Obviously, λc𝗎𝗉𝗉𝖾𝗋(β,γ)λc𝗅𝗈𝗐𝖾𝗋(β,γ)>0. And λc𝗎𝗉𝗉𝖾𝗋(β,γ)λ1(d0(β,γ)) is also finite.

    Due to Claim 23 and the monotonicity and unboundedness of λ2(d) by Claim 22, the regime for -uniqueness, namely dΔ¯((0,λ1(d))(λ2(d),)), is sandwiched as follows:

    (0,λc𝗅𝗈𝗐𝖾𝗋(β,γ)) dΔ¯((0,λ1(d))(λ2(d),))
    dd0((0,λ1(d))(λ2(d),))
    =(0,λc𝗎𝗉𝗉𝖾𝗋(β,γ))(λ2(),)
    =(0,λc𝗎𝗉𝗉𝖾𝗋(β,γ)).

    Therefore, (β,γ,λ) is universally unique, if λ<λc𝗅𝗈𝗐𝖾𝗋(β,γ) and only if λ<λc𝗎𝗉𝗉𝖾𝗋(β,γ).

It now remains to prove Claim 23 and Claim 24.

Proof of Claim 23.

First, it is easy to verify that for Δ¯d0<d1, it always holds that

(0,mind0d<d1λ1(d))(maxd0d<d1λ2(d),)d0d<d1((0,λ1(d))(λ2(d),)),

since for all d0d<d1, it always holds that

(0,mind0d<d1λ1(d))(0,λ1(d)) and (maxd0d<d1λ2(d),)(λ2(d),).

Next, we prove that the two regimes are equal if λ1(d+1)λ2(d) for all integers d0d<d11. This is proved by induction.

The equality in the claim is trivial to hold when d0<d1d0+1, because in this case there is at most one integer d satisfying d0d<d1. Then both sides of the equality becomes

((0,λ1(d))(λ2(d),)).

This establishes the induction basis.

For the induction hypothesis, assume that the equality holds for d1>d0:

d0d<d1((0,λ1(d))(λ2(d),))=(0,λ1<d1)(λ2<d1,), (11)

where we denote

λ1<d1mind0d<d1λ1(d) and λ2<d1maxd0d<d1λ2(d).

For the induction step, we want to establish the equality when d1 is increased by 1. Thus in addition, we assume that λ1(d+1)λ2(d) for the unique integer d[d11,d1), that is

λ1(d1)λ2(d11). (12)

To finish the induction, it is then sufficient to verify that

d0d<d1+1((0,λ1(d))(λ2(d),))=(0,λ1<d1+1)(λ2<d1+1,),

assuming the induction hypothesis (11). Therefore, it is sufficient to verify that

((0,λ1<d1)(λ2<d1,))((0,λ1(d1))(λ2(d1),))
= (0,λ1<d1+1)(λ2<d1+1,). (13)

Note that d1 is the unique integer in [d1,d1+1). And due to the monotonicity of λ2(d), we have λ2(d1)λ2<d1=λ2(d11). Therefore, (λ2(d1),)(λ2<d1,), and thus (13) holds as long as λ1(d1)λ2<d1=λ2(d11), which is guaranteed by the assumption (12).

Proof of Claim 24.

Recall function λ^1(d) defined in (9): For integers dΔ¯, let λ^1(d)2γd+1θ(d)(d+1d1)d, where θ(d)1βγ+d(1βγ). By (10), for all dΔ¯, we have λ1(d)λ^1(d).

It also holds that λ^1(d)λ2(d). For all dΔ¯, we have x2(d)=θ(d)+θ(d)24βγ2βθ(d)2β. Since x+γβx+1 is monotonically increasing in x when βγ<1, it holds that x1(d)+γβx1(d)+1θ(d)2β+γθ(d)2+1=1βd1d+1. Meanwhile, for all dΔ¯=1+βγ1βγ, it holds that βγ(d1d+1)2 and θ(d)24βγ. Therefore,

λ2(d) =x2(d)(x2(d)+γβx2(d)+1)dθ(d)2βd+1(d1d+1)d2γd+1θ(d)(d+1d1)d=λ^1(d). (14)

Assume γ1. It is also not hard to observe that λ^1(d) is decreasing in dΔ¯>1. Indeed, if γ1, then 2γd+1θ(d)>0 is monotonically decreasing in dΔ¯; and we can verify the monotonicity of function h(d)(d+1d1)d for dΔ¯>1 as follows:

(lnh(d)) =2dd21+ln(d+1d1)0 as d,
(lnh(d))′′ =4d21>0,

which guarantees that h(d) is monotonically decreasing in d, and hence λ^1(d)=2γd+1θ(d)h(d) is decreasing in d for dΔ¯, since both 2γd+1θ(d) and h(d) are positive and decreasing in d.

Therefore, when γ1, due to (10), the monotonicity of λ^1(d), and (14), for all dΔ¯,

λ1(d+1)λ^1(d+1)λ^1(d)λ2(d).

This proves that λ1(d+1)λ2(d) for all dΔ¯ when γ1.

Finally, when γ>1, by contradiction suppose that λ1(d+1)>λ2(d) for some dΔ¯. We then deduce a finite upper bound d0(β,γ) on such bad d. By (10), we have λ1(d+1)λ^1(d+1)=2γd+2θ(d+1)h(d+1)2γd+2θ(d)h(d), where the last inequality is due to that θ(d) is increasing and h(d) is decreasing; and by (14), we have λ2(d)θ(d)2βd+11h(d). Hence λ1(d+1)>λ2(d) implies that

(1βγ)dθ(d)2h(d)2<4βγ2. (15)

Recall that for dΔ¯=1+βγ1βγ, it always holds that θ(d)24βγ. And without loss of generality, we only consider large enough d2, and thus h(d)h(2)=9 since h(d) is decreasing. Therefore inequality (15) implies that

(1βγ)d<81γ,

which gives us a necessary condition d<log1/βγ(81γ) for λ1(d+1)>λ2(d), among the d’s that dΔ¯ and d2. Let

d0(β,γ)max{2,Δ¯,log1βγ(81γ)}.

It then must hold that λ1(d+1)λ2(d) for all dd0(β,γ).

Appendix B Proof of Lemma 11 (a Mean Value Theorem approach)

We prove Lemma 11. The notations are defined in Section 3.

  1. 1.

    Due to the Mean Value Theorem, there exists an R~[Rv,Rv+δv] such that

    ϵv =φ(Rv+δv)φ(Rv)=φ(R~)δv=Φ(R~)δv.
  2. 2.

    Suppose that each vertex has at most k children. Then due to the assumption either k is bounded or γ>1.

    If vΛ, then δvRv+δv=f(Rv1,,Rvd)f(0,,0)=λγd, where 0dk. If γ>1, then Rv+δvλ=O(1); and if k is finite, then Rv+δvmax{λγk,λ}=O(1).

    Due to the Mean Value Theorem, there exist Ri~[Rvi,Rvi+δvi], 1id, such that

    ϵv =φ(f(Rv1,,Rvd))φ(f(Rv1+δv1,,Rvd+δvd))
    =φ(f(R1~,,Rd~))(δv1,,δvd)
    =(1βγ)(f(R1~,,Rd~))12(βf(R1~,,Rd~)+1)12(f(R1~,,Rd~)+γ)12i=1dδvi(βRi~+1)(Ri~+γ)
    f(0,,0)γi=1dδviγ.

    If viS for all vi, then due to the above argument, δviλγdi for free vi, where 0dik is the number of children of vi; and trivially δvi=0 for fixed vi. Therefore, ϵvλγd+1i=1dλγdi+1λ32dγd+32max{1γk,1}. If γ>1, then it is easy to see that dγd+32 is bounded by a constant, thus it holds that ϵv=O(1); if k is bounded, then dk is also bounded, thus clearly ϵv=O(1).

  3. 3.

    We then analyze the stepwise contraction of ϵv. Define that yv=φ(Rv) and accordingly yvi=φ(Rvi), 1id. Then yv+ϵv=φ(Rv+δv) and yvi+ϵvi=φ(Rvi+δvi), 1id. We have

    yv = φ(f(φ1(yv1+ϵv1),,φ1(yvd+ϵvd))),
    yv+ϵv = φ(f(φ1(yv1),,φ1(yvd))).

    Apply the Mean Value Theorem. There exist yi~[yvi,yvi+ϵvi] and corresponding Ri~[Rvi,Rvi+δvi] satisfying yi~=φ(Ri~),1id, such that

    ϵv =φ(f(φ1(yv1),,φ1(yvd)))φ(f(Φ1(yv1+ϵv1),,φ1(yvd+ϵvd)))
    =φ(f(φ1(y1~),,φ1(yd~)))(ϵv1,,ϵvd)
    =Φ(f(R1~,,Rd~))i=1dfRi1Φ(Ri~)ϵi
    =(1βγ)(λi=1dβRi~+1Ri~+γ)12(βλi=1dβRi~+1Ri~+γ+1)12(λi=1dβRi~+1Ri~+γ+γ)12i=1dϵviRi~12(βRi~+1)12(Ri~+γ)12
    max1id{ϵvi}(1βγ)(λi=1dβRi~+1Ri~+γ)12(βλi=1dβRi~+1Ri~+γ+1)12(λi=1dβRi~+1Ri~+γ+γ)12i=1dRi~12(βRi~+1)12(Ri~+γ)12
    =α(d;R~1,,R~d)max1id{ϵvi}.

Appendix C Finding a good potential function heuristically

Perhaps the most mysterious step in our proof is the choice of the potential function Φ(x)=1x(βx+1)(x+γ). As in many cases where potential analysis is applied, there is no standard routine for searching for a suitable potential function. On the other hand, it is quite unlikely that we can just guess such a fairly complicated formula without any hints. Here we present a heuristic approach which leads us to the discovery of a good potential function. This part is not rigorous and logically unnecessary for the soundness of our result. Nevertheless, this heuristic approach is general and interesting enough and may find its applications in other scenarios, thus deserves an exposition here.

The heuristics consists of three steps:

  1. 1.

    Find a necessary condition for the potential function, which is an equation related to the potential function at one point.

  2. 2.

    (heuristic step) Enhance the condition by assuming that the equation holds for the whole range of the variable, which gives us a differential equation.

  3. 3.

    Solve the differential equation and get a potential function.

We first assume that the system is at the boundary of uniqueness. This means that d is the critical degree and f(x^)=1, where f(x)=λ(βx+1x+γ)d and x^=f(x^) is the positive fixed point.

We have the following identities:

x^=λ(βx^+1x^+γ)d and 1=λd(1βγ)(βx^+1)d1(x^+γ)d+1,

which together implies another identity

d(1βγ)x^ =(βx^+1)(x^+γ). (16)

The goal is to find a potential function such that α(x)=|f(x)|Φ(f(x))Φ(x)1 for all x. On the other hand, we have α(x^)=|f(x^)|Φ(f(x^))Φ(x^)=1Φ(x^)Φ(x^)=1. So α(x) achieves its maximum when x=x^. As a differentiable function, it holds that α(x^)=0, that is

[f(x)Φ(f(x))Φ(x)]|x=x^=0.

We have the following calculation:

[f(x)Φ(f(x))Φ(x)]x=x^=0
[f(x)Φ(f(x))]Φ(x)|x=x^=f(x)Φ(f(x))Φ(x)|x=x^
[f′′(x^)Φ(f(x^))+f(x^)Φ(f(x^))f(x^)]Φ(x^)=f(x^)Φ(f(x^))Φ(x^)
f′′(x^)Φ(x^)+Φ(x^)=Φ(x^)
f′′(x^)2=Φ(x^)Φ(x^)=(ln(Φ(x^))),

where we use the fact that x^=f(x^) and f(x^)=1.

Taking the second derivative of f(x), we have

f′′(x)=λd(βγ1)(βx+1)d2(x+γ)d+2((d1)β(x+γ)(d+1)(βx+1)).

This equation already gives an identity for Φ. But it is too complicated. We may further simplify the above formula at the point x=x^.

f′′(x^) = λd(βγ1)(βx^+1)d2(x^+γ)d+2((d1)β(x^+γ)(d+1)(βx^+1))
= 1(βx^+1)(x^+γ)((d+1)(βx^+1)(d1)β(x^+γ))
= d+1x^+γ(d1)ββx^+1
= d(1βγ)(βx^+1)(x^+γ)+1x^+γ+ββx^+1
= 1x^+1x^+γ+ββx^+1.

So we have

(ln(Φ(x^)))=f′′(x^)2=12(1x^+1x^+γ+ββx^+1). (17)

We apply the heuristics and assume that (17) simply holds for all x. This gives us a differential equation

(ln(Φ(x)))=12(1x+1x+γ+ββx+1).

The solution of this differential equation is

ln(Φ(x))=12ln(x(x+γ)(βx+1))+C1,

which gives that

Φ(x)=C2x(βx+1)(x+γ),

where C1,C2 are some constants. This gives the potential function we used in the paper.

There are also other equations which hold for the fixed point x^. Choosing which equation for x^ to heuristically extend to all x may affect the potential function we obtained. For example, (16) implies that

1dx^=1βγ(βx^+1)(x^+γ)=1x^+γββx^+1,

thus we can rewrite f′′(x^) as

f′′(x^)=1x^+1x^+γ+ββx^+1=d+1dx^+2ββx^+1.

This gives that

(ln(Φ(x^)))=d+12dx^ββx^+1.

We can also treat this equation as a differential equation for variable x and solving it gives us

Φ(x)=cxd+12d(βx+1),

which is the potential function used in [15].