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

Cutoff phenomenon of the Glauber dynamics for the Ising model on complete multipartite graphs in the high temperature regime

Heejune Kim School of Mathematics, University of Minnesota. [email protected]
Abstract.

In this paper, the Glauber dynamics for the Ising model on the complete multipartite graph Knp1,,npmK_{np_{1},\dots,np_{m}} is investigated where 0<pi<10<p_{i}<1 is the proportion of the vertices in the iith component. We show that the dynamics exhibits the cutoff phenomena at tn\colonequals12(1β/βcr)nlnnt_{n}\colonequals\frac{1}{2(1-\beta/\beta_{cr})}n\ln n with window size O(n)O(n) in the high temperature regime β<βcr\beta<\beta_{cr} where βcr\beta_{cr} is a constant only depending on p1,,pmp_{1},\dots,p_{m}. Exponentially slow mixing is shown in the low temperature regime β>βcr\beta>\beta_{cr}.

Key words and phrases:
Markov chains, Ising model, Mixing time, Cutoff, Coupling, Glauber dynamics, Heat-bath dynamics, Mean-field model
2020 Mathematics Subject Classification:
60J10, 60K35, 82C20

1. Introduction and preliminaries

Informally, the cutoff phenomenon is an abrupt transition of a Markov chain to its equilibrium when the system under consideration is sufficiently large (see Section 1.3 for a rigorous definition). To the author’s knowledge, the first rapid mixing result appeared in [4] on the symmetric group while considering random transpositions. Shortly afterward, [2] showed that the top-in-at-random card-shuffle precisely exhibits a cutoff phenomenon, initiating the whole industry of the cutoff phenomenon.

As pointed out in [12], only a few examples of cutoff were known regarding the Glauber dynamics of the Ising model (see Section 1.2 for formal definitions), such as that of \citesdinglevin on complete graphs and of \citesLubetzky_2012lubetzky2012cutofflubetzky2014 on lattices. Recent researches have mainly focused on lattices. A breakthrough paper by [12] showed cutoff with a continuous-time window O(lnlnn)O(\ln\ln n) for this longstanding problem. An improvement on the window size to optimal O(1)O(1) was made by the same authors in [14] with the information percolation framework. By the same technique, the authors illustrated the existence of cutoff in high enough temperatures for the Ising model of any sequence of graphs with a bounded degree in [15]. Mean-field Potts model on complete graphs was comprehensively explored in [3], again verifying the cutoff phenomenon in high temperatures. For the bipartite Potts model, [7] proved the cutoff phenomena in the high temperatures using their aggregate path coupling method.

The purpose of this paper is to investigate the Glauber dynamics for the Ising model on complete multipartite graphs. (Exact definitions are given in the rest of the introduction.) Indeed, we identify the critical temperature and establish cutoff in the high temperature regime. On the other hand, exponentially slow mixing is established in the low temperature regime. The significance of our setting is that complete multipartite graphs have an intermediate geometry between the complete graphs which have no geometry at all (e.g. [10]), and lattices which have a strong geometry (e.g. [12]). Thus, our result serves as a midway example between those two extreme cases. The method of proof hinges on generalizations of the tools in [10], notably the two-coordinate chain thereof.

Due to the nature of complete multipartite graphs, our model can be considered as a block spin Ising model with no interaction inside each block. Such mean-field block models naturally occur in statistical physics when modelling metamagnets (see [8]) and in studies on social interactions (see, e.g., [6]). A recent paper by [9] contains an excellent introduction to this line of work.

When it comes to cutoff phenomenon on finite graphs, it is easy to convert the discrete-time results to that of the continuous-time and vice versa. Hence, we only consider discrete-time chains.

1.1. Notations

Boldface letters are used to denote vectors or matrices. Inequalities between vectors and matrices are defined element-wise. The dependence of any quantities on the number of vertices nn is understood throughout the paper. Some important quantities not depending on nn will be explicitly mentioned. We will write 𝐞j\mathbf{e}_{j} to be the jjth vector in the standard basis of m\mathbb{R}^{m}. The lower case tt will always denote time. Let \circ denote the Hadamard product between matrices. More precisely, BC=(BijCij)B\circ C=(B_{ij}C_{ij}) whenever B=(Bij)B=(B_{ij}) and C=(Cij)C=(C_{ij}) are matrices with the same dimensions.

1.2. Ising model and Glauber dynamics

Let G=(V,E)G=(V,E) be a finite graph with the vertex set VV and the edge set EE. Elements of Ω\colonequals{±1}V\Omega\colonequals\{\pm 1\}^{V} are called configurations. In the absence of external fields, the Ising model on GG is a distribution μ\mu called the Gibbs distribution on Ω\Omega given by

μ(σ)\colonequalseβH(σ)Z(β)\mu(\sigma)\colonequals\frac{e^{-\beta H(\sigma)}}{Z(\beta)}

where σΩ\sigma\in\Omega, β0\beta\geq 0, H(σ)=ijEhijσ(i)σ(j)H(\sigma)=-\sum_{ij\in E}h_{ij}\sigma(i)\sigma(j), and Z(β)Z(\beta) is a normalizing factor. Assuming an isotropic interaction strength between the vertices, we set hij=1/|V|h_{ij}=1/|V|. The physical interpretation of H(σ)H(\sigma) is the energy of the whole spin system with the configuration σ\sigma. We call each σ(v)\sigma(v) the spin at site vv.

The Glauber dynamics for the Ising model is a reversible Markov chain with respect to the Gibbs distribution satisfying the following rule. At each time, choose a site uniformly at random in VV and update the spin at the chosen site according to μ\mu conditioned on the set of configurations having the same spins at all the sites except the chosen one. The Glauber dynamics for the Gibbs distribution μ\mu is irreducible, aperiodic, and reversible with μ\mu as its unique stationary distribution. For the Ising model, it is easy to see that the probability of updating to ±1\pm 1 at the chosen site vv is r±(S)r_{\pm}(S) where

r±(x)\colonequalse±βxeβx+eβx=1±tanh(βx)2;x\displaystyle r_{\pm}(x)\colonequals\frac{e^{\pm\beta x}}{e^{\beta x}+e^{-\beta x}}=\frac{1\pm\tanh(\beta x)}{2};\quad x\in\mathbb{R} (1)

and S=vvEσ(v)/|V|S=\sum_{vv^{\prime}\in E}\sigma(v^{\prime})/|V| is the mean-field at vv.

1.3. Markov chain mixing and cutoff phenomenon

The total variation distance between two probability measures ν1\nu_{1} and ν2\nu_{2} on Ω\Omega is defined by

ν1ν2TV\colonequalssupAΩ|ν1(A)ν2(A)|=12xΩ|ν1(x)ν2(x)|.\|\nu_{1}-\nu_{2}\|_{TV}\colonequals\sup_{A\subseteq\Omega}|\nu_{1}(A)-\nu_{2}(A)|=\frac{1}{2}\sum_{x\in\Omega}|\nu_{1}(x)-\nu_{2}(x)|.

The total variation distance is half of the L1L^{1}-distance between the probability measures.

Let (σt)({\sigma_{t}}) be the Markov chain of the Glauber dynamics for the Ising model. Define the worst-case total variation distance of the chains to the stationary distribution μ\mu at time tt by

d(t)\colonequalsmaxσΩσ(σt)μTVd(t)\colonequals\max_{\sigma\in\Omega}\|\mathbb{P}_{\sigma}({\sigma_{t}}\in\cdot)-\mu\|_{TV}

where here and thereafter σ\mathbb{P}_{\sigma} denotes the probability given σ0=σ\sigma_{0}=\sigma. The mixing time is defined by

tmix(ε)\colonequalsmin{t:d(t)ε};ε(0,1).t_{\mathrm{mix}}(\varepsilon)\colonequals\min\{t:d(t)\leq\varepsilon\};\quad\varepsilon\in(0,1).

We say a sequence of Markov chains with corresponding mixing times tmix(n)(ε)t_{\mathrm{mix}}^{(n)}(\varepsilon) exhibit a cutoff phenomenon if for every 0<ε<1/20<\varepsilon<1/2,

limntmix(n)(ε)tmix(n)(1ε)=1.\lim_{n\to\infty}\frac{t_{\mathrm{mix}}^{(n)}(\varepsilon)}{t_{\mathrm{m}ix}^{(n)}(1-\varepsilon)}=1.

Furthermore, we say that the cutoff occurs at tmix(n)t_{\mathrm{mix}}^{(n)} with window size O(wn)O(w_{n}) if wn=o(tmix(n))w_{n}=o(t_{\mathrm{mix}}^{(n)}) and

limγlim infndn(tmix(n)γwn)=1,limγlim supndn(tmix(n)+γwn)=0.\displaystyle\lim_{\gamma\to\infty}\liminf_{n\to\infty}d_{n}(t_{\mathrm{mix}}^{(n)}-\gamma w_{n})=1,\quad\lim_{\gamma\to\infty}\limsup_{n\to\infty}d_{n}(t_{\mathrm{mix}}^{(n)}+\gamma w_{n})=0.

1.4. Magnetization chain on complete multipartite graphs

Now, we are in a place to consider a complete mm-partite graph, a graph whose vertices are partitioned into mm different independent sets, and every pair of vertices from different independent sets is connected by an edge. Each edge represents an interaction between the vertices. Denote this graph by Knp1,np2,,npmK_{np_{1},np_{2},\dots,np_{m}} which has nn vertices and mm partitions where i=1mpi=1\sum_{i=1}^{m}p_{i}=1 and pi>0p_{i}>0 for i=1,2,,mi=1,2,\dots,m. We fix the parameters mm and pip_{i}’s hereafter. Without loss of generality, we assume p1p2pmp_{1}\leq p_{2}\leq\dots\leq p_{m}. We may also assume that npinp_{i}\in\mathbb{N} for every ii so that Knp1,np2,,npmK_{np_{1},np_{2},\dots,np_{m}} is well defined whenever such considerations are required. Let V=i=1mJiV=\bigcup_{i=1}^{m}J_{i} be the set of all vertices where JiJ_{i} denotes the set of the iith partition of the vertices. Note npi=|Ji|np_{i}=|J_{i}|.

We define Ωi\colonequals{±1}Ji\Omega_{i}\colonequals\{\pm 1\}^{J_{i}} for i=1,,mi=1,\dots,m so that Ω=i=1mΩi\Omega=\prod_{i=1}^{m}\Omega_{i} is our configuration space. Each configuration σΩ\sigma\in\Omega has a unique representation (σ(1),,σ(m))i=1mΩi(\sigma^{(1)},\dots,\sigma^{(m)})\in\prod_{i=1}^{m}\Omega_{i} and both representations are understood throughout this paper.

For each σΩ\sigma\in\Omega, define the magnetization on JiJ_{i} by S(i)(σ)\colonequalsvJiσ(v)/nS^{(i)}(\sigma)\colonequals\sum_{v\in J_{i}}\sigma(v)/n, i=1,,mi=1,\dots,m. For the Markov chain (σt)t0=(σt(1),,σt(m))t0(\sigma_{t})_{t\geq 0}=(\sigma_{t}^{(1)},\dots,\sigma_{t}^{(m)})_{t\geq 0} starting at σ=(σ(1),,σ(m))i=1mΩi\sigma=(\sigma^{(1)},\dots,\sigma^{(m)})\in\prod_{i=1}^{m}\Omega_{i}, we define the corresponding magnetization on JiJ_{i} by

St(i)\colonequals1nvViσt(i)(v)fori{1,,m},t0.S_{t}^{(i)}\colonequals\frac{1}{n}\sum_{v\in V_{i}}\sigma_{t}^{(i)}(v)\ \text{for}\ i\in\{1,\dots,m\},\ t\geq 0.

We sometimes use the vector notation 𝐒t\colonequals(St(1),,St(m))\mathbf{S}_{t}\colonequals(S^{(1)}_{t},\dots,S^{(m)}_{t}) for t0t\geq 0. We call the process (𝐒t)t0(\mathbf{S}_{t})_{t\geq 0} a magnetization chain. \threfmagmarkov shows that (𝐒t)t0(\mathbf{S}_{t})_{t\geq 0} is in fact a Markov chain. Note that it is a projection of the whole Markov chain (σt)t0({\sigma_{t}})_{t\geq 0}, so mixing of the whole chain (σt)t0({\sigma_{t}})_{t\geq 0} implies the mixing of the chain (𝐒t)t0(\mathbf{S}_{t})_{t\geq 0}. Our aim is to show the converse in a certain sense.

1.5. Main results

Given the above definitions and notations, our main result establishes the cutoff phenomenon on complete multipartite graphs.

Theorem 1.1 (Main result).

For mm\in\mathbb{N} and pi>0p_{i}>0 such that i=1mpi=1\sum_{i=1}^{m}p_{i}=1, the Glauber dynamics for the Ising model on the complete multipartite graph Knp1,,npmK_{np_{1},\dots,np_{m}} exhibits a cutoff at 12(1β/βcr)nlnn\frac{1}{2(1-\beta/\beta_{cr})}n\ln n with window size O(n)O(n) in the high temperature regime β<βcr\beta<\beta_{cr} where βcr=βcr(p1,,pm)\beta_{cr}=\beta_{cr}(p_{1},\dots,p_{m}) is a constant defined in equation (3).

Theorem 1.2.
\thlabel

lowtempresult In the low temperature regime β>βcr\beta>\beta_{cr}, the dynamics is exponentially slow mixing, i.e., tmixC1exp(C2n)t_{\mathrm{mix}}\geq C_{1}\exp(C_{2}n) for some constants C1,C2>0C_{1},\ C_{2}>0 not depending on nn.

A few remarks are in order. Our main result is obtained as a consequence of \threfupperbound and \threflowerbound. In the low temperature regime β>βcr\beta>\beta_{cr}, the mixing time is exponentially slow, therefore identifying the critical temperature βcr\beta_{cr}. In the m=1m=1 case, there are no spin interactions so the chain is equivalent to the lazy random walk on an nn-dimensional hypercube, which has a cutoff at (nlnn)/2(n\ln n)/2 with window size O(n)O(n) (see [1] or [11, Chapter 18]). This result can be seen as a consequence of our main result since m=1m=1 implies βcr=\beta_{cr}=\infty (see equation (3)).

1.6. Organization of the article

As mentioned earlier, our proof is based on the ideas of [10]. We assume high temperatures until Section 6. We first observe that the magnetization chain is a Markov chain in its own right (\threfmagmarkov). A suitable scaling of the magnetization chain leads to a contraction property (\threfnormcontraction). This in turn gives a uniform variance bound of magnetizations in time (Sections 2 and 3). In Section 4, we construct a coupling of the magnetization chain so that it couples in 12(1β/βcr)nlnn+O(n)\frac{1}{2(1-\beta/\beta_{cr})}n\ln n+O(n) steps with high probability. After the magnetization coupling phase, by considering the "2m2m-coordinate chain" inspired by [10], we can construct a post magnetization coupling to reach the full-mixing in another O(n)O(n) steps. This proves the upper bound (\threfupperbound). We construct a suitable distinguishing-statistic of the magnetization chain [[, see]Chapter 7.3]Peres to obtain the lower bound (\threflowerbound). These upper and lower bound results establish the cutoff in the high temperature regime. Exponentially slow mixing in the low temperature regime is shown in Section 6.

2. Contraction of the magnetization chain in high temperatures

We describe the monotone coupling. Let II and UU be independent uniform random variables over VV and [0,1][0,1], respectively. We consider the collection of Markov chains with starting configurations σΩ\sigma\in\Omega. Simultaneously define the next configurations at time t=1t=1 by

σ1(i)={σ(i)if Ii𝟙U<r+(jkS(j)(σ))𝟙Ur+(jkS(j)(σ))if I=iJk\sigma_{1}(i)=\begin{cases}\sigma(i)&\text{if $I\neq i$}\\ \mathbbm{1}_{U<r_{+}(\sum_{j\neq k}S^{(j)}(\sigma))}-\mathbbm{1}_{U\geq r_{+}(\sum_{j\neq k}S^{(j)}(\sigma))}&\text{if $I=i\in J_{k}$}\end{cases}\quad

where r+r_{+} is defined in equation (1). Repeat this procedure independently for each time. It is clear that each Markov chain (σt)t0(\sigma_{t})_{t\geq 0} above is a version of the Glauber dynamics on the complete multipartite graph with starting state σ\sigma’s, defined on a common probability space. The above coupling is called a monotone coupling in the sense that if σσ~\sigma\leq\tilde{\sigma} are starting states for (σt)t0(\sigma_{t})_{t\geq 0} and (σ~t)t0(\tilde{\sigma}_{t})_{t\geq 0}, respectively, then S(i)(σ)S(i)(σ~)S^{(i)}(\sigma)\leq S^{(i)}(\tilde{\sigma}) for i=1,,mi=1,\dots,m so that σ1σ~1\sigma_{1}\leq\tilde{\sigma}_{1}, and σtσ~t\sigma_{t}\leq\tilde{\sigma}_{t} for any t0t\geq 0 accordingly.

Define

𝒮\colonequalsi=1m{pi,pi+2/n,,pi}.\mathcal{S}\colonequals\prod_{i=1}^{m}\{-p_{i},-p_{i}+2/n,\dots,p_{i}\}.
Proposition 2.1 (Magnetization chain).
\thlabel

magmarkov The process (St(1),,St(m))t0(S^{(1)}_{t},\dots,{S^{(m)}_{t}})_{t\geq 0} is a
Markov chain on the magnetization state space 𝒮\mathcal{S}.

Proof.

Note that

((St+1(1),,St+1(m))=(St(1)2n,,St(m)))\displaystyle\mathbb{P}\bigl{(}({S^{(1)}_{t+1}},\dots,{S^{(m)}_{t+1}})=(S^{(1)}_{t}-\frac{2}{n},\dots,{S^{(m)}_{t}})\bigr{)} =p1nSt(1)+|J1|2|J1|r(j1St(j))\displaystyle=p_{1}\frac{n{S^{(1)}_{t}}+|J_{1}|}{2|J_{1}|}r_{-}\Bigl{(}\sum_{j\neq 1}{S^{(j)}_{t}}\Bigr{)}
=p1+St(1)2r(j1St(j))\displaystyle=\frac{p_{1}+{S^{(1)}_{t}}}{2}r_{-}\Bigl{(}\sum_{j\neq 1}{S^{(j)}_{t}}\Bigr{)}

is measurable with respect to the σ\sigma-algebra generated by (St(1),,St(m))({S^{(1)}_{t}},\dots,{S^{(m)}_{t}}). Other cases can be dealt with similarly. ∎

Remark.

By symmetry, (St(1),,St(m))({S^{(1)}_{t}},\dots,{S^{(m)}_{t}}) starting from σ\sigma and (St(1),,St(m))(-{S^{(1)}_{t}},\dots,-{S^{(m)}_{t}}) starting from σ-\sigma have the same distributions. This can also be seen by the physical fact that the map σσ\sigma\mapsto-\sigma just corresponds to flipping the reference axis to which we are measuring the spins of each site. This does not change the dynamics of the spin system.

Definition (Hamming distance).

For two configurations σ\sigma and σ\sigma^{\prime}, denote the Hamming distance by dist(σ,σ)\colonequals12kV|σ(k)σ(k)|\mathrm{dist}(\sigma,\sigma^{\prime})\colonequals\frac{1}{2}\sum_{k\in V}|\sigma(k)-\sigma^{\prime}(k)|.

Remark.

This is a metric on Ω\Omega, which is equal to the number of sites with different spins for two configurations. Similarly, we can define disti\mathrm{dist}_{i} on Ωi\Omega_{i}, respectively, but disti\mathrm{dist}_{i}’s merely satisfy the triangle inequality.

Lemma 2.2 (Contraction in mean for monotone coupling).
\thlabel

monotonecontraction For a monotone coupling (σt,σt)t0({\sigma_{t}},{\sigma^{\prime}_{t}})_{t\geq 0} starting at (σ,σ)=((σ(1),,σ(2)),(σ(1),,σ(2)))(\sigma,\sigma^{\prime})=((\sigma^{(1)},\dots,\sigma^{(2)}),(\sigma^{\prime(1)},\dots,\sigma^{\prime(2)})), we have

(𝔼dist1(σt(1),σt(1))𝔼distm(σt(m),σt(m)))𝐀t(dist1(σ(1),σ0)distm(σ(m),σ(m)))\begin{pmatrix}\mathbb{E}\mathrm{dist}_{1}({\sigma^{(1)}_{t}},{\sigma^{\prime(1)}_{t}})\\ \vdots\\ \mathbb{E}\mathrm{dist}_{m}({\sigma^{(m)}_{t}},{\sigma^{\prime(m)}_{t}})\end{pmatrix}\leq\mathbf{A}^{t}\begin{pmatrix}\mathrm{dist}_{1}(\sigma^{(1)},\sigma_{0}^{\prime})\\ \vdots\\ \mathrm{dist}_{m}(\sigma^{(m)},\sigma^{\prime(m)})\end{pmatrix}

where

𝐀=𝐀n\colonequals(ab1b1b1b2ab2b2b3b3ab3bma)\mathbf{A}=\mathbf{A}_{n}\colonequals\begin{pmatrix}a&b_{1}&b_{1}&\dots&b_{1}\\ b_{2}&a&b_{2}&\dots&b_{2}\\ b_{3}&b_{3}&a&\dots&b_{3}\\ \vdots&\dots&&&\vdots\\ b_{m}&\dots&&\dots&a\end{pmatrix}

with a\colonequals11/na\colonequals 1-1/n, bk\colonequalspkβ/nb_{k}\colonequals p_{k}{\beta}/{n}.

Proof.

Assume d(σ,σ)=1d(\sigma,\sigma^{\prime})=1 with 1=σ(v)=σ(v)-1=\sigma(v)=-\sigma^{\prime}(v) for some vertex vv. Note σσ\sigma\leq\sigma^{\prime}. Since we are considering a monotone coupling, it holds that for each i=1,,mi=1,\dots,m,

disti(σ1(i),σ1(i))=𝟙vJi(1𝟙I=v)+𝟙vJi(𝟙IJi𝟙Bi)\displaystyle\mathrm{dist}_{i}({\sigma^{(i)}_{1}},{\sigma^{\prime(i)}_{1}})=\mathbbm{1}_{v\in J_{i}}(1-\mathbbm{1}_{I=v})+\mathbbm{1}_{v\notin J_{i}}(\mathbbm{1}_{I\in J_{i}}\mathbbm{1}_{B_{i}})

where

Bi={r+(liSl(σ))U<r+(liSl(σ))}.B_{i}=\Biggl{\{}r_{+}\biggl{(}\sum_{l\neq i}S_{l}(\sigma)\biggr{)}\leq U<r_{+}\biggl{(}\sum_{l\neq i}S_{l}(\sigma^{\prime})\biggr{)}\Biggl{\}}.

Note that

(Bi)\displaystyle\mathbb{P}(B_{i}) =12(tanh(βliSl(σ))tanh(βliSl(σ)))\displaystyle=\frac{1}{2}\Biggl{(}\tanh\biggl{(}\beta\sum_{l\neq i}S_{l}(\sigma^{\prime})\biggr{)}-\tanh\biggl{(}\beta\sum_{l\neq i}S_{l}(\sigma)\biggr{)}\Biggr{)}
=12(tanh(β(liSl(σ)+2n))tanh(βliSl(σ)))𝟙vJi\displaystyle=\frac{1}{2}\Biggl{(}\tanh\biggl{(}\beta\biggl{(}\sum_{l\neq i}S_{l}(\sigma)+\frac{2}{n}\biggr{)}\biggr{)}-\tanh\biggl{(}\beta\sum_{l\neq i}S_{l}(\sigma)\biggr{)}\Biggr{)}\mathbbm{1}_{v\notin J_{i}}
tanhβn𝟙vJi.\displaystyle\leq\tanh\frac{\beta}{n}\mathbbm{1}_{v\notin J_{i}}.

Since II and UU are independent, for i=1,,mi=1,\dots,m,

𝔼disti(σ1(i),σ1(i))𝟙vJi(11n)+𝟙vJipitanhβn.\displaystyle\mathbb{E}\mathrm{dist}_{i}({\sigma^{(i)}_{1}},{\sigma^{\prime(i)}_{1}})\leq\mathbbm{1}_{v\in J_{i}}(1-\frac{1}{n})+\mathbbm{1}_{v\notin J_{i}}p_{i}\tanh\frac{\beta}{n}.

Suppose dist(σ,σ)=k>1\mathrm{dist}(\sigma,\sigma^{\prime})=k>1. There exists σ0\colonequalsσ\sigma^{0}\colonequals\sigma, σ1\sigma^{1}, \dots, σk\colonequalsσ\sigma^{k}\colonequals\sigma^{\prime} such that dist(σi,σi+1)=1\mathrm{dist}(\sigma^{i},\sigma^{i+1})=1. By the triangular inequality for disti\mathrm{dist}_{i} and the fact tanh(β/n)β/n\tanh(\beta/n)\leq\beta/n,

𝔼disti(σ1(i),σ1(i))(11n)disti(σ(i),σ(i))+piβnlidistl(σ(l),σ(l)).\displaystyle\mathbb{E}\mathrm{dist}_{i}({\sigma^{(i)}_{1}},{\sigma^{\prime(i)}_{1}})\leq(1-\frac{1}{n})\mathrm{dist}_{i}(\sigma^{(i)},\sigma^{\prime(i)})+p_{i}\frac{\beta}{n}\sum_{l\neq i}\mathrm{dist}_{l}(\sigma^{(l)},\sigma^{\prime(l)}).

Furthermore, by the Markov property,

𝔼[disti(σt+1(i),σt+1(i))|σt,σt](11n)disti(σt(i),σt(i))+piβnlidistl(σt(l),σt(l)).\mathbb{E}[\mathrm{dist}_{i}({\sigma^{(i)}_{t+1}},\sigma^{\prime(i)}_{t+1})|{\sigma_{t}},{\sigma^{\prime}_{t}}]\leq(1-\frac{1}{n})\mathrm{dist}_{i}({\sigma^{(i)}_{t}},{\sigma^{\prime(i)}_{t}})+\frac{p_{i}\beta}{n}\sum_{l\neq i}\mathrm{dist}_{l}({\sigma^{(l)}_{t}},{\sigma^{\prime(l)}_{t}}).

By taking expectation and putting xi,t\colonequals𝔼disti(σt(i),σt(i))x_{i,t}\colonequals\mathbb{E}\mathrm{dist}_{i}({\sigma^{(i)}_{t}},{\sigma^{\prime(i)}_{t}}), we have

(x1,txm,t)𝐀(x1,t1xm,t1).\displaystyle\begin{pmatrix}x_{1,t}\\ \vdots\\ x_{m,t}\end{pmatrix}\leq\mathbf{A}\begin{pmatrix}x_{1,t-1}\\ \vdots\\ x_{m,t-1}\end{pmatrix}.

Iterating gives

(x1,txm,t)𝐀t(dist1(σ(1),σ(1))distm(σ(m),σ(m))).\begin{pmatrix}x_{1,t}\\ \vdots\\ x_{m,t}\end{pmatrix}\leq\mathbf{A}^{t}\begin{pmatrix}\mathrm{dist}_{1}(\sigma^{(1)},\sigma^{\prime(1)})\\ \vdots\\ \mathrm{dist}_{m}(\sigma^{(m)},\sigma^{\prime(m)})\end{pmatrix}.

From now on, 𝐀\mathbf{A} (which depends on the number of vertices nn) always denotes the matrix defined in \threfmonotonecontraction. Note that 𝐀\mathbf{A} is a positive matrix, so by the Perron-Frobenius theorem, there exists the largest eigenvalue g=gn>0g=g_{n}>0 with the left eigenvector 𝐚T\colonequals(a1,,am)>𝟎\mathbf{a}^{T}\colonequals(a_{1},\dots,a_{m})>\mathbf{0} normalized in l1l^{1} norm. Note that gg has algebraic multiplicity 11 (see [16, Section 8.2] for a proof), so 𝐚T\mathbf{a}^{T} is unique.

We fix the following notations

υ\colonequalsn(1g) and \displaystyle\upsilon\colonequals n(1-g)\text{ and } (2)
βcr\colonequals1(m1)i=1maipi\displaystyle\beta_{cr}\colonequals\frac{1}{(m-1)\sum_{i=1}^{m}a_{i}p_{i}} (3)

where gg and (a1,,am)(a_{1},\dots,a_{m}) are defined in the previous paragraph. Another characterization of βcr\beta_{cr} is given in \threfslowmixinglemma. Insuk Seo commented111personal communication that it can also be characterized as the threshold value of β\beta that makes 𝐊\mathbf{K} positive definite where 𝐊\mathbf{K} is defined through the equation 𝐀=𝐈1n𝐊\mathbf{A}=\mathbf{I}-\frac{1}{n}\mathbf{K}, 𝐈\mathbf{I} being the mm-by-mm identity matrix. \threfupsilon connects the quantities υ\upsilon and βcr\beta_{cr}.

Proposition 2.3.
\thlabel

upsilon The left eigenvector 𝐚T\mathbf{a}^{T} only depends on p1,,pmp_{1},\dots,p_{m}. Moreover, υ\upsilon only depends on p1,,pm,p_{1},\dots,p_{m}, and β\beta through the following equation:

υ=1β(m1)i=1maipi.\upsilon=1-\beta(m-1)\sum_{i=1}^{m}a_{i}p_{i}.

Therefore, βcr\beta_{cr} only depends on p1,,pmp_{1},\dots,p_{m}, and we have υ=1β/βcr\upsilon=1-\beta/\beta_{cr}.

Proof.

Since gg satisfies

0\displaystyle 0 =(n/β)mdet(𝐀gI)=det(n𝐀/βngI/β)\displaystyle=(n/\beta)^{m}\det(\mathbf{A}-gI)=\det(n\mathbf{A}/\beta-ngI/\beta)
=det(υ1βp1p1p1p2υ1βp2p2p3p3υ1βp3pmυ1β),\displaystyle=\det\begin{pmatrix}\frac{\upsilon-1}{\beta}&p_{1}&p_{1}&\dots&p_{1}\\ p_{2}&\frac{\upsilon-1}{\beta}&p_{2}&\dots&p_{2}\\ p_{3}&p_{3}&\frac{\upsilon-1}{\beta}&\dots&p_{3}\\ \vdots&\dots&&&\vdots\\ p_{m}&\dots&&\dots&\frac{\upsilon-1}{\beta}\end{pmatrix},

it holds that (υ1)/β(\upsilon-1)/\beta is a root of a polynomial with coefficients only depending on p1,,pmp_{1},\dots,p_{m}. Since 𝐚\mathbf{a} is in the kernel of the transpose of the above matrix, it only depends on p1,,pmp_{1},\dots,p_{m}.

Finally, g=𝐀T𝐚1=11/n+βn(m1)kaipig=\|\mathbf{A}^{T}\mathbf{a}\|_{1}=1-1/n+\frac{\beta}{n}(m-1)\sum_{k}a_{i}p_{i} implies υ=1β(m1)iaipi\upsilon=1-\beta(m-1)\sum_{i}a_{i}p_{i}.

We collect further properties of the matrix 𝐀\mathbf{A} and its left eigenvector 𝐚T\mathbf{a}^{T} in the next two lemmas.

Lemma 2.4.
\thlabel

finalfinallemma We have

a1amandi=1maipi1m.a_{1}\geq\dots\geq a_{m}\ \text{and}\ \sum_{i=1}^{m}a_{i}p_{i}\leq\frac{1}{m}.

The equality in the latter holds if and only if p1==pmp_{1}=\dots=p_{m}.

Proof.

Recall that we assumed p1pmp_{1}\leq\dots\leq p_{m}.

We claim that a1ama_{1}\geq\dots\geq a_{m}. To that end, fix i<ji<j. From 𝐚T𝐀=g𝐚T\mathbf{a}^{T}\mathbf{A}=g\mathbf{a}^{T}, we have (11n)ai+βnkiakpkgai=0=(11n)aj+βnkjakpkgaj(1-\frac{1}{n})a_{i}+\frac{\beta}{n}\sum_{k\neq i}a_{k}p_{k}-ga_{i}=0=(1-\frac{1}{n})a_{j}+\frac{\beta}{n}\sum_{k\neq j}a_{k}p_{k}-ga_{j}. Then (11ngβpin)ai=(11ngβpjn)aj(1-\frac{1}{n}-g-\frac{\beta p_{i}}{n})a_{i}=(1-\frac{1}{n}-g-\frac{\beta p_{j}}{n})a_{j}, i.e., (βpi+1υ)ai=(βpj+1υ)aj(\beta p_{i}+1-\upsilon)a_{i}=(\beta p_{j}+1-\upsilon)a_{j}. Thus, pipjp_{i}\leq p_{j} implies aiaja_{i}\geq a_{j}, proving the claim.

By Chebyshev’s sum inequality, since aiaja_{i}\geq a_{j} and pipjp_{i}\leq p_{j} whenever i<ji<j,

i=1maipi1m(i=1mai)(i=1mpi)=1m.\displaystyle\sum_{i=1}^{m}a_{i}p_{i}\leq\frac{1}{m}\left(\sum_{i=1}^{m}a_{i}\right)\left(\sum_{i=1}^{m}p_{i}\right)=\frac{1}{m}.

The equality holds if and only if a1==ama_{1}=\dots=a_{m} or p1==pmp_{1}=\dots=p_{m}. The proof is now complete by noticing the fact that (βpi+1υ)ai=(βpj+1υ)aj(\beta p_{i}+1-\upsilon)a_{i}=(\beta p_{j}+1-\upsilon)a_{j} and a1==am=1/ma_{1}=\dots=a_{m}=1/m imply p1==pmp_{1}=\dots=p_{m}. ∎

Remark.

As a consequence, we obtain a lower bound βcrm/(m1)\beta_{cr}\geq m/(m-1).

Lemma 2.5.
\thlabel

newlemma For 𝟎𝐬𝒮\mathbf{0}\leq\mathbf{s}\in\mathcal{S} and 𝐩\colonequals(p1,,pm)T\mathbf{p}\colonequals(p_{1},\dots,p_{m})^{T}, we have

𝐀t𝐬1gt(i=1m(s(i))2pi)1/2,𝐞jT𝐀t𝐬pjgt(i=1m(s(i))2pi)1/2.\|\mathbf{A}^{t}\mathbf{s}\|_{1}\leq g^{t}\left(\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}\right)^{1/2},\quad\mathbf{e}_{j}^{T}\mathbf{A}^{t}\mathbf{s}\leq\sqrt{p_{j}}g^{t}\left(\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}\right)^{1/2}.

In particular, it holds that

𝐀t𝐩1gt,𝐞jT𝐀t𝐩pjgt.\|\mathbf{A}^{t}\mathbf{p}\|_{1}\leq g^{t},\quad\mathbf{e}_{j}^{T}\mathbf{A}^{t}\mathbf{p}\leq\sqrt{p_{j}}g^{t}.
Proof.

We want to find a symmetric matrix 𝐂\mathbf{C} which is similar to 𝐀\mathbf{A}. To that end, suppose that there exists an invertible diagonal matrix 𝐃=diag(d1,,dm)\mathbf{D}=\mathrm{diag}(d_{1},\dots,d_{m}) and a symmetric matrix 𝐂\mathbf{C} such that 𝐂=𝐃1𝐀𝐃\mathbf{C}=\mathbf{D}^{-1}\mathbf{A}\mathbf{D}. Then 𝐃𝐀T𝐃1=𝐂T=𝐂=𝐃1𝐀𝐃\mathbf{D}\mathbf{A}^{T}\mathbf{D}^{-1}=\mathbf{C}^{T}=\mathbf{C}=\mathbf{D}^{-1}\mathbf{A}\mathbf{D}, so 𝐃2𝐀T=𝐀𝐃2\mathbf{D}^{2}\mathbf{A}^{T}=\mathbf{A}\mathbf{D}^{2}, which leads to di2pj=pidj2d_{i}^{2}p_{j}=p_{i}d_{j}^{2} for i,j{1,2,,m}i,j\in\{1,2,\dots,m\}. With the above in mind, let 𝐃\colonequalsdiag(p1,,pm)\mathbf{D}\colonequals\mathrm{diag}(\sqrt{p_{1}},\dots,\sqrt{p_{m}}) and 𝐂\colonequals(cij)\mathbf{C}\colonequals(c_{ij}) where cii=11/nc_{ii}=1-1/n and cij=pipjβ/nc_{ij}=\sqrt{p_{i}p_{j}}{\beta}/{n} for iji\neq j. Note that 𝐂\mathbf{C} is real-symmetric and 𝐂=𝐃1𝐀𝐃\mathbf{C}=\mathbf{D}^{-1}\mathbf{A}\mathbf{D}. Then, by the spectral theorem for real symmetric matrices, 𝐂2=g\|\mathbf{C}\|_{2}=g. Note that 𝐂\mathbf{C} and 𝐀\mathbf{A} have the same real eigenvalues since they are similar.

Observe that for 𝐱,𝐲m\mathbf{x},\mathbf{y}\in\mathbb{R}^{m}, 𝐱𝐲T2=𝐱2𝐲2\|\mathbf{x}\mathbf{y}^{T}\|_{2}=\|\mathbf{x}\|_{2}\|\mathbf{y}\|_{2}. This can be easily checked by the equalities

𝐱𝐲T2=sup𝐳2=1𝐱𝐲T𝐳2=sup𝐳2=1|𝐲T𝐳|𝐱2=𝐲2𝐱2.\|\mathbf{x}\mathbf{y}^{T}\|_{2}=\sup_{\|\mathbf{z}\|_{2}=1}\|\mathbf{x}\mathbf{y}^{T}\mathbf{z}\|_{2}=\sup_{\|\mathbf{z}\|_{2}=1}|\mathbf{y}^{T}\mathbf{z}|\|\mathbf{x}\|_{2}=\|\mathbf{y}\|_{2}\|\mathbf{x}\|_{2}.

Let 𝟙\colonequals(1,,1)T\mathbbm{1}\colonequals(1,\dots,1)^{T}. The case 𝐬=𝟎\mathbf{s}=\mathbf{0} is trivial, so assume 𝐬>𝟎\mathbf{s}>\mathbf{0}. Since 𝐬𝟙T\mathbf{s}\mathbbm{1}^{T} has rank 1, 𝐂t𝐃1𝐬𝟙T𝐃\mathbf{C}^{t}\mathbf{D}^{-1}\mathbf{s}\mathbbm{1}^{T}\mathbf{D} has rank 11 . Also, its elements are positive, so it has a positive eigenvalue by the Perron-Frobenius theorem. Thus, Tr(𝐂t𝐃1𝐬𝟙T𝐃)\mathrm{Tr}(\mathbf{C}^{t}\mathbf{D}^{-1}\mathbf{s}\mathbbm{1}^{T}\mathbf{D}) is equal to its spectral radius, from which the following inequality follows:

𝐀t𝐬1\displaystyle\|\mathbf{A}^{t}\mathbf{s}\|_{1} =𝟙T𝐀t𝐬=𝟙T𝐃𝐂t𝐃1𝐬=Tr(𝐂t𝐃1𝐬𝟙T𝐃)𝐂t𝐃1𝐬𝟙T𝐃2\displaystyle=\mathbbm{1}^{T}\mathbf{A}^{t}\mathbf{s}=\mathbbm{1}^{T}\mathbf{D}\cdot\mathbf{C}^{t}\mathbf{D}^{-1}\mathbf{s}=\mathrm{Tr}(\mathbf{C}^{t}\mathbf{D}^{-1}\mathbf{s}\mathbbm{1}^{T}\mathbf{D})\leq\|\mathbf{C}^{t}\mathbf{D}^{-1}\mathbf{s}\mathbbm{1}^{T}\mathbf{D}\|_{2}
𝐂2t𝐃1𝐬𝟙T𝐃2=gt𝐃1𝐬2𝐃𝟙2=gt(i=1m(s(i))2pi)1/2.\displaystyle\leq\|\mathbf{C}\|_{2}^{t}\|\mathbf{D}^{-1}\mathbf{s}\mathbbm{1}^{T}\mathbf{D}\|_{2}=g^{t}\|\mathbf{D}^{-1}\mathbf{s}\|_{2}\|\mathbf{D}\mathbbm{1}\|_{2}=g^{t}\left(\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}\right)^{1/2}.

Similarly,

𝐞jT𝐀t𝐬𝐂2t𝐃1𝐬𝐞jT𝐃2=gt𝐃1𝐬2𝐃𝐞j2=pjgt(i=1m(s(i))2pi)1/2.\mathbf{e}_{j}^{T}\mathbf{A}^{t}\mathbf{s}\leq\|\mathbf{C}\|_{2}^{t}\|\mathbf{D}^{-1}\mathbf{s}\mathbf{e}_{j}^{T}\mathbf{D}\|_{2}=g^{t}\|\mathbf{D}^{-1}\mathbf{s}\|_{2}\|\mathbf{D}\mathbf{e}_{j}\|_{2}=\sqrt{p_{j}}g^{t}\Biggl{(}\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}\Biggr{)}^{1/2}.

Remark.

Another relatively simple proof of 𝐀t𝐬1gt(i=1m(s(i))2pi)1/2\|\mathbf{A}^{t}\mathbf{s}\|_{1}\leq g^{t}\left(\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}\right)^{1/2} can be given as follows. By the Cauchy-Schwartz inequality, we have isi2/piisi\sqrt{\sum_{i}s_{i}^{2}/p_{i}}\geq\sum_{i}s_{i}. Then 𝐀t𝐬1𝐃1𝐀t𝐬2=𝐃1𝐀t𝐃𝐃1𝐬2=𝐂t𝐃1𝐬2gt𝐃1𝐬2\|\mathbf{A}^{t}\mathbf{s}\|_{1}\leq\|\mathbf{D}^{-1}\mathbf{A}^{t}\mathbf{s}\|_{2}=\|\mathbf{D}^{-1}\mathbf{A}^{t}\mathbf{D}\mathbf{D}^{-1}\mathbf{s}\|_{2}=\|\mathbf{C}^{t}\mathbf{D}^{-1}\mathbf{s}\|_{2}\leq g^{t}\|\mathbf{D}^{-1}\mathbf{s}\|_{2}.

From now on, for brevity, we use the notation

𝐩\colonequals(p1,,pm)T.\mathbf{p}\colonequals(p_{1},\dots,p_{m})^{T}.
Lemma 2.6.
\thlabel

lemma For a monotone coupling (σt,σt)t0({\sigma_{t}},{\sigma^{\prime}_{t}})_{t\geq 0} starting at (σ,σ)(\sigma,\sigma^{\prime}), we have

𝔼i=1maidisti(σt,σt)gti=1maidisti(σ,σ).\mathbb{E}\sum_{i=1}^{m}a_{i}\mathrm{dist}_{i}(\sigma_{t},\sigma^{\prime}_{t})\leq g^{t}\sum_{i=1}^{m}a_{i}\mathrm{dist}_{i}(\sigma,\sigma^{\prime}).

Moreover, for i=1,,mi=1,\dots,m,

𝔼disti(σt(i),σt(i))npigt.\mathbb{E}\mathrm{dist}_{i}({\sigma^{(i)}_{t}},{\sigma^{\prime(i)}_{t}})\leq n\sqrt{p_{i}}g^{t}.
Proof.

From \threfmonotonecontraction,

𝔼i=1maidisti(σt,σt)\displaystyle\mathbb{E}\sum_{i=1}^{m}a_{i}\mathrm{dist}_{i}(\sigma_{t},\sigma^{\prime}_{t}) =𝐚T(𝔼dist1(σt(1),σt(1))𝔼distm(σt(m),σt(m)))𝐚T𝐀t(dist1(σ(1),σ(1))distm(σ(m),σ(m)))\displaystyle=\mathbf{a}^{T}\begin{pmatrix}\mathbb{E}\mathrm{dist}_{1}({\sigma^{(1)}_{t}},{\sigma^{\prime(1)}_{t}})\\ \vdots\\ \mathbb{E}\mathrm{dist}_{m}({\sigma^{(m)}_{t}},{\sigma^{\prime(m)}_{t}})\end{pmatrix}\leq\mathbf{a}^{T}\mathbf{A}^{t}\begin{pmatrix}\mathrm{dist}_{1}(\sigma^{(1)},\sigma^{\prime(1)})\\ \vdots\\ \mathrm{dist}_{m}(\sigma^{(m)},\sigma^{\prime(m)})\end{pmatrix}
gt𝐚T(dist1(σ(1),σ(1))distm(σ(m),σ(m)))gti=1maidisti(σ,σ).\displaystyle\leq g^{t}\mathbf{a}^{T}\begin{pmatrix}\mathrm{dist}_{1}(\sigma^{(1)},\sigma^{\prime(1)})\\ \vdots\\ \mathrm{dist}_{m}(\sigma^{(m)},\sigma^{\prime(m)})\end{pmatrix}\leq g^{t}\sum_{i=1}^{m}a_{i}\mathrm{dist}_{i}(\sigma,\sigma^{\prime}).

Notice that distk(σt(k),σt(k))npk\mathrm{dist}_{k}({\sigma^{(k)}_{t}},{\sigma^{\prime(k)}_{t}})\leq np_{k} for each kk, so \threfnewlemma implies

𝔼disti(σt(i),σt(i))n𝐞iT𝐀t𝐩npigt.\mathbb{E}\mathrm{dist}_{i}({\sigma^{(i)}_{t}},{\sigma^{\prime(i)}_{t}})\leq n\mathbf{e}_{i}^{T}\mathbf{A}^{t}\mathbf{p}\leq n\sqrt{p_{i}}g^{t}.

We would like to translate \threflemma to the case of magnetization chains, which is done in \threfnormcontraction.

Lemma 2.7.
\thlabel

generalcontraction For starting magnetizations 𝐬=(s(1),,s(m))(s(1),,s(m))=𝐬\mathbf{s}=(s^{(1)},\dots,s^{(m)})\geq(s^{\prime(1)},\dots,s^{\prime(m)})=\mathbf{s}^{\prime}, the magnetization chains satisfy

𝟎(𝔼𝐬St(1)𝔼𝐬St(1)𝔼𝐬St(m)𝔼𝐬St(m))𝐀t(s(1)s(1)s(m)s(m)).\mathbf{0}\leq\begin{pmatrix}\mathbb{E}_{\mathbf{s}}{S^{(1)}_{t}}-\mathbb{E}_{\mathbf{s}^{\prime}}{S^{\prime(1)}_{t}}\\ \vdots\\ \mathbb{E}_{\mathbf{s}}{S^{(m)}_{t}}-\mathbb{E}_{\mathbf{s}^{\prime}}{S^{\prime(m)}_{t}}\end{pmatrix}\leq\mathbf{A}^{t}\begin{pmatrix}s^{(1)}-s^{\prime(1)}\\ \vdots\\ s^{(m)}-s^{\prime(m)}\end{pmatrix}.
Remark.

We say such pairs of starting magnetizations are monotone pairs.

Proof.

Let (σt,σt)({\sigma_{t}},{\sigma^{\prime}_{t}}) be a monotone coupling starting from (σ,σ)(\sigma,\sigma^{\prime}) where σσ\sigma\geq\sigma^{\prime} and S(i)(σ)=siS^{(i)}(\sigma)=s_{i}, S(i)(σ)=siS^{\prime(i)}(\sigma^{\prime})=s_{i}^{\prime} for i=1,,mi=1,\dots,m. Such a monotone coupling exists because of the given condition sisis_{i}\geq s_{i}^{\prime} for each ii. Since σiσi\sigma_{i}\geq\sigma_{i}^{\prime}, we have sisi=2ndisti(σi,σi)s_{i}-s_{i}^{\prime}=\frac{2}{n}\mathrm{dist}_{i}(\sigma_{i},\sigma_{i}^{\prime}) for each ii. By monotonicity, σt(i)σt(i){\sigma^{(i)}_{t}}\geq{\sigma^{\prime(i)}_{t}} for each ii. Thus, St(i)St(i)=|St(i)St(i)|=2ndisti(σt(i),σt(i))0{S^{(i)}_{t}}-{S^{\prime(i)}_{t}}=|{S^{(i)}_{t}}-{S^{\prime(i)}_{t}}|=\frac{2}{n}\mathrm{dist}_{i}({\sigma^{(i)}_{t}},{\sigma^{\prime(i)}_{t}})\geq 0 for each ii. Then, by \threfmonotonecontraction,

𝟎(𝔼σSt(1)𝔼σSt(1)𝔼σSt(m)𝔼σSt(m))=(𝔼σ,σ|St(1)St(1)|𝔼σ,σ|St(m)St(m)|)𝐀t(s(1)s(1)s(m)s(m)).\mathbf{0}\leq\begin{pmatrix}\mathbb{E}_{\sigma}{S^{(1)}_{t}}-\mathbb{E}_{\sigma^{\prime}}{S^{\prime(1)}_{t}}\\ \vdots\\ \mathbb{E}_{\sigma}{S^{(m)}_{t}}-\mathbb{E}_{\sigma^{\prime}}{S^{\prime(m)}_{t}}\end{pmatrix}=\begin{pmatrix}\mathbb{E}_{\sigma,\sigma^{\prime}}|{S^{(1)}_{t}}-{S^{\prime(1)}_{t}}|\\ \vdots\\ \mathbb{E}_{\sigma,\sigma^{\prime}}|{S^{(m)}_{t}}-{S^{\prime(m)}_{t}}|\end{pmatrix}\leq\mathbf{A}^{t}\begin{pmatrix}s^{(1)}-s^{\prime(1)}\\ \vdots\\ s^{(m)}-s^{\prime(m)}\end{pmatrix}.

Now, we can complete the proof since we have 𝔼σSt(i)𝔼σSt(i)=𝔼𝐬St(i)𝔼𝐬St(i)\mathbb{E}_{\sigma}{S^{(i)}_{t}}-\mathbb{E}_{\sigma^{\prime}}{S^{\prime(i)}_{t}}=\mathbb{E}_{\mathbf{s}}{S^{(i)}_{t}}-\mathbb{E}_{\mathbf{s}^{\prime}}{S^{\prime(i)}_{t}} for each ii by \threfmagmarkov. ∎

Recall that \circ denotes a Hadamard product.

Proposition 2.8.
\thlabel

normcontraction For a monotone coupling (σt,σt)t0(\sigma_{t},\sigma_{t}^{\prime})_{t\geq 0} starting at (σ,σ)(\sigma,\sigma^{\prime}) with magnetizations (𝐬,𝐬)(\mathbf{s},\mathbf{s}^{\prime}), we have

𝔼σ,σ𝐚𝐒t𝐚𝐒t1gt𝐚𝐬𝐚𝐬1.\displaystyle\mathbb{E}_{\sigma,\sigma^{\prime}}\|\mathbf{a}\circ\mathbf{S}_{t}-\mathbf{a}\circ\mathbf{S}_{t}^{\prime}\|_{1}\leq g^{t}\|\mathbf{a}\circ\mathbf{s}-\mathbf{a}\circ\mathbf{s}^{\prime}\|_{1}.

Moreover, not depending on the coupling, we have

𝔼𝐬𝐚𝐒t𝔼𝐬𝐚𝐒t1gt𝐚𝐬𝐚𝐬1.\displaystyle\|\mathbb{E}_{\mathbf{s}}\mathbf{a}\circ\mathbf{S}_{t}-\mathbb{E}_{\mathbf{s}^{\prime}}\mathbf{a}\circ\mathbf{S}_{t}^{\prime}\|_{1}\leq g^{t}\|\mathbf{a}\circ\mathbf{s}-\mathbf{a}\circ\mathbf{s}^{\prime}\|_{1}.
Proof.

For any magnetizations 𝐬𝐬(0)\mathbf{s}\equiv\mathbf{s}_{(0)} and 𝐬𝐬(m)\mathbf{s}^{\prime}\equiv\mathbf{s}_{(m)}, there exists 𝐬(1),,𝐬(m1)𝒮m\mathbf{s}_{(1)},\dots,\mathbf{s}_{(m-1)}\in\mathcal{S}\subset\mathbb{R}^{m} such that 𝐬(i1)𝐬(i)=𝐞i(s(i)s(i))\mathbf{s}_{(i-1)}-\mathbf{s}_{(i)}=\mathbf{e}_{i}(s^{(i)}-s^{\prime(i)}) for i=1,,mi=1,\dots,m. In particular, 𝐬(i1)\mathbf{s}_{(i-1)} and 𝐬(i)\mathbf{s}_{(i)} are a monotone pair for each ii. Then we can consider a monotone coupling (σ(0),t,,σ(m),t)t0(\sigma_{(0),t},\dots,\sigma_{(m),t})_{t\geq 0} with starting states (σ(0),,σ(m))(\sigma_{(0)},\dots,\sigma_{(m)}) such that σt=σ(0),t\sigma_{t}=\sigma_{(0),t}, σt=σ(m),t\sigma_{t}^{\prime}=\sigma_{(m),t} for t0t\geq 0, and the magnetization of the starting configuration σ(i)\sigma_{(i)} is 𝐬(i)\mathbf{s}_{(i)} for i=0,,mi=0,\dots,m.

Let 𝐒(j),t\mathbf{S}_{(j),t} be the magnetization chain corresponding to σ(j),t\sigma_{(j),t} for j=0,,mj=0,\dots,m. By telescoping, \threfgeneralcontraction gives

𝔼σ,σ𝐚𝐒t𝐚𝐒t1j=1m𝔼σ(j1),σ(j)𝐚𝐒(j1),t𝐚𝐒(j),t1\displaystyle\mathbb{E}_{\sigma,\sigma^{\prime}}\|\mathbf{a}\circ\mathbf{S}_{t}-\mathbf{a}\circ\mathbf{S}_{t}^{\prime}\|_{1}\leq\sum_{j=1}^{m}\mathbb{E}_{\sigma_{(j-1)},\sigma_{(j)}}\|\mathbf{a}\circ\mathbf{S}_{(j-1),t}-\mathbf{a}\circ\mathbf{S}_{(j),t}\|_{1}
j=1m𝐚T𝐀t𝐞j|s(j)s(j)|=gtj=1maj|s(j)s(j)|=gt𝐚𝐬𝐚𝐬1.\displaystyle\leq\sum_{j=1}^{m}\mathbf{a}^{T}\mathbf{A}^{t}\mathbf{e}_{j}|s^{(j)}-s^{\prime(j)}|=g^{t}\sum_{j=1}^{m}a_{j}|s^{(j)}-s^{\prime(j)}|=g^{t}\|\mathbf{a}\circ\mathbf{s}-\mathbf{a}\circ\mathbf{s}^{\prime}\|_{1}.

Then, the triangle inequality and \threfmagmarkov imply

𝔼𝐬𝐚𝐒t𝔼𝐬𝐚𝐒t1gt𝐚𝐬𝐚𝐬1.\|\mathbb{E}_{\mathbf{s}}\mathbf{a}\circ\mathbf{S}_{t}-\mathbb{E}_{\mathbf{s}^{\prime}}\mathbf{a}\circ\mathbf{S}_{t}^{\prime}\|_{1}\leq g^{t}\|\mathbf{a}\circ\mathbf{s}-\mathbf{a}\circ\mathbf{s}^{\prime}\|_{1}.

3. Variance bound of the magnetization in high temperatures

The next lemma is a generalization of Lemma 2.6 in [10] to Markov chains with a finite state space in m\mathbb{R}^{m}. Observe that for square-integrable m\mathbb{R}^{m}-valued i.i.d. random vectors X,YX,Y, we have 𝕍arX=12𝔼XY22\mathbb{V}\mathrm{ar}X=\frac{1}{2}\mathbb{E}\|X-Y\|_{2}^{2}.

Lemma 3.1.
\thlabel

variancebound Let (𝐙t)t0(\mathbf{Z}_{t})_{t\geq 0} be a Markov chain in a finite state space 𝒮~m\tilde{\mathcal{S}}\subseteq\mathbb{R}^{m}. Suppose that there exists 0<r<10<r<1 such that for any θ,θ𝒮~\theta,\theta^{\prime}\in\tilde{\mathcal{S}},

𝔼θ𝐙t𝔼θ𝐙t1rtθθ1.\|\mathbb{E}_{\theta}\mathbf{Z}_{t}-\mathbb{E}_{\theta^{\prime}}\mathbf{Z}_{t}^{\prime}\|_{1}\leq r^{t}\|\theta-\theta^{\prime}\|_{1}.

Then, for the l2l^{2} norm variance,

supθ𝒮𝕍arθ𝐙tmsupθ𝒮𝕍arθ𝐙1min{t,(1r2)1}.\sup_{\theta\in\mathcal{S}}\mathbb{V}\mathrm{ar}_{\theta}\mathbf{Z}_{t}\leq m\sup_{\theta\in\mathcal{S}}\mathbb{V}\mathrm{ar}_{\theta}\mathbf{Z}_{1}\>\min\{t,(1-r^{2})^{-1}\}.
Proof.

Put vt\colonequalssupθ𝒮𝕍arθ𝐙tv_{t}\colonequals\sup_{\theta\in\mathcal{S}}{\mathbb{V}\mathrm{ar}}_{\theta}\mathbf{Z}_{t}. Let (𝐙t)(\mathbf{Z}_{t}) and (𝐙t)(\mathbf{Z}_{t}^{\prime}) be independent copies of the chain starting from θ𝒮~\theta\in\tilde{\mathcal{S}}. The idea is to condition on the first step. Note that 𝐱2𝐱1m𝐱2\|\mathbf{x}\|_{2}\leq\|\mathbf{x}\|_{1}\leq\sqrt{m}\|\mathbf{x}\|_{2} for 𝐱m\mathbf{x}\in\mathbb{R}^{m}. Then by the observation right before the statement of this lemma,

12𝔼θ𝐙1𝐙112m12𝔼θ𝐙1𝐙122mv1.\frac{1}{2}\mathbb{E}_{\theta}\|\mathbf{Z}_{1}-\mathbf{Z}_{1}^{\prime}\|_{1}^{2}\leq m\frac{1}{2}\mathbb{E}_{\theta}\|\mathbf{Z}_{1}-\mathbf{Z}_{1}^{\prime}\|_{2}^{2}\leq mv_{1}.

By the assumption and Markov property, we have

𝔼θ[𝐙t|𝐙1]𝔼θ[𝐙t|𝐙1]1=𝔼𝐙1[𝐙t1]𝔼𝐙1[𝐙t1]1rt1𝐙1𝐙11.\|\mathbb{E}_{\theta}[\mathbf{Z}_{t}|\mathbf{Z}_{1}]-\mathbb{E}_{\theta}[\mathbf{Z}_{t}^{\prime}|\mathbf{Z}_{1}^{\prime}]\|_{1}=\|\mathbb{E}_{\mathbf{Z}_{1}}[\mathbf{Z}_{t-1}]-\mathbb{E}_{\mathbf{Z}_{1}^{\prime}}[\mathbf{Z}_{t-1}^{\prime}]\|_{1}\leq r^{t-1}\|\mathbf{Z}_{1}-\mathbf{Z}_{1}^{\prime}\|_{1}.

Thus, for θ𝒮~\theta\in\tilde{\mathcal{S}},

𝕍arθ[𝔼θ(𝐙t|𝐙1)]\displaystyle\mathbb{V}\mathrm{ar}_{\theta}[\mathbb{E}_{\theta}(\mathbf{Z}_{t}|\mathbf{Z}_{1})] =12𝔼θ𝔼𝐙1𝐙t1𝔼𝐙1𝐙t12212𝔼θ𝔼𝐙1𝐙t1𝔼𝐙1𝐙t112\displaystyle=\frac{1}{2}\mathbb{E}_{\theta}\|\mathbb{E}_{\mathbf{Z}_{1}}\mathbf{Z}_{t-1}-\mathbb{E}_{\mathbf{Z}_{1}^{\prime}}\mathbf{Z}_{t-1}^{\prime}\|_{2}^{2}\leq\frac{1}{2}\mathbb{E}_{\theta}\|\mathbb{E}_{\mathbf{Z}_{1}}\mathbf{Z}_{t-1}-\mathbb{E}_{\mathbf{Z}_{1}^{\prime}}\mathbf{Z}_{t-1}^{\prime}\|_{1}^{2}
12𝔼θ[r2(t1)𝐙1𝐙112]mv1r2(t1).\displaystyle\leq\frac{1}{2}\mathbb{E}_{\theta}\Big{[}r^{2(t-1)}\|\mathbf{Z}_{1}-\mathbf{Z}_{1}^{\prime}\|_{1}^{2}\Big{]}\leq mv_{1}r^{2(t-1)}.

By the Markov property, for every θ𝒮~\theta\in\tilde{\mathcal{S}}, 𝕍arθ[𝐙t|𝐙1]vt1\mathbb{V}\mathrm{ar}_{\theta}[\mathbf{Z}_{t}|\mathbf{Z}_{1}]\leq v_{t-1}, so

supθ𝒮𝔼θ[𝕍arθ[𝐙t|𝐙1]]vt1.\sup_{\theta\in\mathcal{S}}\mathbb{E}_{\theta}[\mathbb{V}\mathrm{ar}_{\theta}[\mathbf{Z}_{t}|\mathbf{Z}_{1}]]\leq v_{t-1}.

The total variance formula holds since we are using the l2l^{2} norm. Thus, taking supremum over θ𝒮~\theta\in\tilde{\mathcal{S}} in the total variance formula 𝕍arθ𝐙t=𝔼θ[𝕍arθ[𝐙t|𝐙1]]+𝕍arθ[𝔼θ[𝐙t|𝐙1]]\mathbb{V}\mathrm{ar}_{\theta}\mathbf{Z}_{t}=\mathbb{E}_{\theta}\big{[}\mathbb{V}\mathrm{ar}_{\theta}[\mathbf{Z}_{t}|\mathbf{Z}_{1}]\big{]}+\mathbb{V}\mathrm{ar}_{\theta}\big{[}\mathbb{E}_{\theta}[\mathbf{Z}_{t}|\mathbf{Z}_{1}]\big{]}, we have vtvt1+mv1r2(t1)v_{t}\leq v_{t-1}+mv_{1}r^{2(t-1)}. Upon iterating,

vtmv1t=1tr2(t1)mv1min{t,(1r2)1}.v_{t}\leq mv_{1}\sum_{t=1}^{t}r^{2(t-1)}\leq mv_{1}\min\big{\{}t,(1-r^{2})^{-1}\big{\}}.

The following proposition is an important result bounding the variance of magnetization chains uniformly in time.

Proposition 3.2.
\thlabel

magvariancebound Let β<βcr\beta<\beta_{cr}. For an arbitrary starting configuration 𝐬\mathbf{s} and t0t\geq 0, we have

i=1m𝕍ar𝐬(St(i))=C/n\sum_{i=1}^{m}\mathbb{V}\mathrm{ar}_{\mathbf{s}}({S^{(i)}_{t}})=C/n

where C>0C>0 only depends on p1,,pmp_{1},\dots,p_{m}, and β\beta.

Proof.

Observe that i=1m𝕍ar𝐬(aiSt(i))=𝕍ar𝐬(𝐚𝐒t)\sum_{i=1}^{m}\mathbb{V}\mathrm{ar}_{\mathbf{s}}(a_{i}S^{(i)}_{t})=\mathbb{V}\mathrm{ar}_{\mathbf{s}}(\mathbf{a}\circ\mathbf{S}_{t}). Note that increments of 𝐒t\mathbf{S}_{t} are bounded by 2/n2/n in absolute value. Then, from \threffinalfinallemma, we have

i=1m𝕍ar𝐬aiS1(i)a12(2/n)2.\sum_{i=1}^{m}\mathbb{V}\mathrm{ar}_{\mathbf{s}}{a_{i}S^{(i)}_{1}}\leq a_{1}^{2}(2/n)^{2}.

By \threffinalfinallemma, \threfnormcontraction, and \threfvariancebound, we have

am2i=1m𝕍ar𝐬(St(i))i=1m𝕍ar𝐬(aiSt(i))m4a12n211g2=4ma12υn(1+g)4ma12υn.a_{m}^{2}\sum_{i=1}^{m}\mathbb{V}\mathrm{ar}_{\mathbf{s}}({S^{(i)}_{t}})\leq\sum_{i=1}^{m}\mathbb{V}\mathrm{ar}_{\mathbf{s}}(a_{i}{S^{(i)}_{t}})\leq m\frac{4a_{1}^{2}}{n^{2}}\frac{1}{1-g^{2}}=\frac{4ma_{1}^{2}}{\upsilon n(1+g)}\leq\frac{4ma_{1}^{2}}{\upsilon n}.

Note that \threfupsilon assures υ>0\upsilon>0. ∎

We also establish a bound for the expected magnetization on subsets of partitions. To that end, we need the following observation.

Lemma 3.3.
\thlabel

zerospin For each iVi\in V, 𝔼μ(σ(i))=0\mathbb{E}_{\mu}(\sigma(i))=0 where μ\mu is the Gibbs distribution. In particular, we have 𝔼μ(S(i))=0\mathbb{E}_{\mu}(S^{(i)})=0.

Proof.

Since μ(σ)=μ(σ)\mu(\sigma)=\mu(-\sigma) for each configuration σ\sigma and σσ\sigma\mapsto-\sigma is a bijection from Ω\Omega into itself, we have 𝔼μ(σ(i))=σσ(i)μ(σ)=σ:σ(i)=1μ(σ)σ:σ(i)=1μ(σ)=0\mathbb{E}_{\mu}(\sigma(i))=\sum_{\sigma}\sigma(i)\mu(\sigma)=\sum_{\sigma:\sigma(i)=1}\mu(\sigma)-\sum_{\sigma:\sigma(i)=-1}\mu(\sigma)=0. ∎

Proposition 3.4 (Expected magnetization bound).
\thlabel

expectedmagbound Let β<βcr\beta<\beta_{cr} and 1im1\leq i\leq m. For any BJiB\subseteq J_{i} and a chain (σt)t0(\sigma_{t})_{t\geq 0} starting at σΩ\sigma\in\Omega, define Mt(B)\colonequals12kBσt(k)M_{t}(B)\colonequals\frac{1}{2}\sum_{k\in B}{\sigma_{t}}(k). Then

|𝔼σMt(B)||B|gt/pi.|\mathbb{E}_{\sigma}M_{t}(B)|\leq|B|g^{t}/\sqrt{p_{i}}.

Furthermore, for t12(1β/βcr)nlnnt\geq\frac{1}{2(1-\beta/\beta_{cr})}n\ln n, we have

𝕍arσ(Mt(B))=O(n),𝔼σ|Mt(B)|=O(n).\mathbb{V}\mathrm{ar}_{\sigma}(M_{t}(B))=O(n)\ ,\quad\mathbb{E}_{\sigma}|M_{t}(B)|=O(\sqrt{n}).
Proof.

Let "+" denote the configuration such that all spins are 11 and "-" denote the configuration with all spins 1-1. Let (σt+,σtμ,σt)(\sigma_{t}^{+},\sigma_{t}^{\mu},\sigma_{t}^{-}) be a monotone coupling with starting configuration (+,μ,)(+,\mu,-) where μ\mu is the stationary distribution. Let i{1,,m}i\in\{1,\dots,m\}. By \threflemma and \threfzerospin,

𝔼+[Mt(Ji)+]𝔼+,μ|Mt(Ji)+Mt(Ji)μ|+𝔼μ[Mt(Ji)μ]npigt.\displaystyle\mathbb{E}_{+}[M_{t}(J_{i})^{+}]\leq\mathbb{E}_{+,\mu}|M_{t}(J_{i})^{+}-M_{t}(J_{i})^{\mu}|+\mathbb{E}_{\mu}[M_{t}(J_{i})^{\mu}]\leq n\sqrt{p_{i}}g^{t}.

Then, by symmetry, for vJiv\in J_{i}, 𝔼+[Mt(v)]npigt/|Ji|=gt/pi\mathbb{E}_{+}[M_{t}(v)]\leq n\sqrt{p_{i}}g^{t}/|J_{i}|=g^{t}/\sqrt{p_{i}}. Thus, by summing over sites in BB, 𝔼+[Mt(B)+]|B|gt/pi\mathbb{E}_{+}[M_{t}(B)^{+}]\leq|B|g^{t}/\sqrt{p_{i}}. However, for any configuration σ\sigma, by monotonicity, 𝔼+[Mt(B)+]𝔼σ[Mt(B)]𝔼[Mt(B)]\mathbb{E}_{+}[M_{t}(B)^{+}]\geq\mathbb{E}_{\sigma}[M_{t}(B)]\geq\mathbb{E}_{-}[M_{t}(B)^{-}]. Considering the remark after \threfmagmarkov, 𝔼[Mt(B)]=𝔼+[Mt(B)+]\mathbb{E}_{-}[M_{t}(B)^{-}]=-\mathbb{E}_{+}[M_{t}(B)^{+}]. Thus, |𝔼σ[Mt(B)]||𝔼+[Mt(B)+]||B|gt/pi|\mathbb{E}_{\sigma}[M_{t}(B)]|\leq|\mathbb{E}_{+}[M_{t}(B)^{+}]|\leq|B|g^{t}/\sqrt{p_{i}} for any σ\sigma.

Now, by \threfmagvariancebound, O(1/n)=𝕍arSt(i)=𝕍ar(Mt(Ji)2/n)O(1/n)=\mathbb{V}\mathrm{ar}{S^{(i)}_{t}}=\mathbb{V}\mathrm{ar}(M_{t}(J_{i})2/n), so

𝕍ar+(Mt(Ji))=O(n).\mathbb{V}\mathrm{ar}_{+}(M_{t}(J_{i}))=O(n).

Thus, for t12(1β/βcr)nlnnt\geq\frac{1}{2(1-\beta/\beta_{cr})}n\ln n,

𝔼+(Mt(Ji)2)=𝕍ar+(Mt(Ji))+(𝔼+Mt(Ji))2=O(n)\mathbb{E}_{+}(M_{t}(J_{i})^{2})=\mathbb{V}\mathrm{ar}_{+}(M_{t}(J_{i}))+(\mathbb{E}_{+}M_{t}(J_{i}))^{2}=O(n)

However, by symmetry, for any fixed v1,v2Jiv_{1},v_{2}\in J_{i},

𝔼+(Mt(Ji)2)=npi+(npi2)𝔼+(σt+(v1)σt+(v2)).\mathbb{E}_{+}(M_{t}(J_{i})^{2})=np_{i}+\binom{np_{i}}{2}\mathbb{E}_{+}(\sigma_{t}^{+}(v_{1})\sigma_{t}^{+}(v_{2})).

Thus,

|𝔼+σt+(v1)σt+(v2)|=O(1/n).|\mathbb{E}_{+}\sigma_{t}^{+}(v_{1})\sigma_{t}^{+}(v_{2})|=O(1/n).

Likewise, for BJiB\subseteq J_{i},

𝔼+(Mt(B)2)=|B|+(|B|2)𝔼+(σt+(v1)σt+(v2))O(n).\mathbb{E}_{+}(M_{t}(B)^{2})=|B|+\binom{|B|}{2}\mathbb{E}_{+}(\sigma_{t}^{+}(v_{1})\sigma_{t}^{+}(v_{2}))\leq O(n).

Similarly, 𝔼Mt(B)2O(n)\mathbb{E}_{-}M_{t}(B)^{2}\leq O(n), so from (Mt(B))2(Mt(B)+)2+(Mt(B))2(M_{t}(B))^{2}\leq(M_{t}(B)^{+})^{2}+(M_{t}(B)^{-})^{2},

𝔼(Mt(B)2)=O(n)\mathbb{E}(M_{t}(B)^{2})=O(n)

whenever t12(1β/βcr)nlnnt\geq\frac{1}{2(1-\beta/\beta_{cr})}n\ln n. Thus, for t12(1β/βcr)nlnnt\geq\frac{1}{2(1-\beta/\beta_{cr})}n\ln n,

𝕍arσ(Mt(B))=O(n).\mathbb{V}\mathrm{ar}_{\sigma}(M_{t}(B))=O(n).

Lastly, for t12(1β/βcr)nlnnt\geq\frac{1}{2(1-\beta/\beta_{cr})}n\ln n, from Jensen’s inequality,

𝔼σ|Mt(B)|\displaystyle\mathbb{E}_{\sigma}|M_{t}(B)| 𝔼σ|Mt(B)|2=(𝔼σ[Mt(B)])2+𝕍arσ(Mt(B))\displaystyle\leq\sqrt{\mathbb{E}_{\sigma}|M_{t}(B)|^{2}}=\sqrt{(\mathbb{E}_{\sigma}[M_{t}(B)])^{2}+\mathbb{V}\mathrm{ar}_{\sigma}(M_{t}(B))}
|𝔼σ[Mt(B)]|+𝕍arσ(Mt(B))=O(n).\displaystyle\leq|\mathbb{E}_{\sigma}[M_{t}(B)]|+\sqrt{\mathbb{V}\mathrm{ar}_{\sigma}(M_{t}(B))}=O(\sqrt{n}).

4. Couplings

Fix the notation

tn\colonequals12(1β/βcr)nlnn.t_{n}\colonequals\frac{1}{2(1-\beta/\beta_{cr})}n\ln n.
Definition (Modified matching).

Let σΩ\sigma\in\Omega and σΩ\sigma^{\prime}\in\Omega have magnetizations 𝐬𝒮\mathbf{s}\in\mathcal{S} and 𝐬𝒮\mathbf{s}^{\prime}\in\mathcal{S}, respectively. Consider two copies of the graph, V=iJiV=\bigcup_{i}J_{i} and V=iJiV^{\prime}=\bigcup_{i}J_{i}^{\prime}. Let i{1,,m}i\in\{1,\dots,m\}. If s(i)s(i)s^{(i)}\geq s^{\prime(i)}, then it is possible to match each site in JiJ_{i}^{\prime} with +1+1 spin to a site in JiJ_{i} with +1+1 spin. Any leftover sites in JiJ_{i}^{\prime} are arbitrarily matched to the leftover sites in JiJ_{i}. We match the sites in a similar way whenever s(i)s(i)s^{(i)}\leq s^{\prime(i)}. This defines a bijection fσ,σ:VVf_{\sigma,\sigma^{\prime}}\colon V\to V^{\prime}.

We call this bijection a modified matching of σ\sigma and σ\sigma^{\prime}.

Definition (Modified monotone update and coupling).

Let fσ,σ:VVf_{\sigma,\sigma^{\prime}}\colon V\to V^{\prime} be a modified matching of σ,σΩ\sigma,\sigma^{\prime}\in\Omega. Let II and UU be uniformly distributed over V=i=1mJiV=\bigcup_{i=1}^{m}J_{i} and [0,1][0,1]\subseteq\mathbb{R}, respectively, and be independent. Suppose IJηI\in J_{\eta} for some η{1,,m}\eta\in\{1,\dots,m\} is the chosen site in VV. Consider the case vJησ(v)vJησ(v)\sum_{v\notin J_{\eta}}\sigma(v)\leq\sum_{v\notin J_{\eta}}\sigma^{\prime}(v). If

U<1+tanh(βvJησ(v))2,U<\frac{1+\tanh\left(\beta\sum_{v\notin J_{\eta}}\sigma(v)\right)}{2},

then update the chosen site II of VV by +1 and fσ,σ(I)f_{\sigma,\sigma^{\prime}}(I) of VV^{\prime} by +1. If

U1+tanh(βvJησ(v))2,U\geq\frac{1+\tanh\left(\beta\sum_{v\notin J_{\eta}}\sigma^{\prime}(v)\right)}{2},

then update the chosen site II of VV by -1 and fσ,σ(I)f_{\sigma,\sigma^{\prime}}(I) of VV^{\prime} by -1. Otherwise, if

1+tanh(βvJησ(v))2U<1+tanh(βvJησ(v))2,\frac{1+\tanh\left(\beta\sum_{v\notin J_{\eta}}\sigma(v)\right)}{2}\leq U<\frac{1+\tanh\left(\beta\sum_{v\notin J_{\eta}}\sigma^{\prime}(v)\right)}{2},

then update the chosen site II of VV by -1 and fσ,σ(I)f_{\sigma,\sigma^{\prime}}(I) of VV^{\prime} by +1. The other case vJησ(v)>vJησ(v)\sum_{v\notin J_{\eta}}\sigma(v)>\sum_{v\notin J_{\eta}}\sigma^{\prime}(v) can similarly be updated.

Given the chosen site II, we call the above procedure of deciding the updating spin in the two chains a modified monotone update with respect to the given modified matching.

Now, fix a modified matching fσ,σf_{\sigma,\sigma^{\prime}} of σ\sigma and σ\sigma^{\prime}. Let σt{\sigma_{t}} and σt{\sigma^{\prime}_{t}} be chains starting at σ\sigma and σ\sigma^{\prime}, respectively. Repeating the above procedure independently for each step with respect to fσ,σf_{\sigma,\sigma^{\prime}} gives a coupling of the Glauber dynamics. We call this coupling a modified monotone coupling with respect to the given modified matching.

Remark.
\thref

monotonecontraction and its consequences hold with a suitable distance function for a modified coupling with respect to a given modified matching.

We first construct a coupling such that the magnetizations agree after tn+O(n)t_{n}+O(n) steps in the next two lemmas.

Lemma 4.1 (Lemma 2.4, [10]).
\thlabel

supermartingale Let (Wt)t0(W_{t})_{t\geq 0} be a non-negative supermartingale with a stopping time τ\tau satisfying
(i)  W0=kW_{0}=k
(ii)  Wt+1WtB<W_{t+1}-W_{t}\leq B<\infty
(iii)   𝕍ar(Wt+1|t)>σ2>0on the event{τ>t}\mathbb{V}\mathrm{ar}(W_{t+1}|\mathcal{F}_{t})>\sigma^{2}>0\ \text{on the event}\ \{\tau>t\}. Then for u>4B23σ2u>\frac{4B^{2}}{3\sigma^{2}},

k(τ>u)4kσu.\mathbb{P}_{k}(\tau>u)\leq\frac{4k}{\sigma\sqrt{u}}.
Lemma 4.2 (Magnetization coupling).
\thlabel

magcoupling Let β<βcr\beta<\beta_{cr}. For any configurations σ\sigma and σ\sigma^{\prime}, there exists a coupling (σt,σt)({\sigma_{t}},{\sigma^{\prime}_{t}}) with starting states (σ,σ)(\sigma,\sigma^{\prime}) satisfying the following condition. If τmag\colonequalsmin{t0:𝐒t=𝐒t}\tau_{mag}\colonequals\min\{t\geq 0:\mathbf{S}_{t}=\mathbf{S}_{t}^{\prime}\}, then for large γn\gamma n,

σ,σ(τmag>tn+γn)cγ\mathbb{P}_{\sigma,\sigma^{\prime}}(\tau_{mag}>t_{n}+\gamma n)\leq\frac{c}{\sqrt{\gamma}}

where c>0c>0 is a constant not depending on σ\sigma, σ\sigma^{\prime}, or nn.

Proof.

Let (σt,σt)({\sigma_{t}},{\sigma^{\prime}_{t}}) be a monotone coupling with starting states (σ,σ)(\sigma,\sigma^{\prime}). Put Yi,t\colonequalsn2ai|St(i)St(i)|Y_{i,t}\colonequals\frac{n}{2}a_{i}|{S^{(i)}_{t}}-{S^{\prime(i)}_{t}}| for i=1,,mi=1,\dots,m and Ytot,t\colonequalsi=1mYi,tY_{tot,t}\colonequals\sum_{i=1}^{m}Y_{i,t}. Define

τ\colonequalsmin{ttn:max1imYi,t/ai1}.\tau\colonequals\min\{t\geq t_{n}:\max_{1\leq i\leq m}Y_{i,t}/a_{i}\leq 1\}.

By \threfnormcontraction,

𝔼σ,σ[Ytot,tn]cn\mathbb{E}_{\sigma,\sigma^{\prime}}[Y_{tot,t_{n}}]\leq c\sqrt{n}

for some c>0c>0.

We construct a coupling such that (Ytot,t)tnt<τ(Y_{tot,t})_{t_{n}\leq t<\tau} is a positive supermartingale with bounded increments and the conditional probability of not being lazy is bounded away from zero uniformly in time and nn.

To that end, consider a time tnt<τt_{n}\leq t<\tau. Define Kt\colonequalsi:Yi,t/ai1JiK_{t}\colonequals\bigcup_{i:Y_{i,t}/a_{i}\leq 1}J_{i}, Lt\colonequalsi:Yi,t/ai>1JiL_{t}\colonequals\bigcup_{i:Y_{i,t}/a_{i}>1}J_{i}, and Lt\colonequalsi:Yi,t/ai>1JiL_{t}^{\prime}\colonequals\bigcup_{i:Y_{i,t}/a_{i}>1}J_{i}^{\prime}. Note that LtL_{t}\neq\emptyset since t<τt<\tau. Choose a site equiprobably over V=Kt˙LtV=K_{t}\dot{\cup}L_{t}. Let ftf_{t} be the modified matching of σt{\sigma_{t}} and σt{\sigma^{\prime}_{t}}. If a site in KtK_{t} is chosen, then use the modified monotone update with respect to ftf_{t} to update (σt,σt)({\sigma_{t}},{\sigma^{\prime}_{t}}). If a site in LtL_{t} is chosen, then independently choose another site equiprobably over LtL_{t}^{\prime} (which can be the same site) to update σt{\sigma^{\prime}_{t}} independent of σt{\sigma_{t}}. It is easy to check that the above is a coupling of the Glauber dynamics.

Clearly, Ytot,tY_{tot,t} has bounded increment with the above coupling. Let II be a random variable uniformly distributed over VV which is independent of t\mathcal{F}_{t}. Let E={ILt,σt(I)=+1,σt+1(I)=1,σt+1(ft(I))=1}E=\{I\in L_{t},{\sigma_{t}}(I)=+1,\sigma_{t+1}(I)=-1,\sigma_{t+1}^{\prime}(f_{t}(I))=1\} and F={ILt,σt(I)=1,σt+1(I)=+1,σt+1(ft(I))=1}F=\{I\in L_{t},{\sigma_{t}}(I)=-1,\sigma_{t+1}(I)=+1,\sigma_{t+1}^{\prime}(f_{t}(I))=-1\}. Since LtL_{t}\neq\emptyset implies |Lt|/np1|L_{t}|/n\geq p_{1}, we obtain that (Ytot,t+1Ytot,t|t)\mathbb{P}(Y_{tot,t+1}\neq Y_{tot,t}|\mathcal{F}_{t}) is bounded below by

(Ytot,t+1Ytot,t,ILt|t)(E˙F|t)\displaystyle\geq\mathbb{P}(Y_{tot,t+1}\neq Y_{tot,t},I\in L_{t}|\mathcal{F}_{t})\geq\mathbb{P}(E\dot{\cup}F|\mathcal{F}_{t})
|Lt|+iLtσt(i)2n(1tanh(β(1p1))2)2\displaystyle\geq\frac{|L_{t}|+\sum_{i\in L_{t}}{\sigma_{t}}(i)}{2n}\biggl{(}\frac{1-\tanh(\beta(1-p_{1}))}{2}\biggr{)}^{2}
+|Lt|iLtσt(i)2n(1tanh(β(1p1))2)2\displaystyle\enspace+\frac{|L_{t}|-\sum_{i\in L_{t}}{\sigma_{t}}(i)}{2n}\biggl{(}\frac{1-\tanh(\beta(1-p_{1}))}{2}\biggr{)}^{2}
p1(1tanh(β(1p1))2)2>0.\displaystyle\geq p_{1}\biggl{(}\frac{1-\tanh(\beta(1-p_{1}))}{2}\biggr{)}^{2}>0.

Finally, we need to show the supermartingale property. Consider Y1,t+1/a1Y1,t/a1Y_{1,t+1}/a_{1}-Y_{1,t}/a_{1}. Suppose J1KtJ_{1}\subseteq K_{t}. Then by a direct calculation, on the event {J1Kt}\{J_{1}\subseteq K_{t}\}, it holds that 𝔼(Y1,t+1/a1Y1,t/a1|t)\mathbb{E}(Y_{1,t+1}/a_{1}-Y_{1,t}/a_{1}|\mathcal{F}_{t}) is bounded above by

(p1|St(1)St(1)|2)|tanh(βj1St(j))tanh(βj1St(j))|2\displaystyle\leq\biggl{(}p_{1}-\frac{|{S^{(1)}_{t}}-{S^{\prime(1)}_{t}}|}{2}\biggr{)}\frac{|\tanh(\beta\sum_{j\neq 1}{S^{(j)}_{t}})-\tanh(\beta\sum_{j\neq 1}{S^{\prime(j)}_{t}})|}{2}
|St(1)St(1)|2(1|tanh(βj1St(j))tanh(βj1St(j))|2)\displaystyle\enspace-\frac{|{S^{(1)}_{t}}-{S^{\prime(1)}_{t}}|}{2}\biggl{(}1-\frac{|\tanh(\beta\sum_{j\neq 1}{S^{(j)}_{t}})-\tanh(\beta\sum_{j\neq 1}{S^{\prime(j)}_{t}})|}{2}\biggr{)}
12(|St(1)St(1)|+p1tanh(β|j1St(j)j1St(j)|)).\displaystyle\leq\frac{1}{2}\biggl{(}-|{S^{(1)}_{t}}-{S^{\prime(1)}_{t}}|+p_{1}\tanh\biggl{(}\beta\Bigl{|}\sum_{j\neq 1}{S^{(j)}_{t}}-\sum_{j\neq 1}{S^{\prime(j)}_{t}}\Bigl{|}\biggr{)}\biggr{)}.

Suppose J1LtJ_{1}\subseteq L_{t}. Note that Y1,t>1Y_{1,t}>1 implies (St+1(1)St+1(1))(St(1)St(1))0({S^{(1)}_{t+1}}-{S^{\prime(1)}_{t+1}})({S^{(1)}_{t}}-{S^{\prime(1)}_{t}})\geq 0 and |St(1)St(1)|>0|{S^{(1)}_{t}}-{S^{\prime(1)}_{t}}|>0. Let ξ=(St(1)St(1))/|St(1)St(1)|{±1}\xi=({S^{(1)}_{t}}-{S^{\prime(1)}_{t}})/|{S^{(1)}_{t}}-{S^{\prime(1)}_{t}}|\in\{\pm 1\}. Then by equation (5) in Section 5.2, on the event {J1Lt}\{J_{1}\subseteq L_{t}\}, 𝔼(Y1,t+1/a1Y1,t/a1|t)\mathbb{E}(Y_{1,t+1}/a_{1}-Y_{1,t}/a_{1}|\mathcal{F}_{t}) is equal to

=ξn2(𝔼(St+1(1)St(1)|σt)𝔼(St+1(1)St(1)|σt))\displaystyle=\xi\frac{n}{2}\biggl{(}\mathbb{E}({S^{(1)}_{t+1}}-{S^{(1)}_{t}}|{\sigma_{t}})-\mathbb{E}({S^{\prime(1)}_{t+1}}-{S^{\prime(1)}_{t}}|{\sigma^{\prime}_{t}})\biggr{)}
=ξn21n(St(1)+p1tanh(βj1St(j)))\displaystyle=\xi\frac{n}{2}\frac{1}{n}\biggl{(}-{S^{(1)}_{t}}+p_{1}\tanh(\beta\sum_{j\neq 1}{S^{(j)}_{t}})\biggr{)}
ξn21n(St(1)+p1tanh(βj1St(j)))\displaystyle\quad-\xi\frac{n}{2}\frac{1}{n}\biggl{(}-{S^{\prime(1)}_{t}}+p_{1}\tanh(\beta\sum_{j\neq 1}{S^{\prime(j)}_{t}})\biggr{)}
=ξ2((St(1)St(1))+p1(tanh(βj1St(j))tanh(βj1St(j))))\displaystyle=\frac{\xi}{2}\biggl{(}-({S^{(1)}_{t}}-{S^{\prime(1)}_{t}})+p_{1}\biggl{(}\tanh(\beta\sum_{j\neq 1}{S^{(j)}_{t}})-\tanh(\beta\sum_{j\neq 1}{S^{\prime(j)}_{t}})\biggr{)}\biggr{)}
12(|St(1)St(1)|+p1tanh(β|j1St(j)j1St(j)|)).\displaystyle\leq\frac{1}{2}\biggl{(}-|{S^{(1)}_{t}}-{S^{\prime(1)}_{t}}|+p_{1}\tanh\biggl{(}\beta\Bigl{|}\sum_{j\neq 1}{S^{(j)}_{t}}-\sum_{j\neq 1}{S^{\prime(j)}_{t}}\Bigl{|}\biggr{)}\biggr{)}.

Since either J1LtJ_{1}\subseteq L_{t} or J1KtJ_{1}\subseteq K_{t} must hold, 𝔼(Y1,t+1/a1Y1,t/a1|t)\mathbb{E}(Y_{1,t+1}/a_{1}-Y_{1,t}/a_{1}|\mathcal{F}_{t}) is equal to

=𝟙J1Kt𝔼(Y1,t+1Y1,t|t)+𝟙J1Lt𝔼(Y1,t+1Y1,t|t)\displaystyle=\mathbbm{1}_{J_{1}\subseteq K_{t}}\mathbb{E}(Y_{1,t+1}-Y_{1,t}|\mathcal{F}_{t})+\mathbbm{1}_{J_{1}\subseteq L_{t}}\mathbb{E}(Y_{1,t+1}-Y_{1,t}|\mathcal{F}_{t})
12(|St(1)St(1)|+p1tanh(β|j1St(j)j1St(j)|))\displaystyle\leq\frac{1}{2}\biggl{(}-|{S^{(1)}_{t}}-{S^{\prime(1)}_{t}}|+p_{1}\tanh\biggl{(}\beta\Bigl{|}\sum_{j\neq 1}{S^{(j)}_{t}}-\sum_{j\neq 1}{S^{\prime(j)}_{t}}\Bigr{|}\biggr{)}\biggr{)}
12(|St(1)St(1)|+p1βj1|St(j)St(j)|).\displaystyle\leq\frac{1}{2}\biggl{(}-|{S^{(1)}_{t}}-{S^{\prime(1)}_{t}}|+p_{1}\beta\sum_{j\neq 1}\Bigl{|}{S^{(j)}_{t}}-{S^{\prime(j)}_{t}}\Bigl{|}\biggr{)}.

Thus,

𝔼(Y1,t+1/a1|t)(11n)Y1,t/a1+βp1nj1Yj,t/aj.\mathbb{E}(Y_{1,t+1}/a_{1}|\mathcal{F}_{t})\leq(1-\frac{1}{n})Y_{1,t}/a_{1}+\frac{\beta p_{1}}{n}\sum_{j\neq 1}Y_{j,t}/a_{j}.

Putting in the matrix form with 𝐘~t\colonequals(Y1,t/a1,,Ym,t/am)T\tilde{\mathbf{Y}}_{t}\colonequals(Y_{1,t}/a_{1},\dots,Y_{m,t}/a_{m})^{T}, we have

𝔼(Ytot,t+1|t)=𝐚T𝔼(𝐘~t+1|t)𝐚T𝐀𝐘~t=g𝐚T𝐘~t=gYtot,t.\displaystyle\mathbb{E}(Y_{tot,t+1}|\mathcal{F}_{t})=\mathbf{a}^{T}\mathbb{E}(\tilde{\mathbf{Y}}_{t+1}|\mathcal{F}_{t})\leq\mathbf{a}^{T}\mathbf{A}\tilde{\mathbf{Y}}_{t}=g\mathbf{a}^{T}\tilde{\mathbf{Y}}_{t}=gY_{tot,t}.

Since β<βcr\beta<\beta_{cr} implies g<1g<1 by \threfupsilon, the supermartingale property is established.

With the above coupling, by \threfsupermartingale, for large γn\gamma n,

σ,σ(τ>tn+γn|σtn,σtn)cn(Stn(1),,Stn(m))(Stn(1),,Stn(m))1γn\mathbb{P}_{\sigma,\sigma^{\prime}}(\tau>t_{n}+\gamma n|\sigma_{t_{n}},\sigma^{\prime}_{t_{n}})\leq c^{\prime}\frac{n\|(S^{(1)}_{t_{n}},\dots,S^{(m)}_{t_{n}})-(S^{\prime(1)}_{t_{n}},\dots,S^{\prime(m)}_{t_{n}})\|_{1}}{\sqrt{\gamma n}}

for some c>0c^{\prime}>0 not depending on nn. Taking expectation,

σ,σ(τ>tn+γn)O(γ1/2).\mathbb{P}_{\sigma,\sigma^{\prime}}(\tau>t_{n}+\gamma n)\leq O(\gamma^{-1/2}).

Note στ\sigma_{\tau} has at most mm more +1 spin sites than στ\sigma_{\tau}^{\prime}, so 0Ytot,τa1m0\leq Y_{tot,\tau}\leq a_{1}m by \threffinalfinallemma. At τ\tau, construct a modified matching of στ\sigma_{\tau} and στ\sigma_{\tau}^{\prime}, and use the modified monotone coupling with respect to this modified matching from then on. At τmag\tau_{mag}, we construct another modified matching of the sites to do a new modified monotone coupling so that (St(1),,St(m))=(St(1),,St(m))({S^{(1)}_{t}},\dots,{S^{(m)}_{t}})=({S^{\prime(1)}_{t}},\dots,{S^{\prime(m)}_{t}}) forever after τmag\tau_{mag}.

By \threffinalfinallemma, a modified version of \threfnormcontraction, and the strong Markov property, we have

σ,σ(τmag>τ+γn|στ,στ)\displaystyle\mathbb{P}_{\sigma,\sigma^{\prime}}(\tau_{mag}>\tau+\gamma^{\prime}n|\sigma_{\tau},\sigma_{\tau}^{\prime}) σ,σ(Ytot,τ+γnam|στ,στ)\displaystyle\leq\mathbb{P}_{\sigma,\sigma^{\prime}}(Y_{tot,\tau+\gamma^{\prime}n}\geq a_{m}|\sigma_{\tau},\sigma_{\tau}^{\prime})
𝔼σ,σ[Ytot,τ+γn|στ,στ]/am\displaystyle\leq\mathbb{E}_{\sigma,\sigma^{\prime}}[Y_{tot,\tau+\gamma^{\prime}n}|\sigma_{\tau},\sigma_{\tau}^{\prime}]/a_{m}
gγnYtot,τ/amgγna1m/ameυγa1m/am.\displaystyle\leq g^{\gamma^{\prime}n}Y_{tot,\tau}/a_{m}\leq g^{\gamma^{\prime}n}a_{1}m/a_{m}\leq e^{-\upsilon\gamma^{\prime}}a_{1}m/a_{m}.

Thus,

σ,σ(τmag>tn+(γ+γ)n)\displaystyle\mathbb{P}_{\sigma,\sigma^{\prime}}(\tau_{mag}>t_{n}+(\gamma+\gamma^{\prime})n) O(γ1/2)+eυγa1m/am,\displaystyle\leq O(\gamma^{-1/2})+e^{-\upsilon\gamma^{\prime}}a_{1}m/a_{m},

and putting γ=γ\gamma=\gamma^{\prime} yields

σ,σ(τmag>tn+γn)O(γ1/2).\displaystyle\mathbb{P}_{\sigma,\sigma^{\prime}}(\tau_{mag}>t_{n}+\gamma n)\leq O(\gamma^{-1/2}).

Definition (Good configurations).

Define the set of "good" configurations by

Ω~\colonequals{σΩ:|S(i)(σ)|pi/2,i=1,,m}.\tilde{\Omega}\colonequals\{\sigma\in\Omega:|S^{(i)}(\sigma)|\leq p_{i}/2,\ i=1,\dots,m\}.

For σ=(σ(1),,σ(m))Ω~\sigma=(\sigma^{(1)},\dots,\sigma^{(m)})\in\tilde{\Omega} and each ii, define

uiσ\colonequals|{vJi:σ(i)(v)=1}|,viσ\colonequals|{vJi:σ(i)(v)=1}|.\displaystyle u_{i}^{\sigma}\colonequals|\{v\in J_{i}:\sigma^{(i)}(v)=1\}|,\enspace v_{i}^{\sigma}\colonequals|\{v\in J_{i}:\sigma^{(i)}(v)=-1\}|.

Define

Λ~\colonequals{(u1,v1,u2,v2,,um,vm)2m:|Ji|/4uivi,i=1,,m}.\tilde{\Lambda}\colonequals\{(u_{1},v_{1},u_{2},v_{2},\dots,u_{m},v_{m})\in\mathbb{N}^{2m}:{|J_{i}|}/{4}\leq u_{i}\wedge v_{i},\ i=1,\dots,m\}.
Remark.

Note that σΩ~(u1σ,v1σ,,umσ,vmσ)Λ~\sigma\in\tilde{\Omega}\iff(u_{1}^{\sigma},v_{1}^{\sigma},\dots,u_{m}^{\sigma},v_{m}^{\sigma})\in\tilde{\Lambda}. In other words, Λ~\tilde{\Lambda} is another representation of good configurations Ω~\tilde{\Omega}. We omit the starting state and write uiu_{i} instead of uiσu_{i}^{\sigma} for convenience.

Lemma 4.3 (Lemma 3.3, [10]).
\thlabel

goodstartingstatesdistance For any subset AΩA\subseteq\Omega and stationary distribution π\pi,

dn(t0+t)\displaystyle d_{n}(t_{0}+t) =maxσΩσ(σt0+t)πTV\displaystyle=\max_{\sigma\in\Omega}\|\mathbb{P}_{\sigma}(\sigma_{t_{0}+t}\in\cdot)-\pi\|_{TV}
maxσAσ(σt)πTV+maxσΩσ(σt0A).\displaystyle\leq\max_{\sigma\in A}\|\mathbb{P}_{\sigma}({\sigma_{t}}\in\cdot)-\pi\|_{TV}+\max_{\sigma\in\Omega}\mathbb{P}_{\sigma}(\sigma_{t_{0}}\notin A).

Recall that we are assuming the high temperature regime. By \threfexpectedmagbound, there exists δ>0\delta>0 such that maxσΩ,1im|𝔼σSδn(i)|p1/4\max_{\sigma\in\Omega,1\leq i\leq m}|\mathbb{E}_{\sigma}S^{(i)}_{\delta n}|\leq p_{1}/4. Hence, by \threfmagvariancebound, for large nn,

σ(σδnΩ~)\displaystyle\mathbb{P}_{\sigma}(\sigma_{\delta n}\notin\tilde{\Omega}) i=1mσ(|Sδn(i)|>pi/2)i=1mσ(|Sδn(i)𝔼σSδn(i)|>pi/4)\displaystyle\leq\sum_{i=1}^{m}\mathbb{P}_{\sigma}(|S^{(i)}_{\delta n}|>p_{i}/2)\leq\sum_{i=1}^{m}\mathbb{P}_{\sigma}(|S^{(i)}_{\delta n}-\mathbb{E}_{\sigma}S^{(i)}_{\delta n}|>p_{i}/4)
16p12i=1m𝕍arσSδn(i)=O(1/n).\displaystyle\leq\frac{16}{p_{1}^{2}}\sum_{i=1}^{m}\mathbb{V}\mathrm{ar}_{\sigma}S^{(i)}_{\delta n}=O(1/n).

Combining with \threfgoodstartingstatesdistance,

dn(δn+t)maxσΩ~σ(σt)μTV+O(1/n).d_{n}(\delta n+t)\leq\max_{\sigma\in\tilde{\Omega}}\|\mathbb{P}_{\sigma}({\sigma_{t}}\in\cdot)-\mu\|_{TV}+O(1/n). (4)
Definition (2m2m-coordinate chain).

Let σ~Ω\tilde{\sigma}\in\Omega be a reference configuration. For σΩ\sigma\in\Omega and each ii, define

Ui(σ)\colonequals|{vJi:σ(i)(v)=σ~(i)(v)=1}|,\displaystyle U_{i}(\sigma)\colonequals|\{v\in J_{i}:\sigma^{(i)}(v)=\tilde{\sigma}^{(i)}(v)=1\}|,
Vi(σ)\colonequals|{vJi:σ(i)(v)=σ~(i)(v)=1}|.\displaystyle V_{i}(\sigma)\colonequals|\{v\in J_{i}:\sigma^{(i)}(v)=\tilde{\sigma}^{(i)}(v)=-1\}|.

For a chain (σt)({\sigma_{t}}) with the starting configuration σ0Ω\sigma_{0}\in\Omega, define the 2m2m-coordinate chain with respect to σ~\tilde{\sigma} by

𝐔t\colonequals(Ut(1),Vt(1),,Ut(m),Vt(m))\colonequals(U1(σt),V1(σt),,Um(σt),Vm(σt)).\displaystyle\mathbf{U}_{t}\colonequals({U^{(1)}_{t}},{V^{(1)}_{t}},\dots,{U^{(m)}_{t}},{V^{(m)}_{t}})\colonequals(U_{1}({\sigma_{t}}),V_{1}({\sigma_{t}}),\dots,U_{m}({\sigma_{t}}),V_{m}({\sigma_{t}})).

It is easy to see that the 2m2m-coordinate chain is again a Markov chain in its state space 𝒰2m\mathcal{U}\subseteq\mathbb{N}^{2m} and determines the magnetization chain (St(1),,St(m))({S^{(1)}_{t}},\dots,{S^{(m)}_{t}}) through the relation St(i)=2(Ut(i)Vt(i))/n(u~iv~i)/n{S^{(i)}_{t}}=2({U^{(i)}_{t}}-V^{(i)}_{t})/n-(\tilde{u}_{i}-\tilde{v}_{i})/{n} for i=1,,mi=1,\dots,m.

Symmetry gives us the following lemma which is an adaptation of Lemma 3.4 in [10].

Lemma 4.4.
\thlabel

totalvariationdistance Let (σt)({\sigma_{t}}) be a chain starting at σΩ\sigma\in\Omega. Consider the corresponding 2m2m-coordinate chain starting at 𝐮𝒰\mathbf{u}\in\mathcal{U}. Then

σ(σt)μTV=𝐮((Ut(1),Vt(1),,Ut(m),Vt(m)))νTV\|\mathbb{P}_{\sigma}({\sigma_{t}}\in\cdot)-\mu\|_{TV}=\|\mathbb{P}_{\mathbf{u}}(({U^{(1)}_{t}},{V^{(1)}_{t}},\dots,{U^{(m)}_{t}},{V^{(m)}_{t}})\in\cdot)-\nu\|_{TV}

where ν\nu is the stationary distribution of the 2m2m-coordinate chain.

Proof.

Since μ(σ)=eβnijS(i)(σ)S(j)(σ)/Z(β)\mu(\sigma)=e^{\beta n\sum_{i\neq j}S^{(i)}(\sigma)S^{(j)}(\sigma)}/Z(\beta), given the 2m2m-coordinate 𝐮𝒰\mathbf{u}^{\prime}\in\mathcal{U}, the conditional μ\mu-probability of the configurations is equiprobable. In other words, μ(|Ω(𝐮))\mu(\cdot|\Omega(\mathbf{u}^{\prime})) is uniform where Ω(𝐮)\Omega(\mathbf{u}^{\prime}) is the set of configurations having the 2m2m-coordinate 𝐮\mathbf{u}^{\prime}. Also, by symmetry,

σ(σt|𝐔t=𝐮)\mathbb{P}_{\sigma}({\sigma_{t}}\in\cdot\ |\mathbf{U}_{t}=\mathbf{u}^{\prime})

is uniform over Ω(𝐮)\Omega(\mathbf{u}^{\prime}). Thus,

σ(σt=η)μ(η)=𝐮𝒰𝟙{ηΩ(𝐮)}|Ω(𝐮)|(𝐮(𝐔t=𝐮)μ(Ω(𝐮))).\displaystyle\mathbb{P}_{\sigma}({\sigma_{t}}=\eta)-\mu(\eta)=\sum_{\mathbf{u}^{\prime}\in\mathcal{U}}\frac{\mathbbm{1}\{\eta\in\Omega(\mathbf{u}^{\prime})\}}{|\Omega(\mathbf{u}^{\prime})|}\left(\mathbb{P}_{\mathbf{u}^{\prime}}\left(\mathbf{U}_{t}=\mathbf{u}^{\prime}\right)-\mu(\Omega(\mathbf{u}^{\prime}))\right).

Taking absolute values, applying the triangular inequality, summing over η\eta, and changing the order of summation shows

σ(σt)μTV𝐮((Ut(1),Vt(1),,Ut(m),Vt(m)))νTV.\|\mathbb{P}_{\sigma}({\sigma_{t}}\in\cdot)-\mu\|_{TV}\leq\|\mathbb{P}_{\mathbf{u}}(({U^{(1)}_{t}},{V^{(1)}_{t}},\dots,{U^{(m)}_{t}},{V^{(m)}_{t}})\in\cdot)-\nu\|_{TV}.

The reverse inequality holds since the 2m2m-coordinate chain is a function of the original chain (σt)({\sigma_{t}}). ∎

Remark.

This lemma lets us look at the 2m2m-coordinate chain instead of the original chain when considering the total variation distance.

Fix a good configuration σ~Ω~\tilde{\sigma}\in\tilde{\Omega}. Recall τmag\tau_{mag} defined in \threfmagcoupling. We use the following coupling after τmag\tau_{mag}, which is a generalization of Lemma 3.5 of [10].

Lemma 4.5 (Post magnetization coupling).
\thlabel

postmagcoupling Let σ~Ω~\tilde{\sigma}\in\tilde{\Omega} be a good configuration. Suppose that two configurations σ0,σ0\sigma_{0},\sigma_{0}^{\prime} satisfy S(i)(σ0)=S(i)(σ0)S^{(i)}(\sigma_{0})=S^{(i)}(\sigma_{0}^{\prime}) for i=1,,mi=1,\dots,m. With respect to the good configuration σ~\tilde{\sigma}, define

Θi\colonequals{σΩ:min{Ui(σ),u~iUi(σ),Vi(σ),v~iVi(σ)}|Ji|16},Θ\colonequalsi=1mΘi\displaystyle\Theta_{i}\colonequals\Big{\{}\sigma\in\Omega:\min\{U_{i}(\sigma),\tilde{u}_{i}-U_{i}(\sigma),V_{i}(\sigma),\tilde{v}_{i}-V_{i}(\sigma)\}\geq\frac{|J_{i}|}{16}\Big{\}},\enspace\Theta\colonequals\bigcap_{i=1}^{m}\Theta_{i}

for each ii. Then there exists a coupling (σt,σt)({\sigma_{t}},{\sigma^{\prime}_{t}}) of the Glauber dynamics with starting states (σ0,σ0)(\sigma_{0},\sigma_{0}^{\prime}) satisfying:

(i)𝐒t=𝐒tfor allt0\displaystyle\mathrm{(i)}\;\mathbf{S}_{t}=\mathbf{S}_{t}^{\prime}\ \text{for all}\ t\geq 0
(ii)If Rt(i)\colonequalsUt(i)Ut(i), then 𝔼σ0,σ0(Rt+1(i)Rt(i)|σt,σt)=Rt(i)n,\displaystyle\mathrm{(ii)}\;\text{If ${R^{(i)}_{t}}\colonequals{U^{\prime(i)}_{t}}-{U^{(i)}_{t}}$, then }\mathbb{E}_{\sigma_{0},\sigma_{0}^{\prime}}\left({R^{(i)}_{t+1}}-{R^{(i)}_{t}}|{\sigma_{t}},{\sigma^{\prime}_{t}}\right)=\frac{-{R^{(i)}_{t}}}{n},
i=1,,m\displaystyle\qquad i=1,\dots,m
(iii)There exists c>0 not depending on n such that on the event{σt,σtΘ},\displaystyle\mathrm{(iii)}\;\text{There exists $c>0$ not depending on $n$ such that on the event}\ \{{\sigma_{t}},{\sigma^{\prime}_{t}}\in\Theta\},
σ0,σ0(Rt+1(i)Rt(i)0|σt,σt)c>0for alli=1,,m.\displaystyle\qquad\mathbb{P}_{\sigma_{0},\sigma_{0}^{\prime}}\left({R^{(i)}_{t+1}}-{R^{(i)}_{t}}\neq 0|{\sigma_{t}},{\sigma^{\prime}_{t}}\right)\geq c>0\;\text{for all}\ i=1,\dots,m.
Proof.

We inductively define the coupling. The random spin SS determined by the randomness II and UU is

S=i=1m(𝟙IJi,Ur+(jiSt(j))𝟙IJi,U>r+(jiSt(j))).\displaystyle S=\sum_{i=1}^{m}(\mathbbm{1}_{I\in J_{i},\;U\leq r_{+}(\sum_{j\neq i}{S^{(j)}_{t}})}-\mathbbm{1}_{I\in J_{i},\;U>r_{+}(\sum_{j\neq i}{S^{(j)}_{t}})}).

Suppose that (σt,σt)({\sigma_{t}},{\sigma^{\prime}_{t}}) is given such that the statements hold for some t0t\geq 0. Let σt+1\sigma_{t+1} be determined II and UU. If IJiI\in J_{i} for some ii, then choose II^{\prime} randomly from {vJi:σt(v)=σt(I)}\{v\in J_{i}^{\prime}:{\sigma^{\prime}_{t}}(v)={\sigma_{t}}(I)\}. Update the primed chain by

σt+1(v)={σt(v)if vISif v=I.\sigma_{t+1}^{\prime}(v)=\begin{cases}{\sigma^{\prime}_{t}}(v)&\text{if $v\neq I^{\prime}$}\\ S&\text{if $v=I^{\prime}$}\end{cases}\quad.

By the induction hypothesis 𝐒t=𝐒t\mathbf{S}_{t}=\mathbf{S}_{t}^{\prime}, we have {vJi:σt(v)=σt(I)}\{v\in J_{i}^{\prime}:{\sigma^{\prime}_{t}}(v)={\sigma_{t}}(I)\}\neq\emptyset and (σt)({\sigma^{\prime}_{t}}) satisfies the Glauber dynamics. Also, 𝐒t+1=𝐒t+1\mathbf{S}_{t+1}=\mathbf{S}_{t+1}^{\prime} with this coupling.

For i=1,,mi=1,\dots,m, put

Ai(σ)\colonequals{vJi:σ(v)=σ~(v)=1},\displaystyle A_{i}(\sigma)\colonequals\{v\in J_{i}:\sigma(v)=\tilde{\sigma}(v)=1\},
Bi(σ)\colonequals{vJi:σ(v)=1,σ~(v)=1},\displaystyle B_{i}(\sigma)\colonequals\{v\in J_{i}:\sigma(v)=-1,\ \tilde{\sigma}(v)=1\},
Ci(σ)\colonequals{vJi:σ(v)=1,σ~(v)=1},\displaystyle C_{i}(\sigma)\colonequals\{v\in J_{i}:\sigma(v)=1,\ \tilde{\sigma}(v)=-1\},
Di(σ)\colonequals{vJi:σ(v)=σ~(v)=1},\displaystyle D_{i}(\sigma)\colonequals\{v\in J_{i}:\sigma(v)=\tilde{\sigma}(v)=-1\},

so |Ai(σ)|=Ui(σ)|A_{i}(\sigma)|=U_{i}(\sigma), |Bi(σ)|=u~iUi(σ)|B_{i}(\sigma)|=\tilde{u}_{i}-U_{i}(\sigma), |Ci(σ)|=v~iVi(σ)|C_{i}(\sigma)|=\tilde{v}_{i}-V_{i}(\sigma), and |Di(σ)|=Vi(σ)|D_{i}(\sigma)|=V_{i}(\sigma).

Now we calculate Rt+1(1)Rt(i){R^{(1)}_{t+1}}-{R^{(i)}_{t}} with the above coupling. The following table shows the one-step dynamics of Rt(1){R^{(1)}_{t}}.

II II^{\prime} SS Rt+1(1)Rt(1){R^{(1)}_{t+1}}-{R^{(1)}_{t}}
B1(σt)B_{1}({\sigma_{t}}) D1(σt)D_{1}({\sigma^{\prime}_{t}}) 1 -1
C1(σt)C_{1}({\sigma_{t}}) A1(σt)A_{1}({\sigma^{\prime}_{t}}) -1 -1
A1(σt)A_{1}({\sigma_{t}}) C1(σt)C_{1}({\sigma^{\prime}_{t}}) -1 1
D1(σt)D_{1}({\sigma_{t}}) B1(σt)B_{1}({\sigma^{\prime}_{t}}) 1 1
otherwise otherwise otherwise 0

Since St(1)=St(1){S^{(1)}_{t}}={S^{\prime(1)}_{t}} implies Rt(1)Ut(1)Ut(1)=Vt(1)Vt(1){R^{(1)}_{t}}\equiv{U^{\prime(1)}_{t}}-{U^{(1)}_{t}}={V^{\prime(1)}_{t}}-{V^{(1)}_{t}},

σ0,σ0(Rt+1(1)Rt(1)=1|σt,σt)\equalscolona(Ut(1),Vt(1),U2,t,V2,t)\displaystyle\mathbb{P}_{\sigma_{0},\sigma_{0}^{\prime}}({R^{(1)}_{t+1}}-{R^{(1)}_{t}}=-1|{\sigma_{t}},{\sigma^{\prime}_{t}})\equalscolon a({U^{(1)}_{t}},{V^{(1)}_{t}},U_{2,t},V_{2,t})
=u~1Ut(1)nVt(1)u~1Ut(1)+Vt(1)r+(j1St(j))+v~1Vt(1)nUt(1)v~1Vt(1)+Ut(1)r(j1St(j))\displaystyle=\frac{\tilde{u}_{1}-{U^{(1)}_{t}}}{n}\frac{{V^{\prime(1)}_{t}}}{\tilde{u}_{1}-{U^{\prime(1)}_{t}}+{V^{\prime(1)}_{t}}}r_{+}(\sum_{j\neq 1}{S^{(j)}_{t}})+\frac{\tilde{v}_{1}-{V^{(1)}_{t}}}{n}\frac{{U^{\prime(1)}_{t}}}{\tilde{v}_{1}-{V^{\prime(1)}_{t}}+{U^{\prime(1)}_{t}}}r_{-}(\sum_{j\neq 1}{S^{(j)}_{t}})
=u~1Ut(1)nVt(1)+Rt(1)u~1Ut(1)+Vt(1)r+(j1St(j))+v~1Vt(1)nUt(1)+Rt(1)v~1Vt(1)+Ut(1)r(j1St(j)).\displaystyle=\frac{\tilde{u}_{1}-{U^{(1)}_{t}}}{n}\frac{{V^{(1)}_{t}}+{R^{(1)}_{t}}}{\tilde{u}_{1}-{U^{(1)}_{t}}+{V^{(1)}_{t}}}r_{+}(\sum_{j\neq 1}{S^{(j)}_{t}})+\frac{\tilde{v}_{1}-{V^{(1)}_{t}}}{n}\frac{{U^{(1)}_{t}}+{R^{(1)}_{t}}}{\tilde{v}_{1}-{V^{(1)}_{t}}+{U^{(1)}_{t}}}r_{-}(\sum_{j\neq 1}{S^{(j)}_{t}}).

Likewise,

σ0,σ0(Rt+1(1)Rt(1)=1|σt,σt)\equalscolonb(Ut(1),Vt(1),U2,t,V2,t)\displaystyle\mathbb{P}_{\sigma_{0},\sigma_{0}^{\prime}}({R^{(1)}_{t+1}}-{R^{(1)}_{t}}=1|{\sigma_{t}},{\sigma^{\prime}_{t}})\equalscolon b({U^{(1)}_{t}},{V^{(1)}_{t}},U_{2,t},V_{2,t})
=Ut(1)nv~1Vt(1)Ut(1)+v~1Vt(1)r(j1St(j))+Vt(1)nu~1Ut(1)u~1Ut(1)+Vt(1)r+(j1St(j))\displaystyle=\frac{{U^{(1)}_{t}}}{n}\frac{\tilde{v}_{1}-{V^{\prime(1)}_{t}}}{{U^{\prime(1)}_{t}}+\tilde{v}_{1}-{V^{\prime(1)}_{t}}}r_{-}(\sum_{j\neq 1}{S^{(j)}_{t}})+\frac{{V^{(1)}_{t}}}{n}\frac{\tilde{u}_{1}-{U^{\prime(1)}_{t}}}{\tilde{u}_{1}-{U^{\prime(1)}_{t}}+{V^{\prime(1)}_{t}}}r_{+}(\sum_{j\neq 1}{S^{(j)}_{t}})
=Ut(1)nv~1(Vt(1)+Rt(1))Ut(1)+v~1Vt(1)r(j1St(j))+Vt(1)nu~1(Ut(1)+Rt(1))u~1Ut(1)+Vt(1)r+(j1St(j)).\displaystyle=\frac{{U^{(1)}_{t}}}{n}\frac{\tilde{v}_{1}-({V^{(1)}_{t}}+{R^{(1)}_{t}})}{{U^{(1)}_{t}}+\tilde{v}_{1}-{V^{(1)}_{t}}}r_{-}(\sum_{j\neq 1}{S^{(j)}_{t}})+\frac{{V^{(1)}_{t}}}{n}\frac{\tilde{u}_{1}-({U^{(1)}_{t}}+{R^{(1)}_{t}})}{\tilde{u}_{1}-{U^{(1)}_{t}}+{V^{(1)}_{t}}}r_{+}(\sum_{j\neq 1}{S^{(j)}_{t}}).

Thus, by a direct calculation,

𝔼σ0,σ0(Rt+1(1)Rt(1)|σt,σt)=ba\displaystyle\mathbb{E}_{\sigma_{0},\sigma_{0}^{\prime}}({R^{(1)}_{t+1}}-{R^{(1)}_{t}}|{\sigma_{t}},{\sigma^{\prime}_{t}})=b-a
=Rt(1)n(r+(j1St(j))+r(j1St(j)))=Rt(1)n.\displaystyle=\frac{-{R^{(1)}_{t}}}{n}\biggl{(}r_{+}(\sum_{j\neq 1}{S^{(j)}_{t}})+r_{-}(\sum_{j\neq 1}{S^{(j)}_{t}})\biggr{)}=\frac{-{R^{(1)}_{t}}}{n}.

Moreover, on the event {σt,σtΘ}\{{\sigma_{t}},{\sigma^{\prime}_{t}}\in\Theta\}, (u~1,v~1,,u~m,v~m)Λ~(\tilde{u}_{1},\tilde{v}_{1},\dots,\tilde{u}_{m},\tilde{v}_{m})\in\tilde{\Lambda} implies Ut(1)u~1|J1|/163|J1|/4|J1|/16=11|J1|/16{U^{(1)}_{t}}\leq\tilde{u}_{1}-|J_{1}|/16\leq 3|J_{1}|/4-|J_{1}|/16=11|J_{1}|/16, and u~1Ut(1)3|J1|/4|J1|/16=11|J1|/16\tilde{u}_{1}-{U^{(1)}_{t}}\leq 3|J_{1}|/4-|J_{1}|/16=11|J_{1}|/16. The same upper bound holds for v~1Vt(1)\tilde{v}_{1}-{V^{(1)}_{t}} and Vt(1){V^{(1)}_{t}}. Thus, on the event {σt,σtΘ}\{{\sigma_{t}},{\sigma^{\prime}_{t}}\in\Theta\},

σ0,σ0(Rt+1(1)Rt(1)0|σt,σt)\displaystyle\mathbb{P}_{\sigma_{0},\sigma_{0}^{\prime}}({R^{(1)}_{t+1}}-{R^{(1)}_{t}}\neq 0|{\sigma_{t}},{\sigma^{\prime}_{t}}) bp116116r(j1St(j))1116+1116+p116116r+(j1St(j))1116+1116\displaystyle\geq b\geq\frac{p_{1}}{16}\frac{\frac{1}{16}r_{-}(\sum_{j\neq 1}{S^{(j)}_{t}})}{\frac{11}{16}+\frac{11}{16}}+\frac{p_{1}}{16}\frac{\frac{1}{16}r_{+}(\sum_{j\neq 1}{S^{(j)}_{t}})}{\frac{11}{16}+\frac{11}{16}}
=p1352.\displaystyle=\frac{p_{1}}{352}.

Similarly, for i>1i>1, σ0,σ0(Rt+1(i)Rt(i)0|σt,σt)pi/352p1/352>0\mathbb{P}_{\sigma_{0},\sigma_{0}^{\prime}}({R^{(i)}_{t+1}}-{R^{(i)}_{t}}\neq 0|{\sigma_{t}},{\sigma^{\prime}_{t}})\geq{p_{i}}/{352}\geq{p_{1}}/{352}>0, which concludes the induction. ∎

5. Upper and Lower Bounds in the high temperature regime

5.1. Upper Bound

Theorem 5.1.
\thlabel

upperbound For β<βcr\beta<\beta_{cr}, we have

limγlim supndn(tn+γn)=0.\lim_{\gamma\to\infty}\limsup_{n\to\infty}d_{n}(t_{n}+\gamma n)=0.
Proof.

Let ν\nu be the stationary measure for the 2m2m-coordinate chain. For any A𝒰A\subseteq\mathcal{U},

|𝐮(𝐔tA)ν(A)|\displaystyle|\mathbb{P}_{\mathbf{u}}(\mathbf{U}_{t}\in A)-\nu(A)| =|𝐮𝒰ν(𝐮)(𝐮(𝐔tA)𝐮(𝐔tA))|\displaystyle=\Big{|}\sum_{\mathbf{u}^{\prime}\in\mathcal{U}}\nu(\mathbf{u}^{\prime})\left(\mathbb{P}_{\mathbf{u}}(\mathbf{U}_{t}\in A)-\mathbb{P}_{\mathbf{u}^{\prime}}(\mathbf{U}_{t}^{\prime}\in A)\right)\Big{|}
𝐮𝒰ν(𝐮)𝐮(𝐔t)𝐮(𝐔t)TV\displaystyle\leq\sum_{\mathbf{u}^{\prime}\in\mathcal{U}}\nu(\mathbf{u}^{\prime})\|\mathbb{P}_{\mathbf{u}}(\mathbf{U}_{t}\in\cdot)-\mathbb{P}_{\mathbf{u}^{\prime}}(\mathbf{U}_{t}^{\prime}\in\cdot)\|_{TV}
max𝐮𝒰𝐮(𝐔t)𝐮(𝐔t)TV.\displaystyle\leq\max_{\mathbf{u}^{\prime}\in\mathcal{U}}\|\mathbb{P}_{\mathbf{u}}(\mathbf{U}_{t}\in\cdot)-\mathbb{P}_{\mathbf{u}^{\prime}}(\mathbf{U}_{t}^{\prime}\in\cdot)\|_{TV}.

Thus, taking supremum over A𝒰A\subseteq\mathcal{U} and 𝐮Λ~\mathbf{u}\in\tilde{\Lambda},

max𝐮Λ~𝐮(𝐔t)νTVmax𝐮Λ~,𝐮𝒰𝐮(𝐔t)𝐮(𝐔t)TV.\displaystyle\max_{\mathbf{u}\in\tilde{\Lambda}}\|\mathbb{P}_{\mathbf{u}}\left(\mathbf{U}_{t}\in\cdot\right)-\nu\|_{TV}\leq\max_{\begin{subarray}{c}\mathbf{u}\in\tilde{\Lambda},\\ \mathbf{u}^{\prime}\in\mathcal{U}\end{subarray}}\|\mathbb{P}_{\mathbf{u}}\left(\mathbf{U}_{t}\in\cdot\right)-\mathbb{P}_{\mathbf{u}^{\prime}}\left(\mathbf{U}_{t}^{\prime}\in\cdot\right)\|_{TV}.

Also, from inequality (4) and \threftotalvariationdistance,

dn(δn+t)\displaystyle d_{n}(\delta n+t) maxσΩ~σ(σt)μ+O(1/n)\displaystyle\leq\max_{\sigma\in\tilde{\Omega}}\|\mathbb{P}_{\sigma}({\sigma_{t}}\in\cdot)-\mu\|+O(1/n)
=max𝐮Λ~𝐮(𝐔t)νTV+O(1/n).\displaystyle=\max_{\mathbf{u}\in\tilde{\Lambda}}\|\mathbb{P}_{\mathbf{u}}(\mathbf{U}_{t}\in\cdot)-\nu\|_{TV}+O(1/n).

For 2m2m-coordinate chains 𝐔t\mathbf{U}_{t} and 𝐔t\mathbf{U}_{t}^{\prime} with respect to a fixed σ~Ω~\tilde{\sigma}\in\tilde{\Omega} starting at 𝐮𝒰\mathbf{u}\in\mathcal{U} and 𝐮𝒰\mathbf{u}^{\prime}\in\mathcal{U}, respectively, put

τtot,c\colonequalsmin{t0:𝐔t=𝐔t}.\tau_{tot,c}\colonequals\min\{t\geq 0:\mathbf{U}_{t}=\mathbf{U}_{t}^{\prime}\}.

It is a standard fact [11, Section 5.2] that

𝐮(𝐔t)𝐮(𝐔t)TV𝐮,𝐮(τtot,c>t).\displaystyle\|\mathbb{P}_{\mathbf{u}}(\mathbf{U}_{t}\in\cdot)-\mathbb{P}_{\mathbf{u}^{\prime}}(\mathbf{U}_{t}^{\prime}\in\cdot)\|_{TV}\leq\mathbb{P}_{\mathbf{u},\mathbf{u}^{\prime}}(\tau_{tot,c}>t).

Combining all the above results, it suffices to bound

max𝐮Λ~,𝐮𝒰𝐮,𝐮(τtot,c>t).\max_{\begin{subarray}{c}\mathbf{u}\in\tilde{\Lambda},\\ \mathbf{u}^{\prime}\in\mathcal{U}\end{subarray}}\mathbb{P}_{\mathbf{u},\mathbf{u}^{\prime}}(\tau_{tot,c}>t).

With the above considerations, fix a good starting configuration σ~Ω~\tilde{\sigma}\in\tilde{\Omega} with the associated 2m2m-coordinates 𝐮~=(u~1,v~1,,u~m,v~m)Λ~\tilde{\mathbf{u}}=(\tilde{u}_{1},\tilde{v}_{1},\dots,\tilde{u}_{m},\tilde{v}_{m})\in\tilde{\Lambda} and an arbitrary starting configuration σΩ\sigma^{\prime}\in\Omega. Put

tn(γ)\colonequalstn+γn,HM\colonequals{τmagtn(γ)}.t_{n}(\gamma)\colonequals t_{n}+\gamma n,\enspace H_{M}\colonequals\{\tau_{mag}\leq t_{n}(\gamma)\}.

The first step is the magnetization coupling phase. By \threfmagcoupling, there exists a coupling (σt,σt)({\sigma_{t}},{\sigma^{\prime}_{t}}) for ttn(γ)t\leq t_{n}(\gamma) with starting configurations (σ~,σ)(\tilde{\sigma},\sigma^{\prime}) such that

σ~,σ(HMc)O(1/γ).\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(H_{M}^{c})\leq O(1/\sqrt{\gamma}).

The next step is the 2m2m-coordinate chain coupling phase. For i=1,,mi=1,\dots,m, define

τi,c\colonequalsmin{t0:(Ut(i),Vt(i))=(Ut(i),Vt(i))},\displaystyle\tau_{i,c}\colonequals\min\{t\geq 0:({U^{(i)}_{t}},V^{(i)}_{t})=({U^{\prime(i)}_{t}},V^{\prime(i)}_{t})\},
Θi\colonequals{σΩ:min{Ui(σ),u~iUi(σ),Vi(σ),v~iVi(σ)}|Ji|16},\displaystyle\Theta_{i}\colonequals\Big{\{}\sigma\in\Omega:\min\{U_{i}(\sigma),\tilde{u}_{i}-U_{i}(\sigma),V_{i}(\sigma),\tilde{v}_{i}-V_{i}(\sigma)\}\geq\frac{|J_{i}|}{16}\Big{\}},
Hi(t)\colonequals{σt(i),σt(i)Θi},Hi\colonequalst[tn(γ),tn(2γ)]Hi(t),Htot\colonequalsi=1mHi.\displaystyle H_{i}(t)\colonequals\{{\sigma^{(i)}_{t}},{\sigma^{\prime(i)}_{t}}\in\Theta_{i}\},\quad H_{i}\colonequals\bigcap_{t\in[t_{n}(\gamma),t_{n}(2\gamma)]}H_{i}(t),\quad H_{tot}\colonequals\bigcap_{i=1}^{m}H_{i}.

We have defined the two coordinate chains with respect to σ~\tilde{\sigma}. On the event HMH_{M}, for ttn(γ)t\geq t_{n}(\gamma), we use the coupling in \threfpostmagcoupling, while on the event HMcH_{M}^{c}, we let the chains run independently for ttn(γ)t\geq t_{n}(\gamma) since we do not care about this un-probable event.

Our first claim is that

σ~,σ(Hic)γO(1/n),i=1,,m.\displaystyle\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(H_{i}^{c})\leq\gamma O(1/n),\enspace i=1,\dots,m.

To that end, observe that

{σt(i)Θi}{Ut(i)<|Ji|/16}{u~iUt(i)<|Ji|/16}{Vt(i)<|Ji|/16}{v~iVt(i)<|Ji|/16}.\begin{split}\{{\sigma^{(i)}_{t}}\notin\Theta_{i}\}\subseteq\{{U^{(i)}_{t}}<|J_{i}|/16\}&\cup\{\tilde{u}_{i}-{U^{(i)}_{t}}<|J_{i}|/16\}\\ &\cup\{V^{(i)}_{t}<|J_{i}|/16\}\cup\{\tilde{v}_{i}-V^{(i)}_{t}<|J_{i}|/16\}.\end{split}

Notice u~i|Ji|/4\tilde{u}_{i}\geq|J_{i}|/4 implies

{Ut(i)<|Ji|/16}\displaystyle\{{U^{(i)}_{t}}<|J_{i}|/16\} {u~iUt(i)>3|Ji|/16},\displaystyle\subseteq\{\tilde{u}_{i}-{U^{(i)}_{t}}>3|J_{i}|/16\},
{u~iUt(i)<|Ji|/16}\displaystyle\{\tilde{u}_{i}-{U^{(i)}_{t}}<|J_{i}|/16\} {Ut(i)>3|Ji|/16}.\displaystyle\subseteq\{{U^{(i)}_{t}}>3|J_{i}|/16\}.

Similarly, v~i|Ji|/4\tilde{v}_{i}\geq|J_{i}|/4 implies

{Vt(i)<|Ji|/16}{v~iVt(i)>3|Ji|/16},{v~iVt(i)<|Ji|/16}{Vt(i)>3|Ji|/16}.\begin{split}\{V^{(i)}_{t}<|J_{i}|/16\}&\subseteq\{\tilde{v}_{i}-V^{(i)}_{t}>3|J_{i}|/16\},\\ \{\tilde{v}_{i}-V^{(i)}_{t}<|J_{i}|/16\}&\subseteq\{V^{(i)}_{t}>3|J_{i}|/16\}.\end{split}

Put

A~i\colonequals{kJi:σ~(k)=1},i=1,,m.\tilde{A}_{i}\colonequals\{k\in J_{i}:\tilde{\sigma}(k)=1\},\enspace i=1,\dots,m.

Then, following the notation in \threfexpectedmagbound, |Mt(A~i)|=|Ut(i)(u~iUt(i))||M_{t}(\tilde{A}_{i})|=|{U^{(i)}_{t}}-(\tilde{u}_{i}-{U^{(i)}_{t}})| implies

{Ut(i)<|Ji|/16}{u~iUt(i)<|Ji|/16}{|Mt(A~i)||Ji|/8}.\{{U^{(i)}_{t}}<|J_{i}|/16\}\cup\{\tilde{u}_{i}-{U^{(i)}_{t}}<|J_{i}|/16\}\subseteq\{|M_{t}(\tilde{A}_{i})|\geq|J_{i}|/8\}.

Similarly, |Mt(JiA~i)|=|Vt(i)(v~iVt(i))||M_{t}(J_{i}\setminus\tilde{A}_{i})|=|V^{(i)}_{t}-(\tilde{v}_{i}-V^{(i)}_{t})| implies

{Vt(i)<|Ji|/16}{v~iVt(i)<|Ji|/16}{|Mt(JiA~i)||Ji|/8}.\{V^{(i)}_{t}<|J_{i}|/16\}\cup\{\tilde{v}_{i}-V^{(i)}_{t}<|J_{i}|/16\}\subseteq\{|M_{t}(J_{i}\setminus\tilde{A}_{i})|\geq|J_{i}|/8\}.

Combining all the above results, we obtain

{σt(i)Θi}{|Mt(A~i)||Ji|/8}{|Mt(JiA~i)||Ji|/8}.\{{\sigma^{(i)}_{t}}\notin\Theta_{i}\}\subseteq\{|M_{t}(\tilde{A}_{i})|\geq|J_{i}|/8\}\cup\{|M_{t}(J_{i}\setminus\tilde{A}_{i})|\geq|J_{i}|/8\}.

A parallel argument for the primed chain shows

{σt(i)Θi}{|Mt(A~i)||Ji|/8}{|Mt(JiA~i)||Ji|/8}.\{{\sigma^{\prime(i)}_{t}}\notin\Theta_{i}\}\subseteq\{|M_{t}^{\prime}(\tilde{A}_{i})|\geq|J_{i}|/8\}\cup\{|M_{t}^{\prime}(J_{i}\setminus\tilde{A}_{i})|\geq|J_{i}|/8\}.

In conclusion,

Hi(t)c\displaystyle H_{i}(t)^{c} ={σt(i)Θi}{σt(i)Θi}\displaystyle=\{{\sigma^{(i)}_{t}}\notin\Theta_{i}\}\cup\{{\sigma^{\prime(i)}_{t}}\notin\Theta_{i}\}
{|Mt(A~i)||Ji|/8}{|Mt(JiA~i)||Ji|/8}\displaystyle\subseteq\{|M_{t}(\tilde{A}_{i})|\geq|J_{i}|/8\}\cup\{|M_{t}(J_{i}\setminus\tilde{A}_{i})|\geq|J_{i}|/8\}
{|Mt(A~i)||Ji|/8}{|Mt(JiA~i)||Ji|/8}.\displaystyle\qquad\qquad\qquad\qquad\quad\enspace\;\cup\{|M_{t}^{\prime}(\tilde{A}_{i})|\geq|J_{i}|/8\}\cup\{|M_{t}^{\prime}(J_{i}\setminus\tilde{A}_{i})|\geq|J_{i}|/8\}.

Define

B\colonequalst[tn(γ),tn(2γ)]{|Mt(A~i)||Ji|/8},Y\colonequalst[tn(γ),tn(2γ)]𝟙{|Mt(A~i)||Ji|/16}.B\colonequals\bigcup_{t\in[t_{n}(\gamma),t_{n}(2\gamma)]}\{|M_{t}(\tilde{A}_{i})|\geq|J_{i}|/8\},\quad Y\colonequals\sum_{t\in[t_{n}(\gamma),t_{n}(2\gamma)]}\mathbbm{1}_{\{|M_{t}(\tilde{A}_{i})|\geq|J_{i}|/16\}}.

Since Mt(A~i)M_{t}(\tilde{A}_{i}) has increments in {1,0,1}\{-1,0,1\}, we have B{Y|Ji|/16}B\subseteq\{Y\geq|J_{i}|/16\}. By Chebyshev’s inequality, σ~,σ(B)c𝔼σ~,σ(Y)/n\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(B)\leq c\mathbb{E}_{\tilde{\sigma},\sigma^{\prime}}(Y)/n for some constant c>0c>0. From \threfexpectedmagbound, for ttnt\geq t_{n}, σ~,σ(|Mt(A~i)||Ji|/16)=O(1/n)\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(|M_{t}(\tilde{A}_{i})|\geq|J_{i}|/16)=O(1/n), so 𝔼σ~,σ(Y)=γO(1)\mathbb{E}_{\tilde{\sigma},\sigma^{\prime}}(Y)=\gamma O(1). Thus, σ~,σ(B)=γO(1/n)\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(B)=\gamma O(1/n). Similar results hold for t[tn(γ),tn(2γ)]{|Mt(JiA~i)||Ji|/8}\bigcup_{t\in[t_{n}(\gamma),t_{n}(2\gamma)]}\{|M_{t}(J_{i}\setminus\tilde{A}_{i})|\geq|J_{i}|/8\}, t[tn(γ),tn(2γ)]{|Mt(A~i)||Ji|/8}\bigcup_{t\in[t_{n}(\gamma),t_{n}(2\gamma)]}\{|M_{t}^{\prime}(\tilde{A}_{i})|\geq|J_{i}|/8\}, and t[tn(γ),tn(2γ)]{|Mt(JiA~i)||Ji|/8}\bigcup_{t\in[t_{n}(\gamma),t_{n}(2\gamma)]}\{|M_{t}^{\prime}(J_{i}\setminus\tilde{A}_{i})|\geq|J_{i}|/8\}. In conclusion,

σ~,σ(Hic)=σ~,σ(t[tn(γ),tn(2γ)]Hi(t)c)4γO(1/n),\displaystyle\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(H_{i}^{c})=\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}\Biggl{(}\bigcup_{t\in[t_{n}(\gamma),t_{n}(2\gamma)]}H_{i}(t)^{c}\Biggr{)}\leq 4\gamma O(1/n),

which proves our first claim.

From the first claim,

σ~,σ(Htotc)i=1mσ,σ(Hic)=γO(1/n).\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(H_{tot}^{c})\leq\sum_{i=1}^{m}\mathbb{P}_{\sigma,\sigma^{\prime}}(H_{i}^{c})=\gamma O(1/n).

Now, condition on the event HMH_{M}. Recalling the fact that \threfpostmagcoupling assures 𝐒t=𝐒t\mathbf{S}_{t}=\mathbf{S}_{t}^{\prime} for ttn(γ)t\geq t_{n}(\gamma) on the event HMH_{M}, we can make Rt(i){R^{(i)}_{t}} stay zero after τi,c\tau_{i,c}, using the modified monotone update on JiJ_{i} whenever a site in JiJ_{i} is chosen to be updated. Thus, on HMH_{M},

τtot,c=max1imτi,c.\tau_{tot,c}=\max_{1\leq i\leq m}\tau_{i,c}.

Our second claim is that

σ~,σ(τi,c>tn(2γ),Hi,HM)=O(1/γ),i=1,,m.\displaystyle\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(\tau_{i,c}>t_{n}(2\gamma),H_{i},H_{M})=O(1/\sqrt{\gamma}),\enspace i=1,\dots,m.

From \threfsupermartingale and \threfpostmagcoupling, σ~,σ(τi,c>tn(2γ),Hi,HM|σtn(γ),σtn(γ))c|Rtn(γ)(i)|/nγ\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(\tau_{i,c}>t_{n}(2\gamma),H_{i},H_{M}|\sigma_{t_{n}(\gamma)},\sigma_{t_{n}(\gamma)}^{\prime})\leq{c|R^{(i)}_{t_{n}(\gamma)}|}/{\sqrt{n\gamma}} for some c>0c>0. Taking expectation yields,

σ~,σ(τi,c>tn(2γ),Hi,HM)c𝔼σ~,σ|Rtn(γ)(i)|nγ\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(\tau_{i,c}>t_{n}(2\gamma),H_{i},H_{M})\leq\frac{c\mathbb{E}_{\tilde{\sigma},\sigma^{\prime}}|R^{(i)}_{t_{n}(\gamma)}|}{\sqrt{n\gamma}}

However, for any t>0t>0, |Rt(i)|=|UtUt|=|Mt(A~i)Mt(A~i)||{R^{(i)}_{t}}|=|U_{t}^{\prime}-U_{t}|=|M_{t}^{\prime}(\tilde{A}_{i})-M_{t}(\tilde{A}_{i})|, so from \threfexpectedmagbound, 𝔼σ~,σ|Rtn(γ)(i)|𝔼σ|Mtn(γ)(A~i)|+𝔼σ~|Mtn(γ)(A~i)|=O(n)\mathbb{E}_{\tilde{\sigma},\sigma^{\prime}}|R^{(i)}_{t_{n}(\gamma)}|\leq\mathbb{E}_{\sigma^{\prime}}|M_{t_{n}(\gamma)}^{\prime}(\tilde{A}_{i})|+\mathbb{E}_{\tilde{\sigma}}|M_{t_{n}(\gamma)}(\tilde{A}_{i})|=O(\sqrt{n}), which proves our second claim.

From the second claim,

σ~,σ(τtot,c>tn(2γ),Htot,HM)i=1mσ~,σ(τi,c>tn(2γ),Htot,HM)\displaystyle\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(\tau_{tot,c}>t_{n}(2\gamma),H_{tot},H_{M})\leq\sum_{i=1}^{m}\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(\tau_{i,c}>t_{n}(2\gamma),H_{tot},H_{M})
i=1mσ~,σ(τi,c>tn(2γ),Hi,HM)=O(1/γ).\displaystyle\leq\sum_{i=1}^{m}\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(\tau_{i,c}>t_{n}(2\gamma),H_{i},H_{M})=O(1/\sqrt{\gamma}).

Combining all the above results,

σ~,σ(τtot,c>tn(2γ))\displaystyle\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(\tau_{tot,c}>t_{n}(2\gamma))
σ~,σ(τtot,c>tn(2γ),Htot,HM)+σ~,σ(Htotc)+σ~,σ(HMc)\displaystyle\leq\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(\tau_{tot,c}>t_{n}(2\gamma),H_{tot},H_{M})+\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(H_{tot}^{c})+\mathbb{P}_{\tilde{\sigma},\sigma^{\prime}}(H_{M}^{c})
=O(1/γ)+γO(1/n)+O(1/γ).\displaystyle=O(1/\sqrt{\gamma})+\gamma O(1/n)+O(1/\sqrt{\gamma}).

Finally,

dn(tn+(2γ+δ)n)O(1/γ)+γO(1/n)+O(1/n),\displaystyle d_{n}(t_{n}+(2\gamma+\delta)n)\leq O(1/\sqrt{\gamma})+\gamma O(1/n)+O(1/n),

which gives us the result upon taking limits. ∎

5.2. Lower Bound

We first analyze the drift of magnetization chains. Let 1im1\leq i\leq m and t\mathcal{F}_{t} be the σ\sigma-algebra generated by St(1),,St(m){S^{(1)}_{t}},\dots,{S^{(m)}_{t}}. By a direct calculation,

𝔼[St+1(i)St(i)|t]\displaystyle\mathbb{E}[{S^{(i)}_{t+1}}-{S^{(i)}_{t}}|\mathcal{F}_{t}] =2npi|Ji|nSt(i)2|Ji|r+(jiSt(j))2npi|Ji|+nSt(i)2|Ji|r(jiSt(j))\displaystyle=\frac{2}{n}p_{i}\frac{|J_{i}|-n{S^{(i)}_{t}}}{2|J_{i}|}r_{+}(\sum_{j\neq i}{S^{(j)}_{t}})-\frac{2}{n}p_{i}\frac{|J_{i}|+n{S^{(i)}_{t}}}{2|J_{i}|}r_{-}(\sum_{j\neq i}{S^{(j)}_{t}})
=2npiSt(i)2r+(jiSt(j))2npi+St(i)2r(jiSt(j))\displaystyle=\frac{2}{n}\frac{p_{i}-{S^{(i)}_{t}}}{2}r_{+}(\sum_{j\neq i}{S^{(j)}_{t}})-\frac{2}{n}\frac{p_{i}+{S^{(i)}_{t}}}{2}r_{-}(\sum_{j\neq i}{S^{(j)}_{t}})
=1n(St(i)+pitanh(βjiSt(j))).\displaystyle=\frac{1}{n}\biggl{(}-{S^{(i)}_{t}}+p_{i}\tanh({\beta}\sum_{j\neq i}{S^{(j)}_{t}})\biggr{)}. (5)

The following simple lemma is the main tool to get the lower bound in \threflowerbound.

Lemma 5.2 (Proposition 7.9, [11]).
\thlabel

statistics Let f:𝒮f\colon\mathcal{S}\to\mathbb{R} be a measurable function and ν1\nu_{1}, ν2\nu_{2} be two probability measures on 𝒮\mathcal{S}. Let σ2\colonequalsmax{𝕍arν1f,𝕍arν2f}\sigma_{*}^{2}\colonequals\max\{\mathbb{V}\mathrm{ar}_{\nu_{1}}f,\ \mathbb{V}\mathrm{ar}_{\nu_{2}}f\}. If |𝔼ν1f𝔼ν2f|rσ|\mathbb{E}_{\nu_{1}}f-\mathbb{E}_{\nu_{2}}f|\geq r\sigma_{*}, then

ν1ν2TV18r2\|\nu_{1}-\nu_{2}\|_{TV}\geq 1-\frac{8}{r^{2}}

Positive starting configurations give us the following result.

Lemma 5.3.
\thlabel

finallemma Let 𝐬𝟎\mathbf{s}\geq\mathbf{0} be the starting magentization. Then, for t0t\geq 0,

𝔼𝐬𝐒t1gt(i=1m(s(i))2pi)1/2+O(1/n).\mathbb{E}_{\mathbf{s}}\|\mathbf{S}_{t}\|_{1}\leq g^{t}\left(\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}\right)^{1/2}+O(1/\sqrt{n}).
Proof.

Consider the case that |Ji||J_{i}| is odd for each i=1,,mi=1,\dots,m. Let ν\nu be the starting distribution such that 𝐬+=(1n,,1n)\mathbf{s}_{+}^{\prime}=(\frac{1}{n},\dots,\frac{1}{n}) with probability 12\frac{1}{2} and 𝐬=(1n,,1n)\mathbf{s}_{-}^{\prime}=(-\frac{1}{n},\dots,-\frac{1}{n}) with probability 12\frac{1}{2}.

By \threfgeneralcontraction, since 𝐬𝐬+\mathbf{s}\geq\mathbf{s}_{+}^{\prime} in this case,

𝟎\displaystyle\mathbf{0} 𝔼𝐬,ν(𝐒t𝐒t)12𝐀t(𝐬𝐬+)+12𝐀t(𝐬𝐬)=𝐀t𝐬.\displaystyle\leq\mathbb{E}_{\mathbf{s},\nu}(\mathbf{S}_{t}-\mathbf{S}_{t}^{\prime})\leq\frac{1}{2}\mathbf{A}^{t}(\mathbf{s}-\mathbf{s}_{+}^{\prime})+\frac{1}{2}\mathbf{A}^{t}(\mathbf{s}-\mathbf{s}_{-}^{\prime})=\mathbf{A}^{t}\mathbf{s}.

However, 𝔼νSt(i)=0\mathbb{E}_{\nu}{S^{\prime(i)}_{t}}=0 for i=1,,mi=1,\dots,m by the remark after \threfmagmarkov. Thus, 𝟎𝔼𝐬𝐒t𝐀t𝐬\mathbf{0}\leq\mathbb{E}_{\mathbf{s}}\mathbf{S}_{t}\leq\mathbf{A}^{t}\mathbf{s}, so by \threfnewlemma,

0i=1m𝔼𝐬St(i)𝐀t𝐬1gt(i=1m(s(i))2pi)1/2.\displaystyle 0\leq\sum_{i=1}^{m}\mathbb{E}_{\mathbf{s}}{S^{(i)}_{t}}\leq\|\mathbf{A}^{t}\mathbf{s}\|_{1}\leq g^{t}\left(\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}\right)^{1/2}.

From \threfmagvariancebound and Cauchy-Schwartz inequality, since 0𝔼𝐬St(i)0\leq\mathbb{E}_{\mathbf{s}}{S^{(i)}_{t}} for i=1,,mi=1,\dots,m,

𝔼𝐬𝐒t1=i=1m𝔼𝐬|St(i)|i=1m(|𝔼𝐬St(i)|+𝕍ar𝐬St(i))=i=1m𝔼𝐬St(i)+i=1m𝕍ar𝐬St(i)\displaystyle\mathbb{E}_{\mathbf{s}}\|\mathbf{S}_{t}\|_{1}=\sum_{i=1}^{m}\mathbb{E}_{\mathbf{s}}|{S^{(i)}_{t}}|\leq\sum_{i=1}^{m}\left(|\mathbb{E}_{\mathbf{s}}{S^{(i)}_{t}}|+\sqrt{\mathbb{V}\mathrm{ar}_{\mathbf{s}}{S^{(i)}_{t}}}\right)=\sum_{i=1}^{m}\mathbb{E}_{\mathbf{s}}{S^{(i)}_{t}}+\sum_{i=1}^{m}\sqrt{\mathbb{V}\mathrm{ar}_{\mathbf{s}}{S^{(i)}_{t}}}
gt(i=1m(s(i))2pi)1/2+(mi=1m𝕍arSt(i))1/2=gt(i=1m(s(i))2pi)1/2+O(1n).\displaystyle\leq g^{t}\Biggl{(}\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}\Biggr{)}^{1/2}+\Biggl{(}m\sum_{i=1}^{m}\mathbb{V}\mathrm{ar}{S^{(i)}_{t}}\Biggr{)}^{1/2}=g^{t}\Biggl{(}\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}\Biggr{)}^{1/2}+O(\frac{1}{\sqrt{n}}).

Other cases of |Ji||J_{i}| can similarly be shown by considering 0 instead of 1n\frac{1}{n} whenever the partition has even number of sites. ∎

Finally, we prove the lower bound.

Theorem 5.4.
\thlabel

lowerbound For β<βcr\beta<\beta_{cr}, we have

limγlim infndn(tnγn)=1.\lim_{\gamma\to\infty}\liminf_{n\to\infty}d_{n}(t_{n}-\gamma n)=1.
Proof.

Since the magnetization chain is a projection of the original chain, it suffices to provide a lower bound on the total variation norm of the magnetization chain. Using tanhxxx2/3\tanh x\geq x-x^{2}/3 for xx\in\mathbb{R}, from equations (5), we have

𝔼(St+1(i)|t)\displaystyle\mathbb{E}({S^{(i)}_{t+1}}|\mathcal{F}_{t}) (11n)St(i)+pin(βjiSt(j)β2(jiSt(j))23)\displaystyle\geq(1-\frac{1}{n}){S^{(i)}_{t}}+\frac{p_{i}}{n}\Biggl{(}\beta\sum_{j\neq i}{S^{(j)}_{t}}-\frac{\beta^{2}(\sum_{j\neq i}{S^{(j)}_{t}})^{2}}{3}\Biggr{)}

for each i=1,,mi=1,\dots,m. In the matrix form,

𝔼(𝐒t+1|t)\displaystyle\mathbb{E}(\mathbf{S}_{t+1}|\mathcal{F}_{t}) 𝐀𝐒t𝐱\displaystyle\geq\mathbf{A}\mathbf{S}_{t}-\mathbf{x}

where 𝐱=β23n(p1(j1St(j))2,,pm(jmSt(j))2)T\mathbf{x}=\frac{\beta^{2}}{3n}(p_{1}(\sum_{j\neq 1}{S^{(j)}_{t}})^{2},\dots,p_{m}(\sum_{j\neq m}{S^{(j)}_{t}})^{2})^{T}. Recall the definition of 𝐚T\colonequals(a1,,am)>𝟎\mathbf{a}^{T}\colonequals(a_{1},\dots,a_{m})>\mathbf{0} with 𝐚1=1\|\mathbf{a}\|_{1}=1 being the left eigenvector of 𝐀\mathbf{A} with eigenvalue gg. Then 𝔼(𝐚T𝐒t+1|t)𝐚T𝐀𝐒t𝐚T𝐱=g𝐚T𝐒t𝐚T𝐱\mathbb{E}(\mathbf{a}^{T}\mathbf{S}_{t+1}|\mathcal{F}_{t})\geq\mathbf{a}^{T}\mathbf{A}\mathbf{S}_{t}-\mathbf{a}^{T}\mathbf{x}=g\mathbf{a}^{T}\mathbf{S}_{t}-\mathbf{a}^{T}\mathbf{x}, i.e.,

𝔼(i=1maiSt+1(i)|t)gi=1maiSt(i)β23ni=1maipi(jiSt(j))2.\mathbb{E}\Bigl{(}\sum_{i=1}^{m}a_{i}{S^{(i)}_{t+1}}|\mathcal{F}_{t}\Bigr{)}\geq g\sum_{i=1}^{m}a_{i}{S^{(i)}_{t}}-\frac{\beta^{2}}{3n}\sum_{i=1}^{m}a_{i}p_{i}\biggl{(}\sum_{j\neq i}{S^{(j)}_{t}}\biggr{)}^{2}. (6)

Observe that

i=1maipi(jiSt(j))2k=1makpk(j=1m|St(j)|)2=(k=1makpk)𝐒t12.\displaystyle\sum_{i=1}^{m}a_{i}p_{i}\biggl{(}\sum_{j\neq i}{S^{(j)}_{t}}\biggr{)}^{2}\leq\sum_{k=1}^{m}a_{k}p_{k}\bigg{(}\sum_{j=1}^{m}|S^{(j)}_{t}|\bigg{)}^{2}=\biggl{(}\sum_{k=1}^{m}a_{k}p_{k}\biggr{)}\|\mathbf{S}_{t}\|_{1}^{2}.

Thus, upon taking expectation in equation (6),

𝔼(i=1maiSt+1(i))g𝔼(i=1maiSt(i))β23n(i=1maipi)𝔼𝐒t12.\displaystyle\mathbb{E}\left(\sum_{i=1}^{m}a_{i}{S^{(i)}_{t+1}}\right)\geq g\mathbb{E}\left(\sum_{i=1}^{m}a_{i}{S^{(i)}_{t}}\right)-\frac{\beta^{2}}{3n}\left(\sum_{i=1}^{m}a_{i}p_{i}\right)\mathbb{E}\|\mathbf{S}_{t}\|_{1}^{2}.

We claim that,

𝔼𝐒t12(𝔼𝐒t1)2+O(1/n).\displaystyle\mathbb{E}\|\mathbf{S}_{t}\|_{1}^{2}\leq(\mathbb{E}\|\mathbf{S}_{t}\|_{1})^{2}+O(1/n).

Since 𝔼𝐒t12=(𝔼𝐒t1)2+𝕍ar𝐒t1\mathbb{E}\|\mathbf{S}_{t}\|_{1}^{2}=(\mathbb{E}\|\mathbf{S}_{t}\|_{1})^{2}+\mathbb{V}\mathrm{ar}\|\mathbf{S}_{t}\|_{1}, it suffices to show 𝕍ar𝐒t1O(1/n)\mathbb{V}\mathrm{ar}\|\mathbf{S}_{t}\|_{1}\leq O(1/n). However, from \threfmagvariancebound,

𝕍ar𝐒t1\displaystyle\mathbb{V}\mathrm{ar}\|\mathbf{S}_{t}\|_{1} =i=1m𝕍ar|St(i)|+2i>jCov(|St(i)|,|St(j)|)\displaystyle=\sum_{i=1}^{m}\mathbb{V}\mathrm{ar}|{S^{(i)}_{t}}|+2\sum_{i>j}\mathrm{Cov}(|{S^{(i)}_{t}}|,|{S^{(j)}_{t}}|)
i=1m𝕍arSt(i)+2i>j𝕍arSt(i)𝕍arSt(j)\displaystyle\leq\sum_{i=1}^{m}\mathbb{V}\mathrm{ar}{S^{(i)}_{t}}+2\sum_{i>j}\sqrt{\mathbb{V}\mathrm{ar}{S^{(i)}_{t}}}\sqrt{\mathbb{V}\mathrm{ar}{S^{(j)}_{t}}}
i=1m𝕍arSt(i)+i>j(𝕍arSt(i)+𝕍arSt(j))=mi=1m𝕍arSt(i)=O(1/n),\displaystyle\leq\sum_{i=1}^{m}\mathbb{V}\mathrm{ar}{S^{(i)}_{t}}+\sum_{i>j}(\mathbb{V}\mathrm{ar}{S^{(i)}_{t}}+\mathbb{V}\mathrm{ar}{S^{(j)}_{t}})=m\sum_{i=1}^{m}\mathbb{V}\mathrm{ar}{S^{(i)}_{t}}=O(1/n),

which proves the claim.

Put Zt\colonequalsi=1maiSt(i)/gtZ_{t}\colonequals\sum_{i=1}^{m}a_{i}{S^{(i)}_{t}}/g^{t}. Then, from the claim above,

𝔼Zt+1𝔼Ztβ2iaipi3ngt+1((𝔼𝐒t1)2+O(1/n)).\displaystyle\mathbb{E}Z_{t+1}-\mathbb{E}Z_{t}\geq-\frac{\beta^{2}\sum_{i}a_{i}p_{i}}{3ng^{t+1}}\left((\mathbb{E}\|\mathbf{S}_{t}\|_{1})^{2}+O(1/n)\right).

Assume that 𝐬𝟎\mathbf{s}\geq\mathbf{0} is a non-negative starting magnetization. Recalling the definition υ\colonequalsn(1g)\upsilon\colonequals n(1-g), from \threffinallemma and the fact i(s(i))2/pi1\sum_{i}{(s^{(i)})^{2}}/{p_{i}}\leq 1,

𝔼𝐬Zt+1𝔼𝐬Zt\displaystyle\mathbb{E}_{\mathbf{s}}Z_{t+1}-\mathbb{E}_{\mathbf{s}}Z_{t} β2iaipi3ngt+1((gt(i(s(i))2/pi)1/2+O(1/n))2+O(1/n))\displaystyle\geq-\frac{\beta^{2}\sum_{i}a_{i}p_{i}}{3ng^{t+1}}\Biggl{(}\biggl{(}g^{t}\biggl{(}\sum_{i}{(s^{(i)})^{2}}/{p_{i}}\biggr{)}^{1/2}+O(1/\sqrt{n})\biggr{)}^{2}+O(1/n)\Biggr{)}
β2iaipi3(nυ)(gti(s(i))2/pi+O(1/n)+1gtO(1/n)).\displaystyle\geq-\frac{\beta^{2}\sum_{i}a_{i}p_{i}}{3(n-\upsilon)}\Biggl{(}g^{t}\sum_{i}{(s^{(i)})^{2}}/{p_{i}}+O(1/\sqrt{n})+\frac{1}{g^{t}}O(1/n)\Biggr{)}.

Iterating from 0 to t1t-1,

𝔼𝐬ZtZ0\displaystyle\mathbb{E}_{\mathbf{s}}Z_{t}-Z_{0} β2iaipi3(nυ)(1gtυ/ni=1m(s(i))2pi+tO(1/n)+nυυ(1gt1)O(1/n))\displaystyle\geq-\frac{\beta^{2}\sum_{i}a_{i}p_{i}}{3(n-\upsilon)}\left(\frac{1-g^{t}}{\upsilon/n}\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}+tO(1/\sqrt{n})+\frac{n-\upsilon}{\upsilon}(\frac{1}{g^{t}}-1)O(1/n)\right)
=β2iaipi3υ(1υ/n)(1gt)i=1m(s(i))2piβ2iaipi3(nυ)tO(1/n)\displaystyle=-\frac{\beta^{2}\sum_{i}a_{i}p_{i}}{3\upsilon(1-\upsilon/n)}(1-g^{t})\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}-\frac{\beta^{2}\sum_{i}a_{i}p_{i}}{3(n-\upsilon)}tO(1/\sqrt{n})
β2iaipi3υ(1gt1)O(1/n).\displaystyle\quad\;-\frac{\beta^{2}\sum_{i}a_{i}p_{i}}{3\upsilon}(\frac{1}{g^{t}}-1)O(1/n).

For brevity, let us prefer to use υ\upsilon rather than use βcr\beta_{cr} in view of \threfupsilon. Consider the step t\colonequalstnγn/υ=12υnlnnγnυt_{*}\colonequals t_{n}-\gamma n/\upsilon=\frac{1}{2\upsilon}n\ln n-\frac{\gamma n}{\upsilon}. Observe that 11/xe1/(x1)1-1/x\geq e^{-1/(x-1)} for x>1x>1 implies

gteγnn/(2(nυ)).g^{t_{*}}\geq\frac{e^{\gamma}}{n^{n/(2(n-\upsilon))}}.

Then

𝔼𝐬Zti=1maisi\displaystyle\mathbb{E}_{\mathbf{s}}Z_{t_{*}}-\sum_{i=1}^{m}a_{i}s_{i}\geq β2iaipi3υ(1υ/n)(1eγnn/(2(nυ)))i=1m(s(i))2pi\displaystyle-\frac{\beta^{2}\sum_{i}a_{i}p_{i}}{3\upsilon(1-{\upsilon}/{n})}\left(1-\frac{e^{\gamma}}{n^{{n}/{(2(n-\upsilon))}}}\right)\sum_{i=1}^{m}\frac{(s^{(i)})^{2}}{p_{i}}
β2iaipi3(nυ)(12υnlnnγnυ)O(1/n)\displaystyle-\frac{\beta^{2}\sum_{i}a_{i}p_{i}}{3(n-\upsilon)}\left(\frac{1}{2\upsilon}n\ln n-\frac{\gamma n}{\upsilon}\right)O({1}/{\sqrt{n}})
β2iaipi3υ(nn/(2(nυ))eγ1)O(1/n).\displaystyle-\frac{\beta^{2}\sum_{i}a_{i}p_{i}}{3\upsilon}\left(\frac{n^{n/(2(n-\upsilon))}}{e^{\gamma}}-1\right)O({1}/{n}).

The right-hand side of the above inequality converges to β2iaipii(s(i))2/pi3υ-\frac{\beta^{2}\sum_{i}a_{i}p_{i}\sum_{i}{(s^{(i)})^{2}}/{p_{i}}}{3\upsilon} as nn\to\infty for every γ>0\gamma>0.

We claim that if nn is large enough, then there exists 𝐬>𝟎\mathbf{s}>\mathbf{0} such that

i=1maisiβ2iaipii(s(i))2/pi3υ>0.\sum_{i=1}^{m}a_{i}s_{i}-\frac{\beta^{2}\sum_{i}a_{i}p_{i}\sum_{i}{(s^{(i)})^{2}}/{p_{i}}}{3\upsilon}>0.

Consider 𝐬=ζ𝐩\mathbf{s}=\zeta\mathbf{p} where 0<ζ<10<\zeta<1 is a constant to be determined. We want to find ζ\zeta such that

i=1maipiζβ2iaipii(piζ)2/pi3υ>0,\sum_{i=1}^{m}a_{i}p_{i}\zeta-\frac{\beta^{2}\sum_{i}a_{i}p_{i}\sum_{i}{(p_{i}\zeta)^{2}}/{p_{i}}}{3\upsilon}>0,

which is equivalent to

3υ>β2ζ.3\upsilon>\beta^{2}\zeta.

From \threfupsilon, υ>0\upsilon>0, so 3υβ2𝐩>𝐬>𝟎\frac{3\upsilon}{\beta^{2}}\mathbf{p}>\mathbf{s}>\mathbf{0} assures that the inequality in the claim holds, and such a positive magnetization 𝐬𝒮\mathbf{s}\in\mathcal{S} exists since nn is large and 0β<βcr0\leq\beta<\beta_{cr} (if β=0\beta=0, choose 𝐬=𝐩\mathbf{s}=\mathbf{p}).

By the last claim, for large nn, there exists 𝐬𝒮\mathbf{s}\in\mathcal{S} and ε>0\varepsilon>0 such that

𝔼𝐬(i=1maiSt(i))2εgt2εeγnn/(2(nυ))εeγn.\mathbb{E}_{\mathbf{s}}(\sum_{i=1}^{m}a_{i}S^{(i)}_{t_{*}})\geq 2\varepsilon g^{t_{*}}\geq 2\varepsilon\frac{e^{\gamma}}{n^{n/(2(n-\upsilon))}}\geq\varepsilon\frac{e^{\gamma}}{\sqrt{n}}.
\thref

magvariancebound and the Cauchy-Schwartz inequality imply 𝕍ar(i=1maiSt(i))=O(1n)\mathbb{V}\mathrm{ar}(\sum_{i=1}^{m}a_{i}S^{(i)}_{t_{*}})=O(\frac{1}{n}) as nn\to\infty. Thus, by \threfzerospin and \threfstatistics, for some c>0c>0,

limγlim infndn(tnγnυ)limγ1cε2e2γ=1.\displaystyle\lim_{\gamma\to\infty}\liminf_{n\to\infty}d_{n}(t_{n}-\frac{\gamma n}{\upsilon})\geq\lim_{\gamma\to\infty}1-\frac{c}{\varepsilon^{2}e^{2\gamma}}=1.

6. Exponentially slow mixing in the low temperature regime

Using a standard bottleneck ratio argument, we can show that the mixing time for the Glauber dynamics is exponential in the low temperature regime. The bottleneck ratio is defined as

Φ\colonequalsminA:μ(A)1/2xA,yAμ(x)P(x,y)μ(A)\Phi\colonequals\min_{A:\mu(A)\leq 1/2}\frac{\sum_{x\in A,y\notin A}\mu(x)P(x,y)}{\mu(A)}

where PP is the transition matrix of the Glauber dynamics. The bottleneck ratio gives a lower bound of the mixing time (see [11, Theorem 7.4]):

tmix14Φ.t_{\mathrm{mix}}\geq\frac{1}{4\Phi}.

We need another characterization of the critical temperature βcr\beta_{cr}.

Lemma 6.1.
\thlabel

slowmixinglemma We have that

βcr=iai2pi(iaipi)2iai2pi2\beta_{cr}=\frac{\sum_{i}a_{i}^{2}p_{i}}{(\sum_{i}a_{i}p_{i})^{2}-\sum_{i}a_{i}^{2}p_{i}^{2}}
Proof.

From 𝐚T𝐀=g𝐚T\mathbf{a}^{T}\mathbf{A}=g\mathbf{a}^{T}, equation (3), and \threfupsilon, we have

iaipi=(pk+1βcr)ak\sum_{i}a_{i}p_{i}=\Big{(}p_{k}+\frac{1}{\beta_{cr}}\Big{)}a_{k}

for each k=1,,mk=1,\dots,m. Multiplying akpka_{k}p_{k} to both sides and summing over kk yields the result. ∎

Proof of \threflowtempresult.

It suffices to show that Φc1exp(c2n)\Phi\leq c_{1}\exp(-c_{2}n) for some positive constants c1,c2>0c_{1},c_{2}>0. By symmetry of the Hamiltonian, we have that μ(A)1/2\mu(A)\leq 1/2 where A\colonequals{σ:iS(i)(σ)>0}A\colonequals\{\sigma:\sum_{i}S^{(i)}(\sigma)>0\}. Since the only way to go from AA to AcA^{c} is to go through B\colonequals{σ:|iS(i)(σ)|1/n}B\colonequals\{\sigma:|\sum_{i}S^{(i)}(\sigma)|\leq 1/n\}, it holds that

xA,yAμ(x)P(x,y)μ(B).\sum_{x\in A,y\notin A}\mu(x)P(x,y)\leq\mu(B).

Note that for any σΩ\sigma\in\Omega,

μ(σ)=exp(βn2((iS(i)(σ))2i(S(i)(σ))2))Z(β).\mu(\sigma)=\frac{\exp\bigg{(}\frac{\beta n}{2}\Big{(}\big{(}\sum_{i}S^{(i)}(\sigma)\big{)}^{2}-\sum_{i}\big{(}S^{(i)}(\sigma)\big{)}^{2}\Big{)}\bigg{)}}{Z(\beta)}.

By the Cauchy-Schwartz inequality,

μ(B)(nn/2)exp(βn2(11m)(1n)2)Z(β)(nn/2)/Z(β)\mu(B)\leq\binom{n}{\lceil n/2\rceil}\frac{\exp\Big{(}\frac{\beta n}{2}\big{(}1-\frac{1}{m}\big{)}\big{(}\frac{1}{n}\big{)}^{2}\Big{)}}{Z(\beta)}\lesssim\binom{n}{\lceil n/2\rceil}/Z(\beta)

where \lesssim denotes that the inequality holds for sufficiently large nn up to a constant not depending on nn. Using Stirling’s formula,

Φexp(nln2)Z(β)μ(A).\Phi\lesssim\frac{\exp(n\ln 2)}{Z(\beta)\mu(A)}.

Now, consider the configurations with exactly kinpik_{i}np_{i} many "++" spins in JiJ_{i} where 1/2ki11/2\leq k_{i}\leq 1 for each i=1,,mi=1,\dots,m and there exists at least one ii such that 1/2<ki1/2<k_{i}. These configurations are members of AA and there are at least i=1m(npikinpi)\prod_{i=1}^{m}\binom{np_{i}}{k_{i}np_{i}} many such configurations. Using Stirling’s formula again, we obtain

Z(β)μ(A)(1i=1m(1ki)pi(1ki)kipiki)neβn2((i(2ki1)pi)2i(2ki1)2pi2).Z(\beta)\mu(A)\gtrsim\Bigg{(}\frac{1}{\prod_{i=1}^{m}(1-k_{i})^{p_{i}(1-k_{i})}k_{i}^{p_{i}k_{i}}}\Bigg{)}^{n}e^{\frac{\beta n}{2}\big{(}(\sum_{i}(2k_{i}-1)p_{i})^{2}-\sum_{i}(2k_{i}-1)^{2}p_{i}^{2}\big{)}}.

Define a function ff through the equation

enf(k1,,km)\colonequals(1i=1m(1ki)pi(1ki)kipiki)neβn2((i(2ki1)pi)2i(2ki1)2pi2).e^{nf(k_{1},\dots,k_{m})}\colonequals\Bigg{(}\frac{1}{\prod_{i=1}^{m}(1-k_{i})^{p_{i}(1-k_{i})}k_{i}^{p_{i}k_{i}}}\Bigg{)}^{n}e^{\frac{\beta n}{2}\big{(}(\sum_{i}(2k_{i}-1)p_{i})^{2}-\sum_{i}(2k_{i}-1)^{2}p_{i}^{2}\big{)}}.

Put (k1,,km)=(1/2,,1/2)+γ(v1,,vm)(k_{1},\dots,k_{m})=(1/2,\dots,1/2)+\gamma(v_{1},\dots,v_{m}) where vi0v_{i}\geq 0 for each i=1,,mi=1,\dots,m, γ\gamma\in\mathbb{R}, and ivi20\sum_{i}v_{i}^{2}\neq 0. Fixing viv_{i}’s, we can regard ff as a one-variable function of γ\gamma, say f=f(γ)f=f(\gamma), and this is equivalent to fixing a direction in m\mathbb{R}^{m}. A little calculation shows that

f(γ)\displaystyle f(\gamma) =2βγ2((ivipi)2ivi2pi2)\displaystyle=2\beta\gamma^{2}\bigg{(}\Big{(}\sum_{i}v_{i}p_{i}\Big{)}^{2}-\sum_{i}v_{i}^{2}p_{i}^{2}\bigg{)}
ipi((1/2γvi)ln(1/2γvi)+(1/2+γvi)ln(1/2+γvi))\displaystyle\quad-\sum_{i}p_{i}\big{(}(1/2-\gamma v_{i})\ln(1/2-\gamma v_{i})+(1/2+\gamma v_{i})\ln(1/2+\gamma v_{i})\big{)}
f(γ)\displaystyle f^{\prime}(\gamma) =4βγ((ivipi)2ivi2pi2)ipivi(ln(1/2γvi)+ln(1/2+γvi))\displaystyle=4\beta\gamma\bigg{(}\Big{(}\sum_{i}v_{i}p_{i}\Big{)}^{2}-\sum_{i}v_{i}^{2}p_{i}^{2}\bigg{)}-\sum_{i}p_{i}v_{i}\big{(}-\ln(1/2-\gamma v_{i})+\ln(1/2+\gamma v_{i})\big{)}
f′′(γ)\displaystyle f^{\prime\prime}(\gamma) =4β((ivipi)2ivi2pi2)ipivi2(11/2γvi+11/2+γvi)\displaystyle=4\beta\bigg{(}\Big{(}\sum_{i}v_{i}p_{i}\Big{)}^{2}-\sum_{i}v_{i}^{2}p_{i}^{2}\bigg{)}-\sum_{i}p_{i}v_{i}^{2}\bigg{(}\frac{1}{1/2-\gamma v_{i}}+\frac{1}{1/2+\gamma v_{i}}\bigg{)}

where denotes a differentiation in γ\gamma. Note that f(0)=ln2f(0)=\ln 2 and f(0)=0f^{\prime}(0)=0. Thus, it suffices to show that there is a direction (v1,,vm)(v_{1},\dots,v_{m}) such that f′′(0)>0f^{\prime\prime}(0)>0. \threfslowmixinglemma shows that the direction (v1,,vm)=(a1,,am)(v_{1},\dots,v_{m})=(a_{1},\dots,a_{m}) satisfies f′′(0)>0f^{\prime\prime}(0)>0 whenever β>βcr\beta>\beta_{cr}, which completes the proof. ∎

Remark.

Combined with the non-exponential mixing time of O(nlnn)O(n\ln n) whenever β<βcr\beta<\beta_{cr}, the above proof shows that inf𝐯𝟎,𝐯𝟎ivi2pi(ivipi)2ivi2pi2\inf_{\mathbf{v}\geq\mathbf{0},\mathbf{v}\neq\mathbf{0}}\frac{\sum_{i}v_{i}^{2}p_{i}}{(\sum_{i}v_{i}p_{i})^{2}-\sum_{i}v_{i}^{2}p_{i}^{2}} is achieved with the direction (v1,,vm)=𝐚T(v_{1},\dots,v_{m})=\mathbf{a}^{T}.

Acknowledgments.

The author would like to thank Professor Insuk Seo for introducing the problem and sharing his limitless insight through numerous discussions. The author also acknowledges an anonymous user at math.stackexchange.com 222https://math.stackexchange.com/q/3553425 for the main idea of the proof in \threfnewlemma. Finally, the author acknowledges the anonymous reviewers for their helpful comments and careful reading of the paper.

References

  • [1] David Aldous “Random walks on finite groups and rapidly mixing markov chains” In Séminaire de Probabilités XVII 1981/82 Berlin, Heidelberg: Springer Berlin Heidelberg, 1983, pp. 243–297
  • [2] David Aldous and Persi Diaconis “Shuffling Cards and Stopping Times” In The American Mathematical Monthly 93.5 Mathematical Association of America, 1986, pp. 333–348
  • [3] P. Cuff et al. “Glauber Dynamics for the Mean-Field Potts Model” In Journal of Statistical Physics 149.3, 2012, pp. 432–477
  • [4] Persi Diaconis and Mehrdad Shahshahani “Generating a random permutation with random transpositions” In Zeitschrift für Wahrscheinlichkeitstheorie und Verwandte Gebiete 57.2, 1981, pp. 159–179
  • [5] Jian Ding, Eyal Lubetzky and Yuval Peres “The mixing time evolution of Glauber Dynamics for the mean-field Ising Model” In Communications in Mathematical Physics 289.2, 2009, pp. 725–764
  • [6] Ignacio Gallo, Adriano Barra and Pierluigi Contucci “Parameter Evaluation of a Simple Mean-Field Model of Social Interaction” In Mathematical Models and Methods in Applied Sciences 19, 2008
  • [7] José C. Hernández, Yevgeniy Kovchegov and Peter T. Otto “The aggregate path coupling method for the Potts model on bipartite graph” In Journal of Mathematical Physics 58.2, 2017, pp. 023303 DOI: 10.1063/1.4976502
  • [8] J.M. Kincaid and E.G.D. Cohen “Phase diagrams of liquid helium mixtures and metamagnets: Experiment and mean field theory” In Physics Reports 22.2, 1975, pp. 57–143
  • [9] Holger Knöpfel, Matthias Löwe, Kristina Schubert and Arthur Sinulis “Fluctuation Results for General Block Spin Ising Models” In Journal of Statistical Physics 178.5, 2020, pp. 1175–1200
  • [10] David A. Levin, Malwina J. Luczak and Yuval Peres “Glauber dynamics for the mean-field Ising model: cut-off, critical power law, and metastability” In Probability Theory and Related Fields, 2010, pp. 146–223
  • [11] David A. Levin and Yuval Peres “Markov Chains and Mixing Times: Second Edition” American Mathematical Society, 2017
  • [12] Eyal Lubetzky and Allan Sly “Cutoff for the Ising model on the lattice” In Inventiones mathematicae 191.3 Springer ScienceBusiness Media LLC, 2012, pp. 719–755
  • [13] Eyal Lubetzky and Allan Sly “Cutoff for General Spin Systems with Arbitrary Boundary Conditions” In Communications on Pure and Applied Mathematics 67.6, 2014, pp. 982–1027
  • [14] Eyal Lubetzky and Allan Sly “Information percolation and cutoff for the stochastic Ising model” In Journal of the American Mathematical Society 29.3 American Mathematical Society, 2016, pp. 729–774
  • [15] Eyal Lubetzky and Allan Sly “Universality of cutoff for the Ising model” In Annals of Probability 45.6A The Institute of Mathematical Statistics, 2017, pp. 3664–3696
  • [16] Carl D. Meyer “Matrix Analysis and Applied Linear Algebra” USA: Society for IndustrialApplied Mathematics, 2000