Cutoff phenomenon of the Glauber dynamics for the Ising model on complete multipartite graphs in the high temperature regime
Abstract.
In this paper, the Glauber dynamics for the Ising model on the complete multipartite graph is investigated where is the proportion of the vertices in the th component. We show that the dynamics exhibits the cutoff phenomena at with window size in the high temperature regime where is a constant only depending on . Exponentially slow mixing is shown in the low temperature regime .
Key words and phrases:
Markov chains, Ising model, Mixing time, Cutoff, Coupling, Glauber dynamics, Heat-bath dynamics, Mean-field model2020 Mathematics Subject Classification:
60J10, 60K35, 82C201. 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 for this longstanding problem. An improvement on the window size to optimal 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 is understood throughout the paper. Some important quantities not depending on will be explicitly mentioned. We will write to be the th vector in the standard basis of . The lower case will always denote time. Let denote the Hadamard product between matrices. More precisely, whenever and are matrices with the same dimensions.
1.2. Ising model and Glauber dynamics
Let be a finite graph with the vertex set and the edge set . Elements of are called configurations. In the absence of external fields, the Ising model on is a distribution called the Gibbs distribution on given by
where , , , and is a normalizing factor. Assuming an isotropic interaction strength between the vertices, we set . The physical interpretation of is the energy of the whole spin system with the configuration . We call each the spin at site .
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 and update the spin at the chosen site according to 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 is irreducible, aperiodic, and reversible with as its unique stationary distribution. For the Ising model, it is easy to see that the probability of updating to at the chosen site is where
(1) |
and is the mean-field at .
1.3. Markov chain mixing and cutoff phenomenon
The total variation distance between two probability measures and on is defined by
The total variation distance is half of the -distance between the probability measures.
Let 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 at time by
where here and thereafter denotes the probability given . The mixing time is defined by
We say a sequence of Markov chains with corresponding mixing times exhibit a cutoff phenomenon if for every ,
Furthermore, we say that the cutoff occurs at with window size if and
1.4. Magnetization chain on complete multipartite graphs
Now, we are in a place to consider a complete -partite graph, a graph whose vertices are partitioned into 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 which has vertices and partitions where and for . We fix the parameters and ’s hereafter. Without loss of generality, we assume . We may also assume that for every so that is well defined whenever such considerations are required. Let be the set of all vertices where denotes the set of the th partition of the vertices. Note .
We define for so that is our configuration space. Each configuration has a unique representation and both representations are understood throughout this paper.
For each , define the magnetization on by , . For the Markov chain starting at , we define the corresponding magnetization on by
We sometimes use the vector notation for . We call the process a magnetization chain. \threfmagmarkov shows that is in fact a Markov chain. Note that it is a projection of the whole Markov chain , so mixing of the whole chain implies the mixing of the chain . 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 and such that , the Glauber dynamics for the Ising model on the complete multipartite graph exhibits a cutoff at with window size in the high temperature regime where is a constant defined in equation (3).
Theorem 1.2.
lowtempresult In the low temperature regime , the dynamics is exponentially slow mixing, i.e., for some constants not depending on .
A few remarks are in order. Our main result is obtained as a consequence of \threfupperbound and \threflowerbound. In the low temperature regime , the mixing time is exponentially slow, therefore identifying the critical temperature . In the case, there are no spin interactions so the chain is equivalent to the lazy random walk on an -dimensional hypercube, which has a cutoff at with window size (see [1] or [11, Chapter 18]). This result can be seen as a consequence of our main result since implies (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 steps with high probability. After the magnetization coupling phase, by considering the "-coordinate chain" inspired by [10], we can construct a post magnetization coupling to reach the full-mixing in another 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 and be independent uniform random variables over and , respectively. We consider the collection of Markov chains with starting configurations . Simultaneously define the next configurations at time by
where is defined in equation (1). Repeat this procedure independently for each time. It is clear that each Markov chain above is a version of the Glauber dynamics on the complete multipartite graph with starting state ’s, defined on a common probability space. The above coupling is called a monotone coupling in the sense that if are starting states for and , respectively, then for so that , and for any accordingly.
Define
Proposition 2.1 (Magnetization chain).
magmarkov
The process is a
Markov chain on the magnetization state space .
Proof.
Note that
is measurable with respect to the -algebra generated by . Other cases can be dealt with similarly. ∎
Remark.
By symmetry, starting from and starting from have the same distributions. This can also be seen by the physical fact that the map 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 and , denote the Hamming distance by .
Remark.
This is a metric on , which is equal to the number of sites with different spins for two configurations. Similarly, we can define on , respectively, but ’s merely satisfy the triangle inequality.
Lemma 2.2 (Contraction in mean for monotone coupling).
monotonecontraction For a monotone coupling starting at , we have
where
with , .
Proof.
Assume with for some vertex . Note . Since we are considering a monotone coupling, it holds that for each ,
where
Note that
Since and are independent, for ,
Suppose . There exists , , , such that . By the triangular inequality for and the fact ,
Furthermore, by the Markov property,
By taking expectation and putting , we have
Iterating gives
∎
From now on, (which depends on the number of vertices ) always denotes the matrix defined in \threfmonotonecontraction. Note that is a positive matrix, so by the Perron-Frobenius theorem, there exists the largest eigenvalue with the left eigenvector normalized in norm. Note that has algebraic multiplicity (see [16, Section 8.2] for a proof), so is unique.
We fix the following notations
(2) | |||
(3) |
where and are defined in the previous paragraph. Another characterization of is given in \threfslowmixinglemma. Insuk Seo commented111personal communication that it can also be characterized as the threshold value of that makes positive definite where is defined through the equation , being the -by- identity matrix. \threfupsilon connects the quantities and .
Proposition 2.3.
upsilon The left eigenvector only depends on . Moreover, only depends on and through the following equation:
Therefore, only depends on , and we have .
Proof.
Since satisfies
it holds that is a root of a polynomial with coefficients only depending on . Since is in the kernel of the transpose of the above matrix, it only depends on .
Finally, implies .
∎
We collect further properties of the matrix and its left eigenvector in the next two lemmas.
Lemma 2.4.
finalfinallemma We have
The equality in the latter holds if and only if .
Proof.
Recall that we assumed .
We claim that . To that end, fix . From , we have . Then , i.e., . Thus, implies , proving the claim.
By Chebyshev’s sum inequality, since and whenever ,
The equality holds if and only if or . The proof is now complete by noticing the fact that and imply . ∎
Remark.
As a consequence, we obtain a lower bound .
Lemma 2.5.
newlemma For and , we have
In particular, it holds that
Proof.
We want to find a symmetric matrix which is similar to . To that end, suppose that there exists an invertible diagonal matrix and a symmetric matrix such that . Then , so , which leads to for . With the above in mind, let and where and for . Note that is real-symmetric and . Then, by the spectral theorem for real symmetric matrices, . Note that and have the same real eigenvalues since they are similar.
Observe that for , . This can be easily checked by the equalities
Let . The case is trivial, so assume . Since has rank 1, has rank . Also, its elements are positive, so it has a positive eigenvalue by the Perron-Frobenius theorem. Thus, is equal to its spectral radius, from which the following inequality follows:
Similarly,
∎
Remark.
Another relatively simple proof of can be given as follows. By the Cauchy-Schwartz inequality, we have . Then .
From now on, for brevity, we use the notation
Lemma 2.6.
lemma For a monotone coupling starting at , we have
Moreover, for ,
Proof.
From \threfmonotonecontraction,
Notice that for each , so \threfnewlemma implies
∎
We would like to translate \threflemma to the case of magnetization chains, which is done in \threfnormcontraction.
Lemma 2.7.
generalcontraction For starting magnetizations , the magnetization chains satisfy
Remark.
We say such pairs of starting magnetizations are monotone pairs.
Proof.
Let be a monotone coupling starting from where and , for . Such a monotone coupling exists because of the given condition for each . Since , we have for each . By monotonicity, for each . Thus, for each . Then, by \threfmonotonecontraction,
Now, we can complete the proof since we have for each by \threfmagmarkov. ∎
Recall that denotes a Hadamard product.
Proposition 2.8.
normcontraction For a monotone coupling starting at with magnetizations , we have
Moreover, not depending on the coupling, we have
Proof.
For any magnetizations and , there exists such that for . In particular, and are a monotone pair for each . Then we can consider a monotone coupling with starting states such that , for , and the magnetization of the starting configuration is for .
Let be the magnetization chain corresponding to for . By telescoping, \threfgeneralcontraction gives
Then, the triangle inequality and \threfmagmarkov imply
∎
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 . Observe that for square-integrable -valued i.i.d. random vectors , we have .
Lemma 3.1.
variancebound Let be a Markov chain in a finite state space . Suppose that there exists such that for any ,
Then, for the norm variance,
Proof.
Put . Let and be independent copies of the chain starting from . The idea is to condition on the first step. Note that for . Then by the observation right before the statement of this lemma,
By the assumption and Markov property, we have
Thus, for ,
By the Markov property, for every , , so
The total variance formula holds since we are using the norm. Thus, taking supremum over in the total variance formula , we have . Upon iterating,
∎
The following proposition is an important result bounding the variance of magnetization chains uniformly in time.
Proposition 3.2.
magvariancebound Let . For an arbitrary starting configuration and , we have
where only depends on , and .
Proof.
Observe that . Note that increments of are bounded by in absolute value. Then, from \threffinalfinallemma, we have
By \threffinalfinallemma, \threfnormcontraction, and \threfvariancebound, we have
Note that \threfupsilon assures . ∎
We also establish a bound for the expected magnetization on subsets of partitions. To that end, we need the following observation.
Lemma 3.3.
zerospin For each , where is the Gibbs distribution. In particular, we have .
Proof.
Since for each configuration and is a bijection from into itself, we have . ∎
Proposition 3.4 (Expected magnetization bound).
expectedmagbound Let and . For any and a chain starting at , define . Then
Furthermore, for , we have
Proof.
Let "+" denote the configuration such that all spins are and "-" denote the configuration with all spins . Let be a monotone coupling with starting configuration where is the stationary distribution. Let . By \threflemma and \threfzerospin,
Then, by symmetry, for , . Thus, by summing over sites in , . However, for any configuration , by monotonicity, . Considering the remark after \threfmagmarkov, . Thus, for any .
Now, by \threfmagvariancebound, , so
Thus, for ,
However, by symmetry, for any fixed ,
Thus,
Likewise, for ,
Similarly, , so from ,
whenever . Thus, for ,
Lastly, for , from Jensen’s inequality,
∎
4. Couplings
Fix the notation
Definition (Modified matching).
Let and have magnetizations and , respectively. Consider two copies of the graph, and . Let . If , then it is possible to match each site in with spin to a site in with spin. Any leftover sites in are arbitrarily matched to the leftover sites in . We match the sites in a similar way whenever . This defines a bijection .
We call this bijection a modified matching of and .
Definition (Modified monotone update and coupling).
Let be a modified matching of . Let and be uniformly distributed over and , respectively, and be independent. Suppose for some is the chosen site in . Consider the case . If
then update the chosen site of by +1 and of by +1. If
then update the chosen site of by -1 and of by -1. Otherwise, if
then update the chosen site of by -1 and of by +1. The other case can similarly be updated.
Given the chosen site , 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 of and . Let and be chains starting at and , respectively. Repeating the above procedure independently for each step with respect to gives a coupling of the Glauber dynamics. We call this coupling a modified monotone coupling with respect to the given modified matching.
Remark.
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 steps in the next two lemmas.
Lemma 4.1 (Lemma 2.4, [10]).
supermartingale
Let be a non-negative supermartingale with a stopping time satisfying
(i)
(ii)
(iii) .
Then for ,
Lemma 4.2 (Magnetization coupling).
magcoupling Let . For any configurations and , there exists a coupling with starting states satisfying the following condition. If , then for large ,
where is a constant not depending on , , or .
Proof.
Let be a monotone coupling with starting states . Put for and . Define
By \threfnormcontraction,
for some .
We construct a coupling such that is a positive supermartingale with bounded increments and the conditional probability of not being lazy is bounded away from zero uniformly in time and .
To that end, consider a time . Define , , and . Note that since . Choose a site equiprobably over . Let be the modified matching of and . If a site in is chosen, then use the modified monotone update with respect to to update . If a site in is chosen, then independently choose another site equiprobably over (which can be the same site) to update independent of . It is easy to check that the above is a coupling of the Glauber dynamics.
Clearly, has bounded increment with the above coupling. Let be a random variable uniformly distributed over which is independent of . Let and . Since implies , we obtain that is bounded below by
Finally, we need to show the supermartingale property. Consider . Suppose . Then by a direct calculation, on the event , it holds that is bounded above by
Suppose . Note that implies and . Let . Then by equation (5) in Section 5.2, on the event , is equal to
Since either or must hold, is equal to
Thus,
Putting in the matrix form with , we have
Since implies by \threfupsilon, the supermartingale property is established.
With the above coupling, by \threfsupermartingale, for large ,
for some not depending on . Taking expectation,
Note has at most more +1 spin sites than , so by \threffinalfinallemma. At , construct a modified matching of and , and use the modified monotone coupling with respect to this modified matching from then on. At , we construct another modified matching of the sites to do a new modified monotone coupling so that forever after .
By \threffinalfinallemma, a modified version of \threfnormcontraction, and the strong Markov property, we have
Thus,
and putting yields
∎
Definition (Good configurations).
Define the set of "good" configurations by
For and each , define
Define
Remark.
Note that . In other words, is another representation of good configurations . We omit the starting state and write instead of for convenience.
Lemma 4.3 (Lemma 3.3, [10]).
goodstartingstatesdistance For any subset and stationary distribution ,
Recall that we are assuming the high temperature regime. By \threfexpectedmagbound, there exists such that . Hence, by \threfmagvariancebound, for large ,
Combining with \threfgoodstartingstatesdistance,
(4) |
Definition (-coordinate chain).
Let be a reference configuration. For and each , define
For a chain with the starting configuration , define the -coordinate chain with respect to by
It is easy to see that the -coordinate chain is again a Markov chain in its state space and determines the magnetization chain through the relation for .
Symmetry gives us the following lemma which is an adaptation of Lemma 3.4 in [10].
Lemma 4.4.
totalvariationdistance Let be a chain starting at . Consider the corresponding -coordinate chain starting at . Then
where is the stationary distribution of the -coordinate chain.
Proof.
Since , given the -coordinate , the conditional -probability of the configurations is equiprobable. In other words, is uniform where is the set of configurations having the -coordinate . Also, by symmetry,
is uniform over . Thus,
Taking absolute values, applying the triangular inequality, summing over , and changing the order of summation shows
The reverse inequality holds since the -coordinate chain is a function of the original chain . ∎
Remark.
This lemma lets us look at the -coordinate chain instead of the original chain when considering the total variation distance.
Fix a good configuration . Recall defined in \threfmagcoupling. We use the following coupling after , which is a generalization of Lemma 3.5 of [10].
Lemma 4.5 (Post magnetization coupling).
postmagcoupling Let be a good configuration. Suppose that two configurations satisfy for . With respect to the good configuration , define
for each . Then there exists a coupling of the Glauber dynamics with starting states satisfying:
Proof.
We inductively define the coupling. The random spin determined by the randomness and is
Suppose that is given such that the statements hold for some . Let be determined and . If for some , then choose randomly from . Update the primed chain by
By the induction hypothesis , we have and satisfies the Glauber dynamics. Also, with this coupling.
For , put
so , , , and .
Now we calculate with the above coupling. The following table shows the one-step dynamics of .
1 | -1 | ||
-1 | -1 | ||
-1 | 1 | ||
1 | 1 | ||
otherwise | otherwise | otherwise | 0 |
Since implies ,
Likewise,
Thus, by a direct calculation,
Moreover, on the event , implies , and . The same upper bound holds for and . Thus, on the event ,
Similarly, for , , which concludes the induction. ∎
5. Upper and Lower Bounds in the high temperature regime
5.1. Upper Bound
Theorem 5.1.
upperbound For , we have
Proof.
Let be the stationary measure for the -coordinate chain. For any ,
Thus, taking supremum over and ,
Also, from inequality (4) and \threftotalvariationdistance,
For -coordinate chains and with respect to a fixed starting at and , respectively, put
It is a standard fact [11, Section 5.2] that
Combining all the above results, it suffices to bound
With the above considerations, fix a good starting configuration with the associated -coordinates and an arbitrary starting configuration . Put
The first step is the magnetization coupling phase. By \threfmagcoupling, there exists a coupling for with starting configurations such that
The next step is the -coordinate chain coupling phase. For , define
We have defined the two coordinate chains with respect to . On the event , for , we use the coupling in \threfpostmagcoupling, while on the event , we let the chains run independently for since we do not care about this un-probable event.
Our first claim is that
To that end, observe that
Notice implies
Similarly, implies
Put
Then, following the notation in \threfexpectedmagbound, implies
Similarly, implies
Combining all the above results, we obtain
A parallel argument for the primed chain shows
In conclusion,
Define
Since has increments in , we have . By Chebyshev’s inequality, for some constant . From \threfexpectedmagbound, for , , so . Thus, . Similar results hold for , , and . In conclusion,
which proves our first claim.
From the first claim,
Now, condition on the event . Recalling the fact that \threfpostmagcoupling assures for on the event , we can make stay zero after , using the modified monotone update on whenever a site in is chosen to be updated. Thus, on ,
Our second claim is that
From \threfsupermartingale and \threfpostmagcoupling, for some . Taking expectation yields,
However, for any , , so from \threfexpectedmagbound, , which proves our second claim.
From the second claim,
Combining all the above results,
Finally,
which gives us the result upon taking limits. ∎
5.2. Lower Bound
We first analyze the drift of magnetization chains. Let and be the -algebra generated by . By a direct calculation,
(5) |
The following simple lemma is the main tool to get the lower bound in \threflowerbound.
Lemma 5.2 (Proposition 7.9, [11]).
statistics Let be a measurable function and , be two probability measures on . Let . If , then
Positive starting configurations give us the following result.
Lemma 5.3.
finallemma Let be the starting magentization. Then, for ,
Proof.
Consider the case that is odd for each . Let be the starting distribution such that with probability and with probability .
By \threfgeneralcontraction, since in this case,
However, for by the remark after \threfmagmarkov. Thus, , so by \threfnewlemma,
From \threfmagvariancebound and Cauchy-Schwartz inequality, since for ,
Other cases of can similarly be shown by considering instead of whenever the partition has even number of sites. ∎
Finally, we prove the lower bound.
Theorem 5.4.
lowerbound For , we have
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 for , from equations (5), we have
for each . In the matrix form,
where . Recall the definition of with being the left eigenvector of with eigenvalue . Then , i.e.,
(6) |
Observe that
Thus, upon taking expectation in equation (6),
We claim that,
Since , it suffices to show . However, from \threfmagvariancebound,
which proves the claim.
Put . Then, from the claim above,
Assume that is a non-negative starting magnetization. Recalling the definition , from \threffinallemma and the fact ,
Iterating from to ,
For brevity, let us prefer to use rather than use in view of \threfupsilon. Consider the step . Observe that for implies
Then
The right-hand side of the above inequality converges to as for every .
We claim that if is large enough, then there exists such that
Consider where is a constant to be determined. We want to find such that
which is equivalent to
From \threfupsilon, , so assures that the inequality in the claim holds, and such a positive magnetization exists since is large and (if , choose ).
By the last claim, for large , there exists and such that
magvariancebound and the Cauchy-Schwartz inequality imply as . Thus, by \threfzerospin and \threfstatistics, for some ,
∎
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
where is the transition matrix of the Glauber dynamics. The bottleneck ratio gives a lower bound of the mixing time (see [11, Theorem 7.4]):
We need another characterization of the critical temperature .
Lemma 6.1.
slowmixinglemma We have that
Proof.
From , equation (3), and \threfupsilon, we have
for each . Multiplying to both sides and summing over yields the result. ∎
Proof of \threflowtempresult.
It suffices to show that for some positive constants . By symmetry of the Hamiltonian, we have that where . Since the only way to go from to is to go through , it holds that
Note that for any ,
By the Cauchy-Schwartz inequality,
where denotes that the inequality holds for sufficiently large up to a constant not depending on . Using Stirling’s formula,
Now, consider the configurations with exactly many "" spins in where for each and there exists at least one such that . These configurations are members of and there are at least many such configurations. Using Stirling’s formula again, we obtain
Define a function through the equation
Put where for each , , and . Fixing ’s, we can regard as a one-variable function of , say , and this is equivalent to fixing a direction in . A little calculation shows that
where ′ denotes a differentiation in . Note that and . Thus, it suffices to show that there is a direction such that . \threfslowmixinglemma shows that the direction satisfies whenever , which completes the proof. ∎
Remark.
Combined with the non-exponential mixing time of whenever , the above proof shows that is achieved with the direction .
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