Skip to main content
Log in

Fixation for Two-Dimensional \({\mathcal {U}}\)-Ising and \({\mathcal {U}}\)-Voter Dynamics

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

Given a finite family \({\mathcal {U}}\) of finite subsets of \({\mathbb {Z}}^d\setminus \{0\}\), the \({\mathcal {U}}\)-voter dynamics in the space of configurations \(\{+,-\}^{{\mathbb {Z}}^d}\) is defined as follows: every \(v\in {\mathbb {Z}}^d\) has an independent exponential random clock, and when the clock at v rings, the vertex v chooses \(X\in {\mathcal {U}}\) uniformly at random. If the set \(v+X\) is entirely in state \(+\) (resp. −), then the state of v updates to \(+\) (resp. −), otherwise nothing happens. The critical probability \(p_c^{\mathrm{vot}}({\mathbb {Z}}^d,{\mathcal {U}})\) for this model is the infimum over p such that this system almost surely fixates at \(+\) when the initial states for the vertices are chosen independently to be \(+\) with probability p and to be − with probability \(1-p\). We prove that \(p_c^{\mathrm{vot}}({\mathbb {Z}}^2,{\mathcal {U}})<1\) for a wide class of families \({\mathcal {U}}\). We moreover consider the \({\mathcal {U}}\)-Ising dynamics and show that its corresponding critical probability \(p_c^{\mathrm{Is}}({\mathbb {Z}}^2,{\mathcal {U}})\) is also less than 1, for many families \({\mathcal {U}}\), so that this model exhibits the same phase transition.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4

Similar content being viewed by others

References

  1. Arratia, R.: Site recurrence for annihilating random walks on \({\mathbb{Z}}^{d}\). Ann. Probab. 11(3), 706–713 (1983)

    Article  MathSciNet  Google Scholar 

  2. Balister, P., Bollobás, B., Przykucki, M.J., Smith, P.J.: Subcritical \(\cal{U}\)-bootstrap percolation models have non-trivial phase transitions. Trans. Am. Math. Soc. 368(10), 7385–7411 (2016)

    Article  MathSciNet  Google Scholar 

  3. Blanquicett, D.: Anisotropic bootstrap percolation in three dimensions. Ann. Probab. 48(5), 2591–2614 (2020)

    Article  MathSciNet  Google Scholar 

  4. Bollobás, B., Duminil-Copin, H., Morris, R., Smith, P.: Universality of two-dimensional critical cellular automata. Proc. Lond. Math. Soc. (to appear). arXiv:1406.6680

  5. Bollobás, B., Smith, P.J., Uzzell, A.J.: Monotone cellular automata in a random environment. Combin. Probab. Comput. 24(4), 687–722 (2015)

    Article  MathSciNet  Google Scholar 

  6. Cerf, R., Manzo, F.: The threshold regime of finite volume bootstrap percolation. Stoch. Proc. Appl. 101(1), 69–82 (2002)

    Article  MathSciNet  Google Scholar 

  7. Cox, J.T., Griffeath, D.: Occupation time limit theorems for the voter model. Ann. Probab. 11(4), 876–893 (1983)

    Article  MathSciNet  Google Scholar 

  8. Damron, M., Kogan, H., Newman, C.M., Sidoravicius, V.: Fixation for coarsening dynamics in 2D slabs. Electron. J. Probab. 18(105), 20 (2013)

    MathSciNet  MATH  Google Scholar 

  9. Durrett, R., Liu, X.-F.: The contact process on a finite set. Ann. Probab. 16(3), 1158–1173 (1988)

    Article  MathSciNet  Google Scholar 

  10. van Enter, A.C.D., Fey, A.: Metastability thresholds for anisotropic bootstrap percolation in three dimensions. J. Stat. Phys. 147(1), 97–112 (2012)

    Article  MathSciNet  ADS  Google Scholar 

  11. Fontes, L.R., Schonmann, R.H., Sidoravicius, V.: Stretched exponential fixation in stochastic Ising models at zero temperature. Commun. Math. Phys. 228(3), 495–518 (2002)

    Article  MathSciNet  ADS  Google Scholar 

  12. Hartarsky, I., Marêché, L., Toninelli, C.: Universality for critical KCM: infinite number of stable directions. Prob. Theory Rel. Fields 178(1–2), 289–326 (2020)

    Article  MathSciNet  Google Scholar 

  13. Liggett, T.M.: Stochastic interacting systems: contact, voter and exclusion processes. volume 324 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer, Berlin (1999)

  14. Martinelli, F.: Lectures on Glauber dynamics for discrete spin models. In: Lectures on Probability Theory and Statistics (Saint-Flour, 1997), volume 1717 of Lecture Notes in Math., pp. 93–191. Springer, Berlin (1999)

  15. Morris, R.: Zero-temperature Glauber dynamics on \({\mathbb{Z}}^d\). Prob. Theory Rel. Fields 149(3–4), 417–434 (2011)

    Article  Google Scholar 

  16. Morris, R.: Bootstrap percolation, and other automata. Eur. J. Combin. 66, 250–263 (2017)

    Article  MathSciNet  Google Scholar 

  17. Mountford, T.S.: Critical length for semi-oriented bootstrap percolation. Stoch. Process. Appl. 56(2), 185–205 (1995)

    Article  MathSciNet  Google Scholar 

  18. Nanda, S., Newman, C.M., Stein, D.: Dynamics of Ising spin systems at zero temperature. In: On Dobrushin’s Way. From Probability Theory to Statistical Physics, volume 198 of Am. Math. Soc. Transl. Ser., vol. 2, pp. 183–194 (2000)

Download references

Acknowledgements

The author is very thankful to Rob Morris for his guidance and feedback throughout this project, and Janko Gravner for his invaluable insight and suggestions on the final version of this paper. The author would like to thank the Instituto Nacional de Matemática Pura e Aplicada (IMPA) for the time and space to create, research and write in this strong academic environment, and to the anonymous referees for their careful reading of this manuscript; their detailed comments helped to improve the presentation of this paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Daniel Blanquicett.

Additional information

Communicated by Alessandro Giuliani.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

The author was partially supported by CAPES, Brasil.

Appendix A: On the Unfair Directions

Appendix A: On the Unfair Directions

In this appendix we provide many examples of families with (and without) unfair directions. By using Proposition 3.12, we are able to construct a large class of families having \(e_2\) (wlog) as an unfair direction, as follows.

Example A.1

Fix a two-dimensional family \({\mathcal {U}}'\) satisfying

  1. (I)

    \(\forall X\in {\mathcal {U}}', X\cap {\mathbb {H}}_{-e_2}\ne \emptyset \) and 0 is in the convex hull of X,

and let \({\mathcal {V}},{\mathcal {W}}\) be 1-dimensional families such that

  1. (II)

    \(0<\nu _- \leqslant w_+\) and \(w_-\leqslant \nu _+\),

where \(\nu _\pm \) (resp. \(w_\pm \)) denotes the number of rules of \({\mathcal {V}}\) (resp. \({\mathcal {W}}\)) entirely contained in \({\mathbb {Z}}_\pm \) (\({\mathbb {Z}}_+={\mathbb {N}}\) and \({\mathbb {Z}}_-=-{\mathbb {N}}\)). Then, for every \(i\in {\mathbb {Z}}_+\), the following induced two-dimensional family is voter-eroding:

$$\begin{aligned} {\mathcal {U}}_i({\mathcal {U}}',{\mathcal {V}},{\mathcal {W}}) :={\mathcal {U}}' \cup \bigcup _{R\in {\mathcal {V}}}\{X(i,R)\} \cup \bigcup _{R\in {\mathcal {W}}}\{X(-i,R)\}, \end{aligned}$$

where \(X(\pm i,R)\) denotes the rule \(\{\pm ie_1\}\cup \{re_2:r\in R\}\).

In fact, \({\mathcal {U}}_i({\mathcal {U}}',{\mathcal {V}},{\mathcal {W}})\) satisfies (a), since either \(ie_1\), or \(-ie_1\), is the only vertex in \(l_{e_2}\) and in some rule \(X(\pm ie_1, R)\) at the same time. Condition (b) is also satisfied, since we can take \(g(X)=X\) for \(X\in {\mathcal {U}}'\), and it is easy to see that condition (II) ensures that we can make an one to one assignment to those rules \(X\notin {\mathcal {U}}'\).

One specific example of a family satisfying (I) and (II) is

$$\begin{aligned} {\mathcal {U}}_1(\{\{e_2,-e_2\}\},\ \{\{1\},\{-1\}\},\ \{\{1\},\{-1\}\}) ={\mathcal {N}}_2^2\setminus \{\{e_1,-e_1\}\}. \end{aligned}$$

Moreover, note that Proposition 3.12 also implies that given \(i_1,\dots ,i_k\in {\mathbb {Z}}_+\), the family

$$\begin{aligned} {\mathcal {U}}_{i_1}({\mathcal {U}}'_1,{\mathcal {V}}_1,{\mathcal {W}}_1) \cup \cdots \cup {\mathcal {U}}_{i_k}({\mathcal {U}}'_k,{\mathcal {V}}_k,{\mathcal {W}}_k) \end{aligned}$$

is voter-eroding, as long as \({\mathcal {U}}'_j,{\mathcal {V}}_j,{\mathcal {W}}_j\) satisfy (I) and (II) for each \(j\leqslant k\). For instance, the family \({\mathcal {U}}_2\) considered in Example 1.3 is voter-eroding, since

$$\begin{aligned} {\mathcal {U}}_2={\mathcal {U}}_1(\{(1,1), (-1,4), -e_2\},\{\{-3,-2,-1\}\},\{\{1\}\}) \cup {\mathcal {U}}_2(\emptyset ,\{\{3,5\}\},\{\{-3\}\}). \end{aligned}$$

On the other hand, we are free to construct plenty of examples of different nature which satisfy conditions (a) and (b) of Proposition 3.12, by using the following trivial observation.

Remark A.2

(Adding rules) Infinitely many families with an unfair direction y can be constructed from any family \({\mathcal {U}}\), just by properly adding new rules \(X\subset {\mathbb {H}}_{-y}\). Moreover, if \({\mathcal {U}}\) is critical, then the new families can be chosen critical as well, we just need to add new rules carefully, without modifying (too much) the stable set; an example is going to be given below (see Example A.3).

Fig. 5
figure 5

The 1-dimensional dynamics associated to \(e_2\) for the family \({\mathcal {U}}_{3,8}\), with \(L=100\) and run time 1000. Time is oriented downwards

1.1 No Unfair Directions

For a concrete example of critical families without unfair directions, consider the collection \({\mathcal {U}}_{3,8}\) of all subsets of size 3 of

$$\begin{aligned} \{\pm e_1,\pm e_2,\pm 2e_1,\pm 2e_2\}. \end{aligned}$$

In Example 3.5 we realized that \({\mathcal {S}}({\mathcal {U}}_{3,8}) =\{\pm e_1,\pm e_2\}\). In order to check that no direction in \({\mathcal {S}}({\mathcal {U}}_{3,8})\) is unfair, by symmetry it is enough to consider \(y=e_2\); in the 1-dimensional setting, straightforward calculations show that configurations \(\eta '\) of the form

$$\begin{aligned} \eta '=[-,+,+,+,+,-,+,+,+,+,-,+,+,+,+,-, \cdots , +,+,+,+,-], \end{aligned}$$

which alternate four \(+\)s and one −, do not satisfy (11). However, simulations (see Fig. 5) indicate that the 1-dimensional dynamics look like a diffusive process (configurations like \(\eta '\) are very unlikely), and we could erode the segment \(Y(e_2,L)\) in time \(O(L^{2})\), which would mean that \({\mathcal {U}}_{3,8}\) is voter-eroding. This family could be a good starting point for future research, in order to prove that Conjecture 1.14 holds for symmetric critical families.

Another such family, which is special since its droplets are triangular, is

$$\begin{aligned} {\mathcal {U}}_\rhd =\{\{(-1,1),(-1,-1)\}, \{(0,1),(1,1)\}, \{(0,-1),(1,-1)\}\}, \end{aligned}$$

with \({\mathcal {S}}({\mathcal {U}}_\rhd )=\left\{ -e_1, \frac{1}{\sqrt{2}}(1,1), \frac{1}{\sqrt{2}}(1,-1)\right\} \) (see Fig. 6).

Fig. 6
figure 6

A voter-eroding critical family without unfair directions, its stable set and a failing configuration

This family only has 3 candidates to be y and all of them fail Condition (11). In fact, say we choose \(y=-e_1\) and take \(L=2n+1\) for some fixed n. The configuration \(\eta \) given by

$$\begin{aligned} \eta ^+=\{(0,2k):k=0,\ldots ,n\}, \end{aligned}$$

yields \(\sum _{u\in \eta ^-}r_u(\eta )=n\), while \(\sum _{v\in \eta ^+}r_v(\eta )=2n\). An analogous situation happens for the other 2 candidates. However, \({\mathcal {U}}_{\rhd }\) is voter-eroding (see Proposition A.4).

Now, we illustrate how to construct a voter-eroding family from \({\mathcal {U}}_\rhd \), in the sense of Remark A.2.

Example A.3

The direction \(y=-e_1\) is unfair for the family

$$\begin{aligned} {\mathcal {U}}_\rhd \cup \{\{(-1,2),(-1,-1)\}\}, \end{aligned}$$

this is just because we added a rule in the + side to enforce inequality (15); since it has the same stable set as \({\mathcal {U}}_\rhd \), then it is critical and voter-eroding so our main theorem holds for this new family.

In general, if we consider any rule \(X_0\subset {\mathbb {H}}_{e_1}\), \(X_0\ne \{(-1,1), (-1,-1)\}\) such that there exist \(x,x'\in X_0\) with \(x_2\geqslant -x_1\) and \(x_2'\leqslant x_1'\), we can check that \(-e_1\) is an unfair direction for the family \({\mathcal {U}}_\rhd \cup \{X_0\}\), and the latter has the same stable set as \({\mathcal {U}}_\rhd \). Of course, we can construct infinitely many such families in this way.

As a last example, we will show that \({\mathcal {U}}_{\rhd }\) is \((2+\varepsilon ,{\mathcal {S}}({\mathcal {U}}_\rhd ))\)-eroding for any \(\varepsilon >0\), by using an argument similar to the one given in the proof of Proposition 3.4, and coupling with the 1-dimensional contact process.

Proposition A.4

The family \({\mathcal {U}}_\rhd \) is voter-eroding.

Proof

The droplet erosion time \(\tau \) in the 1-dimensional dynamics associated to \(y=-e_1\) is stochastically dominated by the extinction time \(\tau _{CP}\) of a (slowed down by a 1/3 factor) contact process on \(\{1,\dots , L\}\) with parameter 1. This process is subcritical, so by Theorem 1 in [9], \(\tau _{CP}/\log L\) converges in probability to a positive constant, as \(L\rightarrow \infty \). Now, by following the same lines along the proof of Proposition 3.4, (10) and (9) are straightforward. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Blanquicett, D. Fixation for Two-Dimensional \({\mathcal {U}}\)-Ising and \({\mathcal {U}}\)-Voter Dynamics. J Stat Phys 182, 21 (2021). https://doi.org/10.1007/s10955-020-02697-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10955-020-02697-8

Keywords

Mathematics Subject Classification

Navigation