Skip to main content
Log in

Planelike Interfaces in Long-Range Ising Models and Connections with Nonlocal Minimal Surfaces

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

Abstract

This paper contains three types of results:

  • the construction of ground state solutions for a long-range Ising model whose interfaces stay at a bounded distance from any given hyperplane,

  • the construction of nonlocal minimal surfaces which stay at a bounded distance from any given hyperplane,

  • the reciprocal approximation of ground states for long-range Ising models and nonlocal minimal surfaces.

In particular, we establish the existence of ground state solutions for long-range Ising models with planelike interfaces, which possess scale invariant properties with respect to the periodicity size of the environment. The range of interaction of the Hamiltonian is not necessarily assumed to be finite and also polynomial tails are taken into account (i.e. particles can interact even if they are very far apart the one from the other). In addition, we provide a rigorous bridge between the theory of long-range Ising models and that of nonlocal minimal surfaces, via some precise limit result.

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.

Similar content being viewed by others

Notes

  1. Unfortunately, there seems to be no common agreement in scientific papers about a coherent use of the terminology “phase transition” and “phase coexistence”. The present paper is, in a sense, related to the study of phase coexistence phenomena, in which the ambient spaces is separated, roughly speaking, into two regions with different phases. In most of the mathematical literature, however, the description of similar models based on partial differential equations (such as the Allen–Cahn or Ginzburg–Landau equations) is referred to with the (perhaps rather inaccurate) name of phase transition.

  2. As a historical remark, we also observe that the idea that simple discrete models at the atomic scale could lead to qualitative macroscopic modifications may go back, in its embryonic stages, at least to Democritus, who is alleged to claim that “by convention sweet is sweet, bitter is bitter, hot is hot, cold is cold, color is color; but in truth there are only atoms and the void”, see [31].

  3. In our setting, the terminology “ground state” is used to denote minimizers of the Ising energy functional with respect to compact perturbations. From the point of view of statistical physics, we recall that the support of the equilibrium measures at zero temperature does not necessarily coincide with the ground state configurations, and this is indeed a rather delicate issue, for which we refer the reader to [29] and Appendix B in [59].

  4. We think that the restriction of the exponent of the power to be between d and \(d+1\) is not merely technical but reflects the different behavior that the problem presents at large scale: as a matter of fact, above a certain threshold, the interface of the problem “localizes” when seen from infinity, while below such threshold the interface keeps its nonlocal behavior at any scale. In this sense, the different treatment that we propose for different exponents is not a mere complication, but it reflects a real characteristic of the model considered. See [54] for an analogy with partial differential equations.

  5. Here we adopt a partially misleading terminology, as the boundary \(\partial E\), and not the set E, should be regarded as the minimal surface, in conformity with the classical geometrical notion of perimeter. However, we have \({{\mathrm{Per}}}_K(E; \Omega ) = {{\mathrm{Per}}}_K(\mathbb {R}^d \setminus E; \Omega )\), for any set E, and thus no confusion should arise from this slightly improper notation.

  6. Throughout the whole paper, \(Q_R\) denotes the closed cube of \(\mathbb {R}^d\) having sides of length 2R and centered at the origin, i.e.

    $$\begin{aligned} Q_R := \Big \{ x \in \mathbb {R}^d : | x |_\infty \le R \Big \}. \end{aligned}$$

    We use the same notation for cubes in \(\mathbb {Z}^d\). That is, for \(\ell \in \mathbb {N}\cup \{ 0 \}\), we write

    $$\begin{aligned} Q_\ell := \Big \{ i \in \mathbb {Z}^d : | i |_\infty \le \ell \Big \} = \big \{ -\ell , \ldots , -1, 0, 1, \ldots , \ell \big \}^d. \end{aligned}$$

    Cubes not centered at the origin are indicated with \(Q_R(x) := x + Q_R\) and \(Q_\ell (q) := q + Q_\ell \), with \(x \in \mathbb {R}^d\) and \(q \in \mathbb {Z}^d\).

  7. As usual, we will denote by \(\lceil x \rceil \) the smallest integer greater than or equal to x, and by \(\lfloor x\rfloor \) the largest integer less than or equal to x.

  8. Actually, the Lipschitz regularity assumption on the boundary of \(\Omega \) can be omitted for the deduction of the \(\Gamma \)-\(\liminf \) inequality.

  9. Note that \(\widetilde{H}_{\mathcal {F}_{m, \omega }^{A, B}}\) differs from \(H_{\mathcal {F}_{m, \omega }}\) only with respect to the region over which the magnetic term B is extended. We take into account this slight modification, since \(B_{\mathcal {F}_{m, \omega }}\) might not be well-defined even under assumption (1.8), as the set \(\mathcal {F}_{m, \omega }\) is not finite.

    The main reason to consider the auxiliary functional \(\widetilde{H}_{\mathcal {F}_{m, \omega }^{A, B}}\) is due to the presence of the magnetic term, which is not null in general but it has zero average on a particular domain. To take advantage of this feature, one can either select an appropriate order of summation, or perform a truncation argument. We chose to follow this latter strategy.

  10. In this regard, observe that the family of configurations appearing in the definition of v is actually finite, thanks to the periodicity of \(u_{m, \omega }^{A, B}\).

  11. As usual, \(\mathring{Q}\) denotes the interior of Q.

  12. To be extremely precise, Lemma 5.3 gives a sequence of sets \(\{ \widetilde{E}_k \}_{k \in \mathbb {N}}\) with smooth boundaries such that

    $$\begin{aligned} \left| [( \widetilde{E}_k \cap \Omega ) \cup (E \setminus \Omega ) ] \Delta E \right| \rightarrow 0 \quad \text{ and } \quad {{\mathrm{Per}}}_K \left( ( \widetilde{E}_k \cap \Omega ) \cup (E \setminus \Omega ); \Omega \right) \rightarrow {{\mathrm{Per}}}_K(E; \Omega ), \quad \text{ as } k \rightarrow +\infty . \end{aligned}$$

    Then, it is not hard to check that the sets \(E_k := (\widetilde{E}_k \cap \Omega _{1/k}) \cup (E \setminus \Omega _{1/k})\) fulfill (8.4) and (8.5).

References

  1. Aubry, S., Le Daeron, P.Y.: The discrete Frenkel–Kontorova model and its extensions. I. Exact results for the ground-states. Phys. D 8(3), 381–422 (1983)

    Article  MathSciNet  MATH  Google Scholar 

  2. Auer, F., Bangert, V.: Minimising currents and the stable norm in codimension one. C. R. Acad. Sci. Paris Sér. I Math. 333(12), 1095–1100 (2001)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Bai, F., Branch, R.W., Nicolau, D.V., Pilizota, T., Steel, B.C., Maini, P.K., Berry, R.M.: Conformational spread as a mechanism for cooperativity in the bacterial flagellar switch. Science 327(5966), 685–689 (2010)

    Article  ADS  Google Scholar 

  4. Bessi, U.: Slope-changing solutions of elliptic problems on \({\mathbb{R}}^n\). Nonlinear Anal. 68(12), 3923–3947 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  5. Birindelli, I., Valdinoci, E.: The Ginzburg–Landau equation in the Heisenberg group. Commun. Contemp. Math. 10(5), 671–719 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  6. Blanchard, T., Picco, M., Rajapbour, M.A.: Influence of long-range interactions on the critical behavior of the Ising model. Europhys. Lett. 101(5), 56003 (2013)

    Article  ADS  Google Scholar 

  7. Bourgain, J., Brezis, H., Mironescu, P.: Limiting embedding theorems for \(W^{s, p}\) when \(s\uparrow 1\) and applications. J. Anal. Math. 87, 77–101 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  8. Braides, A.: An example of non-existence of plane-like minimizers for an almost-periodic Ising system. J. Stat. Phys. 157(2), 295–302 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Bucur, C., Valdinoci, E.: Nonlocal Diffusion and Applications. Lecture Notes of the Unione Matematica Italiana. Lecture Notes of the Unione Matematica Italiana, Zurich (2016)

    Book  MATH  Google Scholar 

  10. Caffarelli, L.: Surfaces minimizing nonlocal energies. Atti. Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur. Rend. Lincei Mat. Appl. 20(3), 281–299 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  11. Caffarelli, L., de la Llave, R.: Planelike minimizers in periodic media. Commun. Pure Appl. Math. 54(12), 1403–1441 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  12. Caffarelli, L., de la Llave, R.: Interfaces of ground states in Ising models with periodic coefficients. J. Stat. Phys. 118(3-4), 687–719 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. Caffarelli, L., Roquejoffre, J.-M., Savin, O.: Nonlocal minimal surfaces. Commun. Pure Appl. Math. 63(9), 1111–1144 (2010)

    MathSciNet  MATH  Google Scholar 

  14. Caffarelli, L., Valdinoci, E.: Regularity properties of nonlocal minimal surfaces via limiting arguments. Adv. Math. 248, 843–871 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  15. Campa, A., Dauxois, T., Ruffo, S.: Statistical mechanics and dynamics of solvable models with long-range interactions. Phys. Rep. 480, 57–159 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  16. Cassandro, M., Orlandi, E., Picco, P.: The optimal interface profile for a non-local model of phase separation. Nonlinearity 15(5), 1621–1651 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Chierchia, L., Falcolini, C.: A note on quasi-periodic solutions of some elliptic systems. Z. Angew. Math. Phys. 47(2), 210–220 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  18. Chikazumi, S.: Physics of Ferromagnetism. Oxford University Press, New York (1997)

    Google Scholar 

  19. Cinti, E., Serra, J., Valdinoci, E.: Quantitative flatness results and \(BV\)-estimates for stable nonlocal minimal surfaces, arXiv preprint, arXiv:1602.00540 (2016)

  20. Cozzi, M., Valdinoci, E.: Plane-like minimizers for a non-local Ginzburg–Landau–type energy in a periodic medium. J. Éc. Polytech. Math. 4, 337–388 (2017)

  21. Cozzi, M., Valdinoci, E.: Planelike minimizers of nonlocal Ginzburg–Landau energies and fractional perimeters in periodic media. (2017)

  22. Dauxois, T., Ruffo, S., Arimondo, E., Wilkens, M.: Dynamics and Thermodynamics of Systems with Long-Range Interactions. Lecture Notes in Physics. Springer, New York (2002)

    Book  Google Scholar 

  23. Dávila, J.: On an open question about functions of bounded variation. Calc. Var. Partial Differ. Eq. 15(4), 519–527 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  24. de la Llave, R., Valdinoci, E.: Ground states and critical points for generalized Frenkel–Kontorova models in \({\mathbb{Z}}^d\). Nonlinearity 20(10), 2409–2424 (2007)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  25. de la Llave, R., Valdinoci, E.: Ground states and critical points for Aubry–Mather theory in statistical mechanics. J. Nonlinear Sci. 20(2), 153–218 (2010)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  26. Di Castro, A., Novaga, M., Ruffini, B., Valdinoci, E.: Nonlocal quantitative isoperimetric inequalities. Calc. Var. Partial Differ. Eq. 54(3), 2421–2464 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  27. Di Nezza, E., Palatucci, G., Valdinoci, E.: Hitchhiker’s guide to the fractional Sobolev spaces. Bull. Sci. Math. 136(5), 521–573 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  28. Dipierro, S., Valdinoci, E.: Nonlocal minimal surfaces: interior regularity, quantitative estimates and boundary stickiness. In: Kuusi, T., Palatucci, G. (eds.) Recent Developments in the Nonlocal Theory. Book Series on Measure Theory. De Gruyter, Berlin (2017)

    Google Scholar 

  29. Dobrushin, R.L., Shlosman, S.B.: The problem of translation invariance of Gibbs states at low temperatures. Math. Phys. Rev. 5, 53–195 (1985)

    MathSciNet  MATH  Google Scholar 

  30. Duke, T.A.J., Bray, D.: Heightened sensitivity of a lattice of membrane receptors. Proc. Natl. Acad. Sci. 96(18), 10104–10108 (1999)

    Article  ADS  Google Scholar 

  31. Durant, W.: The Story of Civilization: Part II: The Life of Greece. Simon and Schuster, New York (1939)

    Google Scholar 

  32. Figalli, A., Fusco, N., Maggi, F., Millot, V., Morini, M.: Isoperimetry and stability properties of balls with respect to nonlocal energies. Commun. Math. Phys. 336(1), 441–507 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  33. Frank, R., Seiringer, R.: Non-linear ground state representations and sharp Hardy inequalities. J. Funct. Anal. 255(12), 3407–3430 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  34. Gallavotti, G.: Statistical Mechanics. Springer, Berlin (1999)

    Book  MATH  Google Scholar 

  35. Giusti, E.: Minimal Surfaces and Functions of Bounded Variation. Monographs in Mathematics. Birkhäuser, Basel (1984)

    Book  MATH  Google Scholar 

  36. Grisvard, P.: Elliptic Problems in Nonsmooth Domains. Monographs and Studies in Mathematics. Pitman (Advanced Publishing Program), Boston (1985)

    MATH  Google Scholar 

  37. Hedlund, G.A.: Geodesics on a two-dimensional Riemannian manifold with periodic coefficients. Ann. Math. 33(4), 719–739 (1932)

    Article  MathSciNet  MATH  Google Scholar 

  38. Hopfield, J.J.: Neural networks and physical systems with emergent collective computational abilities. Proc. Natl. Acad. Sci. 79(8), 2554–2558 (1982)

    Article  ADS  MathSciNet  Google Scholar 

  39. Ising, E.: Beitrag zur theorie des Ferromagnetismus. Z. Phys. 31(1), 253–258 (1925)

    Article  ADS  Google Scholar 

  40. Lenz, W.: Beiträge zum Verständnis der magnetischen Eigenschaften in festen Körpern. Physik. Zeit. 21, 613–615 (1920)

    Google Scholar 

  41. Lombardini, L.: Approximation of sets of finite fractional perimeter by smooth sets and comparison of local and global \(s\)-minimal surfaces, preprint (2016)

  42. Mather, J.N.: Existence of quasiperiodic orbits for twist homeomorphisms. Topology 21(4), 457–467 (1989)

    Article  MathSciNet  MATH  Google Scholar 

  43. Mather, J.N.: Action minimizing invariant measures for positive definite Lagrangian systems. Math. Z. 207(2), 169–207 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  44. McLean, W.: Strongly Elliptic Systems and Boundary Integral Equations. Cambridge University Press, Cambridge (2000)

    MATH  Google Scholar 

  45. Morse, H.M.: A fundamental class of geodesics on any closed surface of genus greater than one. Trans. Am. Math. Soc. 26(1), 25–60 (1924)

    Article  MathSciNet  MATH  Google Scholar 

  46. Moser, J.: Minimal solutions of variational problems on a torus. Ann. Inst. H. Poincaré Anal. Non Linéaire 3(3), 229–272 (1986)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  47. Newman, M.E.J., Barkema, G.T.: Monte Carlo Methods in Statistical Physics. Oxford University Press, New York (1999)

    MATH  Google Scholar 

  48. Onsager, L.: Crystal statistics. I. A two-dimensional model with an order-disorder transition. Phys. Rev. 2(65), 117–149 (1944)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  49. Peierls, R.: On Ising’s model of ferromagnetism. Math. Proc. Camb. Philosoph. Soc. 32(3), 477–481 (1936)

    Article  ADS  MATH  Google Scholar 

  50. Petrosyan, A., Valdinoci, E.: Density estimates for a degenerate/singular phase-transition model. SIAM J. Math. Anal. 36(4), 1057–1079 (2005)

    Article  MathSciNet  MATH  Google Scholar 

  51. Picco, M.: Critical behavior of the Ising model with long range interactions, arXiv preprint, arXiv:1207.1018v1 (2012)

  52. Rabinowitz, P.H., Stredulinsky, E.: On some results of Moser and of Bangert. Ann. Inst. H. Poincaré Anal. Non Linéaire 21(5), 673–688 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  53. Ruelle, D.: Statistical Mechanics. World Scientific Publishing, River Edge (1999)

    Book  MATH  Google Scholar 

  54. Savin, O., Valdinoci, E.: \(\Gamma \)-convergence for nonlocal phase transitions. Ann. Inst. H. Poincaré Anal. Non Linéaire 29(4), 479–500 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  55. Teif, V.B.: General transfer matrix formalism to calculate DNA-protein-drug binding in gene regulation: application to \(O_R\) operator of phage \(\lambda \). Nucleic Acids Res. 35(11), e80 (2007)

    Article  Google Scholar 

  56. Torres, M.: Plane-like minimal surfaces in periodic media with exclusions. SIAM J. Math. Anal. 36(2), 523–551 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  57. Valdinoci, E.: Plane-like minimizers in periodic media: jet flows and Ginzburg–Landau-type functionals. J. Reine Angew. Math. 574, 147–185 (2004)

    MathSciNet  MATH  Google Scholar 

  58. Visintin, A.: Generalized coarea formula and fractal sets. Jpn. J. Indust. Appl. Math. 8(2), 175–201 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  59. van Enter, A.C.D., Fernàndez, R., Sokal, A.D.: Regularity properties and pathologies of position-space renormalization-group transformations: scope and limitations of Gibbsian theory. J. Stat. Phys. 72(5–6), 879–1167 (1993)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  60. Wang, J.-S., Selke, W., Andreichenko, V.B., Dotsenko, V.S.: The critical behaviour of the two-dimensional dilute model. Phys. A 164(2), 221–239 (1990)

    Article  Google Scholar 

Download references

Acknowledgements

This work has been supported by the Alexander von Humboldt Foundation, the ERC grant 277749 E.P.S.I.L.O.N. “Elliptic Pde’s and Symmetry of Interfaces and Layers for Odd Nonlinearities”, the PRIN grant 201274FYK7 “Aspetti variazionali e perturbativi nei problemi differenziali nonlineari”, the “María de Maeztu” MINECO Grant MDM-2014-0445 and the MINECO Grant MTM2014-52402-C3-1-P.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Serena Dipierro.

Appendices

Appendix 1: Proof of Lemma 5.3

In the present appendix, we provide a proof of Lemma 5.3 in full details. As mentioned right after its statement in Sect. 5, our argument is based on the strategies already followed in e.g. [19, 35, 41].

Throughout the section, we implicitly suppose conditions (1.18) and (1.19) to be in force. Although the result may in fact hold under weaker hypotheses, we always suppose for simplicity that K satisfies both these assumptions. However, we stress that none of the steps of the proof require the periodicity hypothesis (1.20) to be valid, that we therefore do not suppose to hold.

After these introductory remarks, we may now head to the proof of Lemma 5.3.

Proof of Lemma 5.3

First, notice that, by (1.19) and the fact that F has finite K-perimeter, the characteristic function \(\chi _F\) belongs to the fractional Sobolev space \(W^{s, 1}(\Omega )\). Hence, by standard density results (see e.g. [36, Theorem 1.4.2.1]), there exists a sequence \(\{ \varphi _n \}_{n \in \mathbb {N}} \subset W^{s, 1}(\Omega ) \cap C^\infty (\overline{\Omega })\) such that

$$\begin{aligned} {{\varphi _n \rightarrow \chi _F\; \mathrm{in}~W^{s, 1}(\Omega ), \mathrm{as}~n \rightarrow +\infty .}} \end{aligned}$$
(9.1)

By using again (1.19), this ensures that

$$\begin{aligned} \lim _{n \rightarrow +\infty } \mathscr {K}_K(\varphi _n; \Omega , \Omega ) = \mathscr {K}_K(\chi _F; \Omega , \Omega ). \end{aligned}$$
(9.2)

For \(t \in (0, 1)\), we let

$$\begin{aligned} F_n := \left( \{ \varphi _n > t \} \cap \overline{\Omega } \right) \cup (F \setminus \overline{\Omega }). \end{aligned}$$

Clearly, \(F_n \setminus \overline{\Omega } = F \setminus \overline{\Omega }\), which proves (5.9).

Also, Morse-Sard’s Theorem tells that, for a.e. \(t \in (0, 1)\), the boundary of the level set \(\{ \varphi _n > t \}\) is a smooth hypersurface. Hence \(\partial F_n\) is smooth inside \(\overline{\Omega }\), which gives (5.8).

We now claim that for a.e. \(t \in (0, 1)\) fixed,

$$\begin{aligned} \lim _{n \rightarrow +\infty } |F_n \Delta F| = 0, \end{aligned}$$
(9.3)

and

$$\begin{aligned} \lim _{n \rightarrow +\infty } {{\mathrm{Per}}}_K(F_n; \Omega ) = {{\mathrm{Per}}}_K(F; \Omega ), \end{aligned}$$
(9.4)

up to a subsequence, that is (5.10) and (5.11), respectively.

We begin by checking (9.3). Let \(\tau \in (0, 1)\) and notice that

$$\begin{aligned} \varphi _n - \chi _F&> \tau \quad&\text{ in } (\{ \varphi _n> \tau \} \setminus F) \cap \Omega \\ {\text{ and } }\quad \chi _F - \varphi _n&\ge 1 - \tau \quad&\text{ in } (F \setminus \{ \varphi _n > \tau \}) \cap \Omega . \end{aligned}$$

From this, we deduce that

$$\begin{aligned} \Vert \varphi _n - \chi _F \Vert _{L^1(\Omega )}&\ge \int _{(\{ \varphi _n> \tau \} \setminus F) \cap \Omega } (\varphi _n(x) - \chi _F(x)) \, dx + \int _{(F \setminus \{ \varphi _n> \tau \}) \cap \Omega } (\chi _F(x) - \varphi _n(x)) \, dx \\&\ge \tau |(\{ \varphi _n> \tau \} \setminus F) \cap \Omega | + (1 - \tau ) |(F \setminus \{ \varphi _n> \tau \}) \cap \Omega | \\&\ge \min \{ \tau , 1 - \tau \} \, |(\{ \varphi _n > \tau \} \Delta F) \cap \Omega |. \end{aligned}$$

Therefore, using this and (9.1),

$$\begin{aligned} \{ \varphi _n > \tau \} \longrightarrow F \quad \text{ in } L^1(\Omega ), \text{ for } \text{ a.e. } \tau \in (0, 1). \end{aligned}$$
(9.5)

Claim (9.3) follows as a particular case by taking \(\tau = t\) in formula (9.5) above and recalling that \(F_n \setminus \overline{\Omega } = F \setminus \overline{\Omega }\).

Next, we address the convergence of the perimeters stated in (9.4). Thanks to (9.5) and Lemma 4.1, we have

$$\begin{aligned} \mathcal {L}_K(F \cap \Omega , \Omega \setminus F) \le \liminf _{n \rightarrow +\infty } \mathcal {L}_K(\{ \varphi _n> \tau \} \cap \Omega , \Omega \setminus \{ \varphi _n > \tau \}) \quad \text{ for } \text{ a.e. } \tau \in (0, 1), \end{aligned}$$

or, equivalently,

$$\begin{aligned} \mathscr {K}_K(\chi _F; \Omega , \Omega ) \le \liminf _{n \rightarrow +\infty } \mathscr {K}_K(\chi _{\{ \varphi _n > \tau \}}; \Omega , \Omega ) \quad \text{ for } \text{ a.e. } \tau \in (0, 1). \end{aligned}$$
(9.6)

By applying, in sequence, (9.2), the generalized Coarea Formula of Lemma 4.2, Fatou’s Lemma and (9.6), we compute

$$\begin{aligned} \mathscr {K}_K(\chi _F; \Omega , \Omega )= & {} \lim _{n \rightarrow +\infty } \mathscr {K}_K(\varphi _n; \Omega , \Omega ) = \lim _{n \rightarrow +\infty } \int _{-\infty }^{+\infty } \mathscr {K}_K(\chi _{\{\varphi _n> \tau \}}; \Omega , \Omega ) \, d\tau \\\ge & {} \int _0^1 \liminf _{n \rightarrow +\infty } \mathscr {K}_K(\chi _{\{\varphi _n > \tau \}}; \Omega , \Omega ) \, d\tau \\\ge & {} \int _0^1 \mathscr {K}_K(\chi _F; \Omega , \Omega ) \, d\tau = \mathscr {K}_K(\chi _F; \Omega , \Omega ). \end{aligned}$$

By this and, again, (9.6) we conclude that

$$\begin{aligned} \liminf _{n \rightarrow +\infty } \mathscr {K}_K(\chi _{\{\varphi _n > \tau \}}; \Omega , \Omega ) = \mathscr {K}_K(\chi _F; \Omega , \Omega ) \quad \text{ for } \text{ a.e. } \tau \in (0, 1), \end{aligned}$$

and thence

$$\begin{aligned} \lim _{n \rightarrow +\infty } \mathcal {L}_K(F_n \cap \Omega , \Omega \setminus F_n) = \mathcal {L}_K(F \cap \Omega , \Omega \setminus F). \end{aligned}$$
(9.7)

On the other hand, we claim that

$$\begin{aligned}&\lim _{n \rightarrow +\infty } \mathcal {L}_K(F_n \setminus \Omega , \Omega \setminus F_n) = \mathcal {L}_K(F \setminus \Omega , \Omega \setminus F)\nonumber \\ \text{ and } \quad&\lim _{n \rightarrow +\infty } \mathcal {L}_K(F_n \cap \Omega , \mathbb {R}^d \setminus (F_n \cup \Omega )) = \mathcal {L}_K(F \cap \Omega , \mathbb {R}^d \setminus (F \cup \Omega )), \end{aligned}$$
(9.8)

up to subsequences. To check the validity of (9.8), we first notice that, by (9.3), \(\chi _{F_n} \rightarrow \chi _F\) a.e. in \(\mathbb {R}^d\) (up to extracting a subsequence), as \(n\rightarrow +\infty \). Therefore, in view of Lemma 4.3 we may apply the Lebesgue’s Dominated Convergence Theorem to get

$$\begin{aligned} \lim _{n \rightarrow +\infty } \mathcal {L}_K(F_n \setminus \Omega , \Omega \setminus F_n)&= \lim _{n \rightarrow +\infty } \int _{\Omega } \chi _{\mathbb {R}^d \setminus F_n}(x) \left( \int _{\mathbb {R}^d \setminus \Omega } \chi _{F_n}(y) K(x, y) \, dy \right) dx \\&= \int _{\Omega } \chi _{\mathbb {R}^d \setminus F}(x) \left( \int _{\mathbb {R}^d \setminus \Omega } \chi _{F}(y) K(x, y) \, dy \right) dx \\&= \mathcal {L}_K(F \setminus \Omega , \Omega \setminus F), \end{aligned}$$

and similarly for the limit on the second line of (9.8). The combination of (9.7) and (9.8) yields the convergence of the K-perimeters claimed in (9.4).

The proof of Lemma 5.3 is thus finished. \(\square \)

Appendix 2: Optimality of the Width of the Strip Given in (1.17)

The goal of this appendix is to show that, for large values of the periodicity scale \(\tau \), the interface of the planelike ground states for powerlike interactions, as in (1.9), oscillates, in general, by a quantity proportional to \(\tau \) (i.e., the conclusion in (1.17) of Theorem 1.4 cannot be improved).

Of course, one needs to construct an ad-hoc example to check this optimality. The idea to construct this counterexample comes from similar phenomena in minimal surfaces and minimal foliations, in which the oscillation is produced by the fact that the metric is nonflat. For simplicity, we present here a two-dimensional explicit example, which goes as follows.

Given \(\tau \in 4\mathbb {N}+1\) (to be taken large in the subsequent construction), we define

$$\begin{aligned}&Q:=\left\{ -\frac{\tau -1}{2},\dots ,0,\dots ,\frac{\tau -1}{2}\right\} ^2,\\&\widehat{Q}:=\left\{ -\frac{\tau -1}{4},\dots ,0,\dots ,\frac{\tau -1}{4}\right\} ^2 \end{aligned}$$

and

$$\begin{aligned} \widehat{\mathcal {Q}} :=\left\{ (i,j)\in \mathbb {Z}^2\times \mathbb {Z}^2 { \text{ s.t. } \text{ there } \text{ exists } }(i^{\prime },j^{\prime })\in \widehat{Q}\times \widehat{Q} { \text{ for } \text{ which } } i-i^{\prime }=j-j^{\prime }\in \tau \mathbb {Z}^2 \right\} . \end{aligned}$$

We set

$$\begin{aligned} J_{ij}:= {\left\{ \begin{array}{ll} \displaystyle \frac{\Lambda }{|i-j|^{2+s}} &{} { \text{ if } \, (i,j)\in \widehat{\mathcal {Q}} \text{ and } i \ne j}\\ \quad \, \, \, \, 0 &{} { \text{ if }\, i = j,} \\ \displaystyle \frac{1}{|i-j|^{2+s}} &{}{ \text{ otherwise, }} \end{array}\right. } \end{aligned}$$

for \(\Lambda >1\) (to be chosen conveniently large in the sequel).

We claim that any planelike ground state with rationally independent slope \(\omega \in \mathbb {R}^2\) (with \(\omega \cdot n\ne 0\) for any \(n \in \mathbb {Z}^2\)) for the Hamiltonian associated to this case with vanishing magnetic field (i.e. \(h_i:=0\) in (1.15)) possesses oscillations of order \(\tau \), for large \(\tau \).

For this, we argue by contradiction and suppose that (1.16) holds true with \(M=M(\tau )\) sublinear in \(\tau \), namely there exists a ground state \(u=u_{\omega ,\tau }\) such that

$$\begin{aligned} \partial u \subset \left\{ i \in \mathbb {Z}^2 : \frac{\omega }{|\omega |} \cdot i \in [0, M(\tau )] \right\} , \end{aligned}$$
(10.1)

and

$$\begin{aligned} \lim _{\tau \rightarrow +\infty }\frac{ M(\tau )}{\tau }=0. \end{aligned}$$
(10.2)

Since \(\omega \) is irrational, any straight line \(r_\omega \) with direction normal to \(\omega \) will get arbitrarily close to \(\tau \mathbb {Z}^2\). In particular, up to a translation, we may assume that the origin lies in a \(\frac{\tau }{64}\)-neighborhood of \(r_\omega \), and, from (10.1) and (10.2), we can write

$$\begin{aligned}&u_{i}=-1 \hbox { for any } i \hbox { for which } \frac{\omega }{|\omega |} \cdot i\ge \frac{\tau }{32}\nonumber \\ {\text{ and } } \quad&u_{i}=1 {\hbox { for any } i \hbox { for } \mathrm{which}} \, \frac{\omega }{|\omega |} \cdot i\le - \frac{\tau }{32}, \end{aligned}$$
(10.3)

as long as \(\tau \) is large enough.

We now reach a contradiction with the minimality of u by constructing a suitable competitor v with less energy. To this aim, we define

$$\begin{aligned} v_i:= {\left\{ \begin{array}{ll} -1 &{}{ \text{ if } }i\in \widehat{Q},\\ u_i &{} { \text{ otherwise }}. \end{array}\right. } \end{aligned}$$

Since u is supposed to be minimal, we have that

$$\begin{aligned} 0 \,\ge H_{\widehat{Q}}(u)-H_{\widehat{Q}}(v)= & {} \sum _{(i,j)\in \mathbb {Z}^4\setminus (\mathbb {Z}^2\setminus \widehat{Q})^2 } J_{ij} (v_iv_j -u_iu_j)\nonumber \\= & {} \Lambda \sum _{i,j\in \widehat{Q}} \frac{1-u_i u_j}{|i-j|^{2+s}}-2 \sum \limits _{\begin{array}{c} {i\in {\widehat{Q}}} \\ {j \not \in {\widehat{Q}}} \end{array}} \frac{(1+u_i)\, u_j}{|i-j|^{2+s}}\nonumber \\\ge & {} 4\Lambda \sum \limits _{\begin{array}{c} {i,j\in \widehat{Q}}\\ {\{u_i=1\}}\\ {\{u_j=-1\}} \end{array}} \frac{1}{|i-j|^{2+s}}-4 \sum \limits _{\begin{array}{c} {i\in \widehat{Q}}\\ {j\not \in \widehat{Q} } \end{array}} \frac{1}{|i-j|^{2+s}}. \end{aligned}$$
(10.4)

Now, from (10.3), we know that the number of sites \(i\in \widehat{Q}\) for which \(u_i=1\) is at least of the order \(c\tau ^2\), and similarly that the number of sites \(j\in \widehat{Q}\) for which \(u_j=-1\) is at least of the order \(c\tau ^2\), with \(c>0\) universal. Consequently, we have that

$$\begin{aligned} \sum \limits _{\begin{array}{c} {i,j\in {\widehat{Q}}}\\ {\{u_{i}=1\}} \\ {\{u_j=-1\}} \end{array}} \frac{1}{|i-j|^{2+s}} \ge \frac{c^{\prime }\,\tau ^4}{\tau ^{2+s}}=c^{\prime }\,\tau ^{2-s}, \end{aligned}$$
(10.5)

for some \(c^{\prime }>0\). On the other hand, using the index \(k:=j-i\),

$$\begin{aligned} \sum \limits _{\begin{array}{c} {i\in {\widehat{Q}}}\\ {j\not \in {\widehat{Q}}} \end{array}} \frac{1}{|i-j|^{2+s}}&\le C\, \sum \limits _{\begin{array}{c} {|i|_\infty \le \frac{\tau -1}{2}}\\ {|j|_\infty \ge \frac{\tau +1}{2} } \end{array}} \frac{1}{|i-j|_\infty ^{2+s}}\le C\, \sum \limits _{\begin{array}{c} {|i|_\infty \le \frac{\tau -1}{2}}\\ {|k|_\infty \ge \frac{\tau +1}{2} -|i|_\infty } \end{array}} \frac{1}{|k|_\infty ^{2+s}} \\&\le C^{\prime }\, \sum _{|i|_\infty \le \frac{\tau -1}{2}} \left( \frac{\tau +1}{2} -|i|_\infty \right) ^{-s} = C^{\prime \prime }\, \sum _{\ell =0}^{\frac{\tau -1}{2}} \left( \frac{\tau +1}{2} -\ell \right) ^{-s} \,\ell \\&\le C^{\prime \prime }\,\tau \, \sum _{\ell =0}^{\frac{\tau -1}{2}} \left( \frac{\tau +1}{2} -\ell \right) ^{-s} \le C^{\prime \prime \prime }\,\tau ^{2-s}, \end{aligned}$$

for some C, \(C^{\prime }\), \(C^{\prime \prime }\), \(C^{\prime \prime \prime }>0\).

Thus, we insert this and (10.5) into (10.4) and we find that

$$\begin{aligned} 0 \ge 4\tau ^{2-s}\,(c^{\prime }\, \Lambda -C^{\prime \prime \prime }), \end{aligned}$$

which is a contradiction if \(\Lambda \) is sufficiently large.

Conclusions

After a short review of the classical Ising model, we considered in this paper a spin system with long-range interactions. We gave rigorous proofs of three types of results:

  • the construction of ground state solutions whose phase separation stays at a bounded distance from any given hyperplane,

  • the construction of nonlocal minimal surfaces which stay at a bounded distance from any given hyperplane,

  • the asymptotic link between ground states of long-range Ising models and nonlocal minimal surfaces.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Cozzi, M., Dipierro, S. & Valdinoci, E. Planelike Interfaces in Long-Range Ising Models and Connections with Nonlocal Minimal Surfaces. J Stat Phys 167, 1401–1451 (2017). https://doi.org/10.1007/s10955-017-1783-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-017-1783-1

Keywords

Mathematics Subject Classification

Navigation