Skip to main content
Log in

Localization, Big-Jump Regime and the Effect of Disorder for a Class of Generalized Pinning Models

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

Abstract

One dimensional pinning models have been widely studied in the physical and mathematical literature, also in presence of disorder. Roughly speaking, they undergo a transition between a delocalized phase and a localized one. In mathematical terms these models are obtained by modifying the distribution of a discrete renewal process via a Boltzmann factor with an energy that contains only one body potentials. For some more complex models, notably pinning models based on higher dimensional renewals, other phases may be present. We study a generalization of the one dimensional pinning model in which the energy may depend in a nonlinear way on the contact fraction: this class of models contains the circular DNA case considered for example in Bar et al. (Phys Rev E 86:061904, 2012). We give a full solution of this generalized pinning model in absence of disorder and show that another transition appears. In fact the systems may display up to three different regimes: delocalization, partial localization and full localization. What happens in the partially localized regime can be explained in terms of the “big-jump” phenomenon for sums of heavy tail random variables under conditioning. We then show that disorder completely smears this second transition and we are back to the delocalization versus localization scenario. In fact we show that the disorder, even if arbitrarily weak, is incompatible with the presence of a big-jump.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5

Similar content being viewed by others

References

  1. Alexander, K.S.: The effect of disorder on polymer depinning transitions. Commun. Math. Phys. 279, 117–146 (2008)

    ADS  MathSciNet  MATH  Google Scholar 

  2. Alexander, K.S., Sidoravicius, V.: Pinning of polymers and interfaces by random potentials. Ann. Appl. Probab. 16, 636–669 (2006)

    MathSciNet  MATH  Google Scholar 

  3. Alexander, K.S., Zygouras, N.: Quenched and annealed critical points in polymer pinning models. Commun. Math. Phys. 291, 659–689 (2009)

    ADS  MathSciNet  MATH  Google Scholar 

  4. Alexander, K.S., Zygouras, N.: Path properties of the disordered pinning model in the delocalized regime. Ann. Appl. Prob. 24, 599–615 (2014)

    MathSciNet  MATH  Google Scholar 

  5. Armendáriz, I., Loulakis, M.: Conditional distribution of heavy tailed random variables on large deviations of their sum. Stoch. Proc. Appl. 121, 1138–1147 (2011)

    MathSciNet  MATH  Google Scholar 

  6. Bar, A., Kabakçıoğlu, A., Mukamel, D.: Denaturation of circular DNA: supercoil mechanism. Phys. Rev. E 84, 041935 (2011)

    ADS  Google Scholar 

  7. Bar, A., Kabakçıoğlu, A., Mukamel, D.: Denaturation of circular DNA: supercoils and overtwist. Phys. Rev. E 86, 061904 (2012)

    ADS  Google Scholar 

  8. Berger, Q., Giacomin, G., Khatib, M.: DNA melting structures in the generalized Poland-Scheraga model, ALEA. Lat. Am. J. Probab. Math. Stat. 15, 993–1025 (2018)

    MathSciNet  MATH  Google Scholar 

  9. Berger, Q., Giacomin, G., Khatib, M.: Disorder and denaturation transition in the generalized Poland-Scheraga model. Ann. H. Lebesgue 3, 299–339 (2020)

    MathSciNet  MATH  Google Scholar 

  10. Berger, Q., Giacomin, G., Lacoin, H.: Disorder and critical phenomena: the \(\alpha =0\) copolymer model. Probab. Theory Rel. Fields 174, 787–819 (2019)

    MathSciNet  MATH  Google Scholar 

  11. Berger, Q., Lacoin, H.: Pinning on a defect line: characterization of marginal disorder relevance and sharp asymptotics for the critical point shift. J. Inst. Math. Jussieu 17(2), 305–346 (2018)

    MathSciNet  MATH  Google Scholar 

  12. Borovkov, A.A., Borovkov, K.A.: On probabilities of large deviations for random walks. I. Regularly varying distribution tails. Theory Probab. Appl. 46, 193–213 (2000)

    MATH  Google Scholar 

  13. Borovkov, A.A., Mogul’skiĭ, A.A.: On large deviations of sums of independent random vectors on the boundary and outside of the cramér zone I. Theory Probab. Appl. 53(2), 301–311 (2009)

    MathSciNet  MATH  Google Scholar 

  14. Caravenna, F., den Hollander, F.: A general smoothing inequality for disordered polymers. Electron. Commun. Probab. 18, 1–15 (2013)

    MathSciNet  MATH  Google Scholar 

  15. Caravenna, F., Toninelli, F.L., Torri, N.: Universality for the pinning model in the weak coupling regime. Ann. Probab. 45, 2154–2209 (2017)

    MathSciNet  MATH  Google Scholar 

  16. Chen, X., Dagard, V., Derrida, B., Hu, Y., Lifshits, M., Shi, Z.: The Derrida–Retaux conjecture on recursive models. arXiv:1907.01601

  17. Dasgupta, C., Ma, S.-K.: Low-temperature properties of the random Heisenberg anti-ferromagnetic chain. Phys. Rev. B 22, 1305–1319 (1980)

    ADS  Google Scholar 

  18. Davis, B., McDonald, D.: An elementary proof of the local central limit theorem. J. Theor. Probab. 8, 693–701 (1995)

    MathSciNet  MATH  Google Scholar 

  19. Denisov, D., Dieker, A.B., Shneer, V.: Large deviations for random walks under subexponentiality: the big-jump domain. Ann. Probab. 36(5), 1946–199 (2008)

    MathSciNet  MATH  Google Scholar 

  20. Derrida, B., Giacomin, G., Lacoin, H., Toninelli, F.L.: Fractional moment bounds and disorder relevance for pinning models. Commun. Math. Phys. 287, 867–887 (2009)

    ADS  MathSciNet  MATH  Google Scholar 

  21. Derrida, B., Hakim, V., Vannimenus, J.: Effect of disorder on two-dimensional wetting. J. Stat. Phys. 66, 1189–1213 (1992)

    ADS  MathSciNet  MATH  Google Scholar 

  22. Derrida, B., Retaux, M.: The depinning transition in presence of disorder: a toy model. J. Stat. Phys. 156, 268–290 (2014)

    ADS  MATH  Google Scholar 

  23. Einert, T.R., Orland, H., Netz, R.R.: Secondary structure formation of homopolymeric single-stranded nucleic acids including force and loop entropy: implications for DNA hybridization. Eur. Phys. J. E 34, 55 (2011)

    Google Scholar 

  24. Ferrari, P.A., Landim, C., Sisko, V.V.: Condensation for a fixed number of independent random variables. J. Stat. Phys. 128, 1153–1158 (2007)

    ADS  MathSciNet  MATH  Google Scholar 

  25. Fisher, M.E.: Walks, walls, wetting, and melting. J. Stat. Phys. 34, 667–729 (1984)

    ADS  MathSciNet  MATH  Google Scholar 

  26. Fisher, D.S.: Critical behavior of random transverse-field Ising spin chains. Phys. Rev. B 51, 6411–6461 (1995)

    ADS  Google Scholar 

  27. Garel, T., Orland, H.: On the role of mismatches in DNA denaturation. arXiv:cond-mat/0304080

  28. Garel, T., Orland, H.: Generalized Poland-Scheraga model for DNA hybridization. Biopolymers 75, 453–467 (2004)

    Google Scholar 

  29. Giacomin, G.: Random Polymer Models. Imperial College Press, World Scientific (2007)

    MATH  Google Scholar 

  30. Giacomin, G.: Disorder and critical phenomena through basic probability models, École d’été de probablités de Saint-Flour XL-2010. Lecture Notes in Mathematics, vol. 2025. Springer (2011)

  31. Giacomin, G., Khatib, M.: Generalized Poland Sheraga denaturation model and two dimensional renewal processes. Stoch. Proc. Appl. 127, 526–573 (2017)

    MATH  Google Scholar 

  32. Giacomin, G., Lacoin, H.: The disordered lattice free field pinning model approaching criticality. arXiv:1912.10538

  33. Giacomin, G., Lacoin, H., Toninelli, F.L.: Marginal relevance of disorder for pinning models. Commun. Pure Appl. Math. 63, 233–265 (2010)

    MathSciNet  MATH  Google Scholar 

  34. Giacomin, G., Toninelli, F.L.: Estimates on path delocalization for copolymers at selective interfaces. Probab. Theory Rel. Fields 133, 464–482 (2005)

    MathSciNet  MATH  Google Scholar 

  35. Giacomin, G., Toninelli, F.L.: Smoothing of depinning transitions for directed polymers with quenched disorder. Phys. Rev. Lett. 96, 070602 (2006)

    ADS  Google Scholar 

  36. Giacomin, G., Toninelli, F.L.: Smoothing effect of quenched disorder on polymer depinning transitions. Commun. Math. Phys. 266, 1–16 (2006)

    ADS  MathSciNet  MATH  Google Scholar 

  37. Giacomin, G., Toninelli, F.L.: The localized phase of disordered copolymers with adsorption, ALEA. Lat. Am. J. Probab. Math. Stat. 1, 149–180 (2006)

    MathSciNet  MATH  Google Scholar 

  38. Giacomin, G., Toninelli, F.L.: On the irrelevant disorder regime of pinning models. Ann. Probab. 37, 1841–1875 (2009)

    MathSciNet  MATH  Google Scholar 

  39. Godrèche, C.: Condensation for random variables conditioned by the value of their sum. J. Stat. Mech. Theory Exp. 6, 063207 (2019)

    MathSciNet  MATH  Google Scholar 

  40. Grosskinsky, S., Chleboun, P., Schütz, G.M.: Instability of condensation in the zero-range process with random interaction. Phys. Rev. E 78, 030101 (2008)

    ADS  Google Scholar 

  41. Harris, A.B.: Effect of random defects on the critical behaviour of Ising models. J. Phys. C 7, 1671–1692 (1974)

    ADS  Google Scholar 

  42. Havret, B.: On the Lyapunov exponent of random transfer matrices and on pinning models with constraints, PhD thesis, Université de Paris (2019), https://tel.archives-ouvertes.fr/tel-02478078

  43. den Hollander, F.: Random polymers, Lectures from the 37th Probability Summer School held in Saint-Flour, 2007. Lecture Notes in Mathematics, vol. 1974. Springer (2009)

  44. Iglói, F., Monthus, C.: Strong disorder RG approach of random systems. Phys. Rep. 412, 277–431 (2005)

    ADS  MathSciNet  Google Scholar 

  45. Kingman, J.F.C.: Subadditive Ergodic Theory. Ann. Probab. 1, 882–909 (1973)

    MathSciNet  MATH  Google Scholar 

  46. Lacoin, H.: The martingale approach to disorder irrelevance for pinning models. Electron. Commun. Probab. 15, 418–427 (2010)

    MathSciNet  MATH  Google Scholar 

  47. del Molino, L.C.G., Chleboun, P., Grosskinsky, S.: Condensation in randomly perturbed zero-range processes. J. Phys. A 45, 205001 (2012)

    ADS  MathSciNet  MATH  Google Scholar 

  48. Neher, R.A., Gerland, U.: Intermediate phase in DNA melting. Phys. Rev. E 73, 030902R (2006)

    ADS  Google Scholar 

  49. Poland, D., Scheraga, H.A.: Theory of Helix-Coil Transitions in Biopolymers;: Statistical Mechanical Theory of Order-Disorder Transitions in Biological Macromolecules. Academic Press, New York (1970)

    Google Scholar 

  50. Rudnick, J., Bruinsma, R.: Effects of torsional strain on thermal denaturation of DNA. Phys. Rev. E 65, 030902(R) (2002)

    ADS  Google Scholar 

  51. Toninelli, F.L.: A replica-coupling approach to disordered pinning models. Commun. Math. Phys. 280, 389–401 (2008)

    ADS  MathSciNet  MATH  Google Scholar 

  52. Velenik, Y.: Localization and delocalization of random interfaces. Probab. Surv. 3, 112–169 (2006)

    MathSciNet  MATH  Google Scholar 

  53. Vezzani, A., Barkai, E., Burioni, R.: Single-big-jump principle in physical modeling. Phys. Rev. E 100, 012108 (2019)

    ADS  Google Scholar 

Download references

Acknowledgements

We are grateful to Quentin Berger and Hubert Lacoin for several exchanges. We thank in particular Hubert Lacoin for suggesting the argument of proof of Theorem B.1. G.G. also acknowledges support from Grant ANR-15-CE40-0020.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Giambattista Giacomin.

Additional information

Communicated by Simone Warzel.

Publisher's Note

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

Appendices

Appendix A: Circular DNA Models

In [6, 7] (to which we refer also for a more complete literature) the problem of modeling circular DNA is considered: circular DNA corresponds notably to the genetic structures called plasmids that are present in cells. Plasmids are also used for genetic manipulations. If a doubled stranded DNA has a circular structure, that is if the strands are not free at their ends but form an ring, then the separation of the two strands – even just locally, global separation may not be possible – generates a conflict with the double helix structure and the physical properties of the DNA polymer. In fact, in a double stranded DNA with free ends, the local separation of the two strands just induces a rotation in the chain. But in the circular case (Fig. 6) this rotation cannot take place: the ring of the two strands has a winding number that can change if the backbone of DNA polymer can absorb, typically with an energetic cost, this (over)twist. Another way of absorbing the winding number is by forming nontrivial spatial structures called supercoils at locations where the two strands are attached (this induces a strain on the base pairs involved, so the relative contact energy changes): everybody has experienced the formation of supercoils when trying to disentangle two ropes, or even just one rope.

Fig. 6
figure 6

A schematic image of a circular DNA with three loops and four supercoil sections. The thick line represents the segments of DNA on which the two stands are in contact. The thin line represents a single strand portion, and this happens only in loops

The physical phenomena we just described are very complex: the free end case is already of great complexity! Nonetheless the (simple!) Poland–Scheraga model turns out to be a very relevant model for DNA denaturation study in the free end case (see references in [7, 29]). The circular case is tackled in [6, 7]: a model is built from the Poland–Scheraga and the partition function of the model, with our notations, is (1.9). The function \(\log \Psi (m,N)\), cf. Def. 1.1, that enters the definition can be seen as a nonlocal energy. Note that the functional dependence in \(\Psi \) is only on the length of the polymer and on the number of contacts m. Moreover, with m contacts the total length of the loops is \(N-m\).

Let us take a closer look at the models considered in [6, 7]:

  1. (1)

    Model with overtwist. This first model is particularly simple:

    $$\begin{aligned} \Psi (m, N)\, =\, \exp \left( -\chi (N-m)^2/m\right) \, , \end{aligned}$$
    (A.1)

    with \(\chi >0\). In this case \(Q(m, N)\equiv 1\) and \(H(\rho )= -\chi (1-\rho )^2/\rho \). We refer to [6, 7] for explanations about the choice of the precise shape of this energy term. Simply put, the smaller the number m of contacts is, the less likely the configuration is. In other words, the opening of loops is penalized. The rationale is that if the loop length increases, then overtwist is produced in the backbone, which is costly from the energetic viewpoint.

  2. (2)

    Model with overtwist and supercoils. This is less straightforward and it involves choosing n supercoils among the m contacts. This is done by fair coin flipping: there is no loss of generality in this choice because a bias corresponds simply to an energetic change for supercoil contacts and in the model there is a real parameter w that accounts for that. Here is the choice in [7]:

    $$\begin{aligned} \Psi (m, N)\, =\, \sum _{n=0}^m \frac{1}{2^m} \left( {\begin{array}{c}m\\ n\end{array}}\right) \exp \left( n w - \chi \frac{(N-m-n)^2}{m}\right) \, , \end{aligned}$$
    (A.2)

    where \(\chi > 0\) and w are constants. The fair coin structure and the energetic term for supercoils are clear: added to that there is a penalization term that favors \(N-n \approx m\). Recall that in the simple overtwist case m close to N is favored. This is simply because n supercoils are formed and the twist that remains has to be absorbed by the portion of DNA, of length \(N-n\), which is not in supercoil form.

    The choice (A.2) is in the framework of Definition 1.1 with

    $$\begin{aligned} H(\rho ) = \sup _{\zeta \in [0, \rho ] } \psi (\zeta , \rho ) = \psi (\zeta _0(\rho ), \rho ), \end{aligned}$$
    (A.3)

    where

    $$\begin{aligned} \psi (\zeta , \rho ) := \zeta w -\rho \log 2 - \chi \frac{(1 - \rho -\zeta )^2}{\rho } + \rho \log \rho - \zeta \log ( \zeta ) - (\rho -\zeta ) \log (\rho -\zeta ). \end{aligned}$$
    (A.4)

    Q(mN) is therefore (implicitly) defined and one can derive the asymptotic behavior for \(N\rightarrow \infty \) and m/N asymptotically constant:

    $$\begin{aligned} Q(m,N) \sim q\left( \frac{m}{N}\right) ,\ \text { with }\ q(\rho ) := \sqrt{\frac{\rho }{ \zeta _0( \rho )(\rho - \zeta _0( \rho ))|\partial _\zeta ^2 \psi (\zeta _0( \rho ),\rho )|} }. \end{aligned}$$
    (A.5)

    The function \(\psi \) is concave on the convex domain \(\{0 \le \zeta \le \rho \le 1\}\). Thus, H is also concave. Moreover H is analytic on (0, 1).

  3. (3)

    Model with supercoils. This is the \(\chi =\infty \) limit of (A.2). In this limit \(N-n=m\) and, since \(n\le m\), we have that \(m\ge N/2\) and that the opening of a loop must be compensated by at least as many supercoils. Explicitly we obtain

    $$\begin{aligned} \Psi (m,N)\, =\, \frac{1}{2^m} \left( {\begin{array}{c}m\\ N-m\end{array}}\right) e^{ N w -mw} {\mathbf {1}}_{[1/2,1]}(m/N)\, . \end{aligned}$$
    (A.6)

    Since \(\Psi (m,N)=0\) for \(m<N/2\), this limit case does not fall into the framework of Definition 1.1. A wider set-up that encompasses all models in in [6, 7] can be found in [42, Ch. 4].

We remark that in both examples (1) and (2) \(H(0) = -\infty \) and (of course) \(H'(0)= \infty \). So, with the convention we have chosen to consider localized both the partly and fully localized cases, the circular DNA models are localized for all values of the parameters and they display a big-jump transition if \(\beta =0\) and \(\alpha >1\).

Appendix B: On the Strict Convexity of the Disordered Pinning Free Energy

Theorem B.1

Consider the \(\Psi \equiv 1\) model, that is the disordered pinning model. For every \(\beta \ge 0\) and every \(h >h_c(\beta )\) we have \(\partial ^2_h \textsc {f}(\beta , h) >0\).

Proof

Let us remark that for \(\beta =0\) the result can be established by explicit computations, but the proof that we give here works for the \(\beta =0\) case as well. In this proof \({{\mathbf {P}}} _{N, \omega }= {{\mathbf {P}}} _{N, \omega , \beta , h}\) and \( \mathrm {Var}_{N, \omega }\) is the variance with respect to \({{\mathbf {P}}} _{N, \omega }\). We know from [37, Proof of Theorem 2.1] that for \(h >h_c(\beta )\)

$$\begin{aligned} \partial ^2_h \textsc {f}(\beta , h) \, =\, \lim _{N \rightarrow \infty } \frac{1}{N} {{\mathbb {E}}} \, \mathrm {Var}_{N, \omega } \left( \sum _{j=1}^{N-1} \delta _j \right) \,. \end{aligned}$$
(B.1)

We are going to condition on even sites, so let us replace N with 2N and let us denote by \({{\mathcal {F}}} _e\) the \(\sigma \)-algebra generated by \(\delta _j\) with j even: \(\mathrm {Var}_{N, \omega , e} (\cdot )\) is going to denote the variance with respect to \({{\mathbf {P}}} _{N, \omega }(\, \cdot \, \vert {{\mathcal {F}}} _e)\). By Jensen’s inequality

$$\begin{aligned} \mathrm {Var}_{2N, \omega } \left( \sum _{j=1}^{2N-1} \delta _j \right) \, \ge \, {{\mathbf {E}}} _{2N, \omega } \left[ \mathrm {Var}_{2N, \omega , e} \left( \sum _{j=1}^{N} \delta _{2j-1} \right) \right] \, . \end{aligned}$$
(B.2)

We know consider the conditional variance on the set \(E_\sigma := \{ \tau : \, \delta _{2j} =\sigma _j \) for \(j =1,2, \ldots , N-1\}\) for every \(\sigma \in \{0,1\}^{\{1,\ldots , N-1\}}\). We set \(n(\sigma ):= \sum _{j=1}^{N-1} \sigma _j\) and \(\ell _0:=0\) and we define iteratively \(\ell _{j+1}= \min \{ \ell > \ell _j: \, \sigma _\ell =1\}\) for \(j \le n(\sigma )-1\). We then redifine \(\ell _j\) to be \(2\ell _j\) and set also \(\ell _{n(\sigma )+1}:=2N\). Therefore \(\ell _0, \ell _1, \ldots , \ell _{n(\sigma )+1}\) are the \(n(\sigma )\) pinned even sites, plus 0 and 2N that are pinned from the start. Note that

$$\begin{aligned} \sum _{j=1}^N \delta _{2j-1} \, =\, \sum _{k=1}^{n(\sigma )+1} \sum _{j=1+\ell _{k-1}/2}^{\ell _k/2} \delta _{2j-1}\, , \end{aligned}$$
(B.3)

and remark that, under \({{\mathbf {P}}} _{N, \omega }(\, \cdot \, \vert {{\mathcal {F}}} _e)(\tau )\) with \(\tau \in E_\sigma \), the random variables

$$\begin{aligned} \left( \sum _{j=1+\ell _{k-1}/2}^{\ell _k/2} \delta _{2j-1} \right) _{k=1, \ldots , n(\sigma )+1}\, , \end{aligned}$$
(B.4)

are independent. Therefore on \(E_\sigma \)

$$\begin{aligned}&\mathrm {Var}_{2N, \omega , e} \left( \sum _{j=1}^{N} \delta _{2j-1} \right) \, =\, \nonumber \\&\quad \sum _{k=1}^{n(\sigma )+1} \mathrm {Var}_{2N, \omega , e} \left( \sum _{j=1+\ell _{k-1}/2}^{\ell _k/2} \delta _{2j-1} \right) \, \ge \, \sum _{\begin{array}{c} k=1, \ldots , n(\sigma )+1 \\ \ell _k -\ell _{k-1}=2 \end{array}} \mathrm {Var}_{2N, \omega , e} \left( \delta _{\ell _k-1} \right) \,. \end{aligned}$$
(B.5)

Since \(\delta _{\ell _k-1}\), under the conditional measure we are considering, is just a Bernoulli random variable with parameter (we use the short-cut notation \(\omega =\omega _{\ell _k-1}\))

$$\begin{aligned} p( \omega )\, :=\, \frac{K(1)^2 \exp (h + \beta \omega )}{K(1)^2 \exp (h + \beta \omega ) +K(2)}\, , \end{aligned}$$
(B.6)

we see that for k such that \(\ell _k-\ell _{k-1}=2\)

$$\begin{aligned} \mathrm {Var}_{2N, \omega , e} \left( \delta _{\ell _k-1} \right) \, =\, p\left( \omega _{\ell _k-1} \right) \left( 1- p\left( \omega _{\ell _k-1} \right) \right) \, =: \, \sigma ^2\left( \omega _{\ell _k-1} \right) , \end{aligned}$$
(B.7)

and therefore

$$\begin{aligned} \mathrm {Var}_{2N, \omega , e} \left( \sum _{j=1}^{N} \delta _{2j-1} \right) \, \ge \, {{\mathbf {E}}} _{2N, \omega } \left[ \sum _{k=0}^{N-1} \delta _{2k} \delta _{2k+2}\sigma ^2\left( \omega _{\ell _k-1} \right) \right] \, . \end{aligned}$$
(B.8)

Now we set \(\sigma _\star ^2(L):= \inf \{ \sigma ^2(\omega ):\, \vert \omega \vert \le L\}>0\). We remark that \(\sigma _\star ^2(L)>0\) for every \(L>0\), but in what follows we are forced to work with L such that \({{\mathbb {P}}} (\vert \omega \vert < L)>0\), that is for L above a threshold. With this notation

$$\begin{aligned} \begin{aligned} {{\mathbb {E}}} \mathrm {Var}_{2N, \omega , e} \left( \sum _{j=1}^{N} \delta _{2j-1} \right) \,&\ge \, \sigma _\star ^2(L) {{\mathbb {E}}} {{\mathbf {E}}} _{2N, \omega } \left[ \sum _{k=0}^{N-1} \delta _{2k}\delta _{2k+2} {\mathbf {1}}_{\vert \omega _{2k+1}\vert \le L} \right] \\&\ge \, \sigma _\star ^2(L) \left( {{\mathbb {E}}} {{\mathbf {E}}} _{2N, \omega } \left[ \sum _{k=0}^{N-1} \delta _{2k}\delta _{2k+2} \right] - N {{\mathbb {P}}} \left( \vert \omega \vert >L\right) \right) \, , \end{aligned} \end{aligned}$$
(B.9)

and we are left with showing that

$$\begin{aligned} q\, :=\, \liminf _N \frac{1}{N} {{\mathbb {E}}} {{\mathbf {E}}} _{2N, \omega } \left[ \sum _{k=0}^{N-1} \delta _{2k}\delta _{2k+2} \right] \, >0\, , \end{aligned}$$
(B.10)

because it suffices to choose L so that \({{\mathbb {P}}} \left( \vert \omega \vert >L\right) \le q/2\) to obtain, see (B.1)–(B.2), that \(\partial _h^2 \textsc {f}(\beta , h)\ge \sigma _\star ^2(L) q/4>0\).

In order to establish (B.10) we want to show that the quantity under analysis is bounded below by \(\lim _NN^{-1} {{\mathbb {E}}} {{\mathbf {E}}} _{2N, \omega }[\sum _{j=1}^{2N} \delta _j]\), which is equal to \(2 \partial \textsc {f}(\beta , h)>0\), times a positive constant. This can be done by explicit estimates, but for sake of conciseness we use that, for any choice of a sequence \((b_N)_{N \in {{\mathbb {N}}} }\) of positive integer numbers satisfying \(\lim _N b_N= \infty \) and \(\lim _N{b_N}/N=0\), by [37, Theorem 2.2] we have uniformly on \(k\in [b_N, N-b_n]\cap {{\mathbb {N}}} \)

$$\begin{aligned} \lim _{N \rightarrow \infty } {{\mathbb {E}}} {{\mathbf {E}}} _{2N, \omega }[\delta _k] \, =\, \partial _h\textsc {f}(\beta , h) \ \ \ \ \text { and } \ \ \ \lim _{N \rightarrow \infty } {{\mathbb {E}}} {{\mathbf {E}}} _{2N, \omega }[\delta _k\delta _{k+2}]\, =:\, \textsc {l}(\beta ,h)\, , \end{aligned}$$
(B.11)

where the second statement is just the existence of the limit and (B.10) follows once \( \textsc {l}(\beta ,h)>0\) is shown. For this we write \({{\mathbf {E}}} _{2N, \omega }[\delta _k\delta _{k+2}]= {{\mathbf {E}}} _{2N, \omega }[\delta _k] {{\mathbf {E}}} _{2N, \omega }[\delta _{k+2}\vert \delta _{k}=1 ]\) and

$$\begin{aligned} {{\mathbf {E}}} _{2N, \omega }[\delta _{k+2}\vert \delta _{k}=1 ] \, =\, \frac{Z_{2, \theta ^{k}\omega , \beta ,h} Z_{N-k-2, \theta ^{k+2}\omega , \beta ,h} }{Z_{N-k, \theta ^{k}\omega , \beta ,h} }\, \ge \, C \exp \left( - \beta \left( \vert \omega _{k+1}\vert + \vert \omega _{k+2}\vert \right) \right) \, , \end{aligned}$$
(B.12)

where the constant \(C>0\) depends on h and on \(K(\cdot )\): this estimate is a standard surgery procedure ( [29, Ch. 2], full details can be found in [42, Sec. 5.5]) for which one uses notably the regularly varying character of \(K(\cdot )\). Therefore

$$\begin{aligned} {{\mathbf {E}}} _{2N, \omega }[\delta _k\delta _{k+2}]\, \ge Ce^{-2 \beta L} {{\mathbf {E}}} _{2N, \omega }[\delta _k] \left( 1-{\mathbf {1}}_{\vert \omega _{k+1}\vert + \vert \omega _{k+2}\vert > L } \right) \, , \end{aligned}$$
(B.13)

and, in turn, we have

$$\begin{aligned} {{\mathbb {E}}} {{\mathbf {E}}} _{2N, \omega }[\delta _k\delta _{k+2}]\, \ge \, Ce^{-2 \beta L} \left( {{\mathbb {E}}} {{\mathbf {E}}} _{2N, \omega }[\delta _k]- {{\mathbb {P}}} \left( \vert \omega _{1}\vert > L \right) \right) \, . \end{aligned}$$
(B.14)

It suffices now to choose L so that \({{\mathbb {P}}} \left( \vert \omega _{1}\vert > L \right) \le \partial _h\textsc {f}(\beta , h)/2\) to obtain that, uniformly in k like in (B.11), we have

$$\begin{aligned} \liminf _N {{\mathbb {E}}} {{\mathbf {E}}} _{2N, \omega }\left[ \delta _k\delta _{k+2}\right] \, \ge \, \frac{1}{2} Ce^{-2 \beta L} \partial _h\textsc {f}(\beta , h) \, >\, 0\, , \end{aligned}$$
(B.15)

and we are done. \(\square \)

We include here the result proved under restrictive conditions in Remark 1.5.

Proposition B.2

For every \(\beta \ge 0\) and every h

$$\begin{aligned} \textsc {f}_H(\beta , h)\, \ge \, H(0)\, . \end{aligned}$$
(B.16)

Proof

We can assume \(H(0)> -\infty \) and, with \(b>0\) and u, v and c(b) like in (1.4) of Definition 1.1, we obtain that

$$\begin{aligned}&Z^\Psi _{N, \omega \beta ,h} \, \ge \, c(b) \exp \left( N\min _ {\rho \in (0, b]} H(\rho ) - bN\right) \nonumber \\&\quad {{\mathbf {E}}} \left[ \exp \left( \beta \sum _{j=1}^N (\beta \omega _j +h) \delta _j\right) ; \, \tau _{\lfloor b N \rfloor } =N, \, \frac{\tau }{N}\cap \left( (0, b)\cup (1-b, 1)\right) = \emptyset \right] , \nonumber \\&\, \end{aligned}$$
(B.17)

a bound that is obtained simply by restricting the partition function to the renewals with \(\lfloor b N \rfloor \) contacts and all at distance at least bN from the boundary. Therefore

$$\begin{aligned} \liminf _{N \rightarrow \infty } \frac{1}{N} {{\mathbb {E}}} \log Z^\Psi _{N, \omega \beta ,h} \, \ge \, H(0) + \liminf _{N \rightarrow \infty } \frac{1}{N} {{\mathbb {E}}} \log {{\mathbf {E}}} \left[ \exp \left( \sum _{j=1}^N (\beta \omega _j +h) \delta _j\right) ; \, E_{N, b} \right] \, , \end{aligned}$$
(B.18)

with \(E_{N, b}:= \{ \tau _{\lfloor b N \rfloor } =N, \, (\tau /N)\cap \left( (0, b)\cup (1-b, 1)\right) = \emptyset \}\). Using \({{\mathbf {P}}} '(\cdot ):= {{\mathbf {P}}} (\cdot \vert E_{N, b})\) we see that by Jensen’s inequality the quantity of which we take inferior limit in the right-hand side of the last expression is bounded below by

$$\begin{aligned} \frac{1}{N} {{\mathbb {E}}} \left[ \sum _{j=1}^N (\beta \omega _j +h) {{\mathbf {E}}} '[\delta _j] \right] + \frac{1}{N} \log {{\mathbf {P}}} \left( E_{N, b}\right) \, =\, h \frac{ \lfloor bN\rfloor }{N}+ \frac{1}{N} \log {{\mathbf {P}}} \left( E_{N, b}\right) . \end{aligned}$$
(B.19)

Of course the limit of the first term is hb, which can be made arbitrarily small by choosing b small. The remaining term is bounded below, for \(N \rightarrow \infty \), by a (negative) quantity that vanishes as \(b \searrow 0\) because \( {{\mathbf {P}}} (E_{N, b})\) is bounded below by \(K(\lceil bN\rceil )^2\) times \({{\mathbf {P}}} ( \tau _{\lfloor b N \rfloor -2}= N- 2\lceil bN\rceil )\), so, by Proposition 3.1, \(\liminf _N (1/N) \log {{\mathbf {P}}} (E_{N, b})=0\) for \(\alpha >1\) and it vanishes as \(b \searrow 0\) for \(\alpha \in (0,1]\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Giacomin, G., Havret, B. Localization, Big-Jump Regime and the Effect of Disorder for a Class of Generalized Pinning Models. J Stat Phys 181, 2015–2049 (2020). https://doi.org/10.1007/s10955-020-02653-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-020-02653-6

Keywords

Mathematics Subject Classification

Navigation