Skip to main content
Log in

Continuum Limit of Random Matrix Products in Statistical Mechanics of Disordered Systems

  • Published:
Communications in Mathematical Physics Aims and scope Submit manuscript

Abstract

We consider a particular weak disorder limit (continuum limit) of matrix products that arise in the analysis of disordered statistical mechanics systems, with a particular focus on random transfer matrices. The limit system is a diffusion model for which the leading Lyapunov exponent can be expressed explicitly in terms of modified Bessel functions, a formula that appears in the physical literature on these disordered systems. We provide an analysis of the diffusion system as well as of the link with the matrix products. We then apply the results to the framework considered by Derrida and Hilhorst (J Phys A 16:2641–2654, 1983), which deals in particular with the strong interaction limit for disordered Ising model in one dimension and that identifies a singular behavior of the Lyapunov exponent (of the transfer matrix), and to the two dimensional Ising model with columnar disorder (McCoy–Wu model). We show that the continuum limit sharply captures the Derrida and Hilhorst singularity. Moreover we revisit the analysis by McCoy and Wu (Phys Rev 176:631–643, 1968) and remark that it can be interpreted in terms of the continuum limit approximation. We provide a mathematical analysis of the continuum approximation of the free energy of the McCoy–Wu model, clarifying the prediction (by McCoy and Wu) that, in this approximation, the free energy of the two dimensional Ising model with columnar disorder is \(C^\infty \) but not analytic at the critical temperature.

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

Similar content being viewed by others

References

  1. Alberts, T., Khanin, K., Quastel, J.: The intermediate disorder regime for directed polymers in dimension \(1 + 1\). Ann. Probab. 42, 1212–1256 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  2. Baxter, R.J.: Exactly Solved Models in Statistical Mechanics. Academic Press, Cambridge (1982)

    MATH  Google Scholar 

  3. Bougerol, P., Lacroix, J.: Products of Random Matrices with Applications to Schrödinger Operators. Progress in Probability and Statistics, vol. 8. Birkhäuser Boston, Inc., Boston (1985)

    MATH  Google Scholar 

  4. Bouchaud, J.-P., Comtet, A., Georges, A., Le Doussal, P.: Classical diffusion of a particle in a one-dimensional random force field. Ann. Phys. 201, 285–341 (1990)

    Article  ADS  MathSciNet  Google Scholar 

  5. Caravenna, F., Sun, R., Zygouras, N.: Polynomial chaos and scaling limits of disordered systems. J. Eur. Math. Soc. 19, 1–65 (2017)

    Article  MathSciNet  MATH  Google Scholar 

  6. Cattiaux, P., Chafaï, D., Guillin, A.: Central limit theorems for additive functionals of ergodic Markov diffusions processes. ALEA Lat. Am. J. Probab. Math. Stat. 9, 337–382 (2012)

    MathSciNet  MATH  Google Scholar 

  7. Cohen, J.E., Kesten, H., Newman, C.M. (eds.): Random matrices and their applications. In: Proceedings of the AMS-IMS-SIAM Joint Summer Research Conference Held at Bowdoin College, Brunswick, Maine, June 17–23, 1984. Contemporary Mathematics, vol. 50, American Mathematical Society (1986)

  8. Comtet, A., Desbois, J., Monthus, C.: Localization properties in one-dimensional disordered supersymmetric quantum mechanics. Ann. Phys. 239, 312–350 (1995)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Comtet, A., Luck, J.M., Texier, C., Tourigny, Y.: The lyapunov exponent of products of random \(2\times 2\) matrices close to the identity. J. Stat. Phys. 150, 13–65 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Comtet, A., Texier, C., Tourigny, Y.: Lyapunov exponents, one-dimensional Anderson localization and products of random matrices. J. Phys. A 46, 254003 (2013). 20 pp

    Article  ADS  MathSciNet  MATH  Google Scholar 

  11. Crisanti, A., Paladin, G., Vulpiani, A.: Products of random matrices in statistical physics, volume 104 of Springer Series in Solid-State Sciences. Springer, Berlin (1993)

  12. Derrida, B., Hilhorst, H.J.: Singular behaviour of certain infinite products of random \(2\times 2\) matrices. J. Phys. A 16, 2641–2654 (1983)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. Dubois, L.: Real cone contractions and analyticity properties of the characteristic exponents. Nonlinearity 21, 2519–2536 (2008)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Dunster, T.M.: Bessel functions of purely imaginary order, with an application to second-order linear differential equations having a large parameter. SIAM J. Math. Anal. 21, 995–1018 (1990)

    Article  MathSciNet  MATH  Google Scholar 

  15. Ethier, S., Kurtz, T.: Markov Processes. Characterization and convergence. Wiley Series in Probability and Mathematical Statistics: Probability and Mathematical Statistics (1986)

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

    Article  ADS  Google Scholar 

  17. Freidlin, M.I., Wentzell, A.D.: Random Perturbations of Dynamical Systems. Grundlehren der Mathematischen Wissenschaften, vol. 260, 3rd edn. Springer, Berlin (2012)

    Book  Google Scholar 

  18. Friedlander, F.G.: Diffraction of pulses by a circular cylinder. Commun. Pure Appl. Math. 7, 705–732 (1954)

    Article  MathSciNet  MATH  Google Scholar 

  19. Frisch, H.L., Lloyd, S.P.: Electron levels in a one-dimensional random lattice. Phys. Rev. 120, 1175–1189 (1960)

    Article  ADS  MATH  Google Scholar 

  20. Genovese, G., Giacomin, G., Greenblatt, R.L.: Singular behavior of the leading Lyapunov exponent of a product of random \(2\times 2\) matrices. Commun. Math. Phys. 351, 923–958 (2017)

    Article  ADS  MATH  Google Scholar 

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

    Book  MATH  Google Scholar 

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

  23. Grabsch, A., Texier, C., Tourigny, Y.: One-dimensional disordered quantum mechanics and Sinai diffusion with random absorbers. J. Stat. Phys. 155, 237–276 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  24. Havret, B.: Regular expansion for the characteristic exponent of a product of \(2 \times 2\) random matrices. arXiv:1804.06166 (2018), Math. Phys. Anal. Geom. (to appear)

  25. Karatzas, I., Shreve, S.E.: Brownian Motion and Stochastic Calculus. Graduate Texts in Mathematics, vol. 113. Springer, Berlin (1998)

    Book  Google Scholar 

  26. Krantz, S.G., Parks, H.R.: A Primer of Real Analytic Functions. Birkhüser Verlag, Basel (1992)

    Book  MATH  Google Scholar 

  27. Luck, J.M.: Critical behavior of the aperiodic quantum Ising chain in a transverse magnetic field. J. Stat. Phys. 72, 417–458 (1993)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  28. Maruyama, G., Tanaka, H.: Ergodic property of \(N\)-dimensional recurrent Markov processes. Mem. Fac. Sci. Kyushu Univ. Ser. A 13, 157–172 (1959)

    MathSciNet  MATH  Google Scholar 

  29. McCoy, B.M., Wu, T.T.: The Two-Dimensional Ising Model. Harvard University Press, Cambridge (1973)

    Book  MATH  Google Scholar 

  30. McCoy, B.M., Wu, T.T.: Theory of a two-dimensional Ising model with random impurities. I. Thermodynamics. Phys. Rev. 176, 631–643 (1968)

    Article  ADS  MathSciNet  Google Scholar 

  31. Olver, F.W.J., Olde Daalhuis, A.B., Lozier, D.W., Schneider, B.I., Boisvert, R.F., Clark, C.W., Miller, B.R., Saunders, B.V. (Editors): NIST Digital Library of Mathematical Functions. http://dlmf.nist.gov/, Release 1.0.16 of 2017-09-18 (2017)

  32. Nieuwenhuizen, ThM, Luck, J.M.: Exactly soluble random field Ising models in one dimension. J. Phys. A 19, 1207–1227 (1986)

    Article  ADS  MathSciNet  Google Scholar 

  33. Pardoux, E., Veretennikov, A.Yu.: On the Poisson equation and diffusion approximation. I. Ann. Probab. 29, 1061–1085 (2001)

  34. Ramola, K., Texier, C.: Fluctuations of random matrix products and 1D Dirac equation with random mass. J. Stat. Phys. 157, 497–514 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Rogers, L., Williams, D.: Diffusions, Markov processes, and martingales, vol. 2. Itô calculus. Cambridge University Press, Cambridge (2000)

  36. Ruelle, D.: Analycity properties of the characteristic exponents of random matrix products. Adv. Math. 32, 68–80 (1979)

    Article  MathSciNet  MATH  Google Scholar 

  37. Ruelle, D.: Thermodynamic Formalism. The Mathematical Structures of Equilibrium Statistical Mechanics. 2nd edition. Cambridge Mathematical Library. Cambridge University Press, Cambridge (2004)

  38. Sadel, C., Schulz-Baldes, H.: Random Lie group actions on compact manifolds: a perturbative analysis. Ann. Probab. 36, 2224–2257 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  39. Schomerus, H., Titov, M.: Statistics of finite-time Lyapunov exponents in a random time-dependent potential. Phts. Rev. E 66, 066207 (2002). (11 pages)

    Article  ADS  MathSciNet  Google Scholar 

  40. Shankar, R., Murthy, G.: Nearest-neighbor frustrated random-bond model in \(d={2}\): some exact results. Phys. Rev. B 36, 536–545 (1987)

    Article  ADS  MathSciNet  Google Scholar 

  41. Stroock, D., Varadhan, S.: Multidimensional Diffusion Processes. Classics in Mathematics. Springer, Berlin (2006)

    Google Scholar 

  42. Taleb, M.: Large deviations for a Brownian motion in a drifted Brownian potential. Ann. Probab. 29, 1173–1204 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  43. Zanon, N., Derrida, B.: Weak disorder expansion of Lyapunov exponents in a degenerate case. J. Stat. Phys. 50, 509–528 (1988)

    Article  ADS  MATH  Google Scholar 

Download references

Acknowledgements

We are very grateful to Bernard Derrida for very insightful discussions, and to an anonymous referee for pointing out significant errors in the statement and proof of Proposition 1.5 in an earlier manuscript. Part of this work was developed while G.G. and R.L.G. were visiting IHP (Paris) during the spring-summer 2017 trimester. G.G. acknowledges the support of grant ANR-15-CE40-0020. The work of R.L.G. was funded by the European Research Council under the European Union’s Horizon 2020 Programme, ERC Consolidator Grant UniCoSM (Grant Agreement No. 724939).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Giambattista Giacomin.

Additional information

Communicated by H. Spohn

Publisher's Note

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

Appendix A: The McCoy–Wu Model

Appendix A: The McCoy–Wu Model

In [30], McCoy and Wu examined a two-dimensional Ising model with bond disorder of a particular type (subsequently known as the McCoy–Wu model): the couplings between sites in neighboring columns have a constant strength \(E_1\), while the couplings between neighboring sites in the same column take a random value \(E_2(n)\) which is fixed within each row but varies independently—keeping the same distribution—between different rows (Fig. 1). They showed that in the thermodynamic limit the free energy per site of this model is given (up to the subtraction of an analytic function of \(\beta \)) by

$$\begin{aligned} \textsc {f}_{{\text {MW}}}(\beta )\, :=\, \frac{1}{4\pi }\int _{-\pi }^\pi {\mathcal {L}} ^{{\text {MW}}}_{\beta }(\theta ) \,\text {d}\theta \,, \end{aligned}$$
(A.1)

where \({\mathcal {L}} ^{{\text {MW}}}_{\beta }(\theta )\) is the Lyapunov exponent of the random matrix

$$\begin{aligned} M_\beta (\theta ) :=\begin{pmatrix} 1 &{}\quad \frac{a}{a^2+b^2} \\ \frac{a}{a^2+b^2} \lambda &{}\quad \frac{\lambda }{a^2+b^2} \end{pmatrix}, \end{aligned}$$
(A.2)

with

$$\begin{aligned} a(\theta )\, =\, -2z_1 \frac{\sin (\theta )}{\left| 1+z_1 \exp (i\theta ) \right| ^2}\ \ \text { and } \ \ b(\theta )\, =\, \frac{1-z_1^2}{\left| 1+z_1 \exp (i\theta ) \right| ^2}, \end{aligned}$$
(A.3)

where

$$\begin{aligned} z_1\, =\, \tanh \left( \beta E_1\right) , \ \ z_2(n)\, =\, \tanh \left( \beta E_2(n)\right) \ \ \text { and } \ \ \lambda =\lambda (n)= z_2^2(n). \end{aligned}$$
(A.4)

In [40] a different version of the model has been considered: vertical bounds are random in the horizontal direction and randomness is repeated in each line. This model, that allows frustration, is richer, but the features that are novel with respect to the McCoy–Wu model cannot be appreciated in the weak disorder limit: our analysis applies to [40] as well, but we will not develop this issue here.

Fig. 1
figure 1

The McCoy–Wu disordered version of the two dimensional Ising model: the disordered interactions are in the vertical direction and they are distributed in an IID fashion within one column. This disorder is just copied to all the other columns and the horizontal interactions are non random. The disorder enters the free energy formula via independent copies of the random variable \(\lambda = \tanh ^2 \left( \beta E_2\right) \)

To avoid trivialities we assume that \(E_1\ne 0\) as well as that \(E_2\) is a non degenerate random variable: it is immediately clear that the sign of \(E_2\) does not matter and just a little thought reveals that the sign of \(E_1\) is irrelevant too. Therefore we assume that \(E_1\in (0, \infty )\) and that \(E_2\) is a random variable taking values in \((0, \infty )\). It is helpful (mostly to simplify the presentation) to assume that \(E_2\) takes values in \([E_2^{-}, E_2^+]\), with \(0<E_2^-<E_2^+< \infty \).

Moreover one directly sees that \(a(\cdot )\) is odd and \(b(\cdot )\) is even, which yields that \({\mathcal {L}} ^{{\text {MW}}}_{\beta }(\cdot )\) is even: in fact \(D_\pm M_\beta (\theta ) D_\pm = M_\beta (-\theta )\), with \(D_\pm \) the diagonal matrix with \((+1,-1)\) on the diagonal. Therefore:

$$\begin{aligned} \textsc {f}_{{\text {MW}}}(\beta )\, :=\, \frac{1}{2\pi }\int _{0}^\pi {\mathcal {L}} ^{{\text {MW}}}_{\beta } (\theta ) \,\text {d}\theta . \end{aligned}$$
(A.5)

McCoy and Wu claim that for every \(\upsilon \in (0, \pi )\)—our focus is on \(\upsilon \) small—the function

$$\begin{aligned} \beta \mapsto \frac{1}{2\pi }\int _{\upsilon }^\pi {\mathcal {L}} ^{{\text {MW}}}_{\beta }(\theta ) \,\text {d}\theta \end{aligned}$$
(A.6)

is real analytic on \((0, \infty )\). This can be proven by applying the main result in [36] (see also [13]). We sketch the argument here by considering separately the case \(\theta \) bounded away from 0 and \(\pi \) and the case of \(\theta \) near \(\pi \): with \(\delta >0\) small

  • For \(\theta \in [\delta , \pi -\delta ]\) the matrix \(M_\beta (\theta )\) (with positive entries) maps the closure of the cone Q—here Q is first quadrant without the axes, that is the set of vectors with positive entries—to \(Q\cup \{0\}\). More precisely, by the hypothesis we have made on the support of Z, for every \(\delta \in (0, \pi /2)\) and every \(\varrho \in (0,1)\) the closure of Q is mapped into a cone whose closure is a subset of \(Q\cup \{0\}\) and this subset is the same for every choice of \(\theta \in [\delta , \pi -\delta ]\) and every \(\beta \in [\varrho ,1/\varrho ]\). This uniform cone property implies the real analyticity of \(\beta \mapsto {\mathcal {L}} ^{{\text {MW}}}_{\beta }(\theta )\) with a convergence radius that is bounded away from zero uniformly in \(\theta \in [\delta , \pi -\delta ]\) and \(\beta \in [\varrho ,1/\varrho ]\).

  • For \(\theta \in [\pi -\delta ,\pi ]\) we argue by observing first that

    $$\begin{aligned} a(\pi ) = 0, \quad b(\pi ) = \frac{1-z_1^2}{(1-z_1)^2} = \frac{1+ \tanh \beta E_1}{1 - \tanh \beta E_1} = e^{2 \beta E_1}, \end{aligned}$$
    (A.7)

    so

    $$\begin{aligned} M_\beta (\pi ) \,=\,\begin{pmatrix} 1 &{}\quad 0 \\ 0 &{}\quad e^{-4 \beta E_1} \tanh ^2 \beta E_2 \end{pmatrix}. \end{aligned}$$
    (A.8)

    Since \(e^{-4 \beta E_1} \tanh ^2 \beta E_2<1\) the action of \(M_\beta (\pi )\) contracts uniformly any cone of the form \(\{(x,y): \, y \ge \vert x\vert \}\), in the sense there exists \(\varrho >0\) such that \(M_\beta (\pi )\) sends \(\{(x,y): \, y \ge \vert x\vert \}\) into \(\{(x,y): \, y \ge (1+ \varrho ) \vert x\vert \}\), uniformly in \(\beta >0\) and \(E_2\). Elementary arguments show that this result is only slightly perturbed if we consider \(\theta \in [\pi -\delta ,\pi ]\) with \(\delta \) sufficiently small. This uniform cone property implies the real analyticity of \(\beta \mapsto {\mathcal {L}} ^{{\text {MW}}}_{\beta }(\theta )\) with a convergence radius that is bounded away from zero uniformly in \(\theta \in [\pi -\delta ,\pi ]\) and \(\beta >0\).

Therefore the true issue is the regularity (or lack of it) of

$$\begin{aligned} \beta \mapsto \frac{1}{2\pi }\int _{0}^\upsilon {\mathcal {L}} ^{{\text {MW}}}_{\beta }(\theta ) \,\text {d}\theta \,, \end{aligned}$$
(A.9)

for a \(\upsilon >0\) that can be chosen as small as one wishes. At this point McCoy and Wu claim that the only non analytic point of the map in (A.9) can be at \(\beta _c\) defined by

$$\begin{aligned} 2 \beta _c E_1 + {\mathbb {E}}\left[ \log \tanh \beta _c E_2 \right] \, =\, 0. \end{aligned}$$
(A.10)

To see that this is the only possible candidate, McCoy and Wu point out that

$$\begin{aligned} a(0) = 0, \quad b(0) = \frac{1-z_1^2}{(1+z_1)^2} =\frac{1- \tanh \beta E_1}{1 + \tanh \beta E_2} = e^{-2 \beta E_1}, \end{aligned}$$
(A.11)

so

$$\begin{aligned} M_\beta (0) =\begin{pmatrix} 1 &{} \quad 0 \\ 0 &{} \quad e^{4 \beta E_1} \tanh ^2 \beta E_2 \end{pmatrix}, \end{aligned}$$
(A.12)

and so

$$\begin{aligned} {\mathcal {L}} ^{MW}_\beta (0) = \max \left( 0, 4 \beta E_1 + 2 {\mathbb {E}}\left[ \log \tanh \beta E_2 \right] \right) \end{aligned}$$
(A.13)

for \(\beta \) real. This admits an analytic extension in a neighborhood of any positive \(\beta \) except for \(\beta _c\).

This is of course far from being close to a proof, since one has to control the integral over \(\theta \in (0, \upsilon )\) and not the value in zero. But McCoy and Wu perform also a more subtle analysis that can be understood precisely via the diffusion limit of matrix products that is at the center of our analysis. To explain this let us make a further manipulation to match more sharply our framework.

In fact, as it stands, \(M_\beta (\theta )\), cf. (A.2), is not of the form (1.1). But by noting that

$$\begin{aligned} \frac{1}{a^2(\theta )+b^2(\theta )} = \frac{(1+z_1)^4}{(1-z_1^2)^2} +O(\theta ^2) = \left( \frac{1+z_1}{1-z_1} \right) ^2+ O(\theta ^2) =e^{4 \beta E_1}+ O(\theta ^2), \end{aligned}$$
(A.14)

and

$$\begin{aligned} \frac{a}{a^2(\theta )+b^2(\theta )} = -2 \left( \frac{1+z_1}{1-z_1} \right) ^2 \theta + O(\theta ^2) = -2 e^{4 \beta E_1}+ O(\theta ^2), \end{aligned}$$
(A.15)

if we let

$$\begin{aligned} \widetilde{\varepsilon } := \frac{2 z_1}{(1-z_1)^2} \theta , \end{aligned}$$
(A.16)

we see that to leading order as \(\theta \searrow 0\)

$$\begin{aligned} \begin{pmatrix} 1 &{} \quad -\widetilde{\varepsilon } \\ -\widetilde{\varepsilon } \lambda &{}\quad e^{4 \beta E_1} \lambda \end{pmatrix} \end{aligned}$$
(A.17)

is \(M_\beta (\theta )\). The matrix in (A.17) is of the form (1.1) up to a conjugation and a change of variables: in fact

$$\begin{aligned} \begin{pmatrix} 1&{}\quad \varepsilon \\ \varepsilon Z &{}\quad Z \end{pmatrix} \,:=\, \begin{pmatrix} 1&{}\quad \widetilde{\varepsilon } e^{-2 \beta E_1}\\ \widetilde{\varepsilon } e^{-2 \beta E_1} \lambda &{}\quad e^{4 \beta E_1} \lambda \end{pmatrix} \, =\, \begin{pmatrix} 1 &{}\quad 0 \\ 0 &{}\quad - e^{2 \beta E_1} \end{pmatrix} \begin{pmatrix} 1 &{}\quad -\widetilde{\varepsilon } \\ - \widetilde{\varepsilon } \lambda &{}\quad e^{4 \beta E_1} \lambda \end{pmatrix} \begin{pmatrix} 1 &{}\quad 0 \\ 0 &{}\quad - e^{-2 \beta E_1} \end{pmatrix}, \end{aligned}$$
(A.18)

and we observe—recall (A.16)—that \(\varepsilon = c_\beta \theta \), with \(c_\beta = 2 \sinh (2\beta E_1)\).

Remark A.1

It is important to remark at this stage that the inverse temperature \(\beta \) and our fundamental parameter \(\alpha \)—we recall that \(\alpha \) is the unique non zero real solution to \({\mathbb {E}}Z^\alpha =1\) (\(Z=e^{4\beta E_1} \tanh ^2(\beta E_2)\) depends on \(\beta \)!) when such a solution exists and otherwise \(\alpha =0\)—should be seen as an analytic change of variable: this is treated in detail in Lemma A.2. In particular \(\alpha (\beta _c)=0\) and therefore \(\alpha (\beta )= (\beta -\beta _c) \alpha ' (\beta _c)+ O((\beta -\beta _c)^2)\), but the constant\(\alpha ' (\beta _c)\) depends of the law of Z (with \(\beta =\beta _c\)) and this expansion should be done more carefully when the disorder is weak because, as we will see, \(\alpha ' (\beta _c)\) becomes large in this limit: this is treated in (A.21)–(A.29).

What McCoy and Wu do at this point is

  • making a specific choice of \(Z=Z^\Delta =Z^\Delta _\beta \) that satisfies the hypotheses of Theorem 6.1 (say, with \(\sigma =1\) for simplicity); this actually implements two choices:

    1. 1.

      the first is evident and it is the fact that disorder can be made weak by making \(\Delta \) small;

    2. 2.

      the second is that \(\beta -\beta _c\) is chosen small and, precisely, of the order of \(\Delta \). As we will explain, if we set \(y=(\beta -\beta _c)/ \Delta \) and we keep \(y\in {\mathbb {R}}\) fixed, then \(\alpha (\beta ) \sim -C_{\beta _c} y \), and the constant \(C_{\beta _c}>0\) will be given explicit in the specific case that we are going to develop, see (A.29).

  • they choose also \(\upsilon \propto \Delta \): let us fix in an arbitrary fashion \(\upsilon = \Delta \).

In physical terms these choices correspond to focusing on the critical window in the limit of weak disorder. Cutting the integral at \(\theta = \Delta \) is harmless (as we have discussed before), but of course only as far as \(\Delta \) is kept fixed.

McCoy and Wu are in the end just dealing (recall (A.18)) with the Lyapunov exponent \(\widehat{{\mathcal {L}} }_{\Delta , \beta _c + y\Delta } (c_{\beta _c}x \Delta )\) of the matrix (we perform the change of variable \(\theta =x \Delta \))

$$\begin{aligned} \begin{pmatrix} 1 &{}\quad c_{\beta _c} x \Delta \\ c_{\beta _c} x \Delta Z_{\beta _c + y\Delta }^\Delta &{}\quad Z_{\beta _c + y\Delta }^\Delta \end{pmatrix}. \end{aligned}$$
(A.19)

But Theorem 6.1 (see also Theorem 1.6) tells us that \(\widehat{{\mathcal {L}} }_{\Delta , \beta _c + y\Delta } (c_{\beta _c }x)\) is asymptotically equivalent for \(\Delta \) small to \( \Delta {\mathcal {L}} _{1, C_{\beta _c}\alpha }( c_{\beta _c}x)\) so that

$$\begin{aligned} \int _{0}^\Delta {\mathcal {L}} ^{{\text {MW}}}_{\beta _c + y \Delta }(\theta ) \,\text {d}\theta \, \sim \, \Delta \int _{0}^1 \widehat{{\mathcal {L}} }_{\Delta , \beta _c + y\Delta } (c_{\beta _c }x)\,\text {d}x\, \sim \, \Delta ^2 \int _{0}^1 {\mathcal {L}} _{1,C_{\beta _c}\alpha } (c_{\beta _c}x)\,\text {d}x \end{aligned}$$
(A.20)

and we remind the reader that \({\mathcal {L}} _{1,C_{\beta _c}\alpha }( c_{\beta _c}x)\) has the explicit expression (1.6). Therefore, up to two inessential constants we arrived at (1.28). We did not fully justify the equivalences in (A.20), but this is not really the main problem: the main unresolved mathematical issue is that what we are after is proving that, for a fixed (possibly extremely small) value of \(\Delta \), the leftmost term in (A.20) is a \(C^\infty \) function of y at 0 and that the same expression is not analytic at zero. McCoy and Wu instead argue (and we prove in Theorem 1.7) that \(\alpha \mapsto \int _{0}^1 {\mathcal {L}} _{1,C_{\beta _c}\alpha }( c_{\beta _c}x)\,\text {d}x\) has these properties: but this second statement does not imply the first.

We now complement our discussion with the analysis of the specific distribution chosen for the disorder law in [29, 30]. We also discuss more in detail the change of variable \(\alpha (\beta )\).

1.1 Analysis of the Distribution Chosen by McCoy and Wu [29, 30]

McCoy and Wu consider the disordered variable \(\lambda =\tanh ^2( \beta E_2)\) that depends on a parameter that they call \(\mathtt {N}\) and it is large: in fact

$$\begin{aligned} \Delta \, =\, \mathtt {N}^{-2}. \end{aligned}$$
(A.21)

The density of \(\lambda \) is supported on \((0, \lambda _0)\) and equal to \(\mathtt {N}\lambda _0^{-\mathtt {N}}y^{\mathtt {N}-1}\) for \(y\in (0, \lambda _0)\). Necessarily \(\lambda _0=\lambda _0(\beta )= \tanh ^2(\beta E_2^*)\), with \(E_2^*\) the maximum value that the random variable \(E_2\) can reach. The density of \( Z= Z_\mathtt {N}\) (recall that Z is defined in (A.18)) is therefore

$$\begin{aligned} \mathtt {N}\lambda _1^{-\mathtt {N}}y^{\mathtt {N}-1} \mathbf {1}_{(0, \lambda _1)}(y)\ \ \ \ \text { with } \ \lambda _1(\beta )\, =\, e^{4\beta E_1}\lambda _0(\beta ). \end{aligned}$$
(A.22)

Note that for every \(\nu \in (-\mathtt {N}, \infty )\)

$$\begin{aligned} {\mathbb {E}}\left[ \left( Z_\mathtt {N}\right) ^\nu \right] \, =\, \frac{\lambda _1(\beta )^\nu }{1+ \frac{\nu }{\mathtt {N}}}, \end{aligned}$$
(A.23)

and we want to solve for \(\alpha =\alpha (\beta )\ne 0\) the equation

$$\begin{aligned} {{\mathbb {E}}\left[ \left( Z_\mathtt {N}\right) ^\alpha \right] \, -1} \, =\, \frac{\lambda _1(\beta )^\alpha - 1- \frac{\alpha }{\mathtt {N}}}{1+ \frac{\alpha }{\mathtt {N}}}\, =\, 0. \end{aligned}$$
(A.24)

On one hand we compute

$$\begin{aligned}&\log (\lambda _1(\beta ))\, \nonumber \\&\quad =\log \tanh ^2 (\beta _c E_2^*) - {\mathbb {E}}\left[ \log \tanh ^2 (\beta _c E_2)\right] \nonumber \\&\qquad + 4 (\beta -\beta _c)+ \log \tanh ^2 (\beta E_2^*) -\log \tanh ^2 (\beta _c E_2^*), \end{aligned}$$
(A.25)

and a straightforward computation yields

$$\begin{aligned} {\mathbb {E}}\left[ \log \tanh ^2 (\beta E_2)\right] \, =\, \log \tanh ^2 (\beta E_2^*) - \frac{1}{\mathtt {N}}, \end{aligned}$$
(A.26)

so for \(\beta \) close \(\beta _c\) we have

$$\begin{aligned} \log \lambda _1(\beta )\, =\, \frac{1}{\mathtt {N}}+ (\beta -\beta _c) \left( 4E_1 + \frac{1}{\sinh (2 \beta _c E_2^*)}\right) +O\left( ( \beta - \beta _c)^2\right) . \end{aligned}$$
(A.27)

On the other hand from (A.24) we see that if \(\alpha \) is fixed (so we look at \(\beta \) as a function of \(\alpha \)) we have

$$\begin{aligned} \log \lambda _1(\beta ) \,=\, \frac{\log \left( 1+ \frac{\alpha }{\mathtt {N}}\right) }{\alpha }{\mathop {=}\limits ^{\mathtt {N}\rightarrow \infty }}\, \frac{1}{\mathtt {N}}- \frac{\alpha }{2\mathtt {N}^2}+ O\left( \frac{1}{\mathtt {N}^3}\right) . \end{aligned}$$
(A.28)

By comparing (A.27) and (A.28) we see that if \((\beta -\beta _c) \mathtt {N}^2=O(1)\) then

$$\begin{aligned} \alpha (\beta ) \, =\, -(\beta -\beta _c) {\mathtt {N}^2} \left( 8E_1 + \frac{2}{\sinh (2 \beta _c E_2^*)}\right) + O\left( \mathtt {N}^{-1}\right) . \end{aligned}$$
(A.29)

1.2 On the Relation Between \(\beta \) and \(\alpha \)

Here are the details of the important map that relates \(\beta \) and \(\alpha \):

Lemma A.2

Assume that the support of the random variable \(E_2\) is bounded away from zero, so \(Z= \exp (4 \beta E_2) \tanh ^2(\beta E_2)\) is supported on a compact subinterval of \((0, \infty )\). Assume also that \(E_2\) is not constant. Then the equation

$$\begin{aligned} \frac{{\mathbb {E}}\left[ Z^\alpha \right] -1}{\alpha }\, =\, 0, \end{aligned}$$
(A.30)

has a unique real solution \(\alpha \) for every \(\beta >0\). This defines a map \(\beta \mapsto \alpha (\beta )\) from \((0, \infty )\) to \({\mathbb {R}}\). This map is decreasing, hence it is a bijection, and it is real analytic.

Proof

Let \(f: {\mathbb {R}}\times (0,\infty ) \rightarrow {\mathbb {R}}\) be the function defined by \(f(\alpha , \beta ):=\frac{{\mathbb {E}}\left[ Z^\alpha \right] -1}{\alpha }\), for \(\alpha \in {\mathbb {R}}{{\setminus }} \{0\}\) for \(\alpha \ne 0\), and \(f(0, \beta ):= {\mathbb {E}}[\log Z]\). It is straightforward to see, using the support properties of \(E_2\), that f is real analytic on its entire domain. Then we observe that, for fixed \(\alpha \), \(Z^\alpha \) is an increasing function of \(\beta \) and, by the support properties, this implies that \(\partial _\beta f(\alpha , \beta )>0\) for every \(\beta >0\) and \(\alpha \in {\mathbb {R}}\). On the other hand if we set \(g_\beta (\alpha )= {\mathbb {E}}\left[ Z^\alpha \right] -1\) we have that \(\partial _\alpha f(\alpha , \beta )=( \alpha g'_\beta (\alpha )- g_\beta (\alpha )) / \alpha ^2\). But \(g_\beta (\cdot )\) is (strictly) convex and \(g_\beta (0)=0\): so \( \alpha g'_\beta (\alpha )- g_\beta (\alpha )>0\) for \(\alpha \ne 0\) and therefore \(\partial _\alpha f(\alpha , \beta )>0\) for \(\alpha \ne 0\). For \(\alpha =0\) it suffices to perform a Taylor expansion of \(g_\beta (\alpha )\) at \(\alpha =0\) to see that \(\partial _\alpha f(\alpha , \beta )\vert _{\alpha =0}= g^{\prime \prime }_\beta (0)/2>0\). The proof is completed by applying the Implicit Function Theorem for real analytic functions [26]. \(\quad \square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Comets, F., Giacomin, G. & Greenblatt, R.L. Continuum Limit of Random Matrix Products in Statistical Mechanics of Disordered Systems. Commun. Math. Phys. 369, 171–219 (2019). https://doi.org/10.1007/s00220-019-03466-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-019-03466-9

Navigation