Skip to main content
Log in

Truncated Linear Statistics Associated with the Eigenvalues of Random Matrices II. Partial Sums over Proper Time Delays for Chaotic Quantum Dots

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

Abstract

Invariant ensembles of random matrices are characterized by the distribution of their eigenvalues \(\{\lambda _1,\ldots ,\lambda _N\}\). We study the distribution of truncated linear statistics of the form \(\tilde{L}=\sum _{i=1}^p f(\lambda _i)\) with \(p<N\). This problem has been considered by us in a previous paper when the p eigenvalues are further constrained to be the largest ones (or the smallest). In this second paper we consider the same problem without this restriction which leads to a rather different analysis. We introduce a new ensemble which is related, but not equivalent, to the “thinned ensembles” introduced by Bohigas and Pato. This question is motivated by the study of partial sums of proper time delays in chaotic quantum dots, which are characteristic times of the scattering process. Using the Coulomb gas technique, we derive the large deviation function for \(\tilde{L}\). Large deviations of linear statistics \(L=\sum _{i=1}^N f(\lambda _i)\) are usually dominated by the energy of the Coulomb gas, which scales as \(\sim N^2\), implying that the relative fluctuations are of order 1 / N. For the truncated linear statistics considered here, there is a whole region (including the typical fluctuations region), where the energy of the Coulomb gas is frozen and the large deviation function is purely controlled by an entropic effect. Because the entropy scales as \(\sim N\), the relative fluctuations are of order \(1/\sqrt{N}\). Our analysis relies on the mapping on a problem of p fictitious non-interacting fermions in N energy levels, which can exhibit both positive and negative effective (absolute) temperatures. We determine the large deviation function characterizing the distribution of the truncated linear statistics, and show that, for the case considered here (\(f(\lambda )=1/\lambda \)), the corresponding phase diagram is separated in three different phases.

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
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9

Similar content being viewed by others

Notes

  1. Expressions of the type \(P_{N,\kappa }(s) \underset{N\rightarrow \infty }{\sim } \exp (-N^q \Phi )\) must be understood as \(\lim _{N \rightarrow \infty } (-1/N^q) \ln P_{N,\kappa }(s) = \Phi \).

  2. References [6, 7] has introduced the symmetrized Wigner–Smith matrix \(\mathcal {Q}_s=-\mathrm{i}\mathcal {S}^{-1/2}\,\partial _\varepsilon \mathcal {S}\,\mathcal {S}^{-1/2}\), with the same spectrum of eigenvalues than \(\mathcal {Q}\). The precise statement of these references is that \(1/\mathcal {Q}_s\) is a Wishart matrix. Because \(\mathcal {S}\) and \(\mathcal {Q}\) are only independent in the unitary case, \(1/\mathcal {Q}\) is a Wishart matrix only in this case, strictly speaking. However its eigenvalues are always given by the Laguerre distribution (2.3).

  3. We have used

    $$\begin{aligned} \int _{U(N)} \mathrm{d}U\, U_{i_1 j_1} U_{i_2 j_2} U^*_{k_1 l_1} U^*_{k_2 l_2}&= W(N,1^2) \left( \delta _{i_1 k_1} \delta _{i_2 k_2} \delta _{j_1 l_1} \delta _{j_2 l_2} + \delta _{i_1 k_2} \delta _{i_2 k_1} \delta _{j_1 l_2} \delta _{j_2 l_1} \right) \\ \nonumber&\quad + W(N,2) \left( \delta _{i_1 k_1} \delta _{i_2 k_2} \delta _{j_1 l_2} \delta _{j_2 l_1} + \delta _{i_1 k_2} \delta _{i_2 k_1} \delta _{j_1 l_1} \delta _{j_2 l_2} \right) , \end{aligned}$$

    where the Weingarten functions \(W(N,\sigma )\) is a function of the matrix size and the permutation of indices. In particular: \( W(N,1^2) = \frac{1}{N^2-1} \) and \( W(N,2) = -\frac{1}{N(N^2-1)} \).

  4. The more general statement of Ref. [46] concerns sub-block of matrix \(\mathcal {Q}_s\), introduced in the previous footnote. The distributions of the two matrices \(\mathcal {Q}\) and \(\mathcal {Q}_s\) coincide for \(\beta =2\).

  5. The numerical analysis of Sect. 4.2.2 has shown that the large N result describes very well the distribution already for \(N\gtrsim 50\).

References

  1. Beenakker, C.W.J.: Random-matrix theory of quantum transport. Rev. Mod. Phys. 69(3), 731–808 (1997)

    Article  ADS  MathSciNet  Google Scholar 

  2. Berggren, T., Duits, M.: Mesoscopic fluctuations for the thinned Circular Unitary Ensemble. Preprint math. PR (2016). arXiv:1611.00991

  3. Bohigas, O., Pato, M.P.: Randomly incomplete spectra and intermediate statistics. Phys. Rev. E 74, 036212 (2006)

    Article  ADS  MathSciNet  Google Scholar 

  4. Borot, G., Eynard, B., Majumdar, S.N., Nadal, C.: Large deviations of the maximal eigenvalue of random matrices. J. Stat. Mech. 1111, P11024 (2011)

    Article  MathSciNet  Google Scholar 

  5. Brouwer, P.W., Büttiker, M.: Charge-relaxation and dwell time in the fluctuating admittance of a chaotic cavity. Europhys. Lett. 37(7), 441 (1997)

    Article  ADS  Google Scholar 

  6. Brouwer, P.W., Frahm, K.M., Beenakker, C.W.J.: Quantum mechanical time-delay matrix in chaotic scattering. Phys. Rev. Lett. 78(25), 4737 (1997)

    Article  ADS  Google Scholar 

  7. Brouwer, P.W., Frahm, K.M., Beenakker, C.W.J.: Distribution of the quantum mechanical time-delay matrix for a chaotic cavity. Waves Random Media 9, 91–104 (1999)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  8. Büttiker, M., Prêtre, A., Thomas, H.: Dynamic conductance and the scattering matrix of small conductors. Phys. Rev. Lett. 70(26), 4114 (1993)

    Article  ADS  Google Scholar 

  9. Charlier, C., Claeys, T.: Thinning and conditioning of the Circular Unitary Ensemble. Preprint math-ph (2016). arXiv:1604.08399

  10. Christen, T., Büttiker, M.: Gauge invariant nonlinear electric transport in mesoscopic conductors. Europhys. Lett. 35(7), 523–528 (1996)

    Article  ADS  Google Scholar 

  11. Christen, T., Büttiker, M.: Low-frequency admittance of quantized Hall conductors. Phys. Rev. B 53, 2064–2072 (1996)

    Article  ADS  Google Scholar 

  12. Collins, B.: Moments and cumulants of polynomial random variables on unitary groups, the Itzykson–Zuber integral, and free probability. Int. Math. Res. Notices 2003(17), 953–982 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  13. Cunden, F.D., Facchi, P., Vivo, P.: A shortcut through the Coulomb gas method for spectral linear statistics on random matrices. J. Phys. A 49(13), 135202 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Cunden, F.D., Maltsev, A., Mezzadri, F.: Density and spacings for the energy levels of quadratic Fermi operators. Preprint (2016). arXiv:1609.09777

  15. Cunden, F.D., Mezzadri, F., Simm, N., Vivo, P.: Correlators for the Wigner–Smith time-delay matrix of chaotic cavities. J. Phys. A 49(18), 18LT01 (2016)

  16. Cunden, F.D., Mezzadri, F., Simm, N., Vivo, P.: Large-\(N\) expansion for the time-delay matrix of ballistic chaotic cavities. J. Math. Phys. 57(11), 111901 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. de Carvalho, C.A.A., Nussenzveig, H.M.: Time delay. Phys. Rep. 364, 83–174 (2002)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  18. De Pasquale, A., Facchi, P., Parisi, G., Pascazio, S., Scardicchio, A.: Phase transitions and metastability in the distribution of the bipartite entanglement of a large quantum system. Phys. Rev. A 81, 052324 (2010)

    Article  ADS  Google Scholar 

  19. Dean, D.S., Majumdar, S.N.: Large deviations of extreme eigenvalues of random matrices. Phys. Rev. Lett. 97, 160201 (2006)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Dean, D.S., Majumdar, S.N.: Extreme value statistics of eigenvalues of Gaussian random matrices. Phys. Rev. E 77, 041108 (2008)

    Article  ADS  MathSciNet  Google Scholar 

  21. Dyson, F.J.: Statistical theory of the energy levels of complex systems. I. J. Math. Phys. 3(1), 140–156 (1962)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  22. Dyson, F.J.: Statistical theory of the energy levels of complex systems. II. J. Math. Phys. 3(1), 157–165 (1962)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  23. Fyodorov, Y.V., Sommers, H.-J.: Statistics of resonance poles, phase shift and time delays in quantum chaotic scattering: Random matrix approach for systems with broken time-reversal invariance. J. Math. Phys. 38(4), 1918–1981 (1997)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  24. Gasparian, V., Christen, T., Büttiker, M.: Partial densities of states, scattering matrices and Green’s functions. Phys. Rev. A 54(5), 4022–4031 (1996)

    Article  ADS  Google Scholar 

  25. Grabsch, A., Majumdar, S.N., Texier, C.: Truncated linear statistics associated with the top eigenvalues of random matrices. J. Stat. Phys. 167(2), 234–259 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  26. Grabsch, A., Majumdar, S.N., Texier, C.: Time delays and coherent AC transport in chaotic cavities—random matrices and Coulomb gas (in preparation)

  27. Grabsch, A., Texier, C.: Capacitance and charge relaxation resistance of chaotic cavities—joint distribution of two linear statistics in the Laguerre ensemble of random matrices. Europhys. Lett. 109(5), 50004 (2015)

    Article  ADS  Google Scholar 

  28. Grabsch, A., Texier, C.: Distribution of spectral linear statistics on random matrices beyond the large deviation function—Wigner time delay in multichannel disordered wires. J. Phys. A 49, 465002 (2016)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  29. Khoruzhenko, B.A., Savin, D.V., Sommers, H.-J.: Systematic approach to statistics of conductance and shot-noise in chaotic cavities. Phys. Rev. B 80, 125301 (2009)

    Article  ADS  Google Scholar 

  30. Kottos, T.: Statistics of resonances and delay times in random media: beyond random matrix theory. J. Phys. A 38, 10761–10786 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  31. Kuipers, J., Savin, D.V., Sieber, M.: Efficient semiclassical approach for time delays. New J. Phys. 16, 123018 (2014)

    Article  ADS  Google Scholar 

  32. Lambert, G.: Incomplete determinantal processes: from random matrix to Poisson statistics. Preprint math.PR (2016). arXiv:1612.00806

  33. Lehmann, N., Savin, D.V., Sokolov, V.V., Sommers, H.-J.: Time delay correlations in chaotic scattering: random matrix approach. Physica D 86, 572–585 (1995)

    Article  ADS  MATH  Google Scholar 

  34. Majumdar, S.N., Vergassola, M.: Large deviations of the maximum eigenvalue for Wishart and Gaussian random matrices. Phys. Rev. Lett. 102, 060601 (2009)

    Article  ADS  Google Scholar 

  35. Majumdar, S.N., Schehr, G.: Top eigenvalue of a random matrix: large deviations and third order phase transition. J. Stat. Mech. 1, P01012 (2014)

    Article  MathSciNet  Google Scholar 

  36. Majumdar, S.N., Schehr, G., Villamaina, D., Vivo, P.: Large deviations of the top eigenvalue of large Cauchy random matrices. J. Phys. A 46, 022001 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  37. Marčenko, V.A., Pastur, L.A.: Distribution of eigenvalues for some sets of random matrices. Math. USSR 1(4), 457 (1967)

    Article  Google Scholar 

  38. Mehta, M.L.: Random Matrices, 3rd edn. Elsevier/Academic Press, New York (2004)

    MATH  Google Scholar 

  39. Mello, P.A., Kumar, N.: Quantum Transport in Mesoscopic Systems—Complexity and Statistical Fluctuations. Oxford University Press, New York (2004)

    Book  Google Scholar 

  40. Mezzadri, F., Simm, N.J.: Tau-function theory of quantum chaotic transport with beta = 1, 2, 4. Commun. Math. Phys. 324, 465–513 (2013)

    Article  ADS  MATH  Google Scholar 

  41. Nadal, C., Majumdar, S.N., Vergassola, M.: Phase transitions in the distribution of bipartite entanglement of a random pure state. Phys. Rev. Lett. 104, 110501 (2010)

    Article  ADS  MATH  Google Scholar 

  42. Nadal, C., Majumdar, S.N., Vergassola, M.: Statistical distribution of quantum entanglement for a random bipartite state. J. Stat. Phys. 142(2), 403–438 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  43. Novaes, M.: Statistics of time delay and scattering correlation functions in chaotic systems. I. Random matrix theory. J. Math. Phys. 56, 062110 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  44. Pathria, R.K., Beale, P.D.: Statistical Mechanics, 3rd edn. Elsevier Academic Press, Amsterdam (2011)

    MATH  Google Scholar 

  45. Savin, D.V.: Private communication

  46. Savin, D.V., Fyodorov, Y.V., Sommers, H.-J.: Reducing nonideal to ideal coupling in random matrix description of chaotic scattering: application to the time-delay problem. Phys. Rev. E 63, 035202 (2001)

    Article  ADS  Google Scholar 

  47. Texier, C.: Wigner time delay and related concepts–application to transport in coherent conductors. Physica E 82, 16–33 (2016) (see preprint cond-mat. arXiv:1507.00075 for an updated version)

  48. Texier, C., Majumdar, S.N.: Wigner time-delay distribution in chaotic cavities and freezing transition. Phys. Rev. Lett. 110, 250602 (2013)

    Article  ADS  Google Scholar 

  49. Texier, C., Roux, G.: Physique statistique : des processus élémentaires aux phénomènes collectifs. Dunod, Paris (2017)

    Google Scholar 

  50. Tracy, C.A., Widom, H.: Level-spacing distribution and the Airy kernel. Commun. Math. Phys. 159, 151–174 (1994)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  51. Tracy, C.A., Widom, H.: On orthogonal and symplectic matrix ensembles. Commun. Math. Phys. 177, 727–754 (1996)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  52. Tricomi, F.G.: Integral Equations. Interscience, London (1957)

    MATH  Google Scholar 

  53. Vivo, P., Majumdar, S.N., Bohigas, O.: Large deviations of the maximum eigenvalue in Wishart random matrices. J. Phys. A 40, 4317–4337 (2007)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  54. Vivo, P., Majumdar, S.N., Bohigas, O.: Distributions of conductance and shot noise and associated phase transitions. Phys. Rev. Lett. 101, 216809 (2008)

    Article  ADS  Google Scholar 

  55. Vivo, P., Majumdar, S.N., Bohigas, O.: Probability distributions of linear statistics in chaotic cavities and associated phase transitions. Phys. Rev. B 81, 104202 (2010)

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We are indebted to Dmitry Savin for many useful discussions on chaotic scattering.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Christophe Texier.

Appendix 1: Distribution of the Injectance for \(\beta =2\)

Appendix 1: Distribution of the Injectance for \(\beta =2\)

In this section, we derive the distribution of the injectance of a quantum dot defined in Sect. 2.1:

$$\begin{aligned} \overline{\nu }_\alpha = \frac{1}{2\pi } \sum _{i\in \mathrm {contact}\,\alpha }\mathcal {Q}_{ii}. \end{aligned}$$
(7.1)

We will restrict our study to the situation where \(\beta =2\) because the joint distribution of the \(\mathcal {Q}_{ii}\)’s is known only in this case. We will derive the distribution of \(\overline{\nu }_\alpha \) using a Coulomb gas approach. The computation is very similar to the one described in [48] where the Wigner time delay (case where all contacts contribute) is studied. We start from the distribution of a subblock \(\mathcal {Q}_{p}\) of size \( p\times p\) on the diagonal of \(\mathcal {Q}\) given in [46] for \(\beta =2\):

$$\begin{aligned} P(\mathcal {Q}_{p}^{-1}) \propto (\det \mathcal {Q}_{p})^{-N} \mathrm {e}^{- \mathop {\mathrm {tr}}\nolimits \left\{ \mathcal {Q}_{p}^{-1} \right\} }. \end{aligned}$$
(7.2)

If we suppose that the contact \(\alpha \) has \(p\) open channels, we can rewrite the injectance as:

$$\begin{aligned} \overline{\nu }_\alpha = \frac{1}{2\pi } \mathop {\mathrm {tr}}\nolimits \left\{ \mathcal {Q}_{p} \right\} . \end{aligned}$$
(7.3)

Denoting \(\lbrace \lambda _i \rbrace \) the eigenvalues of \(\mathcal {Q}_{p}^{-1}\), we can write

$$\begin{aligned} \overline{\nu }_\alpha = \frac{1}{2\pi } \sum _{i=1}^{p} \frac{1}{\lambda _i}, \end{aligned}$$
(7.4)

where the joint probability density function of the \(\lbrace \lambda _i \rbrace \) is given by

$$\begin{aligned} P(\lambda _1 ,\ldots , \lambda _{p}) \propto \prod _{i<j} \left| \lambda _i - \lambda _j \right| ^2 \prod _{i=1}^{p} \lambda _i^{N} \mathrm {e}^{-\lambda _i}. \end{aligned}$$
(7.5)

Rescale the eigenvalues as \(\lambda _i = px_i\) and introduce the empirical density

$$\begin{aligned} \rho (x) = \frac{1}{p} \sum _{i=1}^{p} \delta (x - x_i). \end{aligned}$$
(7.6)

The distribution of

$$\begin{aligned} s = 2\pi \overline{\nu }_\alpha = \frac{1}{p} \sum _{i=1}^{p} \frac{1}{x_i} = \int \frac{\rho (x)}{x} \mathrm{d}x \end{aligned}$$
(7.7)

is given by

$$\begin{aligned} P_{N,\kappa }(s) \simeq \frac{ \displaystyle \int \mathcal {D}\rho \, \mathrm {e}^{-\kappa ^2 N^2 \mathscr {E}[\rho ]} \, \delta \left( \int \rho (x) \mathrm{d}x - 1 \right) \delta \left( \int \frac{\rho (x)}{x} \mathrm{d}x - s \right) }{ \displaystyle \int \mathcal {D}\rho \, \mathrm {e}^{- \kappa ^2 N^2 \mathscr {E}[\rho ]} \, \delta \left( \int \rho (x) \mathrm{d}x - 1 \right) }, \end{aligned}$$
(7.8)

where we denoted \(\kappa = p/N\) and introduced the energy (see Sect. 3)

$$\begin{aligned} \mathscr {E}[\rho ] = - \int \mathrm{d}x \int \mathrm{d}y \, \rho (x) \rho (y) \ln \left| x-y \right| + \int \mathrm{d}x \left( x - \frac{1}{\kappa } \ln x \right) \rho (x). \end{aligned}$$
(7.9)

The integrals in (7.8) are dominated by the minimum of the energy under the constraints imposed by the \(\delta \)-functions. Therefore, we introduce the functional

$$\begin{aligned} \mathscr {F}[\rho ,\mu _0,\mu _1] = \mathscr {E}[\rho ] + \mu _0 \left( \int \rho - 1 \right) + \mu _1 \left( \int \frac{\rho (x)}{x} \mathrm{d}x - s \right) , \end{aligned}$$
(7.10)

where \(\mu _0\) and \(\mu _1\) are Lagrange multipliers. Denote \(\rho ^\star (x;\kappa ,s)\) the density which dominates the numerator of (7.8). It is given by \(\left. \frac{\delta \mathscr {F}}{\delta \rho } \right| _{\rho ^\star } = 0\), which reads:

$$\begin{aligned} 2 \int \rho ^\star (y;\kappa ,s) \ln \left| x-y \right| \mathrm{d}y = x - \frac{1}{\kappa } \ln x + \mu _0 + \frac{\mu _1}{x} \quad \text {for} \quad x \in \mathrm {Supp}(\rho ^\star ). \end{aligned}$$
(7.11)

It is more convenient to derive this relation with respect to x:

(7.12)

The values \(\mu _0^\star \) and \(\mu _1^\star \) taken by the Lagrange multipliers are fixed by imposing the constraints

$$\begin{aligned} \int \rho ^\star (x;\kappa ,s) \mathrm{d}x = 1, \quad \int \frac{\rho ^\star (x;\kappa ,s)}{x} \mathrm{d}x = s. \end{aligned}$$
(7.13)

Similarly, denote \(\rho _0^\star (x)\) the density which dominates the denominator. The distribution of s is then given by:

$$\begin{aligned} P_{N,\kappa }(s) \sim \exp \left\{ - \kappa ^2 N^2 \left( \mathscr {E}[\rho ^\star (x;\kappa ,s)]-\mathscr {E}[\rho _0^\star (x)] \right) \right\} . \end{aligned}$$
(7.14)

The thermodynamic identity (3.21) mentioned in the body of the paper allows a direct computation of the energy via the relation:

$$\begin{aligned} \frac{\mathrm {d}\mathscr {E}[\rho ^\star (x;\kappa ,s)]}{\mathrm {d}s} = - \mu _1^\star (\kappa ;s). \end{aligned}$$
(7.15)

Our aim is now to compute the density \(\rho ^\star \). As in the body of the paper, depending on the values of the parameters \(\kappa \) and s, we will find different types of densities \(\rho ^\star \), which we interpret as different “phases”.

1.1 Phase I: Solution with One Compact Support

Let us assume that the solution \(\rho ^\star \) has one compact support [ab]. Then, Eq. (7.12) can be solved using Tricomi’s theorem [52]. We obtain:

$$\begin{aligned} \rho ^\star (x;\kappa ,s) = \frac{x+c}{2\pi x^2} \sqrt{(x-a)(b-x)}. \end{aligned}$$
(7.16)

It is convenient to parametrize this solution in terms of \(u = \sqrt{a/b}\). Then, imposing \(\rho ^\star (a;\kappa ,s)= \rho ^\star (b;\kappa ,s) = 0\), along with the constraints (7.13) yields:

$$\begin{aligned}&\displaystyle v = \sqrt{a b} = \frac{2u}{\kappa } \frac{(2\kappa +1)u^2 -2u + (2\kappa +1)}{(1-u^2)^2}, \end{aligned}$$
(7.17)
$$\begin{aligned}&\displaystyle \mu _1^\star =- \frac{4u^2}{\kappa ^2} \frac{((2\kappa +1)u^2 -2u + (2\kappa +1))(u^2 -2(2\kappa +1)u +1)}{(1-u^2)^4}, \end{aligned}$$
(7.18)
$$\begin{aligned}&\displaystyle c = \frac{\mu _1^\star }{v}, \end{aligned}$$
(7.19)
$$\begin{aligned}&\displaystyle = \sigma _\kappa (u) = (1-u)^2 \frac{-u^4 + 4 (3\kappa +1)u^3 + 2(4\kappa -3)u^2 + 4(3\kappa +1)u-1}{16u^2((2\kappa +1)u^2 -2u + (2\kappa +1))}. \qquad \end{aligned}$$
(7.20)

Given \(\kappa \) and s, the last equation allows to compute u, from which all the other parameters are deduced. One can check that in the limit \(\kappa \rightarrow 1\), one recovers the equations given in [48].

The optimal density \(\rho _0^\star \) for the denominator can be deduced from \(\rho ^\star \) by releasing the constraint (setting \(\mu _1^\star = 0\)).

1.1.1 Domain of Validity

This solution exists as long as \(\rho ^\star \) is positive, which corresponds to \(x+c \ge 0\). This gives the condition

$$\begin{aligned} \kappa \ge \frac{1 - 3 u + 3 u^2 - u^3}{2 u (3 + u^2)}, \end{aligned}$$
(7.21)

which can be rewritten as \(s < s_{I}(\kappa )\). This corresponds to the lower domain delimited by the upper solid line on Fig. 10 (left).

1.1.2 Typical Fluctuations

The typical fluctuations are controlled by the minimum of the energy, which is given by \(\mu _1^\star =0\). The density is then

$$\begin{aligned} \rho _0^\star (x) = \frac{\sqrt{(x-a_0)(b_0-x)}}{2\pi x}, \end{aligned}$$
(7.22)

where

$$\begin{aligned} a_0 = \frac{2 \kappa +1 - 2\sqrt{\kappa (\kappa +1)}}{\kappa }, \quad b_0 = \frac{2 \kappa +1 + 2\sqrt{\kappa (\kappa +1)}}{\kappa }. \end{aligned}$$
(7.23)

The corresponding value of s is \(\kappa \), and expanding Eqs. (7.18, 7.20) for s close to \(\kappa \) yields:

$$\begin{aligned} \mu _1^\star = - \frac{s-\kappa }{\kappa ^3(\kappa +1)} + O((s-\kappa )^2). \end{aligned}$$
(7.24)

The energy is obtained by simple integration, via Eq. (7.15):

$$\begin{aligned} \mathscr {E}[\rho ^\star (x;\kappa ,s)]-\mathscr {E}[\rho _0^\star (x)] = \frac{(s-\kappa )^2}{2\kappa ^3(\kappa +1)} + \mathcal {O}((s-\kappa )^3). \end{aligned}$$
(7.25)

Thus, the distribution of s near \(\kappa \) is given by:

$$\begin{aligned} P_{N,\kappa }(s) \sim \exp \left\{ - N^2 \frac{(s-\kappa )^2}{2\kappa (\kappa +1)} \right\} . \end{aligned}$$
(7.26)

We recover the leading term of the variance given in the introduction, Eq. (2.10):

$$\begin{aligned} \mathrm {Var}(s) \simeq \frac{\kappa (1+\kappa )}{N^2}. \end{aligned}$$
(7.27)

1.1.3 Limiting Behaviour \(s \rightarrow 0\)

The limit \(s \rightarrow 0\) corresponds to \(\mu _1^\star \rightarrow +\infty \). Expanding Eqs. (7.17, 7.18, 7.19, 7.20) in this limit gives:

$$\begin{aligned} \mu _1^\star = \frac{1}{s^2} - \frac{\kappa +2}{2\kappa } \frac{1}{s} + \mathcal {O}(1). \end{aligned}$$
(7.28)

Using again Eq. (7.15), we obtain:

$$\begin{aligned} \mathscr {E}[\rho ^\star (x;\kappa ,s)] = \frac{1}{s} + \frac{\kappa +2}{2\kappa } \ln s + \mathcal {O}(1), \end{aligned}$$
(7.29)

thus:

$$\begin{aligned} P_{N,\kappa }(s) \sim s^{-N^2\kappa (\kappa +2)/2 } e^{- \kappa ^2 N^2/s} \quad \text {for}\quad s \rightarrow 0. \end{aligned}$$
(7.30)

1.2 Phase II: Solution with an Isolated Eigenvalue

As for the Wigner time delay [48] (\(\kappa =1\)) and in Section 5, we look for a solution with an isolated eigenvalue:

$$\begin{aligned} \rho ^\star (x;\kappa ,s) = \frac{1}{p} \delta (x-x_1) + \tilde{\rho }(x), \end{aligned}$$
(7.31)

with now \(\int \tilde{\rho } = 1-1/p\). The minimization of \(\mathscr {F}\) with respect to \(\tilde{\rho }\) reads:

(7.32)

And minimization with respect to \(x_1\) gives

$$\begin{aligned} 2 \int \frac{\tilde{\rho }(y)}{x_1 - y} = 1 - \frac{1}{\kappa x_1} - \frac{\mu _1}{x_1^2}, \quad \text {for}\; x \in \mathrm {Supp}(\tilde{\rho }). \end{aligned}$$
(7.33)

In addition, the constraint becomes:

$$\begin{aligned} s = \frac{1}{px_1} + \int \frac{\tilde{\rho }(x)}{x} \mathrm{d}x. \end{aligned}$$
(7.34)

For \(x_1\) to give a “macroscopic” contribution to s, we must have \(x_1 = \mathcal {O}(N^{-1})\). Then equation (7.33) imposes \(\mu _1 = -x_1/\kappa + \mathcal {O}(p^{-2})\). Finally, at leading order in \(p\), we get:

$$\begin{aligned}&\displaystyle \tilde{\rho }(x) = \rho _0^\star (x) + \mathcal {O}(p^{-1}),\end{aligned}$$
(7.35)
$$\begin{aligned}&\displaystyle x_1 = \frac{1}{p(s-\kappa )} + \mathcal {O}(p^{-2}),\end{aligned}$$
(7.36)
$$\begin{aligned}&\displaystyle \mu _1^\star = - \frac{1}{\kappa p} \frac{1}{s - \kappa } + \mathcal {O}(p^{-2}). \end{aligned}$$
(7.37)

1.2.1 Domain of Validity

This solution remains valid as long as the separate eigenvalue is away from the bulk, namely \(x_1 < a_0\). This gives the condition

$$\begin{aligned} s > \kappa + \frac{1}{pa_0} = s_{II}(\kappa ). \end{aligned}$$
(7.38)

This corresponds to the upper domain represented on Fig. 10 (left). Note that \(s_{II}(\kappa ) \rightarrow \kappa \) as \(N \rightarrow \infty \).

1.2.2 Large Deviation Function

The energy can be computed analytically at leading order:

$$\begin{aligned} \mathcal {E}[\rho ^\star (x;\kappa ,s)] - \mathcal {E}[\rho _0^\star ] = \frac{1}{\kappa p} \ln [p(s - \kappa )] + \frac{\mathscr {C}}{p} + \mathcal {O}(p^{-2}). \end{aligned}$$
(7.39)

where \(\mathscr {C}=-1 - 2 \ln 2\) was introduced above. The constant term is the same as the one appearing in Sect. 5 and Ref. [48]. It arises from corrections of order \(p^{-1}\) to the density \(\rho ^\star \) [26]. From the expression of the energy, we deduce the expression of the distribution of s, for \(s > s_c\):

$$\begin{aligned} P_{N,\kappa }(s) \sim (s-\kappa )^{- N}. \end{aligned}$$
(7.40)

The simplification \(p^2/(\kappa p)=N\) has thus produced the same exponent as for the tail of the distribution of the sum (truncated or not) of proper times. This is explained from the interpretation of Ref. [46], where it was shown that \(\mathcal {Q}_{ii}\) coincides with the partial time delay for \(\beta =2\).

Fig. 10
figure 10

Left Domains of existence of each solution, for \(N=200\). The upper solid line corresponds to \(s = s_{I}(\kappa )\) and delimits the domain of existence of phase I. The lower solid line is \(s = s_{II}(\kappa )\). Phase II exists above this line. As \(N \rightarrow \infty \), it goes to the dashed line \(s = \kappa \), where occurs the transition in the thermodynamic limit. Right phase diagram for the Coulomb gas

1.3 Summary

We obtained the following scalings for the distribution:

$$\begin{aligned} P_{N,\kappa }(s) \underset{N \rightarrow \infty }{\sim } {\left\{ \begin{array}{ll} \exp \left\{ - \kappa ^2 N^2 \varPsi _-(\kappa ;s) \right\} &{} \quad \text {for} \; s < s_{I}(\kappa ) \\ (\kappa N)^{-N} \exp \left\{ - N \varPsi _+(\kappa ;s) \right\} &{} \quad \text {for}\; s > s_{II}(\kappa ) \end{array}\right. } \end{aligned}$$
(7.41)

where the large deviation \(\varPsi _+\) is given by:

$$\begin{aligned} \varPsi _+(\kappa ;s) = \ln (s - \kappa ) - \kappa (1 + 2 \ln 2), \end{aligned}$$
(7.42)

and \(\varPsi _-\) has the following limiting behaviours:

$$\begin{aligned} \varPsi _-(\kappa ;s) \simeq {\left\{ \begin{array}{ll} \displaystyle \frac{1}{s} + \frac{\kappa +2}{2\kappa } \ln s + \mathcal {O}(1) &{} \quad \text {as}\; s \rightarrow 0 \\ \displaystyle \frac{(s-\kappa )^2}{2\kappa ^3(\kappa +1)} + \mathcal {O}((s-\kappa )^3) &{}\quad \text {as}\; s \rightarrow \kappa \end{array}\right. } \end{aligned}$$
(7.43)

The precise point \(s_t\) where the transition between the two phases occurs can be obtained by matching the two large deviation functions. One can show that, for \(N \rightarrow \infty \), \(s_t \rightarrow \kappa \). Therefore for large N we have, for s close to \(\kappa \):

$$\begin{aligned} \lim _{N \rightarrow \infty } - \frac{1}{N^2} \ln P_{N,\kappa }(s) \simeq \left\{ \begin{array}{cl} \displaystyle \frac{(s-\kappa )^2}{2\kappa ^3(\kappa +1)} &{} \text { for } s < \kappa \\ 0 &{} \text { for } s > \kappa \end{array} \right. \end{aligned}$$
(7.44)

This corresponds to a second order phase transition. This was already the case in Ref. [48] in the study of the full linear statistics (\(\kappa =1\)).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Grabsch, A., Majumdar, S.N. & Texier, C. Truncated Linear Statistics Associated with the Eigenvalues of Random Matrices II. Partial Sums over Proper Time Delays for Chaotic Quantum Dots. J Stat Phys 167, 1452–1488 (2017). https://doi.org/10.1007/s10955-017-1780-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-017-1780-4

Keywords

Navigation