Skip to main content

Advertisement

Log in

Limit Theorems for Orthogonal Polynomials Related to Circular Ensembles

  • Published:
Journal of Theoretical Probability Aims and scope Submit manuscript

Abstract

For a natural extension of the circular unitary ensemble of order n, we study as \(n\rightarrow \infty \) the asymptotic behavior of the sequence of monic orthogonal polynomials \((\varPhi _{k,n}, k=0, \ldots , n)\) with respect to the spectral measure associated with a fixed vector, the last term being the characteristic polynomial. We show that, as \(n\rightarrow \infty \), the sequence of processes \((\log \varPhi _{\lfloor nt\rfloor ,n}(1), t \in [0,1])\) converges to a deterministic limit, and we describe the fluctuations and the large deviations.

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

Notes

  1. By a slight abuse of language, “ensemble” will mean such a distribution.

References

  1. Abramowitz, M., Stegun, I.A.: Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, 9th edn, pp. 258–259. Dover, New York (1972)

    MATH  Google Scholar 

  2. Borot, G., Guionnet, A.: Asymptotic expansion of \(\beta \) matrix models in the one-cut regime. Comm. Math. Phys. 317(2), 447–483 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  3. Bourgade, P., Hughes, C.P., Nikeghbali, A., Yor, M.: The characteristic polynomial of a random unitary matrix: a probabilistic approach. Duke Math. J. 145(1), 45–69 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  4. Bourgade, P., Nikeghbali, A., Rouault, A.: The characteristic polynomial on compact groups with Haar measure: some equalities in law. C. R. Math. Acad. Sci. Paris 345(4), 229–232 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  5. Bourgade, P., Nikeghbali, A., Rouault, A.: Circular Jacobi ensembles and deformed Verblunsky coefficients. Int. Math. Res. Not. IMRN 23, 4357–4394 (2009)

    MathSciNet  MATH  Google Scholar 

  6. Bourgade, P., Nikeghbali, A., Rouault, A.: Ewens Measures on Compact Groups and Hypergeometric Kernels. In: Séminaire de probabilités XLIII, volume 2006 of Lecture Notes in Mathematics, pp. 351–378. Springer, Berlin (2010)

  7. Dembo, A., Zeitouni, O.: Large Deviations Techniques and Applications, 2nd edn. Springer, Berlin (1998)

    Book  MATH  Google Scholar 

  8. Dette, H., Gamboa, F.: Asymptotic properties of the algebraic moment range process. Acta Math. Hung. 116, 247–264 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  9. Erdelyi, A., Magnus, W., Oberhettinger, F., Tricomi, F.G.: Higher Transcendental Functions, vol. I. Krieger, New York (1981)

    MATH  Google Scholar 

  10. Forrester, P.J.: Log-Gases and Random Matrices, volume 34 of London Mathematical Society Monographs Series. Princeton University Press, Princeton (2010)

    Google Scholar 

  11. Hiai, F., Petz, D.: A large deviation theorem for the empirical eigenvalue distribution of random unitary matrices. Ann. Inst. H. Poincaré Probab. Stat. 36(1), 71–85 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  12. Hughes, C.P., Keating, J.P., O’Connell, N.: On the characteristic polynomial of a random unitary matrix. Comm. Math. Phys. 220(2), 429–451 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  13. Jacod, J., Shiryaev, A.N.: Limit Theorems for Stochastic Processes. Springer, Berlin (1987)

    Book  MATH  Google Scholar 

  14. Keating, J.P., Snaith, N.C.: Random matrix theory and \(\zeta (1/2+it)\). Comm. Math. Phys. 214(1), 57–89 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  15. Killip, R.: Gaussian fluctuations for \(\beta \) ensembles. Int. Math. Res. Not. 2008(8) (2008); Art. ID rnn007, 19

  16. Killip, R., Nenciu, I.: Matrix models for circular ensembles. Int. Math. Res. Not. 50, 2665–2701 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  17. Killip, R., Stoiciu, M.: Eigenvalue statistics for CMV matrices: from Poisson to clock via random matrix ensembles. Duke Math. J. 146(3), 361–399 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  18. Neretin, Y.A.: Hua-type integrals over unitary groups and over projective limits of unitary groups. Duke Math. J. 114(2), 239–266 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  19. Neretin, YuA: Matrix analogues of the \(B\)-function, and the Plancherel formula for Berezin kernel representations. Math. Sb. 191(5), 67–100 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  20. Olver, F.W.J.: Asymptotics and Special Functions. Academic Press, New York (1974)

    MATH  Google Scholar 

  21. Rockafellar, R.T.: Integrals which are convex functionals, II. Pac. J. Math. 39(2), 439–469 (1971)

    Article  MathSciNet  MATH  Google Scholar 

  22. Rouault, A.: Asymptotic behavior of random determinants in the Laguerre, Gram and Jacobi ensembles. ALEA Lat. Am. J. Probab. Math. Stat. 3, 181–230 (2007). (electronic)

    MathSciNet  MATH  Google Scholar 

  23. Ryckman, E.: Linear statistics of point processes via orthogonal polynomials. J. Stat. Phys. 132(3), 473–486 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  24. Saff, E.B., Totik, V.: Logarithmic Potentials with External Fields, Volume 316 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer, Berlin (1997). Appendix B by Thomas Bloom

  25. Simon, B.: Orthogonal Polynomials on the Unit Circle. Part 1: Classical Theory. Colloquium Publications. American Mathematical Society 54, Part 1. American Mathematical Society (AMS), Providence, RI (2005)

  26. Simon, B.: CMV matrices: five years after. J. Comput. Appl. Math. 208, 120–154 (2007)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Alain Rouault.

Appendix

Appendix

1.1 Some Properties of \(\ell = \log \varGamma \) and \(\varPsi = \ell '\)

From the Binet formula (Abramowitz and Stegun [1] or Erdélyi et al. [9], p. 21), we have for \({\mathfrak {Re }\,}\ x > 0\)

$$\begin{aligned} \ell (x)= & {} \left( x-\frac{1}{2}\right) \log x -x +1 + \int _0^\infty f(s)[\mathrm{e}^{-sx} - \mathrm{e}^{-s}]\, \mathrm{d}s\nonumber \\= & {} \left( x-\frac{1}{2}\right) \log x -x + \frac{1}{2} \log (2\pi ) + \int _0^\infty f(s)\mathrm{e}^{-sx}\, \mathrm{d}s, \end{aligned}$$
(5.1)

where the function f is defined by

$$\begin{aligned} f(s) = \left[ \frac{1}{2}-\frac{1}{s}+ \frac{1}{\mathrm{e}^s -1}\right] \frac{1}{s} = 2\sum _{k=1}^\infty \frac{1}{s^2 + 4\pi ^2 k^2}, \end{aligned}$$
(5.2)

and satisfies for every \(s \ge 0\):

$$\begin{aligned} 0 < f(s) \le f(0)=1/12,\quad 0<\left( sf(s)+ \frac{1}{2}\right) <1. \end{aligned}$$
(5.3)

By differentiation (recall that \(\varPsi \) is the digamma function \(\varGamma '/\varGamma \)),

$$\begin{aligned} \log x - \varPsi (x) = \frac{1}{2x} +\int _0 ^\infty s f(s) \mathrm{e}^{-sx}\, \mathrm{d}s = \int _0 ^\infty \mathrm{e}^{-sx} \left( sf(s)+ \frac{1}{2}\right) \mathrm{d}s.\qquad \end{aligned}$$
(5.4)

Moreover, since \(\varPsi (x+1) = \frac{1}{x} + \varPsi (x)\), we have a variation of (5.4):

$$\begin{aligned} \log x-\varPsi (x+1)= & {} - \frac{1}{2x} +\int _0 ^\infty s f(s) \mathrm{e}^{-sx}\, \mathrm{d}s\nonumber \\= & {} \int _0 ^\infty \mathrm{e}^{-sx} \left( sf(s)- \frac{1}{2}\right) \mathrm{d}s. \end{aligned}$$
(5.5)

As easy consequences, we have, for every \(x > 0\),

$$\begin{aligned} 0< & {} x\left( \log x - \varPsi (x)\right) \le 1, \end{aligned}$$
(5.6)
$$\begin{aligned} 0< & {} \log x - \varPsi (x) - \frac{1}{2x} \le \frac{1}{12x^2}. \end{aligned}$$
(5.7)

Differentiating again, we see that for \(q\ge 1\), \({\mathfrak {Re }\,}\ x > 0\),

$$\begin{aligned} \varPsi ^{(q)}(x) = (-1)^{q-1} (q-1)! x ^{-q} + (-1)^{q-1} \int _0^\infty \mathrm{e}^{-sx} s^q \left( sf(s)+ \frac{1}{2}\right) \,\mathrm{d}s \end{aligned}$$

and then

$$\begin{aligned} |\varPsi ^{(q)}(x) - (-1)^{q-1} (q-1)! x ^{-q}| \le ({\mathfrak {Re }\,}\ x)^{-q-1} q!. \end{aligned}$$
(5.8)

Another useful formula is

$$\begin{aligned} \varPsi (z+1) = -\gamma - \sum _{k=1}^\infty \left( \frac{1}{k+z}-\frac{1}{k}\right) , \end{aligned}$$
(5.9)

for \(z \notin (-\infty , -1]\).

1.2 The Density \(g_r^{(\delta )}\) and Some Moments Related to It

Recall that for \(r >0\) and \(\delta \) such that \(r + 2{\mathfrak {Re }\,}\ \delta +1>0\), the density \(g_{r}^{(\delta )}\) on the unit disk \({\mathbb {D}}\) is given by

$$\begin{aligned} g_{r}^{(\delta )}(z) = c_{r, \delta } (1 -|z|^2)^{r -1} (1-z)^{{\bar{\delta }}} (1- \bar{z})^\delta \end{aligned}$$

where \( c_{r, \delta }\) is the normalization constant. The following lemma is the key to compute \( c_{r, \delta }\) and the moments of \(g_r^{(\delta )}\).

Lemma 3

Let \(s,t,\ell \) be complex numbers such that: \({\mathfrak {Re }\,}\ \ell \), \({\mathfrak {Re }\,}(s + \ell +1)\), \({\mathfrak {Re }\,}( t +\ell + 1 )\) and \({\mathfrak {Re }\,}( s + t +\ell + 1 )\) are strictly positive. Then, the following identity holds:

$$\begin{aligned} \int _{{\mathbb {D}}} (1 -|z|^2)^{\ell -1} (1-z)^s (1- \bar{z})^t {\mathrm {d}}^2z = \pi \varGamma \left[ \begin{matrix} \ell \ ,\ \ell +1+s+t\\ \ell +1+s\ , \ \ell +1+t \end{matrix}\right] , \end{aligned}$$
(5.10)

where for the sake of simplicity we use the polygamma symbol

$$\begin{aligned} \varGamma \left[ \begin{matrix} a, b, \ldots \\ c,d, \ldots \end{matrix}\right] := \frac{\varGamma (a)\varGamma (b)\cdots }{\varGamma (c)\varGamma (d)\cdots }. \end{aligned}$$

A proof of this result is given in [5]. A first consequence is that

$$\begin{aligned} c_{r, \delta }= \pi ^{-1} \varGamma \left[ \begin{matrix} r +1+\delta \ , \ r+1+{\bar{\delta }}\\ r\ ,\ r+1+\delta + {\bar{\delta }}\\ \end{matrix}\right] . \end{aligned}$$
(5.11)

A second consequence is that if \(\gamma \) has the density \(g_{r}^{(\delta )}\), then we have

$$\begin{aligned} {\mathbb {E}} (1- \gamma )^a (1- \bar{\gamma })^b = \varGamma \left[ \begin{matrix} r+1+\delta +{\bar{\delta }}+a+b,\,r+1+\bar{\delta },\,r+1+\delta \\ r+1+\delta +{\bar{\delta }},\,r+1+{\bar{\delta }}+ a,\,r+1+\delta +b \end{matrix}\right] \qquad \quad \end{aligned}$$
(5.12)

as soon as all the real parts of the arguments of the gamma functions are strictly positive.

Let us notice that for \(r=0\), the RHS of (5.12) is the Mellin–Fourier transform of \(1-\gamma \) when \(\gamma \in {\mathbb {T}}\) is distributed according to \(\lambda ^{(\delta )}\).

In this paper, we need the following computations, in order to deduce the moments of \(\log (1-\gamma )\). The quantities involved below are all well-defined as soon as \(s > s_0\), where \(s_0\) is some strictly negative quantity depending on r and \(\delta \), and in particular, for (st) in the neighborhood of (0, 0), one can write:

$$\begin{aligned} \varLambda (s,t):= & {} \log {\mathbb {E}} \exp \left( 2s {\mathfrak {Re }\,}\log (1-\gamma ) +2 t {\mathfrak {Im }\,}\log (1-\gamma )\right) \nonumber \\= & {} \log {\mathbb {E}} \exp \left( {\mathfrak {Re }\,}\ (2(s-{\mathrm {i}}t )\log (1-\gamma )\right) \nonumber \\= & {} \log {\mathbb {E}} (1-\gamma )^{s-{\mathrm {i}}t} (1-\bar{\gamma })^{s+{\mathrm {i}}t}\nonumber \\= & {} \ell \big (r +1+\delta +\overline{\delta }+ 2s\big )- \ell \big (r+1+ \delta +\overline{\delta }\big )\nonumber \\&-\ell \left( r +1+ \overline{\delta }+s-{\mathrm {i}}t\Big ) - \ell \Big (r +1+ \delta + s+{\mathrm {i}}t\right) \nonumber \\&+\,\ell \big (r +1 +\overline{\delta }\big )+ \ell \big (r +1 +\delta \big ). \end{aligned}$$
(5.13)

To compute moments, we need differentiation. First we have:

$$\begin{aligned} \frac{\partial }{\partial s}\varLambda (s, t)= & {} 2 \varPsi \left( r +1+\delta +\overline{\delta }+2s\right) \nonumber \\&-\varPsi \left( r +1+ \delta + s+ {\mathrm {i}}t\right) -\varPsi \left( r +1+ \overline{\delta }+s- {\mathrm {i}}t\right) \nonumber \\ \frac{\partial }{\partial t}\varLambda (s, t)= & {} {\mathrm {i}}\varPsi \Big (r +1+{\bar{\delta }}+ s -{\mathrm {i}}t\Big ) - {\mathrm {i}}\varPsi \Big (r +1+\delta +s + {\mathrm {i}}t\Big ). \end{aligned}$$
(5.14)

The first moment is then:

$$\begin{aligned} {\mathbb {E}}\,{\mathfrak {Re }\,}\log (1-\gamma )= & {} \varPsi (r+1 + \delta + {\bar{\delta }}) - \frac{1}{2}\varPsi (r+1+\delta ) -\frac{1}{2}\varPsi (r+1+{\bar{\delta }})\\ {\mathbb {E}}\,{\mathfrak {Im }\,}\log (1-\gamma )= & {} \frac{1}{2{\mathrm {i}}}\varPsi (r+1+\delta ) -\frac{1}{2{\mathrm {i}}}\varPsi (r+1+{\bar{\delta }}) \end{aligned}$$

or

$$\begin{aligned} {\mathbb {E}} \log (1-\gamma ) = \varPsi (r+1 + \delta + {\bar{\delta }}) - \varPsi (r+1+{\bar{\delta }}). \end{aligned}$$
(5.15)

Differentiating again (5.14), we get

$$\begin{aligned} \frac{\partial ^2}{\partial s^2} \varLambda (s, t)= & {} 4 \varPsi ' \big (r +1+\delta +\overline{\delta }+2s\big )\nonumber \\&- \varPsi '\Big (r +1+ \delta + s + {\mathrm {i}}t\Big )-\varPsi '\Big (r +1+ \overline{\delta }+s- {\mathrm {i}}t\Big )\nonumber \\ \frac{\partial ^2}{\partial t^2}\varLambda (s, t)= & {} \varPsi ' \Big (r +1+{\bar{\delta }}+s-{\mathrm {i}}t\Big ) +\varPsi ' \Big (r +1+\delta +s +{\mathrm {i}}t\Big )\nonumber \\ \frac{\partial ^2}{\partial s\partial t} \varLambda (s, t)= & {} -{\mathrm {i}}\varPsi '\Big (r+1+\delta +s + {\mathrm {i}}t\Big ) +{\mathrm {i}}\varPsi '\Big (r+1+{\bar{\delta }}+s-{\mathrm {i}}t\Big )\qquad \end{aligned}$$
(5.16)

and the second moments are

$$\begin{aligned}&\hbox {Var}\,{\mathfrak {Re }\,}\log (1-\gamma ) = \varPsi ' \big (r +1+\delta +\overline{\delta }\big ) -\frac{1}{4}\varPsi '\big (r +1+ \delta \big )-\frac{1}{4}\varPsi '\big (r +1+ \overline{\delta }\big )\nonumber \\&\hbox {Var}\,{\mathfrak {Im }\,}\log (1-\gamma ) = \frac{1}{4}\varPsi ' \big (r +1+\delta \big ) + \frac{1}{4}\varPsi ' \big (r +1+\overline{\delta }\big )\nonumber \\&\hbox {Cov}\,({\mathfrak {Re }\,}\log (1-\gamma ), {\mathfrak {Im }\,}\log (1-\gamma )) = \frac{1}{4{\mathrm {i}}}\varPsi '\big (r+1+\delta \big ) - \frac{1}{4{\mathrm {i}}}\varPsi '\big (r+1+{\bar{\delta }}\big ).\nonumber \\ \end{aligned}$$
(5.17)

1.3 Complex Logarithm and Characteristic Polynomial

Let \(E_k\) be the set of the complex \(k \times k\) matrices with no eigenvalue on the interval \([1, \infty )\). For \(V \in E_k\), let us define

$$\begin{aligned} \log {\text {det}} (I_k - V) := \sum _{j=1}^k \log (1 - \lambda _j), \end{aligned}$$

where the \(\lambda _j\)’s are the roots, counted with multiplicity, of the polynomial \(z \mapsto \det (z I_k -V)\), and where in the right-hand side, one considers the principal branch of the logarithm. This definition is meaningful, since by assumption, \(1 - \lambda _j \notin \mathbb {R}_-\) for all \(j \in \{1, \ldots , k\}\). By the continuity of the set of roots of a polynomial with respect to its coefficients, the set \(E_k\) is open and the function \(V \mapsto \log {\text {det}} (I_k - V)\) defined just above is continuous on \(E_k\). In fact, since \(E_k\) is connected (this is easily checked by tridiagonalizing the matrices), this is the unique way to define the logarithm of \({\text {det}} (I_k - V)\) as a continuous function of \(V \in E_k\) if we assume that it should take the value zero at \(V =0\).

Now, with the notation of the beginning of the paper, the matrix \(G_k(U_n)\) is a submatrix of the unitary matrix \(U_n\), and all its eigenvalues have modulus bounded by 1. If we assume \( \gamma _0, \ldots , \gamma _{n-1} \ne 1\) (which holds almost surely under \(\hbox {CJ}_{\beta , \delta }^{(n)}\)), then by (1.6), \(\varPhi _{k,n}(1) \ne 0\), and one easily deduces that \(G_k(U_n) \in E_k\), which allows to define \(\log \varPhi _{k,n}(1)\) without ambiguity. Now, the map from \(\mathbb {D}^{n-1} \times (\mathbb {U} \backslash \{1\})\) to \(\mathbb {R}\), given by

$$\begin{aligned} (\gamma _1, \ldots , \gamma _{n-1}) \mapsto \sum _{j=0}^{k-1} \log (1 -\gamma _j) \end{aligned}$$

is continuous if we take the principal branch of the logarithm, and since \(U_n\) depends continuously on \((\gamma _1, \ldots , \gamma _{n-1}) \in \mathbb {D}^{n-1} \times (\mathbb {U} \backslash \{1\})\), as it can be checked in [5], the map

$$\begin{aligned} (\gamma _1, \ldots , \gamma _{n-1}) \mapsto \log \varPhi _{k,n}(1) \end{aligned}$$

is also continuous. These two maps have the same exponential, and one can check that they are both real if the \(\gamma _j\)’s are all real. Hence, they are equal, which fully justifies the equation

$$\begin{aligned} \log \varPhi _{k,n} (1) = \sum _{j=0}^{k-1} \log (1-\gamma _j). \end{aligned}$$
(5.18)

1.4 Abel–Plana Summation Formula

Theorem 9

Let \(m < n\) be integers and let g be a holomorphic function on the strip \(\{t \in \mathbb {C}, n \le {\mathfrak {Re }\,}\ t \le m\}\) (i.e., g is continuous on this strip and holomorphic in its interior). We assume that \(g(t) = o\left( \exp (2\pi |{\mathfrak {Im }\,}\ t|)\right) \) as \({\mathfrak {Im }\,}\ t \rightarrow \pm \infty \), uniformly with respect to \({\mathfrak {Re }\,}\ t \in [n,m]\). Then,

$$\begin{aligned} \sum _{j= m+1}^n g(j)= & {} \int _m^n g(t)\mathrm{d}t + \frac{g(n) - g(m)}{2}\nonumber \\&+\,\, {\mathrm {i}}\int _0^\infty \frac{g(m+{\mathrm {i}}y) - g(n+ {\mathrm {i}}y) - g(m -{\mathrm {i}}y) + g(n-{\mathrm {i}}y)}{\mathrm{e}^{2\pi y}-1} \mathrm{d}y.\nonumber \\ \end{aligned}$$
(5.19)

For a proof see [20, p. 290].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Najnudel, J., Nikeghbali, A. & Rouault, A. Limit Theorems for Orthogonal Polynomials Related to Circular Ensembles. J Theor Probab 29, 1199–1239 (2016). https://doi.org/10.1007/s10959-015-0632-x

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10959-015-0632-x

Keywords

Mathematics Subject Classification (2010)

Navigation