Skip to main content
Log in

On the Moments of the Moments of the Characteristic Polynomials of Random Unitary Matrices

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

Abstract

Denoting by \(P_N(A,\theta )=\det (I-Ae^{-i\theta })\) the characteristic polynomial on the unit circle in the complex plane of an \(N\times N\) random unitary matrix A, we calculate the kth moment, defined with respect to an average over \(A\in U(N)\), of the random variable corresponding to the \(2\beta \)th moment of \(P_N(A,\theta )\) with respect to the uniform measure \(\frac{d\theta }{2\pi }\), for all \(k, \beta \in \mathbb {N}\) . These moments of moments have played an important role in recent investigations of the extreme value statistics of characteristic polynomials and their connections with log-correlated Gaussian fields. Our approach is based on a new combinatorial representation of the moments using the theory of symmetric functions, and an analysis of a second representation in terms of multiple contour integrals. Our main result is that the moments of moments are polynomials in N of degree \(k^2\beta ^2-k+1\). This resolves a conjecture of Fyodorov and Keating (Philos Trans R Soc A 372(2007):20120503, 2014) concerning the scaling of the moments with N as \(N\rightarrow \infty \), for \(k,\beta \in \mathbb {N}\). Indeed, it goes further in that we give a method for computing these polynomials explicitly and obtain a general formula for the leading coefficient.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Similar content being viewed by others

Notes

  1. By \(A(N)\sim B(N)\), we mean that \(A(N)/B(N)\rightarrow 1\) when \(N\rightarrow \infty \).

  2. Henceforth, whenever we write ‘up to terms of order 1 / N’ followed by a statement of the form \(A(N)\sim B(N)\) we mean that \(A(N)=B(N)(1+O(1/N))\) as \(N\rightarrow \infty \).

References

  1. Arguin, L.-P., Belius, D., Bourgade, P.: Maximum of the characteristic polynomial of random unitary matrices. Commun. Math. Phys. 349(1), 703–751 (2017)

    Article  ADS  MathSciNet  Google Scholar 

  2. Arguin, L.-P., Belius, D., Bourgade, P., Radziwiłł, M., Soundararajan, K.: Maximum of the Riemann zeta function on a short interval of the critical line. Commun. Pure Appl. Math. 72, 500–535 (2019)

    Article  MathSciNet  Google Scholar 

  3. Baker, T.H., Forrester, P.J.: Finite-\(N\) fluctuation formulas for random matrices. J. Stat. Phys. 88, 1371–1386 (1997)

    Article  ADS  MathSciNet  Google Scholar 

  4. Bailey, E.C., Bettin, S., Blower, G., Conrey, J.B., Prokhorov, A., Rubinstein, M.O., Snaith, N.C.: Mixed moments of characteristic polynomials of random unitary matrices. arXiv:1901.07479

  5. Basor, E., Bleher, P., Buckingham, R., Grava, T., Its, A., Its, E., Keating, J.P.: A representation of joint moments of CUE characteristic polynomials in terms of Painlevé functions. arXiv:1811.00064 (Accepted for publication in Nonlinearity)

  6. Berestycki, N.: An elementary approach to Gaussian multiplicative chaos. Electron. Commun. Probab.22, paper no. 27 (2017)

  7. Bump, D., Gamburd, A.: On the averages of characteristic polynomials from classical groups. Commun. Math. Phys. 265(1), 227–274 (2006)

    Article  ADS  MathSciNet  Google Scholar 

  8. Carpentier, D., Le Doussal, P.: Glass transition of a particle in a random potential, front selection in nonlinear renormalization group, and entropic phenomena in Liouville and sinh-Gordon models. Phys. Rev. E 63, 026110 (2001)

    Article  ADS  Google Scholar 

  9. Chhaibi, R., Madaule, T., Najnudel, J.: On the maximum of the C\(\beta \)E field. Duke Math. J. 167, 2243–2345 (2018)

    Article  MathSciNet  Google Scholar 

  10. Claeys, T., Krasovsky, I.: Toeplitz determinants with merging singularities. Duke Math. J. 164(15), 2897–2987 (2015)

    Article  MathSciNet  Google Scholar 

  11. Conrey, J.B., Farmer, D.W., Keating, J.P., Rubinstein, M.O., Snaith, N.C.: Autocorrelation of random matrix polynomials. Commun. Math. Phys. 237(3), 365–395 (2003)

    Article  ADS  MathSciNet  Google Scholar 

  12. Conrey, J.B., Farmer, D.W., Keating, J.P., Rubinstein, M.O., Snaith, N.C.: Integral moments of L-functions. Proc. Lond. Math. Soc. 3(91), 33–104 (2005)

    Article  MathSciNet  Google Scholar 

  13. Conrey, J.B., Farmer, D.W., Zirnbauer, M.R.: Autocorrelation of ratios of \(L\)-functions. Commun. Number Theory Phys. 2, 593–636 (2018)

    Article  MathSciNet  Google Scholar 

  14. Conrey, J.B., Snaith, N.C.: Correlations of eigenvalues and Riemann zeros. Commun. Number Theory Phys. 2, 477–536 (2008)

    Article  MathSciNet  Google Scholar 

  15. Derrida, B.: Random energy model: an exactly solvable model of disordered systems. Phys. Rev. B 24, 2613–2626 (1981)

    Article  ADS  MathSciNet  Google Scholar 

  16. Diaconis, P., Shahshahani, M.: On the eigenvalues of random matrices. J. Appl. Probab. A 31, 49–62 (1994)

    Article  MathSciNet  Google Scholar 

  17. Ding, J., Roy, R., Zeitouni, O.: Convergence of the centered maximum of log-correlated Gaussian fields. Ann. Probab. 45, 3886–3928 (2017)

    Article  MathSciNet  Google Scholar 

  18. Forrester, P.J., Frankel, N.E.: Applications and generalizations of Fisher-Hartwig asymptotics. J. Math. Phys. 45, 2003–2028 (1994)

    Article  ADS  MathSciNet  Google Scholar 

  19. Forrester, P.J., Keating, J.P.: Singularity dominated strong fluctuations for some random matrix averages. Commun. Math. Phys. 250, 119–131 (2004)

    Article  ADS  MathSciNet  Google Scholar 

  20. Fyodorov, Y.V., Bouchaud, J.-P.: Freezing and extreme-value statistics in a random energy model with logarithmically correlated potential. J. Phys. A Math. Theor. 41, 372001 (2008)

    Article  MathSciNet  Google Scholar 

  21. Fyodorov, Y.V., Gnutzmann, S., Keating, J.P.: Extreme values of CUE characteristic polynomials: a numerical study. J. Phys. A 51, 464001 (2018)

    Article  ADS  MathSciNet  Google Scholar 

  22. Fyodorov, Y.V., Hiary, G.A., Keating, J.P.: Freezing transition, characteristic polynomials of random matrices, and the Riemann zeta-function. Phys. Rev. Lett. 108, 170601 (2012)

    Article  ADS  Google Scholar 

  23. Fyodorov, Y.V., Keating, J.P.: Freezing transitions and extreme values: random matrix theory, and disordered landscapes. Philos. Trans. R. Soc. A 372(2007), 20120503 (2014)

    Article  ADS  MathSciNet  Google Scholar 

  24. Fyodorov, Y.V., Khoruzhenko, B.A., Simm, N.J.: Fractional Brownian motion with Hurst index \(H = 0\) and the Gaussian unitary ensemble. Ann. Probab. 44, 2980–3031 (2016)

    Article  MathSciNet  Google Scholar 

  25. Fyodorov, Y.V., Simm, N.J.: On the distribution of the maximum value of the characteristic polynomial of GUE random matrices. Nonlinearity 29, 2837–2855 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  26. Fyodorov, Y.V., Le Doussal, P.: Moments of the position of the maximum for GUE characteristic polynomials and for log-correlated Gaussian processes. J. Stat. Phys. 164, 190–240 (2016)

    Article  ADS  MathSciNet  Google Scholar 

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

    Article  ADS  MathSciNet  Google Scholar 

  28. Kahane, J.-P.: Sur le chaos multiplicatif. Ann. Sci. Math. Québec 9(2), 105–150 (1985)

    MathSciNet  MATH  Google Scholar 

  29. Keating, J.P.: Random matrices and number theory: some recent themes. In: G. Schehr et al. (eds) Stochastic Processes and Random Matrices: Lecture Notes of Les Houches Summer School: July 2015. Session CIV. Oxford University Press (2017)

  30. Keating, J.P., Odgers, B.E.: Symmetry transitions in random matrix theory and \(L\)-functions. Commun. Math. Phys. 281, 499–528 (2008)

    Article  ADS  MathSciNet  Google Scholar 

  31. Keating, J.P., Rodgers, B., Roditty-Gershon, E., Rudnick, Z.: Sums of divisor functions in \({F}_q[t]\) and matrix integrals. Math. Z. 288, 167–198 (2018)

    Article  MathSciNet  Google Scholar 

  32. Keating, J.P., Scott, E.: Undergraduate summer research project. University of Bristol (2015) (unpublished)

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

    Article  ADS  MathSciNet  Google Scholar 

  34. Kistler, N.: Derrida’s random energy models. From spin glasses to the extremes of correlated random fields. In: Gayrard, V., Kistler, N. (eds.) Correlated Random Systems: Five Different Methods. Springer Lecture Notes in Mathematics, vol. 2143. Springer, Berlin (2015)

    Google Scholar 

  35. Kupiainen, A., Rhodes, R., Vargas, V.: Integrability of Liouville theory: proof of the DOZZ Formula. arXiv:1707.08785 (2017)

  36. Macdonald, I.G.: Symmetric Functions and Hall Polynomials. Oxford University Press, Oxford (1998)

    MATH  Google Scholar 

  37. Najnudel, J.: On the extreme values of the Riemann zeta function on random intervals of the critical line. Probab. Theory Relat. Fields. 172, 387–452 (2018)

    Article  MathSciNet  Google Scholar 

  38. Nikula, M., Saksman, E., Webb, C.: Multiplicative chaos and the characteristic polynomial of the CUE: the \(L^1\)-phase arXiv:1806.0183

  39. Paquette, E., Zeitouni, O.: The maximum of the CUE field. Int. Math. Res. Not. 16, 5028–5119 (2017)

    MATH  Google Scholar 

  40. Remy, G.: The Fyodorov–Bouchaud formula and Liouville conformal field theory. arXiv:1710.06897 (2017)

  41. Rhodes, R., Vargas, V.: Lecture notes on Gaussian multiplicative chaos and Liouville quantum gravity. In: Schehr, G. et al. (ed.) Stochastic Processes and Random Matrices: Lecture Notes of Les Houches Summer School: July 2015. Session CIV. Oxford University Press (2017)

  42. Saksman, E., Webb, C.: The Riemann zeta function and Gaussian multiplicative chaos: statistics on the critical line. arXiv:1609.00027

  43. Stanley, R.P.: Theory and application of plane partitions. Part 2. Stud. Appl. Math. 50, 259–279 (1971)

    Article  Google Scholar 

  44. Stanley, R.P.: Enumerative Combinatorics, vol. 2. Cambridge University Press, Cambridge (1999)

    Book  Google Scholar 

  45. Subag, E., Zeitouni, O.: Freezing and decorated Poisson point processes. Commun. Math. Phys. 337, 55–92 (2015)

    Article  ADS  MathSciNet  Google Scholar 

  46. Webb, C.: The characteristic polynomial of a random unitary matrix and Gaussian multiplicative chaos—the \(L^2\)-phase. Electron. J. Probab., 20(104) 1–21 (2015)

Download references

Acknowledgements

We thank Edva Roditty-Gershon and Scott Harper for helpful discussions, and Euan Scott for contributing to preliminary computations of the moments using the method outlined in the Appendix. ECB is grateful to the Heilbronn Institute for Mathematical Research for support. JPK is pleased to acknowledge support from a Royal Society Wolfson Research Merit Award and ERC Advanced Grant 740900 (LogCorRM). We are most grateful to the referees for their careful reading of the manuscript and for a number of helpful questions and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to J. P. Keating.

Additional information

Communicated by L. Erdos

Publisher's Note

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

Appendix

Appendix

1.1 Examples

Here we give explicit examples of the polynomials \({{\,\mathrm{MoM}\,}}_N(k,\beta )\) for small values of \(k,\beta \). The formulae we record extend the results of preliminary calculations due to Keating and Scott [32] (c.f. [29]), which formed the basis for some of the numerical computations in [21].

First, the general technique is described and then explicit forms of the polynomials are given in the cases of \(\beta =1\), \(k\in \{1,2,3,4\}\) and \(\beta =2\), \(k\in \{1,2\}\). We should remark that the moment formula of Keating and Snaith [33] gives the full polynomials for the case \(k=1\), \(\beta \in \mathbb {N}\); see (10).

The technique we use is in a slightly more general form than is needed here, because we see it as having other potential applications; we then specialise back to the actual formula required for our calculations. The more general form we start with was first derived by Conrey, Farmer and Zirnbauer [13], and later by Bump and Gamburd using symmetric function theory [7]. Note that we used a special case of this result to prove Theorem 1.3 in Sect. 4. First, define for finite sets ABCD,

$$\begin{aligned} R(A,B;C,D)\,{:=}\,\int _{U(N)}\frac{\prod _{\alpha \in A}\det (I-X^*e^{-\alpha })\prod _{\beta \in B}\det (I-Xe^{-\beta })}{\prod _{\gamma \in C}\det (I-X^*e^{-\gamma })\prod _{\delta \in D}\det (I-Xe^{-\delta })}dX. \end{aligned}$$

Further if

$$\begin{aligned} Z(A,B)\,{:=}\, \prod _{\begin{array}{c} \alpha \in A,\\ \beta \in B \end{array}}\frac{1}{(1-e^{-(\alpha +\beta )})}, \end{aligned}$$

then define

$$\begin{aligned} Z(A,B;C,D)\,{:=}\,\frac{Z(A,B)Z(C,D)}{Z(A,D)Z(B,C)}. \end{aligned}$$

Finally, if \(S\subset A\) and \(T\subset B\) then \(\overline{S}=A-S\), \(\overline{T}=B-T\), \(S^-=\{-\hat{\alpha }:\hat{\alpha }\in S\}\) and similarly for T. Note that here we are using the notation \(U+V\), \(U-V\) (to be interpreted as \(U\cup V\) and \(U\backslash V\) respectively for sets UV) to be consistent with the statement of the Theorem in [14].

Theorem

([11]). With \(N\ge 0\) and \({{\,\mathrm{Re}\,}}(\gamma )>0, {{\,\mathrm{Re}\,}}(\delta )>0\) for \(\gamma \in C\), \(\delta \in D\), \(|C|\le |A|+N\), \(|D|\le |B|+N\), we have

$$\begin{aligned} R(A,B;C,D)=\sum _{\begin{array}{c} S\subset A,T\subset B\\ |S|=|T| \end{array}}e^{-N(\sum _{\hat{\alpha }\in S}\hat{\alpha }+\sum _{\hat{\beta }\in T}\hat{\beta })}Z(\overline{S}+T^-,\overline{T}+S^-;C,D), \end{aligned}$$

where \(A=S+\overline{S}\) and \(B=T+\overline{T}\).

To see how this is used to give the full polynomials for \({{\,\mathrm{MoM}\,}}_N(k,\beta )\), we outline the simplest case with \(k=\beta =1\). We note that

$$\begin{aligned} {{\,\mathrm{MoM}\,}}_N(1,1)&=\frac{1}{2\pi }\int _0^{2\pi }\mathbb {E}_{A\in U(N)}\left( |P_N(A,\theta )|^2\right) d\theta \\&=\frac{1}{2\pi }\int _0^{2\pi }\int _{U(N)}P_N(A,\theta )P_N(A^*,-\theta )dAd\theta , \end{aligned}$$

so we apply the above theorem with \(A=\{i\alpha \}, B=\{i\beta \}\), \(C,D=\emptyset \). This gives us that

$$\begin{aligned} {{\,\mathrm{MoM}\,}}_N(1,1)&=\frac{1}{2\pi }\int _0^{2\pi }\lim _{\beta \rightarrow -\alpha }Z(A,B)+e^{-iN(\alpha +\beta )}Z(B^-,A^-)d\alpha \\&=\frac{1}{2\pi }\int _0^{2\pi }\lim _{\beta \rightarrow -\alpha }\sum _{m=0}^Ne^{-im(\alpha +\beta )}d\alpha \\&=N+1. \end{aligned}$$

Higher values of \(k, \beta \) clearly result in bigger sets AB, and hence many more choices for ST. Nevertheless, small cases of \({{\,\mathrm{MoM}\,}}_N(k,\beta )\) can be computed in the same way. For example

$$\begin{aligned} {{\,\mathrm{MoM}\,}}_N(1,1)&=N+1\\ {{\,\mathrm{MoM}\,}}_N(2,1)&=\frac{1}{6}(N+3)(N+2)(N+1)\\ {{\,\mathrm{MoM}\,}}_N(3,1)&=\frac{1}{2520}(N+5)(N+4)(N+3)(N+2)(N+1)(N^2+6N+21)\\ {{\,\mathrm{MoM}\,}}_N(4,1)&=\frac{1}{778377600}(N+7)(N+6)(N+5)(N+4)(N+3)(N+2)(N+1)\\&\quad \times (7N^6+168N^5+1804N^4+10944N^3+41893N^2+99624N+154440)\\ {{\,\mathrm{MoM}\,}}_N(1,2)&=\frac{1}{12}(N+1)(N+2)^2(N+3)\\ {{\,\mathrm{MoM}\,}}_N(2,2)&=\frac{1}{163459296000}(N+7)(N+6)(N+5)(N+4)(N+3)(N+2)(N+1)\\&\quad \times (298N^8+9536N^7+134071N^6+1081640N^5+5494237 N^4+18102224N^3\\&\quad +38466354N^2+50225040N+32432400). \end{aligned}$$

It is worth noting explicitly that this method gives exact information about the moments of the moments at the freezing transition \(\beta =1\).

1.2 Vandermonde determinant coefficients

Recall that we are interested in determining the coefficient of terms of the form \((x_1\cdots x_n)^{n-1}\) in the square of the Vandermonde determinant,

$$\begin{aligned} \Delta (x_1,\dots ,x_n)^2=\sum _{\sigma ,\tau \in S_n}{{\,\mathrm{sgn}\,}}(\sigma ){{\,\mathrm{sgn}\,}}(\tau )\prod _{i=1}^nx_i^{\sigma (i)+\tau (i)-2}. \end{aligned}$$
(121)

Thus, we require that \(\sigma (i)+\tau (i)=n+1\) for all \(i\in \{1,\dots ,n\}\) and in particular we want to show that this coefficient is strictly positive.

Immediately, we see that there will be n! terms of the required form since fixing \(\sigma (i)\) completely determines \(\tau (i)\). Consider the bijection

$$\begin{aligned} \phi :\{1,\dots ,n\}\rightarrow \{1,\dots ,n\}, \quad i\mapsto n+1-i. \end{aligned}$$

The order of \(\phi \) is 2 and if n is even, it has no fixed point, whereas if n is odd there is a unique fixed point \((n+1)/2\). Thus, \(\phi \in S_n\) and it consists of n / 2 transpositions if n is even, and \((n-1)/2\) transpositions if n is odd. Now set \(\tau =\phi \circ \sigma \), so \(\tau \in S_n\), and \(\tau (i)=n+1-\sigma (i)\). Given \(\sigma \), we have found our unique permutation. To determine the sign of \(\tau \), note that \({{\,\mathrm{sgn}\,}}(\tau )={{\,\mathrm{sgn}\,}}(\phi ){{\,\mathrm{sgn}\,}}(\sigma )\), and

$$\begin{aligned} {{\,\mathrm{sgn}\,}}(\phi )=(-1)^{\left\lfloor \tfrac{n}{2}\right\rfloor }={\left\{ \begin{array}{ll}+1&{}\text {if }n\equiv 0,1\mod 4\\ -1&{}\text {if }n\equiv 2,3\mod 4.\end{array}\right. } \end{aligned}$$

Thus, the coefficient of \((x_1\cdots x_n)^{n-1}\) in \(\Delta (x_1,\dots ,x_n)^2\) is \({{\,\mathrm{sgn}\,}}(\phi )n!\). It now follows that the coefficient of

$$\begin{aligned} \prod _{i=1}^{l_1}v_i^{l_1-1}\prod _{i=l_1+1}^{l_1+l_2}v_i^{l_2-1}\cdots \prod _{i=\sum _{j=1}^{k-1}l_j+1}^{k\beta }v_i^{l_k-1}\prod _{i=k\beta +1}^{\sum _{j=1}^{k-1}l_j+2\beta }v_i^{2\beta -l_k-1}\cdots \prod _{i=2(k-1)\beta +1+l_1}^{2k\beta }v_i^{2\beta -l_1-1} \end{aligned}$$

in

$$\begin{aligned} \prod _{n=1}^{k}\Delta (v_{\sum _{j=1}^{n-1}l_j+1},\dots ,v_{\sum _{j=1}^{n}l_j})^2 \prod _{n=1}^{k}\Delta (v_{\sum _{j=1}^{n}l_j+2(k-n)\beta +1},\dots ,v_{\sum _{j=1}^{n-1}l_j+2(k-(n-1))\beta })^2 \end{aligned}$$

is given by

$$\begin{aligned} (-1)^{\sum _{j=1}^k\left( \big \lfloor \tfrac{l_j}{2}\big \rfloor +\big \lfloor \tfrac{2\beta -l_j}{2}\big \rfloor \right) }\prod _{j=1}^kl_j!(2\beta -l_j)!&=(-1)^{k\beta +\sum _{j=1}^k\left( \big \lfloor \tfrac{l_j}{2}\big \rfloor +\big \lfloor \tfrac{-l_j}{2}\big \rfloor \right) }\prod _{j=1}^kl_j!(2\beta -l_j)!\\&=(-1)^{k\beta }(-1)^{\sum _{j=1}^k\delta _{\{l_j\text { is odd\}}}}\prod _{j=1}^kl_j!(2\beta -l_j)!\\&=(-1)^{k\beta }(-1)^{\#\{j:l_j\text { is odd}\}}\prod _{j=1}^kl_j!(2\beta -l_j)!. \end{aligned}$$

This proves the result since the parity of \(\#\{j:l_j\text { is odd}\}\) is the same as the parity of \(k\beta \) as \(\sum _{j=1}^kl_j= k\beta \).

1.3 Uniqueness of construction

When trying to construct the term of the form \((v_1\cdots v_{2k\beta })^{2\beta -1}\) in

$$\begin{aligned}&\prod _{q=1}^{k}\Delta (v_{\sum _{j=1}^{q-1}l_j+1},\dots ,v_{\sum _{j=1}^{q}l_j})^2 \prod _{q=1}^{k}\Delta (v_{\sum _{j=1}^{q}l_j+2(k-q)\beta +1},\dots ,v_{\sum _{j=1}^{q-1}l_j+2(k-(q-1))\beta })^2\\&\prod _{\begin{array}{c} m\le k\beta <n\\ \alpha _m=\alpha _n \end{array}}\left( {v_n-v_m}\right) , \end{aligned}$$

first note that the variables \(v_m\), for \(m\in \{1,\dots ,k\beta \}\) only appear in the Vandermonde determinants and the products

$$\begin{aligned} \prod _{\begin{array}{c} m\le k\beta <n\\ \alpha _m=\alpha _n \end{array}}\left( {v_n-v_m}\right) =\prod _{\begin{array}{c} m\in \{1,\dots ,l_1\}\\ n\in \{2(k-1)\beta +1+l_1,\dots ,2k\beta \} \end{array}}(v_n-v_m)\cdots \prod _{\begin{array}{c} m\in \{\sum _{j=1}^{k-1}l_j+1,\dots ,k\beta \}\\ n\in \{k\beta +1,\dots ,\sum _{j=1}^{k-1}l_j+2\beta \} \end{array}}(v_n-v_m). \end{aligned}$$

In particular, after fixing \(q\in \{1,\dots ,k\}\) take \(v_j\) with \(j\in \{\sum _{i=1}^{q-1}l_i+1,\dots ,\sum _{i=1}^ql_i\}\). Then \(v_j\) only appears in the following two terms:

$$\begin{aligned} \Delta (v_{\sum _{i=1}^{q-1}l_i+1},\dots ,v_{\sum _{i=1}^ql_i})^2\text { and }\prod _{\begin{array}{c} m\in \{\sum _{i=1}^{q-1}l_i+1,\dots ,\sum _{i=1}^ql_i\}\\ n\in \{2k\beta -\sum _{i=1}^{q}(2\beta -l_i)+1,\dots ,2k\beta -\sum _{i=1}^{q-1}(2\beta -l_i)\} \end{array}}(v_n-v_m). \end{aligned}$$

In particular these are both homogeneous polynomials: the former of degree \(l_q(l_q-1)\) in \(l_q\) variables and the latter is of degree \(l_q(2\beta -l_q)\) in \(2\beta \) variables. We will show that the only way to construct a term of the form \((v_1\cdots v_{2k\beta })^{2\beta -1}\) is as described following Eq. (106). Without loss of generality, we will set \(q=1\) and assume \(l_1\ge 2\). From the above discussion, the square of the Vandermonde determinant consists of terms of the form

$$\begin{aligned} v_1^{a_1}\cdots v_{l_1}^{a_{l_1}},\text { with }\sum _{i=1}^{l_1}a_i=l_1(l_1-1). \end{aligned}$$
(122)

Similarly, the product term is built of elements of the form

$$\begin{aligned} v_1^{b_1}\cdots v_{l_1}^{b_{l_1}}v_{2(k-1)\beta +1+l_1}^{b_{l_1+1}}\cdots v_{2k\beta }^{b_{2\beta }},\text { with }\sum _{i=1}^{2\beta }b_i=l_1(2\beta -l_1),\ 0\le b_i\le 2\beta -l_1.\nonumber \\ \end{aligned}$$
(123)

Hence, each term of

$$\begin{aligned} \Delta (v_1,\dots ,v_{l_1})^2\prod _{\begin{array}{c} m\in \{1,\dots ,l_1\}\\ n\in \{2(k-1)\beta +1+l_1,\dots ,2k\beta \} \end{array}}(v_n-v_m) \end{aligned}$$

is of the form

$$\begin{aligned} v_1^{a_1+b_1}\cdots v_{l_1}^{a_{l_1}+b_{l_1}}v_{2(k-1)\beta +1+l_1}^{b_{l_1+1}}\cdots v_{2k\beta }^{b_{2\beta }}, \end{aligned}$$

with \(a_i,b_i\) satisfying the homogenous conditions. To reach our goal, we need to find all possibilities for \(a_i, 1\le i\le l_1\) and \(b_i, 1\le i\le 2\beta \) that \(a_i+b_i=2\beta -1\) for \(i\in \{1,\dots ,l_1\}\). This implies that we need \(\sum _{i=1}^{l_1}(a_i+b_i)=l_1(2\beta -1)\). Now note that the ‘homogeneous conditions’ in Eqs. (122) and (123) together mean that

$$\begin{aligned} \sum _{i=1}^{l_1}(a_i+b_i)+\sum _{l_1+1}^{2\beta }b_i=\sum _{i=1}^{l_1}a_i+\sum _{i=1}^{2\beta }b_i=l_1(2\beta -1). \end{aligned}$$

Thus, we must set \(b_{l_1+1},\dots ,b_{2\beta }=0\) if we want to construct the required term. This leaves us with finding all \(a_i,b_i\)\(1\le i\le l_1\) such that all the following are satisfied,

$$\begin{aligned} a_i+b_i&=2\beta -1,\\ \sum _{i=1}^{l_1}a_i&=l_1(l_1-1),\\ \sum _{i=1}^{l_1}b_i&=l_1(2\beta -l_1),\\ 0\le&b_i\le 2\beta -l_1. \end{aligned}$$

However, the latter two conditions imply that we must have \(b_i=2\beta -l_1\) for all \(1\le i\le l_1\) which in turn gives us that \(a_i=l_1-1\) for all \(1\le i\le l_1\), and these are the only possible choices. This exactly matches the construction described following Eq. (106). The case for \(q\in \{2,\dots ,k\}\) follows similarly.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bailey, E.C., Keating, J.P. On the Moments of the Moments of the Characteristic Polynomials of Random Unitary Matrices. Commun. Math. Phys. 371, 689–726 (2019). https://doi.org/10.1007/s00220-019-03503-7

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-019-03503-7

Navigation