Skip to main content
Log in

The Interpolating Airy Kernels for the \(\beta =1\) and \(\beta =4\) Elliptic Ginibre Ensembles

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

Abstract

We consider two families of non-Hermitian Gaussian random matrices, namely the elliptic Ginibre ensembles of asymmetric \(N\)-by-\(N\) matrices with Dyson index \(\beta =1\) (real elements) and with \(\beta =4\) (quaternion-real elements). Both ensembles have already been solved for finite \(N\) using the method of skew-orthogonal polynomials, given for these particular ensembles in terms of Hermite polynomials in the complex plane. In this paper we investigate the microscopic weakly non-Hermitian large-\(N\) limit of each ensemble in the vicinity of the largest or smallest real eigenvalue. Specifically, we derive the limiting matrix-kernels for each case, from which all the eigenvalue correlation functions can be determined. We call these new kernels the “interpolating” Airy kernels, since we can recover—as opposing limiting cases—not only the well-known Airy kernels for the Hermitian ensembles, but also the complementary error function and Poisson kernels for the maximally non-Hermitian ensembles at the edge of the spectrum. Together with the known interpolating Airy kernel for \(\beta =2\), which we rederive here as well, this completes the analysis of all three elliptic Ginibre ensembles in the microscopic scaling limit at the spectral edge.

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

Similar content being viewed by others

Notes

  1. The Bessel kernels arise from the Gaussian chiral ensembles.

  2. This can be demonstrated for the other two ensembles as well, although rigorous proofs are beyond the scope of this paper.

  3. In [4] the \(\beta =4\) case was cast into the same form as for \(\beta =1\).

  4. Note that, in contrast to the present paper, the Hermite polynomials orthogonal with respect to the weight \(\exp [-x^2/2]\) were used in [22].

  5. But see [39] for an alternative approach to this issue.

  6. Intuitively, the scaling parameter \(w\) here can be considered roughly equivalent to \(N^{1/3}\), where \(N\) is the matrix size.

References

  1. Fyodorov, Y.V., Sommers, H.-J.: Random Matrices close to Hermitian or unitary: overview of methods and results. J. Phys. A 36, 3303–3347 (2003). http://arXiv.org/abs/nlin/0207051

  2. Zabrodin, A.: Random matrices and Laplacian growth. In: Akemann, G., Baik, J., Di Francesco, P. (eds.) Chapter 39 of The Oxford Handbook of Random Matrix Theory, pp. 802–823. Oxford University Press (2011). http://arXiv.org/abs/0907.4929/math-ph

  3. Akemann, G.: Matrix models and QCD with chemical potential. Int. J. Mod. Phys. A 22, 1077–1122 (2007). http://arXiv.org/abs/hep-th/0701175

    Google Scholar 

  4. Khoruzhenko, B.A., Sommers, H.-J.: Non-Hermitian ensembles. In: Akemann, G., Baik, J., Di Francesco, P. (eds.) Chapter 18 of The Oxford Handbook of Random Matrix Theory, pp. 376–397. Oxford University Press (2011). http://arXiv.org/abs/0911.5645/math-ph

  5. Fyodorov, Y.V., Khoruzhenko, B.A., Sommers, H.-J.: Almost-Hermitian random matrices: eigenvalue density in the complex plane. Phys. Lett. A 226, 46–52 (1997). http://arXiv.org/abs/cond-mat/9606173

    Google Scholar 

  6. Fyodorov, Y.V., Khoruzhenko, B.A., Sommers, H.-J.: Almost-Hermitian random matrices: crossover from Wigner-Dyson to Ginibre eigenvalue statistics. Phys. Rev. Lett. 79, 557–560 (1997). http://arXiv.org/abs/cond-mat/9703152

    Google Scholar 

  7. Efetov, K.B.: Directed quantum chaos. Phys. Rev. Lett. 79, 491–494 (1997). http://arXiv.org/abs/cond-mat/9702091/cond-mat.dis-nn

    Google Scholar 

  8. Efetov, K.B.: Quantum disordered systems with a direction. Phys. Rev. B 56, 9630–9648 (1997). http://arXiv.org/abs/cond-mat/9706055

    Google Scholar 

  9. Akemann, G., Phillips, M.J.: Universality conjecture for all Airy, sine and Bessel kernels in the complex plane, to appear. In: Deift, P., Forrester, P. (eds.) Mathematical Sciences Research Institute Publications. Cambridge University Press, http://arXiv.org/abs/1204.2740/math-ph

  10. Bender, M.: Edge scaling limits for a family of non-Hermitian random matrix ensembles. Probab. Theory Relat. Fields 147, 241–271 (2010). http://arXiv.org/abs/0808.2608v1/math.PR

  11. Akemann, G., Bender, M.: Interpolation between Airy and Poisson statistics for unitary chiral non-Hermitian random matrix ensembles. J. Math. Phys. 51, 103524–1– 21 (2010). http://arXiv.org/abs/1003.4222/math-ph

  12. Forrester, P.J.: The spectrum edge of random matrix ensembles, Nucl. Phys. B 402 [FS], 709–728 (1993)

  13. Tracy, C.A., Widom, H.: Level-spacing distributions and the Airy kernel. Phys. Lett. B 305, 115–118 (1993)

    Article  ADS  MathSciNet  Google Scholar 

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

    Article  ADS  MATH  MathSciNet  Google Scholar 

  15. Rider, B.: A limit theorem at the edge of a non-Hermitian random matrix ensemble. J. Phys. A 36, 3401–3409 (2003)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  16. Rider, B.: Order statistics and Ginibre’s ensembles. J. Stat. Phys. 114, 1139–1148 (2004)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  17. Rider, B., Sinclair, C.D.: Extremal laws for the real Ginibre ensemble, Preprint (2012). http://arXiv.org/abs/1209.6085/math-ph

  18. Johansson, K.: From Gumbel to Tracy-Widom. Probab. Theory Relat. Fields 138, 75–112 (2007). http://arXiv.org/abs/math/0510181

  19. Ameur, Y., Hedenmalm, H., Makarov, N.: Fluctuations of eigenvalues of random normal matrices. Duke Math. J. 159, 31–81 (2011). http://arXiv.org/abs/0807.0375v3/math.PR

  20. Berman, R.J.: Determinantal point processes and fermions on complex manifolds: Bulk universality, Preprint (2008). http://arXiv.org/abs/0811.3341v1/math.CV

  21. Tao, T., Vu, V.: Random matrices: universality of local spectral statistics of non-Hermitian matrices, Preprint (2012). http://arXiv.org/abs/1206.1893//math.PR

  22. Fyodorov, Y.V., Khoruzhenko, B.A., Sommers, H.-J.: Universality in the random matrix spectra in the regime of weak non-Hermiticity. Ann. Inst. Henri Poincaré 68, 449–489 (1998). http://arXiv.org/abs/chao-dyn/9802025

  23. Akemann, G.: Microscopic universality of complex matrix model correlation functions at weak non-Hermiticity. Phys. Lett. B 547, 100–108 (2002). http://arXiv.org/abs/hep-th/0206086

  24. Akemann, G., Damgaard, P.H., Osborn, J.C., Splittorff, K.: A new chiral two-matrix theory for Dirac spectra with imaginary chemical potential. Nucl. Phys. B 766, 34–67 (2007) Erratum-ibid. B 800 406–407. http://arXiv.org/abs/hep-th/0609059

    Google Scholar 

  25. Forrester, P.J., Nagao, T., Honner, G.: Correlations for the orthogonal-unitary and symplectic-unitary transitions at the hard and soft edges. Nucl. Phys. B 553, 601–643 (1999). http://arXiv.org/abs/cond-mat/9811142

  26. Ma\(\hat{\text{ c }}\)edo, A.M.S.: Universal parametric correlations at the soft edge of the spectrum of random matrix ensembles, EPL 26, 641–646 (1994)

  27. Ginibre, J.: Statistical ensembles of complex, quaternion, and real matrices. J. Math. Phys. 6, 440–449 (1965)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  28. Lehmann, N., Sommers, H.-J.: Eigenvalue statistics of random real matrices. Phys. Rev. Lett. 67, 941–944 (1991)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  29. Edelman, A.: The probability that a random real Gaussian matrix has \(k\) real eigenvalues, related distributions, and the circular law. J. Multivar. Anal. 60, 203–232 (1997)

    Article  MATH  MathSciNet  Google Scholar 

  30. Kanzieper, E.: Eigenvalue correlations in non-Hermitean symplectic random matrices. J. Phys. A 35, 6631–6644 (2002). http://arXiv.org/abs/cond-mat/0109287

    Google Scholar 

  31. Akemann, G., Kieburg, M., Phillips, M.J.: Skew-orthogonal Laguerre polynomials for chiral real asymmetric random matrices. J. Phys. A 43, 375207–1–24 (2010). http://arXiv.org/abs/1005.2983/math-ph

  32. Sommers, H.-J.: Symplectic structure of the real Ginibre ensemble. J. Phys. A 40, F671–F676 (2007). http://arXiv.org/abs/0706.1671/cond-mat.stat-mech

    Google Scholar 

  33. Sommers, H.-J., Wieczorek, W.: General eigenvalue correlations for the real Ginibre ensemble. J. Phys. A 41, 405003–1–24 (2008). http://arXiv.org/abs/0806.2756/cond-mat.stat-mech

    Google Scholar 

  34. Forrester, P.J., Mays, A.: A method to calculate correlation functions for \(\beta =1\) random matrices of odd size. J. Stat. Phys. 134, 443–462 (2009). http://arXiv.org/abs/0809.5116/math-ph

  35. van Eijndhoven, S.J.L., Meyers, J.L.H.: New orthogonality relations for the Hermite polynomials and related Hilbert spaces. J. Math. Ana. Appl. 146, 89–98 (1990)

    Article  MATH  Google Scholar 

  36. Di Francesco, P., Gaudin, M., Itzykson, C., Lesage, F.: Laughlin’s wave functions, Coulomb gases and expansions of the discriminant. Int. J. Mod. Phys. A 9, 4257–4352 (1994). http://arXiv.org/abs/hep-th/9401163

    Google Scholar 

  37. Plancherel, M., Rotach, W.: Sur les valeurs asymptotiques des polynomes d’Hermite. Commentarii Mathematici Helvetici 1, 227–254 (1929)

    Article  MATH  MathSciNet  Google Scholar 

  38. Szegö, G.: Orthogonal Polynomials. American Mathematical Society, Providence (1939)

    Google Scholar 

  39. Bohigas, O., Pato, M.P.: Transition between Hermitian and non-Hermitian Gaussian ensembles. J. Phys. A 46, 115001-1–11 (2013)

    Google Scholar 

  40. Kanzieper, E.: Exact replica treatment of non-Hermitean complex random matrices. In: Kovras, O. (ed.) Chapter 3 of Frontiers in Field Theory, pp. 23–51. Nova Science Publishers, New York (2005). http://arXiv.org/abs/cond-mat/0312006

  41. Mehta, M.L.: Random Matrices, 3rd edn. Elsevier, London (2004)

    MATH  Google Scholar 

  42. Anderson, G.W., Guionnet, A., Zeitouni, O.: An Introduction to Random Matrices. Cambridge University Press, Cambridge (2010)

    MATH  Google Scholar 

  43. Forrester, P.J., Nagao, T.: Eigenvalue statistics of the real Ginibre ensemble. Phys. Rev. Lett. 99, 050603–1–4 (2007). http://arXiv.org/abs/0706.2020/cond-mat.stat-mech

    Google Scholar 

  44. Forrester, P.J., Nagao, T.: Skew orthogonal polynomials and the partly symmetric real Ginibre ensemble. J. Phys. A 41, 375003–1–19 (2008). http://arXiv.org/abs/0806.0055/math-ph

    Google Scholar 

  45. Gradshteyn, I.S., Ryzhik, I.M.: Table of Integrals, Series and Products, 7th edn. Elsevier, Amsterdam (2007)

    MATH  Google Scholar 

  46. Borodin, A., Sinclair, C.D.: The Ginibre ensemble of real random matrices and its scaling limits. Commun. Math. Phys. 291, 177–224 (2009). http://arXiv.org/abs/0805.2986/math-ph

  47. Abramowitz, M., Stegun, I.A.: Handbook of Mathematical Functions. Dover, New York (1965)

    Google Scholar 

  48. Akemann, G., Basile, F.: Massive partition functions and complex eigenvalue correlations in matrix models with symplectic symmetry. Nucl. Phys. B 766, 150–177 (2007). http://arXiv.org/abs/math-ph/0606060

  49. Akemann, G.: The complex Laguerre symplectic ensemble of non-Hermitian matrices. Nucl. Phys. B 730, 253–299 (2005). http://arXiv.org/abs/hep-th/0507156

Download references

Acknowledgments

Partial support by the SFB\(|\)TR12 “Symmetries and Universality in Mesoscopic Systems” of the German research council DFG is acknowledged (G.A.). We would like to thank Queen Mary University of London (G.A.) and Bielefeld University (M.J.P.) for hospitality.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to G. Akemann.

Appendices

Appendix: 1 The Deformed Airy Function

In this appendix we introduce the deformed Airy function

$$\begin{aligned} {{\mathrm{Aid}}}(Z,\sigma ) \equiv \exp \left[ \frac{\sigma ^6}{12} + \frac{\sigma ^2 Z}{2} \right] {{\mathrm{Ai}}}\left( Z + \frac{\sigma ^4}{4} \right) , \end{aligned}$$
(7.1)

and consider three of its scaling limits. Using the large argument asymptotic behaviour of the Airy function (see Eq. 10.4.59 of [47])

$$\begin{aligned} {{\mathrm{Ai}}}(w)&\sim \frac{\exp \left[ {} - \tfrac{2}{3} w^{3/2} \right] }{2\sqrt{\pi }w^{1/4}} \left( 1 - \frac{5}{48w^{3/2}} + \mathcal {O}(w^{-3}) \right) \nonumber \\&= \frac{1}{2\sqrt{\pi }} \, \exp \left[ {} - \tfrac{2}{3} w^{3/2} - \tfrac{1}{4} \log w + \log \left( 1 - \frac{5}{48w^{3/2}} + \mathcal {O}(w^{-3}) \right) \right] \end{aligned}$$
(7.2)

we can easily show that, at large \(\sigma \) and fixed \(Z\) (or, at least, for \(Z\) growing more slowly than \(\sigma ^4\)), we have

$$\begin{aligned} {{\mathrm{Aid}}}(Z,\sigma ) = \frac{1}{\sqrt{2\pi }\,\sigma } \exp \left[ {} - \, \frac{Z^2}{2\sigma ^2} - \frac{Z}{\sigma ^4} + \frac{2Z^3-5}{6\sigma ^6} + \cdots \right] . \end{aligned}$$
(7.3)

From this, we immediately have the first of our scaling limits,

$$\begin{aligned} \lim _{\sigma \rightarrow \infty } \sigma \, {{\mathrm{Aid}}}(\sigma u, \sigma ) = \frac{1}{\sqrt{2\pi }} \, e^{-u^2/2} \end{aligned}$$
(7.4)

where we scale the first argument proportional to \(\sigma \).

We now consider a second scaling limit. In our interpolating Airy kernels, e.g. Eqs. (3.7) and (4.21), the deformed Airy function typically appears in combination with an additional factor:

$$\begin{aligned} f(Z,t,\sigma ) \equiv \exp \left[ - \frac{Y^2}{2\sigma ^2} \right] {{\mathrm{Aid}}}(Z+t, \sigma ). \end{aligned}$$
(7.5)

Consider therefore the behaviour of

$$\begin{aligned} g(Z,t,\sigma )&\equiv f(X + i\sigma Y, t, \sigma ) \nonumber \\&= \exp \left[ - \, \frac{Y^2}{2} \right] \, {{\mathrm{Aid}}}(X + i\sigma Y + t, \sigma ), \end{aligned}$$
(7.6)

where it should be noted that we are scaling the real and imaginary parts of the argument \(Z\) differently. At large \(\sigma \) and fixed \(X\) and \(Y\) (or, at least, with \(X\) and \(Y\) growing sufficiently slowly with \(\sigma \)), we then have from Eq. (7.3)

$$\begin{aligned}&g(Z,t,\sigma )\nonumber \\&\quad = \frac{1}{\sqrt{2\pi }\,\sigma } \, \exp \left[ - \frac{Y^2}{2} \right] \, \exp \left[ - \frac{(X + i\sigma Y + t)^2}{2\sigma ^2} - \frac{X + i\sigma Y + t}{\sigma ^4} + \frac{2(X + i\sigma Y + t)^3-5}{6\sigma ^6} + \ldots \right] \nonumber \\&\quad = \exp \left[ - \frac{iXY}{\sigma } - \frac{iY^3}{3\sigma ^3} \right] \, h(Z,t,\sigma ) \end{aligned}$$
(7.7)

where

$$\begin{aligned} h(Z,t,\sigma )&= \frac{1}{\sqrt{2\pi }\,\sigma } \, \exp \left[ \left( {} - \frac{X^2}{2\sigma ^2} - \frac{iY}{\sigma ^3} - \frac{X}{\sigma ^4} - \frac{Y^2X}{\sigma ^4} + \frac{iX^2Y}{\sigma ^5} + \frac{X^3}{3\sigma ^6} - \frac{5}{6\sigma ^6} + \ldots \right) \right. \nonumber \\&\qquad \qquad \qquad {} - t \left( \frac{iY}{\sigma } + \frac{X}{\sigma ^2} + \frac{1}{\sigma ^4} + \frac{Y^2}{\sigma ^4} - \frac{2iXY}{\sigma ^5} - \frac{X^2}{\sigma ^6} + \ldots \right) \nonumber \\&\qquad \qquad \qquad \qquad \left. {} - t^2 \left( \frac{1}{2\sigma ^2} - \frac{iY}{\sigma ^5} - \frac{X}{\sigma ^6} + \ldots \right) + \frac{t^3}{3\sigma ^6} + \ldots \right] . \end{aligned}$$
(7.8)

In the final expression here, we have only explicitly written those terms that come from expanding the terms in the exponent in Eq. (7.7) up to cubic order. We have ordered the terms in Eq. (7.8) by increasing powers of \(t\), and then by decreasing powers of \(\sigma \), although shortly we will let \(X\), \(Y\) and \(t\) themselves all be dependent on \(\sigma \), and so the relative sizes of the terms will change. Note that we chose to factor out two terms (the difference between \(g\) and \(h\)) for later convenience.

Let us now introduce some new coordinates \(x\) and \(y\), where again we scale the real and imaginary parts in different ways:

$$\begin{aligned} X&= a(\sigma )x + c(\sigma ), \nonumber \\ Y&= b(\sigma )y, \end{aligned}$$
(7.9)

in which the scalings and shift are given by (see [10])

$$\begin{aligned} a(\sigma )&\equiv \frac{\sigma }{\sqrt{6\log \sigma }}, \nonumber \\ b(\sigma )&\equiv \frac{\sigma ^{3/2}}{(6\log \sigma )^{1/4}}, \nonumber \\ c(\sigma )&\equiv a(\sigma )\Big ( 3 \log \sigma - \tfrac{5}{4} \log (6\log \sigma ) - \log (2\pi ) \Big ). \end{aligned}$$
(7.10)

We can choose how to scale \(t\), since this will be an integration variable in the expressions for the limiting kernels. In order to make the analysis as straightforward as possible, it turns out that the following, simple scaling is optimal:

$$\begin{aligned} t = d(\sigma ) u, \end{aligned}$$
(7.11)

where

$$\begin{aligned} d(\sigma ) \equiv \sigma . \end{aligned}$$
(7.12)

So now consider the function

$$\begin{aligned} m(x,y,u,\sigma )&\equiv h(X + iY, t, \sigma ) \nonumber \\&= h\Big (a(\sigma )x + c(\sigma ) + ib(\sigma )y, d(\sigma )u, \sigma \Big ). \end{aligned}$$
(7.13)

On substituting into Eq. (7.8), we get a very large number of terms in the exponent (almost 80, in fact). However, most of these will vanish in the limit \(\sigma \rightarrow \infty \), so we will retain only those that do not. After some simplification, we arrive at our second scaling limit of the deformed Airy function,

$$\begin{aligned} m(x,y,u,\sigma )&\sim \frac{(6 \log \sigma )^{5/8}}{\sigma ^{7/4}} \exp \left[ {} - \frac{x+y^2}{2} - \frac{u^2}{2} - \frac{u\sqrt{6\log \sigma }}{2} - \frac{iuy\sigma ^{3/2}}{(6\log \sigma )^{1/4}} \right] . \end{aligned}$$
(7.14)

For the third scaling limit, we use the asymptotic behaviour of the Airy function Eq. 10.4.60 of [47]

$$\begin{aligned} {{\mathrm{Ai}}}(-w) \sim \frac{\sin \left( \tfrac{2}{3} w^{3/2} + \frac{\pi }{4} \right) }{\sqrt{\pi } \, w^{1/4}} \end{aligned}$$
(7.15)

for large positive \(w\) to show that, for \(\alpha >0\) and at large real \(w\), (with \(z\), \(\sigma \) and \(\alpha \) fixed)

$$\begin{aligned} {{\mathrm{Aid}}}\left( \frac{z}{w} - \alpha w^2, \frac{\sigma }{w} \right)&\sim \frac{\exp [-\alpha \sigma ^2/2]}{\sqrt{\pi } \, \alpha ^{1/4}\sqrt{w}} \, \sin \left( \tfrac{2}{3} \alpha ^{3/2} w^3 \left( 1 - \frac{z}{\alpha w^3} - \frac{\sigma ^4}{4\alpha w^6} \right) ^{3/2} + \frac{\pi }{4} \right) \nonumber \\&\sim \frac{\exp [-\alpha \sigma ^2/2]}{\sqrt{\pi } \, \alpha ^{1/4}\sqrt{w}} \, \sin \left( \tfrac{2}{3} \alpha ^{3/2} w^3 - \sqrt{\alpha } \, z + \frac{\pi }{4} \right) . \end{aligned}$$
(7.16)

Appendix: 2 Hermitian Limit of the \(\beta =4\) Matrix-kernel

As a check in this appendix we will take the Hermitian limit \({\sigma \rightarrow 0}\) of the \(\beta =4\) matrix-kernel elements Eq. (4.21) as they appear in Eq. (4.4). Since all the elements of the matrix-kernel coincide to leading order, this leads to the appearance of first and second order derivatives acting on the Hermitian limit of the kernel. The same problem was encountered previously when taking the Hermitian limit of the \(\beta =4\) kernel [48] in the microscopic hard-edge scaling limit, starting from the (interpolating) Bessel kernel in the complex plane [49]. Because we will follow [48] closely, we can be brief. There it was shown by using properties of the Pfaffian that the matrix-kernel inside Eq. (4.4) can be changed as follows, without changing the value of the Pfaffian:

$$\begin{aligned}&\left( \begin{array}{ll} \mathcal {K}(z_i,z_j) &{} \mathcal {K}(z_i,z_j^*) \\ \mathcal {K}(z_i^*,z_j) &{} \mathcal {K}(z_i^*,z_j^*) \end{array} \right) \rightarrow \nonumber \\&\quad \left( \begin{array}{ll} \mathcal {K}(z_i,z_j) - \mathcal {K}(z_i,z_j^*) - \mathcal {K}(z_i^*,z_j) + \mathcal {K}(z_i^*, z_j^*)) &{} \frac{1}{2}( \mathcal {K}(z_i,z_j)- \mathcal {K}(z_i^*,z_j) + \mathcal {K}(z_i,z_j^*) - \mathcal {K}(z_i^*, z_j^*) ) \\ \frac{1}{2}( \mathcal {K}(z_i,z_j) - \mathcal {K}(z_i,z_j^*) + \mathcal {K}(z_i^*,z_j) - \mathcal {K}(z_i^*, z_j^*)) &{} \frac{1}{4}( \mathcal {K}(z_i,z_j) + \mathcal {K}(z_i,z_j^*) + \mathcal {K}(z_i^*,z_j) + \mathcal {K}(z_i^*, z_j^*)) \end{array} \right) \!.\nonumber \\ \end{aligned}$$
(8.1)

We have suppressed all labels here, partly for simplicity, and partly because the replacement is an identity both before and after taking the large-\(N\) limit. We will now expand the matrix-kernel elements from Eq. (4.21) as in Sect. 4.1, but this time to linear order in both \(Y_1\) and \(Y_2\), with \(Y_{1,2},\sigma \ll 1\).

$$\begin{aligned}&\mathcal {K}_{\text {Ai}}^{(4)}(Z_1,Z_2^*) \approx \frac{-i\sqrt{|Y_1Y_2|}}{4\sqrt{\pi }\,\sigma ^3} \, \exp \left[ - \, \frac{Y_1^2+Y_2^2}{2\sigma ^2} \right] \nonumber \\&\qquad \times \left\{ \int _0^{\infty } ds \int _0^s dt \, \Big ( {{\mathrm{Ai}}}(X_2+s)-iY_2{{\mathrm{Ai}}}'(X_2+s) \Big ) \Big ( {{\mathrm{Ai}}}(X_1+t)+iY_1{{\mathrm{Ai}}}'(X_1+t) \Big ) \right. \nonumber \\&\qquad \left. {} - \Big ( {{\mathrm{Ai}}}(X_1+s)+iY_1{{\mathrm{Ai}}}'(X_1+s) \Big ) \Big ( {{\mathrm{Ai}}}(X_2+t)-iY_2{{\mathrm{Ai}}}'(X_2+t) \Big ) \right\} \nonumber \\&\quad \equiv v(Y_1,Y_2)\, \Big [ T_1(X_1,X_2) + iY_1T_2(X_1,X_2) + iY_2T_3(X_1,X_2) + Y_1Y_2T_4(X_1,X_2) \Big ], \end{aligned}$$
(8.2)

where

$$\begin{aligned} v(Y_1,Y_2)&=\frac{-i\sqrt{|Y_1Y_2|}}{4\sqrt{\pi }\,\sigma ^3} \, \exp \left[ - \, \frac{Y_1^2+Y_2^2}{2\sigma ^2} \right] ,\nonumber \\ T_1(X_1,X_2)&= \int _0^{\infty } ds \int _0^s dt \, \Big [ {{\mathrm{Ai}}}(X_2+s){{\mathrm{Ai}}}(X_1+t) - {{\mathrm{Ai}}}(X_1+s){{\mathrm{Ai}}}(X_2+t) \Big ], \nonumber \\ T_2(X_1,X_2)&= \int _0^{\infty } ds \int _0^s dt \, \Big [ {{\mathrm{Ai}}}(X_2+s){{\mathrm{Ai}}}'(X_1+t) - {{\mathrm{Ai}}}'(X_1+s){{\mathrm{Ai}}}(X_2+t) \Big ] \nonumber \\&= 2\int _0^{\infty } ds \, {{\mathrm{Ai}}}(X_1+s){{\mathrm{Ai}}}(X_2+s) -{{\mathrm{Ai}}}(X_1) \int _0^{\infty } ds \, {{\mathrm{Ai}}}(X_2+s), \nonumber \\ T_3(X_1,X_2)&= \int _0^{\infty } ds \int _0^s dt \, \Big [ -{{\mathrm{Ai}}}'(X_2+s){{\mathrm{Ai}}}(X_1+t) + {{\mathrm{Ai}}}(X_1+s){{\mathrm{Ai}}}'(X_2+t) \Big ] \nonumber \\&= 2\int _0^{\infty } ds \, {{\mathrm{Ai}}}(X_1+s){{\mathrm{Ai}}}(X_2+s) -{{\mathrm{Ai}}}(X_2) \int _0^{\infty } ds \, {{\mathrm{Ai}}}(X_1+s),\nonumber \\ T_4(X_1,X_2)&=\int _0^{\infty } ds \int _0^s dt \, \Big [ {{\mathrm{Ai}}}'(X_2+s){{\mathrm{Ai}}}'(X_1+t) - {{\mathrm{Ai}}}'(X_1+s){{\mathrm{Ai}}}'(X_2+t) \Big ] \nonumber \\&= \int _0^{\infty } ds \, \Big [ {{\mathrm{Ai}}}(X_1+s){{\mathrm{Ai}}}'(X_2+s) - {{\mathrm{Ai}}}(X_2+s){{\mathrm{Ai}}}'(X_1+s) \Big ]. \end{aligned}$$
(8.3)

Whilst the integration in \(T_4\) is straightforward, in \(T_2\) and \(T_3\) we have again used Eq. (4.25). Applying the shift Eq. (8.1) under the Pfaffian we obtain

$$\begin{aligned}&\lim _{\sigma ,Y\ll 1}R_{k,\text {Ai}}^{(4)}(Z_1,\ldots ,Z_k) \approx {{\mathrm{Pf}}}_{i,j=1,\ldots ,k} \left[ \!\left( \! \begin{array}{ll} -4Y_iY_jv(Y_i,Y_j)T_4(X_i,X_j) &{}2iY_iv(Y_i,Y_j)T_2(X_i,X_j)\\ -2iY_jv(Y_i,Y_j)T_3(X_i,X_j) &{}v(Y_i,Y_j)T_1(X_i,X_j)\\ \end{array} \!\right) \!\right] \nonumber \\&\quad =\prod _{j=1}^k\frac{2Y_j^2}{\sqrt{\pi }\,\sigma ^3} \, \exp \left[ - \, \frac{Y_j^2}{\sigma ^2} \right] {{\mathrm{Pf}}}_{i,j=1,\ldots ,k} \left[ \left( \begin{array}{ll} -\tfrac{1}{4} T_4(X_i,X_j) &{}\tfrac{1}{4} T_2(X_i,X_j)\\ -\tfrac{1}{4} T_3(X_i,X_j) &{}-\tfrac{1}{4} T_1(X_i,X_j)\\ \end{array} \right) \right] . \end{aligned}$$
(8.4)

In the second step we have taken out all the \(Y\)-dependent factors, which, using Eq. (4.27), will give rise to the product \(\displaystyle \prod _{j=1}^k\delta (Y_j)\) in the limit \(\sigma \rightarrow 0\) as expected. To see the matching with the known result for \(\beta =4\) with real eigenvalues we note that

$$\begin{aligned} T_2(X_1,X_2)=T_3(X_2,X_1)=2\int _0^{\infty } ds \, {{\mathrm{Ai}}}(X_1+s){{\mathrm{Ai}}}(X_2+s) -{{\mathrm{Ai}}}(X_1) \int _{X_2}^{\infty } ds \, {{\mathrm{Ai}}}(s),\nonumber \\ \end{aligned}$$
(8.5)

where the first part is proportional to the \(\beta =2\) Airy kernel of real eigenvalues, Eq. (3.29). Furthermore we can show that

$$\begin{aligned} T_4(X_1,X_2) = 2\int _0^{\infty } ds \, {{\mathrm{Ai}}}(X_1+s){{\mathrm{Ai}}}'(X_2+s) + {{\mathrm{Ai}}}(X_2){{\mathrm{Ai}}}(X_1) =\frac{\partial }{\partial X_2}T_2(X_1,X_2),\nonumber \\ \end{aligned}$$
(8.6)

and

$$\begin{aligned} T_1(X_1,X_2) =2 \int _{X_2}^{X_1}dt \int _0^{\infty } ds \, {{\mathrm{Ai}}}(t+s){{\mathrm{Ai}}}(X_2+s)- \int _{X_2}^{X_1}dt{{\mathrm{Ai}}}(t) \int _{X_2}^{\infty } ds \, {{\mathrm{Ai}}}(s), \end{aligned}$$
(8.7)

the latter of which can be verified by drawing a sketch of the \(st\)-plane showing the regions where there are contributions to the integrals, and comparing with the original form of \(T_1\) in Eq. (8.3). Our final result for the limiting matrix kernel thus agrees with the literature, see e.g. page 162 of [42].

Appendix: 3 Some Properties of an Elementary Integral

Let us define the function for \(x \in \mathbb {R}\), \(\tau > 0\) and integer \(j \ge 0\)

$$\begin{aligned} I_j(x;\tau ) \equiv \int _{-\infty }^{\infty } dt \, {{\mathrm{sgn}}}(x-t) w^{(1)}(t) H_j \left( \frac{t}{\sqrt{2\tau }} \right) , \end{aligned}$$
(9.1)

where, for real \(x\),

$$\begin{aligned} w^{(1)}(x) \equiv w^{(1)}(x;\tau ) = \exp \left[ - \, \frac{x^2}{2(1+\tau )} \right] , \end{aligned}$$
(9.2)

and \(H_j(x)\) is the Hermite polynomial of degree \(j\) orthogonal with respect to \(\exp [-x^2]\). To determine a useful recurrence relation involving the \(I_j(x;\tau )\), we begin by integrating Eq. (9.1) by parts using Eq. 8.952.1 of [45]:

$$\begin{aligned}&I_j(x;\tau ) = \left[ {{\mathrm{sgn}}}(x-t) w^{(1)}(t) \frac{\sqrt{2\tau }}{2(j+1)} H_{j+1} \left( \frac{t}{\sqrt{2\tau }} \right) \right] _{-\infty }^{\infty } \nonumber \\&\qquad {} - \frac{\sqrt{2\tau }}{2(j+1)} \int _{-\infty }^{\infty } dt \, H_{j+1} \left( \frac{t}{\sqrt{2\tau }} \right) w^{(1)}(t) \left\{ -2\delta (x-t) + {{\mathrm{sgn}}}(x-t) \left( \frac{-t}{1+\tau } \right) \right\} \nonumber \\&\quad = \frac{\sqrt{2\tau }}{j+1} H_{j+1} \left( \frac{x}{\sqrt{2\tau }} \right) w^{(1)}(x) \nonumber \\&\qquad + \frac{\sqrt{\tau }}{\sqrt{2}(1+\tau )(j+1)} \int _{-\infty }^{\infty } dt \, t \, {{\mathrm{sgn}}}(x-t) w^{(1)}(t) H_{j+1} \left( \frac{t}{\sqrt{2\tau }} \right) . \end{aligned}$$
(9.3)

From Eq. 8.952.2 of [45] we have the recurrence relation for Hermite polynomials

$$\begin{aligned} t \, H_n\left( \frac{t}{\sqrt{2\tau }} \right) = \sqrt{2\tau } \left\{ \tfrac{1}{2}H_{n+1}\left( \frac{t}{\sqrt{2\tau }} \right) + n \, H_{n-1}\left( \frac{t}{\sqrt{2\tau }} \right) \right\} , \end{aligned}$$
(9.4)

and hence

$$\begin{aligned} I_j(x;\tau ) \!=\! \frac{\sqrt{2\tau }}{j+1} H_{j+1} \left( \frac{x}{\sqrt{2\tau }} \right) w^{(1)}(x) \!+\! \frac{\tau }{(1\!+\!\tau )(j\!+\!1)} \left\{ \tfrac{1}{2} I_{j+2}(x;\tau ) \!+\! (j+1)I_j(x;\tau ) \right\} \ , \end{aligned}$$
(9.5)

which can be rearranged to give the following recurrence relation for the \(I_j(x;\tau )\) themselves:

$$\begin{aligned} I_j(x;\tau )&= \frac{\sqrt{2\tau }(1+\tau )}{j+1} H_{j+1} \left( \frac{x}{\sqrt{2\tau }} \right) w^{(1)}(x) + \frac{\tau }{2(j+1)} \, I_{j+2}(x;\tau ). \end{aligned}$$
(9.6)

We can evaluate the integral \(I_j(x;\tau )\) by repeated application of Eq. (9.6), to give for odd \(j\):

$$\begin{aligned} I_j(x;\tau ) = - 2(1+\tau ) \left( \frac{2}{\tau } \right) ^{j/2} (j-1)!! \, w^{(1)}(x) \sum _{k=0}^{(j-1)/2} \frac{1}{(2k)!!} \left( \frac{\tau }{2} \right) ^k H_{2k}\left( \frac{x}{\sqrt{2\tau }}\right) ,\qquad \end{aligned}$$
(9.7)

and for even \(j\):

$$\begin{aligned} I_j(x;\tau )&= \left( \frac{2}{\tau } \right) ^{j/2} (j-1)!!\nonumber \\&\times \left\{ I_0(x;\tau ) - 2(1+\tau )w^{(1)}(x) \sum _{k=0}^{j/2-1} \frac{1}{(2k+1)!!} \left( \frac{\tau }{2} \right) ^{k+1/2} H_{2k+1}\left( \frac{x}{\sqrt{2\tau }}\right) \right\} ,\nonumber \\ \end{aligned}$$
(9.8)

where \(I_0(x;\tau )\) can be written in terms of the error function as

$$\begin{aligned} I_0(x;\tau ) = \sqrt{2\pi (1+\tau )} \, {{\mathrm{erf}}}\left( \frac{x}{\sqrt{2(1+\tau )}} \right) . \end{aligned}$$
(9.9)

Appendix: 4 Hermitian Limit of the \(\beta =1\) Matrix-Kernel

As with the \(\beta =4\) case in Appendix 2, we verify that the Hermitian limit \(\sigma \rightarrow 0\) of the matrix-kernel (and hence of all \(k\)-point correlation functions) agrees with known results in the literature. Since we do not have the discontinuity associated with the \(\beta =4\) ensemble, here we can, in all cases, let \(\sigma \rightarrow 0\) at fixed arguments (eigenvalues) \(Z_j=X_j+iY_j\). Indeed, note that no new Dirac delta functions appear in any of the following limits, since these are already present in the non-Hermitian case, see Eqs. (5.34) and (5.35).

It is useful to introduce the following general result for the pointwise limit of the complementary error function:

$$\begin{aligned} \lim _{\sigma \rightarrow 0} {{\mathrm{erfc}}}\left( \frac{|Y|}{\sigma } \right) = \delta _{Y0}, \end{aligned}$$
(10.1)

where, for \(x,y \in \mathbb {R}\), we define the Kronecker delta function as

$$\begin{aligned} \delta _{xy} = {\left\{ \begin{array}{ll} \,\, 1 &{} \text {if }x = y, \\ \,\, 0 &{} \text {otherwise}. \end{array}\right. } \end{aligned}$$
(10.2)
  1. (i)

    We begin with \(\hat{\mathcal {K}}_{\text {Ai}}^{(1)}(Z_1,Z_2)\), given in Eq. (5.38). On expanding the \(e^{\sigma ^2 t}\) inside the integral, and then letting \(\sigma \rightarrow 0\), this gives

    $$\begin{aligned}&\lim _{\sigma \rightarrow 0} \hat{\mathcal {K}}_{\text {Ai}}^{(1)}(Z_1,Z_2)\nonumber \\&\quad = \frac{(X_1-X_2)}{4} \, \int _0^{\infty } dt \, t \, {{\mathrm{Ai}}}(X_1+t) {{\mathrm{Ai}}}(X_2+t) \, \delta _{Y_1 0} \delta _{Y_2 0} \nonumber \\&\quad = \frac{1}{4} \left\{ \int _0^{\infty } dt \, t \, \big [ {{\mathrm{Ai}}}''(X_1+t) - t\,{{\mathrm{Ai}}}(X_1+t) \big ] \, {{\mathrm{Ai}}}(X_2+t) - (X_1 \leftrightarrow X_2) \right\} \, \delta _{Y_1 0} \delta _{Y_2 0}\nonumber \\&\quad = \frac{1}{4} \int _0^{\infty } dt \, t \, \big [ {{\mathrm{Ai}}}''(X_1+t){{\mathrm{Ai}}}(X_2+t) - (X_1 \leftrightarrow X_2) \big ] \, \delta _{Y_1 0} \delta _{Y_2 0}\nonumber \\&\quad = \frac{1}{4} \left\{ - \int _0^{\infty } dt \, {{\mathrm{Ai}}}'(X_1+t) \big [ t\,{{\mathrm{Ai}}}'(X_2+t) + {{\mathrm{Ai}}}(X_2+t) \big ] - (X_1 \leftrightarrow X_2) \right\} \, \delta _{Y_1 0} \delta _{Y_2 0}\nonumber \\&\quad = \frac{1}{4} \left\{ - \int _0^{\infty } dt \, {{\mathrm{Ai}}}'(X_1+t) {{\mathrm{Ai}}}(X_2+t) - (X_1 \leftrightarrow X_2) \right\} \, \delta _{Y_1 0} \delta _{Y_2 0}\nonumber \\&\quad = \frac{1}{2} \left\{ \frac{\partial }{\partial X_2} \mathcal {K}_{\text {Ai,Herm}}^{(2)}(X_1,X_2) + \tfrac{1}{2} \, {{\mathrm{Ai}}}(X_1){{\mathrm{Ai}}}(X_2) \right\} \, \delta _{Y_1 0} \delta _{Y_2 0}, \end{aligned}$$
    (10.3)

    where we used the fact that \({{\mathrm{Ai}}}''(X_1+t)=(X_1+t){{\mathrm{Ai}}}(X_1+t)\) in the first step, cancelled terms, integrated by parts, cancelled more terms, and then integrated only the \(X_1\)-term by parts. We also used the result Eq. (10.1) that the limit of the complementary error function involves the delta function defined in Eq. (10.2). In the final line, \(\mathcal {K}_{\text {Ai,Herm}}^{(2)}(X_1,X_2)=\int _0^{\infty } dt \, {{\mathrm{Ai}}}(X_1+t) {{\mathrm{Ai}}}(X_2+t)\) is the Hermitian limit of the \(\beta =2\) kernel, see Eq. (3.30).

  2. (ii)

    For \(G_{\text {Ai}}^{\mathbb {R}\,(1)}(Z_1,X_2)\), we take \(\sigma \rightarrow 0\) in Eq. (5.49). This gives

    $$\begin{aligned} - \lim _{\sigma \rightarrow \infty } G_{\text {Ai}}^{\mathbb {R}\,(1)}(Z_1,X_2) = \left\{ \mathcal {K}_{\text {Ai,Herm}}^{(2)}(X_1,X_2) + \tfrac{1}{2} \, {{\mathrm{Ai}}}(X_1)\left( 1 - \int _0^{\infty } dt \, {{\mathrm{Ai}}}(X_2 + t) \right) \right\} \, \delta _{Y_1 0}.\nonumber \\ \end{aligned}$$
    (10.4)
  3. (iii)

    For \(G_{\text {Ai}}^{\mathbb {C}\,(1)}(Z_1,Z_2)\), we have

    $$\begin{aligned} G_{\text {Ai}}^{\mathbb {C}\,(1)}(Z_1,Z_2) = 2i \, {{\mathrm{sgn}}}(Y_2) \hat{\mathcal {K}}_{\text {Ai}}^{(1)}(Z_1,Z_2^*). \end{aligned}$$
    (10.5)

    The pre-kernel \(\hat{\mathcal {K}}_{\text {Ai}}^{(1)}(Z_1,Z_2^*)\) only has a non-zero limit when the arguments are both real (i.e. when \(Y_1=Y_2=0\)). But in this case, \({{\mathrm{sgn}}}(Y_2)=0\), and so \(G_{\text {Ai}}^{\mathbb {C}\,(1)}(Z_1,Z_2) \rightarrow 0\) for all arguments.

  4. (iv)

    For \(W_{\text {Ai}}^{\mathbb {R}\mathbb {R}\,(1)}(X_1,X_2)\), we simply set \(\sigma = 0\) in Eq. (5.62):

    $$\begin{aligned}&- \lim _{\sigma \rightarrow 0} W_{\text {Ai}}^{\mathbb {R}\mathbb {R}\,(1)}(X_1,X_2)\nonumber \\&\quad = \lim _{\sigma \rightarrow 0} \Big \{ - 2 A(X_1,X_2) + B(X_1)B(X_2) + B(X_2) - B(X_1) \Big \} \nonumber \\&\quad = 2 \left\{ \int _{X_2}^{X_1}\! dt \, \mathcal {K}_{\text {Ai,Herm}}^{(2)}(t,X_2) + \tfrac{1}{2} \left( \int _{X_2}^{X_1}\! dt \, {{\mathrm{Ai}}}(t) \right) \! \left( 1 - \int _{X_2}^{\infty }\! dt \, {{\mathrm{Ai}}}(t) \right) \! \right\} . \end{aligned}$$
    (10.6)

    The second step here can be proved with a few lines of easy manipulation.

  5. (v)

    For \(W_{\text {Ai}}^{\mathbb {R}\mathbb {C}\,(1)}(X_1,Z_2)\), we have

    $$\begin{aligned} W_{\text {Ai}}^{\mathbb {R}\mathbb {C}\,(1)}(X_1,Z_2) = 2i \, {{\mathrm{sgn}}}(Y_2) G_{\text {Ai}}^{\mathbb {R}\,(1)}(Z_2^*,X_1). \end{aligned}$$
    (10.7)

    The function \(G_{\text {Ai}}^{\mathbb {R}\,(1)}(Z_2^*,X_1)\) only has a non-zero limit when the first argument is real (i.e. when \(Y_2=0\)). But in this case, \({{\mathrm{sgn}}}(Y_2)=0\), and so \(W_{\text {Ai}}^{\mathbb {R}\mathbb {C}\,(1)}(X_1,Z_2) \rightarrow 0\) for all arguments.

  6. (vi)

    For \(W_{\text {Ai}}^{\mathbb {C}\mathbb {C}\,(1)}(Z_1,Z_2)\), we have

    $$\begin{aligned} W_{\text {Ai}}^{\mathbb {C}\mathbb {C}\,(1)}(Z_1,Z_2) = -2i \, {{\mathrm{sgn}}}(Y_1) G_{\text {Ai}}^{\mathbb {C}\,(1)}(Z_1^*,Z_2). \end{aligned}$$
    (10.8)

    But \(G_{\text {Ai}}^{\mathbb {C}\,(1)}(Z_1^*,Z_2)\rightarrow 0\), and so \(W_{\text {Ai}}^{\mathbb {C}\mathbb {C}\,(1)}(Z_1,Z_2)\rightarrow 0\) for all arguments.

  7. (vii)

    The Hermitian limit of the bivariate weight function \(\mathcal {F}_{\infty }^{(1)}(Z_1,Z_2)\) is trivial, since \(\mathcal {F}_{\infty }^{(1)}(Z_1,Z_2)\) does not depend on \(\sigma \). We have simply

    $$\begin{aligned} \lim _{\sigma \rightarrow 0} \mathcal {F}_{\infty }^{(1)}(Z_1,Z_2) = 2i\delta ^{(2)}(Z_1-Z_2^*){{\mathrm{sgn}}}(Y_1) + \delta (Y_1)\delta (Y_2){{\mathrm{sgn}}}(X_2-X_1).\qquad \end{aligned}$$
    (10.9)

In fact, we can drop the first term completely in the Hermitian limit, since it always gets multiplied by \(\hat{\mathcal {K}}_{\text {Ai}}^{(1)}(Z_1,Z_2;\sigma =0) \propto \delta _{Y_1 0}\) when evaluating the Pfaffian to determine the correlation functions.

We expect the \(\sigma \rightarrow 0\) limits to be consistent with those in the literature, e.g. on page 162 of [42]; our limit of \(G_{\text {Ai}}^{(1)}(Z_1,Z_2)\) matches [42] precisely. Our limit of \(\hat{\mathcal {K}}_{\text {Ai}}^{(1)}(Z_1,Z_2)\) is one half of that given in [42], but our limit of \(W_{\text {Ai}}^{(1)}(Z_1,Z_2)\) is twice what is given in [42]. It therefore follows that, once we evaluate the Pfaffian, we will get precisely the same \(k\)-point correlation functions for all values of \(k\).

Rights and permissions

Reprints and permissions

About this article

Cite this article

Akemann, G., Phillips, M.J. The Interpolating Airy Kernels for the \(\beta =1\) and \(\beta =4\) Elliptic Ginibre Ensembles. J Stat Phys 155, 421–465 (2014). https://doi.org/10.1007/s10955-014-0962-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-014-0962-6

Keywords

Navigation