Skip to main content
Log in

Determinantal Structures in the O’Connell-Yor Directed Random Polymer Model

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

Abstract

We study the semi-discrete directed random polymer model introduced by O’Connell and Yor. We obtain a representation for the moment generating function of the polymer partition function in terms of a determinantal measure. This measure is an extension of the probability measure of the eigenvalues for the Gaussian unitary ensemble in random matrix theory. To establish the relation, we introduce another determinantal measure on larger degrees of freedom and consider its few properties, from which the representation above follows immediately.

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

References

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

    Article  MathSciNet  MATH  Google Scholar 

  2. Amir, G., Corwin, I., Quastel, J.: Probability distribution of the free energy of the continuum directed random polymer in 1 + 1 dimensions. Commun. Pure Appl. Math. 64, 466–537 (2011)

    Article  MathSciNet  MATH  Google Scholar 

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

    MATH  Google Scholar 

  4. Andréief, C.: Note sur une relation les integrales de nies des produits des fonctions. Mém. de la Soc. Sci. Bordx. 2, 1–14 (1883)

    Google Scholar 

  5. Andrews, G.E., Askey, R.: Special Functions. Cambridge University Press, Cambridge (2014)

    MATH  Google Scholar 

  6. Barraquand, G., Corwin, I.: Random-walk in Beta-distributed random environment (2015). arXiv:1503.04117

  7. Baryshnikov, Y.: GUEs and queues. Probab. Theory Relat. Fields 119, 256–274 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  8. Bertini, L., Cancrini, N.: The stochastic heat equation: Feynman-Kac formula and intermittence. J. Stat. Phys. 78, 1377–1401 (1995)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Bertini, L., Giacomin, G.: Stochastic Burgers and KPZ equations from particle systems. Commun. Math. Phys. 183, 571–607 (1997)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  10. Borodin, A., Corwin, I.: Macdonald processes. Probab. Theory Relat. Fields 158, 225–400 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  11. Borodin, A., Corwin, I., Ferrari, P.L.: Free energy fluctuations for directed polymers in random media in 1\(+\)1dimension. Commun. Pure Appl. Math. 67, 1129–1214 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  12. Borodin, A., Corwin, I., Ferrari, P.L., Vető, B.: Height fluctuations for the stationary KPZ equation. Math. Phys. Anal. Geom. 18(2015), 20 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  13. Borodin, A., Corwin, I., Remenik, D.: Log-gamma polymer free energy fluctuations via a fredholm determinant identity. Commun. Math. Phys. 324, 215–232 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Borodin, A., Corwin, I., Sasamoto, T.: From duality to determinants for q-TASEP and ASEP. Ann. Probab. 42, 2314–2382 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  15. Borodin, A., Ferrari, P.L.: Large time asymptotics of growth models on space-like paths I: PushASEP. Electron. J. Probab. 13, 1380–1418 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  16. Borodin, A., Ferrari, P.L.: Anisotropic growth of random surfaces in \(2+1\) dimensions. Commun. Math. Phys. 325, 603–684 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  17. Borodin, A., Ferrari, P.L., Prähofer, M.: Fluctuations in the discrete TASEP with periodic initial configurations and the airy \(_1\) process. Int. Math. Res. Pap. 2007, rpm002 (2007)

    MATH  Google Scholar 

  18. Borodin, A., Ferrari, P.L., Prähofer, M., Sasamoto, T.: Fluctuation properties of the TASEP with periodic initial configuration. J. Stat. Phys. 129, 1055–1080 (2007)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Borodin, A., Ferrari, P.L., Sasamoto, T.: Large time asymptotics of growth models on space-like paths II: PNG and parallel TASEP. Commun. Math. Phys. 283, 417–449 (2008)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  20. Borodin, A., Ferrari, P.L., Sasamoto, T.: Transition between Airy\(_1\) and Airy\(_2\) processes and TASEP fluctuations. Commun. Pure Appl. Math. 61, 1603–1629 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  21. Borodin, A., Ferrari, P.L., Sasamoto, T.: Two speed TASEP. J. Stat. Phys. 137, 936–977 (2009)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  22. Bump, D.: The Rankin Selberg Method: A Survey, Number Theory, Trace Formulas and Discrete Groups. Academic Press, New York (1989)

    MATH  Google Scholar 

  23. Calabrese, P., Le Doussal, P.: Exact solution for the KPZ equation with flat initial conditions. Phys. Rev. Lett. 106, 250603 (2011)

    Article  ADS  Google Scholar 

  24. Calabrese, P., Le Doussal, P., Rosso, A.: Free-energy distribution of the directed polymer at high temperature. EPL 90, 20002 (2010)

    Article  ADS  Google Scholar 

  25. Corwin, I., O’Connell, N., Seppäläinen, T., Zygouras, N.: Tropical combinatorics and Whittaker functions. Duke Math. J. 163, 513–563 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  26. Corwin, I., Seppäläinen, T., Shen, H.: The strict-weak lattice polymer. J. Stat. Phys. 160, 1027–1053 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  27. Corwin, I., Tsai, L-C.: KPZ equation limit of higher-spin exclusion processes (2015). arXiv:1505.04158

  28. Dean, D.S., Le Doussal, P., Majumdar, S.N., Schehr, G.: Finite temperature free fermions and the Kardar-Parisi-Zhang equation at finite time. Phys. Rev. Lett. 114, 110402 (2015)

    Article  ADS  Google Scholar 

  29. Dotsenko, V.: Bethe ansatz derivation of the Tracy-Widom distribution for one-dimensional directed polymers. EPL 90, 20003 (2010)

    Article  ADS  Google Scholar 

  30. Dotsenko, V.: Replica Bethe ansatz derivation of the Tracy-Widom distribution of the free energy fluctuations in one-dimensional directed polymers. J. Stat. Mech. 2010, P07010 (2010)

    Google Scholar 

  31. Dyson, F.J.: A Brownian-motion model for the eigenvalues of a random matrix. J. Math. Phys. 3, 1191–1198 (1962)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  32. Ferrari, P.L., Spohn, H., Weiss, T.: Brownian motions with one-sided collisions: the stationary case (2015). arXiv:1502.01468

  33. Ferrari, P.L., Spohn, H., Weiss, T.: Scaling limit for Brownian motions with one-sided collisions. Ann. Appl. Probab. 25, 1349–1382 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  34. Ferrari, P.L., Vető, B.: Tracy-Widom asymptotics for q-TASEP. Ann. Inst. H. Poincaré Probab. Stat. 51, 1465–1485 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Forrester, P.J.: Log-gases and random matrices. Princeton University Press, Princeton (2010)

    MATH  Google Scholar 

  36. Gravner, J., Tracy, C.A., Widom, H.: Limit theorems for height fluctuations in a class of discrete space and time growth models. J. Stat. Phys. 102, 1085–1132 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  37. Hairer, M.: Solving the KPZ equation. Ann. Math. 178, 559–664 (2013)

    Article  MathSciNet  MATH  Google Scholar 

  38. Imamura, T., Sasamoto, T.: Exact solution for the stationary Kardar-Parisi-Zhang equation. Phys. Rev. Lett. 108, 190603 (2012)

    Article  ADS  Google Scholar 

  39. Imamura, T., Sasamoto, T.: Stationary correlations for the 1D KPZ equation. J. Stat. Phys. 150, 908–939 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  40. Itô, K.: Multiple Wiener integral. J. Math. Soc. Jpn. 3, 157–169 (1951)

    Article  MathSciNet  MATH  Google Scholar 

  41. Janjigian, C.: Large deviations of the free energy in the O’Connell-Yor polymer. J. Stat. Phys. 160, 1054–1080 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  42. Johansson, K.: Shape fluctuations and random matrices. Commun. Math. Phys. 209, 437–476 (2000)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  43. Johansson, K.: Discrete polynuclear growth and determinantal processes. Commun. Math. Phys. 242, 277–329 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  44. Kardar, M., Parisi, G., Zhang, Y.C.: Dynamic scaling of growing interfaces. Phys. Rev. Lett. 56, 889–892 (1986)

    Article  ADS  MATH  Google Scholar 

  45. Katori, M.: O’Connell’s process as a vicious Brownian motion. Phys. Rev. E 84, 061144 (2011)

    Article  ADS  Google Scholar 

  46. Katori, M.: Survival probability of mutually killing Brownian motions and the O’Connell process. J. Stat. Phys. 147, 206–223 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  47. Katori, M.: System of complex Brownian motions associated with the O’Connell process. J. Stat. Phys. 149, 411–431 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  48. Kupiainen, A.: Renormalization group and stochastic PDEs. Ann. H. Poincaré 17, 497–535 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  49. Le Doussal, P., Calabrese, P.: The KPZ equation with flat initial condition and the directed polymer with one free end. J. Stat. Mech. 2012, P06001 (2012)

    Article  Google Scholar 

  50. Macdonald, I.G.: Symmetric Functions and Hall Polynomials. Clarendon Press, Oxford (1995)

    MATH  Google Scholar 

  51. Major, P.: Multiple Wiener-Itô Integrals with Applications to Limit Theorems. Lecture Notes in Mathematics, 2nd edn, p. 849. Springer, New York (2013)

    Google Scholar 

  52. Mehta, M.L.: Random Matrices. Elsevier, Amsterdam (2004)

    MATH  Google Scholar 

  53. Moreno Flores, G., Quastel, J.: unpublished

  54. Moriarty, J., O’Connell, N.: On the free energy of a directed polymer in a Brownian environment. Markov Process. Relat. Fields 13, 251–266 (2007)

    MathSciNet  MATH  Google Scholar 

  55. Mueller, C.: On the support of solutions to the heat equation with noise. Stoch. Stoch. Rep. 37, 225–245 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  56. Nagao, T., Sasamoto, T.: Asymmetric simple exclusion process and modified random matrix ensembles. Nucl. Phys. B 699, 487–502 (2004)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  57. O’Connell, N.: Directed polymers and the quantum Toda lattice. Ann. Probab. 40, 437–458 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  58. O’Connell, N., Ortmann, J.: Tracy-Widom asymptotics for a random polymer model with gamma-distributed weights. Electron. J. Probab. 20, 1–18 (2015)

    MathSciNet  MATH  Google Scholar 

  59. O’Connell, N., Yor, M.: Brownian analogues of Burke’s theorem. Stoch. Process. Appl. 96, 285–304 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  60. Okounkov, A., Reshetikhin, N.: Correlation function of Schur process with application to local geometry of a random 3-dimensional Young diagram. J. Am. Math. Soc. 16, 581–603 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  61. O’Connell, N., Seppäläinen, T., Zygouras, N.: Geometric RSK correspondence, Whittaker functions and symmetrized random polymers. Inven. Math. 197, 361–416 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  62. Ortmann, J., Quastel, J., Remenik, D.: A Pfaffian representation for flat ASEP (2015). arXiv:1501.05626

  63. Ortmann, J., Quastel, J., Remenik, D.: Exact formulas for random growth with half-flat initial data. Ann. Appl. Probab. 26, 507–548 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  64. Rakos, A., Schütz, G.M.: Current distribution and random matrix ensembles for an integrable asymmetric fragmentation process. J. Stat. Phys. 118, 511–530 (2005)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  65. Sasamoto, T.: Spatial correlations of the 1D KPZ surface on a flat substrate. J. Phys. A 38, L549–L556 (2005)

    Article  ADS  MathSciNet  Google Scholar 

  66. Sasamoto, T., Spohn, H.: Exact height distributions for the KPZ equation with narrow wedge initial condition. Nucl. Phys. B 834, 523–542 (2010)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  67. Sasamoto, T., Spohn, H.: One-dimensional Kardar-Parisi-Zhang equation: an exact solution and its universality. Phys. Rev. Lett. 104, 230602 (2010)

    Article  ADS  Google Scholar 

  68. Sasamoto, T., Spohn, H.: The \(1+1\)-dimensional Kardar-Parisi-Zhang equation and its universality class. J. Stat. Mech. 2010, P11013 (2010)

    Article  Google Scholar 

  69. Sasamoto, T., Spohn, H.: The crossover regime for the weakly asymmetric simple exclusion process. J. Stat. Phys. 140, 209–231 (2010)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  70. Sasamoto, T., Wadati, M.: Determinant form solution for the derivative nonlinear Schrödinger type model. J. Phys. Soc. Jpn. 67, 784–790 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  71. Schütz, G.M.: Exact solution of the master equation for the asymmetric exclusion process. J. Stat. Phys. 88, 427–445 (1997)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  72. Seppäläinen, T., Valkó, B.: Bounds for scaling exponents for a 1+1 dimensional directed polymer in a Brownian environment. ALEA 7, 451–476 (2010)

    MathSciNet  MATH  Google Scholar 

  73. Soshnikov, A.: Determinantal random point fields. Russ. Math. Surv. 55, 923–975 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  74. Stade, E.: Archimedean \(L\)-factors on GL\((n)\times \)GL(n) and generalized Barnes integrals. Israel J. Math. 127, 201–219 (2002)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  ADS  MathSciNet  MATH  Google Scholar 

  76. Tracy, C.A., Widom, H.: Correlation functions, cluster functions, and spacing distributions for random matrices. J. Stat. Phys. 92, 809–835 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  77. Vető, B.: Tracy-Widom limit of q-Hahn TASEP. Electron. J. Probab. 20, 1–22 (2015)

    MathSciNet  MATH  Google Scholar 

  78. Warren, J.: Dyson’s Brownian motions, intertwining and interlacing. Electron. J. Probab. 12, 573–590 (2007)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

The work of T. I. and T. S. is supported by KAKENHI (25800215) and KAKENHI (25103004, 15K05203, 14510499) respectively.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Tomohiro Sasamoto.

Additional information

This is one of several papers published in JSP comprising the “Special Issue: KPZ” (Volume (169), Issue (4), year (2015)).

Appendices

Appendix 1: Proof of Lemma 5

First we give a proof of (2.29). For this purpose, it is sufficient to show the case of \(m=1,~x=0\),

$$\begin{aligned} \int _{-\infty }^{\infty }dx\, e^{-ax}f_B(x)=\frac{\pi }{\beta }\cot \left( \frac{\pi a}{\beta }\right) . \end{aligned}$$
(7.1)

Furthermore setting \(e^{\beta x}=y\), \(a/\beta =b\), one sees that (7.1) is rewritten as

$$\begin{aligned} \int _0^\infty dy\, h_b(y)=\pi \cot \pi b, \end{aligned}$$
(7.2)

where \(h_b(y)=y^{-b-1}/(y-1)\) and we take the branch cut of \(h_b(y)\) to be the positive real axis. Hence here we prove (7.2). We set the contour C as depicted in Fig. 2 with \(\alpha =1\). From the Cauchy integral theorem, we find

$$\begin{aligned} \int _{C} dy\, h_b(y)=0. \end{aligned}$$
(7.3)

Dividing the contour C into \(C_i,~i=1,\ldots ,6\) as in Fig. 2, we find that by simple calculations,

$$\begin{aligned}&\lim _{\scriptstyle \begin{array}{c} \delta \rightarrow 0, \epsilon \rightarrow 0 \\ R\rightarrow \infty \end{array}} \int _{C_1} dy\, h_b(y) =-\lim _{\scriptstyle \begin{array}{c} \delta \rightarrow 0, \epsilon \rightarrow 0 \\ R\rightarrow \infty \end{array}} e^{2\pi i b}\int _{C_4}dy\, h_b(y) =\int _0^\infty dy\, h_b(y),\nonumber \\&\lim _{\epsilon \rightarrow 0}\int _{C_2dy\,}dy\, h_b(y)=\lim _{\epsilon \rightarrow 0}e^{2\pi i b} \int _{C_5}dy\, h_b(y)=-\pi i,\nonumber \\&\lim _{R\rightarrow \infty }\int _{C_3}dy\, h_b(y)=\lim _{\delta \rightarrow 0} \int _{C_6}dy\, h_b(y)=0, \end{aligned}$$
(7.4)

where note that the factors \(e^{2\pi i b}\)s come from the cut locus of \(y^{-b}\).

From (7.3),(7.4), we get

$$\begin{aligned} 0=\lim _{\scriptstyle \begin{array}{c} \delta \rightarrow 0, \epsilon \rightarrow 0 \\ R\rightarrow \infty \end{array}} \sum _{j=1}^6\int _{C_j}dy\, h_b(y)=(1-e^{-2\pi i b})\int _0^\infty dy\, h_b(y) \, -(1+e^{-2\pi i b})\pi i, \end{aligned}$$
(7.5)

which leads to (7.2).

Fig. 2
figure 2

The contour C on \(\mathbb {C}\), where \(\alpha \in (0,\infty )\). It consists of the paths \(C_1,\ldots ,C_6\)

Next we give a proof of (2.30). For this purpose we first show the following relation. Let \(I_j(x),~j=1,2,\ldots \), \(x\in (0,\infty )\), be

$$\begin{aligned} I_j(x)=\int _0^\infty dw\, \frac{1}{x-w}\frac{(\log w)^{j-1}}{w+1}. \end{aligned}$$
(7.6)

Then we have

$$\begin{aligned} I_j(x)=\frac{1}{x+1}\,r_{j}(\log x), \end{aligned}$$
(7.7)

where \(r_k(x)~(k=0,1,2,\ldots )\) is a kth order polynomial of x where the coefficient of the highest degree is 1 / k. This relation (7.7) can be derived by considering the integration of \(m_j(w;x):=(\log w)^j/((x-w)(w+1)),~x>0\) with respect to w along the contour C in Fig. 2 with \(\alpha =x\) and \(R>1\).

Note that

$$\begin{aligned} \int _{C}dw\, m_j(w;x)=2\pi i\,\frac{(\pi i)^{j}}{x+1}, \end{aligned}$$
(7.8)

where RHS corresponds to the residue of \(m_j(w;x)\) at \(w=-1\). As in the previous case (7.4), one easily gets

$$\begin{aligned} \lim _{\scriptstyle \begin{array}{c} \delta \rightarrow 0, \epsilon \rightarrow 0 \\ R\rightarrow \infty \end{array}} \int _{C_1}dw\,m_j(w;x)&=I_{j+1}(x),\nonumber \\ \lim _{\scriptstyle \begin{array}{c} \delta \rightarrow 0, \epsilon \rightarrow 0 \\ R\rightarrow \infty \end{array}} \int _{C_4}dw\,m_j(w;x)&=-\int _0^\infty dw\, m_j(we^{2\pi i};x) =-\int _0^\infty dw\, \frac{(\log w+2\pi i)^j}{(x-w)(w+1)}\nonumber \\&\quad =-\sum _{k=0}^{j}\left( {\begin{array}{c}j\\ k\end{array}}\right) (2\pi i)^{j-k}I_{k+1}(x),\nonumber \\ \lim _{\epsilon \rightarrow 0}\int _{C_2} dw\, m_j(w;x)&=\pi i\,\frac{(\log x)^j}{x+1},~~ \lim _{\epsilon \rightarrow 0}\int _{C_5}m_j(w;x)dw=\pi i\,\frac{(\log x+2\pi i)^j}{x+1},\nonumber \\ \lim _{R\rightarrow \infty }\int _{C_3} dw\, m_j(w;x)&=\lim _{\delta \rightarrow 0} \int _{C_6}dw\,m_j(w;x)=0. \end{aligned}$$
(7.9)

Substituting (7.9) into (7.8), we find

$$\begin{aligned} 2\pi i\frac{(\pi i)^j}{x+1}&=\lim _{\scriptstyle \begin{array}{c} \delta \rightarrow 0, \epsilon \rightarrow 0 \\ R\rightarrow \infty \end{array}}\sum _{k=1}^6\int _{C_k}dw\, m_j(w;x)\nonumber \\&=I_{j+1}(x)-\sum _{k=0}^j\left( {\begin{array}{c}j\\ k\end{array}}\right) (2\pi i)^{j-k}I_{k+1}(x)+\frac{\pi i}{x+1} \left( (\log x)^j+\left( \log x+2\pi i\right) ^j\right) . \end{aligned}$$
(7.10)

Thus we obtain

$$\begin{aligned} I_{j}(x)=\frac{(\log x)^{j}+(\log x+2\pi i)^j-2(\pi i)^j}{2j(x+1)}-\frac{1}{j}\sum _{k=0}^{j-2}\left( {\begin{array}{c}j\\ k\end{array}}\right) (2\pi i)^{j-1-k}I_{k+1}(x) \end{aligned}$$
(7.11)

which leads to (7.7).

Here we show (2.30). We find that (2.30) is rewritten as

$$\begin{aligned} J_B^{*(m)}J_F(x)=q_m\left( \frac{\log x}{\beta }\right) J_F(x), \end{aligned}$$
(7.12)

where \(q_m(x)\) is defined below (2.30) and the functions \(J_F(x)\) and \(J_B(x)\) on \(\mathbb {R}_+\) are defined by \(J_F(x)=1/(x+1)\) and \(J_B(x)=1/\beta x\).

We prove (7.12) by using (7.7) and by mathematical induction: suppose that (7.12) holds for \(m=N-1\). Then we get

$$\begin{aligned} J^{*(N)}_BJ_F(x)&=\frac{1}{\beta }\int _{0}^\infty dy\,\frac{1}{x-y}J^{*(N-1)}_BJ_F(y) =\frac{1}{\beta ^N(N-1)!}I_{N}(y)+\sum _{k=0}^{N-2}\frac{c_k}{\beta ^{k+1}}I_{k+1}(y) \end{aligned}$$
(7.13)

where \(c_k (k=0,1,\ldots ,N-2)\) is the coefficient of \(x^k\) in \(q_{N-1}(x)\) and in the last equality we used the assumption for the mathematical induction and (7.6). Considering (7.7), we arrive at (7.12). \(\square \)

Appendix 2: Proof of Lemma 10

To show Lemma 10, we will use the following identity. For \((x_1,\ldots , x_{N-1})\in \mathbb {R}^{N-1}\) satisfying \(x_1>\cdots >x_{N-1}\) and \((y_1,\ldots ,y_N)\in \mathbb {R}^N\), we have

$$\begin{aligned}&\sum _{\sigma \in S_N}\mathrm{{sgn}}\sigma \prod _{j=2}^N1_{>0}\left( x_{j-1}-y_{\sigma (j)}\right) \nonumber \\&\qquad \quad =\sum _{\sigma \in S_N}\mathrm{{sgn}}\sigma \prod _{j=2}^N1_{>0}\left( x_{j-1}-y_{\sigma (j)}\right) 1_{>0}\left( y_{\sigma (j-1)}-x_{j-1}\right) , \end{aligned}$$
(8.1)

where \(S_N\) is the permutation of \((1,2,\ldots ,N)\). For the proof of (8.1), it is sufficient to show for \(m=1,2, \cdots , N\),

$$\begin{aligned}&\sum _{\sigma \in S_N}\mathrm{{sgn}}\sigma \prod _{j=2}^N1_{>0}(x_{j-1}-y_{\sigma (j)})\cdot \prod _{k=2}^{m}1_{>0}(y_{\sigma (k-1)}-x_{k-1})\cdot 1_{>0}(x_{m}-y_{\sigma (m)}) =0, \end{aligned}$$
(8.2)

where, as in (8.1), we assume the condition \(x_1>x_2>\cdots >x_{N-1}\). This can easily be obtained by noting that

$$\begin{aligned}&\sum _{\sigma \in S_N}\mathrm{{sgn}}\sigma \prod _{j=2}^N1_{>0}(x_{j-1}-y_{\sigma (j)})\cdot \prod _{k=2}^{m}1_{>0}(y_{\sigma (k-1)}-x_{k-1})\cdot 1_{>0}(x_{m}-y_{\sigma (m)})\nonumber \\&\quad =\sum _{\sigma \in S_N}\mathrm{{sgn}}\sigma \prod _{\begin{array}{c} j=2\\ j\ne m \end{array}}^N1_{>0}(x_{j-1}-y_{\sigma (j)})\cdot \prod _{k=2}^{m}1_{>0}(y_{\sigma (k-1)}-x_{k-1})\cdot 1_{>0}(x_{m}-y_{\sigma (m)})\nonumber \\&\quad =\sum _{\sigma \in S_N}\mathrm{{sgn}}\sigma \prod _{\begin{array}{c} j=2\\ j\ne m,~m+1 \end{array}}^N 1_{>0}(x_{j-1}-y_{\sigma (j)})\nonumber \\&\qquad \times \prod _{k=2}^{m}1_{>0}(y_{\sigma (k-1)}-x_{k-1})\cdot 1_{>0}(x_{m}-y_{\sigma (m)})1_{>0}(x_{m}-y_{\sigma (m+1)})\nonumber \\&\quad =0 \end{aligned}$$
(8.3)

where in the first equality we used the fact that the factor \(1_{>0}(x_{m-1}-y_{\sigma (m)})\) can be omitted in this equation thanks to the factor \(1_{>0}(x_m-y_{\sigma (m)})\) with the condition \(x_{m-1}>x_{m}\) and the last equality follows from the fact that the term with \(\sigma \) cancels the term with \(\sigma '\) where \(\sigma '\) is defined in terms of \(\sigma \) as \(\sigma '(m)=\sigma (m+1)\) and \(\sigma '(m+1)=\sigma (m)\) with \(\sigma '(k)=\sigma (k)\) for \(k\ne m,~m+1\).

Using (8.2), we have for \(x_1>x_2>\cdots >x_{N-1}\)

$$\begin{aligned}&\sum _{\sigma \in S_N}\mathrm{{sgn}}\sigma \prod _{j=2}^N1_{>0}\left( x_{j-1}-y_{\sigma (j)}\right) \nonumber \\&\quad = \sum _{\sigma \in S_N}\mathrm{{sgn}}\sigma \prod _{j=2}^N1_{>0}\left( x_{j-1}-y_{\sigma (j)}\right) \cdot (1-1_{>0}(x_{1}-y_{\sigma (1)}))\nonumber \\&\quad =\sum _{\sigma \in S_N}\mathrm{{sgn}}\sigma \prod _{j=2}^N1_{>0}\left( x_{j-1}-y_{\sigma (j)}\right) \cdot 1_{>0}(y_{\sigma (1)}-x_1)\nonumber \\&\quad =\sum _{\sigma \in S_N}\mathrm{{sgn}}\sigma \prod _{j=2}^N1_{>0}\left( x_{j-1}-y_{\sigma (j)}\right) \cdot 1_{>0}(y_{\sigma (1)}-x_1)(1-1_{>0}(x_{2}-y_{\sigma (2)}))\nonumber \\&\quad =\sum _{\sigma \in S_N}\mathrm{{sgn}}\sigma \prod _{j=2}^N1_{>0}\left( x_{j-1}-y_{\sigma (j)}\right) \cdot \prod _{k=1}^2 1_{>0}(y_{\sigma (k)}-x_k), \end{aligned}$$
(8.4)

where for the first and the third equality, we used (8.2) with \(m=1\) and \(m=2\) respectively. Performing the procedure in (8.4) repeatedly, we arrive at (8.1).

Now we give a proof of the lemma by the mathematical induction. The case \(N=1\) in is trivial. Suppose that it holds for \(N-1\). Then noticing

$$\begin{aligned} 1_{V_N}(\underline{x}_N)=\prod _{k=2}^N\prod _{j=2}^k1_{>0}\left( x^{(k-1)}_{j-1}-x^{(k)}_j\right) =1_{V_{N-1}}(\underline{x}_{N-1})\prod _{j=2}^N1_{>0}\left( x^{(N-1)}_{j-1}-x^{(N)}_j\right) , \end{aligned}$$
(8.5)

we see that LHS of (3.41) is written as

$$\begin{aligned}&\sum _{\sigma ^{(j)}\in S_j,j=1,\ldots ,N}\mathrm{{sgn}}\sigma ^{(N)}\, 1_{V_{N-1}}(\underline{x}_{N-1}^{\sigma }) \prod _{j=2}^N 1_{>0}\left( x^{(N-1)}_{\sigma ^{(N-1)}(j-1)}-x^{(N)}_{\sigma ^{(N)}(j)}\right) \nonumber \\&\qquad =\sum _{\sigma ^{(j)}\in S_j,j=1,\ldots ,N}\mathrm{{sgn}}\sigma ^{(N-1)}\, 1_{V_{N-1}}(\underline{x}_{N-1}^{\sigma })\cdot \mathrm{{sgn}}\sigma ^{(N)} \prod _{j=2}^N 1_{>0}\left( x^{(N-1)}_{j-1}-x^{(N)}_{\sigma ^{(N)}(j)}\right) \nonumber \\&\qquad =\sum _{\sigma ^{(j)}\in S_j,j=1,\ldots ,N}\mathrm{{sgn}}\sigma ^{(N-1)}\, 1_{\text {GT}}(\underline{x}_{N-1}^{\sigma })\cdot \mathrm{{sgn}}\sigma ^{(N)} \prod _{j=2}^N 1_{>0}\left( x^{(N-1)}_{j-1}-x^{(N)}_{\sigma ^{(N)}(j)}\right) \nonumber \\&\qquad =\sum _{\sigma ^{(j)}\in S_j,j=1,\ldots ,N}\mathrm{{sgn}}\sigma ^{(N)}\, 1_{\text {GT}}(\underline{x}_{N-1}^{\sigma }) \prod _{j=2}^N 1_{>0}\left( x^{(N-1)}_{\sigma ^{(N-1)}(j-1)}-x^{(N)}_{\sigma ^{(N)}(j)}\right) , \end{aligned}$$
(8.6)

where in the second equality we used the assumption for \(N-1\). Note that in the rightmost side of (8.6), the condition \(x^{(N-1)}_{\sigma ^{(N-1)}(1)}>x^{(N-1)}_{\sigma ^{(N-1)}(2)}>\cdots > x^{(N-1)}_{\sigma ^{(N-1)}(N-1)}\) holds for the support of \(1_{\text {GT}}(\underline{x}_{N-1}^{\sigma })\). Thus we can apply (8.1) to the rightmost side. We see that it becomes

$$\begin{aligned}&\sum _{\sigma ^{(j)}\in S_j,j=1,\ldots ,N}\mathrm{{sgn}}\sigma ^{(N)}\, 1_{\text {GT}}(\underline{x}_{N-1}^{\sigma })\nonumber \\&\qquad \times \prod _{j=2}^N 1_{>0}\left( x^{(N-1)}_{\sigma ^{(N-1)}(j-1)}-x^{(N)}_{\sigma ^{(N)}(j)}\right) 1_{>0}\left( x^{(N)}_{\sigma ^{(N)}(j-1)}-x^{(N-1)}_{\sigma ^{(N-1)}(j-1)}\right) \nonumber \\&\quad = \sum _{\sigma ^{(j)}\in S_j,j=1,\ldots ,N}\mathrm{{sgn}}\sigma ^{(N)} \, 1_\mathrm{{GT}}(\underline{x}_N^\sigma ), \end{aligned}$$
(8.7)

which completes the proof of Lemma 10. \(\square \)

Appendix 3: The Saddle Point Analysis of \(\psi _k(x;t)\)

In this Appendix, we give a proof of (5.42) based on the saddle point method in a similar way to Sect. 5.4.3 in [10]. Here we deal with the case of general Y while the case of \(Y=0\) was considered in [10]. We focus mainly on the limit about \(\psi _k(x;t)\) (2.1) in (5.42) since the case \(\phi _k(x;t)\) (4.1) can also be estimated in a parallel way. Changing the variable as \(w=-i\sqrt{N}z\), (2.1) becomes

$$\begin{aligned} \psi _k(x,t)=\frac{\sqrt{N}}{2\pi i}\int _{-i\infty }^{i\infty } dz\, e^{f_N(z;t,x)}, \end{aligned}$$
(9.1)

where

$$\begin{aligned} f_N(z;t,x)=-\sqrt{N}zx+N\frac{z^2t}{2}+(k-N)\log (\sqrt{N}z) -N\log \Gamma (\sqrt{N} z). \end{aligned}$$
(9.2)

Substituting (5.41) and (5.43) into (9.2), we arrange the first three terms in ascending order of powers of N as

$$\begin{aligned}&-\sqrt{N}zx_i+N\frac{z^2t}{2}+(k-N)\log (\sqrt{N}z)\nonumber \\&\quad =N^{3/2}\log N \cdot \frac{z}{2} +N^{3/2}\left( -z+\frac{T^{1/2}z^2}{2}-\frac{z}{2}\log T\right) +N\left( \gamma _T^2Yz^2-2\gamma _T^2YT^{-1/2}z -\frac{T^{1/2}z}{2}\right) \nonumber \\&\qquad +N^{1/2}\log N\cdot \frac{\lambda }{2(2\gamma _T)^{1/2}} +N^{1/2}\left( \frac{\gamma _T^3}{12}z-\gamma _T\xi _i z +\gamma _TY^2z-\gamma _T^2Yz +\frac{\lambda }{(2\gamma _T)^{1/2}}\log z\right) . \end{aligned}$$
(9.3)

Using the Stirling formula

$$\begin{aligned} \log \Gamma (n)=n\log n-n-\frac{\log 2\pi n}{2}+\frac{1}{12n}+O(n^{-3}) \end{aligned}$$
(9.4)

for the last term in (9.2), we have

$$\begin{aligned} -N\log \Gamma (\sqrt{N}z)&=-N^{3/2}\log N\cdot \frac{z}{2}+ N^{3/2}\left( z-z\log z\right) +\frac{N}{4}\log N+ N \frac{\log 2\pi z}{2}\nonumber \\&~~~~-N^{1/2}\frac{1}{12z}+O(N^{-1/2}). \end{aligned}$$
(9.5)

Thus from (9.3) and (9.5), \(f_N(z)\) (9.2) can be expressed as

$$\begin{aligned}&f_N(z;t,x)=N^{3/2}f(z)+Ng(z)+N^{1/2}h(z)+C_1+O(N^{-1/2}), \end{aligned}$$
(9.6)
$$\begin{aligned}&f(z)=\frac{T^{1/2}z^2}{2}-z\log z-\frac{z}{2}\log T, \end{aligned}$$
(9.7)
$$\begin{aligned}&g(z)=-\frac{T^{1/2}z}{2}+\frac{\log z}{2}+\gamma _T^2Yz^2-\frac{2\gamma _T^2Y}{T^{1/2}}z, \end{aligned}$$
(9.8)
$$\begin{aligned}&h(z)=-\frac{1}{12z}+\left( \frac{\gamma _T^3}{12}-\gamma _T^2Y-\gamma _T(\xi _i-Y^2) \right) z +\frac{\lambda }{(2\gamma _T)^{1/2}}\log z. \end{aligned}$$
(9.9)

Here \(C_1\), which does not depend on z is written as

$$\begin{aligned} C_1=N^{1/2}\log N\cdot \frac{\lambda }{2(2\gamma _T)^{1/2}}+\frac{N}{2}\log 2\pi \sqrt{N}. \end{aligned}$$
(9.10)

We note that f(z) above has a double saddle point \(z_c=T^{-1/2}\) such that \(f'(z_c)=f''(z_c)=0\). We expand f(z),  g(z),  h(z) around \(z_c\). Noting \(f'''(z_c)=2\gamma _T^3\), \(g'(z_c)=0,~g''(z_c)=2\gamma _T^2Y-\gamma _T^3\), \(h'(z_c)=\gamma _T^3/4-\gamma _T^2Y+\gamma _T(\lambda -\xi _i+Y^2)\), we get

$$\begin{aligned} N^{3/2}f(z)&=N^{3/2}f(z_c)+\frac{N^{3/2}}{3!}f'''(z_c)(z-z_3)^3+ O(N^{3/2}(z-z_c)^4)\nonumber \\&=C_2+N^{3/2}\frac{\gamma _T^3}{3}(z-z_c)^3+O(N^{3/2}(z-z_c)^4), \end{aligned}$$
(9.11)
$$\begin{aligned} Ng(z)&=Ng(z_c)+Ng'(z_c)(z-z_c)+N\frac{g''(z_c)}{2}(z-z_c)^2 +O(N(z-z_c)^3)\nonumber \\&=C_3-N\left( \frac{\gamma _T^3}{2}-\gamma _T^2Y\right) (z-z_c)^2+O(N(z-z_c)^3), \end{aligned}$$
(9.12)
$$\begin{aligned} N^{1/2}h(z)&=N^{1/2}h(z_c)+N^{1/2}h'(z_c)(z-z_c)+C_4+O(N^{1/2}(z-z_c)^2) \nonumber \\&=C_4+N^{1/2}\left( \frac{\gamma _T^3}{4}-\gamma _T^2Y+\gamma _T(\lambda -\xi +Y^2) \right) (z-z_c)+O(N^{1/2}(z-z_c)^2), \end{aligned}$$
(9.13)

where \(C_2,~C_3\) and \(C_4\) are

$$\begin{aligned}&C_2:=N^{3/2}f(z_c)=N^{3/2}T^{-1/2}/2,~~C_3:=Ng(z_c)=-N \left( \frac{1+\log T}{2}+\frac{Y}{2\gamma _T}\right) , \nonumber \\&C_4:=N^{1/2}h(z_c)=-N^{1/2}\left( \frac{T^{1/2}}{24}+(\gamma _T^2Y+\gamma _T(\xi -Y^2))T^{-1/2} -\frac{\lambda \log T^{-1/2}}{(2\gamma _T)^{1/2}} \right) . \end{aligned}$$
(9.14)

Thus from (9.6)–(9.13) and under the scaling \(\frac{z'}{\sqrt{N}}=(z-z_c)\), we have

$$\begin{aligned} f_N(z;t,x_i)=\frac{\gamma _T^3}{3} \left( z'+\frac{Y}{\gamma _T}-\frac{1}{2}\right) ^3-\gamma _T(\xi -\lambda ) \left( z'+\frac{Y}{\gamma _T}-\frac{1}{2}\right) +\sum _{j=1}^5C_j+O\left( N^{-1/2}\right) , \end{aligned}$$
(9.15)

where \(C_1,\ldots ,C_4\) are defined in (9.10) and (9.14) and \(C_5\) is

$$\begin{aligned} C_5=-\frac{1}{3}\left( Y-\frac{\gamma _T}{2}\right) ^3+\left( \xi -\lambda \right) \left( Y-\frac{\gamma _T}{2}\right) . \end{aligned}$$
(9.16)

Further changing the variable \(z'+Y/\gamma _T-1/2=-iv/\gamma _T\), we obtain

$$\begin{aligned}&f_N(z;t,x_i)=\frac{i}{3}v^3+i(\xi _i-\lambda )v +\sum _{j=1}^5C_j+O\left( N^{-1/2}\right) . \end{aligned}$$
(9.17)

Hence from (9.1) and (9.17), we get the limiting form of \(\psi _k(x_i;t)\)

$$\begin{aligned} e^{-\sum _{j=1}^5C_j}\gamma _T \psi _k(x_i;t)\sim \frac{1}{2\pi }\int _{-\infty }^{\infty }dv\,e^{\frac{i}{3}v^3+i(\xi _i-\lambda )v} =\mathrm{Ai}(\xi _i-\lambda ), \end{aligned}$$
(9.18)

which is nothing but (5.42).

As with (9.1), we rewrite \(\phi _k(x;t)\) (4.1) by the change of variable \(v=\sqrt{N}z\),

$$\begin{aligned} \phi _k(x;t)=\frac{1}{2\pi i}\oint dz\, \frac{e^{-f_N(z;t,x)}}{z}, \end{aligned}$$
(9.19)

where \(f_N(z;t,x)\) is given in (9.2). Applying the same techniques as above to this equation, we get the result for \(\phi _k(x;t)\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Imamura, T., Sasamoto, T. Determinantal Structures in the O’Connell-Yor Directed Random Polymer Model. J Stat Phys 163, 675–713 (2016). https://doi.org/10.1007/s10955-016-1492-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-016-1492-1

Keywords

Navigation