Skip to main content
Log in

Stationary Half-Space Last Passage Percolation

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

Abstract

In this paper we study stationary last passage percolation with exponential weights and in half-space geometry. We determine the limiting distribution of the last passage time in a critical window close to the origin. The result is a new two-parameter family of distributions: one parameter for the strength of the diagonal bounding the half-space (strength of the source at the origin in the equivalent TASEP language) and the other for the distance of the point of observation from the origin. It should be compared with the one-parameter family giving the Baik–Rains distributions for full-space geometry. We finally show that far enough away from the characteristic line, our distributions indeed converge to the Baik–Rains family. We derive our results using a related inhomogeneous integrable model having Pfaffian correlations, together with careful analytic continuation, and steepest descent analysis.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3

Similar content being viewed by others

Notes

  1. This holds true around points with smooth limit shape. Around shocks there are some differences, see e.g. [33, 35,36,37, 61].

  2. The random initial condition on the two sides is recovered using boundary sources, by the use of some Burke-type property [28, 32], as shown for the exclusion process in [66].

  3. The results are obtained through a description in terms of Pfaffian structures [8, 14, 16,17,18, 27, 43, 45, 70].

  4. A path of adjacent vertices with down steps \((-1, 0)\) or right steps (0, 1).

  5. The Baik–Rains distribution away from the characteristic line is given in [7], formula (3.35).

  6. Here we use notation as in [6, 40] with one exception: for the functions involved, we keep their contour integral representations instead of rewriting them in terms of (integrals of) Airy and exponential functions. The original expression in [7] looks quite different, but as is shown in [40, Appendix A], gives the same distribution.

  7. The reason for the minus sign in front is that our z contour is oriented downwards.

  8. Baik–Rains [7] treat the Poisson case instead of the exponential one but the shift argument is similar.

  9. Here \(\overline{{\mathsf {K}}}^{(m)} (x_i, x_j)\) is the \(2m \times 2m\) anti-symmetric matrix with \(2 \times 2\) block at (ij) given by \(\overline{{\mathsf {K}}} (x_i, x_j)\) for \(1 \le i, j \le m\).

  10. For more on point processes in general and determinantal ones in particular, see e.g.  [19, 26, 55].

References

  1. Aggarwal, A.: Current fluctuations of the stationary ASEP and six-vertex model. Duke Math. J. 167, 269–384 (2018)

    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)

    MathSciNet  MATH  Google Scholar 

  3. Baik, J., Barraquand, G., Corwin, I., Suidan, T.: Facilitated exclusion process. In: Celledoni, E., Di Nunno, G., Ebrahimi-Fard, K., Munthe-Kaas, H.Z. (eds.) Computation and Combinatorics in Dynamics, Stochastics and Control. Springer, Berlin (2018)

    MATH  Google Scholar 

  4. Baik, J., Barraquand, G., Corwin, I., Suidan, T.: Pfaffian Schur processes and last passage percolation in a half-quadrant. Ann. Probab. 46, 3015–3089 (2018)

    MathSciNet  MATH  Google Scholar 

  5. Baik, J., Deift, P.A., Johansson, K.: On the distribution of the length of the longest increasing subsequence of random permutations. J. Am. Math. Soc. 12, 1119–1178 (1999)

    MathSciNet  MATH  Google Scholar 

  6. Baik, J., Ferrari, P.L., Péché, S.: Limit process of stationary TASEP near the characteristic line. Commun. Pure Appl. Math. 63, 1017–1070 (2010)

    MathSciNet  MATH  Google Scholar 

  7. Baik, J., Rains, E.M.: Limiting distributions for a polynuclear growth model with external sources. J. Stat. Phys. 100, 523–542 (2000)

    MathSciNet  MATH  Google Scholar 

  8. Baik, J., Rains, E.M.: Algebraic aspects of increasing subsequences. Duke Math. J. 109, 1–65 (2001)

    MathSciNet  MATH  Google Scholar 

  9. Baik, J., Rains, E.M.: The asymptotics of monotone subsequences of involutions. Duke Math. J. 109, 205–281 (2001)

    MathSciNet  MATH  Google Scholar 

  10. Baik, J., Rains, E.M.: Symmetrized random permutations. In: Bleher, P. (ed.) Random Matrix Models and Their Applications, vol. 40, pp. 1–19. Cambridge University Press, Cambridge (2001)

    MATH  Google Scholar 

  11. Balázs, M., Cator, E., Seppäläinen, T.: Cube root fluctuations for the corner growth model associated to the exclusion process. Electron. J. Probab. 11, 1094–1132 (2006)

    MathSciNet  MATH  Google Scholar 

  12. Barraquand, G.: A phase transition for \(q\)-TASEP with a few slower particles. Stoch. Proc. Appl. 125, 2674–2699 (2015)

    MathSciNet  MATH  Google Scholar 

  13. Barraquand, G., Borodin, A., Corwin, I.: Half-space Macdonald processes (2018). arXiv:1802.08210

  14. Barraquand, G., Borodin, A., Corwin, I., Wheeler, M.: Stochastic six-vertex model in a half-quadrant and half-line open ASEP. Duke Math. J. 167, 2457–2529 (2018)

    MathSciNet  MATH  Google Scholar 

  15. van Beijeren, H., Kutner, R., Spohn, H.: Excess noise for driven diffusive systems. Phys. Rev. Lett. 54, 2026–2029 (1985)

    ADS  MathSciNet  Google Scholar 

  16. Betea, D., Bouttier, J., Nejjar, P., Vuletić, M.: The free boundary Schur process and applications I. Ann. Henri Poincaré 19, 3663–3742 (2018)

    ADS  MathSciNet  MATH  Google Scholar 

  17. Bisi, E., Zygouras, N.: Point-to-line polymers and orthogonal Whittaker functions. Trans. Amer. Math. Soc. 371, 8339–8379 (2019)

    MathSciNet  MATH  Google Scholar 

  18. Bisi, E., Zygouras, N.: Transition between characters of classical groups, decomposition of Gelfand–Tsetlin patterns and last passage percolation (2019). arXiv:1905.09756

  19. Borodin, A.: Determinantal point processes. In: Akemann, G., Baik, J., Di Francesco, P. (eds.) The Oxford Handbook of Random Matrix Theory, Chapter 11. Oxford University Press, Oxford (2010)

    Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  21. Borodin, A., Corwin, I., Ferrari, P.L., Vető, B.: Height fluctuations for the stationary KPZ equation. Math. Phys. Anal. Geom. 80, 20 (2015)

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

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

    ADS  MathSciNet  MATH  Google Scholar 

  24. 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)

    ADS  MathSciNet  MATH  Google Scholar 

  25. Borodin, A., Gorin, V.: Lectures on integrable probability (2012). arXiv:1212.3351

  26. Borodin, A., Okounkov, A., Olshanski, G.: Asymptotics of Plancherel measures for symmetric groups. J. Am. Math. Soc. 13, 481–515 (2000)

    MathSciNet  MATH  Google Scholar 

  27. Borodin, A., Rains, E.M.: Eynard–Mehta theorem, Schur process, and their Pfaffian analogs. J. Stat. Phys. 121, 291–317 (2006)

    ADS  MathSciNet  MATH  Google Scholar 

  28. Burke, P.J.: The output of a queuing system. Oper. Res. 4, 699–704 (1956)

    MathSciNet  MATH  Google Scholar 

  29. Chhita, S., Ferrari, P.L., Spohn, H.: Limit distributions for KPZ growth models with spatially homogeneous random initial conditions. Ann. Appl. Probab. 28, 1573–1603 (2018)

    MathSciNet  MATH  Google Scholar 

  30. Corwin, I.: The Kardar–Parisi–Zhang equation and universality class. Random Matrices Theory Appl. 01, 1130001 (2012)

    MathSciNet  MATH  Google Scholar 

  31. Corwin, I., Liu, Z., Wang, D.: Fluctuations of TASEP and LPP with general initial data. Ann. Appl. Probab. 26, 2030–2082 (2016)

    MathSciNet  MATH  Google Scholar 

  32. Darief, M., Mairesse, J., O’Connell, N.: Queues, stores, and tableaux. J. Appl. Probab. 4, 1145–1167 (2005)

    MathSciNet  MATH  Google Scholar 

  33. Ferrari, P.A., Fontes, L.: Shock fluctuations in the asymmetric simple exclusion process. Probab. Theory Relat. Fields 99, 305–319 (1994)

    MathSciNet  MATH  Google Scholar 

  34. Ferrari, P.L.: From interacting particle systems to random matrices. J. Stat. Mech. 2010, P10016 (2010)

    MathSciNet  Google Scholar 

  35. Ferrari, P.L., Ghosal, P., Nejjar, P.: Limit law of a second class particle in TASEP with non-random initial condition. Ann. Inst. Henri Poincaré Probab. Statist. 55, 1203–1225 (2019)

    ADS  MathSciNet  MATH  Google Scholar 

  36. Ferrari, P.L., Nejjar, P.: Anomalous shock fluctuations in TASEP and last passage percolation models. Probab. Theory Relat. Fields 161, 61–109 (2015)

    MathSciNet  MATH  Google Scholar 

  37. Ferrari, P.L., Nejjar, P.: Fluctuations of the competition interface in presence of shocks. ALEA Lat. Am. J. Probab. Math. Stat. 14, 299–325 (2017)

    MathSciNet  MATH  Google Scholar 

  38. Ferrari, P.L., Occelli, A.: Universality of the GOE Tracy–Widom distribution for TASEP with arbitrary particle density. Electron. J. Probab. 23(51), 1–24 (2018)

    MathSciNet  MATH  Google Scholar 

  39. Ferrari, P.L., Occelli, A.: Time-time covariance for last passage percolation with generic initial profile. Math. Phys. Anal. Geom. 22, 1 (2019)

    MathSciNet  MATH  Google Scholar 

  40. Ferrari, P.L., Spohn, H.: Scaling limit for the space-time covariance of the stationary totally asymmetric simple exclusion process. Commun. Math. Phys. 265, 1–44 (2006)

    ADS  MathSciNet  MATH  Google Scholar 

  41. Ferrari, P.L., Spohn, H.: Random growth models. In: Akemann, G., Baik, J., Di Francesco, P. (eds.) The Oxford Handbook of Random Matrix Theory, pp. 782–801. Oxford University Press, Oxford (2011)

    MATH  Google Scholar 

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

    ADS  MathSciNet  MATH  Google Scholar 

  43. Forrester, P.J., Rains, E.M.: Symmetrized models of last passage percolation and non-intersecting lattice paths. J. Stat. Phys. 129, 833–855 (2007)

    ADS  MathSciNet  MATH  Google Scholar 

  44. Forster, D., Nelson, D.R., Stephen, M.J.: Large-distance and long-time properties of a randomly stirred fluid. Phys. Rev. A 16, 732–749 (1977)

    ADS  MathSciNet  Google Scholar 

  45. Ghosal, P.: Correlation functions of the Pfaffian Schur process using Macdonald difference operators. SIGMA 15, 092 (2019)

    MathSciNet  MATH  Google Scholar 

  46. 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)

    MathSciNet  MATH  Google Scholar 

  47. Grosskinsky, S.: Phase transitions in nonequilibrium stochastic particle systems with local conservation laws. Ph.D. Thesis, Technische Universität München (2004). http://mediatum.ub.tum.de/doc/602023/document.pdf

  48. Imamura, T., Muccioconi, M., Sasamoto, T.: Stationary stochastic higher spin six vertex model and q-Whittaker measure (2019). arXiv:1901.08381

  49. Imamura, T., Sasamoto, T.: Fluctuations of the one-dimensional polynuclear growth model with external sources. Nucl. Phys. B 699, 503–544 (2004)

    ADS  MathSciNet  MATH  Google Scholar 

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

    ADS  MathSciNet  MATH  Google Scholar 

  51. Imamura, T., Sasamoto, T.: Free energy distribution of the stationary O’Connell–Yor directed random polymer model. J. Phys. A Math. Theor. 50, 285203 (2017)

    MathSciNet  MATH  Google Scholar 

  52. Imamura, T., Sasamoto, T.: Fluctuations for stationary \(q\)-TASEP. Probab. Theory Relat. Fields 174, 647–730 (2019)

    MathSciNet  MATH  Google Scholar 

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

    ADS  MathSciNet  MATH  Google Scholar 

  54. Johansson, K.: Discrete polynuclear growth and determinantal processes. Commun. Math. Phys. 242, 277–329 (2003)

    ADS  MathSciNet  MATH  Google Scholar 

  55. Johansson, K.: Random matrices and determinantal processes. In: Bovier, A., Dunlop, F., van Enter, A., den Hollander, F., Dalibard, J. (eds.) Mathematical Statistical Physics, Session LXXXIII: Lecture Notes of the Les Houches Summer School 2005, pp. 1–56. Elsevier, Amsterdam (2006)

    Google Scholar 

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

    ADS  MATH  Google Scholar 

  57. Krajenbrink, A., Le Doussal, P.: Replica Bethe Ansatz solution to the Kardar–Parisi–Zhang equation on the half-line (2019). arXiv:1905.05718

  58. Liggett, T.M.: Ergodic theorems for the asymmetric simple exclusion process. Trans. Am. Math. Soc. 213, 237–261 (1975)

    MathSciNet  MATH  Google Scholar 

  59. Liggett, T.M.: Ergodic theorems for the asymmetric simple exclusion process II. Ann. Probab. 4, 795–801 (1977)

    MathSciNet  MATH  Google Scholar 

  60. Liggett, T.M.: Stochastic Interacting Systems: Contact, Voter and Exclusion Processes. Springer, Berlin (1999)

    MATH  Google Scholar 

  61. Nejjar, P.: Transition to shocks in TASEP and decoupling of last passage times. ALEA Lat. Am. J. Probab. Math. Stat. 15, 1311–1334 (2018)

    MathSciNet  MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  63. Prähofer, M., Spohn, H.: Universal distributions for growth processes in 1 + 1 dimensions and random matrices. Phys. Rev. Lett. 84, 4882–4885 (2000)

    ADS  Google Scholar 

  64. Prähofer, M., Spohn, H.: Current fluctuations for the totally asymmetric simple exclusion process. In: Sidoravicius, V. (ed.) In and Out of Equilibrium. Progress in Probability. Birkhäuser, Basel (2002)

    MATH  Google Scholar 

  65. Prähofer, M., Spohn, H.: Scale invariance of the PNG droplet and the Airy process. J. Stat. Phys. 108, 1071–1106 (2002)

    MathSciNet  MATH  Google Scholar 

  66. Prähofer, M., Spohn, H.: Exact scaling function for one-dimensional stationary KPZ growth. J. Stat. Phys. 115, 255–279 (2004)

    ADS  MathSciNet  MATH  Google Scholar 

  67. Quastel, J.: Introduction to KPZ. Curr. Dev. Math. 2011, 125–194 (2011)

    MATH  Google Scholar 

  68. Quastel, J., Remenik, D.: How flat is flat in a random interface growth? Trans. Amer. Math. Soc. 371, 6047–6085 (2019)

    MathSciNet  MATH  Google Scholar 

  69. Quastel, J., Spohn, H.: The one-dimensional KPZ equation and its universality class. J. Stat. Phys. 160, 965–984 (2015)

    ADS  MathSciNet  MATH  Google Scholar 

  70. Rains, E.M.: Correlation functions for symmetrized increasing subsequences (2000). arXiv:math.CO/0006097

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

    ADS  MathSciNet  Google Scholar 

  72. Sasamoto, T., Imamura, T.: Fluctuations of a one-dimensional polynuclear growth model in a half space. J. Stat. Phys. 115, 749–803 (2004)

    ADS  MathSciNet  MATH  Google Scholar 

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

    ADS  MathSciNet  MATH  Google Scholar 

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

    ADS  Google Scholar 

  75. Stembridge, J.R.: Nonintersecting paths, Pfaffians, and plane partitions. Adv. Math. 83, 96–131 (1990)

    MathSciNet  MATH  Google Scholar 

  76. Takeuchi, K.A.: An appetizer to modern developments on the Kardar–Parisi–Zhang universality class. Physica A 504, 77–105 (2016)

    ADS  MathSciNet  Google Scholar 

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

    ADS  MathSciNet  MATH  Google Scholar 

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

    ADS  MathSciNet  MATH  Google Scholar 

  79. Zygouras, N.: Some algebraic structures in the KPZ universality (2018). arXiv:1812.07204

Download references

Acknowledgements

This work is supported by the German Research Foundation through the Collaborative Research Center 1060 “The Mathematics of Emergent Effects”, Project B04, and by the Deutsche Forschungs-gemeinschaft (DFG, German Research Foundation) under Germany’s Excellence Strategy—GZ 2047/1, Projekt ID 390685813. The authors would furthermore like to thank G. Barraquand, E. Bisi, J. Bouttier, P. Le Doussal, E. Emrah, S. Grosskinsky, P. Nejjar, T. Sasamoto, H. Spohn, B. Vető, and N. Zygouras for useful comments and suggestions regarding this manuscript and/or their related papers. We are also very grateful to an anonymous referee for careful reading of our manuscript and a number of constructive remarks.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Patrik L. Ferrari.

Additional information

Communicated by A. Borodin

Publisher's Note

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

Appendices

On Pfaffians and Point Processes

Given an anti-symmetric \(2n \times 2n\) matrix \((a_{i,j})\), its Pfaffian is defined as:

$$\begin{aligned} \mathrm {pf}[a_{i,j}]_{1\le i < j\le 2n} = \frac{1}{2^{n} n!}\sum _{\sigma \in S_{2n}}{{\,\mathrm{sgn}\,}}(\sigma )a_{\sigma (1),\sigma (2)} a_{\sigma (3),\sigma (4)} \cdots a_{\sigma (2n-1),\sigma (2n)} \end{aligned}$$
(A.1)

where \(S_{2n}\) is the permutation group on 2n letters. Remark that the Pfaffian is determined entirely by the upper triangular part. One can show that

$$\begin{aligned} \left( \mathrm {pf}[a_{i,j}]_{1\le i < j\le 2n}\right) ^2=\det [a_{i,j}]_{1\le i,j\le 2n}. \end{aligned}$$
(A.2)

Suppose that one has a \(2 \times 2\) anti-symmetric matrix kernel K(xy), i.e. K is a \(2 \times 2\) matrix function of (xy) satisfying \(K(x, y) = -K^t(y, x)\) with t denoting transposition. Notice the interchange of x and y. Given such a kernel and points (variables or numbers) \(x_1, \dots , x_n\), one can define a \(2n \times 2n\) anti-symmetric matrix \(K^{(n)}\) block-wise as follows: its \(2 \times 2\) block (ij) for \(1 \le i, j \le n\) is given by the matrix \(K(x_i, x_j)\). Because \(K(x, y) = -K^t(y, x)\), \(K^{(n)}\) thus defined is even-dimensional anti-symmetric and its Pfaffian well-defined.

Throughout we are interested in probability distributions on configuration spaces \(\varLambda \) equipped with a measure \(d \mu \) as follows: either \(\varLambda = {\mathbb {Z}}\) (or a semi-infinite interval in \({\mathbb {Z}}\)) is discrete and \(d \mu \) is counting measure, or \(\varLambda = {\mathbb {R}}\) (or a semi-infinite interval in \({\mathbb {R}}\)) is continuous and \(d \mu = d x\) is Lebesgue measure. A point process (measure)Footnote 10 on such a configuration space \(\varLambda \) is called Pfaffian with \(2 \times 2\) matrix correlation kernel K if there exists a \(2 \times 2\) matrix kernel K satisfying \(K(x, y) = -K^t(y, x)\) such that the n-point correlation functions \(\rho _n(x_1, x_2, \dots , x_n)\) of the process, for all \(n \ge 1\), are Pfaffians of the associated \(2n \times 2n\) matrix \(K^{(n)}\):

$$\begin{aligned} \rho _n(x_1, x_2, \dots , x_n) = \mathrm {pf}[K^{(n)} (x_i, x_j)]_{1\le i<j\le n}. \end{aligned}$$
(A.3)

We recall that if \(\varLambda \) is discrete, \(\rho _n(x_1, x_2, \dots , x_n) = {\mathbb {P}}(S : x_i \in S,\ \forall \,1 \le i \le n)\) is the probability of S containing all the x’s; if \(\varLambda \) is continuous, \(\rho _n(x_1, x_2, \dots , x_n) d x_1 \dots d x_n\) is the probability of sets S having non-empty intersection with all \([x_i, x_i + d x_i]\). A simple observation states that the one-point function is the \(K_{12}\) entry: \({\mathbb {P}}(S:x \in S) = \rho _1(x) = K_{12}(x, x)\) (in the discrete setting; the obvious modification is needed in the continuous setting).

Given a \(2 \times 2\) anti-symmetric matrix kernel K defined on a configuration space \(\varLambda \) equipped with a measure \(d \mu \) (either counting measure or Lebesgue measure depending on the underlying space), the Fredholm Pfaffian of K restricted to the subspace \(U \subset \varLambda \) is defined as

$$\begin{aligned} \mathrm {pf}(J + \lambda K)_{L^2(U) \times L^2(U)} = \sum _{n=0}^{\infty }\frac{\lambda ^n}{n!} \int _{U^n} \mathrm {pf}[ K^{(n)} (x_i, x_j)]_{1\le i<j\le n} \prod _{i=1}^n d \mu (x_i) \end{aligned}$$
(A.4)

assuming the sum is finite. Here J is the anti-symmetric matrix kernel \(J(x,y)=\delta _{x,y} \left( {\begin{matrix} 0 &{} 1 \\ -1 &{} 0 \end{matrix}} \right) \), but as is oftentimes the case in the literature, this technicality is overlooked and we think of J just as the corresponding \(2 \times 2\) matrix. We note that oftentimes a sufficient condition for the Fredholm Pfaffian \(\mathrm {pf}(J+\lambda K)_{L^2(U) \times L^2(U)}\) to be finite finite is that K is a trace class operator on \(L^2(U) \times L^2(U)\) (or rather it has trace class entries). Moreover, Fredholm Pfaffians are defined up to conjugation, in the following sense. Suppose \({\tilde{K}}\) is the anti-symmetric matrix kernel

$$\begin{aligned} {\tilde{K}}(x,y)= \left( \begin{matrix} e^{f(x)} &{} \quad 0 \\ 0 &{}\quad e^{-f(x)} \end{matrix}\right) K(x,y) \left( \begin{matrix} e^{f(y)} &{}\quad 0 \\ 0 &{} \quad e^{-f(y)} \end{matrix}\right) \end{aligned}$$
(A.5)

for a \(d\mu \)-measurable function f. Then it is not hard to check that \(\mathrm {pf}[K^{(n)} (x_i, x_j)]_{1\le i<j\le 2n} = \mathrm {pf}[{\tilde{K}}^{(n)} (x_i, x_j)]_{1\le i<j\le 2n}\) and so \(\mathrm {pf}(J+\lambda K) = \mathrm {pf}(J+\lambda {\tilde{K}})\) whenever both are defined. Importantly, we can use this to define \(\mathrm {pf}(J+\lambda K)\) even if K is not a trace class operator provided we find an appropriate f which makes \({\tilde{K}}\) trace class.

Owing to identity (A.2), we have the following relation between Fredholm Pfaffians with \(2 \times 2\) matrix kernels K and block Fredholm determinants with related kernel \(J^{-1}K\):

$$\begin{aligned} \mathrm {pf}(J + \lambda K)^2_{L^2(U) \times L^2(U)} = \det (\mathbb {1}+ \lambda J^{-1} K)_{L^2(U) \times L^2(U)} \end{aligned}$$
(A.6)

where we remark the Fredholm determinant on the right hand side is defined as in (A.4) with pf replaced by det and \(K^{(n)}\) by \((J^{-1}K)^{(n)}\).

Finally, for any two bounded \(2 \times 2\) matrix operators \(A:L^2(U) \times L^2(U) \rightarrow L^2(V) \times L^2(V)\), \(B:L^2(V) \times L^2(V) \rightarrow L^2(U) \times L^2(U)\) for which both sides below are defined, we have

$$\begin{aligned} \mathrm {pf}(J+\lambda JAB)_{L^2(V) \times L^2(V)}=\mathrm {pf}(J+\lambda JBA)_{L^2(U) \times L^2(U)}. \end{aligned}$$
(A.7)

For example, this could happen if the entries of either A or B are trace class (as such operators form an ideal inside the bounded ones), or if the entries of both A and B are Hilbert–Schmidt (as products of two Hilbert–Schmidt operators are trace class), but it could happen in more generality too. We note that oftentimes AB is infinite dimensional while BA is finite dimensional. This fact is immediately implied by the corresponding Fredholm determinant identity \(\det (\mathbb {1}+ \lambda AB) =\det (\mathbb {1}+ \lambda BA)\) and (A.6).

For more on Pfaffians, the reader is referred to the the Appendix of [62] for the analytic side and to [75] for the algebraic and combinatorial ones.

Some Determinantal Computations

In this section we prove (3.23). For this we need to show that

$$\begin{aligned} \left\langle Y_1 \left| Z_2 \right\rangle \right. = \left\langle Y_2 \left| Z_1 \right\rangle \right. =0 \text { and } \left\langle Y_1 \left| Z_1 \right\rangle \right. = \left\langle Y_2 \left| Z_2 \right\rangle \right. \end{aligned}$$
(B.1)

where \(Z_i, Y_i, i=1,2\) are defined in (3.18). For brevity and consistency, let us rename

$$\begin{aligned} A=(\mathbb {1}-{{\overline{G}}})^{-1}, \quad h_1= g_1, \quad h_2=g_2, \quad f_1 = f_{+}^\beta . \end{aligned}$$
(B.2)

We have

$$\begin{aligned} \left\langle Y_1 \left| Z_2 \right\rangle \right. = \left\langle f_1 \left| A_{12} f_1 \right\rangle \right. \end{aligned}$$
(B.3)

and

$$\begin{aligned} \left\langle Y_2 \left| Z_1 \right\rangle \right. = \left\langle -h_1 \quad h_2 \left| \begin{array}{c} A_{11}h_2+A_{12}h_1 \\ A_{21}h_2+A_{22} h_1 \end{array} \right\rangle \right. . \end{aligned}$$
(B.4)

To show that both are zero, we need the following result.

Proposition 39

We have:

$$\begin{aligned} \begin{aligned} A_{12}(x,y)=-A_{12}(y,x),\quad A_{21}(x,y)=-A_{21}(y,x),\quad A_{11}(x,y)=A_{22}(y,x). \end{aligned} \end{aligned}$$
(B.5)

Proof

Assume for now that we know the norm of \({\overline{G}}\) is less than 1. Then we can expand A as a Neumann series \((\mathbb {1}-\overline{G})^{-1}=\mathbb {1}+{{\overline{G}}}+{{\overline{G}}}^2+{{\overline{G}}}^3+\cdots \). We will prove by induction that, for any \(n\ge 1\)

$$\begin{aligned} \begin{aligned} {{\overline{G}}}^n_{12}(x,y)=-{{\overline{G}}}^n_{12}(y,x),\quad \overline{G}^n_{21}(x,y)=-{{\overline{G}}}^n_{21}(y,x),\quad \overline{G}^n_{11}(x,y)={{\overline{G}}}^n_{22}(y, x). \end{aligned} \end{aligned}$$
(B.6)

The base case is true by the definition of \({{\overline{G}}}=J^{-1}{{\overline{K}}}\) and by the fact that

$$\begin{aligned} \begin{aligned} {{\overline{K}}}_{12}(x,y)=-{{\overline{K}}}_{21}(y,x),\quad {{\overline{K}}}_{11}(x,y) =-{{\overline{K}}}_{11}(y,x),\quad {{\overline{K}}}_{22}(x,y)=-{{\overline{K}}}_{22}(y,x). \end{aligned} \end{aligned}$$
(B.7)

Now we proceed with the inductive step: assuming that (B.6) holds for \(B={{\overline{G}}}^n\), for \(C={{\overline{G}}}^{n+1}\) we have

$$\begin{aligned} \begin{aligned} C_{11}(x,y)&=\left( B_{11}{{\overline{G}}}_{11}+B_{12}{{\overline{G}}}_{21}\right) (x,y)\\&=\int _s^{\infty } dz B_{11}(x,z)\overline{G}_{11}(z,y)+\int _s^{\infty } dz B_{12}(x,z){{\overline{G}}}_{21}(z,y)\\&=\int _s^{\infty } dz B_{22}(z,x)\overline{G}_{22}(y,z)+\int _s^{\infty } dz B_{12}(z,x){{\overline{G}}}_{21}(y,z)\\&=C_{22}(y,x). \end{aligned} \end{aligned}$$
(B.8)

Moreover,

$$\begin{aligned} \begin{aligned} C_{12}(x,y)&=\left( B_{11}{{\overline{G}}}_{12}+B_{12}{{\overline{G}}}_{22}\right) (x,y)\\&=\int _s^{\infty } dz B_{11}(x,z)\overline{G}_{12}(z,y)+\int _s^{\infty } dz B_{12}(x,z){{\overline{G}}}_{22}(z,y)\\&=-\int _s^{\infty } dz B_{22}(z,x)\overline{G}_{12}(y,z)-\int _s^{\infty } dz B_{12}(z,x){{\overline{G}}}_{11}(y,z)\\&=-C_{12}(y,x) \end{aligned} \end{aligned}$$
(B.9)

where we have used that \(C = B {\overline{G}} = {\overline{G}} B\). An analogous computation holds for \(C_{21}\).

Now it may happen that the norm of \({\overline{G}}\), of which we remain ignorant, is \(\ge 1\). We then replace \({\overline{G}}\) by \(\omega {\overline{G}}\) for \(\omega >0\) small enough to make \(\omega {\overline{G}}\) of subunit norm. The argument above applies to the corresponding \(A(\omega ) = (\mathbb {1}-\omega {\overline{G}})^{-1}\). We then analytically continue in \(\omega \), as the spectrum of any trace class operator (and in particular of \({\overline{G}}\)) is discrete without non-zero accumulation points. \(\quad \square \)

Using Proposition 39 in Eq. (B.3), we have

$$\begin{aligned} \begin{aligned} \left\langle Y_1 \left| Z_2 \right\rangle \right.&=\int _s^{\infty } \int _s^{\infty } dx dy f_1(x)A_{12}(x,y)f_1(y)\\&=-\int _s^{\infty } \int _s^{\infty } dx dy f_1(x)A_{12}(y,x)f_1(y) \\&=0. \end{aligned} \end{aligned}$$
(B.10)

For the same reason, (B.4) becomes:

$$\begin{aligned} \left\langle Y_1 \left| Z_2 \right\rangle \right. = -\left\langle h_1 \left| A_{11}h_2 \right\rangle \right. - \left\langle h_1 \left| A_{12}h_1 \right\rangle \right. + \left\langle h_2 \left| A_{22}h_1 \right\rangle \right. + \left\langle h_2 \left| A_{21}h_2 \right\rangle \right. = 0 \end{aligned}$$
(B.11)

since the last two terms are zero, and the first two are equal by Proposition 39:

$$\begin{aligned} \begin{aligned} \left\langle h_2 \left| A_{22}h_1 \right\rangle \right.&= \int _s^{\infty } \int _s^{\infty } dx dy h_2(x)A_{22}(x,y)h_1(y)\\&=\int _s^{\infty } \int _s^{\infty } dx dy h_2(x)A_{11}(y,x)h_1(y)\\&=\left\langle h_1 \left| A_{11}h_2 \right\rangle \right. . \end{aligned} \end{aligned}$$
(B.12)

Now, we show that \(\left\langle Y_1 \left| Z_1 \right\rangle \right. = \left\langle Y_2 \left| Z_2 \right\rangle \right. \). The two are explicitly given by

$$\begin{aligned} \left\langle Y_1 \left| Z_1 \right\rangle \right. = \left\langle f_1 \left| A_{11} h_2 \right\rangle \right. + \left\langle f_1 \left| A_{12}h_1 \right\rangle \right. , \quad \left\langle Y_2 \left| Z_2 \right\rangle \right. = -\left\langle h_1 \left| A_{12} f_1 \right\rangle \right. + \left\langle h_2 \left| A_{22}f_1 \right\rangle \right. . \end{aligned}$$
(B.13)

By Proposition 39, as \(A_{11}(x, y) = A_{22}(y, x)\) and \(A_{12}(x, y) = - A_{12}(y, x)\), it follows that \(\left\langle f_1 \left| A_{11} h_2 \right\rangle \right. = \left\langle h_2 \left| A_{22} f_1 \right\rangle \right. \) and that \(\left\langle f_1 \left| A_{12} h_1 \right\rangle \right. = -\left\langle h_1 \left| A_{12} f_1 \right\rangle \right. \) proving the desired equality.

Correlations for Geometric Weights

1.1 Correlation kernel for generic geometric weights

In this section we briefly review the main result on integrable last passage percolation with i.i.d. geometric random variables. It is by passing to the exponential limit in this result that we obtain Theorem 13. Consider half-space last passage percolation with i.i.d.  geometric random variables \((W_{i,j})_{1 \le j \le i \le N}\) distributed as follows:

$$\begin{aligned} W_{i,j} = {\left\{ \begin{array}{ll} \mathrm {Geom}(a x_i), &{} i = j, \\ \mathrm {Geom}(x_i x_j), &{} i > j \end{array}\right. } \end{aligned}$$
(C.1)

where \(a, x_1, \dots , x_N\) are real parameters satisfying

$$\begin{aligned} 0 \le a< \min _i \tfrac{1}{x_i}, \quad 0< x_1, \dots , x_N < 1 \end{aligned}$$
(C.2)

and a random variable X is geometric\(\mathrm {Geom}(q)\) if \({\mathbb {P}}(X = k) = (1-q)q^k, \forall k \in {\mathbb {N}}\). We depict this in Fig. 4. It is helpful to visualize the x’s as indexing the rows and columns of the half-space.

Fig. 4
figure 4

A possible LPP path (polymer) from (1, 1) to \((N, N-n)\) for \((N, n) = (8, 2)\). The non-diagonal dots are independent geometric random variables with parameters assigned by multiplying the row and column x parameters; the diagonal has an extra parameter a

We have the following result for the last passage percolation time \(L_{N,N-n}\) from (1, 1) to \((N,N-n)\) (for fixed \(0 \le n \le N-1\)). The precise statement as stated below is first rigorously proven and explicitly stated in [4]. Previously the kernel was derived in [72], but not in a completely rigorous way, specifically for the \(n=0\) case. The Pfaffian structure for \(n=0\) was first derived by Rains [70], and subsequently extended for generic n and/or as multi-point joint distribution in [4, 16, 27, 45, 72]. See also the pioneering algebraic work of Baik–Rains [8, 10] for an alternative but equivalent approach via Toeplitz+Hankel determinants and matrix integrals, as well as the more combinatorial approach of Forrester–Rains [43]. We follow [16] for the exposition hereinafter.

Theorem 40

The distribution of \(L_{N,N-n}\) is a Fredholm Pfaffian

$$\begin{aligned} {\mathbb {P}}(L_{N,N-n} \le s) = \mathrm {pf}(J-K^{\mathrm{geo}})_{\ell ^2 (\{s+1, s+2,\dots \})} \end{aligned}$$
(C.3)

with \(2 \times 2\) matrix correlation kernel \(K^{\mathrm{geo}}: {\mathbb {Z}}^2 \rightarrow \mathrm {Mat}_2({\mathbb {R}})\) given by

$$\begin{aligned} \begin{aligned} K^{\mathrm{geo}}_{11}(k, \ell )&= \frac{1}{(2\pi \mathrm{i})^2}\int \frac{dz}{z^{k}} \int \frac{dw}{w^{\ell }} F(z) F(w) \frac{(z-w) (z-a) (w-a)}{(z^2-1) (w^2-1) (zw-1)}, \\ K^{\mathrm{geo}}_{12}(k, \ell ) = -K^{\mathrm{geo}}_{21}(\ell , k)&= \frac{1}{(2\pi \mathrm{i})^2}\int \frac{dz}{z^{k}} \int \frac{dw}{w^{-\ell +1}} \frac{F(z)}{F(w)} \frac{(zw-1) (z-a)}{(z-w) (z^2-1) (w-a)}, \\ K^{\mathrm{geo}}_{22}(k, \ell )&= \frac{1}{(2\pi \mathrm{i})^2}\int \frac{dz}{z^{-k+1}} \int \frac{dw}{w^{-\ell +1}} \frac{1}{F(z) F(w)} \frac{(z-w)}{ (zw-1) (z-a) (w-a)} \end{aligned} \end{aligned}$$
(C.4)

where

$$\begin{aligned} F(z)= \frac{\prod _{k=1}^N (1-x_k/z)}{\prod _{k=1}^{N-n} (1-x_k z)} \end{aligned}$$
(C.5)

and where the contours are positively oriented circles centered around the origin satisfying the following conditions:

  • for \(K^{\mathrm{geo}}_{11}\), \(1< |z|, |w| < \min _i \frac{1}{x_i}\);

  • for \(K^{\mathrm{geo}}_{12}\), \(\max \{ \max _i x_i, a\}< |w| < |z|\) and \(1< |z| < \min _i \frac{1}{x_i}\);

  • for \(K^{\mathrm{geo}}_{22}\), \(\max \{ \max _i x_i, a\} < |w|, |z|\) and \(1 < |zw|\).

1.2 Reformulation of the correlation kernel for our model

Now consider the case \(x_1=b\in (0,1)\), \(x_2=x_3=\ldots =x_N=\sqrt{q}\) for some \(q\in (0,1)\) and \(a>0\) such that \(a\sqrt{q}<1\) and \(ab<1\). We will manipulate the kernel so that the limit to the exponential case becomes straightforward. In particular, it is not necessary to do any steepest descent analysis (Laplace method) as was done in [4].

Let us define the following functions:

$$\begin{aligned} H(z)=\frac{(1-\sqrt{q}/z)^{N-1}}{(1-\sqrt{q}z)^{N-1}},\quad B(z)= \frac{1-b/z}{1-bz} \end{aligned}$$
(C.6)

which when combined yield

$$\begin{aligned} F(z)=H(z) B(z) (1-\sqrt{q} z)^n. \end{aligned}$$
(C.7)

Notice that

$$\begin{aligned} H(z^{-1}) = H(z)^{-1},\quad B(z^{-1})=B(z)^{-1},\quad F(z^{-1})=H(z)^{-1} B(z)^{-1} (1-\sqrt{q}/z)^n. \end{aligned}$$
(C.8)

Since in the Fredholm Pfaffian (C.3) the smallest value to be considered is \(s=0\), the \(k,\ell \) entries of the kernel will be in \(\{1,2,3,\ldots \}\) only. Below we only consider the case \(b\ne \sqrt{q}\), since the \(b=\sqrt{q}\) case is simpler: just set \(B(z)=0\) and replace \(N-1\) with N in H(z). To see this, take for instance the \(b\rightarrow \sqrt{q}\) limit in (C.13).

Lemma 41

Let \(b\ne \sqrt{q}\) and \(k,\ell \ge 1\). The kernel entry \(K^\mathrm{geo}_{11}\) can be expressed as

$$\begin{aligned} \begin{aligned} K^{\mathrm{geo}}_{11}(k,\ell ) =\,&\frac{-1}{(2\pi \mathrm{i})^2}\oint \limits _{\varGamma _{\sqrt{q}}} dw\!\!\! \oint \limits _{\varGamma _{1/\sqrt{q}}} dz \frac{w^{\ell -1}}{z^{k}}\frac{h^{\mathrm{geo}}_{11}(z,w)}{z-w} \\&+ \frac{-1}{(2\pi \mathrm{i})^2}\oint \limits _{\varGamma _{b}} dw \oint \limits _{\varGamma _{1/\sqrt{q}}} dz \frac{w^{\ell -1}}{z^{k}}\frac{h^{\mathrm{geo}}_{11}(z,w)}{z-w}\\&+\frac{-1}{(2\pi \mathrm{i})^2}\oint \limits _{\varGamma _{\sqrt{q}}} dw \oint \limits _{\varGamma _{1/b}} dz \frac{w^{\ell -1}}{z^{k}}\frac{h^\mathrm{geo}_{11}(z,w)}{z-w} \end{aligned} \end{aligned}$$
(C.9)

with

$$\begin{aligned} h^{\mathrm{geo}}_{11}(z,w)=\frac{H(z)B(z)}{H(w)B(w)}(1-\sqrt{q} z)^n(1-\sqrt{q}/w)^n \frac{(zw-1)(z-a)(1-w a)}{(z^2-1)(1-w^2)}. \end{aligned}$$
(C.10)

Proof

Let us start with

$$\begin{aligned} \begin{aligned} K^{\mathrm{geo}}_{11}(k,\ell )&= \oint \frac{dw}{2\pi \mathrm{i}} \oint \frac{dz}{2\pi \mathrm{i}} \frac{H(z)(1-\sqrt{q} z)^nB(z) \, H(w) (1-\sqrt{q}w)^n B(w)}{z^{k}w^{\ell }}\\&\quad \times \frac{(z-w)(z-a)(w-a)}{(z^2-1)(w^2-1)(zw-1)} \end{aligned} \end{aligned}$$
(C.11)

where the integrals can be taken over circles with radii larger than 1 and smaller than \(\min \{1/b,1/\sqrt{q}\}\). The change of variable \(w\rightarrow w^{-1}\) leads to

$$\begin{aligned} K^{\mathrm{geo}}_{11}(k,\ell ) =\frac{1}{(2\pi \mathrm{i})^2}\oint dw \oint dz \frac{w^{\ell -1}}{z^{k}}\frac{h^{\mathrm{geo}}_{11}(z,w)}{z-w} \end{aligned}$$
(C.12)

where we can take \(|w|\in (\max \{\sqrt{q},b\},1)\) and \(|z|\in (1,\min \{1/b,1/\sqrt{q}\})\).

The integrand behaves like \(w^{N-n+\ell -1}\) as \(w\rightarrow 0\) and thus there is no pole at \(w=0\) for any \(\ell \ge 1\) and \(n\le N\). Also, it behaves like \(1/(z^{N-n+k+1})\lesssim z^{-2}\) as \(z\rightarrow \infty \) for all \(n\le N\) and \(k\ge 1\). Thus there is no pole in z at \(\infty \). This means that we can deform the integration contours to include only the poles at b and \(\sqrt{q}\). Moreover, we can open up the z contour and close it down around the poles at 1/b and \(1/\sqrt{q}\). By doing this we change of orientation of the contour, which then gives a minus sign as all the closed contours we consider as anticlockwise oriented. Thus we get

$$\begin{aligned} K^{\mathrm{geo}}_{11}(k,\ell ) =\frac{-1}{(2\pi \mathrm{i})^2}\oint \limits _{\varGamma _{b,\sqrt{q}}} dw \oint \limits _{\varGamma _{1/b,1/\sqrt{q}}} dz \frac{w^{\ell -1}}{z^{k}}\frac{h^{\mathrm{geo}}_{11}(z,w)}{z-w} \end{aligned}$$
(C.13)

The claimed result follows from the observation that

$$\begin{aligned} \oint \limits _{\varGamma _b} dw \oint \limits _{\varGamma _{1/b}}dz \frac{w^{\ell -1}}{z^{k}}\frac{h^{\mathrm{geo}}_{11}(z,w)}{z-w} =0. \end{aligned}$$
(C.14)

\(\square \)

Lemma 42

Let \(a,b\sqrt{q}\) be all different and \(k,\ell \ge 1\). The kernel entry \(K^{\mathrm{geo}}_{12}\) can be expressed as

$$\begin{aligned} \begin{aligned} K^{\mathrm{geo}}_{12}(k,\ell )&=\frac{-1}{(2\pi \mathrm{i})^2}\oint \limits _{\varGamma _{\sqrt{q},a,b}} dw\oint \limits _{\varGamma _{1/\sqrt{q}}} dz \frac{w^{\ell -1}}{z^{k}}\frac{h^{\mathrm{geo}}_{12}(z,w)}{z-w} \\&\quad + \frac{-1}{(2\pi \mathrm{i})^2}\oint \limits _{\varGamma _{\sqrt{q},a}} dw \oint \limits _{\varGamma _{1/b}} dz \frac{w^{\ell -1}}{z^{k}}\frac{h^\mathrm{geo}_{12}(z,w)}{z-w} \end{aligned} \end{aligned}$$
(C.15)

where

$$\begin{aligned} h^{\mathrm{geo}}_{12}(z,w)=\frac{H(z)B(z)}{H(w)B(w)}\frac{(1-\sqrt{q} z)^n}{(1-\sqrt{q}/w)^n} \frac{(zw-1)(z-a)}{(z^2-1)(w-a)}. \end{aligned}$$
(C.16)

Proof

This time the change of variables is not necessary. Initially we have integration paths such that \(\max \{\sqrt{q},a,b\}<|w|<|z|\) and \(1<|z|<\min \{1/b,1/\sqrt{q}\}\). As before, 0 is not a pole for w and \(\infty \) is not a pole for z. Thus we can deform the w and z contours to include only \(\sqrt{q},a,b\) and \(1/b,1/\sqrt{q}\) respectively. Also in this case the contribution of the poles at \((w=b,z=1/b)\) is equal to zero. We then get the claimed result. \(\square \)

Lemma 43

Let \(a,b,\sqrt{q}\) be all different and \(k,\ell \ge 1\). The kernel entry \(K^{\mathrm{geo}}_{22}\) can be expressed as

$$\begin{aligned} \begin{aligned} K^{\mathrm{geo}}_{22}(k,\ell )&= E(k,\ell )+\frac{-1}{(2\pi \mathrm{i})^2}\oint \limits _{\varGamma _{\sqrt{q}}}\ \,dw \oint \limits _{\varGamma _{1/\sqrt{q},1/a,1/b}} dz \frac{w^{\ell -1}}{z^{k}}\frac{h^{\mathrm{geo}}_{22}(z,w)}{z-w} \\&\quad + \frac{-1}{(2\pi \mathrm{i})^2}\oint \limits _{\varGamma _{a}}\, dw\oint \limits _{\varGamma _{1/\sqrt{q},1/b}} dz \frac{w^{\ell -1}}{z^{k}}\frac{h^{\mathrm{geo}}_{22}(z,w)}{z-w}\\&\quad +\frac{-1}{(2\pi \mathrm{i})^2}\oint \limits _{\varGamma _{b}}\, dw \oint \limits _{\varGamma _{1/\sqrt{q},1/a}} dz \frac{w^{\ell -1}}{z^{k}}\frac{h^{\mathrm{geo}}_{22}(z,w)}{z-w} \end{aligned} \end{aligned}$$
(C.17)

where

$$\begin{aligned} h^\mathrm{geo}_{22}(z,w)=\frac{H(z)B(z)}{H(w)B(w)}\frac{1}{(1-\sqrt{q}/z)^n(1-\sqrt{q}/w)^n} \frac{(zw-1)}{(w-a)(1-az)} \end{aligned}$$
(C.18)

and

$$\begin{aligned} E(k,\ell )=-{{\,\mathrm{sgn}\,}}(k-\ell ) \oint \limits _{\varGamma _{1/a,1/\sqrt{q}}} \frac{dz}{2\pi \mathrm{i}} z^{\ell -k} \frac{z^{-1}-z}{(1-az)(z-a)}\frac{1}{(1-\sqrt{q}/z)^n (1-\sqrt{q} z)^n}. \end{aligned}$$
(C.19)

Remark 44

The condition that \(a,b,\sqrt{q}\) are all different is not really a restriction. If two or more of them are equal, one just has to use a different contour decomposition starting from the integration contours for (wz) as \(\varGamma _{\sqrt{q},a,b}\times \varGamma _{1/\sqrt{q},1/a,1/b}\). For example, if \(a=b\) and recalling that \(b\in (0,1)\), the given contours are already fine. The decomposition we mention is useful since we want to study the case \(b\rightarrow 1/a\) later.

Proof

We start with

$$\begin{aligned} K^{\mathrm{geo}}_{22}(k,\ell )&=\oint \frac{dw}{2\pi \mathrm{i}}\oint \frac{dz}{2\pi \mathrm{i}} \frac{z^{k-1} w^{\ell -1}}{H(z)B(z)H(w)B(w)}\frac{1}{(1-\sqrt{q}z)^n(1-\sqrt{q}w)^n}\nonumber \\&\quad \times \frac{z-w}{(zw-1)(z-a)(w-a)} \end{aligned}$$
(C.20)

where the integration contours can be taken as circles with \(|z|=|w|>\max \{\sqrt{q},a,b,1\}\) (the value \(1/\sqrt{q}\) is not a pole). The change of variable \(z\rightarrow z^{-1}\) leads to

$$\begin{aligned} K^{\mathrm{geo}}_{22}(k,\ell )=\frac{1}{(2\pi \mathrm{i})^2} \oint dw \oint dz \frac{w^{\ell -1}}{z^{k}} \frac{h^{\mathrm{geo}}_{22}(z,w)}{z-w} \end{aligned}$$
(C.21)

where the integration contours need to satisfy \(\max \{\sqrt{q},a,b\}<|z|<|w|<\min \{1/b,1/a,1/\sqrt{q}\}\). Notice however that this time we cannot directly shrink the w contour around \(\sqrt{q},a,b\) since z is also a pole for w and lies inside the w contour. We can exchange the contours, that is get w inside z, by correcting with the residue at \(w=z\). We get

$$\begin{aligned} K^{\mathrm{geo}}_{22}(k,\ell )=E(k,\ell )+\frac{1}{(2\pi \mathrm{i})^2} \oint dw \oint dz \frac{w^{\ell -1}}{z^{k}} \frac{h^{\mathrm{geo}}_{22}(z,w)}{z-w} \end{aligned}$$
(C.22)

where the integration contours now satisfy \(\max \{\sqrt{q},a,b\}<|w|<|z|<\min \{1/b,1/a,1/\sqrt{q}\}\), and where

$$\begin{aligned} E(k,\ell ) = \oint \limits _{\varGamma _{a, \sqrt{q}}} \frac{dz}{2\pi \mathrm{i}} \frac{z^{\ell -k-1}(1-z^2)}{(1-\sqrt{q}/z)^n(1-\sqrt{q} z)^n}\frac{1}{(1-az)(z-a)}. \end{aligned}$$
(C.23)

For the double integral, we verify that \(w=0\) and \(z=\infty \) are not poles and then deform them to include only the poles at \(\sqrt{q},a,b\) for w and \(1/\sqrt{q},1/a,1/b\) for z (the minus sign comes from the exchange of the orientation in the z path). Then we notice that the contributions of the poles at \((w=a,z=1/a)\) and at \((w=b,z=1/b)\) are 0. We are left with \(9-2=7\) contributions which we put together as three integrals in such a way that the paths can be taken as single contours and do not cross (recall the conditions \(ab<1\), \(a\sqrt{q}<1\) and that \(a,b,\sqrt{q}\) are all different).

Concerning \(E(k,\ell )\), it is an easy computation to verify that \(E(k,k)=0\) (the \(n=0\) and \(n \ge 1\) are two separate computations), which is actually a consequence of the anti-symmetry of the entry \(K_{22}(k,l)\). Furthermore, for \(k>\ell \) we have

$$\begin{aligned} E(k,\ell )=-\oint \limits _{\varGamma _{1/a,1/\sqrt{q}}} \frac{dz}{2\pi \mathrm{i}} z^{\ell -k} \frac{z^{-1}-z}{(1-az)(z-a)}\frac{1}{(1-\sqrt{q}/z)^n (1-\sqrt{q} z)^n} \end{aligned}$$
(C.24)

and for \(k<\ell \) we have

$$\begin{aligned} E(k,\ell )=\oint \limits _{\varGamma _{a,\sqrt{q}}} \frac{dz}{2\pi \mathrm{i}} z^{\ell -k} \frac{z^{-1}-z}{(1-az)(z-a)}\frac{1}{(1-\sqrt{q}/z)^n (1-\sqrt{q} z)^n} = -E(\ell ,k). \end{aligned}$$
(C.25)

\(\square \)

1.3 From the geometric to the exponential model: proof of Theorem 13

The exponential model is obtained from the geometric one by taking the \(\varepsilon \rightarrow 0\) limit with the following variables:

$$\begin{aligned} (a,b)=(1-\varepsilon \alpha ,1-\varepsilon \beta ),\quad q=1-\varepsilon ,\quad (k,\ell )=\varepsilon ^{-1}(x,y). \end{aligned}$$
(C.26)

Let us consider the accordingly rescaled kernel and conjugated kernel

$$\begin{aligned} K^\varepsilon _{11}(x,y)&=\varepsilon ^{-2-2n}K^{\mathrm{geo}}_{11}(k,\ell ), \quad K^\varepsilon _{12}(x,y) = \varepsilon ^{-1}K^{\mathrm{geo}}_{12}(k,\ell ), \nonumber \\ K^\varepsilon _{21}(x,y)&=\varepsilon ^{-1}K^{\mathrm{geo}}_{21}(k,\ell ), \quad K^\varepsilon _{22}(x,y) = \varepsilon ^{2n}K^{\mathrm{geo}}_{22}(k,\ell ). \end{aligned}$$
(C.27)

Then the following result is straightforward and does not require complicated asymptotic analysis.

Proposition 45

Uniformly for xy in a bounded set of \({\mathbb {R}}_+\), we have

$$\begin{aligned} \lim _{\varepsilon \rightarrow 0} K^\varepsilon _{ij}(x,y)= K_{ij}(x,y) \end{aligned}$$
(C.28)

where the kernel K is the one of Theorem 13 for the exponential model.

Proof

Let us also change the integration variables as \(z=1+Z \varepsilon \), \(w=1+W \varepsilon \). Notice that all the variables are in an \(\varepsilon \)-neighborhood of 1 and thus the contours for Z and W can be fixed independently of \(\varepsilon \); they just need to pass/be in the correct position with respect to the poles.

Therefore we can easily get the following point-wise limits:

$$\begin{aligned} \begin{aligned} \lim _{\varepsilon \rightarrow 0} H(z) z^{-k}&=\left[ \frac{\tfrac{1}{2}+z}{\tfrac{1}{2}-z} \right] ^{N-1} e^{-x Z}=\varPhi (x,Z),\\ \lim _{\varepsilon \rightarrow 0} (1-\sqrt{q} z)^n \varepsilon ^{-n}&= (\tfrac{1}{2}-Z)^n,\\ \lim _{\varepsilon \rightarrow 0} (1-\sqrt{q} /w)^n \varepsilon ^{-n}&= (\tfrac{1}{2}+W)^n \end{aligned} \end{aligned}$$
(C.29)

and

$$\begin{aligned} \begin{aligned}&\lim _{\varepsilon \rightarrow 0} \frac{1-b/z}{1-b z}\frac{1-b w}{1-b/w}\frac{(z-a)(1-w a)(zw-1)}{(z^2-1)(1-w^2)(z-w)}\\&\quad =\frac{(Z+\alpha )(Z+\beta )(W-\alpha )(W-\beta )(Z+W)}{4Z W (Z-W)(W+\beta )(Z-\beta )},\\&\lim _{\varepsilon \rightarrow 0} \frac{1-b/z}{1-b z}\frac{1-b w}{1-b/w}\frac{(z-a)(zw-1)}{(z^2-1)(w-a)(z-w)} \varepsilon \\&\quad =\frac{(Z+\alpha )(Z+\beta )(W-\beta )(Z+W)}{2Z (Z-W)(W+\alpha )(W+\beta )(Z-\beta )},\\&\lim _{\varepsilon \rightarrow 0} \frac{1-b/z}{1-b z}\frac{1-b w}{1-b/w}\frac{zw-1}{(1-a z)(w-a)(z-w)} \varepsilon ^{2}\\&\quad =\frac{-(Z+\beta )(W-\beta )(Z+W)}{(Z-W)(W+\alpha )(W+\beta )(Z-\alpha )(Z-\beta )},\\&\lim _{\varepsilon \rightarrow 0} z^{\ell -k} \frac{z^{-1}-z}{(1-a z)(z-a)} \varepsilon =e^{-Z(x-y)}\frac{2Z}{Z^2-\alpha ^2}. \end{aligned} \end{aligned}$$
(C.30)

By choosing the paths for ZW so that their distances to any of the poles are bounded away from 0 uniformly for all \(0<\varepsilon \ll 1\), we have that the integrands (appropriately multiplied by some powers of \(\varepsilon \)) are uniformly bounded. Thus we can use dominated convergence to take the \(\varepsilon \rightarrow 0\) limit inside the integration. \(\quad \square \)

It remains for us to find appropriate bounds in order to prove the convergence of the Fredholm Pfaffian to its corresponding limit. We do this next.

Lemma 46

Let \(\beta \in (0,1/2)\), \(\alpha \in (-1/2,1/2)\) with \(\alpha +\beta >0\). Let us set

$$\begin{aligned} \delta =(\beta -\max \{0,-\alpha \})/4,\text { and }\mu =\max \{0,-\alpha \}+2\delta . \end{aligned}$$
(C.31)

Then the following bounds hold uniformly in \(x,y\in {\mathbb {R}}_+\),

$$\begin{aligned} |K^\varepsilon _{11}(x,y) e^{\mu (x+y)}|&\le C e^{-\delta (x+y)}, \quad |K^\varepsilon _{12}(x,y) e^{\mu x} e^{-\mu y}| \le C e^{-\delta (x+y)}, \nonumber \\ |K^\varepsilon _{21}(x,y) e^{-\mu x} e^{\mu y}|&\le C e^{-\delta (x+y)}, \quad |K^\varepsilon _{22}(x,y)e^{-\mu (x+y)}| \le C e^{-\delta (x+y)}. \end{aligned}$$
(C.32)

Proof

The proof is quite simple, since the \(k,\ell \) dependence is only through the term \(w^\ell /z^k\) and the rest of the integrand remains nicely bounded and converges as we saw in Proposition 45.

For \(K^\varepsilon _{11}(k,l)\), choose contours satisfying \(|z|\ge 1+(\mu +\delta )\varepsilon \) and \(|w|\le 1-(\mu +\delta )\varepsilon \). This gives \(|w^\ell /z^k|\le \frac{(1-(\mu +\delta )\varepsilon )^{y/\varepsilon }}{(1+(\mu +\delta )\varepsilon )^{x/\varepsilon }}\simeq e^{-\mu (x+y)} e^{-\delta (x+y)}\).

For \(K^\varepsilon _{12}(k,l)\), choose the contours such that \(|z|\ge 1+(\mu +\delta )\varepsilon \) and \(|w|\le 1+(\mu -\delta )\varepsilon \). This gives \(|w^\ell /z^k|\lesssim e^{-\mu (x-y)} e^{-\delta (x+y)}\).

For the double integrand in \(K^\varepsilon _{22}(k,l)\) there are several terms. In the first and third one, we choose contours satisfying \(|z|\ge 1-(\mu -\delta )\varepsilon \) and \(|w|\le 1-(\mu +\delta )\varepsilon \), while in the second one \(|z|\ge 1+(\mu +\delta ) \varepsilon \) and \(|w|\le 1+(\mu -\delta )\varepsilon \). Combining the cases, we have \(|w^\ell /z^k|\lesssim e^{\mu (x+y)} e^{-\delta (x+y)}\). Finally, for the term coming from \(E(k,\ell )\), for \(k>\ell \) we take a contour with \(|z|\ge 1-(\mu -\delta )\varepsilon \), which gives \(|z^{\ell -k}|\lesssim e^{(\mu -\delta )(x-y)}\le e^{\mu x} e^{-\delta (x+y)}\) by using \(\mu -\delta \ge \delta \). For \(k<\ell \) the bound follows from the anti-symmetry of \(E(k,\ell )\). \(\square \)

Proof of Theorem 13

With Proposition 45 and Lemma 46 in hand, the proof of Theorem 13 is standard. One writes the Fredholm Pfaffian as a series expansion and then uses the Hadamard bound. Exactly the same steps are made in the proof of the convergence of the large time limit of Proposition 20 and thus we omit them here for brevity.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Betea, D., Ferrari, P.L. & Occelli, A. Stationary Half-Space Last Passage Percolation. Commun. Math. Phys. 377, 421–467 (2020). https://doi.org/10.1007/s00220-020-03712-5

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-020-03712-5

Navigation