Skip to main content
Log in

Stochastic PDE Limit of the Six Vertex Model

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

Abstract

We study the stochastic six vertex model and prove that under weak asymmetry scaling (i.e., when the parameter \(\Delta \rightarrow 1^+\) so as to zoom into the ferroelectric/disordered phase critical point) its height function fluctuations converge to the solution to the Kardar–Parisi–Zhang (KPZ) equation. We also prove that the one-dimensional family of stochastic Gibbs states for the symmetric six vertex model converge under the same scaling to the stationary solution to the stochastic Burgers equation. Our proofs rely upon the Markov (self) duality of our model. The starting point is an exact microscopic Hopf–Cole transform for the stochastic six vertex model which follows from the model’s known one-particle Markov self-duality. Given this transform, the crucial step is to establish self-averaging for specific quadratic function of the transformed height function. We use the model’s two-particle self-duality to produce explicit expressions (as Bethe ansatz contour integrals) for conditional expectations from which we extract time-decorrelation and hence self-averaging in time. The crux of our Markov duality method is that the entire convergence result reduces to precise estimates on the one-particle and two-particle transition probabilities. Previous to our work, Markov dualities had only been used to prove convergence of particle systems to linear Gaussian SPDEs (e.g. the stochastic heat equation with additive noise).

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
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14

We’re sorry, something doesn't seem to be working properly.

Please try refreshing the page. If that doesn't work, please contact support so we can address the problem.

Notes

  1. If there is no left-most particle, the dynamics can be still be defined with some care—see Sect. 2.1.

  2. When \(|\Delta |<1\) the conical points in the phase diagram disappear and the two disordered phases merge. When \(\Delta <-1\) a new antiferroelectric phase emerges for HV near zero. The associated Gibbs state is composed of diagonal bands of zig-zags made up only of the c-type vertices.

  3. In fact, (1.12) only gives upper boundary of the lens. The other boundary comes from applying the diagonal symmetry of the symmetric model.

  4. Recall, the symmetric 6V model only depends on (abc) through two parameters b/a and c/a. Also, note that we have used bold symbols here for \({\mathbf {u}}\) and \(\varvec{\eta }\) since later in the text, \({\mathbf {u}}\) and \(\varvec{\eta }\) will be used for occupation variables. Even though the Baxter parameterizations is limited to this discussion, we prefer not to risk confusion here.

  5. In [RS18], \(t=2h-1\) and \(s=2v-1\). There was a transcription error in [RS18, Eq. (34)] (which related a result from [BS95]). What was written there as \(\tanh ({\mathbf {u}}+\varvec{\eta })\) should be \(\tanh ({\mathbf {u}})\) (as stated here) [Pri18].

  6. This indicator function is essentially the occupation variable \(\eta \) which is used later in the text, see (2.2). Note that the meaning of the coordinates in \(\eta \) and u are opposite.

  7. We could have made the choice of parameter \(({\mathbf {u}}_\varepsilon ,\varvec{\eta }_\varepsilon )=(\frac{1}{2}\zeta {\sqrt{\varepsilon }},\frac{1}{2}{\sqrt{\varepsilon }})\) (without the lower order part in \({\mathbf {u}}_\varepsilon \)). This would lead to the parameter \(b_1\) which also depends on \(\varepsilon \), though the relation \(b_2/b_1 =e^{-{\sqrt{\varepsilon }}}\) would still hold. Our proof and result should still go through with extra notational complexity, though we do not pursue this direction here.

  8. In terms of the \(\varepsilon \) and \(\zeta \), the (abc) parameterization has a similar expansion of the form \(a=(1+\zeta )\tfrac{1}{2}\sqrt{\varepsilon }+{\mathcal {O}}(\varepsilon ^{-\frac{3}{2}})\), \(b=\zeta \tfrac{1}{2}\sqrt{\varepsilon }+{\mathcal {O}}(\varepsilon ^{-\frac{3}{2}})\), \(c=\tfrac{1}{2}\sqrt{\varepsilon }+{\mathcal {O}}(\varepsilon ^{-\frac{3}{2}})\).

  9. The positivity and well-posedness of (1.27) follows classical methods, see [Cor12, Qua11] for further details.

  10. Note that \(\eta (t,y)=\eta _{y}(\vec {x}(t))\). We distinguish the notation \(\eta (t,y)\) as a process and the notation \(\eta _{y}\) as a function on particle configurations \(\vec {x}\) merely for convenience.

  11. The computation for \(\lambda \) simply boils down to a geometric series. The computation for \(\mu \) boils down to a sum of the form \(\sum _{n\ge 0} (n+1) (b_2\tau ^{-\rho })^n\); this multiplied by \((1-b_2 \tau ^{-\rho })\) again gives a geometric series.

  12. For reference, see [Par18, Proposition 4.3] where existence, uniqueness and positivity in a more complicated case (i.e., with boundaries) are proved.

  13. This is different from exponentiating the interpolated height function. Nevertheless, under the weak asymmetry scaling \( \tau =\exp (-{\sqrt{\varepsilon }}) \), it is straightforward to verify that the difference between these two interpolation schemes is negligible as \( \varepsilon \rightarrow 0. \)

  14. In fact this is a weak solution. But solving (4.23) in the weak and mild senses are equivalent as shown in [BG97, Proposition 4.11].

  15. Here \( \Gamma (t,\varepsilon )\) does not depend on \( \varepsilon \), but we keep this notation to be consistent throughout all cases.

  16. This follows immediately from (1.29) by a simple tilting and centering.

References

  1. Aggarwal, A., Borodin, A.: Phase transitions in the ASEP and stochastic six-vertex model. Ann. Probab. 47(2), 613–689 (2019)

    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(4), 466–537 (2011)

    MathSciNet  MATH  Google Scholar 

  3. Aggarwal, A.: Current fluctuations of the stationary ASEP and six-vertex model. Duke Math. J. 167, 269–384 (2016)

    MathSciNet  MATH  Google Scholar 

  4. Aggarwal, A.: Convergence of the stochastic six-vertex model to the ASEP. Math. Phys. Anal. Geom. 20(2), 3 (2017)

    MATH  Google Scholar 

  5. Aggarwal, A.: Limit shapes and local statistics for the stochastic six-vertex model (2019). arXiv:1902.10867

  6. Assing, S.: A rigorous equation for the Cole-Hopf solution of the conservative KPZ equation. Stoch. PDEs Anal. Comput. 1(2), 365–388 (2013)

    MathSciNet  MATH  Google Scholar 

  7. Baxter, R.J.: Exactly Solved Models in Statistical Mechanics. Academic Press Inc, London (1989)

    MATH  Google Scholar 

  8. Barraquand, G., Borodin, A., Corwin, I., Wheeler, M.: Stochastic six-vertex model in a half-quadrant and half-line open asymmetric simple exclusion process. Duke Math. J. 167(13), 2457–2529 (2018)

    MathSciNet  MATH  Google Scholar 

  9. Borodin, A., Bufetov, A., Wheeler, M.: Between the stochastic six vertex model and Hall-Littlewood processes (2016). arXiv:1611.09486

  10. Bertini, L., Cancrini, N.: The stochastic heat equation: Feynman-Kac formula and intermittence. J. Statist. Phys. 78(5–6), 1377–1401 (1995)

    ADS  MathSciNet  MATH  Google Scholar 

  11. Borodin, A., Corwin, I., Ferrari, P., Vető, B.: Height fluctuations for the stationary KPZ equation. Math. Phys. Anal. Geom. 18(20), 1–95 (2015)

    MathSciNet  MATH  Google Scholar 

  12. Borodin, A., Corwin, I., Gorin, V.: Stochastic six-vertex model. Duke Math. J. 165(3), 563–624 (2016)

    MathSciNet  MATH  Google Scholar 

  13. Borodin, A., Corwin, I., Sasamoto, T.: From duality to determinants for \(q\)-TASEP and ASEP. Ann. Probab. 42(6), 2314–2382 (2014)

    MathSciNet  MATH  Google Scholar 

  14. Bertini, L., Giacomin, G.: Stochastic Burgers and KPZ equations from particle systems. Commun. Math. Phys. 183(3), 571–607 (1997)

    ADS  MathSciNet  MATH  Google Scholar 

  15. Borodin, A., Gorin, V.: Moments match between the KPZ equation and the Airy point process. SIGMA Sym. Integr. Geom. Meth. Appl. 12, 102 (2016)

    MathSciNet  MATH  Google Scholar 

  16. Borodin, A., Gorin, V.: A stochastic telegraph equation from the six-vertex model (2018). arXiv:1803.09137

  17. Bleher, P., Liechty, K.: Random Matrices and the Six-Vertex Model. CRM Monograph Series. American Mathematical Society, Providence (2014)

    MATH  Google Scholar 

  18. Borodin, A., Olshanski, G.: The ASEP and determinantal point processes. Commun. Math. Phys. 353, 853–903 (2017)

    ADS  MathSciNet  MATH  Google Scholar 

  19. Borodin, A., Petrov, L.: Integrable probability: stochastic vertex models and symmetric functions. In: Stochastic Processes and Random Matrices: Lecture Notes of the Les Houches Summer School, vol. 104, (2015)

  20. Bufetov, A., Petrov, L.: Law of Large Numbers for infinite random matrices over a finite field. Sel. Math. 21, 1271–1338 (2015)

    MathSciNet  MATH  Google Scholar 

  21. Borodin, A., Petrov, L.: Higher spin six vertex model and symmetric rational functions. Sel. Math. 24, 751–874 (2016)

    MathSciNet  MATH  Google Scholar 

  22. Bukman, D.J., Shore, J.D.: The conical point in the ferroelectric six-vertex model. J. Stat. Phys. 78(5), 1277–1309 (1995)

    ADS  MathSciNet  MATH  Google Scholar 

  23. Belitsky, V., Schütz, G.M.: Self-duality for the two-component asymmetric simple exclusion process. J. Math. Phys. 56(8), 083302 (2015)

    ADS  MathSciNet  MATH  Google Scholar 

  24. Chen, Z., de Gier, J., Wheeler, M.: Integrable stochastic dualities and the deformed Knizhnik-Zamolodchikov equation. Int. Math. Res. Not. 1, 59 (2018)

    Google Scholar 

  25. Carinci, G., Giardinà, C., Redig, F., Sasamoto, T.: A generalized asymmetric exclusion process with \(U_q({\mathfrak{sl}}_2)\) stochastic duality. Probab. Theo. Rel. Fields 166(3–4), 887–933 (2016)

    MATH  Google Scholar 

  26. Cannizzaro, G., Matetski, K.: Space-time discrete KPZ equation. Commun. Math. Phys. 358(2), 521–588 (2016)

    ADS  MathSciNet  MATH  Google Scholar 

  27. Corwin, I.: The Kardar-Parisi-Zhang equation and universality class. Random Matrices Theory Appl. 1(1), 1130001 (2012)

    MathSciNet  MATH  Google Scholar 

  28. Corwin, I.: Two ways to solve ASEP. In Springer Proceedings in Mathematics and Statistics of Topics in Percolative and Disordered Systems, vol. 69. Springer, New York (2014)

  29. Corwin, I., Petrov, L.: Stochastic higher spin vertex models on the line. Commun. Math. Phys. 343(2), 651–700 (2016)

    ADS  MathSciNet  MATH  Google Scholar 

  30. Corwin, I., Petrov, L.: Correction to: Stochastic higher spin vertex models on the line. Commun. Math. Phys. 371(1), 353–355 (2019)

    ADS  MathSciNet  MATH  Google Scholar 

  31. Corwin, I., Shen, H.: Open ASEP in the weakly asymmetric regime. Commun. Pure Appl. Math. 71(10), 2065–2128 (2018)

    MathSciNet  MATH  Google Scholar 

  32. Corwin, I., Shen, H., Tsai, L.-C.: \({{\rm ASEP}}(q, j)\) converges to the KPZ equation. Ann. Inst. Henri Poincaré Probab. Stat. 54(2), 995–1012 (2018)

    MathSciNet  MATH  Google Scholar 

  33. Corwin, I., Tsai, L.-C.: KPZ equation limit of higher-spin exclusion processes. Ann. Probab. 45, 1771–1798 (2017)

    MathSciNet  MATH  Google Scholar 

  34. Derrida, B., Appert, C.: Universal large-deviation function of the Kardar-Parisi-Zhang equation in one dimension. J. Statis. Phys. 94(1/2), 1–30 (1999)

    ADS  MathSciNet  MATH  Google Scholar 

  35. Derrida, B., Evans, M., Hakim, V., Pasquier, V.: Exact solution of a 1d asymmetric exclusion model using a matrix formulation. J. Phys. A 26, 1493–1517 (1993)

    ADS  MathSciNet  MATH  Google Scholar 

  36. Derrida, B.: An exactly soluble non-equilibrium system: the asymmetric simple exclusion process, vol. 301, pp. 65–83. Fundamental problems in statistical mechanics (Altenberg, 1997) (1998)

  37. Diehl, J., Gubinelli, M., Perkowski, N.: The Kardar-Parisi-Zhang equation as scaling limit of weakly asymmetric interacting brownian motions. Commun. Math. Phys. 354(2), 549–589 (2017)

    ADS  MathSciNet  MATH  Google Scholar 

  38. Derrida, B., Lebowitz, J.: Exact large deviation function in the asymmetric exclusion process. Phys. Rev. Let. 80(2), 209–213 (1998)

    ADS  MathSciNet  MATH  Google Scholar 

  39. De Masi, A., Presutti, E.: Mathematical Methods for Hydrodynamic Limits. Lecture Notes in Mathematics. Springer (1991)

  40. De Masi, A., Presutti, E., Scacciatelli, E.: The weakly asymmetric simple exclusion process. Ann. Inst. Henri Poincaré Probab. Stat. 25, 10 (1989)

    MathSciNet  MATH  Google Scholar 

  41. Dembo, A., Tsai, L.-C.: Weakly asymmetric non-simple exclusion process and the Kardar-Parisi-Zhang equation. Commun. Math. Phys. 341(1), 219–261 (2016)

    ADS  MathSciNet  MATH  Google Scholar 

  42. Erhard, D., Hairer, M.: Discretisation of regularity structures (2017). arXiv:1705.02836

  43. Faddeev, L.: How algebraic Bethe ansatz works for integrable model (1996). arXiv:hep-th/9605187

  44. Franco, T., Gonçalves, P., Simon, M.: Crossover to the stochastic Burgers equation for the WASEP with a slow bond. Commun. Math. Phys. 346(3), 801–838 (2016)

    ADS  MathSciNet  MATH  Google Scholar 

  45. Funaki, T., Quastel, J.: KPZ equation, its renormalization and invariant measures. Stoch. Partial Differ. Equ. Anal. Comput. 3(2), 159–220 (2015)

    MathSciNet  MATH  Google Scholar 

  46. Gartner, J.: Convergence towards Burgers’ equation and propagation of chaos for weakly asymmetric exclusion processes. Stoch. Process. Appl. 27(2), 233–260 (1988)

    MathSciNet  MATH  Google Scholar 

  47. Gaudin, M.: The Bethe Wavefunction. Cambridge University Press, Cambridge (2014)

    MATH  Google Scholar 

  48. Gerencsér, M., Hairer, M.: A solution theory for quasilinear singular spdes. Commun. Pure Appl. Math. 72(9), 1983–2005 (2019)

    MathSciNet  MATH  Google Scholar 

  49. Ghosal, P.: Hall-Littlewood-PushTASEP and its KPZ limit (2017). arXiv:1701.07308

  50. Gubinelli, M., Imkeller, P., Perkowski, N.: Paracontrolled distributions and singular PDEs. Forum Math. PI 3, 1 (2015)

    MathSciNet  MATH  Google Scholar 

  51. Goncalves, P., Jara, M.: Universality of KPZ equation (2010). arXiv:1003.4478

  52. Gubinelli, M., Jara, M.: Regularization by noise and stochastic Burgers equations. Stoch. Partial Differ. Equ. Anal. Comput. 1(2), 325–350 (2013)

    MathSciNet  MATH  Google Scholar 

  53. Gonçalves, P., Jara, M.: Nonlinear fluctuations of weakly asymmetric interacting particle systems. Arch. Ration. Mech. Anal. 212(2), 597–644 (2014)

    MathSciNet  MATH  Google Scholar 

  54. Gonçalves, P., Jara, M.: Stochastic Burgers equation from long range exclusion interactions. Stoch. Process. Appl. 127(12), 4029–4052 (2017)

    MathSciNet  MATH  Google Scholar 

  55. Gonçalves, P., Jara, M., Sethuraman, S.: A stochastic burgers equation from a class of microscopic interactions. Ann. Probab. 43(1), 286–338 (2015)

    MathSciNet  MATH  Google Scholar 

  56. Gubinelli, M., Perkowski, N.: Energy solutions of KPZ are unique. J. Am. Math. Soc. 31, 427–471 (2017)

    MathSciNet  MATH  Google Scholar 

  57. Gubinelli, M., Perkowski, N.: KPZ reloaded. Commun. Math. Phys. 349, 165–269 (2017)

    ADS  MathSciNet  MATH  Google Scholar 

  58. Goncalves, P., Perkowski, N., Simon, M.: Derivation of the stochastic Burgers equation with Dirichlet boundary conditions from the WASEP (2017). arXiv:1710.11011

  59. Gwa, L.-H., Spohn, H.: Six-vertex model, roughened surfaces, and an asymmetric spin Hamiltonian. Phys. Rev. Lett. 68, 725–728 (1992)

    ADS  MathSciNet  MATH  Google Scholar 

  60. Hairer, M.: Solving the KPZ equation. Ann. Math. 178(2), 559–664 (2013)

    MathSciNet  MATH  Google Scholar 

  61. Hairer, M.: A theory of regularity structures. Invent. Math. 198(2), 269–504 (2014)

    ADS  MathSciNet  MATH  Google Scholar 

  62. Hairer, M., Matetski, K.: Discretisations of rough stochastic PDEs. Ann. Probab. 46(3), 1651–1709 (2018)

    MathSciNet  MATH  Google Scholar 

  63. Hairer, M., Quastel, J.: A class of growth models rescaling to KPZ. Forum Math. PI 6(e3), 112 (2018)

    MathSciNet  MATH  Google Scholar 

  64. Hairer, M., Shen, H.: A central limit theorem for the KPZ equation. Ann. Probab. 45(6B), 4167–4221 (2017)

    MathSciNet  MATH  Google Scholar 

  65. Jimbo, M., Miwa, T.: Algebraic Analysis of Solvable Lattice Models. In: CBMS Regional Conference Series in Mathematics, vol. 85. American Mathematical Soc. (1993)

  66. Jayaprakash, C., Saam, W.F.: Thermal evolution of crystal shapes: the fcc crystal. Phys. Rev. B 30, 3916–3928 (1984)

    ADS  Google Scholar 

  67. Korepin, V.E., Bogoliubov, N.M., Izergin, A.G.: Quantum Inverse Scattering Method and Correlation Functions. Cambridge Monographs on Mathematical Physics. Cambridge University Press, Cambridge (1993)

    MATH  Google Scholar 

  68. Kenyon, R.: Conformal invariance of domino tiling. Ann. Probab. 28, 759–795 (2000)

    MathSciNet  MATH  Google Scholar 

  69. Kenyon, R.: Dominos and the Gaussian free field. Ann. Probab. 29, 1128–1137 (2001)

    MathSciNet  MATH  Google Scholar 

  70. Kenyon, R.: Lectures on dimers. Park City Math Institute Lectures (2009)

  71. Kim, D.: Bethe ansatz solution for crossover scaling functions of the asymmetric xxz chain and the kardar-parisi-zhang-type growth model. Phys. Rev. E 52, 3512–3524 (1995)

    ADS  Google Scholar 

  72. Kenyon, R., Miller, J., Sheffield, S., Wilson, D.B.: The six-vertex model and Schramm-Loewner evolution. Phys. Rev. E 95, 052146 (2017)

    ADS  Google Scholar 

  73. Kozlowski, K. K.: Asymptotic analysis and quantum integrable models (2015). arXiv:1508.06085

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

    ADS  MATH  Google Scholar 

  75. Kuan, J.: Stochastic duality of ASEP with two particle types via symmetry of quantum groups of rank two. J. Phys. A 49(11), 115002 (2016)

    ADS  MathSciNet  MATH  Google Scholar 

  76. Kuan, J.: An algebraic construction of duality functions for the stochastic \({\cal{U}}_q( A_n^{(1)})\) vertex model and its degenerations. Commun. Math. Phys. 359(1), 121–187 (2018)

    MathSciNet  MATH  Google Scholar 

  77. Kunita, H.: Stochastic Flows and Stochastic Differential Equations, vol. 24. Cambridge University Press, Cambridge (1997)

    MATH  Google Scholar 

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

    MathSciNet  MATH  Google Scholar 

  79. Labbé, C.: Weakly asymmetric bridges and the KPZ equation. Commun. Math. Phys. 353(3), 1261–1298 (2017)

    ADS  MathSciNet  MATH  Google Scholar 

  80. Lieb, E.H.: Residual entropy of square ice. Phys. Rev. 162, 162–172 (1967)

    ADS  Google Scholar 

  81. Liggett, T.M.: Interacting Particle Systems. Classics in Mathematics. Springer, Berlin (2005)

    Google Scholar 

  82. Lin, Y.: KPZ equation limit of stochastic higher spin six vertex model (2019). arXiv:1905.11155

  83. Lin, Y.: Markov duality for stochastic six vertex model (2019). arXiv:1901.00764

  84. Lieb, E.H., Wu, F.Y.: Two dimensional ferroelectric models. Phase Transit. Crit. Phenom. 1, 331–490 (1972)

    Google Scholar 

  85. Matetski, K.: Martingale-driven approximations of singular stochastic PDEs (2018). arXiv:1808.09429

  86. Nienhuis, B.: Critical behavior of two-dimensional spin models and charge asymmetry in the Coulomb gas. J. Stat. Phys. 34, 731–761 (1984)

    ADS  MathSciNet  MATH  Google Scholar 

  87. Nolden, I.M.: The asymmetric six-vertex model. J. Stat. Phys. 67(1–2), 155–201 (1992)

    ADS  MathSciNet  MATH  Google Scholar 

  88. Parekh, S.: The KPZ limit of ASEP with boundary. Commun. Math Phys. 365, 569–649 (2018)

    ADS  MathSciNet  MATH  Google Scholar 

  89. Pauling, L.: The structure and entropy of ice and of other crystals with some randomness of atomic arrangement. J. Am. Chem. Soc. 57(12), 2680–2684 (1935)

    Google Scholar 

  90. Perkowski, N., Cornelis Rosati, T.: The KPZ equation on the real line. Electron. J. Probab. 24, 56 (2019)

    MathSciNet  MATH  Google Scholar 

  91. Personal correspondences with Nicolai Reshetikhin and Amol Aggarwal. (2018)

  92. Quastel, J., Spohn, H.: The one-dimensional KPZ equation and its universality class. J. Stat. Phys. 160(4), 965–984 (2015)

    ADS  MathSciNet  MATH  Google Scholar 

  93. Quastel, J.: Introduction to KPZ. Curr. Dev. Math. 1, 10 (2011)

    Google Scholar 

  94. Ravishankar, K.: Fluctuations from the hydrodynamical limit for the symmetric simple exclusion in \({\mathbb{Z}}^d\). Stoch. Proc. Appl. 42, 31–37 (1992)

    MATH  Google Scholar 

  95. Reshetikhin, N.: Lectures on the integrability of the six-vertex model. In: Exact Methods in Low-dimensional Statistical Physics and Quantum Computing, pp. 197–266. Oxford Univ. Press, Oxford (2010)

  96. Reshetikhin, N., Sridhar, A.: Limit shapes of the stochastic six vertex model. Commun. Math. Phys. 363(3), 741–765 (2018)

    ADS  MathSciNet  MATH  Google Scholar 

  97. Rys, F.: Über ein zwidimensionales klassisches konfigurations-modell. Helv. Phys. Acat. 36, 537–559 (1963)

    MATH  Google Scholar 

  98. Schütz, G.M.: Duality relations for asymmetric exclusion processes. J. Stat. Phys. 86(5–6), 1265–1287 (1997)

    ADS  MathSciNet  MATH  Google Scholar 

  99. Seppäläinen, T.: Scaling for a one-dimensional directed polymer with boundary conditions. Ann. Probab. 40(1), 19–73 (2012)

    MathSciNet  MATH  Google Scholar 

  100. Slater, J.C.: Theory of the transition in \(KH_2PO_4\). J. Chem. Phys. 9(16), 16–33 (1941)

    ADS  Google Scholar 

  101. Spohn, H.: KPZ scaling theory and the semidiscrete directed polymer model. In: Random Matrix Theory, Interacting Particle Systems, and Integrable Systems, Math. Sci. Res. Inst. Publ., vol. 65, pp. 483–493. Cambridge Univ. Press, New York (2014)

  102. Sandow, S., Schütz, G.: On \(U_q[SU(2)]\)-symmetric driven diffusion. Euro. Phys. Lett. 26(1), 7 (1994)

    ADS  Google Scholar 

  103. Sasamoto, T., Spohn, H.: Point-interacting brownian motions in the KPZ universality class. Electron. J. Probab. 87, 28 (2015)

    MathSciNet  MATH  Google Scholar 

  104. Shen, H., Tsai, L.-C.: Stochastic telegraph equation limit for the stochastic six vertex model. Proc. Am. Math. Soc. 147(6), 2685–2705 (2019)

    MathSciNet  MATH  Google Scholar 

  105. Sutherland, B.: Exact solution of a two-dimensional model for Hydrogen-bonded crystals. Phys. Rev. Lett. 19, 103–104 (1967)

    ADS  Google Scholar 

  106. Tracy, C.A., Widom, H.: Integral formulas for the asymmetric simple exclusion process. Commun. Math. Phys. 279(3), 815–844 (2008)

    ADS  MathSciNet  MATH  Google Scholar 

  107. Tracy, C.A., Widom, H.: Erratum to: Integral formulas for the asymmetric simple exclusion process. Commun. Math. Phys. 304, 875–878 (2011)

    ADS  MATH  Google Scholar 

  108. Yang, K.: The KPZ equation, non-equilibrium solutions, and weak universality for long-range interactions (2018). arXiv:1810.02836

  109. Yang, C.-N., Yang, C.-P.: One-dimensional chain of anisotropic spin-spin interactions. II. Properties of the ground-state energy per lattice site for an infinite system. Phys. Rev. 150(1), 327 (1966)

    ADS  Google Scholar 

Download references

Acknowledgements

The authors wish to thank J. Dubédat, P. di Francesco, and T. Spencer for discussions about the disordered phase of the 6V model, A. Aggarwal for discussions about his work on this subject, Y. Lin for careful reading of part of the manuscript and K. Ravishankar for helpful historical references about the role of dualities in SPDE limits. We also appreciate helpful comments from an anonymous referee. I. Corwin was partially supported by the NSF through DMS:1811143 and DMS:1664650, and the Packard Foundation through a Packard Fellowship for Science and Engineering. H. Shen is partially supported by the NSF through DMS:1712684. L-C. Tsai is partially supported by the NSF through DMS:1712575 and the Simons Foundation through a Junior Fellowship. Both I. Corwin and P. Ghosal collaborated on this project during the 2017 Park City Mathematics Institute, partially funded by the NSF through DMS:1441467.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Ivan Corwin.

Additional information

Communicated by H. Spohn.

Publisher's Note

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

Appendix A. Quadratic variation in ASEP

Appendix A. Quadratic variation in ASEP

In this appendix we expand upon the brief discussion from Sects. 1.3 and 1.4 and explain how our Markov duality method can be applied to ASEP, which is a simpler limit of the stochastic 6V model. We will not carry out the necessary analysis, but rather just point to the main steps.

Recall that ASEP is an interacting particle system on \( {\mathbb {Z}}\), where particles inhabit sites index by \( {\mathbb {Z}}\) and jump left and right according to continuous-time exponential clocks with rates \( \ell > 0 \) and \( r> 0 \) subject to exclusion (jumps to occupied sites are suppressed). We will assume that \(\ell +r=1\) and set \(\tau := r/\ell \). The ASEP height function \( N_{ASEP }(t,x) \) has 1/0 slopes entering occupied/vacant sites as depicted in Fig. 6. For ASEP with near-stationary initial data of density \( \rho =\frac{1}{2} \) we define a variantFootnote 16 of the Hopf–Cole transform of \( N_{ASEP }(t,x) \) by

$$\begin{aligned} Z_{ASEP }(t,x) := \tau ^{N_{ASEP }(t,x)-\frac{1}{2} x} e^{t(1-2\sqrt{\ell r})}, \quad t\in [0,\infty ), \ x\in {\mathbb {Z}}. \end{aligned}$$

This solves the following microscopic SHE:

$$\begin{aligned} d Z_{ASEP }(t,x) = \sqrt{\ell r} \Delta Z_{ASEP }(t,x) dt + dM(t,x), \end{aligned}$$
(A.1)

where \( \Delta f(x):= f(x+1)+f(x-1)-2f(x) \) denotes the discrete Laplacian, and, for each \( x\in {\mathbb {Z}}\), the process \( M(t,x) \), \( t\in {\mathbb {R}}_+ \), is a martingale.

Under weak asymmetry scaling, i.e., \( \tau =\tau _\varepsilon :=e^{-{\sqrt{\varepsilon }}} \) and \( (t,x)\mapsto (\varepsilon ^{-2}t,\varepsilon ^{-1}x) \), an informal scaling argument applied to (A.1) indicates that the equation should converge to the continuum SHE. Key to establishing this convergence is the identification of the limiting quadratic variation of \( M(t,x) \). Under weak asymmetry scaling, the optional quadratic variation of \( M(t,x) \) reads

$$\begin{aligned} d\langle M(t,x), M(t,x') \rangle = \varepsilon {\mathbf {1}}_{{\{x=x'\}}} \Big ( \big ( \tfrac{1}{4} + \varepsilon ^\frac{1}{2} \mathcal {B}_\varepsilon (t,x) \big ) Z_{ASEP }^2(t,x) + {\widetilde{F}}_\varepsilon (t,x) \Big ) dt, \end{aligned}$$
(A.2)

where, following notations in Sect. 7, \( \mathcal {B}_\varepsilon (t,x) \) is a generic, uniformly bounded process, and

$$\begin{aligned} {\widetilde{F}}_\varepsilon (t,x) := \varepsilon ^{-\frac{1}{2}}\nabla Z_{ASEP }(t,x) \varepsilon ^{-\frac{1}{2}}\nabla Z_{ASEP }(t,x-1). \end{aligned}$$
(A.3)

Referring to the r.h.s. of (A.2), we see that \( \varepsilon ^\frac{1}{2} \mathcal {B}_\varepsilon (t,x) \) is indeed negligible compared to the constant \( \frac{1}{4} \) factor. Key to identifying the limiting behavior is to argue that \( {\widetilde{F}}(t,x) \) is also negligible. With \( \nabla Z_{ASEP }(t,x) = (e^{-{\sqrt{\varepsilon }} \eta (t,x+1)}-1)Z_{ASEP }(t,x) \), we indeed have \( {\widetilde{F}}_\varepsilon (t,x) = \mathcal {B}_\varepsilon (t,x) Z_{ASEP }^2(t,x) \), i.e., pointwise bounded up to a multiplicative factor of \( Z_{ASEP }^2(t,x) \). On the other hand, it is conceivable that this term \( {\widetilde{F}}(t,x) \) does not tend to zero pointwise, i.e., \( {\widetilde{F}}(t,x) \not \rightarrow _\text {P} 0 \). The crux of the convergence result is to prove that this term converges to zero after time-averaging:

$$\begin{aligned} {\mathbb {E}}\Big [ \Big ( \varepsilon ^{2}\int _0^{\varepsilon ^{-2}T} {\widetilde{F}}_\varepsilon (t,x) dt \Big )^2\Big ] \longrightarrow 0. \end{aligned}$$
(A.4)

This is first achieved in [BG97] by showing the decay as t becomes large of the conditional expectation

$$\begin{aligned} {\mathbb {E}}\big [ {\widetilde{F}}_\varepsilon (t+s,x) \big | {\mathscr {F}}(s) \big ], \end{aligned}$$

where \( {\mathscr {F}}\) denotes the canonical filtration of ASEP. Roughly speaking, the estimate starts by using (A.1) to develop a sequence of inequality that bounds the conditional expectation. ‘Closing’ the series of inequality relies crucially on an identity [BG97, (A.6)] for the (semi)-discrete heat kernel. We do not know of a way to generalize this approach from [BG97] to the stochastic 6V model setting.

Here we provide an alternative approach via duality. The Markov duality method also begins with bounding conditional expectations. However, instead of trying to close a sequence of inequalities, this method provides direct access to the conditional expectations. First, the expression \( {\widetilde{F}}_\varepsilon (t,x) \) is not convenient for our purpose. Use \( \nabla Z_{ASEP }(t,x) = (e^{-{\sqrt{\varepsilon }} (\eta ^+(t,x)-\frac{1}{2}) }-1)Z_{ASEP }(t,x) \) where \( \eta ^+(t,x) := \eta (t,x+1) \), and Taylor expand

$$\begin{aligned} \nabla Z_{ASEP }(t,x) ={\sqrt{\varepsilon }} (\tfrac{1}{2} - \eta ^{+}(t,x)) Z_{ASEP }+ \varepsilon \mathcal {B}_\varepsilon (t,x)Z_{ASEP }(t,x) , \end{aligned}$$
(A.5)

where \( \mathcal {B}_\varepsilon (t,x)\) stands for a generic uniformly bounded process as in Sect. 7. We can then write \( {\widetilde{F}}_\varepsilon (t,x) = F_\varepsilon (t,x) + \varepsilon ^{1/2} \mathcal {B}_\varepsilon (t,x)Z_{ASEP }^2(t,x) \), where

$$\begin{aligned} F_\varepsilon (t,x) = \tfrac{1}{2} Z_{\nabla }(t,x,x-1) + \tfrac{1}{2}Z_{\nabla }(t,x-1,x+1) + {\widetilde{Z}}(t,x-1,x), \end{aligned}$$
(A.6)

where, following the notation in Sect. 7,

$$\begin{aligned} Z_{\nabla }(t,x_1,x_2)&:= (\varepsilon ^{-\frac{1}{2}}\nabla Z_{ASEP }(t,x_1)) Z_{ASEP }(t,x_2),\\ {\widetilde{Z}}(t,x_1,t_2)&:= (\eta ^+Z_{ASEP })(t,x_1) (\eta ^+Z_{ASEP })(t,x_2) - \tfrac{1}{4} Z_{ASEP }(t,x_1) Z_{ASEP }(t,x_2). \end{aligned}$$

To see (A.6), we use (A.5) just as in the proof of Lemma 7.3 (for the stochastic 6V model):

$$\begin{aligned} {\widetilde{F}}_\varepsilon (t,x)&= \big ((\tfrac{1}{2} - \eta ^+) Z_{ASEP }\big )(t,x) \big ((\tfrac{1}{2} - \eta ^+) Z_{ASEP }\big )(t,x-1) + \varepsilon ^{1/2} \mathcal {B}_\varepsilon (t,x)Z_{ASEP }^2(t,x) \\&= \tfrac{1}{2} \big ((\tfrac{1}{2} - \eta ^+) Z_{ASEP }\big )(t,x) Z_{ASEP }(t,x-1)\\&\quad \,\, + \tfrac{1}{2} \big ((\tfrac{1}{2} - \eta ^+) Z_{ASEP }\big )(t,x-1) Z_{ASEP }(t,x)\\&\quad \,\, + {\widetilde{Z}}(t,x-1,x) + \varepsilon ^{1/2} \mathcal {B}_\varepsilon (t,x)Z_{ASEP }^2(t,x) \\&= \text{ r.h.s } \text{ of } {(A.6)} + \varepsilon ^{1/2} \mathcal {B}_\varepsilon (t,x)Z_{ASEP }^2(t,x) . \end{aligned}$$

In the last step, we replace \(Z_{ASEP }(t,x)\) with \(Z_{ASEP }(t,x\pm 1)\) costing error of order \(\varepsilon ^{1/2} \mathcal {B}_\varepsilon (t,x)Z_{ASEP }(t,x)\).

As mentioned in Sect. 1.4, ASEP enjoys self-duality via the functions Q and \( {\widetilde{Q}} \) defined therein. Specifically, the \( k=2 \) duality translates (after tilting and centering) into the following statement, in which we used the notation

$$\begin{aligned} {\mathbf {V}}_{ASEP }\big ((y_1,y_2),(x_1,x_2);t\big ) := e^{2t(1-2\sqrt{\ell r})} \tau ^{-\frac{1}{2}(x_1+x_2-y_1-y_2)} {\mathbb {P}}_\text {ASEP}\big ((y_1,y_2)\rightarrow (x_1,y_2) ;t\big ). \end{aligned}$$

Proposition A.1

For all \( x_1<x_2\in {\mathbb {Z}}\) and \( t,s\in [0,\infty ) \), we have

$$\begin{aligned}&{\mathbb {E}}\Big [ Z_{ASEP }(t+s,x_1) Z_{ASEP }(t+s,x_2) \Big \vert {\mathscr {F}}(s) \Big ]\nonumber \\&\quad = \sum _{y_1<y_2\in {\mathbb {Z}}} {\mathbf {V}}_{ASEP }\big ((y_1,y_2),(x_1,x_2);t\big ) Z_{ASEP }(s,y_1) Z_{ASEP }(s,y_2), \end{aligned}$$
(A.7)
$$\begin{aligned}&{\mathbb {E}}\Big [ (\eta ^{+}Z_{ASEP })(t+s,x_1) (\eta ^{+}Z_{ASEP })(t+s,x_2) \Big \vert {\mathscr {F}}(s) \Big ] \nonumber \\&\quad = \sum _{y_1<y_2\in {\mathbb {Z}}} {\mathbf {V}}_{ASEP }\big ((y_1,y_2),(x_1,x_2);t\big ) \big (\eta ^{+}Z_{ASEP }\big )(s,x_1)\big (\eta ^{+}Z_{ASEP }\big )(s,x_2).\quad \end{aligned}$$
(A.8)

Proposition A.1 provides the necessary ingredients for expressing conditional expectations for the relevant quantities. Specifically, with \( {\widetilde{Z}}(t,x-1,x) \) being an linear combination the two observables in (A.7) and in (A.8) at \( (x_1,x_2)=(x-1,x) \) we have

$$\begin{aligned} {\mathbb {E}}\Big [{\widetilde{Z}}(t+s,x-1,x)\Big |{\mathscr {F}}(s)\Big ] = \sum _{y_1<y_2\in {\mathbb {Z}}} {\mathbf {V}}_{ASEP }\big ((y_1,y_2),(x-1,x);t\big ) {\widetilde{Z}}(t,y_1,y_2). \end{aligned}$$
(A.9)

Likewise, \( Z_{\nabla }(t,x,x-1) \) is the difference of \( Z_{ASEP }(t,x+1)Z_{ASEP }(t,x-1) \) and \( Z_{ASEP }(t,x)Z_{ASEP }(t,x-1) \). Taking the difference of (A.7) for \( (x_1,x_2) = (x+1,x-1) \) and for \( (x,x-1) \) gives

$$\begin{aligned}&{\mathbb {E}}\Big [Z_{\nabla }(t+s,x,x-1) \Big |{\mathscr {F}}(s)\Big ] \\&\quad = \sum _{y_1<y_2\in {\mathbb {Z}}} \varepsilon ^{-\frac{1}{2}}\nabla _{x_1}{\mathbf {V}}_{ASEP }\big ((y_1,y_2),(x_1,x_2);t\big ) \big |_{(x_1,x_2)=(x,x-1)} Z_{ASEP }(s,y_1)Z_{ASEP }(s,y_2), \end{aligned}$$

where \( \nabla _{x_1} \) denotes the discrete (forward) gradient acting on the variable \( x_1 \). Similarly,

$$\begin{aligned}&{\mathbb {E}}\Big [Z_{\nabla }(t+s,x-1,x+1)\Big |{\mathscr {F}}(s)\Big ] \\&\quad = \sum _{y_1<y_2\in {\mathbb {Z}}} \varepsilon ^{-\frac{1}{2}}\nabla _{x_1}{\mathbf {V}}_{ASEP }\big ((y_1,y_2),(x_1,x_2); t\big )\big |_{(x_1,x_2)=(x-1,x+1)} Z_{ASEP }(s,y_1)Z_{ASEP }(s,y_2). \end{aligned}$$

From the perspective of duality, roughly speaking, the mechanism of decay in \( t\rightarrow \infty \) arises from the discrete gradient \( \nabla _{x_1} \). The semigroup \( {\mathbf {V}}_{ASEP }\) behaves similar to (two copies of) the heat kernel, so that \(\sum _{y_1<y_2}{\mathbf {V}}_{ASEP }\big ((y_1,y_2),(x_1,x_2);t\big ) = \mathcal O(1)\), and each gradient of \( {\mathbf {V}}_{ASEP }\) effectively produces a factor of \( t^{-1/2} \) for large t. Under the scaling \( \varepsilon ^{-2} \) of time, namely \( t^{-1/2}\approx \varepsilon ^1\), we expect to trade in \( \varepsilon ^{-1/2}\nabla \) for \( \varepsilon ^{-1/2}\varepsilon ^{1}=\varepsilon ^{1/2} \rightarrow 0 \). In other words, the key heuristic is that the l.h.s of (A.4) behaves as

$$\begin{aligned} {\mathbb {E}}\Big [ \Big ( \varepsilon ^{2}\int _0^{\varepsilon ^{-2}T} {\widetilde{F}}_\varepsilon (t,x) dt \Big )^2\Big ] \approx \varepsilon ^4 \int _0^{\varepsilon ^{-2}T} \!\!\!\! \int _0^{\varepsilon ^{-2}T} \!\!\! \frac{\varepsilon ^{-1/2}}{\sqrt{|t_1-t_2|}}\,dt_1dt_2 \approx \varepsilon ^{\frac{1}{2}} \rightarrow 0.\quad \end{aligned}$$
(A.10)

Note that the identity (A.9) in its current form does not involve gradients of \( {\mathbf {V}}_{ASEP }\). This identity can, however, be rewritten via Taylor expansion and summation by parts in a form that exposes the decay in \( t\rightarrow \infty \). We do not perform this procedure here, and direct the readers to Lemma 7.5, where the exact same procedure in carried out for the stochastic 6V model. Specifically, the identity (7.28) therein holds with \( ({\mathbf {V}}_{ASEP },Z_{ASEP },{\mathbb {Z}}) \) in place of \( ({\mathbf {V}}_\varepsilon ,Z,\Xi (s)) \), and with \( s,t\in [0,\infty ) \) instead of \( {\mathbb {Z}}_{\ge 0} \).

Given the preceding discussion, the task for bounding conditional expectations boils down to estimating the semigroup \( {\mathbf {V}}_{ASEP }\) and its gradients. Thanks to Bethe ansatz, \( {\mathbf {V}}_{ASEP }\) permits an explicit, analyzable formula in terms double contour integrals. Under weak asymmetry scaling, we write \( {\mathbf {V}}_{ASEP }={\mathbf {V}}_{\varepsilon ,ASEP }\) and the formula reads

$$\begin{aligned} {\mathbf {V}}_{\varepsilon ,ASEP }\big ((y_1,y_2),(x_1,x_2);t\big ) :=&\oint _{\mathcal {C}_r}\oint _{\mathcal {C}_r} \Big ( z_1^{x_1-y_1}z_2^{x_2-y_2} -{\mathfrak {F}}^ASEP _\varepsilon (z_1,z_2) z_1^{x_2-y_1}z_2^{x_1-y_2} \Big )\\&\prod _{i=1}^2 \frac{e^{t{\mathfrak {E}}^ASEP _\varepsilon (z_i)}dz_i}{2\pi {\mathbf {i}}z_i}, \end{aligned}$$

where \( \mathcal {C}_r \) is a counter-clockwise oriented, circular contour centered at origin, with a large enough radius r so as to include all poles of the integrand, and

$$\begin{aligned} {\mathfrak {F}}^ASEP _\varepsilon (z_1,z_2) := \frac{1+z_1z_2-(e^{-\frac{1}{2}{\sqrt{\varepsilon }}}+e^{\frac{1}{2}{\sqrt{\varepsilon }}}) z_2}{1+z_1z_2-(e^{-\frac{1}{2}{\sqrt{\varepsilon }}}+e^{\frac{1}{2}{\sqrt{\varepsilon }}}) z_1}, \quad {\mathfrak {E}}^ASEP _\varepsilon (z) := \sqrt{\ell r} \big (z+z^{-1}-2\big ). \end{aligned}$$

This contour integral formula is amenable to steepest decent analysis. Careful analysis jointly in \( (x_1,x_2,y_1,y_2,t) \) should produce the relevant estimates on \( {\mathbf {V}}_{\varepsilon ,ASEP }\) and its gradient (the result and proof should be analogous to Proposition 6.1). We do not pursue this analysis here.

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Corwin, I., Ghosal, P., Shen, H. et al. Stochastic PDE Limit of the Six Vertex Model. Commun. Math. Phys. 375, 1945–2038 (2020). https://doi.org/10.1007/s00220-019-03678-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-019-03678-z

Navigation