Skip to main content

Fluctuations in Stochastic Interacting Particle Systems

  • Conference paper
  • First Online:
Stochastic Dynamics Out of Equilibrium (IHPStochDyn 2017)

Part of the book series: Springer Proceedings in Mathematics & Statistics ((PROMS,volume 282))

Included in the following conference series:

  • 1087 Accesses

Abstract

We discuss fluctuations in stochastic lattice gas models from a microscopic and mesoscopic perspective by using techniques from algebra, in particular the use of symmetries and time-reversal. First we present a generic method to derive rigorously duality functions. As applications we obtain detailed information about density fluctuations in the symmetric simple exclusion process on any graph and about the microscopic structure and fluctuations of shocks in the one-dimensional asymmetric simple exclusion process. Then we use time reversal to prove a general current fluctuation theorem from which celebrated fluctuation relations such as the Jarzynski relation and the Gallavotti-Cohen symmetry arise as corollaries and which can be straightforwardly generalized to derive other fluctuation relations. Finally, going beyond rigorous results, we describe briefly how nonlinear fluctuating hydrodynamics yields the Fibonacci family of dynamical universality classes which has the diffusive and Kardar-Parisi-Zhang universality classes as its first two members.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 169.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 219.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 219.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    This equation has the form of a quantum mechanical Schrödinger equation in imaginary time, with H playing formally the role of the quantum Hamiltonian. This fact has given rise to the notion “quantum Hamiltonian formalism”.

  2. 2.

    On other words, detailed balance implies that the eigenvalues of the generator are all real and that the related symmetrized generator obtained from the ground state transformation can be interpreted as Hamiltonian of some quantum system. One sees that the use of the term quantum Hamiltonian formalism is justified by more than the formal analogy between Schrödinger equation and master equation.

  3. 3.

    The measure (2.75) as well as all related measures and functions introduced below depend both on \(L^-\) and \(L^+\). In order to avoid heavy notation we indicate this dependence only by the volume L.

References

  1. Alcaraz, F.C., Rittenberg, V.: Reaction-diffusion processes as physical realizations of Hecke algebras. Phys. Lett. B 314, 377–380 (1993)

    Article  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 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  3. Bahadoran, C.: Hydrodynamics and hydrostatics for a class of asymmetric particle systems with open boundaries. Commun. Math. Phys. 310(1), 1–24 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  4. Balázs, M., Farkas, G., Kovács, P., Rákos, A.: Random walk of second class particles in product shock measures. J. Stat. Phys. 139(2), 252–279 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  5. Baxter, R.J.: Exactly Solved Models in Statistical Mechanics. Academic Press, New York (1982)

    MATH  Google Scholar 

  6. Belitsky, V., Schütz, G.M.: Diffusion and coalescence of shocks in the partially asymmetric exclusion process. Electron. J. Prob. 7, 1–21 (2002). Paper No. 11

    Article  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  8. Belitsky, V., Schütz, G.M.: Self-duality and shock dynamics in the n-species priority ASEP. Stoch. Proc. Appl. 128, 1165–1207 (2018)

    Article  MathSciNet  MATH  Google Scholar 

  9. Ben Arous, G., Corwin, I.: Current fluctuations for TASEP: a proof of the PrähoferSpohn conjecture. Ann. Probab. 39, 104–138 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  10. Bernardin, C., Gonçalves, P., Jara, M.: 3/4-fractional superdiffusion in a system of harmonic oscillators perturbed by a conservative noise. Arch. Ration. Mech. Anal. 220, 505–542 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  11. Bertini, L., De Sole, A., Gabrielli, D., Jona Lasinio, G., Landim, C.: Macroscopic fluctuation theory. Rev. Mod. Phys. 87, 593–636 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  12. Blythe, R.A., Evans, M.R.: Nonequilibrium steady states of matrix product form: a solvers guide. J. Phys. A Math. Theor. 40, R333–R441 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  13. Bochkov, G.N., Kuzovlev, Y.E.: General theory of thermal fluctuations in nonlinear systems. Sov. Phys.—JETP 45, 125–130 (1977)

    Google Scholar 

  14. Bochkov, G.N., Kuzovlev, Y.E.: Fluctuation-dissipation relations for nonequilibrium processes in open systems. Sov. Phys.–JETP 49, 543–551 (1979)

    Google Scholar 

  15. Bochkov, G.N., Kuzovlev, Y.E.: Nonlinear fluctuation-dissipation relations and stochastic models in nonequilibrium thermodynamics I. Generalized fluctuation-dissipation theorem. Phys. A 106, 443–479 (1981)

    Article  MathSciNet  Google Scholar 

  16. Borodin, A., Petrov, L.: Lectures on integrable probability: stochastic vertex models and symmetric functions. In: Schehr, G., Altland, A., Fyodorov, Y.V., O’Connell, N., Cugliandolo, L.F. (eds.) Lecture Notes of the Les Houches Summer School, vol. 104, July 2015

    Google Scholar 

  17. Calabrese, P., Le Doussal, P.: Exact solution for the Kardar-Parisi-Zhang Equation with flat initial conditions. Phys Rev. Lett. 106, 250603 (2011)

    Article  Google Scholar 

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

    Google Scholar 

  19. Cancrini, N., Galves, A.: Approach to equilibrium in the symmetric simple exclusion process. Markov Processes Relat. Fields 1, 175–184 (1995)

    MathSciNet  MATH  Google Scholar 

  20. Cipriani, P., Denisov, S., Politi, A.: From anomalous energy diffusion to Lévy walks and heat conductivity in one-dimensional systems. Phys. Rev. Lett. 94, 244301 (2005)

    Article  Google Scholar 

  21. Chetrite, R., Touchette, H.: Nonequilibrium Markov processes conditioned on large deviations. Ann. Henri Poincaré 16(9), 2005–2057 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  22. Crooks, G.E.: Entropy production fluctuation theorem and the nonequilibrium work relation for free energy differences. Phys. Rev. E 60, 2721–2726 (1999)

    Article  Google Scholar 

  23. Derrida, B., Domany, E., Mukamel, D.: An exact solution of a one-dimensional asymmetric exclusion model with open boundaries. J. Stat. Phys. 69, 667 (1992)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MATH  Google Scholar 

  25. Derrida, B., Janowsky, S.A., Lebowitz, J.L., Speer, E.R.: Exact solution of the totally asymmetric simple exclusion process: shock profiles. J. Stat. Phys. 73, 813–842 (1993)

    Article  MathSciNet  MATH  Google Scholar 

  26. Derrida, B.: An exactly soluble nonequilibrium system: the asymmetric simple exclusion process. Phys. Rep. 301, 65–83 (1998)

    Article  MathSciNet  Google Scholar 

  27. Devillard, P., Spohn, H.: Universality class of interface growth with reflection symmetry. J. Stat. Phys. 66, 1089–1099 (1992)

    Article  MathSciNet  MATH  Google Scholar 

  28. Doi, M.: Second quantization representation for classical many-particle system. J. Phys. A: Math. Gen. 9, 1465–1477 (1976)

    Article  Google Scholar 

  29. Evans, D.J., Cohen, E.G.D., Morriss, G.P.: Probability of second law violations in shearing steady states. Phys. Rev. Lett. 71(15), 2401–2404 (1993)

    Article  MATH  Google Scholar 

  30. Evans, D.J., Searles, D.J.: Equilibrium microstates which generate second law violating steady states. Phys. Rev. E 50(2), 1645–1648 (1994)

    Article  Google Scholar 

  31. Evans, D.J., Searles, D.J.: The fluctuation theorem. Adv. Phys. 51, 1529–1585 (2002)

    Article  Google Scholar 

  32. Ferrari, P.A., Kipnis, C., Saada, E.: Microscopic structure of travelling waves in the asymmetric simple exclusion process. Ann. Probab. 19(1), 226–244 (1991)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  34. Gallavotti, G., Cohen, E.G.D.: Dynamical ensembles in stationary states. J. Stat. Phys. 80, 931–970 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  35. Giardinà, C., Kurchan, J., Redig, F., Vafayi, K.: Duality and hidden symmetries in interacting particle systems. J. Stat. Phys. 135, 25–55 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  36. Gradshteyn, I.S., Ryzhik, I.M.: Tables of Integrals, Series and Products. Academic Press, Orlando (1981)

    Book  Google Scholar 

  37. Grassberger, P., Scheunert, M.: Fock-space methods for identical classical objects. Fortschr. Phys. 28, 547–578 (1980)

    Article  MathSciNet  Google Scholar 

  38. Grisi, R., Schütz, G.M.: Current symmetries for particle systems with several conservation laws. J. Stat. Phys. 145, 1499–1512 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  39. Gwa, L.H., Spohn, H.: Bethe solution for the dynamical-scaling exponent of the noisy Burgers equation. Phys. Rev. A 46(2), 844–854 (1992)

    Article  Google Scholar 

  40. Halpin-Healy, T., Takeuchi, K.A.: A KPZ cocktail- shaken, not stirred: toasting 30 years of kinetically roughened surfaces. J. Stat. Phys. 160(4), 794–814 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  41. Harris, R.J., Rákos, A., Schütz, G.M.: Breakdown of Gallavotti-Cohen symmetry for stochastic dynamics. Europhys. Lett. 75, 227–233 (2006)

    Article  MathSciNet  Google Scholar 

  42. Harris, R.J., Schütz, G.M.: Fluctuation theorems for stochastic dynamics. J. Stat. Mech. P07020 (2007)

    Google Scholar 

  43. Harris, R.J., Popkov, V., Schütz, G.M.: Dynamics of instantaneous condensation in the ZRP conditioned on an atypical current. Entropy 15, 5065–5083 (2013)

    Article  MATH  Google Scholar 

  44. Harris, R.J., Schütz, G.M.: Fluctuation theorems for stochastic interacting particle systems. Markov Processes Relat. Fields 20, 3–44 (2014)

    MathSciNet  MATH  Google Scholar 

  45. Imamura, T., Sasamoto, T.: Current moments of 1D ASEP by duality. J. Stat. Phys. 142, 919–930 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  46. Imamura, T., Sasamoto, T.: Exact Solution for the Stationary Kardar-Parisi-Zhang equation. Phys. Rev. Lett. 108, 190603 (2012)

    Article  Google Scholar 

  47. Isaev, A.P., Pyatov, P.N., Rittenberg, V.: Diffusion algebras. J. Phys. A: Math. Gen. 34, 5815–5834 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  48. Jansen, S., Kurt, N.: On the notion(s) of duality for Markov processes. Probab. Surv. 11, 59–120 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  49. Jarzynski, C.: A non-quilibrium equality for free energy differences. Phys. Rev. Lett. 78, 2690–2693 (1997)

    Article  Google Scholar 

  50. Jarzynski, C.: Hamiltonian derivation of a detailed fluctuation theorem. J. Stat. Phys. 98, 77–102 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  51. Jarzynski, C.: Comparison of far-from-equilibrium work relations. C. R. Phys. 8, 495–506 (2007)

    Article  Google Scholar 

  52. Jimbo, M.: A \(q\)-difference analogue of U(\(\mathfrak{g}\)) and the Yang-Baxter equation. Lett. Math. Phys. 10, 63–69 (1985)

    Article  MathSciNet  MATH  Google Scholar 

  53. Jimbo, M.: A \(q\)-analogue of U(\(\mathfrak{g}\mathfrak{l}\)(N + 1)), Hecke algebra, and the Yang-Baxter equation. Lett. Math. Phys. 11, 247–252 (1986)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  55. Kadanoff, L.P., Swift, J.: Transport coefficients near the critical point: a master-equation approach. Phys. Rev. 165, 310–322 (1968)

    Article  Google Scholar 

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

    Article  MATH  Google Scholar 

  57. Karevski, D., Schütz, G.M.: Conformal invariance in driven diffusive systems at high currents. Phys. Rev. Lett. 118, 030601 (2017)

    Article  Google Scholar 

  58. Kim, D.: Bethe ansatz solution for crossover scaling functions of the asymmetric XXZ chain and the KPZ-type growth model. Phys. Rev. E 52, 3512–3524 (1995)

    Article  Google Scholar 

  59. Kipnis, C., Landim, C., Olla, S.: Hydrodynamic limit for a nongradient system: the generalized symmetric exclusion process. Commun. Pure Appl. Math. 47, 1475–1545 (1994)

    Article  MATH  Google Scholar 

  60. Kipnis, C., Landim, C.: Scaling Limits of Interacting Particle Systems. Springer, Berlin (1999)

    Book  MATH  Google Scholar 

  61. Kolomeisky, A.B., Schütz, G.M., Kolomeisky, E.B., Straley, J.P.: Phase diagram of one-dimensional driven lattice gases with open boundaries. J. Phys. A: Math. Gen. 31, 6911–6919 (1998)

    Article  MATH  Google Scholar 

  62. Krug, J.: Boundary-induced phase transitions in driven diffusive systems. Phys. Rev. Lett. 67, 1882–1885 (1991)

    Article  MathSciNet  Google Scholar 

  63. Lax, P.D.: Hyperbolic Systems of Conservation Laws and the Mathematical Theory of Shock Waves. CBMS 2011. SIAM, Philadelphia (1973)

    Book  MATH  Google Scholar 

  64. Le Doussal, P.: Crossover from droplet to flat initial conditions in the KPZ equation from the replica Bethe ansatz. J. Stat. Mech. 2014, P04018 (2014)

    Article  MathSciNet  Google Scholar 

  65. Lebowitz, J.L., Spohn, H.: A Gallavotti-Cohen-type symmetry in the large deviation functional for stochastic dynamics. J. Stat. Phys. 95, 333–365 (1999)

    Article  MathSciNet  MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  67. Liggett, T.M.: Interacting Particle Systems. Springer, Berlin (1985)

    Book  MATH  Google Scholar 

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

    Book  MATH  Google Scholar 

  69. Lloyd, P., Sudbury, A., Donnelly, P.: Quantum operators in classical probability theory: I. “Quantum spin” techniques and the exclusion model of diffusion. Stoch. Processes Appl. 61(2), 205–221 (1996)

    Article  MathSciNet  MATH  Google Scholar 

  70. MacDonald, J.T., Gibbs, J.H., Pipkin, A.C.: Kinetics of biopolymerization on nucleic acid templates. Biopolymers 6, 1–25 (1968)

    Article  Google Scholar 

  71. Maes, C.: On the origin and the use of fluctuation relations for the entropy. Sém. Poincaré 2, 29–62 (2003)

    Google Scholar 

  72. Minc, H.: Nonnegative Matrices. Wiley, New York (1988)

    MATH  Google Scholar 

  73. Nagel, K., Schreckenberg, M.: A cellular automaton model for freeway traffic. J. Phys. I France 2, 2221–2229 (1992)

    Article  Google Scholar 

  74. Pasquier, V., Saleur, H.: Common structures between finite systems and conformal field theories through quantum groups. Nucl. Phys. B 330, 523–556 (1990)

    Article  MathSciNet  Google Scholar 

  75. Popkov, V., Schütz, G.M.: Steady-state selection in driven diffusive systems with open boundaries. Europhys. Lett. 48, 257–264 (1999)

    Article  Google Scholar 

  76. Popkov, V., Schütz, G.M.: Shocks and excitation dynamics in a driven diffusive two-channel system. J. Stat. Phys. 112, 523–540 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  77. Popkov, V., Salerno, M.: Hydrodynamic limit of multichain driven diffusive models. Phys. Rev. E 69, 046103 (2004)

    Article  Google Scholar 

  78. Popkov, V., Schmidt, J., Schütz, G.M.: Universality classes in two-component driven diffusive systems. J. Stat. Phys. 160, 835–860 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  79. Popkov, V., Schadschneider, A., Schmidt, J., Schütz, G.M.: Fibonacci family of dynamical universality classes. Proc. Natl. Acad. Sci. U.S.A. 112(41), 12645–12650 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  80. Popkov, V., Schadschneider, A., Schmidt, J., Schütz, G.M.: Exact scaling solution of the mode coupling equations for non-linear fluctuating hydrodynamics in one dimension. J. Stat. Mech. 093211 (2016)

    Google Scholar 

  81. Prähofer, M., Spohn, H.: In and out of equilibrium. In: Sidoravicius, V. (ed.) Progress in Probability, vol. 51. Birkhauser, Boston (2002)

    MATH  Google Scholar 

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

    Article  MathSciNet  MATH  Google Scholar 

  83. Rákos, A., Schütz, G.M.: Exact shock measures and steady state selection in a driven diffusive system with two conserved densities. J. Stat. Phys. 117, 55–76 (2004)

    Article  MathSciNet  MATH  Google Scholar 

  84. Rákos, A., Harris, R.J.: On the range of validity of the fluctuation theorem for stochastic Markovian dynamics. J. Stat. Mech. P05005 (2008)

    Google Scholar 

  85. Rezakhanlou, F.: Hydrodynamic limit for attractive particle systems on \(\mathbb{Z}^d\). Commun. Math. Phys. 140, 417–448 (1991)

    Article  MathSciNet  MATH  Google Scholar 

  86. Sandow, S., Schütz, G.: On \(U_q[SU(2)]\)-symmetric driven diffusion. Europhys. Lett. 26, 7–13 (1994)

    Article  MathSciNet  Google Scholar 

  87. Sasamoto, T., Spohn, H.: One-dimensional Kardar-Parisi-Zhang equation: an exact solution and its universality. Phys. Rev. Lett. 834, 523 (2010)

    MATH  Google Scholar 

  88. Schadschneider, A., Chowdhury, D., Nishinari, K.: Stochastic Transport in Complex Systems. Elsevier, Amsterdam (2010)

    MATH  Google Scholar 

  89. Schütz, G., Domany, E.: Phase transitions in an exactly soluble one-dimensional asymmetric exclusion model. J. Stat. Phys. 72, 277–296 (1993)

    Article  MATH  Google Scholar 

  90. Schütz, G., Sandow, S.: Non-abelian symmetries of stochastic processes: derivation of correlation functions for random vertex models and disordered interacting many-particle systems. Phys. Rev. E 49, 2726–2744 (1994)

    Article  Google Scholar 

  91. Schütz, G.M.: The Heisenberg chain as a dynamical model for protein synthesis - some theoretical and experimental results. Int. J. Mod. Phys. B 11, 197–202 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  92. Schütz, G.M.: Duality relations for asymmetric exclusion processes. J. Stat. Phys. 86, 1265–1288 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  93. Schütz, G.M.: Solvable models for many-body systems far from equilibrium. In: Domb, C., Lebowitz, J. (eds.) Phase Transitions and Critical Phenomena, vol. 19, pp. 1–251. Academic, London (2001)

    Chapter  Google Scholar 

  94. Schütz, G.M.: Critical phenomena and universal dynamics in one-dimensional driven diffusive systems with two species of particles. J. Phys. A: Math. Gen. 36, R339–R379 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  95. Schütz, G.M., Wehefritz-Kaufmann, B.: Kardar-Parisi-Zhang modes in \(d\)-dimensional directed polymers. Phys. Rev. E 96, 032119 (2017)

    Article  Google Scholar 

  96. Schütz, G.M.: On the Fibonacci universality classes in nonlinear fluctuating hydrodynamics. In: Gonçalves, P., Soares, A. (eds.) From Particle Systems to Partial Differential Equations. PSPDE V, Braga, Portugal, November 2016. Springer Proceedings in Mathematics & Statistics, vol. 258, pp. 149–167. Springer, Cham (2018)

    Google Scholar 

  97. Searles, D.J., Evans, D.J.: Fluctuation theorem for stochastic systems. Phys. Rev. E 60(1), 159–164 (1999)

    Article  Google Scholar 

  98. Seifert, U.: Entropy production along a stochastic trajectory and an integral fluctuation theorem. Phys. Rev. Lett. 95, 040602 (2005)

    Article  Google Scholar 

  99. Seifert, U.: Stochastic thermodynamics, fluctuation theorems, and molecular machines. Rep. Prog. Phys. 75, 126001 (2012)

    Article  Google Scholar 

  100. Spohn, H.: Large Scale Dynamics of Interacting Particles. Springer, Berlin (1991)

    Book  MATH  Google Scholar 

  101. Spohn, H.: Bosonization, vicinal surfaces, and hydrodynamic fluctuation theory. Phys. Rev. E 60, 6411–6420 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  102. Spohn, H.: Nonlinear fluctuating hydrodynamics for anharmonic chains. J. Stat. Phys. 154, 1191–1227 (2014)

    Article  MathSciNet  MATH  Google Scholar 

  103. Spohn, H., Stoltz, G.: Nonlinear fluctuating hydrodynamics in one dimension: the case of two conserved fields. J. Stat. Phys. 160, 861–884 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  104. Spitzer, F.: Interaction of Markov processes. Adv. Math. 5, 246–290 (1970)

    Article  MathSciNet  MATH  Google Scholar 

  105. Sudbury, A., Lloyd, P.: Quantum operators in classical probability theory. II: the concept of duality in interacting particle systems. Ann. Probab. 23(4), 1816–1830 (1995)

    Article  MathSciNet  MATH  Google Scholar 

  106. Tóth, B., Valkó, B.: Onsager relations and Eulerian hydrodynamic limit for systems with several conservation laws. J. Stat. Phys. 112, 497–521 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  107. Tracy, C.A., Widom, H.: Asymptotics in ASEP with step initial condition. Commun. Math. Phys. 290, 129–154 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  108. van Beijeren, H.: Exact results for anomalous transport in one-dimensional Hamiltonian systems. Phys. Rev. Lett. 108, 108601 (2012)

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Gunter M. Schütz .

Editor information

Editors and Affiliations

A Some Linear and Multilinear Algebra

A Some Linear and Multilinear Algebra

In order to obtain more information about stochastic lattice gas models it will turn out to be convenient to write the generator of the process as a matrix which is called intensity matrix. Many probabilistic operations and notions then have a natural counterpart in linear algebra which we summarize here in elementary form. Indeed, much of what is presented is trivial, but perhaps necessary to write down in order to point out technically important subtleties and to introduce notation for the less common applications such the Kronecker product of matrices, sometimes also called outer product, which is essential for the choice of basis of the intensity matrix for lattice models.

In the following and throughout this work we use the Kronecker-symbol defined by

$$\begin{aligned} \delta _{\alpha ,\beta } = \left\{ \begin{array}{ll} 1 &{} \mathrm{if }~ \alpha =\beta \\ 0 &{} \mathrm{else } \end{array} \right. \end{aligned}$$
(A.1)

for \(\alpha ,\beta \) from any set. Complex conjugation is denoted by a bar as e.g. in \(\overline{z}\).

Matrices and Vectors. A \(m\times n\) matrix A is a number array

$$\begin{aligned} A = \left( \begin{array}{cccc} A_{11} &{} A_{12} &{} A_{13} &{} \dots \\ A_{21} &{} A_{22} &{} A_{23} &{} \dots \\ A_{31} &{} A_{32} &{} A_{33} &{} \dots \\ A_{41} &{} A_{42} &{} A_{43} &{} \dots \\ A_{51} &{} A_{52} &{} A_{53} &{} \dots \\ \vdots &{} \vdots &{} \vdots &{} \ddots \end{array} \right) . \end{aligned}$$

with \(m \ge 1\) rows and \(n\ge 1\) columns and matrix elements \(A_{kl}\) in row k and column l. The matrix elements \(A_{kl}\) will be mostly real numbers, but they can also be complex in certain applications. Hence we shall generally assume . We discuss some special cases.

(a) If \(m=n=1\) the matrix reduces a single number and we shall not differentiate between numbers and \(1\times 1\)-matrices.

(b) If \(n=1\) and \(m>1\) a matrix \(\varPhi \) is a column array of m numbers. We call such a matrix a ket-vector that we denote by the so-called ket-symbol \(| \, {\varPhi }\, \rangle \). The matrix elements \(\varPhi _{1l}\) with \(1\le l \le m\) will be denoted in simplified form by \(\varPhi _{l}\) and called components of the ket-vector. Thus

$$\begin{aligned} | \, {\varPhi }\, \rangle = \left( \begin{array}{c} \varPhi _1 \\ \varPhi _2 \\ \vdots \\ \varPhi _m \end{array} \right) . \end{aligned}$$

The vector with \(\varPhi _i = \delta _{ik}\) is a canonical basis vector of the vector space denoted by \(| \, {e_k}\, \rangle \). The set \(B_m := \{| \, {e_k}\, \rangle : k\in \{1,\dots ,m\}\}\) spans and is called the canonical basis.

(c) If \(m=1\) and \(n>1\) a matrix \(\varPsi \) is a row array of n numbers. We call such a matrix a bra-vector that we denote by the so-called bra-symbol \(\langle \, {\varPsi }\, |\). The matrix elements \(\varPsi _{k1}\) with \(1\le k \le n\) will be denoted in simplified form as \(\varPsi _{k}\) and called components of the bra-vector. Thus

$$\begin{aligned} \langle \, {\varPsi }\, | = \left( \varPsi _1 , \varPsi _2 , \dots , \varPsi _n \right) . \end{aligned}$$

Defining \(\langle \, {e_k}\, |=| \, {e_k}\, \rangle ^T\) one realizes that the set \(B^*:= \{\langle \, {e_k}\, |: k\in \{1,\dots ,n\}\}\) spans . Since any finite-dimensional vector space is isomorphic to its dual, we can think of the bra-vectors \(\mathbb {B}_n^*\) as representing the canonical basis of the dual space .

The letter or number inside the ket-symbol \(| \, {\cdot }\, \rangle \) or the bra-symbol \(\langle \, {\cdot }\, |\) is not to be understood as the argument of some function, but just as a symbol that collectively represents the components of the vector. When we use the term matrix we shall tacitly assume that \(m,n\ge 2\). The distinction between “proper” matrices on the one hand and the two types of vectors or simple numbers on the other hand is useful because many fundamental linear algebra operations can be represented as products involving numbers, bra- and ket-vectors and proper matrices with more than one column or row.

We usually denote proper matrices by capital letters or small letters with circumflex accent as e.g. in \(\hat{a}\). The unit matrix of dimension \(n>2\) with components \(A_{kl}=\delta _{k,l}\) is denoted by \(\mathbf {1}\) and for \(n=2\) we use the notation \(\mathbbm {1}\). Since multiplication of a vector with the unit matrix is the same as multiplication with the scalar unity 1 of the field F we do not usually differentiate between the two operations, i.e., in equations for matrices we often write a multiple \(x \mathbf {1}\) of the unity matrix simply as x.

Addition and Multiplication of Matrices. Any two matrices A and B which have the same number of rows and columns can be multiplied by a number and added to form a matrix \(C=xA+yB\) with the rule that \(C_{kl}=xA_{kl}+yB_{kl}\) where . Square matrices with \(m=n\) form a ring with a multiplication rule that can be generalized to non-square matrices as follows.

Definition 12

(Matrix product). For \(m,n,p \ge 1\) let A be a \(m\times p\)-matrix and B be a \(p\times n\)-matrix, both with matrix elements in some field F. The matrix product AB is an \(m\times n\) matrix C with matrix elements \(C_{kl} \equiv (AB)_{kl} \in F\) given by

$$\begin{aligned} C_{kl} = \sum _{j=1}^p A_{kj} B_{jl}, \quad 1 \le k \le m, \quad 1 \le l \le n. \end{aligned}$$
(A.2)

Square matrices AB of the same dimension \(m=n=p\) satisfying

$$\begin{aligned}{}[{A}\,,\,{B}] := AB - BA = 0 \end{aligned}$$
(A.3)

are said to commute.

Notice that unless \(m=n\) the reverse product BA is not defined since the number of columns in the first factor must be equal to the number of rows in the second factor of any matrix product. For a square matrix A the \(p^{th}\) power of A is denoted \(A^p\) and is defined for strictly positive integers by iteration of (A.2). By convention \(A^0 = \mathbf {1}\).

We discuss separately the special cases where at least one of the three number mnp is equal to one.

(a) If \(n=1\) and \(p,m>1\) the we can write the matrix B as a ket-vector \(| \, {\varPhi }\, \rangle \) with components \(\varPhi _k := B_{k1}\), \(k\in \{1,\dots ,p\}\). Then also the matrix product C is a ket-vector (with m components given by (A.2)) and the matrix product can be interpreted as a linear mapping \(| \, {\varPhi }\, \rangle \mapsto | \, {\tilde{\varPhi }}\, \rangle \) given by \(| \, {\tilde{\varPhi }}\, \rangle = A | \, {\varPhi }\, \rangle \), corresponding to the standard right multiplication of a matrix A with the column vector \(| \, {\varPhi }\, \rangle \).

(b) Likewise, for \(m=1\) and \(p,n>1\) we can write \(A = \langle \, {\varPsi }\, |\) as a bra-vector with p components with components \(\varPsi _l := A_{1l}\) and find that the matrix product is a linear mapping \(\langle \, {\varPsi }\, | \mapsto \langle \, {\tilde{\varPsi }}\, |\) that yields the bra-vector \(\langle \, {\tilde{\varPhi }}\, | = \langle \, {\varPsi }\, | B\) with n components given by (A.2), corresponding to the left multiplication of a matrix B with the row vector \(\langle \, {\varPsi }\, |\).

(c) If \(p=1\) and \(m,n>1\) then the matrix product actually turns into a product of two vectors. It maps a m-component ket-vector \(| \, {\varPhi }\, \rangle \) (=\(m\times 1\)-matrix A) with components \(\varPhi _k := A_{k1}\) and an n-component bra-vector \(\langle \, {\varPsi }\, |\) (=\(1\times n\)-matrix B) with components \(\varPsi _l := B_{1l}\) into a proper \(m\times n\) matrix

$$\begin{aligned} C=| \, {\varPhi }\, \rangle \langle \, {\varPsi }\, | \end{aligned}$$
(A.4)

with matrix elements \(C_{kl} = \varPhi _k \varPsi _l\) as given by (A.2). This mapping, called dyadic product, is a special form of the Kronecker product discussed below.

(d) For \(m=n=1\) the matrix product reduces to a single number \(C=\langle \, {\varPsi }\, || \, {\varPhi }\, \rangle =C_{11} \in F\) with

$$\begin{aligned} \langle \, {\varPsi }\, || \, {\varPhi }\, \rangle = \sum _{i=1}^p \varPsi _i \varPhi _i \equiv \langle {\varPsi }| \, {\varPhi } \rangle . \end{aligned}$$
(A.5)

It defines a bilinear mapping \((\langle \, {\varPsi }\, |,| \, {\varPhi }\, \rangle )\mapsto C_{11}\) which can be interpreted as a dual pairing \(d: \mathfrak {V}^*\times \mathfrak {V} \rightarrow F\), \((\langle \, {\varPsi }\, |,| \, {\varPhi }\, \rangle ) \mapsto \langle {\varPsi }| \, {\varPhi } \rangle \) since it is natural to regard the bra-vector to be an element of the vector space dual to the vector space to which the ket-vector belongs. This motivates the simplified notation \(\langle {\varPsi }| \, {\varPhi } \rangle \) of this matrix product with only one vertical bar.

Specifically, for the basis vectors we obtain from (A.5) the biorthogonality relation

$$\begin{aligned} \langle {e_i}| \, {e_j} \rangle = \delta _{ij}. \end{aligned}$$
(A.6)

Notice the difference between the dual pairing (A.5) and the scalar product \(s: \mathfrak {V} \times \mathfrak {V} \rightarrow F\) defined by the sesquilinear form \((| \, {\varPhi '}\, \rangle ,| \, {\varPhi }\, \rangle )\mapsto \langle {\varPhi ',\varPhi } \rangle :=\sum _{i=1}^p \overline{\varPhi }_i' \varPhi _i\) which is linear in the second argument, but antilinear in the first. When \(\langle \, {\varPhi '}\, |\) has only real components (as is the case in most of our applications) this distinction is irrelevant, but should nevertheless be kept in mind.

The Kronecker Product. The Kronecker product \(A\otimes B\) is defined for arbitrary rectangular matrices (including vectors and numbers) as follows.

Definition 13

(Kronecker product). Let A and B be two finite-dimensional matrices with \(m_A\ge 1\) (\(m_B\ge 1\)) rows and \(n_A\ge 1\) (\(n_B\ge 1\)) columns with matrix elements \(A_{ij}\) and \(B_{kl}\) respectively. The Kronecker product \(A\otimes B\) is a \(m_Am_B \times n_An_B\)-matrix C with matrix elements

$$\begin{aligned} C_{(i-1)m_B+k,(j-1)n_B+l}=A_{ij} B_{kl} \end{aligned}$$
(A.7)

with \(i \in \{1,\dots ,m_A\}, j \in \{1,\dots ,n_A\}, k \in \{1,\dots ,m_B\}, l \in \{1,\dots ,n_B\}\).

Alternatively we can write

$$\begin{aligned} A\otimes B = \left( \begin{array}{cccc} A_{11} B &{} A_{12} B &{} A_{13} B &{} \dots \\ A_{21} B &{} A_{22} B &{} A_{23} B &{} \dots \\ A_{31} B &{} A_{32} B &{} A_{33} B &{} \dots \\ A_{41} B &{} A_{42} B &{} A_{43} B &{} \dots \\ A_{51} B &{} A_{52} B &{} A_{53} B &{} \dots \\ \vdots &{} \vdots &{} \vdots &{} \ddots \end{array} \right) . \end{aligned}$$

Here each matrix “element” is itself a matrix, viz. the matrix B multiplied by the number \(A_{ij}\). In general \(A\otimes B\ne B\otimes A\). For the p-fold Kronecker product of a matrix A with itself is denoted by \(A^{\otimes p}\) with the convention that \(A^{\otimes 1} := A\) and \(A^{\otimes 0} := 1\) where 1 is the unit element of F and not the unit matrix. We discuss special cases.

(a) Consider \(n_A=n_B=1\), i.e., the Kronecker product of ket-vectors \(| \, {\varPhi ^1}\, \rangle , | \, {\varPhi ^2}\, \rangle \) with components \(\varPhi ^1_i\) where \(i \in \{1,\dots ,m_A\}\) and \(\varPhi ^2_k\) where \(k \in \{1,\dots ,m_B\}\). The tensor product \(| \, {\varPhi ^1}\, \rangle \otimes | \, {\varPhi ^2}\, \rangle \) is a column vector of dimension \(m_A m_B\) denoted by \(| \, {\varPhi ^1,\varPhi ^2}\, \rangle \) and has factorized components \((| \, {\varPhi ^1,\varPhi ^2}\, \rangle )_{(i-1)m_B+k} = \varPhi ^1_i \varPhi ^2_k\). Specifically, for the canonical basis vectors one gets \(| \, {e_i}\, \rangle \otimes | \, {e_k}\, \rangle \equiv | \, {e_i,e_k}\, \rangle = | \, {e_{(i-1)m_B+k}}\, \rangle \). Thus the Kronecker product of two canonical basis vectors yields a canonical basis vector. The set \(\mathbb {B}_{m_Am_B} := \{| \, {e_{(i-1)m_B+k}}\, \rangle : (i,k) \in \{1,\dots ,m_A\}\times \{1,\dots ,m_B\}\}\) forms the canonical basis of the tensor space .

(b) Similarly, for \(m_A=m_B=1\), i.e., for bra-vectors \(\langle \, {\varPsi ^1}\, |, \langle \, {\varPsi ^2}\, |\) with components \(\varPsi ^1_j\) where \(j \in \{1,\dots ,n_A\}\) and \(\varPsi ^2_l\) where \(l \in \{1,\dots ,n_B\}\) the tensor product \(\langle \, {\varPsi ^1}\, | \otimes \langle \, {\varPsi ^2}\, |\) is a row vector of dimension \(n_A n_B\) denoted by \(\langle \, {\varPsi ^1,\varPsi ^2}\, |\). It has factorized components \((\langle \, {\varPsi ^1,\varPsi ^2}\, |)_{(j-1)n_B+l} = \varPsi ^1_j \varPsi ^2_l\) and for the canonical basis vectors one gets \(\langle \, {e_j,e_l}\, | = \langle \, {e_{(j-1)n_B+l}}\, |\).

(c) For the Kronecker product of a bra-vector \(\langle \, {\varPsi }\, |\) and a ket-vector \(| \, {\varPhi }\, \rangle \) the Definition 13 yields

$$\begin{aligned} \langle \, {\varPsi }\, |\otimes | \, {\varPhi }\, \rangle = | \, {\varPhi }\, \rangle \otimes \langle \, {\varPsi }\, | = | \, {\varPhi }\, \rangle \langle \, {\varPsi }\, | \end{aligned}$$
(A.8)

with the dyadic product (A.4).

The Kronecker product is associative. Multiple Kronecker products of matrices define multilinear maps of the multiple tensor product of vector spaces defined by iterating the Kronecker product Definition 13. They satisfy the multiplication rule

$$\begin{aligned} (A\otimes B) (C \otimes D) = (AC) \otimes (BD) \end{aligned}$$
(A.9)

where we assume that the matrix products AC and BD are defined by (A.2).

We note an important factorization property of the dual pairing of Kronecker products of vectors which is an immediate consequence of the multilinearity of the Kronecker product encoded in (A.7).

Proposition 5

Let \(\langle \, {\varPsi ^k}\, |\) (\(| \, {\varPhi ^k}\, \rangle \)) be a bra-vector (ket-vector) of dimension \(d_k\) with components () and \(\langle \, {\varPsi ^1,\varPsi ^2,\dots ,\varPsi ^L}\, | = \langle \, {\varPsi ^1}\, | \otimes \langle \, {\varPsi ^2}\, | \otimes \dots \otimes \langle \, {\varPsi ^L}\, |\) (\(| \, {\varPhi ^1,\varPhi ^2,\dots ,\varPhi ^L}\, \rangle = | \, {\varPhi ^1}\, \rangle \otimes | \, {\varPhi ^2}\, \rangle \otimes \dots \otimes | \, {\varPhi ^L}\, \rangle \)) be the L-fold Kronecker product of these vectors. Then the dual pairing factorizes as

$$\begin{aligned} \langle {\varPsi ^1,\varPsi ^2,\dots ,\varPsi ^L}| \, {\varPhi ^1,\varPhi ^2,\dots ,\varPhi ^L} \rangle = \prod _{k=1}^L \langle {\varPsi ^k}| \, {\varPhi ^k} \rangle \end{aligned}$$
(A.10)

with \(\langle {\varPsi ^k}| \, {\varPhi ^k} \rangle \) given by (A.5).

When \(\langle \, {\varPsi ^k}\, | = \langle \, {\varPsi }\, |\) for all \(k \in \{1,\dots ,L\}\) then we write \(\langle \, {\varPsi ^1,\varPsi ^2,\dots ,\varPsi ^L}\, | = \langle \, {\varPsi }\, |^{\otimes L}\) and analogously for ket-vectors and proper matrices A.

Finally we introduce local operators which act non-trivially only on component k in an L-fold tensor space. For simplicity we assume equal dimensions \(d:=d_1=d_2=\dots =d_L\).

Definition 14

(Local operator). Let \(\mathbf {1}\) be the d-dimensional unit matrix and A be an arbitrary square matrix of dimension \(d \ge 1\). The local operator \(A_k\) is the Kronecker product

$$\begin{aligned} A_k := \mathbf {1}^{\otimes (k-1)} \otimes A \otimes \mathbf {1}^{\otimes (L-k)}. \end{aligned}$$
(A.11)

Notice the difference between the number and the unit matrix \(\mathbf {1}\) in this definition. The expression “local operator” come from the fact that when acting on a tensor vector \(| \, {\varPhi ^1,\dots ,\varPhi ^L}\, \rangle \) only the \(k^{th}\) factor is changed by the action of \(A_k\). More precisely,

$$\begin{aligned} A_k \left( | \, {\varPhi ^1}\, \rangle \otimes \dots \otimes | \, {\varPhi ^k}\, \rangle \otimes \dots \otimes | \, {\varPhi ^L}\, \rangle \right) = | \, {\varPhi ^1}\, \rangle \otimes \dots \otimes | \, {\tilde{\varPhi }^k}\, \rangle \otimes \dots \otimes | \, {\varPhi ^L}\, \rangle . \end{aligned}$$
(A.12)

where \(| \, {\tilde{\varPhi }^k}\, \rangle = A| \, {\varPhi ^k}\, \rangle \).

From (A.9) one finds

$$\begin{aligned} A_k B_k = (AB)_k \end{aligned}$$
(A.13)

which is equal to \(B_kA_k\) if and only if \(AB=BA\). On the other hand, by construction one has for two square matrices AB of dimension k the commutation relation

$$\begin{aligned} A_k B_l = B_l A_k ~\mathrm{for}~ k\ne l \end{aligned}$$
(A.14)

even when \(AB\ne BA\). In order to avoid confusion concerning the role of the indices we point out that for \(L=2\) and \([{A}\,,\,{B}]\ne 0\) we have

$$\begin{aligned} A\otimes B = A_1 B_2 = B_2 A_1 \ne B \otimes A = B_1 A_2 = A_2 B_1. \end{aligned}$$
(A.15)

We also note that for matrices \(A^{(k)}\) one has

$$\begin{aligned} A^{(1)}_1 A^{(2)}_2 \dots A^{(L)}_L = A^{(1)} \otimes A^{(2)} \otimes \dots \otimes A^{(L)}. \end{aligned}$$
(A.16)

The upper index defines the matrix while the lower index defines its position in the L-fold Kronecker product. We stress that \(A^{(k)}\) is a matrix of dimension d while \(A^{(k)}_k\) is a matrix of dimension \(d^L\).

From Proposition 5 one finds for d-dimensional square matrices \(A^{(k)}\) the factorization property

$$\begin{aligned} \langle \, {\varPsi ^1,\varPsi ^2,\dots ,\varPsi ^L}\, | A^{(1)}_1 A^{(2)}_2 \dots A^{(L)}_L | \, {\varPhi ^1,\varPhi ^2,\dots ,\varPhi ^L}\, \rangle = \prod _{k=1}^L \langle \, {\varPsi ^k}\, |A^{(k)}| \, {\varPhi ^k}\, \rangle . \end{aligned}$$
(A.17)

We write explicitly two special cases of particular importance:

$$\begin{aligned} \frac{\langle \, {\varPsi ^1,\varPsi ^2,\dots ,\varPsi ^L}\, | A_k | \, {\varPhi ^1,\varPhi ^2,\dots ,\varPhi ^L}\, \rangle }{\langle {\varPsi ^1,\varPsi ^2,\dots ,\varPsi ^L}| \, {\varPhi ^1,\varPhi ^2,\dots ,\varPhi ^L} \rangle }= & {} \frac{\langle \, {\varPsi ^k}\, |A| \, {\varPhi ^k}\, \rangle }{\langle {\varPsi ^k}| \, {\varPhi ^k} \rangle } \end{aligned}$$
(A.18)
$$\begin{aligned} \left( \langle \, {\varPsi }\, |\right) ^{\otimes L} \left( | \, {\varPhi }\, \rangle \right) ^{\otimes L}= & {} \langle {\varPsi }| \, {\varPhi } \rangle ^L. \end{aligned}$$
(A.19)

These computational properties of the matrix product (A.2) and of the Kronecker product defined in Definition 13 will be exploited throughout these notes.

Rights and permissions

Reprints and permissions

Copyright information

© 2019 Springer Nature Switzerland AG

About this paper

Check for updates. Verify currency and authenticity via CrossMark

Cite this paper

Schütz, G.M. (2019). Fluctuations in Stochastic Interacting Particle Systems. In: Giacomin, G., Olla, S., Saada, E., Spohn, H., Stoltz, G. (eds) Stochastic Dynamics Out of Equilibrium. IHPStochDyn 2017. Springer Proceedings in Mathematics & Statistics, vol 282. Springer, Cham. https://doi.org/10.1007/978-3-030-15096-9_3

Download citation

Publish with us

Policies and ethics