Skip to main content
Log in

Heat Flow in a Periodically Forced, Thermostatted Chain

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

Abstract

We investigate the properties of a harmonic chain in contact with a thermal bath at one end and subjected, at its other end, to a periodic force. The particles also undergo a random velocity reversal action, which results in a finite heat conductivity of the system. We prove the approach of the system to a time periodic state and compute the heat current, equal to the time averaged work done on the system, in that state. This work approaches a finite positive value as the length of the chain increases. Rescaling space, the strength and/or the period of the force leads to a macroscopic temperature profile corresponding to the stationary solution of a continuum heat equation with Dirichlet-Neumann boundary conditions.

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.

Similar content being viewed by others

References

  1. Basile, G., Olla, S., Spohn, H.: Energy transport in stochastically perturbed lattice dynamics. Arch. Ration. Mech. Anal. 195(1), 171–203 (2010)

    Article  MathSciNet  MATH  Google Scholar 

  2. Bernardin, C., Kannan, V., Lebowitz, J.L., Lukkarinen, J.: Nonequilibrium stationary states of harmonic chains with bulk noises. Eur. Phys. J. B 84, 685–689 (2011)

    Article  ADS  Google Scholar 

  3. Bernardin, C., Olla, S.: Transport properties of a chain of anharmonic oscillators with random flip of velocities. J. Stat Phys 145, 1224–1255 (2011). https://doi.org/10.1007/s10955-011-0385-6

    Article  ADS  MathSciNet  MATH  Google Scholar 

  4. Bolsterli, M., Rich, M., Visscher, W.M.: Simulation of nonharmonic interactions in a crystal by self-consistent reservoirs. Phys. Rev. A 4, 1086–1088 (1970)

    Article  ADS  Google Scholar 

  5. Bonetto, F., Lebowitz, J.L., Lukkarinen, J.: Fourier’s Law for a Harmonic Crystal with Self-Consistent Stochastic Reservoirs. J. Stat. Phys. Vol. 116, (2004)

  6. Gihman, I.I., Skorohod, A.V.: The theory of stochastic processes. III. Translated from the Russian by Samuel Kotz. With an appendix containing corrections to Volumes I and II. Grundlehren der Mathematischen Wissenschaften, 232. Springer-Verlag, Berlin-New York, (1979)

  7. Jara, M., Komorowski, T., Olla, S.: Superdiffusion of energy in a chain of harmonic oscillators with noise. Comm. Math. Phys. 339(2), 407–453 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  8. Khasminskii, R.: Stochastic Stability of Differential Equations, Stochastic Modelling and Applied Probability 66, Springer

  9. Komorowski, T., Lebowitz, J. L., Olla, S.: Heat flow in a periodically forced, thermostatted chain—with internet supplement, Available at arXiv:2205.03839

  10. Komorowski, T., Lebowitz, J. L., Olla, S.: Heat flow in a periodically forced, thermostatted chain II, (2022) arXiv:2209.12923, to appear in J. Stat. Phys

  11. Komorowski, T., Lebowitz, J.L., Olla, S., Simon, M.: On the Conversion of Work into Heat: Microscopic Models and Macroscopic Equations, (2022), arXiv:2212.00093, to appear in Ensaios Matemáticos for the volume dedicated to Errico Presutti 80th birthday

  12. Komorowski, T., Olla, S., Simon, M.: An open microscopic model of heat conduction: evolution and non-equilibrium stationary states. Commun. Math. Sci. 18(3), 751–780 (2020). https://doi.org/10.4310/CMS.2020.v18.n3.a8

    Article  MathSciNet  MATH  Google Scholar 

  13. Koushik, R.: Green’s function on lattices, available at arXiv:1409.7806

  14. Lebowitz, J.L., Bergmann, P.G.: Irreversible Gibbsian ensembles. Ann. Phys. 1(1), 1–23 (1957). https://doi.org/10.1016/0003-4916(57)90002-7

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Lukkarinen, J.: Thermalization in harmonic particle chains with velocity flips. J. Stat. Phys. 155, 1143–1177 (2014). https://doi.org/10.1007/s10955-014-0930-1

    Article  ADS  MathSciNet  MATH  Google Scholar 

  16. Rich, M., Visscher, W.M.: Disordered harmonic chain with self-consistent reservoirs. Phys. Rev. B 11, 2164–2170 (1975)

    Article  ADS  Google Scholar 

  17. Rieder, Z., Lebowitz, J.L., Lieb, E.: Properties of harmonic crystal in a stationary non-equilibrium state. J. Math. Phys. 8, 1073–1078 (1967)

    Article  ADS  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Stefano Olla.

Additional information

Communicated by H. Spohn

Publisher's Note

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

We warmly thank David Huse for very stimulating discussions on the subject. The work of J.L.L. was supported in part by the A.F.O.S.R. He thanks the Institute for Advanced Studies for its hospitality. T.K. acknowledges the support of the NCN grant 2020/37/B/ST1/00426. S.O. has been partially supported by the ANR-15-CE40-0020-01 grant LSD.

Appendices

Appendix A. The proof of Theorem 1.1

Since the result does not depend on the scaling factor \(n^a\), standing by the force \({\mathcal F}(t)\), and the period size \(\theta _n\) we assume that \(a=0\) and \(\theta _n=\theta \). Given a Borel probability measure \(\mu \) on \({\mathbb R}^{2(n+1)}\) (see (1.1)) we denote by \(\big ({\textbf{q}}_{\mu }(t), {\textbf{p}}_{\mu }(t)\big )\) the solution of (1.3)–(1.4)) such that \(\big ({\textbf{q}}_{\mu }(0), {\textbf{p}}_{\mu }(0)\big )\) is distributed according to \(\mu \). Denote then by \(\big (\overline{{\textbf{q}}}_{\mu }(t), \overline{{\textbf{p}}}_{\mu }(t)\big )\) and \(C_{\mu }(t)\) the vector of averages and matrix of the mixed second moments of the solution, correspondingly. They are defined by formulas (2.1) and a \(2\times 2\) block matrix

$$\begin{aligned} C _{\mu } (t)=\left[ \begin{array}{ll} C^{(q)}_{\mu } (t)&{}C^{(q,p)}_{\mu } (t)\\ C^{(p,q)}_{\mu } (t) &{}C^{(p)}_{\mu } (t) \end{array} \right] \end{aligned}$$

Each block is an \((n+1)\times (n+1)\) matrix

$$\begin{aligned}&C^{(q)}_{\mu } (t)=[ q_x(t) q_{x'}(t) ]_{x,x'=0,\ldots ,n}\quad C^{(p)}_{\mu } (t)={\mathbb E}[ p_x(t) p_{x'}(t) ]_{x,x'=0,\ldots ,n},\\&C^{(q,p)}_{\mu } (t)={\mathbb E}[ q_x(t) p_{x'}(t)]_{x,x'=0,\ldots ,n} \end{aligned}$$

and \(C^{(p,q)}_{\mu } (t)=[C^{(q,p)}_{\mu } (t)]^T\), where the initial state is taken to be \(\mu \).

By similar calculation as done in (6.15), their evolution is described by the system of linear differential equations with periodic forcing

$$\begin{aligned} \begin{aligned}&\frac{\textrm{d}}{\textrm{d}t}\left( \begin{array}{c}\overline{\textbf{q}}_{\mu }(t)\\ \overline{\textbf{p}}_{\mu } (t) \end{array}\right) =-A \left( \begin{array}{c}\overline{\textbf{q}}_{\mu } (t)\\ \overline{\textbf{p}}_{\mu } (t) \end{array}\right) + \;{\mathcal F}(t/\theta )\textrm{e}_{p,n+1},\\&\frac{\textrm{d}}{\textrm{d}t}C_{\mu }(t)=-A C_{\mu }(t)- C_{\mu }(t)A^T+\Sigma _2\big (\textbf{c}_{2,\mu }(t)\big )+ {\mathcal F}(t/\theta ) F(t), \end{aligned} \end{aligned}$$
(A.1)

where

$$\begin{aligned} \textbf{c}_{2,\mu }(t)=\left( \begin{array}{c} C^{(p)}_{0,0,\mu }(t) \\ \vdots \\ C^{(p)}_{n,n,\mu }(t) \end{array} \right) \end{aligned}$$
(A.2)

and

$$\begin{aligned} F(t):=\left[ \begin{array}{cc} 0&{} \overline{\textbf{q}}_\mu (t) \otimes \textrm{e}_{p,n+1} \\ &{}\\ \textrm{e}_{p,n+1}\otimes \overline{\textbf{q}}_\mu (t) &{} \ \textrm{e}_{p,n+1}\otimes \overline{\textbf{p}}_\mu (t)+\overline{\textbf{p}}_\mu (t) \otimes \textrm{e}_{p,n+1} \end{array} \right] \end{aligned}$$
(A.3)

Here \(\textrm{e}_{p,n+1}\) and \(\Sigma _2\) are defined in (2.4) and (6.10) respectively. Suppose that we are given a vector \({\overline{X}}\in {\mathbb R}^{2(n+1)}\) and a symmetric non-negative definite \(2(n+1)\times 2(n+1)\) matrix \(S\geqslant {\overline{X}}\otimes {\overline{X}}\). Then, equations (A.1) describe the evolution of the first two moments of the solution of (1.3)–(1.4)) whose initial distribution is a random vector with the first two moments given by \({\overline{X}}\) and S, respectively.

1.1 A.1. The existence and uniqueness of the periodic mean and second moment

In the first step we show the existence of a periodic solution of (A.1) that corresponds to the mean and covariance of a certain probability evolution.

Proposition A.1

There exists a unique vector \(\overline{\textbf{X}}_\textrm{per}=(\overline{\textbf{q}}_{\textrm{per}}, \overline{\textbf{p}}_{\textrm{per}} ) \in {\mathbb R}^{2(n+1)}\) and a non-negative symmetric matrix \(C_\textrm{per}\geqslant \overline{\textbf{X}}_{\textrm{per}}\otimes \overline{\textbf{X}}_{\textrm{per}}\) such that the solution of (A.1) with

$$\begin{aligned} \Big ((\overline{\textbf{q}}(0), \overline{\textbf{p}} (0) ) ,C(0)\Big )=\Big ((\overline{\textbf{q}}_{\textrm{per}}, \overline{\textbf{p}}_{\textrm{per}} ) ,C_{\textrm{per}}\Big ) \end{aligned}$$

satisfies

$$\begin{aligned} \Big ((\overline{\textbf{q}}(0), \overline{\textbf{p}} (0) ) ,C(0)\Big )=\Big ((\overline{\textbf{q}}(\theta ), \overline{\textbf{p}} (\theta ) ) ,C(\theta )\Big ). \end{aligned}$$
(A.4)

In addition, we have

$$\begin{aligned} {C^{(p)}_{x,x}(t)} \geqslant T_-,\quad x=0,\ldots ,n. \end{aligned}$$
(A.5)

The remaining part of this section is devoted to the proof of this results.

1.1.1 A.1.1. The existence of the periodic first moment

Let

$$\begin{aligned} \left( \begin{array}{c}\overline{\textbf{q}}\\ \overline{\textbf{p}} \end{array}\right) := \int _{-\infty }^0 {\mathcal F}(s /\theta ) e^{As}\textrm{e}_{p,n+1}\textrm{d}s. \end{aligned}$$
(A.6)

Thanks to Proposition 2.1 the vector \((\overline{\textbf{q}}, \overline{\textbf{p}} )\) is well defined. One can easily check that the solution of the first equation of (A.1) starting from the vector is given by

$$\begin{aligned} \overline{\textbf{X}}(t)=\left( \begin{array}{c}\overline{\textbf{q}}(t)\\ \overline{\textbf{p}}(t) \end{array}\right) := \int _{-\infty }^t {\mathcal F}(s/\theta ) e^{-A (t-s)}\textrm{e}_{p,n+1}\textrm{d}s. \end{aligned}$$
(A.7)

and is therefore \(\theta \)-periodic. In fact, thanks to Proposition 2.1 the periodic solution has to be unique. Since the coordinates of \(\overline{\textbf{X}}(t)\) satisfy the first equation of (A.1) we conclude that the matrix \(\overline{\textbf{X}}_2(t):=\overline{\textbf{X}}(t)\otimes \overline{\textbf{X}}(t)\) satisfies

$$\begin{aligned} \frac{\textrm{d}}{\textrm{d}t}\overline{\textbf{X}}_2(t)=-A \overline{\textbf{X}}_2(t)- \overline{\textbf{X}}_2(t) A^T+ {\mathcal F}(t /\theta ) F(t), \end{aligned}$$

and it is given by the formula

$$\begin{aligned} \overline{\textbf{X}}_2(t)= \int _{-\infty }^t{\mathcal F}(s/\theta )e^{-A(t-s)} F(s)e^{-A^T(t-s)}\textrm{d}s ,\quad t\in {\mathbb R}. \end{aligned}$$
(A.8)

1.1.2 A.1.2. The existence of the periodic second moment

Now we are going to establish the existence of a periodic second moment. Suppose that C(t) is a periodic solution of the second equation of (A.1). Using the argument made in the proof of Proposition 6.1 we can conclude that it satisfies the equation

$$\begin{aligned} \begin{aligned} C(t) = \int _{-\infty }^te^{-A(t-s)}\left( \Sigma _2\big (\textbf{c}_{2}(s)\big ) + {\mathcal F}(s/\theta ) F(s)\right) e^{-A^T(t-s)}\textrm{d}s\\ =\int _{-\infty }^te^{-A(t-s)}\Sigma _2\big (\textbf{c}_{2}(s)\big ) e^{-A^T(t-s)}\textrm{d}s +\overline{\textbf{X}}_2(t)\\ = \int _{0}^\infty e^{-As}\Sigma _2\big (\textbf{c}_{2}(t-s)\big ) e^{-A^T s} \textrm{d}s +\overline{\textbf{X}}_2(t)\\ = \sum _{\ell =0}^\infty \int _{0}^\theta e^{-A(s+\ell \theta )} \Sigma _2\big (\textbf{c}_{2}(t-s)\big ) e^{-A^T(s+\ell \theta )} \textrm{d}s +\overline{\textbf{X}}_2(t)\\,\quad t\in {\mathbb R}, \end{aligned} \end{aligned}$$
(A.9)

where the matrix \(\Sigma _2\) is defined by (6.10), \(\textbf{c}_{2}(s)\) relates to C(s) via (A.2) and F(s) is defined by (A.3), using \((\overline{\textbf{q}}(t), \overline{\textbf{p}}(t) )\) instead of \((\overline{\textbf{q}}_\mu (t), \overline{\textbf{p}}_\mu (t) )\). Conversely, any periodic symmetric matrix valued function C(t) satisfying (A.9) is a periodic solution to the second equation of (A.1).

For \(x,x',y=0,\ldots ,n\) define

$$\begin{aligned} {g_{x,x',y}(s):=\sum _{\ell =0}^{+\infty } \Big [e^{-A(s+\ell \theta )}\Big ]_{x+n+1,y+n+1}\Big [e^{-A^T(s+\ell \theta )}\Big ]_{y+n+1,x'+n+1}.} \end{aligned}$$
(A.10)

Consider the following linear mapping: \({\mathcal L}:[C(\mathbb T_\theta )]^{n+1}\rightarrow [C(\mathbb T_\theta )]^{n+1}\), where \(\mathbb T_\theta :=\theta \mathbb T\) is the torus of size \(\theta \), that assigns to a given vector of \(\theta \)-periodic functions \(\textbf{T}(s)=[T_{0}(s),\ldots ,T_n(s)]\) a vector valued function

$$\begin{aligned} {\mathcal L}{} \textbf{T} :=({\mathfrak {G}}_0\textbf{T},\ldots , {\mathfrak {G}}_n\textbf{T}), \end{aligned}$$
(A.11)

where

$$\begin{aligned} {\mathfrak {G}}_x\textbf{T}(t) = \sum _{y=0}^n \int _0^{\theta }{{\mathfrak {G}}_{x,y}(s)} T_y\left( t-s\right) \textrm{d}s. \end{aligned}$$
(A.12)

Here

$$\begin{aligned} {{\mathfrak {G}}_{x,y}(s)}= 4 \gamma g_{x,x,y}(s),\quad y=0,\ldots ,n. \end{aligned}$$
(A.13)

Obviously, from (A.10), we have \({\mathfrak {G}}_{x,y}(s)\geqslant 0\). Note also that although \( g_{x,x',y}(\cdot )\) need not be \(\theta \)-periodic the functions \({\mathfrak {G}}_x\textbf{T}(t)\), \(x=0,\ldots ,n\) are \(\theta \)-periodic. In addition, if C(t) satisfies (A.9), then

$$\begin{aligned} \textbf{c}_{2}(t) ={\mathcal L}{\mathcal T}(\textbf{c}_{2})(t) +\overline{\textbf{p}}^2(t), \end{aligned}$$
(A.14)

where for a given \(\textbf{T}^T=(T_0,\ldots ,T_n)\in {\mathbb R}^{n+1}\)

$$\begin{aligned} {\mathcal T} (\textbf{T})= \left( \begin{array}{c} T_-\\ T_1 \\ \vdots \\ T_n \end{array} \right) ,\quad \overline{\textbf{p}}^2(t) = \left( \begin{array}{c} {\overline{p}}_0^2(t)\\ \vdots \\ {\overline{p}}_n^2(t) \end{array} \right) . \end{aligned}$$

Conversely, by finding a solution \(\textbf{c}_{2}\) of (A.14) one can define then a \(\theta \)-periodic function C(t) by the right hand side of (A.9). The entries of the function corresponding to \(C_{x,x}\), \(x=n+1,\ldots ,2n+1\) coincide with the coordinates of the vector \(\textbf{c}_{2}\), by virtue of (A.14). Thus, the function C(t) solves equation (A.9). We have reduced therefore the problem of finding a periodic solution to the second equation of (A.1) to solving equation (A.14).

1.1.3 A.1.3. Solution of (A.14)

Let

$$\begin{aligned} {\mathcal C}_+:=[\textbf{T}=(T_0,T_1,\ldots ,T_n):\,T_x\in C(\mathbb T_\theta )\,\text{ and } T_x\geqslant T_-,\,x=0,\ldots ,n]. \end{aligned}$$

It is a closed subset of \(\Big (C(\mathbb T_\theta )\Big )^{n+1}\), equipped with the norm

$$\begin{aligned} |\!\Vert \textbf{T}\Vert \!|:=\max \{ \Vert T_x\Vert _\infty ,\,x=0,\ldots ,n\}. \end{aligned}$$

Consider the mapping

$$\begin{aligned} {\mathfrak {T}}=({\mathfrak {T}}_0,\ldots , {\mathfrak {T}}_n): {\mathcal C}_+\rightarrow \Big (C(\mathbb T_\theta )\Big )^{n+1}, \quad \text{ where } \quad {\mathfrak {T}}{} \textbf{T}:={\mathcal L}{\mathcal T}(\textbf{T})+\overline{\textbf{p}}^2. \end{aligned}$$
(A.15)

Using the notation of (A.12) and (A.14) we have

$$\begin{aligned} {\mathfrak {T}}_x(\textbf{T})(t):=T_-\int _0^{\theta } {\mathfrak {G}}_{x,0}(s)\textrm{d}s+\sum _{x'=1}^n\int _0^\theta {\mathfrak {G}}_{x,x'}(s)T_{x'}\left( t-s\right) \textrm{d}s+{\overline{p}}_x^2(t),\quad x=0,\ldots ,n. \end{aligned}$$

Comparing (6.12) with (A.9), after time averaging over a period, it is easy to identify

$$\begin{aligned} \int _0^{\theta } {\mathfrak {G}}_{x,y}(s)\textrm{d}s = M_{x,y} \end{aligned}$$
(A.16)

defined by (7.2). The matrix \([M_{x,y}]_{x,y=0}^n\) is symmetric, bi-stochastic (as can be easily seen from (7.2)). It also follows immediately that

$$\begin{aligned} 1= \sum _{y=0}^n \int _0^{\theta } {\mathfrak {G}}_{x,y}(s)\textrm{d}s,\quad x=0,1,\ldots ,n \end{aligned}$$
(A.17)

and, as a consequence, \({\mathfrak {T}}\big ({\mathcal C}_+\big )\subset {\mathcal C}_+\). Furthermore, we claim that \(M_{x,y}>0\) for all \(x,y=0,\ldots ,n\). Indeed, a simple calculation, using (2.3) and (2.5), yields

$$\begin{aligned}&\big [ (\lambda +A)^{-1}\big ]_{x+n+1,y+n+1}= \sum _{j=0}^n\frac{\lambda \psi _j(x)\psi _j(y) }{\lambda ^2+2\gamma \lambda +\mu _j},\nonumber \\&\big [ (\lambda +A)^{-1}\big ]_{x,y+n+1}= \sum _{j=0}^n\frac{ \psi _j(x)\psi _j(y)}{\lambda ^2+2\gamma \lambda +\mu _j}, \nonumber \\&\big [ (\lambda +A)^{-1}\big ]_{x+n+1,y} =-\sum _{j=0}^n\frac{ \mu _{j}\psi _j(x)\psi _j(y) }{\lambda ^2+2\gamma \lambda +\mu _j}, \nonumber \\&\big [ (\lambda +A)^{-1}\big ]_{x,y+n+1}= \sum _{j=0}^n\frac{ \psi _j(x)\psi _j(y)}{\lambda ^2+2\gamma \lambda +\mu _j}. \end{aligned}$$
(A.18)

The poles of the meromorphic functions appearing in (A.18) are given by

$$\begin{aligned} \lambda _{j,\pm }=-\Big (\gamma \pm \sqrt{\gamma ^2-\mu _j}\Big ). \end{aligned}$$
(A.19)

Suppose that \(M_{x,y}=0\) for some xy. From (A.16) we conclude then that

$$\begin{aligned} 0= M_{x,y}=\int _0^{\theta } {\mathfrak {G}}_{x,y}(s)\textrm{d}s =4\gamma \int _0^{+\infty }\Big [e^{-As}\Big ]^2_{x+n+1,y+n+1}\textrm{d}s, \end{aligned}$$
(A.20)

which in turn would implies that \(\Big [e^{-As}\Big ]_{x+n+1,y+n+1}\equiv 0\) for all \(s\geqslant 0\), thus also

$$\begin{aligned} 0\equiv \big [ (\lambda +A)^{-1}\big ]_{x+n+1,y+n+1}= \sum _{j=0}^n\frac{\lambda \psi _j(x)\psi _j(y) }{\lambda ^2+2\gamma \lambda +\mu _j}. \end{aligned}$$

As a result, we conclude that \( \psi _j(x)\psi _j(y)=0\), for all \(j=0,\ldots ,n\), which is impossible.

We shall show that the mapping \({\mathfrak {T}}\) has a unique fixed point in \({\mathcal C}_+\) by proving that the mapping is a contraction in the norm \(|\!\Vert \cdot \Vert \!|\). Indeed, for \(\textbf{T}_j^T=[T_{j,0},T_{j,1},\ldots ,T_{j,n}]\), \(j=1,2\), we have

$$\begin{aligned}&|{\mathfrak {T}}_x(\textbf{T}_1)(t) -{\mathfrak {T}}_x(\textbf{T}_2)(t)|=\Big | \sum _{x'=1}^n\int _0^\theta {\mathfrak {G}}_{x,x'}(s)\Big [T_{1,x'}\left( t-s\right) - T_{2,x'}\left( t-s\right) \Big ]\textrm{d}s\Big |\\&\quad \leqslant \sum _{x'=1}^n\int _0^\theta {\mathfrak {G}}_{x,x'}(s) \Big | T_{1,x'}\left( t-s\right) - T_{2,x'}\left( t-s\right) \Big |\textrm{d}s\\&\quad \leqslant \sum _{x'=1}^n \left( \int _0^\theta {\mathfrak {G}}_{x,x'}(s)\textrm{d}s\right) \Vert T_{1,x'} - T_{2,x'}\Vert _{\infty }\\&\quad \leqslant \left( 1- \int _0^\theta {\mathfrak {G}}_{x,0}(s)\textrm{d}s\right) |\!\Vert \textbf{T}_1-\textbf{T}_2|\!\Vert ,\quad x=0,\ldots ,n. \end{aligned}$$

Therefore

$$\begin{aligned} |\!\Vert {\mathfrak {T}}(\textbf{T}_1)-{\mathfrak {T}}(\textbf{T}_2)\Vert \!| \leqslant \rho |\!\Vert \textbf{T}_1-\textbf{T}_2|\!\Vert , \end{aligned}$$

where

$$\begin{aligned} \rho :=\max \Big [ 1- \int _0^\theta {\mathfrak {G}}_{x,0}(s)\textrm{d}s ,\,x=0,\ldots ,n\Big ]<1. \end{aligned}$$

We have proven that \( \Vert {\mathfrak {T}}(\textbf{T}_1) -{\mathfrak {T}} (\textbf{T}_2)\Vert _\infty \leqslant \rho \Vert \textbf{T}_1-\textbf{T}_2\Vert _{\infty }\) and the existence of a unique fixed point follows. This ends the proof of Proposition A.1. \(\square \)

1.2 A.2. The end of the proof of Theorem 1.1

Suppose now that \(\nu \) is a probability law whose first and second moments are \(\theta \)-periodic, e.g. it could be a Gaussian distribution with the mean and the second moment given by \(\textbf{P}_{\textrm{per}}\) and \(C_{\textrm{per}}\), respectively. Denote by

$$\begin{aligned} {\mathcal P}_{s,t}F(\textbf{q}, \textbf{p})=\int _{{\mathbb R}^{2(n+1)}}F(\textbf{q}', \textbf{p}'){\mathcal P}_{s,t}(\textbf{q}, \textbf{p}; \textrm{d}\textbf{q}', \textrm{d}\textbf{p}') \end{aligned}$$

the evolution family of transition probability operators corresponding to the dynamics described by (1.3)–(1.4). Consider the event \( E:=[N_x({\theta })=0,\,x=1,\ldots ,n]. \) We have \({\mathbb P}[E]>0\). Suppose that the dynamics starts at \((\textbf{q}, \textbf{p})\). Then, for any \(F\geqslant 0\) we can write

$$\begin{aligned} {\mathcal P}_{0,\theta }F(\textbf{q}, \textbf{p})\geqslant {\mathbb E}\Big [F\big (\textbf{q}(\theta ), \textbf{p}(\theta )\big ),\,E\Big ]={\mathbb P}[E]{\mathcal Q}_{0, \theta }F(\textbf{q}, \textbf{p}), \end{aligned}$$
(A.21)

where \({\mathcal Q}_{s,t}\) is the transition probability operator for the non-homogeneous Ornstein-Uhlenbeck dynamics that corresponds to the generator \( {\mathcal {G}}_t^{(g)} = {\mathcal {A}}_t + 2 \gamma S_{-} \), see (1.7) and (1.8). Using the hypoellipticity of \({\mathcal {G}}_t^{(g)}\), see [9, Section A.3], one can prove that there exist strictly positive transition probability density kernels \(\rho _{s,t}\) corresponding to \({\mathcal Q}_{s,t}\), see [9, Section A.2] for more details. Thanks to (A.21) we conclude that

$$\begin{aligned} {\mathcal P}_{0, \theta }(\textbf{q}, \textbf{p}; \textrm{d}\textbf{q}', \textrm{d}\textbf{p}') \geqslant c_*\rho _{0,\theta }(\textbf{q}, \textbf{p}; \textbf{q}', \textbf{p}') {\textrm{d}} \textbf{q}' \textrm{d}\textbf{p}', \end{aligned}$$
(A.22)

where \(c_*:={\mathbb P}[E]\). Then, \(\nu _{0,t}:=\nu {\mathcal P}_{0,t}\) describes the law of \((\textbf{q}(t), \textbf{p}(t) )\) with the prescribed initial data. Thanks to Proposition A.1 we can see that the total energy \({\mathcal H}(t):=\sum _{x=0}^n{\mathcal E}_x(t)\) (see (1.2)) is a Lyapunov function for the above system, since \({\mathbb E}{\mathcal H}(t)\) is \(\theta \)-periodic. The above implies that the family of laws \(\{\nu _{0,t},\,t\geqslant 0\}\) is tight in \({\mathbb R}^{2(n+1)}\). Thus, also the family \(\mu _N:=N^{-1}\int _0^{N\theta }\nu _{0,s}\textrm{d}s\) is tight. Suppose that \(\mu _\infty \) is its limiting measure, i.e. there exists a sequence \(N'\rightarrow +\infty \) such that \(\mu _{N'}\rightarrow \mu _\infty \), in the topology of weak convergence. Since \({\mathcal P}_{s,t}\) has the Feller property one can easily conclude that \(\mu _\infty {\mathcal P}_{0,\theta }=\mu _\infty \). Hence \(\mu _s^P:=\mu _\infty {\mathcal P}_{0,s}\) , \(s\in [0,+\infty )\) is a periodic stationary state.

Suppose that \(\mu (\textrm{d}\textbf{q}, \textrm{d}\textbf{p})=f (\textbf{q}, \textbf{p}) \textrm{d}\textbf{q} \textrm{d}\textbf{p}\), where f is a \(C^\infty \) smooth probability density. One can show, using the regularity theory of stochastic differential equations, that \(\mu {\mathcal P}_{0,\theta }\) is absolutely continuous w.r.t. the Lebesgue measure and its density is also \(C^\infty \) smooth, see e.g. [6, Corollary III.3.4, p. 303]. This allows us to conclude further that \(\mu {\mathcal P}_{0,\theta }\) is absolutely continuous, provided that \(\mu \) is absolutely continuous. We shall denote by \({\mathcal P}_{0,\theta }\) the corresponding operator induced on \(L^1({\mathbb R}^{2(n+1)})\). The operator \({\mathcal Q}_{0,\theta }\) corresponding to the Gaussian dynamics transforms \(\mu _\infty \) into an absolutely continuous measure. Thanks to (A.22) we conclude that

$$\begin{aligned} \mu _\infty (\textrm{d}\textbf{q}', \textrm{d}\textbf{p}')=\mu _\infty {\mathcal P}_{0,\theta }(\textrm{d}\textbf{q}', \textrm{d}\textbf{p}')\geqslant c_*\mu _\infty {\mathcal Q}_{0,\theta }( \textbf{q}', \textbf{p}')\textrm{d}\textbf{q}' \textrm{d}\textbf{p}'. \end{aligned}$$
(A.23)

Therefore the singular part of \(\mu _\infty \) is of at most mass \(1-c_*\). Since \({\mathcal P}_{0,\theta }\) transforms the space of abolutely continuous measures into itself, both the singular and absolutely continuous parts of \(\mu _\infty \), after normalization, become invariant under \({\mathcal P}_{0,\theta }\). Iterating this procedure we conclude, after m steps, that the singular part can be of at most mass \((1-c_*)^m\), which eventually leads to the conclusion that the measure \(\mu _\infty \) is absolutely continuous. The respective density is positive, due to (A.22). This ends the proof of Theorem 1.1. \(\square \)

Appendix B. Green Functions Convergence

Recall that \(G_{\omega _0}\) and \(G^n_{\omega _0}\) are the Green’s functions corresponding to \(\omega _0^2-\Delta \) and \(\omega _0^2-\Delta _{\textrm{N}}\), where \(\Delta \) is the free lattice laplacian on \({\mathbb {Z}}\) and \(\Delta _{\textrm{N}}\) is the Neumann discrete laplacian on \(\{0,1,\dots , n\}\), see Sects. 2.3 and 2.4, respectively.

1.1 B.1. Estimantes on oscillating sums

Define \(\chi _n(x)\) as the \(n+1\)-periodic extension of \(\chi _n(x):=(1+x)\wedge (n+2-x)\), \(x\in [0,n+1]\). Suppose that \(\Phi :{\mathbb R}^2\rightarrow {\mathbb {C}}\) is a \(\theta ,\theta '\)-periodic function in each variable respectively. Denote

$$\begin{aligned} H_{x,x'}^{(n)} =\frac{1}{(n+1)^2}\sum _{j,j'=0}^n\Phi \left( \frac{\theta j}{n+1}, \frac{\theta ' j'}{n+1}\right) \exp \left\{ \frac{ 2i\pi j x}{n+1}\right\} \exp \left\{ \frac{ 2i\pi j'x'}{n+1}\right\} \end{aligned}$$

for \(x,x'\in {\mathbb Z}\).

Lemma B.1

Suppose that \(\Phi \) is of \(C^k\)-class for some \(k\geqslant 1\). Then, there exists C such that

$$\begin{aligned} \Big |H_{x,x'}^{(n)}\Big |\leqslant \frac{C}{\chi ^\ell _n(x) \chi ^{\ell '}_n(x')},\quad x,x'\in {\mathbb Z},\,n\geqslant 1,\,\ell ,\ell '\geqslant 0,\,\ell +\ell '\leqslant k. \end{aligned}$$
(B.1)

Proof

To simplify the notation we suppose that \(\theta =\theta '=1\). Summation by parts yields

$$\begin{aligned} H_{x,x'}^{(n)}&=\frac{1}{(n+1)^2}\sum _{j,j'=0}^n\Phi \left( \frac{ j}{n+1}, \frac{ j'}{n+1}\right) \frac{\exp \left\{ \frac{ 2 i\pi x}{n+1}\right\} -1}{\exp \left\{ \frac{ 2 i\pi x}{n+1}\right\} -1}\exp \left\{ \frac{ 2 i\pi jx}{n+1}\right\} \exp \left\{ \frac{ 2i\pi j'x'}{n+1}\right\} \\&=\frac{1}{(n+1)^2 [\exp \left\{ \frac{ i\pi x}{n+1}\right\} -1]} \sum _{j,j'=0}^n\Big [\Phi \left( \frac{ j-1}{n+1}, \frac{ j'}{n+1}\right) - \Phi \left( \frac{ j}{n+1}, \frac{ j'}{n+1}\right) \Big ] \\&\quad \times \exp \left\{ \frac{ 2 i\pi jx}{n+1}\right\} \exp \left\{ \frac{ 2 i\pi j'x}{n+1}\right\} \end{aligned}$$

Since \(\Phi \) is of \(C^1\) class

$$\begin{aligned} \Big |\Phi \left( \frac{ j-1}{n+1}, \frac{ j'}{n+1}\right) - \Phi \left( \frac{ j}{n+1}, \frac{ j'}{n+1}\right) \Big |\leqslant \frac{C}{n+1} \end{aligned}$$

for some constant \(C>0\). In addition, there exists \(c>0\) such that

$$\begin{aligned} \left| \exp \left\{ \frac{ 2 i\pi x}{n+1}\right\} -1\right| \geqslant \frac{c\chi _n(x) }{n+1} \end{aligned}$$

for \(x,x'\in {\mathbb Z},\,n\geqslant 1\). Thus, there exists \(C>0\) such that

$$\begin{aligned} |H_{x,x'}^{(n)}|\leqslant \frac{C}{\chi _n(x)},\quad x,x'\in {\mathbb Z}. \end{aligned}$$

Iterating this argument in the regularity degree k of \(\Phi \) we conclude (B.1). \(\square \)

1.2 B.2. Application

An application concerns the approximation of the Green’s function \(G_{\omega _0}\) by \(G_{\omega _0}^n\) along the diagonal.

Lemma B.2

We have

$$\begin{aligned} G_{\omega _0}^n(y,y)=G_{\omega _0}(0)+{\widetilde{H}}^{(n)}(y) +O\Big (\frac{1}{n}\Big ),\quad y=0,\ldots ,n,\,n\geqslant 1. \end{aligned}$$
(B.2)

Here for some constant \(C>0\) we have

$$\begin{aligned} |{\widetilde{H}}^{(n)}(y)|\leqslant \frac{C}{\chi ^2(y)} ,\quad y=0,\ldots ,n,\,n\geqslant 1. \end{aligned}$$
(B.3)

Proof

Using the definition of the Green’s function (2.13) (with \(\ell =0\)) and formulas (2.14) we obtain

$$\begin{aligned} G_{\omega _0}^n(y,y) =\frac{ 1}{n+1}\sum _{j=0}^n \Xi \left( \frac{ j}{n+1}\right) \left[ 1+\cos \left( \frac{\pi j(2y+1)}{n+1}\right) \right] +O\Big (\frac{1}{n}\Big ), \end{aligned}$$

where

$$\begin{aligned} \Xi \left( u \right) =\left\{ 4\sin ^2\left( \frac{\pi u}{2}\right) +\omega _0^2 \right\} ^{-1}. \end{aligned}$$

As a result we write \(G_{\omega _0}^n(y,y)\) in the form (B.2), with

$$\begin{aligned} {\widetilde{H}}_y^{(n)}=\frac{ 1}{2(n+1)}\sum _{j=-n-1}^n\cos \left( \frac{\pi j(2y+1)}{n+1}\right) \Xi \left( \frac{ j}{n+1}\right) . \end{aligned}$$

Estimate (B.3) is then a consequence of Lemma B.1. \(\square \)

Appendix C. Proofs of Lemmas 9.29.3 and 9.5

1.1 C.1. Proof of Lemma 9.2

For \(m\in {\mathbb Z}\) and \(g=(g_0,\ldots ,g_n)\in {\mathbb R}^{n+1}\) define

$$\begin{aligned} S(m):=\int _0^{+\infty }e^{-2\pi i ms/\theta }e^{-As}\Sigma (g) e^{-A^Ts} \textrm{d}s= \left[ \begin{array}{cc} S^{q}(m)&{} S^{qp}(m)\\ &{}\\ S^{pq}(m)&{} S^{p}(m) \end{array} \right] , \end{aligned}$$

where \(\Sigma _2\) is defined in (6.10). Note that

$$\begin{aligned} \sum _{y=0}^nM_{x,y}(m)g_y=S^{p}_{x,x}(m). \end{aligned}$$
(C.1)

Arguing similarly as in the proof of (6.16) we get

$$\begin{aligned} A S(m)+ S(m) A^T +2i\alpha _mS(m)=\Sigma _2(g), \end{aligned}$$
(C.2)

where \(\alpha _m:=\pi m/\theta \). Denote

$$\begin{aligned} \widetilde{ S}^{q,p}_{j,j'}(m)=\sum _{x,x'=0}^nS^{q,p}_{x,x'}(m)\psi _j(x) \psi _{j'}(x') \end{aligned}$$

Following the same manipulations as those leading to (6.20) we obtain

$$\begin{aligned}&{\widetilde{S}}^{p}_{j,j'} (m)=\frac{1}{2} \widetilde{S}^{q}_{j,j'}(m) \Big [\mu _{j'}+\mu _j + 2(\gamma +i\alpha _m) i\alpha _m \Big ] , \nonumber \\&{\widetilde{S}}^{q}_{j,j'}(m)\Big [\mu _{j'}-\mu _j -4 i\alpha _m (\gamma +i\alpha _m)\Big ]+4(\gamma +i\alpha _m) {\widetilde{S}}^{qp}_{j,j'} (m) =0,\nonumber \\&{\widetilde{S}}^{qp}_{j,j'}(m)(\mu _{j}-\mu _{j'} )+ 2 i\alpha _m\widetilde{S}^{q}_{j,j'}(m)\mu _{j'} +2(2\gamma +i\alpha _m){\widetilde{S}}^{p}_{j,j'}(m)= {\widetilde{F}}_{j,j'} \end{aligned}$$
(C.3)

and \( {\widetilde{F}}_{j,j'}=\sum _{x=0}^n\psi _j(x)\psi _{j'}(x)g_x. \) Solving the above system using the procedure used to deal with (6.20) we obtain

$$\begin{aligned} {\widetilde{S}}^{p}_{j,j'}(m)=\sum _{y}\Theta _m(\mu _j,\mu _{j'})\psi _j(y)g_y\psi _{j'}(y), \end{aligned}$$

with

$$\begin{aligned}&\Theta _m(c,c') :=\frac{2\gamma }{2\gamma +i\alpha _m} \Bigg \{\Big [ (2\gamma +i\alpha _m)\Big (c+c' +2 (\gamma +i\alpha _m)i\alpha _m\Big ) \Big ]^{-1} \nonumber \\&\quad \quad \quad \quad \quad \quad \times \Big [\frac{1}{4}(\gamma +i\alpha _m)^{-1}(c-c')^2 + i\alpha _m(c+c') \Big ] +1 \Bigg \}^{-1} . \end{aligned}$$
(C.4)

Note that \(\Theta _0(c,c')=\Theta (c,c')\) defined in (6.24). From (C.1) we conclude that

$$\begin{aligned} M_{x,y}(m)= \sum _{j,j'=0}^n\Theta _m(\mu _j,\mu _{j'}) \psi _j(x)\psi _{j'}(x) \psi _j(y)\psi _{j'}(y) \end{aligned}$$
(C.5)

As in (7.5), for any sequence \((f_x)\in {\mathbb {C}}^{n+1}\) we can write

$$\begin{aligned} \sum _{x,y=0}^n(\delta _{x,y}-M_{x,y}(m))f_y^\star f_x = \sum _{j,j'=0}^n \left( 1- \Theta _m(\mu _j,\mu _{j'})\right) \left| \sum _{x=0}^n \psi _j(x)f_x\psi _{j'}(x)\right| ^2. \end{aligned}$$

We have

$$\begin{aligned}&1- \Theta _m(c,c')\\&\quad = \Bigg \{i\alpha _m+\Big [c+c' +2 (\gamma +i\alpha _m) i\alpha _m\Big ]^{-1} \Big [\frac{1}{4}(\gamma +i\alpha _m)^{-1}(c-c')^2 + i\alpha _m (c+c') \Big ]\Bigg \}\\&\quad \quad \times \Bigg \{\Big [ c+c' +2 (\gamma +i\alpha _m) i\alpha _m \Big ]^{-1} \Big [\frac{1}{4}\Big (\gamma +i\alpha _m\Big )^{-1}(c-c')^2 + i\alpha _m (c+c') \Big ] +(2\gamma +i\alpha _m) \Bigg \}^{-1}. \end{aligned}$$

Thus \( \lim _{m\rightarrow +\infty }\Big (1- \Theta _m(c,c')\Big )=1. \) On the other hand, if \(m\not =0\), an easy calculation shows that \(1- \Theta _m(c,c')=0\) implies that \( (c-c')^2=8\alpha ^2_m(\alpha ^2_m+\gamma ^2)\) and \(\quad c+c'=2\alpha ^2_m\), where \(\alpha _m=\pi m/\theta \). But this would clearly lead to a contradiction, as then we would have \(|c-c'|>\sqrt{2}(c+c')\), which is clearly impossible (remember that \(c,c'>0\)). Hence, there exists \( {\mathfrak {C}}_*>0\) such that

$$\begin{aligned} |1- \Theta _m(c,c')|\geqslant {\mathfrak {C}}_*,\quad |m|\geqslant 1,\,c,c'\in [0,\omega _0^2+4]. \end{aligned}$$

This ends the proof of (9.9). \(\square \)

1.2 C.2. Proof of Lemma  9.3

Using (C.5) we obtain

$$\begin{aligned} \sum _{x=0}^n|M_{x,0}(m)|^2= \sum _{j_1,\ldots ,j_4 =0}^n\Theta _m(\mu _{j_1},\mu _{j_2}) \Theta _m(\mu _{j_3},\mu _{j_4}) \prod _{k=1}^4\psi _{j_k}(0) \sum _{x=0}^n\prod _{k=1}^4\psi _{j_k}(x) . \end{aligned}$$
(C.6)

Applying elementary trigonometric identities we conclude that

$$\begin{aligned}&\sum _{x=0}^n\prod _{k=1}^4\psi _{j_k}(x) =\frac{1}{(n+1)^2} \left\{ \prod _{k=1}^4(2-\delta _{0,j_k})\right\} ^{1/2} \sum _{x=0}^n\prod ^4_{k=1}\cos \left( \frac{\pi j_k(2x+1)}{2(n+1)}\right) \\&\quad =\frac{1}{2^5(n+1)} \left\{ \prod _{k=1}^4(2-\delta _{0,j_k})\right\} ^{1/2} \sum _{\iota _1,\ldots ,\iota _4\in \{-1,1\}}\cos \left( \frac{\pi }{2(n+1)}\sum _{k=1}^4\iota _kj_k\right) 1_{2(n+1){\mathbb Z}}\Big (\sum _{k=1}^4\iota _kj_k\Big ) \end{aligned}$$

Therefore we can write

$$\begin{aligned}&\sum _{x=0}^n|M_{x,0}(m)|^2= \sum _{\iota ,\iota _1\in \{-1,1\}}\sum _{\iota ',\iota _1'\in \{-1,1\}}\int _0^1\textrm{d}u\int _0^1\textrm{d}u' \int _0^1\textrm{d}u_1\int _0^1\textrm{d}u'_1 {\mathfrak {V}}_{m}(u,u'){\mathfrak {V}}_{m}^\star (u_1,u'_1)\end{aligned}$$
(C.7)
$$\begin{aligned}&\quad \times \exp \left\{ \textrm{i}\pi \Big ( \iota u+\iota 'u'+\iota _1 u_1+\iota '_1u'_1\Big )\right\} \sum _{q\in {\mathbb Z}}\delta _q\Big (\iota u+\iota 'u'+\iota _1 u_1+\iota '_1u'_1\Big )+O_m\left( \frac{1}{n}\right) , \end{aligned}$$
(C.8)

where \(O_m\left( \frac{1}{n}\right) \leqslant \frac{C}{n}\) for some constant \(C>0\), independent of n and m, and

$$\begin{aligned} {\mathfrak {V}}_{m}(u,u'):= \Theta _m\Big (\omega _0^2+4\sin ^2\left( \frac{\pi u}{2}\right) , \omega _0^2+4\sin ^2\left( \frac{\pi u'}{2}\right) \Big )\cos \left( \frac{\pi u}{2}\right) \cos \left( \frac{\pi u'}{2}\right) . \end{aligned}$$

We claim that there exists \(C>0\), independent of n and m, such that \({\mathfrak {V}}_{m}(u,u')\leqslant C\) for all \(u,u'\in [0,\omega _0^2+4]\). Indeed, as can be seen directly from (C.4), we have \(\lim _{m\rightarrow +\infty }\Theta _m(c,c')=0\) uniformly in \(c,c'\in [0,\omega _0^2+4]\). On the other hand the function \({\mathbb R}\times [0,\omega _0^2+4]^2\ni (m,c,c')\rightarrow \Theta _m(c,c')\) is bounded on compact set. If otherwise, this would imply that there exist \((m,c,c')\in {\mathbb R}\times [0,\omega _0^2+4]^2\) such that

$$\begin{aligned}&0=\Big [ (2\gamma +i\alpha _m)\Big (c+c' +2 (\gamma +i\alpha _m) i\alpha _m\Big ) \Big ]^{-1} \\&\quad \times \Big [\frac{1}{4}(\gamma +i\alpha _m)^{-1}(c-c')^2 + i\alpha _m (c+c') \Big ] +1. \end{aligned}$$

An easy calculation gives \(c+c'= 2\alpha ^2_m\) and \( (c-c')^2=8(\alpha ^2_m+\gamma ^2) \alpha ^2_m,\) where \(\alpha _m=\pi m/\theta \). This leads to a contradiction, as then \(|c-c'|>\sqrt{2}(c+c')\) (but both \(c,c'>0\)). Thus the conclusion of the lemma follows. \(\square \)

1.3 C.3. Proof of Lemma  9.4

From (A.18) we obtain

$$\begin{aligned} {[}e^{-At}]_{x+n+1,x'+n+1}= \sum _{j=0}^nE_j(t)\psi _j(x)\psi _j(x'), \end{aligned}$$

where (cf (6.18))

$$\begin{aligned} E_j(t):=\frac{1}{2\sqrt{\gamma ^2-\mu _j}}\Big [ -\lambda _{j,+} \exp \left\{ \lambda _{j,+}t\right\} + \lambda _{j,-} \exp \left\{ \lambda _{j,-}t\right\} \Big ],\quad \text{ if } \mu _j\not = \gamma ^2. \end{aligned}$$

In the case \(\mu _j= \gamma ^2\) (then \(\lambda _{j,\pm }=\gamma \), cf (A.19)) we have \( E_j(t):=(1-\gamma t)e^{-\gamma t} . \) Using (A.6) we obtain therefore

$$\begin{aligned} {\overline{p}}_x(t)&= \frac{1}{n^{1/2}} \sum _{j=0}^n\int _{0}^{+\infty } {\mathcal F}((t-s)/\theta ) [e^{-A s}]_{x+n+1,2n+1} \textrm{d}s \nonumber \\&=\frac{1}{n^{1/2}} \sum _{j=0}^n\psi _j(x)\psi _j(n) \int _{0}^{+\infty } {\mathcal F}((t-s)/\theta )E_j(s)\textrm{d}s. \end{aligned}$$
(C.9)

From (C.9) we conclude that there exists \({\mathfrak {p}}_*>0\) such that

$$\begin{aligned} \sup _{t\in {\mathbb R},x=0,\ldots ,n}|{\overline{p}}_x(t)|\leqslant \frac{{\mathfrak {p}}_*}{n^{1/2}},\quad n=1,2,\ldots . \end{aligned}$$
(C.10)

Estimate (9.11) is then a straightforward consequence of (C.10) and (4.11). \(\square \)

1.4 C.4. Proof of Lemma  9.5

Multiplying both sides of (9.4) by \(V_x(t)\) and averaging over time we get

$$\begin{aligned} \langle \langle V_x^2\rangle \rangle&= \sum _{x'=1}^n \int _0^\theta {\mathfrak {g}}_{x,x'}(s) \langle \langle V_x(\cdot )V_{x'}(\cdot -s)\rangle \rangle \textrm{d}s+ \langle \langle V_x v_x\rangle \rangle \nonumber \\&\leqslant \sum _{x'=1}^n M_{x,x'} \langle \langle V_x^2\rangle \rangle ^{1/2}\langle \langle V_{x'}^2\rangle \rangle ^{1/2}+ \langle \langle V_x v_x\rangle \rangle . \end{aligned}$$
(C.11)

Summing up over x we obtain

$$\begin{aligned}&\sum _{x,x'=0}^n \Big (\delta _{x,x'} -M_{x,x'}\Big ) \langle \langle V_x^2\rangle \rangle ^{1/2}\langle \langle V_{x'}^2\rangle \rangle ^{1/2}+\langle \langle V_0^2\rangle \rangle ^{1/2}\sum _{x=0}^n M_{x,0} \langle \langle V_{x}^2\rangle \rangle ^{1/2}\\&\quad \leqslant \sum _{x=0}^n \langle \langle V_x v_x\rangle \rangle . \end{aligned}$$

Using (7.4) and the Cauchy-Schwarz inequality we obtain in particular that

$$\begin{aligned} M_{0,0}\langle \langle V_0^2\rangle \rangle \leqslant \left\{ \sum _{x=0}^n \langle \langle V_x^2 \rangle \rangle \right\} ^{1/2}\left\{ \sum _{x=0}^n \langle \langle v_x^2 \rangle \rangle \right\} ^{1/2}\leqslant \frac{C}{n}\left\{ \sum _{x=0}^n \langle \langle V_x \rangle \rangle ^2\right\} ^{1/2} \end{aligned}$$

for some \(C>0\) independent of n. The last estimate follows from (9.11). To finish the proof note that from (7.2) we have

$$\begin{aligned} M_{0,0}&=\frac{4}{(n+1)^2}\sum _{j,j'=0}^n(1-\delta _{0,j})(1-\delta _{0,j'})\Theta (\mu _j,\mu _{j'})\cos ^2\left( \frac{\pi j}{2(n+1)}\right) \cos ^2\left( \frac{\pi j'}{2(n+1)}\right) \\&\approx 4\gamma ^2\int _0^1\int _0^1 \frac{\Big [\omega _0^2+2\sin ^2(\pi u/2)+2\sin ^2(\pi u'/2)\Big ]\cos ^2(\pi u/2)\cos ^2(\pi u'/2)\textrm{d}u\textrm{d}u'}{\gamma ^2 \Big (\omega _0^2+2\sin ^2(\pi u/2)+2\sin ^2(\pi u'/2) \Big )+ \Big (\sin ^2(\pi u/2)-\sin ^2(\pi u'/2))\Big )^2}, \end{aligned}$$

where the equality holds up to a term of order O(1/n). \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Komorowski, T., Lebowitz, J.L. & Olla, S. Heat Flow in a Periodically Forced, Thermostatted Chain. Commun. Math. Phys. 400, 2181–2225 (2023). https://doi.org/10.1007/s00220-023-04654-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00220-023-04654-4

Navigation