Skip to main content
Log in

The Vlasov–Fokker–Planck equation with high dimensional parametric forcing term

  • Published:
Numerische Mathematik Aims and scope Submit manuscript

Abstract

We consider the Vlasov–Fokker–Planck equation with random electric field where the random field is parametrized by countably many infinite random variables due to uncertainty. At the theoretical level, with suitable assumption on the anisotropy of the randomness, adopting the technique employed in elliptic PDEs (Cohen and DeVore in Acta Numerica 24:1–159, 2015) , we prove the best N approximation in the random space enjoys a convergence rate, which depends on the summability of the coefficients of the random variable, higher than the Monte-Carlo method. For the numerical method, based on the adaptive sparse polynomial interpolation (ASPI) method introduced in Chkifa et al. (Found Comput Math 14:601–603, 2014), we develop a residual based adaptive sparse polynomial interpolation (RASPI) method which is more efficient for multi-scale linear kinetic equation, when using numerical schemes that are time dependent and implicit. Numerical experiments show that the numerical error of the RASPI decays faster than the Monte-Carlo method and is also dimension independent.

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

Similar content being viewed by others

References

  1. Binev, P., Cohen, A., Dahmen, W., DeVore, R., Petrova, G., Wojtaszczyk, P.: Convergence rates for greedy algorithms in reduced basis methods. SIAM J. Math. Anal. 43(3), 1457–1472 (2011)

    Article  MathSciNet  Google Scholar 

  2. Buffa, A., Maday, Y., Patera, A.T., Prud’homme, C., Turinici, G.: A priori convergence of the greedy algorithm for the parametrized reduced basis method. ESAIM: Math. Modell. Numer. Anal. 46(3), 595–603 (2012)

  3. Caflisch, R.E.: Monte carlo and quasi-monte carlo methods. Acta Numerica 7, 1–49 (1998)

    Article  MathSciNet  Google Scholar 

  4. Chkifa, A., Cohen, A., Schwab, C.: High-dimensional adaptive sparse polynomial interpolation and applications to parametric pdes. Found. Comput. Math. 14(4), 601–633 (2014)

    Article  MathSciNet  Google Scholar 

  5. Chkifa, A., Cohen, A., Schwab, C.: Breaking the curse of dimensionality in sparse polynomial approximation of parametric pdes. J. de Mathématiques Pures et Appliquées 103(2), 400–428 (2015)

  6. Chkifa, M.A.: On the lebesgue constant of leja sequences for the complex unit disk and of their real projection. J. Approx. Theory 166, 176–200 (2013)

    Article  MathSciNet  Google Scholar 

  7. Cohen, A., DeVore, R.: Approximation of high-dimensional parametric pdes. Acta Numerica 24, 1–159 (2015)

    Article  MathSciNet  Google Scholar 

  8. Cohen, Albert, DeVore, R., Schwab, C.: Convergence rates of best n-term galerkin approximations for a class of elliptic spdes. Found. Comput. Math. 10(6), 615–646 (2010)

    Article  MathSciNet  Google Scholar 

  9. Cohen, Albert, DeVore, R., Schwab, C.: Analytic regularity and polynomial approximation of parametric and stochastic elliptic pde’s. Anal. Appl. 9(01), 11–47 (2011)

  10. Daus E, Jin S, Liu L: On the multi-species boltzmann equation with uncertainty and its stochastic galerkin approximation. Preprint (2019)

  11. DeVore, R., Petrova, G., Wojtaszczyk, P.: Greedy algorithms for reduced bases in Banach spaces equation with uncertainty and its stochastic galerkin approximation. Constr. Approx. 37(3), 455–466 (2013)

    Article  MathSciNet  Google Scholar 

  12. Duan, R., Fornasier, M., Toscani, G.: A kinetic flocking model with diffusion. Commun. Math. Phys. 300(1), 95–145 (2010)

    Article  MathSciNet  Google Scholar 

  13. Filbet, F., Jin, S.: A class of asymptotic-preserving schemes for kinetic equations and related problems with stiff sources. J. Comput. Phys. 229(20), 7625–7648 (2010)

    Article  MathSciNet  Google Scholar 

  14. Hansen, Markus, Schwab, C.: Analytic regularity and nonlinear approximation of a class of parametric semilinear elliptic pdes. Mathematische Nachrichten 286(8–9), 832–860 (2013)

    Article  MathSciNet  Google Scholar 

  15. Hansen, Markus, Schwab, C.: Sparse adaptive approximation of high dimensional parametric initial value problems. Vietnam J. Math. 41(2), 181–215 (2013)

    Article  MathSciNet  Google Scholar 

  16. Hernández-Santamaría, Víctor, Lazar, Martin, Zuazua, Enrique: Greedy optimal control for elliptic problems and its application to turnpike problems. (2017)

  17. Hu J, Jin S, Shu R: On stochastic galerkin approximation of the nonlinear boltzmann equation with uncertainty in the fluid regime. J. Comp. Phys. 397, 108838 (2019)

  18. Hwang HJ, Jang J: On the vlasov-poisson-fokker-planck equation near maxwellian. Discrete Continuous Dyn. Syst. Ser. B, 18(3) (2013)

  19. Jin, S.: Efficient asymptotic-preserving (AP) schemes for some multiscale kinetic equations. SIAM J. Sci. Comput. 21, 441–454 (1999)

    Article  MathSciNet  Google Scholar 

  20. Jin S: Asymptotic preserving (ap) schemes for multiscale kinetic and hyperbolic equations: a review. Lecture Notes for Summer School on Methods and Models of Kinetic Theory(M&MKT), Porto Ercole (Grosseto, Italy). Rivista di Matematica della Universita di Parma, 3:177–216 (2012)

  21. Jin, S., Liu, J.G., Ma, Z.: Uniform spectral convergence of the stochastic galerkin method for the linear transport equations with random inputs in diffusive regime and a micro-macro decomposition based asymptotic preserving method. Res. Math. Sci. 4(1), 1–25 (2017). https://doi.org/10.1186/s40687-017-0105-1

    Article  MathSciNet  MATH  Google Scholar 

  22. Jin, S., Liu, L.: An asymptotic-preserving stochastic galerkin method for the semiconductor boltzmann equation with random inputs and diffusive scalings. SIAM Multiscale Model. Simult. 15, 157–183 (2017)

    Article  MathSciNet  Google Scholar 

  23. Jin, S., Wang, L.: An asymptotic preserving scheme for the Vlasov-Poisson-Fokker-Planck system in the high field regime. Acta Mathematica Scientia 31(6), 2219–2232 (2011)

    Article  MathSciNet  Google Scholar 

  24. Jin, S., Yan, B.: A class of asymptotic-preserving schemes for the Fokker-Planck-Landau equation. J. Comput. Phys. 230(17), 6420–6437 (2011)

    Article  MathSciNet  Google Scholar 

  25. Jin, S., Zhu, Y.: Hypocoercivity and uniform regularity for the Vlasov-Poisson-Fokker-Planck system with uncertainty and multiple scales. SIAM J. Math. Anal. 50, 1790–1816 (2018)

    Article  MathSciNet  Google Scholar 

  26. Kunoth, A., Schwab, C.: Analytic regularity and gpc approximation for control problems constrained by linear parametric elliptic and parabolic pdes. SIAM J. Control Opt. 51(3), 2442–2471 (2013)

    Article  MathSciNet  Google Scholar 

  27. Kuo, F.Y., Schwab, C., Sloan, I.H.: Quasi-monte carlo finite element methods for a class of elliptic partial differential equations with random coefficients. SIAM J. Numer. Anal. 50(6), 3351–3374 (2012)

    Article  MathSciNet  Google Scholar 

  28. Kuo, F.Y., Schwab, C., Sloan, I.H.: Multi-level quasi-monte carlo finite element methods for a class of elliptic pdes with random coefficients. Found. Comput. Math. 15(2), 411–449 (2015)

    Article  MathSciNet  Google Scholar 

  29. Lazar, M., Zuazua, E.: Greedy controllability of finite dimensional linear systems. Automatica 74, 327–340 (2016)

    Article  MathSciNet  Google Scholar 

  30. Li, Q., Wang, L.: Uniform regularity for linear kinetic equations with random input based on hypocoercivity. SIAM/ASA J. Uncertain. Quantif. 5(1), 1193–1219 (2017)

    Article  MathSciNet  Google Scholar 

  31. Liu, L., Jin, S.: Hypocoercivity based sensitivity analysis and spectral convergence of the stochastic galerkin approximation to collisional kinetic equations with multiple scales and random inputs. Multiscale Model. Simul. 16(3), 1085–1114 (2018)

    Article  MathSciNet  Google Scholar 

  32. Loeve, M.: Probability theory. I. Springer, New York-Heidelberg (1978)

    MATH  Google Scholar 

  33. Loeve, M.: Probability theory. II. Springer-Verlag, New York-Heidelberg (1978)

    Book  Google Scholar 

  34. Morokoff, W.J., Caflisch, R.E.: Quasi-monte carlo integration. J. Comput. Phys. 122(2), 218–230 (1995)

    Article  MathSciNet  Google Scholar 

  35. Nobile, F., Tempone, R., Webster, C.G.: An anisotropic sparse grid stochastic collocation method for partial differential equations with random input data. SIAM J. Numer. Anal. 46(5), 2411–2442 (2008)

    Article  MathSciNet  Google Scholar 

  36. Nobile, F., Tempone, R., Webster, C.G.: A sparse grid stochastic collocation method for partial differential equations with random input data. SIAM J. Numer. Anal. 46(5), 2309–2345 (2008)

    Article  MathSciNet  Google Scholar 

  37. Rozza, G., Huynh, D.B.P., Patera, A.T.: Reduced basis approximation and a posteriori error estimation for affinely parametrized elliptic coercive partial differential equations. Arch. Comput. Methods Eng. 15(3), 1 (2007)

    Article  Google Scholar 

  38. Schillings, C., Schwab, C.: Sparsity in bayesian inversion of parametric operator equations. Inverse Problems 30(6), 065007 (2014)

    Article  MathSciNet  Google Scholar 

  39. Schillings, C., Schwab, C.: Sparse, adaptive smolyak quadratures for bayesian inverse problems. Inverse Problems 29(6), 0655011 (2013)

    Article  MathSciNet  Google Scholar 

  40. Schwab, C., Stuart, A.M.: Sparse deterministic approximation of bayesian inverse problems. Inverse Problems 28(4), 045003 (2012)

    Article  MathSciNet  Google Scholar 

  41. Schwab, C., Todor, R., Radu, A.: Karhunen-loève approximation of random fields by generalized fast multipole methods. J. Comput. Phys. 217(1), 100–122 (2006)

    Article  MathSciNet  Google Scholar 

  42. Shu, R., Jin, S.: Uniform regularity in the random space and spectral accuracy of the stochastic galerkin method for a kinetic-fluid two-phase flow model with random initial inputs in the light particle regime. ESAIM: Math. Modell. Numer. Anal. 52(5), 1651–1678 (2018)

    Article  MathSciNet  Google Scholar 

  43. Todor, R.A.: Robust eigenvalue computation for smoothing operators. SIAM J. Numer Anal. 44(2), 865–878 (2006)

    Article  MathSciNet  Google Scholar 

  44. Zech, J., Dung, D., Schwab, Ch.: Multilevel approximation of parametric and stochastic pdes. Technical Report 2018-05, Seminar for Applied Mathematics, ETH Zürich, Switzerland (2018)

  45. Zhu Y: Sensitivity analysis and uniform regularity for the boltzmann equation with uncertainty and its stochastic galerkin approximation. Preprint

Download references

Acknowledgements

The first author was partially supported by NSF Grants DMS-1522184, DMS-1819012, DMS-1107291: RNMS KI- Net, and NSFC Grant No. 31571071. The work of the third author has been funded by the Alexander von Humboldt-Professorship program, the European Union’s Horizon 2020 research and innovation programme under the Marie Sklodowska-Curie grant agreement No.765579-ConFlex, Grant MTM2017-92996-C2-1-R COSNET of MINECO (Spain), ELKARTEK project KK-2018/00083 ROAD2DC of the Basque Government, ICON of the French ANR and Nonlocal PDEs: Analysis, Control and Beyond, AFOSR Grant FA9550-18-1-0242, and Transregio 154 Project *Mathematical Modelling, Simulation and Optimization using the Example of Gas Networks* of the German DFG.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shi Jin.

Additional information

Publisher's Note

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

Appendices

Appendices

1.1 The proof of lemma 3.7

1.1.1 The proof of the time independent random electric field

We will first prove the case when the random electric field is time independent, that is, \(E(t,x,z)\equiv E(x,z)\).

Lemma A.1

Under Condition 3.2, for all \( \mathbf{z }\in U\), the following estimates hold,

$$\begin{aligned}&|\nu | = 0: \partial _t G^0_i + \frac{\eta }{\epsilon }\left\Vert h\right\Vert _V^2 \le 0; \end{aligned}$$
(A.1)
$$\begin{aligned}&|\nu | > 1: \partial _tG_i^\nu + \frac{\eta }{\epsilon }\left\Vert \partial ^{\nu }h\right\Vert _V^2\le \frac{2}{\lambda \epsilon }\sum _{\nu _j\ne 0} \nu _j^2C_j \left\Vert \partial ^{\nu -e_j}h \right\Vert _V^2, \end{aligned}$$
(A.2)

where \(i = 1,2\), \(\eta = \frac{C_s}{10}\), and

$$\begin{aligned}&\frac{\epsilon }{2\lambda } \left\Vert \partial ^{\nu }h \right\Vert _V^2 \le G_1^\nu \le \frac{3}{2\lambda \epsilon } \left\Vert \partial ^{\nu }h \right\Vert _V^2,\quad \frac{1}{2\lambda \epsilon } \left\Vert \partial ^{\nu }h \right\Vert _V^2 \le G_2^\nu \le \frac{3}{2\lambda \epsilon } \left\Vert \partial ^{\nu }h \right\Vert _V^2 . \end{aligned}$$
(A.3)

Proof

For \(E(t,x,\mathbf{z }) \equiv E(x, \mathbf{z })\), the microscopic equation (2.13) for the perturbative solution \(h(t,x,v, \mathbf{z })\) is simplified to,

$$\begin{aligned} \epsilon \partial _th + v\partial _xh - \frac{1}{\epsilon } {\mathcal {L}}h = E\left( \partial _v - \frac{v}{2}\right) h, \end{aligned}$$
(0.4)

where \({\mathcal {L}}\) is the linearized Fokker Planck operator defined in (2.14) that satisfies the local coercivity property as in (2.15). If one multiplies \(\sqrt{M}\) and \(v\sqrt{M}\) to (0.4) respectively, and integrates them over v, then one gets the macroscopic equations,

figure a

The first equation is the perturbative continuity equation, while the second one is the perturbative momentum equation. Notice the operators \(\Pi \) and \(1-\Pi \) are perpendicular to each other in \(L^2_{x,v}\), that is,

$$\begin{aligned} \left\Vert h\right\Vert ^2 = \left\Vert \Pi h \right\Vert ^2 + \left\Vert (1-\Pi ) h\right\Vert ^2 = \left\Vert \sigma \right\Vert ^2 + \left\Vert (1-\Pi ) h\right\Vert ^2. \end{aligned}$$
(A.7)

If one takes \(\partial ^\nu \) and \(\partial ^\nu \partial _x\) to (0.4), and multiplies \(\partial ^\nu _\mathbf{z }h\) and \(\partial ^\nu \partial _xh\) respectively, then integrates them over xv and adds the two equations together, one has

$$\begin{aligned}&\frac{\epsilon }{2}\partial _t \left\Vert \partial ^{\nu }h \right\Vert _V^2 +\frac{\lambda }{\epsilon } \left\Vert (1-\Pi )\partial ^{\nu }h\right\Vert _{V,\omega }^2 \nonumber \\ \le&\underbrace{\left\langle \partial ^\nu \left( E\left( \partial _v-\frac{v}{2}\right) h \right) , \partial ^\nu h\right\rangle }_{I} +\underbrace{\left\langle \partial ^\nu \partial _x\left( E\left( \partial _v-\frac{v}{2}\right) h \right) , \partial ^\nu \partial _xh\right\rangle }_{II}, \end{aligned}$$
(A.8)

where the second term II comes from the hypocoercivity of \({\mathcal {L}}\) in (2.15). If one takes \(\partial ^\nu \) to (A.6) , and multiplies \(\partial ^\nu \partial _x\sigma \), then integrates it over xv, one has

(A.9)

Here the first term on the LHS and the first term on the RHS come from

$$\begin{aligned}&\left\langle \partial _t \partial ^{\nu }u, \partial ^{\nu }\partial _x\sigma \right\rangle = \partial _t\left\langle \partial ^{\nu }u, \partial ^{\nu }\partial _x\sigma \right\rangle - \left\langle \partial ^{\nu }u, \partial _t \partial ^{\nu }\partial _x\sigma \right\rangle \nonumber \\ =&\partial _t\left\langle \partial ^{\nu }u, \partial ^{\nu }\partial _x\sigma \right\rangle + \frac{1}{\epsilon }\left\langle \partial ^{\nu }u, \partial ^{\nu }\partial _x^2u \right\rangle = \partial _t\left\langle \partial ^{\nu }u, \partial ^{\nu }\partial _x\sigma \right\rangle - \frac{1}{\epsilon }\left\Vert \partial ^{\nu }\partial _xu\right\Vert ^2, \end{aligned}$$
(A.10)

where the second equality is because of the continuity equation (A.5). The second term on the LHS of (A.9) is because of

$$\begin{aligned}&\frac{1}{\epsilon }\left\langle \partial ^{\nu }u, \partial ^{\nu }\partial _x\sigma \right\rangle = \left\langle -\frac{1}{\epsilon } \partial ^{\nu }\partial _xu, \partial ^{\nu }\sigma \right\rangle = \left\langle \partial _t \partial ^{\nu }\sigma , \partial ^{\nu }\sigma \right\rangle = \frac{1}{2}\partial _t\left\Vert \partial ^{\nu }\sigma \right\Vert ^2. \end{aligned}$$
(A.11)

Furthermore since,

$$\begin{aligned}&\left\langle \partial ^{\nu }\partial _x\sigma , \partial ^{\nu }\partial _x\sigma \right\rangle + \left\langle \int v^2(1-\Pi )\partial _x \partial ^{\nu }h\sqrt{M}\,dv , \partial ^{\nu }\partial _x\sigma \right\rangle \nonumber \\&\quad \ge \frac{1}{2}\left\Vert \partial ^{\nu }\partial _x\sigma \right\Vert ^2-\frac{1}{2}\left\Vert (1-\Pi )\partial ^{\nu }\partial _xh \right\Vert ^2_\omega , \end{aligned}$$
(A.12)

this gives the third term on the LHS and the second term on the RHS of (A.9).

Furthermore, integrate (A.5) over x, by the periodic boundary condition, one has,

$$\begin{aligned} \partial _t \int \sigma dx = 0. \end{aligned}$$

Since from (2.12), we know that

$$\begin{aligned} \int \sigma (0,x,z) dx = \int h\sqrt{M}dxdv = 0, \end{aligned}$$

so

$$\begin{aligned} \int \sigma (t,x,z) dx = \int \sigma (0,x,z) dx = 0. \end{aligned}$$

Similar equality can be obtained for \(\partial ^\nu \sigma \). Therefore, one can apply the Poincare inequality (1.1) to \(\left\Vert \partial _x\sigma \right\Vert ^2\) to get,

$$\begin{aligned} \left\Vert \partial ^\nu \partial _x\sigma \right\Vert ^2 \ge C_s\left\Vert \partial ^\nu \sigma \right\Vert _V^2. \end{aligned}$$

By adding \(\theta _i\)(A.8) \(+ \frac{1}{2\epsilon }\) (A.9), and the fact that

$$\begin{aligned} \left\Vert \partial ^{\nu }\partial _xu\right\Vert ^2 \le \left\Vert (1-\Pi )\partial ^{\nu }\partial _xh \right\Vert ^2\le \left\Vert (1-\Pi )\partial ^{\nu }h \right\Vert _{V,\omega }^2, \end{aligned}$$

one has,

$$\begin{aligned}&\partial _t G_i^{\nu } +\frac{\lambda \theta _i}{\epsilon } \left\Vert (1-\Pi )\partial ^{\nu }h\right\Vert _{V,\omega }^2 + \frac{ C_s}{4\epsilon }\left\Vert \partial ^{\nu }\sigma \right\Vert _V^2 \nonumber \\&\quad \le \frac{3}{4\epsilon }\left\Vert (1-\Pi )\partial ^{\nu }h \right\Vert _{V,\omega }^2 + \theta _i\left( I+II\right) -\frac{1}{2\epsilon }III, \end{aligned}$$
(A.13)

where \(C_s\) comes from the Poincare inequality (1.1). In order to find an estimate for \(G^\nu _i\), one needs to estimate terms \(\theta _i\left( I+II\right) -\frac{1}{2\epsilon }III\). First notice that

$$\begin{aligned}&\left( \partial _v + \frac{v}{2}\right) \partial ^\nu \partial _x^i (\Pi h) = -\frac{v}{2}\left( \partial ^\nu \partial _x^i \sigma \right) \sqrt{M}+ \frac{v}{2}\left( \partial ^\nu \partial _x^i \sigma \right) \sqrt{M}= 0, \quad \text {for }i = 0, 1, \end{aligned}$$
(A.14)

which implies that terms I and II can be simplified to, for \(i = 0,1\),

$$\begin{aligned} I, II&= -\left\langle \partial ^\nu \partial _x^i(Eh), \left( \partial _v + \frac{v}{2}\right) \partial ^\nu \partial _x^ih\right\rangle \nonumber \\&= -\left\langle \partial ^\nu \partial _x^i(Eh), \left( \partial _v + \frac{v}{2}\right) \left( \partial ^\nu \partial _x^i( \Pi h) + (1-\Pi )\partial ^\nu \partial _x^ih\right) \right\rangle \nonumber \\&= -\left\langle \partial ^\nu \partial _x^i(Eh), \left( \partial _v + \frac{v}{2}\right) (1-\Pi )\partial ^\nu \partial _x^i h\right\rangle . \end{aligned}$$
(A.15)

Another inequality that will be used frequently later is that for \(i = 0,1\),

$$\begin{aligned}&\left\Vert \left( \partial _v + \frac{v}{2}\right) \left( 1-\Pi \right) \partial ^\nu \partial _x^i h\right\Vert ^2 \nonumber \\&\quad = \left\Vert \partial _v(1-\Pi )\partial ^\nu \partial _x^i h\right\Vert ^2 + \frac{1}{4}\left\Vert v (1-\Pi )\partial ^\nu \partial _x^i h \right\Vert ^2 + \int _\Omega \frac{v}{2}\partial _v\left( (1-\Pi )\partial ^\nu \partial _x^i h\right) ^2dxdv \nonumber \\&\quad = \left\Vert \partial _v(1-\Pi )\partial ^\nu \partial _x^i h\right\Vert ^2 + \frac{1}{4}\left\Vert v (1-\Pi )\partial ^\nu \partial _x^i h \right\Vert ^2 \nonumber \\&\qquad - \frac{1}{2} \left\Vert (1-\Pi )\partial ^\nu \partial _x^i h \right\Vert ^2 \le \left\Vert (1-\Pi )\partial ^\nu \partial _x^i h \right\Vert ^2_\omega . \end{aligned}$$
(A.16)

Based on (A.15) and (A.16), we will bound the term \(\theta _i\left( I+II\right) -\frac{1}{2\epsilon }III\) for the cases \(|\nu | = 0, |\nu |>1\) respectively. Firstly, for the case \(|\nu | = 0\),

$$\begin{aligned}&\theta _i\left( I+II\right) -\frac{1}{2\epsilon }III \nonumber \\&\quad = \theta _i\left\langle Eh, \left( \partial _v + \frac{v}{2}\right) (1-\Pi ) h\right\rangle + \theta _i\left\langle \partial _x\left( Eh\right) , \left( \partial _v + \frac{v}{2}\right) (1-\Pi ) h\right\rangle \nonumber \\&\qquad - \frac{1}{2\epsilon }\left\langle E\sigma , \partial _x\sigma \right\rangle \nonumber \\&\quad \le \frac{\theta _i}{2}\left\Vert E \right\Vert _{L^\infty _x}\left( \epsilon \left\Vert h \right\Vert _V^2 + \frac{1}{\epsilon }\left\Vert (1-\Pi ) h\right\Vert _{V,\omega }^2\right) \nonumber \\&\quad + \frac{\theta _i}{2}\left\Vert \partial _xE \right\Vert _{L^\infty _x}\left( \epsilon \left\Vert h \right\Vert ^2 + \frac{1}{\epsilon }\left\Vert (1-\Pi ) \partial _xh\right\Vert _\omega ^2\right) \nonumber \\&+ \frac{1}{4\epsilon } \left\Vert E \right\Vert _{L^\infty _x}\left( \left\Vert \sigma \right\Vert ^2 + \left\Vert \partial _x\sigma \right\Vert ^2\right) . \end{aligned}$$
(A.17)

Since \(\left\Vert E \right\Vert _{L^\infty (U,W^\infty _x)} \le C_E\), and \(\left\Vert h \right\Vert ^2_V = \left\Vert \sigma \right\Vert _V^2 + \left\Vert (1-\Pi )h \right\Vert _V^2\), one can further simplify (A.17) to

$$\begin{aligned}&(A.17) \le C_E\left( \left( \frac{\epsilon \theta _i}{2} + \frac{1}{4\epsilon }\right) \left\Vert \sigma \right\Vert _V^2 + \frac{\theta _i}{\epsilon }\left\Vert (1-\Pi ) h\right\Vert _{V,\omega }^2\right) . \end{aligned}$$
(A.18)

Plug the above estimate into (A.13), one has,

$$\begin{aligned}&\partial _t G_i^0 +\frac{\lambda \theta _i}{\epsilon } \left\Vert (1-\Pi )h\right\Vert _{V,\omega }^2 + \frac{ C_s}{4\epsilon }\left\Vert \sigma \right\Vert _V^2 \nonumber \\ \le&\frac{3}{4\epsilon }\left\Vert (1-\Pi )h \right\Vert _{V,\omega }^2 + C_E\left( \left( \frac{\epsilon \theta _i}{2} + \frac{1}{4\epsilon }\right) \left\Vert \sigma \right\Vert _V^2 + \frac{\theta _i}{\epsilon }\left\Vert (1-\Pi ) h\right\Vert _{V,\omega }^2\right) , \end{aligned}$$
(A.19)

which implies,

$$\begin{aligned}&\partial _t G_i^{0} + C^0_h \left\Vert (1-\Pi )h\right\Vert _{V,\omega }^2 + C^0_\sigma \left\Vert \sigma \right\Vert _V^2 \le 0, \end{aligned}$$
(A.20)

where

$$\begin{aligned}&C^0_h = \frac{\lambda \theta _i}{\epsilon } -\frac{3}{4\epsilon } - \frac{C_E\theta _i}{\epsilon },\quad C^0_\sigma = \frac{ C_s}{4\epsilon } - C_E\left( \frac{\epsilon \theta _i}{2} + \frac{1}{4\epsilon }\right) . \end{aligned}$$
(A.21)

Since \(C_E\le \frac{\lambda C_s}{8} \le \frac{\lambda }{16}\), so \(\lambda - C_E \ge \frac{15}{16}\lambda \), which implies \(\left( \lambda - C_E\right) \frac{\theta _i}{\epsilon } \ge \frac{4}{\lambda }\) for both \(i = 1,2\). Therefore \(C_h^0 \ge \frac{3}{\epsilon }\). Since \(C_s - C_E \ge \frac{7}{8}C_s\), and \(\epsilon C_E\theta _i\le \frac{\epsilon C_s}{7}\) for both \(i = 1,2\), therefore \(C_\sigma ^0 \ge \frac{7C_s}{32\epsilon } - \frac{\epsilon C_s}{14} \ge \frac{C_s}{10\epsilon }\). Hence plug back \(C^0_h, C^0_\sigma \ge \frac{C_s}{10\epsilon }\) to (A.20), one has,

(A.22)

which is exactly (A.1) in Lemma A.1.

For the case where \(|\nu | > 0 \), based on (A.15), (A.16) and the definition of \(E = {\bar{E}}(x) + \sum _{j\ge 1}E_j(x)z_j\), one can bound \(\theta _i\left( I+II\right) -\frac{1}{2\epsilon }III\) by

$$\begin{aligned}&\theta _i\left( I+II\right) -\frac{1}{2\epsilon }III\nonumber \\&\quad \le \theta _i\left\langle \partial _x\left( E\partial ^{\nu }h\right) , \left( \partial _v+\frac{v}{2}\right) (1-\Pi )\partial ^{\nu }\partial _xh\right\rangle - \frac{1}{2\epsilon } \left\langle E\partial ^{\nu }\sigma , \partial ^{\nu }\partial _x\sigma \right\rangle \nonumber \\&\qquad +\sum _{\nu _j\ne 0}\left( \theta _i\left\langle \nu _jE_j\partial ^{\nu -e_j}h, (\partial _v+\frac{v}{2})(1-\Pi )\partial ^{\nu }h\right\rangle \right. \nonumber \\&\qquad \left. + \theta _i\left\langle \partial _x\left( \nu _jE_j\partial ^{\nu -e_j}h\right) , (\partial _v+\frac{v}{2})(1-\Pi )\partial ^{\nu }\partial _xh\right\rangle \right. \nonumber \\&\qquad \left. -\frac{1}{2\epsilon } \left\langle \nu _jE_j\partial ^{\nu -e_j}\sigma , \partial ^{\nu }\partial _x\sigma \right\rangle \right) . \end{aligned}$$
(A.23)

Since the three terms in the second line of the above equation is similar to the case where \(|\nu | = 0\), hence one can get similar estimates as in (A.18). In addition, by assumption \(\left\Vert E_j \right\Vert _{L^\infty (U,W^{1,\infty }_x)} \le C_j\), one has

$$\begin{aligned}&(A.23)\nonumber \\&\quad \le C_E\left( \left( \frac{\epsilon \theta _i}{2}+\frac{1}{4\epsilon }\right) \left\Vert \partial ^{\nu }\sigma \right\Vert _V^2+ \frac{\theta _i}{\epsilon }\left\Vert (1-\Pi )\partial ^{\nu }h \right\Vert _{V,\omega } ^2\right) \nonumber \\&\qquad +\frac{\theta _i}{2}\sum _{\nu _j\ne 0} C_j\left( \epsilon \nu _j^2\left\Vert \partial ^{\nu -e_j}h \right\Vert _V^2 + \frac{1}{\epsilon }\left\Vert (1-\Pi )\partial ^{\nu }h \right\Vert _{V,\omega }^2\right) \nonumber \\&\qquad +\frac{1}{4\epsilon }\sum _{\nu _j\ne 0}C_j\left( \nu _j^2\left\Vert \partial ^{\nu -e_j}\sigma \right\Vert ^2 + \left\Vert \partial ^{\nu }\partial _x\sigma \right\Vert ^2 \right) \nonumber \\&\quad \le C_E\left( \frac{1}{2}\left( \epsilon \theta _i+\frac{1}{\epsilon }\right) \left\Vert \partial ^{\nu }\sigma \right\Vert _V^2+ \frac{3\theta _i}{2\epsilon }\left\Vert (1-\Pi )\partial ^{\nu }h \right\Vert _{V,\omega } ^2\right) \nonumber \\&\qquad +\sum _{\nu _j\ne 0} \nu _j^2C_j\left( \frac{\epsilon \theta _i}{2} + \frac{1}{4\epsilon }\right) \left\Vert \partial ^{\nu -e_j}h \right\Vert _V^2 , \end{aligned}$$
(A.24)

where the second inequality is because that \(\sum _{j\ge 1}C_j \le C_E\). Hence plug (A.24) into (A.13), one has,

$$\begin{aligned}&\partial _t G^\nu _i + C^\nu _h \left\Vert (1-\Pi )\partial ^\nu h\right\Vert _{V,\omega }^2 + C^\nu _\sigma \left\Vert \partial ^\nu \sigma \right\Vert _V^2 \le \sum _{\nu _j\ne 0} \nu _j^2C_j\left( \frac{\epsilon \theta _i}{2} + \frac{1}{4\epsilon }\right) \left\Vert \partial ^{\nu -e_j}h \right\Vert _V^2, \end{aligned}$$
(A.25)

where

$$\begin{aligned}&C^\nu _h = \frac{\lambda \theta _i}{\epsilon } -\frac{3}{4\epsilon } - \frac{3C_E\theta _i}{2\epsilon } \ge \frac{\eta }{\epsilon },\quad C^\nu _\sigma = \frac{ C_s}{4\epsilon } - \frac{C_E}{2}\left( \epsilon \theta _i+ \frac{1}{\epsilon }\right) \ge \frac{\eta }{\epsilon } \end{aligned}$$
(A.26)

and

$$\begin{aligned}&\frac{\epsilon \theta _i}{2} + \frac{1}{4\epsilon } \le \frac{2}{\lambda \epsilon }, \end{aligned}$$
(A.27)

for both \(i = 1,2\), which gives (A.2) in Lemma A.1. \(\square \)

1.1.2 The proof of lemma 3.7

If one takes \(\partial ^\nu \) and \(\partial ^\nu \partial _x\) to (2.13), and multiplies \(\partial ^\nu _\mathbf{z }h\) and \(\partial ^\nu \partial _xh\) respectively, then integrates them over xv and adds the two equations together, one has

$$\begin{aligned}&\frac{\epsilon }{2}\partial _t \left\Vert \partial ^{\nu }h \right\Vert _V^2 +\frac{\lambda }{\epsilon } \left\Vert (1-\Pi )\partial ^{\nu }h\right\Vert _{V,\omega }^2 \nonumber \\ \le&\left\langle \partial ^\nu \left( E\left( \partial _v-\frac{v}{2}\right) h \right) , \partial ^\nu h\right\rangle +\left\langle \partial ^\nu \partial _x\left( E\left( \partial _v-\frac{v}{2}\right) h \right) , \partial ^\nu \partial _xh\right\rangle , \nonumber \\&-\underbrace{\left\langle v\sqrt{M}\partial ^\nu \left( \left( E - E^\infty \right) e^{-\phi ^\infty }\right) , \partial ^\nu h\right\rangle }_{IV} -\underbrace{\left\langle v\sqrt{M}\partial ^\nu \partial _x\left( \left( E - E^\infty \right) e^{-\phi ^\infty } \right) , \partial ^\nu \partial _xh\right\rangle }_{V}. \end{aligned}$$
(A.28)

If one multiplies \(\sqrt{M}\) and \(v\sqrt{M}\) to (2.13), and integrates it over v, then one gets

figure b

Then if one takes \(\partial ^\nu \) to (A.30), and multiplies \(\partial ^\nu \partial _x\sigma \), then integrates it over xv, one has

(A.31)

By comparing (A.28) and (A.31) with (A.8) and (A.9), we actually only need to bound terms IVVVI. Similar to (A.13), one has the following estimates for \(G_i^\nu \),

(A.32)

First notice, for \(i = 0, 1\),

(A.33)

So one can break

in \((IV+ V)\), therefore

(A.34)

For the term VI, one can bound it by

(A.35)

For \(|\nu | = 0\), plug (A.34) and (A.35) into (A.32), and based on the estimates (A.19) we have already got in Appendices 0.1, one has

(A.36)

which implies,

(A.37)

where

(A.38)

Since \(C_E\le \frac{\lambda C_s}{8} \le \frac{\lambda }{16}\), so \(\lambda - 2C_E \ge \frac{7}{8}\lambda \), which implies \(\left( \lambda - C_E\right) \frac{\theta _i}{\epsilon } \ge 1\) for both \(i = 1,2\). Therefore \(C_h^0 \ge \frac{1}{4\epsilon }\). Since \(\frac{C_s}{4} - \frac{3C_E}{8} \ge \frac{11}{64}C_s\), and \(\epsilon C_E\theta _i\le \frac{\epsilon C_s}{7}\) for both \(i = 1,2\), therefore \(C_\sigma ^0 \ge \frac{11C_s}{64\epsilon } - \frac{\epsilon C_s}{14} \ge \frac{C_s}{10\epsilon }\). Hence plug back \(C^0_h, C^0_\sigma \ge \frac{C_s}{10\epsilon }\) to (A.37), one has,

(A.39)

For \(|\nu | > 0\), plug (A.34) and (A.35) into (A.32), and based on the estimates (A.24), one has

(A.40)

which implies

(A.41)

for the same \(C^\nu _h, C^\nu _h\) defined in (A.38). This completes the proof.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Jin, S., Zhu, Y. & Zuazua, E. The Vlasov–Fokker–Planck equation with high dimensional parametric forcing term. Numer. Math. 150, 479–519 (2022). https://doi.org/10.1007/s00211-021-01257-w

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00211-021-01257-w

Mathematics Subject Classification

Navigation