Skip to main content
Log in

Entropy Dissipation Semi-Discretization Schemes for Fokker–Planck Equations

  • Published:
Journal of Dynamics and Differential Equations Aims and scope Submit manuscript

Abstract

We propose a new semi-discretization scheme to approximate nonlinear Fokker–Planck equations, by exploiting the gradient flow structures with respect to the 2-Wasserstein metric in the space of probability densities. We discretize the underlying state by a finite graph and define a discrete 2-Wasserstein metric in the discrete probability space. Based on such metric, we introduce a gradient flow of the discrete free energy as semi discretization scheme. We prove that the scheme maintains dissipativity of the free energy and converges to a discrete Gibbs measure at exponential dissipation rate. We exhibit these properties on several numerical examples.

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

Similar content being viewed by others

Notes

  1. Here the eigenvector of A corresponding to the smallest eigenvalue satisfies the assumption (19).

References

  1. Ambrosio, L.: Lecture Notes on Optimal Transport Problems. Springer, Berlin (2003)

    Book  MATH  Google Scholar 

  2. Ambrosio, L., Gigli, N., Savaré, G.: Gradient Flows: In Metric Spaces and in the Space of Probability Measures. Springer Science and Business Media, Berlin (2006)

    MATH  Google Scholar 

  3. Benamou, J.D., Brenier, Y.: A computational fluid mechanics solution to the Monge-Kantorovich mass transfer problem. Numer. Math. 84(3), 375–393 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  4. Benedetto, D., Caglioti, E., Carrillo, J., Pulvirenti, M.: A non-Maxwellian steady distribution for one-dimensional granular media. J. Stat. Phys. 91(5–6), 979–990 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  5. Buet, C., Cordier, S.: Numerical analysis of conservative and entropy schemes for the Fokker–Planck–Landau equation. SIAM J. Numer. Anal. 36, 953–973 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  6. Chang, J., Cooper, G.: A practical difference scheme for Fokker–Planck equations. J. Comput. Phys. 6(1), 1–16 (1970)

    Article  MATH  Google Scholar 

  7. Che, R., Huang, W., Li, Y., Tetali, P.: Convergence to global equilibrium for Fokker–Planck equations on a graph and Talagrand-type inequalities. J. Differ. Equ. 261, 2552–2583 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  8. Carrillo, J., Chertock, A., Huang, Y.: A finite-volume method for nonlinear nonlocal equations with a gradient flow structure. Commun. Comput. Phys. 17(01), 233–258 (2015)

    Article  MathSciNet  MATH  Google Scholar 

  9. Carrillo, J.A., Huang, Y., Patacchini, F.S., Wolansky, G.: Numerical Study of a Particle Method for Gradient Flows. arXiv:1512.03029, (2015)

  10. Carrillo, J., McCann, R., Villani, C.: Kinetic equilibration rates for granular media and related equations: entropy dissipation and mass transportation estimates. Revista Matematica Iberoamericana 19(3), 971–1018 (2003)

    Article  MathSciNet  MATH  Google Scholar 

  11. Carrillo, J.A., JuEngel, A., Markowich, P.A., Toscani, G., Unterreiter, A.: Entropy dissipation methods for degenerate parabolic problems and generalized sobolev inequalities. Mon. Hefte. Math. 133(1), 1–82 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  12. Chow, S.N., Huang, W., Li, Y., Zhou, H.: Fokker–Planck equations for a free energy functional or Markov process on a graph. Arch. Ration. Mech. Anal. 203(3), 969–1008 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  13. Chow, S.N., Li, W., Zhou, H.: Entropy dissipation of Fokker–Planck equations on graphs arXiv:1701.04841, (2017)

  14. Dieci, L., Li, W., Zhou, H.: A new model for realistic random perturbations of stochastic oscillators. J. Differ. Equ. 261(4), 2502–2527 (2016)

    Article  MathSciNet  MATH  Google Scholar 

  15. Erbar, M., Maas, J.: Ricci curvature of finite Markov chains via convexity of the entropy. Arch. Ration. Mech. Anal. 206(3), 997–1038 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  16. Jordan, R., Kinderlehrer, D., Otto, F.: The variational formulation of the Fokker–Planck equation. SIAM J. Math. Anal. 29(1), 1–17 (1998)

    Article  MathSciNet  MATH  Google Scholar 

  17. Li, W.: A study of stochastic differential equations and Fokker–Planck equations with applications. PhD thesis, Georgia tech, (2016)

  18. Maas, J.: Gradient flows of the entropy for finite Markov chains. J. Funct. Anal. 261(8), 2250–2292 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  19. Markowich, P.A., Villani, C.: On the trend to equilibrium for the Fokker–Planck equation: an interplay between physics and functional analysis. Mat. Contemp 19, 1–29 (2000)

    MathSciNet  MATH  Google Scholar 

  20. Mielke, A.: A gradient structure for reaction-diffusion systems and for energy-drift-diffusion. Nonlinearity 24(4), 13–29 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  21. Otto, F.: The geometry of dissipative evolution equations: the porous medium equation. Commun. Partial Differ. Equ. 26(1–2), 101–174 (2001)

    Article  MathSciNet  MATH  Google Scholar 

  22. Otto, F., Villani, C.: Generalization of an inequality by Talagrand and links with the logarithmic Sobolev inequality. J. Funct. Anal. 173(2), 361–400 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  23. Shu, C.-W., Osher, S.: Efficient implementation of essentially non-oscillatory shock-capturing schemes. J. Comput. Phys. 77, 439–471 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  24. Villani, C.: A review of mathematical topics in collisional kinetic theory. In: Handbook of Mathematical Fluid Dynamics, vol. 1, pp. 71–305 (2002)

  25. Villani, C.: Topics in optimal transportation. Number 58. American Mathematical Soc., (2003)

  26. Villani, C.: Optimal Transport: Old and New, vol. 338. Springer Science and Business Media, Berlin (2008)

    MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Wuchen Li.

Additional information

This work is partially supported by NSF Awards DMS–1042998, DMS–1419027, and ONR Award N000141310408.

Appendix

Appendix

Generally, to obtain \(\lambda _{\mathcal {F}}(\rho )\) in Definition 6 is not easy. Below, we give simple 1-d model example to illustrate situations in which \(\lambda _{\mathcal {F}}(\rho )\) can be explicitly obtained, and its dependence on the graph structure (the boundary conditions of the PDE).

A 1-d model problem Suppose that the free energy contains only the linear entropy term, so that the gradient flow is the heat equation:

$$\begin{aligned} \frac{\partial \rho }{\partial t}=\Delta \rho , \quad x\in (a,b). \end{aligned}$$
(16)

Here, we consider either (i) Neumann boundary conditions (zero flux) \(\frac{\partial \rho }{\partial x}|_{x=a}=\frac{\partial \rho }{\partial x}|_{x=b}=0\), or (ii) periodic boundary conditions \(\rho (t,a)=\rho (t,b)\).

We approximate the solution of (16) by (3), with a uniform discretization \(\Delta x=\frac{b-a}{n-1}\):

$$\begin{aligned} \frac{d\rho _i}{dt}=\frac{1}{\Delta x^2}\left\{ \sum _{j\in N(i)}\rho _j(\log \rho _j-\log \rho _i)_+ -\sum _{j\in N(i)}\rho _i(\log \rho _i-\log \rho _j)_+\right\} . \end{aligned}$$
(17)

The above two types of boundary conditions lead to distinct graph structures.

(i) A lattice graph \(L_{n}\):

figure a

(ii) A cycle graph \(C_n\):

figure b

In both cases, (17) is the gradient flow of the discrete linear entropy

$$\begin{aligned} \mathcal {H}(\rho )=\sum _{i=1}^n\rho _i\log \rho _i, \end{aligned}$$

and the unique Gibbs measure is \(\rho ^{\infty }=(\frac{1}{n}, \ldots , \frac{1}{n})\). We are going to estimate how fast the solution \(\rho (t)\) of the semi-discretization scheme (17) converges to the equilibrium \(\rho ^{\infty }\).

As we have seen in Theorem 7, the asymptotic convergence rates are determined by \(\lambda _{\mathcal {F}}(\rho )\):

$$\begin{aligned} \begin{aligned}&\lambda _{\mathcal {H}}(\rho ^{\infty })=\min _{\Phi \in \mathbb {R}^n} \left\{ \frac{1}{\Delta x^4}\sum _{(i,j)\in E}\sum _{(k,l)\in E}h_{ij, kl}(\Phi _i-\Phi _j)_+(\Phi _k-\Phi _l)_+~:\right. \\&\quad \left. \sum _{(i,j)\in E}\left( \frac{\Phi _i-\Phi _j}{\Delta x}\right) ^2_+\rho _i=1\right\} , \end{aligned} \end{aligned}$$
(18)

where

$$\begin{aligned} h_{ij, kl}=f_{ik}+f_{jl}-f_{il}-f_{jk}, \quad \text {and} \quad f_{ij}(\rho ^\infty )=\frac{\partial ^2}{\partial \rho _i\partial \rho _j} \mathcal {H}(\rho )|_{\rho =\rho ^{\infty }}={\left\{ \begin{array}{ll} \frac{1}{\rho _i^{\infty }} \quad &{}\text {if}~ i=j;\\ 0\quad &{}\text {if}~i\ne j.\quad \end{array}\right. } \end{aligned}$$

For the present model, we can find exact values of (18) for the above two graphs.

Theorem 13

We have

$$\begin{aligned} \begin{aligned} \lambda _{\mathcal {H}}(\rho ^{\infty })=\frac{\pi ^2}{(b-a)^2}+o(1), \quad (L_n) \end{aligned} \end{aligned}$$

and

$$\begin{aligned} \begin{aligned} \lambda _{\mathcal {H}}(\rho ^{\infty })=\frac{4\pi ^2}{(b-a)^2}+o(1).\quad (C_n) \end{aligned} \end{aligned}$$

Proof

First, consider the lattice graph \(L_n\). Without loss of generality, let \((\Phi _i)_{i=1}^n\) in (18) satisfy the relation

$$\begin{aligned} \Phi _1\ge \Phi _2 \ge \cdots \ge \Phi _n. \end{aligned}$$
(19)

Denote \(\xi :=(\xi _i)_{i=1}^{n-1}\in \mathbb {R}^{n-1}_+\) by

$$\begin{aligned} \xi _i:=\frac{\Phi _{i+1}-\Phi _{i}}{\sqrt{n}\Delta x}, \quad 1\le i\le n, \end{aligned}$$
(20)

and substitute \(\rho ^{\infty }\) into (18), to obtain

$$\begin{aligned} \lambda _{\mathcal {H}}(\rho ^{\infty })= \min _{\xi \in \mathbb {R}^{n-1}_+}\left\{ \frac{1}{\Delta x^2}\xi ^TA\xi \quad :\quad \xi ^T\xi =1\right\} , \end{aligned}$$

where

$$\begin{aligned} A=\begin{pmatrix} 2 &{} -\,1 &{} &{} \\ -\,1 &{} 2 &{} -\,1 &{} \\ \\ &{}\ddots &{} \ddots &{}\ddots &{}\\ \\ &{} &{} -\,1&{} 2 &{} -\,1\\ &{} &{} &{}-\,1 &{} 2 \end{pmatrix}\in \mathbb {R}^{(n-1)\times (n-1)}. \end{aligned}$$

It is simple to observe that A is positive definite and thatFootnote 1

$$\begin{aligned} \begin{aligned} \lambda _{\mathcal {H}}(\rho ^{\infty })=&\frac{1}{\Delta x^2}\times (\text {the smallest eigenvalue of}~ A)=\frac{1}{\frac{(b-a)^2}{(n-1)^2}}\left[ 2-2\cos \left( \frac{\pi }{n-1}\right) \right] \\ =&\frac{\pi ^2}{(b-a)^2}+o(1). \end{aligned} \end{aligned}$$

Next, we analyze the convergence rate for the cycle graph \(C_n\). Again we assume the relation (19) and let \(\xi \) as in (20). Since \(C_n\) has one more edge than \(L_n\), we let \(\eta \in \mathbb {R}\):

$$\begin{aligned} \eta :=\frac{\Phi _{1}-\Phi _{n}}{\sqrt{n}\Delta x}=\sum _{i=1}^{n-1} \xi _i. \end{aligned}$$

Substituting \(\rho ^{\infty }\) into (18), we have

$$\begin{aligned} \begin{aligned} \lambda _{\mathcal {H}}(\rho ^{\infty })=&\min _{(\xi , \eta )\in \mathbb {R}^{n}_+} \left\{ \frac{1}{\Delta x^2} [\xi ^TA\xi +2\xi _1\eta +2\xi _{n-1}\eta +2\eta ^2]\, :\right. \\&\quad \left. \xi ^T\xi +\eta ^2=1, ~\eta =\sum _{i=1}^{n-1}\xi _i\right\} . \end{aligned} \end{aligned}$$
(21)

The following transformations reduce (21) to a simpler eigenvalue problem. Let

$$\begin{aligned} \begin{pmatrix}\xi \\ \eta \end{pmatrix}=P\xi ,\quad \text {where}\quad P=\begin{pmatrix} I \\ \mathbf 1 \end{pmatrix}\in \mathbb {R}^{n\times (n-1)} \end{aligned}$$

with the identity matrix \(I\in \mathbb {R}^{(n-1)\times (n-1)}\) and \(\mathbf 1 \in \mathbb {R}^{n-1}\) being the vector of all 1’s. Then, (21) becomes

$$\begin{aligned} \lambda _{\mathcal {H}}(\rho ^{\infty })= \min _{\xi \in \mathbb {R}^{n-1}_+}\left\{ \frac{1}{\Delta x^2}(P\xi )^T B (P\xi ){:}(P\xi )^T(P\xi )=1\right\} , \end{aligned}$$
(22)

where

$$\begin{aligned} B=\begin{pmatrix} A &{} b^{\text {T}} \\ b &{} 2 \end{pmatrix}\in \mathbb {R}^{n\times n} \quad \text {with}\quad b^T\in \mathbb {R}^{n-1},\ b=(1,0,\ldots , 0, 1), \end{aligned}$$

and A is as above.

Below, we compute (22). First, we give explicit formulas for the eigenvalues and eigenvectors of B. \(\square \)

Lemma 14

Let \(n\ge 3\). For each \(k=0,1,\ldots , n-1\), the eigenvalues of B are

$$\begin{aligned} \lambda _k=2- 2\cos \left( \frac{2k\pi }{n}\right) . \end{aligned}$$

For \(k=0, 1,\ldots , n-1\), the associated eigenvectors in un-normalized form are:

$$\begin{aligned} v_k=(v_k(j))_{j=1}^{n}, \quad w_k=(w_k(j))_{j=1}^{n}, \end{aligned}$$

where, for \(j=1,\ldots , n-1\),

$$\begin{aligned} v_k(j)=\sin \left( \frac{2\pi k j}{n}\right) ,\quad w_k(j)=\cos \left( \frac{2\pi k j}{n}\right) ; \end{aligned}$$

and when \(j=n\),

$$\begin{aligned} v_k(n)=-\sin \left( \frac{2\pi k j}{n}\right) ,\quad w_k(j)=-\cos \left( \frac{2\pi k j}{n}\right) . \end{aligned}$$

Proof

The proof is by direct computation. We just show the details for the case of \(j=1\). We have

$$\begin{aligned} \begin{aligned} (Bv_k)(1)&=2v_k(1)-v_{k}(2)+v_{k}(n)\\&=2\sin \left( \frac{2\pi k }{n}\right) -\sin \left( \frac{2\cdot 2\pi k}{n}\right) -0\quad \text {By double angle formula}\\&=\left( 1-2\cos \frac{2k\pi }{n}\right) v_k(1). \end{aligned} \end{aligned}$$

And

$$\begin{aligned} \begin{aligned} (Bw_k)(1)&= 2w_k(1)-w_{k}(2)+w_{k}(n)\\&=2\cos \left( \frac{2\pi k }{n}\right) -\cos \left( \frac{2\cdot 2\pi k}{n}\right) +1\quad \text {By double angle formula}\\&=\left( 1-2\cos \frac{2k\pi }{n}\right) w_k(1). \end{aligned} \end{aligned}$$

\(\square \)

Note that in Lemma 14, many eigenvalues are repeated. As a consequence, obviously there are only two eigenvectors associated to each repeated eigenvalues, and not four; the repeating eigenvalues, in fact, have identical pairs \(v_k\), \(w_k\), up to sign. However, the eigenvalue equal to 0 is simple, with associated eigenvector \(w_0=(1,\ldots , 1, -\,1)^T\). Moreover, aside from this 0 eigenvalue, all other eigenvalues are positive.

Now, observe that \(P^Tw_0=0\), and therefore the matrix \(V=[w_0,P]\) is invertible and

$$\begin{aligned} BV=V\begin{bmatrix} 0&0 \\ 0&C\end{bmatrix}, \end{aligned}$$

where \(C\in \mathbb {R}^{n-1,n-1}\). Further, notice that \(P^TP\) is positive definite and thus it has a unique positive definite square root \((P^TP)^{1/2}\). Thus, \(\xi ^TP^TBP\xi \), subject to \((P\xi )^TP\xi =1\), can be rewritten as

$$\begin{aligned} \xi ^TP^TBP\xi =\xi ^TP^TPC\xi =\xi ^T(P^TP)^{1/2}(P^TP)^{1/2}C(P^TP)^{-1/2}(P^TP)^{1/2}\xi \end{aligned}$$

and thus, with \(x=(P^TP)^{1/2}\xi \), we end up with the problem

$$\begin{aligned} \min _{x: \ x^Tx=1} x^T\left[ (P^TP)^{1/2}C(P^TP)^{-1/2}\right] x\,. \end{aligned}$$

Finally, we notice that the matrix \(\left[ (P^TP)^{1/2}C(P^TP)^{-1/2}\right] \) is symmetric, and it is obviously similar to C, so that indeed

$$\begin{aligned} \begin{aligned}&\min _{x: \ x^Tx=1} x^T\left[ (P^TP)^{1/2}C(P^TP)^{-1/2}\right] x \\&\quad =\min _{\xi \in \mathbb {R}^{n-1}_+}\left\{ (P\xi )^T B (P\xi )~:~ (P\xi )^T(P\xi )=1\right\} =\text {The second smallest eigenvalue of}~ B. \end{aligned} \end{aligned}$$
(23)

Putting it all together, (22) gives

$$\begin{aligned} \begin{aligned} \lambda _{\mathcal {H}}(\rho ^{\infty })=&\frac{1}{\Delta x^2}(\text {the second smallest eigenvalue of}~ B)\\ =&\frac{1}{\frac{(b-a)^2}{(n-1)^2}}\left[ 2-2\cos \left( \frac{2\pi }{n}\right) \right] =\frac{4\pi ^2}{(b-a)^2}+o(1), \end{aligned} \end{aligned}$$

and the proof of Theorem 13 is completed. \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Chow, SN., Dieci, L., Li, W. et al. Entropy Dissipation Semi-Discretization Schemes for Fokker–Planck Equations. J Dyn Diff Equat 31, 765–792 (2019). https://doi.org/10.1007/s10884-018-9659-x

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10884-018-9659-x

Keywords

Mathematics Subject Classification

Navigation