Entropy-dissipating finite-difference schemes for nonlinear fourth-order parabolic equations
Abstract.
Structure-preserving finite-difference schemes for general nonlinear fourth-order parabolic equations on the one-dimensional torus are derived. Examples include the thin-film and the Derrida–Lebowitz–Speer–Spohn equations. The schemes conserve the mass and dissipate the entropy. The scheme associated to the logarithmic entropy also preserves the positivity. The idea of the derivation is to reformulate the equations in such a way that the chain rule is avoided. A central finite-difference discretization is then applied to the reformulation. In this way, the same dissipation rates as in the continuous case are recovered. The strategy can be extended to a multi-dimensional thin-film equation. Numerical examples in one and two space dimensions illustrate the dissipation properties.
Key words and phrases:
Entropy, finite differences, thin-film equation, DLSS equation, discrete chain rule, denoising.2000 Mathematics Subject Classification:
35K30, 35Q68, 65M06, 65M12.1. Introduction
The design of numerical schemes that preserve the structure of the associated partial differential equations is an important task in numerical mathematics. In this paper, we develop new finite-difference approximations conserving the mass, preserving the positivity, and dissipating the entropy of nonlinear fourth-order parabolic equations of the type
(1) |
where is the one-dimensional torus, , , and , together with the initial condition in . We also discuss briefly the discretization of multi-dimensional equations; see Section 4.
A special case of (1) (with ) is the thin-film equation
(2) |
which models the flow of a thin liquid along a solid surface with film height or the thin neck of a Hele-Shaw flow in the lubrication approximation [1]. The multi-dimensional version is given by and its discretization will be discussed in Section 4. Another example (with , , ) is the Derrida–Lebowitz–Speer–Spohn (DLSS) equation
(3) |
which arises as a scaling limit of interface fluctuations in spin systems [7] and describes the evolution of the electron density in a quantum semiconductor under simplifying assumptions.
Equations (2) and (3) conserve the mass and preserve the positivity of the solution, even for the corresponding multi-dimensional versions [6, 14]. They also dissipate the entropy111Strictly speaking, equation (1) produces entropy and dissipates energy, but the notion entropy dissipation seems to be common in numerical schemes. in the sense that there exists a Lyapunov functional with entropy density such that the entropy production provides gradient estimates. For instance, (2) and (3) dissipate the entropy density , where
(4) | ||||
if (thin-film equation) and (DLSS equation); see [15, Section 4]. These bounds are optimal. We call the Shannon entropy (density), the logarithmic entropy, and for a Rényi entropy. It holds that pointwise for and pointwise for . We prefer to define as in (4) to avoid the additional term in which would complicate the subsequent computations.
The proof of the dissipation of the entropy is based on suitable integrations by parts and the chain rule [15]. On the discrete level, we face the problem that the chain rule is generally not available. We are aware of two general strategies.
The first strategy is to exploit the gradient-flow structure of the parabolic equation (if it exists). It involves only one integration by parts, and the discrete chain rule can be formulated by means of suitable mean functions. This idea was elaborated as the Discrete Variational Derivative Method for finite-difference approximations [11]. The gradient-flow formulation (with respect to the -Wasserstein metric) yields a natural semi-discretization in time of the evolution using the minimizing movement scheme in finite-dimensional spaces from finite-volume or finite-difference approximations. These techniques were used in [8, 13, 19, 20] for the multi-dimensional DLSS equation. It allows for the proof of entropy dissipation of the Shannon entropy and the Fisher information , but not for general Rényi entropies, since no Wasserstein gradient-flow formulation seems to be available for these functionals. The thin-film equation with is shown in [18] to be a gradient flow with respect to a weighted Wasserstein metric. In the work [22], a finite-difference scheme that dissipates the discrete norm of the solution to the one-dimensional thin-film equation was analyzed.
The minimizing movement scheme is based on the implicit Euler method. We mention that higher-order time discretizations were investigated too, in the framework of semi-discrete problems; see [4] using the two-step BDF method and [17] using one-leg multistep generalizations. A generic framework for Galerkin methods in space and discontinuous Galerkin methods in time was presented in [9].
The second strategy uses time-continuous Markov chains on finite state spaces. Birth-death processes that define the Markov chain can be interpreted as a finite-volume discretization of a one-dimensional Fokker–Planck equation, and the dissipation of the discrete Shannon entropy can be proved. The nonlinear integrations by parts are reduced to a discrete Bochner-type inequality [5, 10, 16], which is obtained by identifying the Radon–Nikodym derivative of a measure involving the jump rates of the Markov chain [3, Section 2]. It seems that this idea is restricted to linear diffusion equations.
In this paper, we suggest a third strategy. The idea is to write the flux as a combination of derivatives of the function . This allows for integrations by parts that can be extended to the discrete level and it avoids the application of the chain rule. More precisely, we determine two functions and depending on and its derivatives such that . The function is known in thermodynamics as the chemical potential, and the formulation of the flux in terms of the chemical potential seems to be natural from a thermodynamic viewpoint. We apply this idea to fourth-order parabolic equations for the first time. It turns out that for with , we can write
where , , and the coefficients depend on , , , and ; see (10) and (11). Integrating by parts twice and using equation (1) gives for :
The task is to show that the integrand, written as a polynomial in , is nonnegative for all values of . It follows from the product rule that, for ,
(5) |
Under certain conditions on the parameters, we expect that the integrand is bounded from below by , up to a factor, which yields some gradient estimates.
On the discrete level, we imitate this idea: The flux and the variables , of the polynomials and is discretized by central finite differences. For this, let be a discrete torus with grid size and define the scheme
(6) |
with the initial condition for , where , , and and are the polynomials and , evaluated at , respectively; see (19). We show below that with the discrete entropy , the discrete analog of (5) becomes
where , approximate , , respectively. This yields exactly the polynomial of the continuous case. Thus, we obtain the same conditions on the parameters as for the continuous equation.
We still need a discrete analog of the product rule to conclude. This is done by carefully choosing and . Definition (20) ensures that which approximates . This choice is used in the central scheme (6) for . Noncentral schemes require different definitions of and ; see Remark 7.
A drawback of our technique is that scheme (6) depends on the entropy to be dissipated. The scheme does not dissipate all admissible entropies. In applications, however, one usually wants to dissipate only that entropy which is physically relevant.
Our main results can be sketched as follows:
- •
- •
-
•
Case : For the case , we need the formulation instead of , where and , since generally does not depend on the logarithm. For details, see Proposition 8.
-
•
Case : We show that scheme (6) with possesses global positive solutions. This result is a consequence of the discrete entropy inequality and mass conservation, which imply that is bounded for all . Consequently, is bounded for all and , and since the function tends to infinity if either or , this proves that is bounded from below and above. We refer to Proposition 6 for details.
-
•
Multi-dimensional case: In principle, the multi-dimensional case can be treated using functions and with many variables. Practically, however, the computations are becoming too involved and it may be unclear how to discretize mixed derivatives. One idea to overcome this issue is to use scalar variables only, like and . This allows us to treat the multi-dimensional thin-film equation; see Proposition 10.
The paper is organized as follows. We prove the continuous entropy inequality (8) in Section 2 and the discrete entropy inequality (9) in Section 3. A scheme for the multi-dimensional thin-film equation is proposed and analyzed in Section 4. Numerical simulations for the thin-film and DLSS equations in one space dimension and for a thin-film equation in two space dimensions are presented in Section 5.
2. General continuous equation
To prepare the discretization, we need to analyze the entropy dissipation properties of the continuous equation (1). We show first that can be written as with and functions and which depend on , , and .
Lemma 1.
The lemma shows in particular that the formulation is not unique. This fact is used to optimize later the range of admissible parameters , , , and .
Proof.
For the following theorem, we recall definition (4) of and set .
Theorem 2.
Proof.
Let first with . We calculate the time derivative of the entropy, using integration by parts twice:
(12) |
where
(13) |
The right-hand side of (12) is nonpositive if for all . Taking into account that , this is the case if and only if
In view of definition (10) of and , we may interpret as a free parameter und and as affine functions in . Therefore, we require that
(14) |
We choose the optimal value of by computing the critical value of :
This yields
(15) |
Inserting in (14) leads to
which is equivalent to , see (7).
If , there exists such that for all , . Inserting this information in (12), we infer that
The discriminant equals
and it is positive for all . Therefore, there exists such that
and this gives the conclusion for and when .
It remains to analyze the case . Here, we cannot formulate the flux as with , since does not contain logarithmic terms. Our idea is to write with and functions and that depend on , , and . The time derivative of the entropy can be written in terms of and its derivatives only, since the logarithmic term only appears with its derivatives.
The formulation corresponds to the expression used for . In fact, we have
where are given by (11) and is a free parameter. As before, the time derivative becomes
Set and . Since and , we obtain
where
(16) |
Observe that . Thus, if the polynomial is nonnegative for all , which is the case if and only if
We choose the optimal value from , i.e.
Then is equivalent to .
Finally, if , we have and consequently,
The discriminant is positive and there exists such that . We infer that
which concludes the proof with and . ∎
Remark 3 (Examples).
Remark 4 (Systematic integration by parts).
The result of Theorem 2 can be also derived by the method of systematic integration by parts of [15]. Our proof is taylored in such a way that it can be directly “translated” to the discrete level. Indeed, the method of [15] needs several chain rules that are not available on the discrete level and our technique needs to be used.
Still, systematic integration by parts and our strategy are strongly related. Systematic integration by parts means that we are adding so-called integration-by-parts formulas whose derivatives vanishe. In this way, we can derive (12),
by choosing . In the method of systematic integration by parts, we are adding a term of the type and optimize . By contrast, the constant is fixed, but we optimize in the formulation of and . In both cases, just one parameter needs to be optimized. ∎
3. General discretized equation
We “translate” the computations of the previous section to the discrete level. For this, we use the discrete entropy
(17) |
where for and is defined in (4). We recall scheme (6):
(18) |
where the functions and are given by
(19) |
is an arbitrary average of around the point , and and are discrete analogs of and :
(20) |
The parameters for are given by (10) and (15). For later use, we note that
implies that
(21) |
Theorem 5.
Proof.
Let , . We compute the time derivative of the discrete entropy, using summation by parts twice:
In the last step, we recognize the discrete analog of the chain rule . Inserting (21), we find that
where is the same polynomial as in (13). The proof of Theorem 2 shows that if . Moreover, if the strict inequality holds, with the same constant as in Theorem 2, which translates into the inequality
finishing the proof. ∎
In Theorem 2, the existence of positive solutions is assumed. We show that such solutions exist globally, at least in case .
Proposition 6.
Proof.
Scheme (18) is a system of ordinary differential equations. According to the Picard–Lindelöf theorem, there exists a unique local positive differentiable solution on the maximal time interval for some . This solution can be extended to if the functions are uniformly positive and bounded. By Theorem 5,
Moreover, scheme (18) conserves the mass, . This shows that
Since diverges for and , there exist constants and such that for all and . Therefore, we can extend the solution globally. ∎
Remark 7 (Noncentral scheme for ).
A more direct discrete analog of and is given by
In this situation, the scheme for needs to be noncentral,
where and are as before. Indeed, it follows from summation by parts for , that
and we can conclude as before. ∎
The case has to be treated in a slightly different way. Since , we consider scheme
(22) |
with and defined via
(23) |
and
(24) |
Proposition 8.
Proof.
We have with and :
By definition of and ,
Therefore,
where is the same polynomial as in (16). It is nonnegative if and only if holds. ∎
4. Discretized multi-dimensional thin-film equation
The ideas of the previous section cannot be adapted in a straightforward way to the multi-dimensional setting, since there are many possibilities to choose the finite differences and the discrete variables. One idea is to employ only scalar variables like , , etc., similarly as for the method of systematic integration by parts of [15]. Still, there does not exist a general approach to define the scalar discrete variables, but we show in this section that the multi-dimensional case can be treated at least in principle. As an example, we consider the thin-film equation
where is the multi-dimensional torus and , and the logarithmic entropy
We show first that we can write , where and , are functions depending on and .
Lemma 9.
It holds that , where
and the parameters are defined by
(25) |
Proof.
We compute
Here, denotes the Hessian matrix of . Identifying the coefficients of and gives the linear system
The unique solution is given by (25). ∎
Proposition 10.
Let . Then for all .
Proof.
The time derivative of becomes
The polynomial
(26) |
is nonnegative in if and only if . Hence, we need to assume that . ∎
Remark 11 (Discussion).
The restriction in the previous lemma comes from the fact that we do not have a free parameter to optimize the inequalities. One may overcome this issue by allowing and to depend on more variables or by assuming that and are matrix-valued with variables like and and to formulate . However, this leads to several parameters that need to be determined and eventually to complicated numerical schemes which seem to be less interesting in practice.
Another way to understand the restriction is from systematic integration by parts. Indeed, computing
we see that is a Lyapunov functional if the last term vanishes, which is the case if . This computation suggests the following generalization: Let and consider the Rényi entropy . Then
and the last term vanishes if . As for the case and , which is discussed below, we expect that the generalization and can be “translated” to the discrete case, but we leave the details to the interested reader. ∎
We turn to the discrete setting. Let be the discrete multi-dimensional torus, be a function, and be the -th unit vector of . We write for and we introduce the finite differences
where and . The discrete divergence of and the discrete gradient and Laplacian of are defined by, respectively,
where . The discrete analogs of and are
where . The numerical scheme reads as
(27) |
and the coefficients are as in (25).
Lemma 12.
Proof.
By the Picard–Lindelöf theorem, there exists a local smooth positive solution . We use the summation-by-parts formula
to compute the time derivative of the entropy:
The product rule reads as
Therefore,
This gives the polynomial (26) which is nonnegative in if and only if . Then for all , and we conclude as in the proof of Proposition 6 that for all and , where are some constants. Thus, the solution can be extended to a global one. ∎
5. Numerical tests
We apply our scheme to the thin-film and DLSS equations on the torus in one and two space dimensions. The system of ordinary differential equations is solved by the command
scipy.integrate.solve_ivp
from the SciPy library, which uses the
Backward Differentiation Formula (BDF) method of variable order or the
implicit Runge–Kutta method of the Radau IIA family of order 5.
We used the default values tol = 1e-3} for the
bsolute tolerance
and tol = 1e-6} fo
the relative tolerance. The local erroris computed
according to tol + rtol *
bs(u).
5.1. DLSS equation
The DLSS equation is solved by scheme (18) using the logarithmic entropy:
where . Figure 1 shows the solution to the DLSS equation at various time steps using the initial datum and the space grid size . We see that the solution is not monotone, since it possesses at and a local maximum. After some time, it approaches the constant steady state given by .

The entropy decay for is illustrated in Figure 2 (left). We used the initial datum for and for . We observe in the semi-logarithmic plot that the decay is exponential, as expected. The rate degrades for larger times when the error dominates, i.e., when the grid is rather coarse.
The error (in space and time) at time is shown in Figure 2 (right), using the initial datum for . As an explicit solution is not known, we use a numerical solution with as the reference solution. As expected, the convergence rate is roughly of second order.


5.2. Thin-film equation
The thin-film equation is solved by scheme (18) using the logarithmic entropy:
where . The solutions at different times, emanating from the initial datum , are shown in Figure 3, where we have chosen . Again, the solutions converge to the constant steady state. The decay of the logarithmic entropy is illustrated in Figure 4, using and the initial datum for . The decay rate is exponential over a large time interval.



Finally, we present a numerical example in two space dimensions. As the initial datum, we choose a lantern picture with pixels in gray scale; see Figure 5 (top left). The evolution of the discrete solution is shown in the remaining panels of Figure 5 for various times. The values and correspond to black and white, respectively. Because of the periodic boundary conditions, we observe a small gray band at the lower right boundary. Interestingly, the solution shows a denoising effect, especially for . For larger times, the diffusion drives the solution to the constant steady state. These results are not surprising, as fourth-order parabolic equations have been suggested in the literature for image denoising. For instance, Bertozzi and Greer [2] analyzed
where is a diffusivity function, while Wei [21] considered
This model was generalized to fractional derivatives; see, e.g., [12]. An example is the equation
which formally reduces to a general thin-film equation in the limiting case . We do not claim that the thin-film equation is a good image denoising model; our numerical example is just a nice illustration.




Finally, we show the entropy decay of the two-dimensional example in Figure 6. The decay rate is exponential until approximately . For later times, the numerical error dominates. Observe, however, that we obtain denoising for very small times, like , where the decay rate is still exponential.

References
- [1] F. Bernis. Viscous flows, fourth order nonlinear degenerate parabolic equations and singular elliptic problems. In: J.I. Díaz et al. (eds.). Free Boundary Problems: Theory and Applications. Longman Sci. Tech., Pitman Res. Notes Math. Ser. 323 (1995), 40–56.
- [2] A. Bertozzi and J. Greer. Low-curvature image simplifiers: global regularity of smooth solutions and Laplacian limiting schemes. Commun. Pure Appl. Math. 57 (2004), 764–790.
- [3] A.-S. Boudou, P. Caputo, P. Dai Pra, and G. Posta. Spectral gap estimates for interacting particle systems via a Bochner-type identity. J. Funct. Anal. 232 (2006), 222–258.
- [4] M. Bukal, E. Emmrich, and A. Jüngel. Entropy-stable and entropy-dissipative approximations of a fourth-order quantum diffusion equation. Numer. Math. 127 (2014), 365–396.
- [5] P. Caputo, P. Dai Pra, and G. Posta. Convex entropy decay via the Bochner–Bakry–Emery approach. Ann. Inst. H. Poincaré Prob. Stat. 45 (2009), 734–753.
- [6] R. Dal Passo, H. Garcke, and G. Grün. On a fourth order degenerate parabolic equation: global entropy estimates and qualitative behaviour of solutions. SIAM J. Math. Anal. 29 (1998), 321–342.
- [7] B. Derrida, J. Lebowitz, E. Speer, and H. Spohn. Fluctuations of a stationary nonequilibrium interface. Phys. Rev. Lett. 67 (1991), 165–168.
- [8] B. Düring, D. Matthes, and J.-P. Milišić. A gradient flow scheme for nonlinear fourth order equations. Discrete Cont. Dyn. Sys. B 14 (2010), 935–959.
- [9] H. Egger. Structure preserving approximation of dissipative evolution problems. Numer. Math. 143 (2019), 85–106.
- [10] M. Fathi and J. Maas. Entropic Ricci curvature bounds for discrete interacting systems. Ann. Appl. Prob. 26 (2016), 1774–1806.
- [11] D. Furihata and T. Matsuo. Discrete Variational Derivative Method. Chapman and Hall/CRC Press, Boca Raton, Florida, 2010.
- [12] P. Guidotti and K. Longo. Well-posedness for a class of fourth order diffusions for image processing. Nonlin. Diff. Eqs. Appl. NoDEA 18 (2011), 407–425.
- [13] X. Huo and H. Liu. A positivity-preserving and energy stable scheme for a quantum diffusion equation. Submitted for publication, 2019. arXiv:1912.00813.
- [14] A. Jüngel and D. Matthes. The Derrida–Lebowitz–Speer–Spohn equation: existence, non-uniqueness, and decay rates of the solutions. SIAM J. Math. Anal. 39 (2008), 1996–2015.
- [15] A. Jüngel and D. Matthes. An algorithmic construction of entropies in higher-order nonlinear PDEs. Nonlinearity 19 (2006), 633–659.
- [16] A. Jüngel and W. Yue. Discrete Bochner inequalities via the Bochner–Bakry–Emery approach for Markov chains. Ann. Appl. Prob. 27 (2017), 2238–2269.
- [17] A. Jüngel and J.-P. Miličić. Entropy dissipative one-leg multistep time approximations of nonlinear diffusive equations. Numer. Meth. Partial Diff. Eqs. 31 (2015), 1119–1149.
- [18] S. Lisini, D. Matthes, and G. Savaré. Cahn–Hilliard and thin film equations with nonlinear mobility as gradient flows in weighted-Wasserstein metrics. J. Diff. Eqs. 253 (2012), 814–850.
- [19] J. Maas and D. Matthes. Long-time behavior of a finite volume discretization for a fourth order diffusion equation. Nonlinearity 29 (2016), 1992–2023.
- [20] D. Matthes and H. Osberger. A convergent Lagrangian discretization for a nonlinear fourth-order equation. Found. Comput. Math. 17 (2017), 73–126.
- [21] G. Wei. Generalized Perona–Malik equation for image restoration. IEEE Signal Process. Lett. 6 (1999), 165–167.
- [22] L. Zhornitskaya and A. Bertozzi. Positivity-preserving numerical schemes for lubrication-type equations. SIAM J. Numer. Anal. 37 (2000), 523–555.