SOME VIEWS ON GLOBAL REGULARITY OF THE THIN FILM EQUATION

Similar documents
THE LUBRICATION APPROXIMATION FOR THIN VISCOUS FILMS: REGULARITY AND LONG TIME BEHAVIOR OF WEAK SOLUTIONS A.L. BERTOZZI AND M. PUGH.

Recent progress on the vanishing viscosity limit in the presence of a boundary

MATH 124B Solution Key HW 05

Global well-posedness and decay for the viscous surface wave problem without surface tension

The Dirichlet s P rinciple. In this lecture we discuss an alternative formulation of the Dirichlet problem for the Laplace equation:

Getting started: CFD notation

We begin by considering the following three sequences:

Class Meeting # 2: The Diffusion (aka Heat) Equation

Boundary Conditions in Fluid Mechanics

Existence and Uniqueness

Math background. Physics. Simulation. Related phenomena. Frontiers in graphics. Rigid fluids

Course Description for Real Analysis, Math 156

13 PDEs on spatially bounded domains: initial boundary value problems (IBVPs)

CHAPTER 7 SEVERAL FORMS OF THE EQUATIONS OF MOTION

A LOWER BOUND ON BLOWUP RATES FOR THE 3D INCOMPRESSIBLE EULER EQUATION AND A SINGLE EXPONENTIAL BEALE-KATO-MAJDA ESTIMATE. 1.

Optimality Conditions for Constrained Optimization

2. Conservation of Mass

Weak solutions for the Cahn-Hilliard equation with degenerate mobility

The Skorokhod reflection problem for functions with discontinuities (contractive case)

A dual form of the sharp Nash inequality and its weighted generalization

Low Froude Number Limit of the Rotating Shallow Water and Euler Equations

CHAPTER 3. Sequences. 1. Basic Properties

Solution of Partial Differential Equations

DYNAMIC BIFURCATION THEORY OF RAYLEIGH-BÉNARD CONVECTION WITH INFINITE PRANDTL NUMBER

Math 5587 Lecture 2. Jeff Calder. August 31, Initial/boundary conditions and well-posedness

EULERIAN DERIVATIONS OF NON-INERTIAL NAVIER-STOKES EQUATIONS

Boundary value problems for partial differential equations

REMARKS ON THE VANISHING OBSTACLE LIMIT FOR A 3D VISCOUS INCOMPRESSIBLE FLUID

Blowup dynamics of an unstable thin-film equation

1 The Observability Canonical Form

A VERY BRIEF REVIEW OF MEASURE THEORY

MATH Final Project Mean Curvature Flows

PDE Solvers for Fluid Flow

REAL ANALYSIS LECTURE NOTES: 1.4 OUTER MEASURE

Math 541 Fall 2008 Connectivity Transition from Math 453/503 to Math 541 Ross E. Staffeldt-August 2008

FOURTH ORDER CONSERVATIVE TWIST SYSTEMS: SIMPLE CLOSED CHARACTERISTICS

n v molecules will pass per unit time through the area from left to

Introduction to Real Analysis Alternative Chapter 1

Chapter 9: Differential Analysis

Part III. 10 Topological Space Basics. Topological Spaces

2. FLUID-FLOW EQUATIONS SPRING 2019

arxiv: v2 [math.ap] 28 Nov 2016

The first order quasi-linear PDEs

We are going to discuss what it means for a sequence to converge in three stages: First, we define what it means for a sequence to converge to zero

Continuum Mechanics Fundamentals

Chapter 1 The Real Numbers

( ε > 0) ( N N) ( n N) : n > N a n A < ε.

Problem 3. Give an example of a sequence of continuous functions on a compact domain converging pointwise but not uniformly to a continuous function

Damped harmonic motion

On 2 d incompressible Euler equations with partial damping.

Lecture 1: Introduction to Linear and Non-Linear Waves

5 Birkhoff s Ergodic Theorem

Chapter 9: Differential Analysis of Fluid Flow

Nonlinear Analysis 71 (2009) Contents lists available at ScienceDirect. Nonlinear Analysis. journal homepage:

arxiv: v1 [math.ap] 21 Dec 2016

Équation de Burgers avec particule ponctuelle

CONNECTIONS BETWEEN A CONJECTURE OF SCHIFFER S AND INCOMPRESSIBLE FLUID MECHANICS

Long-wave instabilities and saturation in thin film equations

u xx + u yy = 0. (5.1)

Example 1. Hamilton-Jacobi equation. In particular, the eikonal equation. for some n( x) > 0 in Ω. Here 1 / 2

Multiple integrals: Sufficient conditions for a local minimum, Jacobi and Weierstrass-type conditions

Introductory Analysis I Fall 2014 Homework #9 Due: Wednesday, November 19

arxiv: v1 [math.ap] 10 Apr 2013

Fundamentals of Fluid Dynamics: Ideal Flow Theory & Basic Aerodynamics

Navier-Stokes Equation: Principle of Conservation of Momentum

The Hopf equation. The Hopf equation A toy model of fluid mechanics

NONSTANDARD NONCONFORMING APPROXIMATION OF THE STOKES PROBLEM, I: PERIODIC BOUNDARY CONDITIONS

AA210A Fundamentals of Compressible Flow. Chapter 5 -The conservation equations

Math 342 Partial Differential Equations «Viktor Grigoryan

(x k ) sequence in F, lim x k = x x F. If F : R n R is a function, level sets and sublevel sets of F are any sets of the form (respectively);

TRAVELING WAVES IN 2D REACTIVE BOUSSINESQ SYSTEMS WITH NO-SLIP BOUNDARY CONDITIONS

RECOVERY OF NON-LINEAR CONDUCTIVITIES FOR CIRCULAR PLANAR GRAPHS

7 The Navier-Stokes Equations

Behaviour of Lipschitz functions on negligible sets. Non-differentiability in R. Outline

Relevant sections from AMATH 351 Course Notes (Wainwright): 1.3 Relevant sections from AMATH 351 Course Notes (Poulin and Ingalls): 1.1.

PDEs, part 1: Introduction and elliptic PDEs

1 Lyapunov theory of stability

TMA 4180 Optimeringsteori KARUSH-KUHN-TUCKER THEOREM

Some Background Material

On the Optimal Insulation of Conductors 1

L p -boundedness of the Hilbert transform

Iowa State University. Instructor: Alex Roitershtein Summer Homework #5. Solutions

Math Tune-Up Louisiana State University August, Lectures on Partial Differential Equations and Hilbert Space

Shell Balances in Fluid Mechanics

Applied Mathematics Letters

On Friedrichs inequality, Helmholtz decomposition, vector potentials, and the div-curl lemma. Ben Schweizer 1

are harmonic functions so by superposition

Diffusion on the half-line. The Dirichlet problem

P(E t, Ω)dt, (2) 4t has an advantage with respect. to the compactly supported mollifiers, i.e., the function W (t)f satisfies a semigroup law:

Divergence Formulation of Source Term

Statistics 612: L p spaces, metrics on spaces of probabilites, and connections to estimation

Convex Functions and Optimization

1 Introduction. or equivalently f(z) =

Waves in Flows. Global Existence of Solutions with non-decaying initial data 2d(3d)-Navier-Stokes ibvp in half-plane(space)

Metric Spaces and Topology

Singularities and Laplacians in Boundary Value Problems for Nonlinear Ordinary Differential Equations

Spanning and Independence Properties of Finite Frames

ON THE REGULARITY OF SAMPLE PATHS OF SUB-ELLIPTIC DIFFUSIONS ON MANIFOLDS

Math 4606, Summer 2004: Inductive sets, N, the Peano Axioms, Recursive Sequences Page 1 of 10

Exercises: Brunn, Minkowski and convex pie

Transcription:

SOME VIEWS ON GLOBAL REGULARITY OF THE THIN FILM EQUATION STAN PALASEK Abstract. We introduce the thin film equation and the problem of proving positivity and global regularity on a periodic domain with a slip parameter n > 1. Several ideas in the direction of a proof are described. Some are recounted only briefly and qualitatively due to a lack of either an essential ingredient of the proof or enough time to carry it through. This document should be considered as both a record of the author s effort up to this point and a source of inspiration for future work on the problem. 1. Introduction We consider the equation for the evolution of the height of a thin film of viscous fluid with the simplifying assumption that the effect of surface tension is very strong compared to advection: { ht + (f(h)h xxx ) x = 0 (1). h(x, 0) = h 0 (x) We will derive this in the next section. Here f : R R is given, h : T 1 [0, T ) R and h 0 : T 1 (0, ), h 0 H 1 (T ) where T 1 is the one-dimensional 1 torus or, equivalently, a periodic arrangement on the whole real line. The choice of f is based on the most appropriate physical model for the interaction between the fluid and the substrate, see for example Greenspan [4]. For simplicity of the analysis it is standard to consider f(h) = h n with n > 0 determining the extent of the slip. Starting from positive initial data, it is straightforward to observe that as long as a smooth solution exists and stays positive, f(h) is uniformly away from zero by compactness, so the equation is uniformly parabolic. Thus at least for short times, the solution is smooth and positive. One introduces a monotonic quantity called the entropy (to be discussed in greater detail later) which controls the thinness of the film and allows one to deduce that solutions, and in fact weak solutions, are nonnegative (see Bernis-Friedman [1]). Thus we will tend to omit the absolute value and consider f(h) = h n. The essential conclusion of the above discussion is that to prove global wellposedness of (1), it suffices to prove that the equation preserves the strict positivity of the data. The purpose of this report is to recount several good ideas in this direction that are not successful in their current forms. We consider the problem only for n > 1 because the numerical results of Bertozzi et al. strongly indicate Date: 22 March 2018. This work was done for a course on thin films taught by Andrea Bertozzi. 1 The system could just as well be posed in several dimensions, but the one dimensional problem is already hard enough. The torus could also be replaced with an interval in R with the Neumann boundary data h x = h xxx = 0. 1

2 STAN PALASEK breakdown of smoothness and positivity for smaller values [3]. On the other hand, it is quite straightforward to prove well-posedness for n 4. Indeed, approximating the equation with a uniformly parabolic nonlinearity (modifying f), the approximate solutions inherit the Cx 1/2 Hölder property from the data, which is passed on to the exact solution upon taking a uniform limit. If h = 0 at a point, it has to stay near zero in a Hölder sense, which would force the entropy to be infinite unless n < 4. The same idea can be refined to prove global regularity for n 7/2 (see [2, 3]). What follows are attempts to lower this bound further still. 2. Almost-transported quantities in the fluid interior Since the dynamics of the free interface are ultimately determined by the dynamics of the fluid in the interior, it seems natural to understand the equations for that motion. A sensible place to start is by examining the derivation of (1) from the Navier-Stokes equations. What follows is known as the lubrication approximation and can be found in [4]. For t [0, T ), let (u(t), v(t)) : Ω(t) = {(x, y) T 1 (0, ) : y < h(x, t)} R 2 be the velocity field of the fluid. First we have the kinetic equation to assert that the interface moves with the fluid: (2) h t + u y=h h x = v y=h. Next, the surface tension γ which balances the fluid pressure p: (3) p = p 0 γh xx. We assume the shear stress at the boundary vanishes, (4) u y y=h = 0, the fluid is incompressible, (5) u x + v y = 0, and the fluid slips against the substrate at a velocity depending on the interface: (6) u y=0 = u y y=h κ(h), v y=0 = 0. The choice of κ in (6), the so-called slip condition, is more a matter of physics than of analysis and there is a great deal of variability among authors. One finds that κ(h) = 1 3 (hn 2 h) produces (1) with f(h) = h n in the nonlinearity, which will be our choice in the interest of convenience. A significant note is that n = 3 produces κ(h) = 0 which is known as the no-slip condition. Of course we are still yet to give the law of motion for the fluid. Here we invoke the assumption that pressure and viscosity are the only relevant forces in the fluid interior, and that the film is thin enough that the pressure gradient in the vertical direction is negligible. One arrives at the Stokes equation in the horizontal direction, (7) p x = µu yy, p y = 0

SOME VIEWS ON GLOBAL REGULARITY OF THE THIN FILM EQUATION 3 where µ is the viscosity. It is straightforward to integrate the equations (3) through (7) to find the velocity profiles u = γ µ (hκ(h) + yh 1 2 y2 )h xxx v = γ µ y(κ(h) + hκ (h) + 1 2 y)h xh xxx + γ µ y( hκ(h) 1 2 yh + 1 6 y2 )h xxxx Putting these into (2), we arrive at the equation (8) h t + γ µ (h2 ( 1 3 h + κ(h))h xxx) x = 0 and finally normalize γ/µ = 3. Taking f(h) = h 2 ( 1 3h + κ(h)), we recover the thin film equation (1) for an arbitrary slip condition. Next let us rewrite the velocity field with the κ that corresponds to f(h) = h n : u = (h n 1 h 2 + 3yh 3 2 y2 )h xxx v = y(2h (n 1)h n 2 3 2 y)h xh xxx + y(h 2 h n 1 3 2 yh + 1 2 y2 )h xxxx. One notices that defining the potential on Ω(t) (9) φ = y(h 2 h n 1 3 2 yh + 1 2 y2 )h xxx, the flow in the interior is given exactly by 2 (10) (u, v) = ( y φ, x φ) = φ. One can go further and find several infinite families of potentials with different particle dynamics but induce the same boundary dynamics, that is to say equations (1) and (10) (2) are equivalent. For example, taking n = 3 in (9), one finds φ = 1 2 y2 (y 3h)h xxx which suggests the potential (11) φ = 1 n 1 yn 1 (y nh)h xxx for a general n. It is trivial to compute that this too induces the correct boundary dynamics. It is this description of the internal fluid dynamics that leads to a very nice but incorrect proof that (1) preserves strict positivity of the data for a wider range of n. Conjecture. If n > 3/2 and u C 1/8 t C 1/2 x L 2 t H 1 x(t 1 R + ) is a weak solution 3 of (1) with data u 0 H 1 (T 1 ) satisfying u 0 > 0, then u > 0 for all time. 2 As a result, (1) can be rewritten as th = ν φ y=h where ν is the outward normal scaled by (1 + h 2 x) 1/2, the length element of the interface. This is in direct analogy to the Zakharov formulation of the water wave equations using the Dirichlet-to- Neumann operator [5]. 3 There are several notions of weak solution, see eg. [2]. The ideas discussed here could be applied at a sequence of times approaching the blowup time to derive a contradiction. Thus one can imagine all the solutions to be in the classical sense as long as it is kept in mind that the high derivatives may not stay uniformly bounded as t T.

4 STAN PALASEK Sketch of an interesting argument which fails. Suppose the contrary, that is there exists a smallest T > 0 such that there exists x 0 T 1 with u(x 0, T ) = 0. Recall from the discussion in the introduction that u is C 1/2 in space, say for simplicity with norm 1. Let us identify T 1 with the real interval [ 1/2, 1/2] so that u(x 0, T ) = 0 implies u(x, T ) x x 0 1/2 for all x T 1. The next observation is that according to (10), in a Lagrangian frame each particle moves along level sets of φ (as φ φ = 0 for any potential) so φ is constant along characteristics. This suggests we might reach a contradiction by keeping track of the number of particles with very small φ and proving that there are more at the time of blowup than there were to begin with. (Note this is indeed a contradiction because of the incompressibility assumption.) Consider the fluid region at time T which lies within a horizontal distance of 0 < δ 1 from x 0. Thanks to the 1/2-Hölder estimate, this region has volume on the order of δ 3/2. Since the bulk of the particles in this region have height on the order of δ 1/2, the potential defined in (11) is on the order δ n/2 h xxx. At time t = 0 when the interface is uniformly away from h = 0, φ is smallest for particles near the substrate. The same volume δ 3/2 can be found in the region bounded above by height δ 3/2, but now φ is on the order of (δ 3/2 ) n 1. Since φ is only transported in a divergence-free manner, the volume of fluid with φ of a certain order is conserved, so it must be that δ n/2 h xxx (δ 3/2 ) n 1 as δ 0, that is h xxx δ n 3/2. It is well-known that h xxx is unbounded near a singularity (see eg. [3]) so it must be that n 3/2. These heuristic estimates can be made more precise. The range of n can perhaps be widened even further by using estimates to the effect that h xxx grows at least with a power of h 1 near a singularity. Unfortunately, this method falls short in its current form. It is tempting to say that because particles follow level curves of the potential, it must be constant in Lagrangian coordinates. But this cannot be the case, as φ at a fixed Eulerian point changes with respect with time. In fact, as can be deduced from [2], h xxx 0 (in, say, L 2 ) and h h 0 uniformly as t so φ 0 everywhere. This argument might be completed by showing that over the short time scale on which blowup can occur (thanks to, for example, the uniform convergence of h to its average), the change in φ along a characteristic is higher order than δ (3/2)(n 1) so it can be neglected in the δ 0 limit. 4 3. Incompressibility near a singularity As before, suppose blowup first occurs at time T, located at x 0. One can imagine two ways a contradiction can be derived by examining the local movement of volume in the fluid region near the singularity. First, by Hölder continuity (again assuming the constant is 1), there is a region U = {(x, y) I (0, A) : y > x x 0 1/2 } where A > 0 and I is an interval containing x 0, such that U was full of fluid at 4 What follows is a strange idea far outside the scope of this paper: framing the thin film equation as the evolution of the height of a free interface above an incompressible vector field is physically natural because that is the precise scenario the equation was derived to model. But interestingly, there are infinitely-many qualitatively different choices of particle dynamics that yield the same equation (in addition to (9) which is derived from the physics). It is certainly the case that other evolution equations can be formulated this way even ones that do not literally model a fluid interface. Plausibly, blowup can be globally excluded by exhibiting a geometric contradiction to the existence of a vector field satisfying the equation lying underneath a singularity.

SOME VIEWS ON GLOBAL REGULARITY OF THE THIN FILM EQUATION 5 some t < T but is empty by T. Thus a particle in U at a time t < T would have T t time units to travel the distance to the complement of U. Second, one could consider V (t) = {(x, y) I (0, ) : y < h(x, t)}. As t T, the volume of this region tends to decrease, so by incompressibility, the net fluid flow must be outward. One might derive a contradiction by showing the rate at which volume can exit the region is slower than the rate at which space is being taken away by the shrinking of V (t). Or, that the only way to escape is to a thin layer above the substrate outside of I T 1, and bound how quickly volume can leave that layer (since the only way out is in the vertical direction). How close this method gets to a contradiction depends heavily on the choice of φ. For example, one could choose φ = h n h xxx, so that t h + φ x + φ y h x y=h = 0 trivially gives (1), but then particles freely move in the vertical direction exactly in parallel with the boundary so there is no hope of a contradiction. On the other hand, it is straightforward compute that the family of potentials (12) φ = y k h n k h xxx, k R produce the correct interface dynamics, as do convex combinations thereof (of course if one desires that the trajectories stay in the upper-half plane, k > 0 is necessary). Recalling (10), the u and v scale like y k 1 and y k respectively for particles with small y that are uniformly away from the interface. Therefore by taking k very large, the velocities in thin strip above the substrate but away from a singularity can be made arbitrarily small. Moreover, writing φ = (y/h) k h n h xxx, it is clear that choosing k large has little effect near a singularity where y is close to h. It follows that as t T, V (t) loses volume, leaving such a thin strip for the fluid to escape. But the fluid which started in that thin strip can barely leave by the choice of k large. This hints of a potential obstruction to a singularity which should be explored. 4. Exploiting global monotonic quantities The ideas discussed so far have not directly made use of the monotonic and conserved quantities enjoyed by (1). For classical solutions it is elementary to check that we have conservation of mass, d (13) h = 0, dt T 1 conservation of energy, d 1 (14) h 2 x + h n h 2 xxx = 0, dt 2 T 1 and a monotonic quantity called entropy that was originally introduced in [1], d 1 (15) + (n 2)(n 1) h 2 dt T hn 2 xx = 0. 1 T 1 The last equality is only true and useful when n > 2. An analogous statement holds for n = 2 when 1/h n 2 is replaced by a logarithm. These bounds offer quite a bit of control, and it is not absurd to suppose that they might be almost sufficient to prove global well-posedness for n 2. Indeed, (15) was sufficient for n 4 [1], and a refinement of it was sufficient for n 7/2 [3]. An argument that additionally uses (13) may proceed as follows: in the event that the interface touches down, 1/h n 2 is larger in a neighborhood of the singularity,

6 STAN PALASEK which forces 1/h n 2 to increase elsewhere. But too great an increase elsewhere would fail to conserve mass, violating (13). One can formulate increasingly careful estimates to this effect, but to date the results are only local. That is to say we can conclude the desired positivity if any of the following is true: the data is small (in Ḣ 1 ), the minimum of the data is not too small, the torus is small, or we consider only small times. In particular, it is not difficult to obtain global well-posedness under the assumption (16) L 1/2 h 0 Ḣ1 c n min h 0 where c n is a constant depending on n (2, 4) and L = T 1. One could hope to show that the small time when positivity is guaranteed contradicts upper bounds for T. For example, applying (15) along with the interpolation inequality u 2 L 2 t Ḣ2 u 4 /M where M = h L 4 t Ẇ x 1,4/3 0 and Grönwall s inequality, we find for a constant d n (17) T which easily implies (16). ( 1 d n min h n 2 0 ( L h 0 Ḣ1) n 2 ) L 6 M 2 5. Shape of the interface at the singularity It is straightforward to see for n > 2 that due to Hölder continuity of h and the finiteness of the entropy (15), h cannot have finite left or right derivatives at a singularity; otherwise h will be too small over too much space for 1/h n 2 to be integrable. Thus we expect the liminf of the the left derivative to be and the limsup of the right derivative to be + as t T. But by smoothness of h before T, it has a global minimum where h x vanishes and h xx > 0. From all these considerations, one may very informally expect that h should have its third derivative large and positive approaching the minimum from the left, and large and negative from the right. This might be in direct conflict with the observation that at the local minimum from h x = 0, (1) reduces to h t = h n h xxxx, so h xxxx > 0 on a set of positive measure. Although this discussion hints at a substantial obstruction to a singularity for n > 2, there are many difficulties in making this reasoning rigorous. For example, we must contend with a severe lack of control over the high derivatives of h near T. Furthermore we must settle for only knowing h xxxx > 0 on a potentially small subset of times leading of to T. 6. Conclusion This is just a sample of the author s (so far) fruitless approaches to the thin film global regularity problem. More work is certainly warranted to push these ideas to their conclusions. References [1] F. Bernis and A. Friedman. Higher order nonlinear degenerate parabolic equations. Journal of Differential Equations, 83(1):179 206, 1990. [2] A. Bertozzi and M. Pugh. The lubrication approximation for thin viscous films: regularity and long-time behavior of weak solutions. Communications on pure and applied mathematics, 49(2):85 123, 1996.

SOME VIEWS ON GLOBAL REGULARITY OF THE THIN FILM EQUATION 7 [3] A. L. Bertozzi, M. P. Brenner, T. F. Dupont, and L. P. Kadanoff. Singularities and similarities in interface flows. In Trends and perspectives in applied mathematics, pages 155 208. Springer, 1994. [4] H. P. Greenspan. On the motion of a small viscous droplet that wets a surface. Journal of Fluid Mechanics, 84(1):125 143, 1978. [5] V. E. Zakharov. Stability of periodic waves of finite amplitude on the surface of a deep fluid. Journal of Applied Mechanics and Technical Physics, 9(2):190 194, 1968. UCLA Department of Mathematics E-mail address: palasek@math.ucla.edu