The gradient-flow structure of non-Newtonian thin-film equations

Oberseminar Nichtlineare Differentialgleichungen, Universität Stuttgart 2023

This is the manuscript of a blackboard talk. For the full result see https://arxiv.org/abs/2301.10300

Introduction

We consider a thin liquid film of an

  • incompressible \(\operatorname{div} u =0\)
  • viscous
  • non-Newtonian \(\mu = \mu(\epsilon)\) fluid on a solid bottom such that
  • capillary forces dominate. We assume that the free boundary is given by the graph of a function \(h=h(t,x)\)

Assumptions

  • We assume that the aspect ratio \(\varepsilon = \frac{H}{L} \ll 1\) is very small. This will allow us to derive a closed equation for the film height \(u\) via asymptotic analysis.
  • We assume that the problem is one-dimensional, that is that the fluid is homogenenous in \(y\)-direction.

Governing equations

The dynamics of the full problem is given by the Navier-Stokes equations in the moving domain \(\Omega(t)\):

\[\left\{ \begin{array}{rcl} \rho(\partial_t \vec{u} + (\vec{u}\cdot \nabla) \vec{u}) & = & \operatorname{div}\cdot T(p,\vec{u}) \\ \operatorname{div} \vec{u} & = & 0 \\ & \vdots & \\ \text{B.C.} \end{array} \right.\]

For simplicity we omit gravitational forces by assuming they are much smaller than the capillary forces.

Rheology

  • here: \(T(p,\vec{u}) = -p\mathrm{Id} + \mu(D)D\) is Cauchy stress tensor (\(D\) is the symmetric gradient)
  • shear dependent viscosity \(\mu\)
    • \(\mu\equiv c\) -> Newtonian fluid
    • \(\mu' >0\) -> shear-thickening fluid
    • \(\mu'<0\) -> shear-thinning fluid
  • we will assume that the viscosity is explicitly given by the Ostwald-deWaele law \(\mu(\epsilon) = \mu_0|\epsilon|^{\frac{1}{\alpha}-1}\)

Boundary conditions

  • kinematic boundary condition: \(\partial_t h + u_h \partial_x h = u_v\) (particles that are on the boundary remain on the boundary)
  • slip condition at solid bottom (e.g. \(\vec{u}=0\))
  • stress-balance condition: \(T(p,\vec{u}) \cdot \vec{n} = \gamma \kappa \vec{n}\), where \(\gamma\) is the strength of surface tension, \(\kappa\) is the curvature and \(\vec{n}\) is the unit outer normal to the free boundary Idea: surface tension tries to equilibrate surface area (=length) of the surface
\[E_{surface}[h] = \int_{\Omega} \sqrt{1+|\partial_xh(x)}\, \mathrm{d}x \sim |\Omega| + \int_{\Omega}\tfrac{1}{2} |\partial_x h(x)|^2 \, \mathrm{d}x\]

Lubrication approximation

Asymptotic expansion in aspect ratio \(\varepsilon\) and sending \(\varepsilon \searrow 0\), we obtain a closed equation for the film height

\[\partial_t h + \partial_x\bigl( m(h)|\partial_x^3h|^{\alpha-1} \partial_x^3h \bigr) = 0\]

Gradient flows

Gradient flows are evolutionary systems driven by an energy, in the sense that the energy decreases along solutions, as fast as possible. Clasically,

\[\dot{x} = - \nabla E[x]\]

and we find that \(E\) decreases along solutions

\[\frac{d}{dt} E[x(t)] = \dot{x}(t) \nabla E[x(t)] = -|\nabla E[x(t)]|^2\]

What does as fast as possible mean? This is what is guaranteed by the dissipation mechanism.

Typically, equations have various gradient-flow structures, depending on the choice of energy and dissipation.

The thin-film equation

\[\left\{ \begin{array}{rcll} \partial_t h + \partial_x\bigl( m(h)|\partial_x^3h|^{\alpha-1} \partial_x^3h \bigr) & = & 0, & t>0,x\in \Omega\\ \partial_x h = m(h)|\partial_x^3 h|^{\alpha-1}\partial_x^3h & = & 0, & t>0,x\in \partial\Omega\\ h(0,x) & = & h_0(x), & x\in \Omega \end{array} \right.\]

Note that solutions conserve their mass (this becomes evident either from modelling or from testing this equation with the function constantly equal to one):

\[\frac{d}{dt} \int_{\Omega} h(t,x) \, \mathrm{d}x = 0.\]

Then \(h\) solves a continuity equation, i.e. there is a flux \(j\) such that

\[\left\{\begin{array}{rcll} \partial_t h + \partial_x j & = & 0, &t>0,x\in\Omega \\ \partial_x h = j & = & 0, & t>0, x\in\partial\Omega. \end{array}\right.\]

We have hence transformed the problem to finding a pair \((h,j)\) with \(j=m(h)\partial_x^3h\). The capillary effects forces the film to reduce surface energy. This leads to the choice of the energy

\(E[h] = \int_{\Omega} \tfrac{1}{2} |\partial_x h|^2 \,\mathrm{d} x.\) Then \(\frac{d}{dt} E[h](t) = \int_{\Omega}\partial_{tx} h \partial_x h\, \mathrm{d}x = \underbrace{- \int_{\Omega} \partial_t h\partial_x^2 h \, \mathrm{d} x}_{=\langle DE[h],-\partial_x j\rangle} = \int_{\Omega} \partial_x^2 \partial_x j\, \mathrm{d} x = -\int_{\Omega} \partial_x^3 h j \, \mathrm{d}x.\)

But how does \(j\) dissipate energy? We need to choose a dissipation potential. We make the educated guess that \(j\) is choosen as the minimiser of the functional

\[\langle DE[h],-\partial_x j\rangle + \tfrac{\alpha}{\alpha+1} \int_{\Omega} \frac{|j|^{\frac{\alpha+1}{\alpha}}}{m(h)^{\frac{1}{\alpha}}}\,\mathrm{d} x.\]

Since if \((h,j)\) solve the continuity equation, this implies that (using that \(j\) must be a critical point of this equation)

\[\frac{|j|^{\frac{\alpha+1}{\alpha}-1}j}{m(h)^{\frac{1}{\alpha}}} = -\partial_x DE[h] = \partial_x^3h,\]

and hence

\[j = m(h)|\partial_x^3 h|^{\alpha-1}\partial_x^3h.\]

Hence, we look for pairs \((h,j)\) that solve the continuity equation and that minimise the dissipation relation.

Excurse: Positive solutions

Eventually, we want to use this structure to find solutions. But: while for Newtonian thin-film equations, one can show non-negativity of solutions by using entropy methods, this method fails in general for non-Newtonian fluids.

For shear-thinning fluids, Ansini and Giacomelli introduced an additional term in the dissipation to obtain entropy methods for a modified equation. We choose a different approach: since we are working in an energy-driven scheme anyway, we modify the energy to force solutions away from zero: therefore, we introduce a strongly singular potential \(G\) with

  • \(G_{\sigma} \in C^2((0,\infty))\)with \(G\geq 0\)
  • \(G_{\sigma}\) is convex
  • \(G_{\sigma}(s)= 0\) for \(s\geq 2\sigma\) and \(G_{\sigma}(s) \geq s^{-2}\) for \(s\leq \sigma\)

Then, we introduce the new energy

\[E^{\sigma}[h] = \int_{\Omega} \tfrac{1}{2} |\partial_x h|^2 +G_{\sigma}(h) \, \mathrm{d}x.\]

Note that the control of \(h\) in \(H^1(\Omega)\) gives \(h\in C^{1/2}(\Omega)\) and hence \(E^{\sigma}[h]<\infty\) implies that \(h\) must be strictly positive, since if \(h(x_0) = 0\), we would get

\[\int_{x_0}^{x} G_{\sigma}(h)\, \mathrm{d}y \geq \int_{x_0}^{x} |h(y)-h(x_0)|^{-2} \,\mathrm{d} y \geq \int_{x_0}^{x} |x-x_0|^{-1} \,\mathrm{d}y = +\infty.\]

In fact, families with uniformly bounded energy are uniformly bounded below.

We then obtain \(DE^{\sigma}[h] = \partial_x^2 h - G_{\sigma}'(h)\) and

\[j = m(h) |\partial_x^3h - G_{\sigma}''(h)\partial_xh|^{\alpha-1}(\partial_x^3h - G_{\sigma}''(h)\partial_xh).\]

Minimising-movement scheme

We may use this method to construct solutions. Recall that we look for a pair of solutions \((h,j)\) that satisfies the continuity equation

\[\left\{\begin{array}{rcll} \partial_t h + \partial_x j & = & 0, &t>0,x\in\Omega \\ \partial_x h = j & = & 0, & t>0, x\in\partial\Omega. \end{array}\right.\]

and such that infinitesimally \(j\) is chosen as the minimiser of the energy-dissipation functional

\[\frac{d}{dt}E^{\sigma}[h] + \frac{\alpha}{\alpha+1} \int_{\Omega} \frac{|j|^{\frac{\alpha+1}{\alpha}}}{m(h)^{\frac{1}{\alpha}}}\, \mathrm{d} x\]

Integrating in time, we get

\[E^{\sigma}[h](t_1) + \frac{\alpha}{\alpha+1} \int_{t_0}^{t_1}\int_{\Omega} \frac{|j|^{\frac{\alpha+1}{\alpha}}}{m(h)^{\frac{1}{\alpha}}}\,\mathrm{d}x \,\mathrm{d}s = E^{\sigma}[h](t_0).\tag{1}\]

One typically discretises the time, that is we introduce a minimising-movement scheme. The idea is, that if the system at time \(t\) is in a state \(h^*\), then after some time \(\tau\) the system is in the state \(h(t+\tau)\) chosen by following the curve \((h,j)\) solving the discretised continuity equation

\[\left\{\begin{array}{rcll} \frac{h-h^*}{\tau} + \partial_x j & = & 0, &t>0,x\in\Omega \\ j & = & 0, & t>0, x\in\partial\Omega. \tag{2} \end{array}\right.\]

and minimising the left-hand side of \((1)\). That is, we look for a minimiser \((h,j)\) of

\[\mathcal{F}^{\sigma}_{\tau,h^*}[h,j] = E^{\sigma}[h] + \tau\tfrac{\alpha}{\alpha+1} \int_{\Omega} \frac{|j|^{\frac{\alpha+1}{\alpha}}}{m(h^*)^{\frac{1}{\alpha}}} \, \mathrm{d} x.\tag{3}\]

Note that we have changed the denominator of the dissipation potential to be evaluated at the previous state \(h^*\) to make minimisation easier. It will turn out that solutions are continuous in time, hence this will not cause additional complications.

Then, we can define the minimising-movement scheme with time step \(\tau>0\):

\[\begin{split} h^{\sigma}_{\tau,0} & = 0\\ (h^{\sigma}_{\tau,k},j^{\sigma}_{\tau,k}) & = \text{minimiser of }\mathcal{F}_{\tau,h^{\sigma}_{\tau,k-1}} \text{ among all pairs } (h,j) \text{ solving }(2). \end{split}\]

This is a minimisation problem of a strictly convex functional with linear constraint. Unique minimisers exist by the direct method of the calculus of variations. The Euler-Lagrange equation is given by

\[\left\{ \begin{array}{rcll} \frac{h^{\sigma}_{\tau,k+1} - h^{\sigma}_{\tau,k}}{h} + \partial_x j^{\sigma}_{\tau,k+1} & = & 0, & x\in \Omega \\ \partial_x h^{\sigma}_{\tau,k+1} = j^{\sigma}_{\tau,k+1} & = & 0, & x\in \partial\Omega \\ j^{\sigma}_{\tau,k+1} & = & m(h^{\sigma}_{\tau,k})|\partial_x^3 h^{\sigma}_{\tau,k+1} - G_{\sigma}''(h^{\sigma}_{\tau,k+1})\partial_x h^{\sigma}_{\tau,k+1}|^{\alpha-1}(\partial_x^3 h^{\sigma}_{\tau,k+1} - G_{\sigma}''(h^{\sigma}_{\tau,k+1})\partial_x h^{\sigma}_{\tau,k+1}), & x\in \Omega \end{array} \right.\]

Energy-dissipation inequality

Naively, one obtains the inequality

\[E^{\sigma}[h^{\sigma}_{\tau,k+1}] + \tau\tfrac{\alpha}{\alpha+1} \int_{\Omega} \frac{|j^{\sigma}_{\tau,k+1}|^{\frac{\alpha+1}{\alpha}}}{m(h^{\sigma}_{\tau,k})^{\frac{1}{\alpha}}} \, \mathrm{d} x = \mathcal{F}^{\sigma}_{\tau,h^{\sigma}_{\tau,k}}[h^{\sigma}_{\tau,k+1},j^{\sigma}_{\tau,k+1}] \leq \mathcal{F}^{\sigma}_{\tau,h^{\sigma}_{\tau,k}}[h^{\sigma}_{\tau,k},0] = E^{\sigma}[h^{\sigma}_{\tau,k}].\]

But we can do better using a trick due to de Giorgi. By convexity of both terms in the energy, we can improve this inequality to give

\[E^{\sigma}[h^{\sigma}_{\tau,k+1}] + \tau \int_{\Omega} \frac{|j^{\sigma}_{\tau,k+1}|^{\frac{\alpha+1}{\alpha}}}{m(h^{\sigma}_{\tau,k})^{\frac{1}{\alpha}}} \, \mathrm{d} x = \mathcal{F}^{\sigma}_{\tau,h^{\sigma}_{\tau,k}}[h^{\sigma}_{\tau,k+1},j^{\sigma}_{\tau,k+1}] \leq \mathcal{F}^{\sigma}_{\tau,h^{\sigma}_{\tau,k}}[h^{\sigma}_{\tau,k},0] = E^{\sigma}[h^{\sigma}_{\tau,k}].\]

Denote by

\[\begin{split} \hat{h}^{\sigma}(t) & = \text{linear interpolation between } h^{\sigma}_{\tau,k} \text{ and } h^{\sigma}_{\tau,k},\ t\in [\tau k,\tau(k+1)],\\ j^{\sigma}(t) & = j^{\sigma}_{\tau,k+1},\ t\in [\tau k,\tau(k+1)]. \end{split}\]

We are now in the position to send \(\tau \searrow 0\), since we can obtain the required a-priori estimates from the energy-dissipation inequality.

Theorem The limit $\tau \searrow 0$.

Let $h_0\in H^1(\Omega)$ with finite energy $E^{\sigma}[h_0]<+\infty$. Then every accumulation point $(h^{\sigma},j^{\sigma})$ satisfisies

\[\left\{ \begin{array}{rcll} \partial_ t h^{\sigma} + \partial_x j^{\sigma} & = & 0, & x\in \Omega \\ \partial_x h^{\sigma} = j^{\sigma} & = & 0, & x\in \partial\Omega \\ j^{\sigma} & = & m(h^{\sigma})|\partial_x^3 h^{\sigma} - G_{\sigma}''(h^{\sigma})\partial_x h^{\sigma}|^{\alpha-1}(\partial_x^3h^{\sigma}-G''_{\sigma}(h^{\sigma})\partial_x h^{\sigma}), & x\in \Omega\\ h^{\sigma}(0,x) & = h_0(x). \end{array} \right.\]

and

\[E^{\sigma}[h^{\sigma}](t_1) + \frac{\alpha}{\alpha+1}\int_{t_0}^{t_1}\int_{\Omega} \frac{|j^{\sigma}|^{\frac{\alpha+1}{\alpha}}}{m(h)^{\frac{1}{\alpha}}} \,\mathrm{d}x \,\mathrm{d}s + \frac{1}{\alpha+1} \int_{t_0}^{t_1}\int_{\Omega} m(h)|\partial_x^3h^{\sigma}-G''_{\sigma}(h^{\sigma})\partial_x h^{\sigma}|^{\alpha+1}\,\mathrm{d}x \,\mathrm{d}s = E^{\sigma}[h^{\sigma}](t_0). \tag{4}\]

Note that every solution actually satisfies the energy-dissipation equality for a.e. \(t_1>t_2\geq 0\).

In total, we have constructed positive weak solutions to a modified equation.

Results without modification

In a second step, we want to remove the modification \(G_{\sigma}\) from the equation. This can again be obtained by using a-priori estimates due to the energy-dissipation identity \((4)\). Then we can send \(\sigma\searrow 0\). This result is summarised in the following theorem.

Theorem The limit \(\sigma\searrow 0\).

Let \(h_0\in H^1(\Omega)\). Then there exists an accumulation point

\[h\in L_{\infty}\bigl([0,\infty);H^1(\Omega)\bigr)\cap C^{\frac{1}{5\alpha+3},\frac{1}{2}}_{\mathrm{loc}}([0,\infty)\times\bar\Omega)\]

with \(\partial_x^3 h\in L_{\alpha+1,\mathrm{loc}}(\{h>0\})\) and \(\partial_t h \in L_{\alpha+1}\bigl((0,\infty);(W^1_{\alpha+1}(\Omega))'\bigr)\) of \((h^{\sigma})_{\sigma}\) and \(h\) is a global non-negative weak solution to

\[\left\{ \begin{array}{rcll} \partial_t h + \partial_x\bigl( m(h)|\partial_x^3h|^{\alpha-1} \partial_x^3h \bigr) & = & 0, & t>0,x\in \Omega\\ \partial_x h = m(h)\partial_x^3 h & = & 0, & t>0,x\in \partial\Omega\\ h(0,x) & = & h_0(x), & x\in \Omega \end{array} \right.\]

that satisfy and energy-dissipation inequality of the form

\(\int_{\Omega}\tfrac{1}{2} |\partial_x h(t,x)|^2 \, \mathrm{d}x + \int_0^{t} \int_{\Omega} m(h) |\partial_x^3 h(s,x)|^{\alpha+1} \, \mathrm{d}x \,\mathrm{d}t \leq \int_{\Omega}\tfrac{1}{2} |\partial_x h_0(x)|^2 \, \mathrm{d}x.\) We may lose equality precisely when the solution becomes zero.

Strategy of proof

  • Obtain uniform Hölder continuity for \((h^{\sigma})_{\sigma}\) from Sobolev embedding in space and equation
  • Use a-priori bounds from energy-dissipation inequality
  • Identify limit of non-linear flux by localised version of Minty’s trick

In particular, the thin-film equation is a gradient-flow for positive film heights.