Asymptotic behaviour of
aggregation-diffusion equations with non-linear mobility

Meeting of the Spanish “Nonlocal PDEs and Applications” Network, Valencia

Author
Affiliation

David Gómez-Castro

Universidad Autónoma de Madrid - ICMAT

Published

October 2, 2024

Collaborators

José A. Carrillo and J.L. Vázquez

Alejandro Fernández-Jiménez

Aggregation-Diffusion

\[\begin{equation*} \label{eq:ADE} \tag{ADE} \frac{\partial \rho}{\partial t} = \diver \Big( \mob(\rho) \nabla ( U'(\rho) + V + W * \rho ) \Big). \end{equation*}\] The mobility \(\mob\) is usually linear, i.e., \(\mob(\rho) = \rho\).

Modelling

For the introductory part we will follow [Gómez-Castro, 2024].

Continuity equations

Let \(\rho\) be a density and \(\omega \subset \Rd\) any control volume, if \(\mathbf F\) is the out-going flux \[\begin{equation} \label{eq:conservation control volume} \frac{\diff }{\diff t} \int_\omega \rho \diff x =- \int_{\partial \omega} \mathbf F \cdot \mathbf n \diff S= -\int_{ \omega} \diver \mathbf F \diff x \end{equation}\] Hence we arrive at the \[\begin{equation} \label{eq:conservation law} \frac{\partial \rho}{\partial t} = -\diver \mathbf F. \end{equation}\]

Heat equation

For the transport of heat, we use Fourier’s law to model the flux \(\mathbf F = -D \nabla \rho\) and yields the heat equation \[\begin{equation*} \tag{HE} \label{eq:HE} \frac{\partial \rho}{\partial t} = D \Delta \rho . \end{equation*}\]

Porous Medium Equation

A generalisation of this problem which is used to the flow a gas through porous media:

  • Flux in terms of a velocity: \(\mathbf F = -\rho \mathbf v\);

  • Darcy’s law to relate the velocity with the pressure \(p\) as \(\mathbf v = - \frac{k} \mu \nabla p\);

  • general state equation \(p = \phi(\rho)\). For perfect gases \(\phi (\rho) = p_0 \rho^\gamma\).

This yields the so-called Porous Medium Equation \[\begin{equation*} \tag{PME} \label{eq:PME} \frac{\partial \rho}{\partial t} = \Delta \rho^m. \end{equation*}\]

Particle systems

We can also understand transport from the particle perspective. Consider a known velocity field \(\mathbf v(t,x)\). Consider \(N\) particles with positions \(X_i (t)\) of equal masses \(1/N\) moving in the velocity field \[\begin{equation*} % \tag{AE$_N$} \frac{\diff X_i}{\diff t} = \mathbf v(t, X_i(t)) , \qquad i = 1, \cdots, N. \end{equation*}\] Define the empirical distribution \[\begin{equation} \label{eq:empirical distribution} \mu_t^N = \sum_{j=1}^N \frac 1 N \delta_{X_j(t)} , \end{equation}\] where \(\delta_x\) is the Dirac delta at a point \(x\). This is a weak solution to the continuity equation \[\begin{equation*} \partial_t \mu^N + \diver (\mu^N \mathbf v ) = 0. \end{equation*}\]

Aggregation and Confinement

Consider the particles interacting non-locally by \[\begin{equation*} % \tag{AE$_N$} \frac{\diff X_i}{\diff t} = - \sum_{\substack{ j=1 \\ j \ne i}}^N \frac 1 N \nabla W (X_i - X_j) - \nabla V (X_i) , \qquad i = 1, \cdots, N. \end{equation*}\] Notice that \[ \sum_{\substack{ j=1 \\ j \ne i}}^N \frac 1 N \nabla W (x - X_j(t)) = \int_\Rd \nabla W (x-y) \diff \mu^N_t (y). \] The empirical measure \(\mu^N\) is a distributional solution of the Aggregation-Confinement Equation \[\begin{equation} \partial_t \mu = \diver (\mu \nabla ( W * \mu + V ) ). \end{equation}\] Linear diffusion can be added to the particle system by introducing noise and working with Stochastic ODEs and a mean-field approximation.


Special attention has been paid to the case \(V = 0\) which is simply called Aggregation Equation \[\begin{equation*} \tag{AE} \label{eq:AE} \frac{\partial \rho}{\partial t} = \diver (\rho \nabla W * \rho). \end{equation*}\] As a toy model we will also discuss the confinement problem \[\begin{equation*} \tag{CE} \label{eq:ADE U,W = 0} \frac{\partial \rho}{\partial t} = \diver(\rho \nabla V ). \end{equation*}\]

No-flux conditions

In a bounded domain \(\Omega\), we preserve mass if we set \(\mathbf F \cdot \mathbf n = 0\) on \(\partial \Omega\), i.e., \[\begin{equation*} \tag{BC} \label{eq:BC Omega} \rho \nabla ( U'(\rho) + V + W*\rho ) \cdot \mathbf n = 0 \qquad \text{ for } t > 0 \text{ and } x \in \partial \Omega \end{equation*}\] (and set \(\rho = 0\) in \(\Rd \setminus \Omega\) for the convolution).

Non-convolutional setting

It is sometimes preferable to discuss the generalisation \[\begin{equation*} \tag{ADE$^*$} \label{eq:ADE Omega} \begin{aligned} &\frac{\partial \rho}{\partial t} = \diver \Big( \rho \nabla ( U'(\rho) + V + \mathcal K \rho ) \Big), \\ &\quad \text{where } \mathcal K \rho (x) = \int_\Omega K(x,y) \rho(y) \diff y. \end{aligned} \end{equation*}\]

Special cases

Heat equation \(\quad \partial_t \rho = \Delta \rho_t\)

As explained above, the heat equation \(\eqref{eq:HE}\) is the most classical example in our family of equations. It corresponds to \(\eqref{eq:ADE}\) with the choices \(U = \rho \log \rho\) and \(V = W = 0\).
In \(\Rd\) It is well known that \(\eqref{eq:HE}\) admits the Gaussian solution \[ K_t (x) = (4 \pi t)^{-\frac d 2} \exp \left( - \frac{|x|^2}{4t} \right). \] All solutions (with suitable initial data) are precisely of the form \[ \rho_t = K_t * \rho_0, \] and it is not difficult to check that \[ \|\rho_t\|_{L^1} = M , \qquad \| \rho_t - M K_t \|_{L^1} \to 0 \quad{ as }\, t \to \infty. \] A good survey on the matter can be found in [Vázquez, 2017].


The fundamental solution \(K_t\) is of the so-called type \[\begin{equation*} \tag{SS} \label{eq:self-similar} K_t (x) = A(t) F\left( \frac x {\sigma(t)} \right). \end{equation*}\] In this expression, \(F\) is usually called self-similar profile and \(\sigma(t), A(t)\) the scaling parameters. Due to mass conservation the natural choice is \(A(t) = \sigma(t)^{-d}\).

Fokker-Planck \(\quad \partial_t u = \Delta u_t + \nabla \cdot (x u_t)\)

If \(\rho\) solve \(\eqref{eq:HE}\), the self-similar change of variable \[ \rho(x,t) = u(y,\tau) (1 + t)^{-\frac d 2}, \qquad 2\tau = \log(1 + t), % \qquad y = (1+t)^{-\frac 1 2} x, % \] leads to the so-called linear Fokker-Planck.

This equation corresponds to \[ U_1 (\rho) = \rho \log \rho, V = \frac{|x|^2}{2} \text{ and }W = 0. \]

Stationary state: \(\log u + \frac{|x|^2}{2} = -h\) the Gaussian profile \[\begin{equation} \label{eq:Gaussian} \widehat u(y) = AG(y), \qquad G(y) = \frac{1}{\sqrt{2\pi}} e^{-\frac{|y|^2}{2}}. \end{equation}\] The constant \(A\) is the unique value such that \(\int \rho_t = M\). In original variable, we recover \(M K_t\).

Porous medium equation \(\quad \frac{\partial \rho}{\partial t} = \Delta \rho^m\)

The range \(m \in (0,1)\) is sometimes called . The \(\eqref{eq:PME}\) corresponds, for \(m \ne 1\), to \[ U_m (\rho) = \frac{\rho^m}{m-1}, \qquad V = W = 0. \] \(\eqref{eq:PME}\) admits a self-similar solution \[\begin{equation*} \tag{ZKB} \label{eq:Barenblatt} \begin{aligned} &B_t = t^{-\alpha} (C - k|x|^2t^{-2\beta})^{\frac 1 {m-1}}_+, \end{aligned} \end{equation*}\] where \(\alpha = \tfrac{d}{d(m-1)+2},\, \beta = \tfrac{\alpha}{d},\, k = \tfrac{\alpha(m-1)}{d}\).

This solution is usually denoted ZKB after Zel’dovich and Kompaneets, and Barenblatt

(see [Vázquez, 2006b] and the references therein)

Keller-Segel model

The Keller-Segel [Keller & Segel, 1970] (or Patlak-Keller-Segel model [Patlak, 1953]) model for describes the motion of cells by chemotactical attraction by means of the coupled system \[\begin{equation*} \tag{KSE} \label{eq:Keller-Segel un-normalised} \begin{dcases} \frac{\partial \rho}{\partial t} = \Delta \rho - \diver (\rho \nabla V_t ) \\ -\Delta V_t= \rho_t . \end{dcases} \end{equation*}\] This model was first studied for \(d = 2\).

Some authors replace the second equation by a more general evolutionary problem \(\ee \frac{\partial V_t}{\partial t} - \Delta V_t + \alpha V_t = \rho_t\). We will discuss \(\eqref{eq:Keller-Segel un-normalised}\), which can be thought as the limit \(\ee, \alpha \to 0\) [Nanjundiah, 1973].


To write this model as \(\eqref{eq:ADE}\) notice that \(\Rd\) we can solve in Fourier transform \[ V_t = W_{\rm N} * \rho \qquad \text{where } W_{\mathrm N} = \mathrm{F}^{-1}[|\xi|^{-2}]. \], i.e., \[\begin{equation} \label{eq:Newtonian potential} W_{\mathrm N} (x) = \begin{dcases} -\frac{1}{2\pi} \log|x| & \text{if } d = 2, \\ \frac{1}{d(d-2)\omega_d} |x|^{2-d} & \text{if } d > 2. \end{dcases} \end{equation}\] Notice that \(-\Delta W_{\rm N} = \delta_0\). The so-called corresponds to \(\rm N = -W_{\rm N}\).

Eventually, we can write this problem as \(\eqref{eq:ADE}\) where \(U = \rho \log \rho\), \(V = 0\) and \(W = -W_{\rm N}\).


Sometimes it is convenient to express the problem in a single equation using operator notation: \[ \frac{\partial \rho}{\partial t} = \Delta \rho - \diver (\rho \nabla (-\Delta)^{-1} \rho). \] When written in this form, the authors usually allow for any initial mass \(\|\rho_0\|_{L^1}\).

To set unit mass let \(\chi = \|\rho_0\|_{L^1}\) and \(u(t,x) = \rho(t,x) / \chi\) we arrive at the equation \[\begin{equation*} \tag{KSE$_\chi$} \label{eq:Keller chi} \frac{\partial u}{\partial t} = \Delta u - \chi \diver (u \nabla (-\Delta)^{-1} u). \end{equation*}\]

Some authors replace \(\Delta \rho\) by \(\Delta \rho^m\): [Calvez & Carrillo, 2006], [Bedrossian, Rodríguez & Bertozzi, 2011],[Luckhaus & Sugiyama, 2007],[Sugiyama & Kunii, 2006],[Sugiyama, 2007a],[Sugiyama, 2007b].

Chapman-Rubinstein-Schatzman

In the context of superconductivity [Chapman, Rubinstein & Schatzman, 1996] introduced a mean-field model for “the motion of rectilinear vortices in the mixed state of a type-II semiconductor”. Ginzburg-Landau with parameter \(\kappa \to +\infty\) and the number of vortices becomes large \[\begin{equation*} \tag{CRSE} \label{eq:CRSE} \begin{dcases} \frac{\partial \rho}{\partial t} = \diver( |\rho| \nabla V_t ) \\ -\Delta V_t + V_t = \rho_t . \end{dcases} \end{equation*}\]

As long as \(\rho \ge 0\), this model corresponds to \(\eqref{eq:ADE}\) where \(U, V = 0\).

In \(\Rd\) we can solve as before to recover \(V_t = W * \rho_t\) where \(W = \mathrm{F^{-1}}[ (1 + |\xi|^2)^{-1} ]\).

There is extensive literature on this problem. [Huang & Svobodny, 1998] [Schätzle & Styles, 1999] [Ambrosio & Serfaty, 2008] [Ambrosio, Mainini & Serfaty, 2011]

This sparked broader interest in the family of aggregation equations \(\eqref{eq:AE}\).

Newtonian vortex problem

In [Lin & Zhang, 1999] the authors justify that a different limit from Gizburg-Landau equations leads to the problem without zero-order term \[\begin{equation*} \tag{NVE} \label{eq:Newtonian vortex} \begin{dcases} \frac{\partial \rho}{\partial t} = \diver (\rho \nabla V_t ) \\ -\Delta V_t = \rho. \end{dcases} \end{equation*}\] Again, this can be set in \(\Rd\) or in a bounded domain (with suitable boundary conditions).

This problem also attracted several authors, e.g.,
[Masmoudi & Zhang, 2005],[Bertozzi, Laurent & Léger, 2012].

The Caffarelli-Vázquez problem

In [Caffarelli & Vazquez, 2011], Caffarelli and Vázquez introduced a non-local porous medium-type equation given by \[\begin{equation*} \begin{dcases} \frac{\partial \rho}{\partial t} = \diver (\rho \nabla V_t ) \\ (-\Delta)^s V_t = \rho , \end{dcases} \end{equation*}\] where \((-\Delta)^s\) denotes the fractional Laplacian.

In \(\Rd\) this operator is to the operational fractional power of the Laplacian, the operator with Fourier symbol \(|\xi|^{2s}\). \[\begin{equation*} \label{eq:Caffarelli-Vazquez} \tag{CVE} \frac{\partial \rho}{\partial t} = \diver (\rho \nabla (-\Delta)^{-s} \rho ). \end{equation*}\] To solve the fractional Laplace equation we take the Riesz potential \(W = \mathrm F^{-1}[|\xi|^{-2s}]\).

There has been a lot of analysis for this equation [Caffarelli & Vázquez, 2011], [Biler, Imbert & Karch, 2015], [Caffarelli, Soria & Vázquez, 2013],[Caffarelli & Vázquez, 2015], [Serfaty & Vázquez, 2014], [Lisini, Mainini & Segatti, 2018].


Caffarelli and Vázquez proposed the so-called Porous Medium Equation with non-local pressure \[\begin{equation*} \label{eq:PME non-local pressure} \tag{PMEn} \frac{\partial \rho}{\partial t} = \diver (\rho^{m-1} \nabla (-\Delta)^{-s} \rho ). \end{equation*}\]

More examples

The power-type family of Aggregation-Diffusion Equations

Many authors have devoted their attention to the power-type family of non-linearities given by \[ \begin{gathered} U = \frac{\rho^m}{m-1}, \qquad \qquad V_\lambda (x) = \begin{dcases} \frac{|x|^\lambda}{\lambda} & \text{if } \lambda \ne 0 \\ \log|x| & \text{if } \lambda = 0 , \end{dcases}, \\ W_k (x) = \begin{dcases} \frac{|x|^k}{k} & \text{if } k \in (-d,d) \setminus \{0\} \\ \log|x| & \text{if } k = 0 , \end{dcases} \end{gathered} \]

Asymptotic behaviour: the famous cases

  • Long-time asymptotic behaviour: \(\rho_t \to \widehat \rho\) as \(t \to \infty\)

  • Blow-up: \(\rho_t \rightharpoonup \widehat \mu \not \in L^1(\Omega)\) as \(t \nearrow T \in (0,\infty]\).

    • We say there is concentration if \(\mu\) contains a Dirac
  • Extinction: \(\rho_t \searrow 0\) as \(t \nearrow T \in (0,\infty)\)

    • If, up to a rescaling \(\frac{\| \rho_t - U_t \|_{L^p}}{\|\rho_t\|_{L^p}}\) we say there are intermediate asymptotics

Formal steady states

Anything of form \[ \rho \nabla (U'(\rho) + V + W*\rho) = 0. \]

This means that if \(\rho\) is continuous, in any connected component \(\omega\) of its support \[ U'(\rho) + V + W * \rho = C_\omega. \]

The constant can naturally differ between components.

Intermediate asymptotics

For \(\eqref{eq:HE}\) and \(\eqref{eq:PME}\) we know that solutions approximate the self-similar solution “faster” than they vanish, for example \[\begin{equation} \label{eq:intermediate asymptotics} \frac{\| \rho_t - B_t \|_{L^p}}{\| \rho_t \|_{L^p}} \to 0, \qquad \text{as } t \to \infty. \end{equation}\] See [Vázquez, 2006b], [Vázquez, 2006a].

For \(\eqref{eq:HE}\) see [Vázquez, 2017].

\(\eqref{eq:ADE U,W = 0}\): blow-up

Going back to the simplest example \(\eqref{eq:ADE U,W = 0}\) we can look at the examples \(V = \frac{|x|^\alpha}{\alpha}\) with \(\alpha > 0\). Then \(v = -\nabla V = - |x|^{\alpha-2} x\) and \(\rho_t = (X_t)_\sharp \rho_0\) with corresponding characteristic field is obtained by solving \[ \frac{\partial X_t}{\partial t} = -|X_t|^{\alpha-2} X_t, \] with \(X_t(0) = y\). We conclude that \[ X_t (y) = \begin{dcases} y e^{-t} & \text{if }\alpha=2, \\ ( |y|^{2-\alpha} - (2-\alpha) t )^{\frac 1 {2-\alpha}} \frac{y}{|y|} & \text{if } \alpha \ne 2. \end{dcases} \] For \(\alpha < 2\) the characteristics arrive at \(0\) at finite time and hence we get a Dirac delta at \(0\) at a certain finite time for any \(\rho_0 \not\equiv 0\). This is not problematic in the distributional sense in this case. However, this behaviour happens also for other problems.

Keller-Segel problem: conditional blow-up

Some authors noticed by direct techniques that the Keller-Segel model has finite-time blow-up for initial data with large enough mass:

  • [Jäger & Luckhaus, 1992] showed for \(\eqref{eq:Keller-Segel un-normalised}\) in a bounded domain with no-flux the existence of a critical mass \(M^*\) such that,
    if \(\| \rho_0 \|_{L^1} > M^*\), then \(\rho(T^*)\) contains a Dirac delta.

  • Later this result was obtained also in \(\Rd\) in [Herrero & Velázquez, 1996] using radial arguments, where the authors characterise the critical mass is \(M^* = 8 \pi\).
    In more generality in [Dolbeault & Perthame, 2004].

  • Conversely, for \(\| \rho_0 \|_{L^1} < M^*\) global existence for \(\eqref{eq:Keller-Segel un-normalised}\) is known.
    In \(d = 2\) it was shown in [Blanchet, Dolbeault & Perthame, 2006].


A nice proof through the second-order moment.

For \(\eqref{eq:Keller-Segel un-normalised}\) in \(d = 2\) it was first noticed by [Dolbeault & Perthame, 2004] that, working with solutions of initial mass \(\int_\Rd \rho_0 = M\), \[ \frac{\diff}{\diff t} \int_\Rd |x|^2 \rho(t,x) \diff x = 4 M \Big( 1- \frac{M}{8\pi }\Big). \]

Hence, if \(M > {8 \pi}\) necessarily the second-order moment becomes negative in finite time. This is incompatible with our non-negative solutions. There is complete concentration to a Dirac delta.

This was later extended by [Blanchet, Carrillo & Laurençot, 2009] to \(d > 2\), where the evolution is more involved.

Free energy and gradient-flow

Free energy: a Lyapunov functional

\(\eqref{eq:ADE Omega}\) is related to the free energy \[\begin{equation*} \tag{FE$^*$} \label{eq:free energy star} \mathcal F(\rho) =\int_\Omega U(\rho) + \int_\Omega V \rho + \frac 1 2 \iint_{\Omega \times \Omega } K(x,y) \rho(x) \rho(y) \diff x \diff y . \end{equation*}\] For convenience, let us define the first variation \[\begin{equation} \label{eq:first variation} \frac{\delta \mathcal F}{\delta \rho} [\rho] \defeq U'(\rho) + V + \int_\Omega K(\cdot,y)\rho(y) \diff y. \end{equation}\]

We write \(\eqref{eq:ADE Omega}\) as \[\begin{equation*} \partial_t \rho = \diver\left( \mob(\rho) \nabla \frac{\delta \mathcal F}{\delta \rho} [\rho]\right) \end{equation*}\]

Then, \[\begin{equation*} \frac{d}{dt} \mathcal F[\rho_t] = \int \frac{\delta \mathcal F}{\delta \rho} [\rho] \partial_t \rho_t = - \int \mob(\rho_t) \left|\frac{\delta \mathcal F}{\delta \rho} [\rho] \right|^2. \end{equation*}\]

Gradient-flow structure

When \(\mob(\rho) = \rho\) there a nice gradient-flow theory:

  • Otto calculus (see, e.g., [Ambrosio, Brué & Semola, 2021]) allows us to show that \(\eqref{eq:ADE}\) in \(\Rd\) is formally the \(2\)-Wasserstein gradient flow in the sense that \[ \frac{\partial \rho}{\partial t} = - \nabla_{\Wass_2} \mathcal F[\rho_t] \]

  • There is a good notion of convexity (McCann’s displacement convexity). Under certain condition we have \[ \Wass_2(\mu_t, \widehat \mu_t) \le e^{-\lambda t} \Wass_2(\mu_0, \widehat \mu_0). \]

For general \(\mob(\rho)\), there is a generalisation [Dolbeault, Nazaret & Savaré, 2009]

  • This is only a distance if \(\mob\) is concave.

  • It is (almost) never displacement convex in our settings.

Euler-Lagrange condition: non-negative functions of given mass

If \(\calF(\widehat \rho) \le \calF(\rho)\) for all \(\rho \ge 0\) such that \(\int \rho = \int \widehat \rho\) we have that, for some \(C \in \mathbb R\) \[\begin{equation*} \begin{dcases} \frac{\delta \calF}{\delta\rho} [\widehat \rho] = C & \text{if } \rho > 0, \\ \frac{\delta \calF}{\delta\rho} [\widehat \rho] \ge C & \text{if } \rho = 0. \\ \end{dcases} \end{equation*}\]

For \(\eqref{eq:ADE}\) we have \(\frac{\delta \calF}{\delta\rho} [\rho] = U'(\rho) + V + W*\rho\).

The free-energy minimiser is the steady state with same constant in all connected components.

\[ \widehat \rho(x) = \max\Bigg\{0, (U')^{-1} \big(C-V(x) - W*\widehat \rho\big)\Bigg\}. \]

Notice that when \(W = 0\) we the solution is parametrised by \(C\).

Many steady states, only one free-energy minimiser

An example with \(W = 0\)

You don’t always converge to the global minimizer

An example with \(W = 0\)

Case 1: Newtonian vortex with power-type non-linear mobility

\[\begin{equation} \tag{NVE} \begin{dcases} \frac{\partial \rho}{\partial t} = \diver (\rho^\alpha \nabla V_t ) \\ -\Delta V_t = \rho. \end{dcases} \end{equation}\]

[Carrillo, Gómez-Castro & Vázquez, 2022a],[Carrillo, Gómez-Castro & Vázquez, 2022c]

Self-similar analysis

We can look for solutions of the form \[ U(t,x)= t^{-\gamma} F(|x|\,t^{-\beta}). \]

Plugging this in the equation we recover the self-similar solution of mass 1 for \(\alpha \in (0,1)\) \[ \scriptsize U(t,x) = t^{-\frac 1 \alpha}\left( \alpha + \left( \frac{ \omega_d |x|^dt^{-\frac 1 {\alpha}} } { \alpha} \right)^{\frac {\alpha} {1-\alpha }} \right)^{-1/\alpha }. \]

The case \(\alpha = 1\) was already studied in [Serfaty & Vázquez, 2014].

The same algebra works for \(\alpha > 1\) but gives no finite-mass solutions

Mass variable

Let us define \[ m(t,x) = \int_{-\infty}^x \rho(t,y) \diff y \]

Integrating the equation for \(\rho\) in \(x\) \[ \frac{\partial m}{\partial t} = \rho^\alpha \frac{\partial V_t}{\partial x}. \]

The equation for \(V_t\) is \(-\frac{\partial^2 V}{\partial x^2} = \rho = \frac{\partial m}{\partial x}\). Setting \(\frac{\partial V}{\partial x} (-\infty) = 0\) we get \(-\frac{\partial V}{\partial x} = m.\), so \[ \frac{\partial m}{\partial t} = -\left( \frac{\partial m}{\partial x} \right)^\alpha m \]

If \(d > 1\) and \(\rho\) is radially symmetric, we get to the same equation for \[ m(t,v) = \int_{A_v} \rho(t,x) \diff x , \quad A_v = B(0,r) \text{ such that }|A_v| = v. \]


Due to the conservation of mass, we can write the boundary value problem \[\begin{equation*} \begin{dcases} \frac{\partial m}{\partial t} + \left( \frac{\partial m}{\partial v} \right)^\alpha m = 0 \\ m(0,v) = m_0(v), \\ m(t,0) = 0, \\ m(t,\infty) = m_0(\infty). \end{dcases} \end{equation*}\]

This problem admits solutions by generalised characteristics, and the natural is viscosity solutions. We develop a well-posedness theory.

Time asymptotics for \(\alpha \in (0,1)\)

By using generalised characteristics in the equation for the mass, we can prove

Let \(\alpha \in (0,1)\) and \(\rho_0\) is radially non-increasing. Then, we have that \[ \sup_{y \in \mathbb R^d} \left| \frac{ \rho(t,y) - F_M \left ( {|y|} \right ) }{F_M \left ( {|y|} \right )}\right| \longrightarrow 0 , \] as \(t \to +\infty\).

Time asymptotics for \(m\)

Assume that \(0 \le \rho_0 \in L^\infty_c(\Rd)\) is radially symmetric, and let \(M = \| \rho_0 \|_{L^1}\).

Let \(\alpha \in (0,1)\) and denote \[ G_M(\kappa)=\int \limits _{\omega_d |y|^d \le \kappa } F_M (|y|) \diff y \] the mass function of the selfsimilar solution with total mass \(M\). Then, \[\begin{equation} \label{eq:asymp mass} \sup_{\kappa \ge \ee } \left| \dfrac{ m (t, t^{\frac 1 \alpha} \kappa ) } {G_M (\kappa)} - 1 \right| \longrightarrow 0 , \end{equation}\] as \(t \to \infty\) for any \(\ee > 0\).

Let \(\alpha > 1\) and define \[\begin{equation} G( y ) = \begin{dcases} y & y\le 1 \\ 1 & y > 1. \end{dcases} \end{equation}\] Then, \[\begin{equation} \sup_ { y \ge \varepsilon } \left| \frac{m\left( t , M (\alpha t)^{\frac 1 \alpha} y \right) }{M G (y)} - 1\right| \to 0, \end{equation}\] as \(t \to +\infty\) for any \(\varepsilon > 0\).

Numerical results

Waiting time for \(\alpha > 1\)

We call “waiting time” to the time it takes the support to start evolving.

A subsolution with waiting time

\[\begin{equation} \label{eq:Ansatz} \scriptsize \underline m(t, v) = \begin{dcases} \left( M^{\frac{\alpha}{\alpha - 1}} - \alpha^{\frac 1 {\alpha - 1}} \frac{(c_0 - v)_+^{\frac{\alpha}{\alpha - 1}}}{(T-t)^{\frac{1}{\alpha - 1}}} \right)_+^{\frac{\alpha - 1}{\alpha}}, & \text{if } v < 1, \\ M & \text{if } v > 1 \end{dcases} \end{equation}\]

Let \(\alpha > 1\), \(m_0 \in BUC ( [0,+\infty) )\), and let \(c_0= \max \mathrm{supp} \rho_0\).

There is waiting time if and only if \[\begin{equation} \limsup_{v \to c_0^-} \frac{M - m_0(v)}{(c_0 - v)^{\frac{\alpha}{\alpha-1}} } < +\infty, \end{equation}\]

Case 2: Asymptotic concentration with fast diffusion

[Carrillo, Gómez-Castro & Vázquez, 2022b] [Carrillo, Fernández-Jiménez & Gómez-Castro, 2024]

Fast Diffusion

[Carrillo, Gómez-Castro & Vázquez, 2022b] case \(W = 0\).

The Euler-Lagrange equation when \(U'(\rho) = \frac{\rho^{m}}{m-1}\) for \(0 < m < 1\) becomes \[ \widehat \rho(x; C) = \left( \tfrac{1-m}{m} (C + V(x)) \right)^{-\frac 1 {1-m}}. \]

Notice that \(\mathfrak M (C) = \int \rho(x; C) dx\) decreases with \(C\).

We consider \(m \in (0,1)\), \(V\) radially increasing, and \(\rho_0\) radially symmetric.

W.l.o.g. \(V(0) = 0\) so \(C \ge 0\).

\[ \rho_t \longrightarrow \begin{dcases} \widehat \rho(x; C) & \text{if } M = \mathfrak M(C), \\ \widehat \rho (x; 0) + (\mathfrak M(0) - M) \delta_0 & \text{if } M > \mathfrak M(0) . \end{dcases} \]

In [Carrillo, Fernández-Jiménez & Gómez-Castro, 2024] we deal with \(W \ne 0\).

Sketch of proof of [Carrillo, Gómez-Castro & Vázquez, 2022b]

For technical convenience, we work in a ball \(B_R\), \(\nabla u \cdot n = \nabla V \cdot n = 0\) on \(\partial B_R\).

  • Analysis of the mass function for radial solution \(M(t,r) = \int_{B_r} \rho(t,x) dx\).

Taking volume variable \(v = \omega_n r^d\), \(M\) satisfies \[\begin{equation} \tag{$\star$} \label{eq:mass} \frac{\partial M}{\partial t} = (n \omega_n^{\frac 1 n } v ^{\frac{n-1} n})^2 \left\{ \frac{\partial }{\partial v} \left[ \left( \frac{\partial M }{\partial v} \right)^m \right] + \frac{\partial M }{\partial v} \frac{\partial V}{\partial v} \right\} \end{equation}\] We develop a theory of viscosity solutions with comparison principle.

  • To check the limit \(t \to \infty\) we consider \(M_n (t,v) = M( t - n, v)\) as \([0,1] \times B_R\) functions.

  • By uniform continuity away from \(0\) (see [DiBenedetto, 1993]), \(M_{n_k} \to M_\infty\) uniformly over compacts of \([0,1] \times (0,R]\).

  • By stability of viscosity solutions \(M_\infty\) is a solution of \(\eqref{eq:mass}\).

  • To check that \(M_\infty = M_\infty (x)\), use free energy dissipation \[ \int_{t}^{t+1} \left\| \frac{\partial \rho}{\partial t} \right\|_{W^{-1,1}}^2 \le \int_{t}^{t+1} \left\| \rho \nabla (U'(\rho) + V ) \right\|_{L^1}^2 \le C (\calF[\rho_{t}] - \calF[\rho_{t+1}]). \]

  • So \(M_\infty\) is a viscosity solution of \(\frac{\partial }{\partial v} \left[ \left( \frac{\partial M }{\partial v} \right)^m \right] + \frac{\partial M }{\partial v} \frac{\partial V}{\partial v} = 0\)

  • So \(M_\infty (r) = M_\infty(0^+) + \int_{B_r} \widehat \rho (x, C) dx\).

  • Build special subsolution to characterise the constant.

Case 3: Asymptotic simplification with linear diffusion

[Carrillo, Gómez-Castro, Yao & Zeng, 2023]

Asymptotic simplication for linear diffusion

We start by a classical observation.

If \(U(\rho) = \rho \log \rho\), \(V = 0\) and \(W \in L^\infty (\Rd)\), then there are no continuous steady states of finite positive mass.

If \(W \in L^\infty (\Rd)\) and \(\rho \in L^1(\Rd)\), then \(W * \rho \in L^\infty (\Rd)\).

The Euler-Lagrange equation is \(\log \rho + W * \rho = c\), so

\[ \rho = e^{c} e^{W*\rho } \ge e^{c} e^{ - \| W * \rho \|_{L^\infty}} > 0. \]

Therefore, in this range we always expect diffusion.

In the case of linear diffusion \[\begin{equation*} \partial_t \rho = \Delta \rho + \diver (\rho \nabla W * \rho ). \end{equation*}\]

[Cañizo, Carrillo & Schonbek, 2012]: for small \(W\):

\[\begin{equation} \label{eq:asymptotic simplification} \tag{$\star$} \| \rho(t, \cdot) - K(t, \cdot) \|_{L^1} \to 0 \qquad \text{as } t \to \infty. \end{equation}\] where \(K\) is the heat kernel.

Let \(n\ge 2\), and assume \(W(x) = W(-x)\)

  • \(W \in \mathcal W^{1,\infty} (\Rd)\)

  • \(\nabla W \in L^{n-\ee} (\Rd)\)

  • \(\Delta W \in L^{\frac n 2} (\Rd)\) (and also \(\Delta W \in L^{\frac n 2 - \ee} (\Rd)\) if \(n\geq 3\))

Then \(\eqref{eq:asymptotic simplification}\).

Notice that this hypothesis work for \(W(x) \sim |x|^{-\varepsilon}\) for any \(\ee > 0\),

but not for the critical case \(W(x) \sim \log |x|\).

Sketch of proof

First, we prove well-posedness by Duhamel’s formula and that, in rescaled variable

  • If \(\nabla W \in L^n (\Rd)\) then \(\sup_{\tau \ge 1} \| \widetilde \rho(\tau, \cdot) \|_{H^1 } < \infty.\)

  • If \(n \ge 2\), \(\nabla W \in L^n (\Rd)\) and \(\Delta W \in L^{\frac n2} (\Rd)\) then \(\sup_{\tau \ge 1} \| \widetilde \rho(\tau, \cdot) \|_{C^\alpha } < \infty\) (modulus of continuity arguments, e.g. [Kiselev, Nazarov & Volberg, 2007])

Using a smart change in variable \[\mathcal F [\rho(t)] \le -\frac{n}{2}\ln t + C(n, \|W\|_{L^\infty}).\]

Thus, we prove that \(\int_{\Rd} \widetilde \rho (\tau, y) |\log \widetilde \rho(\tau, y) | \diff y \le C .\)

We study the \(L^1\) relative entropy \(E_1 (\widetilde \rho \| G) = \int_{\Rd} \widetilde \rho \log \frac{\widetilde \rho}{G} \diff y.\)

Like for \(\eqref{eq:HE}\), we can use logarithmic Sobolev inequality to recover an Ordinary Differential Inequality for \(E_1\) (where the terms from \(W\) are a controlled errors)

Lastly, we apply the Csiszar-Kullback inequality \[\begin{equation*} { \| \rho(t,\cdot) - K(t,\cdot) \|_{L^1} = \| \widetilde \rho(t, \cdot) - G (\cdot) \|_{L^1} } { \le 2 \sqrt{ E_1 (\widetilde \rho \| G)} } { \to 0, \qquad \text{as } t \to \infty. } \end{equation*}\]

Case 4: Saturation and free boundaries

We consider \(\mob(0) = \mob(\beta) = 0\).

Free boundaries for two level sets:

Assume \(\mob(0) = \mob(\beta) = 0\).

Numerical results obtained in [Bailo, Carrillo & Hu, 2023] suggested the formation of free boundaries at levels \(0, \beta\).

Furthermore, their numerical results suggest that the steady states are the doubly-truncated Barenblatts \[ \widehat \rho = \min \Bigg\{ \beta, \max\Big\{ 0, (U')^{-1}(C - V - W*\widehat \rho) \Big\} \Bigg \} \]


In an upcoming paper with Carrillo and Fernández-Jiménez we justify this behaviour.

Keep an eye on arxiv.

Questions?

Bibliography

Ambrosio, L., Brué, E. & Semola, D. (2021) Lectures on optimal transport. Springer International Publishing. doi:10.1007/978-3-030-72162-6.
Ambrosio, L., Mainini, E. & Serfaty, S. (2011) Gradient flow of the Chapman-Rubinstein-Schatzman model for signed vortices. Annales de l’I.H.P. Analyse non linéaire. 28 (2), 217–246. doi:10.1016/j.anihpc.2010.11.006.
Ambrosio, L. & Serfaty, S. (2008) A gradient flow approach to an evolution problem arising in superconductivity. Communications on Pure and Applied Mathematics. 61 (11), 1495–1539. doi:10.1002/cpa.20223.
Bailo, R., Carrillo, J.A. & Hu, J. (2023) Bound-Preserving Finite-Volume Schemes for Systems of Continuity Equations with Saturation. SIAM Journal on Applied Mathematics. 83 (3), 1315–1339. doi:10.1137/22M1488703.
Bedrossian, J., Rodríguez, N. & Bertozzi, A.L. (2011) Local and global well-posedness for aggregation equations and Patlak-Keller-Segel models with degenerate diffusion. Nonlinearity. 24 (6), 1683–1714. doi:10.1088/0951-7715/24/6/001.
Bertozzi, A.L., Laurent, T. & Léger, F. (2012) Aggregation via Newtonian potential and aggregation patches. Mathematical Models and Methods in Applied Sciences. 22.
Biler, P., Imbert, C. & Karch, G. (2015) The Nonlocal Porous Medium Equation: Barenblatt Profiles and Other Weak Solutions. Archive for Rational Mechanics and Analysis. 215 (2), 497–529. doi:10.1007/s00205-014-0786-1.
Blanchet, A., Carrillo, J.A. & Laurençot, P. (2009) Critical mass for a patlak-keller-segel model with degenerate diffusion in higher dimensions. Calculus of Variations and Partial Differential Equations. 35 (2), 133–168. doi:10.1007/s00526-008-0200-7.
Blanchet, A., Dolbeault, J. & Perthame, B. (2006) Two-dimensional Keller-Segel model: Optimal critical mass and qualitative properties of the solutions. Electron. J. Differential Equations. No. 44, 32.
Caffarelli, L.A., Soria, F. & Vázquez, J.L. (2013) Regularity of solutions of the fractional porous medium flow. J. Eur. Math. Soc. (JEMS). 15 (5), 1701–1746. doi:10.4171/JEMS/401.
Caffarelli, L.A. & Vazquez, J.L. (2011) Nonlinear porous medium flow with fractional potential pressure. Arch. Ration. Mech. Anal. 202 (2), 537–565. doi:10.1007/s00205-011-0420-4.
Caffarelli, L.A. & Vázquez, J.L. (2011) Asymptotic behaviour of a porous medium equation with fractional diffusion. Discrete Contin. Dyn. Syst. 29 (4), 1393–1404. doi:10.3934/dcds.2011.29.1393.
Caffarelli, L.A. & Vázquez, J.L. (2015) Regularity of solutions of the fractional porous medium flow with exponent 1/2. Algebra i Analiz. 27 (3), 125–156. doi:10.1090/spmj/1397.
Calvez, V. & Carrillo, J.A. (2006) Volume effects in the Keller-Segel model: Energy estimates preventing blow-up. J. des Math. Pures Appl. 86 (2), 155–175. doi:10.1016/j.matpur.2006.04.002.
Cañizo, J.A., Carrillo, J.A. & Schonbek, M.E. (2012) Decay rates for a class of diffusive-dominated interaction equations. Journal of Mathematical Analysis and Applications. 389 (1), 541–557. doi:10.1016/j.jmaa.2011.12.006.
Carrillo, J.A., Fernández-Jiménez, A. & Gómez-Castro, D. (2024) Partial mass concentration for fast-diffusions with non-local aggregation terms. Journal of Differential Equations. 409, 700–773. doi:10.1016/j.jde.2024.08.013.
Carrillo, J.A., Gómez-Castro, D. & Vázquez, J.L. (2022a) A fast regularisation of a Newtonian vortex equation. Annales de l’Institut Henri Poincaré C, Analyse non linéaire. http://arxiv.org/abs/1912.00912.
Carrillo, J.A., Gómez-Castro, D. & Vázquez, J.L. (2022b) Infinite-time concentration in aggregation–diffusion equations with a given potential. J. Math. Pures Appl. 157, 346–398. doi:10.1016/j.matpur.2021.11.002.
Carrillo, J.A., Gómez-Castro, D. & Vázquez, J.L. (2022c) Vortex formation for a non-local interaction model with Newtonian repulsion and superlinear mobility. Advances in Nonlinear Analysis. 11 (1), 937–967. doi:10.1515/anona-2021-0231.
Carrillo, J.A., Gómez-Castro, D., Yao, Y. & Zeng, C. (2023) Asymptotic simplification of Aggregation-Diffusion equations towards the heat kernel. Archive for Rational Mechanics and Analysis. 247. doi:10.1007/s00205-022-01838-5.
Chapman, S.J., Rubinstein, J. & Schatzman, M. (1996) A mean-field model of superconducting vortices. European Journal of Applied Mathematics. 7 (2), 97–111.
DiBenedetto, E. (1993) Degenerate parabolic equations. Universitext. New York, NY, Springer New York. doi:10.1007/978-1-4612-0895-2.
Dolbeault, J., Nazaret, B. & Savaré, G. (2009) A new class of transport distances between measures. Calc. Var. Partial Differ. Equ. 34 (2), 193–231. doi:10.1007/s00526-008-0182-5.
Dolbeault, J. & Perthame, B. (2004) Optimal critical mass in the two dimensional keller-segel model in. Comptes Rendus Mathematique. 339 (9), 611–616. doi:10.1016/j.crma.2004.08.011.
Fernández-Real, X. & Figalli, A. (2022) The continuous formulation of shallow neural networks as wasserstein-type gradient flows. In: Analysis at Large: Dedicated to the Life and Work of Jean Bourgain. Springer. pp. 29–57.
Gómez-Castro, D. (2024) Beginner’s guide to aggregation-diffusion equations. SeMA Journal. doi:10.1007/s40324-024-00350-y.
Herrero, M.A. & Velázquez, J.J.L. (1996) Chemotactic collapse for the Keller-Segel model. J. Math. Biol. 35 (2), 177–194. doi:10.1007/s002850050049.
Huang, C. & Svobodny, T. (1998) Evolution of Mixed-State Regions in Type-II Superconductors. SIAM Journal on Mathematical Analysis. 29 (4), 1002–1021. doi:10.1137/S003614109731504X.
Jäger, W. & Luckhaus, S. (1992) On explosions of solutions to a system of partial differential equations modelling chemotaxis. Transactions of the American Mathematical Society. 329 (2), 819–824. doi:10.1090/s0002-9947-1992-1046835-6.
Keller, E.F. & Segel, L.A. (1970) Initiation of slime mold aggregation viewed as an instability. Journal of Theoretical Biology. 26 (3), 399–415. doi:10.1016/0022-5193(70)90092-5.
Kiselev, A., Nazarov, F. & Volberg, A. (2007) Global well-posedness for the critical 2D dissipative quasi-geostrophic equation. Inventiones Mathematicae. 167 (3), 445–453. doi:10.1007/s00222-006-0020-3.
Lin, F. & Zhang, P. (1999) On the hydrodynamic limit of Ginzburg-Landau vortices. Discrete and Continuous Dynamical Systems. 6 (1), 121–142. doi:10.3934/dcds.2000.6.121.
Lisini, S., Mainini, E. & Segatti, A. (2018) A Gradient Flow Approach to the Porous Medium Equation with Fractional Pressure. Archive for Rational Mechanics and Analysis. 227 (2), 567–606. doi:10.1007/s00205-017-1168-2.
Luckhaus, S. & Sugiyama, Y. (2007) Asymptotic Profile with the Optimal Convergence Rate for A Parabolic Equation of Chemotaxis in Super-critical Cases. Indiana University Mathematics Journal. 56 (3), 1279–1297. https://www.jstor.org/stable/24902731.
Masmoudi, N. & Zhang, P. (2005) Global solutions to vortex density equations arising from sup-conductivity. Annales de l’IHP Analyse non linéaire. 22 (4), 441–458. doi:10.1016/j.anihpc.2004.07.002.
Nanjundiah, V. (1973) Chemotaxis, signal relaying and aggregation morphology. Journal of Theoretical Biology. 42 (1), 63–105.
Patlak, C.S. (1953) Random walk with persistence and external bias. The Bulletin of Mathematical Biophysics. 15 (3), 311–338. doi:10.1007/BF02476407.
Schätzle, R. & Styles, V. (1999) Analysis of a mean field model of superconducting vortices. European Journal of Applied Mathematics. 10 (4), 319–352.
Serfaty, S. & Vázquez, J.L. (2014) A mean field equation as limit of nonlinear diffusions with fractional Laplacian operators. Calc. Var. Partial Differ. Equ. 49 (3-4), 1091–1120. doi:10.1007/s00526-013-0613-9.
Sugiyama, Y. (2007a) Application of the best constant of the Sobolev inequality to degenerate Keller-Segel models. Advances in Differential Equations. 12 (2). doi:10.57262/ade/1355867472.
Sugiyama, Y. (2007b) Time global existence and asymptotic behavior of solutions to degenerate quasi-linear parabolic systems of chemotaxis. Differential and Integral Equations. 20 (2). doi:10.57262/die/1356039511.
Sugiyama, Y. & Kunii, H. (2006) Global existence and decay properties for a degenerate KellerSegel model with a power factor in drift term. Journal of Differential Equations. 227 (1), 333–364. doi:10.1016/j.jde.2006.03.003.
Vázquez, J.L. (2017) Asymptotic behaviour methods for the Heat Equation. Convergence to the Gaussian. 2017. arXiv. https://arxiv.org/abs/1706.10034.
Vázquez, J.L. (2006a) Smoothing and decay estimates for nonlinear diffusion equations. Oxford University Press. doi:10.1093/acprof:oso/9780199202973.001.0001.
Vázquez, J.L. (2006b) The Porous Medium Equation. Oxford, United Kingdom, Oxford University Press. doi:10.1093/acprof:oso/9780198569039.001.0001.