Seminar, NTNU, Trondheim
February 19, 2025
\[ \newcommand{\diff}{d} \newcommand{\dr}{\diff r} \newcommnand{\R}{{\mathbb R}} \newcommand{\Rd}{{\mathbb R^d}} \newcommand{\diver}{\nabla \cdot} \newcommand{\calF}{\mathcal{F}} \newcommand{\calU}{\mathcal{U}} \newcommand{\defeq}{\overset{\mathrm{def}}{=}} \newcommand{\ee}{\varepsilon} \newcommand{\mob}{\mathrm{m}} \newcommand{\Wass}{\mathfrak{W}} \]
\[\begin{equation*} \label{eq:ADE} \tag{ADE} \frac{\partial \rho}{\partial t} = \diver \Big( \mob(\rho) \nabla ( U'(\rho) + V + W * \rho ) \Big). \end{equation*}\] The mobility \(\mob\) is often linear, i.e., \(\mob(\rho) = \rho\).
For the introductory part we will follow [Gómez-Castro, 2024].
Let \(\rho\) be a density and \(\omega \subset \Rd\) any control volume, if \(\mathbf F\) is the out-going flux \(\frac{\diff }{\diff t} \int_\omega \rho \diff x =- \int_{\partial \omega} \mathbf F \cdot \mathbf n \diff S= -\int_{ \omega} \diver \mathbf F \diff x\).
Hence we arrive at the \(\frac{\partial \rho}{\partial t} = -\diver \mathbf F.\)
Heat equation: Fourier’s law to model the flux \(\mathbf F = -D \nabla \rho\) \[ \frac{\partial \rho}{\partial t} = D\Delta \rho . \]
Porous-Medium Equation: \(\mathbf F = -\rho \mathbf v\); Darcy’s law \(\mathbf v = - \frac{k} \mu \nabla p\);
general state equation \(p = \phi(\rho)\). For perfect gases \(\phi (\rho) = p_0 \rho^\gamma\). \[
\frac{\partial \rho}{\partial t} = \Delta \rho^m.
\]
Non-interacting particles
Consider \(N\) particles moving in a velocity field \(\frac{\diff X_i}{\diff t} = \mathbf v(t, X_i(t))\)
Define the empirical distribution of equal masses \(1/N: \quad\) \(\mu_t^N = \sum_{j=1}^N \frac 1 N \delta_{X_j(t)}\)
This is a weak solution to the continuity equation
\[ \partial_t \mu^N + \diver (\mu^N \mathbf v ) = 0. \]
Interacting particles: Aggregation and Confinement
Non-local interactions: \(\frac{\diff X_i}{\diff t} = -\sum_{\substack{ j=1 \\ j \ne i}}^N \frac 1 N \nabla W (X_i - X_j)- \nabla V (X_i)\).
We recover Aggregation-Confinement Equation
\[ \partial_t \mu^N = \diver (\mu^N \nabla ( W * \mu^N + V ) ). \]
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).
Heat equation \(\quad \partial_t \rho = \Delta \rho_t\)
It corresponds to \(\eqref{eq:ADE}\): \(U = \rho \log \rho\) and \(V = W = 0\).
In \(\Rd\) admits the Gaussian solution \(\rho_t = K_t * \rho_0,\) \[ \rho_t \to 0, \qquad \|\rho_t\|_{L^1} = M , \qquad \| \rho_t - M K_t \|_{L^1} \to 0 \quad\text{ as }\, t \to \infty. \] A good survey on the matter can be found in [Vázquez, 2017].
Linear Fokker-Planck. A self-similar change of variable leads to the linear Fokker-Planck equation \[ \partial_t u = \Delta u_t + \nabla \cdot (x u_t) \]
ADE: \(U_1 (\rho) = \rho \log \rho, V = \frac{|x|^2}{2} \text{ and }W = 0.\)
Asymptotic behaviour: \(u_t \overset{t \to \infty}{\longrightarrow} G\).
The range \(m \in (0,1)\) is sometimes called Fast Diffusion Equation.
ADE: \(U_m (\rho) = \frac{\rho^m}{m-1}, \quad V = W = 0.\)
Self-similar solution \[ B_t = t^{-\alpha} (C - k|x|^2t^{-2\beta})^{\frac 1 {m-1}}_+, \] where \(\alpha = \tfrac{d}{d(m-1)+2},\, \beta = \tfrac{\alpha}{d},\, k = \tfrac{\alpha(m-1)}{d}\).
(see [Vázquez, 2006b])
We get the self-similar solution from a change to self-similar variables \[ \frac{\partial u}{\partial t} = \Delta u^m + \nabla\cdot(x u). \]
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 \(d = 2\).
ADE using Fourier transform: \(V_t = W_{\rm N} * \rho\) where \(W_{\mathrm N} = \mathrm{F}^{-1}[|\xi|^{-2}].\)
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
[Chapman, Rubinstein & Schatzman, 1996] introduced a mean-field model from Gizburg-Landau in the context of superconductivity \[\begin{equation*} \frac{\partial \rho}{\partial t} = \diver( |\rho| \nabla V_t ) \qquad \qquad -\Delta V_t + V_t = \rho_t . \end{equation*}\]
As long as \(\rho \ge 0\) it is ADE: \(U, V = 0, W = \mathrm{F^{-1}}[ (1 + |\xi|^2)^{-1} ]\).
[Huang & Svobodny, 1998], [Schätzle & Styles, 1999] [Ambrosio & Serfaty, 2008], [Ambrosio, Mainini & Serfaty, 2011]
Newtonian vortex problem
In [Lin & Zhang, 1999] taking a different limit from Gizburg-Landau equations \[\begin{equation*} \frac{\partial \rho}{\partial t} = \diver (\rho \nabla V_t ) \qquad \qquad -\Delta V_t = \rho. \end{equation*}\]
[Masmoudi & Zhang, 2005], [Bertozzi, Laurent & Léger, 2012].
Caffarelli-Vázquez
[Caffarelli & Vazquez, 2011] introduces 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. ADE: Riesz potential \(W = \mathrm F^{-1}[|\xi|^{-2s}]\).
[Caffarelli & Vázquez, 2011], [Biler, Imbert & Karch, 2015], [Caffarelli, Soria & Vázquez, 2013],[Caffarelli & Vázquez, 2015], [Serfaty & Vázquez, 2014], [Lisini, Mainini & Segatti, 2018].
Porous-Medium Equation with non-local pressure
\[ \frac{\partial \rho}{\partial t} = \diver (\mob(\rho) \nabla V_t ) \qquad \text{where } \mob(\rho)=\rho^{m-1} \]
[Stan, Del Teso & Vázquez, 2016], [Stan, Del Teso & Vázquez, 2019],
[del Teso & Jakobsen, 2023]
Many authors have devoted their attention to the power-type family of non-linearities given by \[ \begin{gathered} U = \frac{\rho^m}{m-1}, \qquad \qquad V_\lambda (x) = \begin{dcases} \frac{|x|^\lambda}{\lambda} & \text{if } \lambda \ne 0 \\ \log|x| & \text{if } \lambda = 0 , \end{dcases}, \\ W_k (x) = \begin{dcases} \frac{|x|^k}{k} & \text{if } k \in (-d,d) \setminus \{0\} \\ \log|x| & \text{if } k = 0 , \end{dcases} \end{gathered} \]
In bounded domains we can replace \(W*\rho\) by \(\int_{\Omega} K(x,y) \rho(y) dy\).
More examples
See [Gómez-Castro, 2024].
Write linear-mobility ADE \(\mob(\rho) = \rho\) as \[ \partial_t \rho = - \nabla \cdot F, \qquad F = \rho v, \qquad v = -\nabla \xi, \qquad \xi = U'(\rho) + V + \int_\Omega K(\cdot, y)\rho(y)dy. \]
Consider the implicit up-winding finite-volume scheme [Bailo, Carrillo & Hu, 2020] \[ \begin{aligned} \frac{\rho_i^{ n+1} - \rho_i^{n}}{\Delta t} &= - \frac{F^{}_{i+\frac 1 2} (\rho^{n+1}) - F^{}_{i-\frac 1 2} (\rho^{ n+1})}{\Delta x} , \qquad i \in \{1, \cdots, N\}, n \in \mathbb N , \\ F_{i+\frac 1 2}^{} (\rho) &= \rho_i (v_{i+\frac 1 2}^{} (\rho))^+ + \rho_{i+1} (v_{i+\frac 1 2}^{}(\rho))^- , \\ v_{i+\frac 1 2}^{} (\rho) &= -\frac{\xi_{i+1}^{}(\rho) - \xi_{i}^{}(\rho)}{\Delta x} , \\ \xi_i^{}(\rho) &= U_{}'(\rho_i) + V (x_i) + \sum_{j} K_{ij} \rho_j, \\ F_{\frac 1 2}^{} (\rho) &= F_{N + \frac 1 2}^{} (\rho) = 0 . \end{aligned} \]
Key properties
(Some) Non-linear mobilities
When \(F = \mob(\rho) v\) and up-winding is more difficult.
When \(\mob(\rho) = \mob_{\uparrow}(\rho) \mob_{\downarrow}(\rho)\) with \(\mob_{\uparrow}\) non-decreasing and \(\mob_{\downarrow}\) non-increasing.
[Bailo, Carrillo & Hu, 2023] proposed:
\[ F_{i+\frac 1 2}^{} (\rho) = \mob_{\uparrow}(\rho_i) \mob_{\downarrow} (\rho_{i+1}) (v_{i+\frac 1 2}^{} (\rho))^+ + \mob_{\uparrow} (\rho_{i+1}) \mob_{\downarrow} (\rho_{i}) (v_{i+\frac 1 2}^{}(\rho))^- . \\ \]
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 \(t \nearrow T \in (0,\infty]\).
Extinction: \(\rho_t \searrow 0\) as \(t \nearrow T \in (0,\infty]\)
ADE: \(\qquad \partial\rho_t = \nabla \cdot( \rho \nabla (U'(\rho) + V + W*\rho) )\)
Steady state: \(\qquad \rho \nabla (U'(\rho) + V + W*\rho) = 0.\)
This suggests that, for steady states, \[ U'(\rho) + V + W * \rho = C_D \qquad \forall D \subset \mathrm{supp} \rho \text{ open and connected}. \]
The constant can naturally differ between components.
Example. Linear Fokker-Planck. \(\mathrm{supp} \rho=\Rd\)
Example. PME with \(m > 1\) in self-similar variables. Barenblatt profile.
Heat equation:
in \(\Rd\): Gaussian self-similar solution [Vázquez, 2017]
In bounded domains: from spectral decomposicion \(a_1 e^{-\lambda_1 t} \varphi_1\).
Porous-Medium Equation: Barenblatt solution \[\begin{equation} \label{eq:intermediate asymptotics} \frac{\| \rho_t - B_t \|_{L^p}}{\| \rho_t \|_{L^p}} \to 0, \qquad \text{as } t \to \infty. \end{equation}\] See [Vázquez, 2006b], [Vázquez, 2006a].
We can look at the examples \(V = \frac{|x|^\alpha}{\alpha}\) with \(\alpha > 0\), i.e. \(v = -\nabla V = - |x|^{\alpha-2} x\).
Then, \(\rho_t = (X_t)_\sharp \rho_0\) with characteristic field the solutions to \[ \frac{\partial X_t}{\partial t} = -|X_t|^{\alpha-2} X_t, \qquad \qquad X_t(0) = y. \]
This can be explicitly solved as \[ X_t (y) = \begin{dcases} y e^{-t} & \text{if }\alpha=2, \\ ( |y|^{2-\alpha} - (2-\alpha) t )^{\frac 1 {2-\alpha}} \frac{y}{|y|} & \text{if } \alpha \ne 2. \end{dcases} \]
For \(\alpha < 2\) the characteristics arrive at \(0\) at finite time.
We get a Dirac delta at \(0\) at a certain finite time for any \(\rho_0 \not\equiv 0\).
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
\(\exists M^*\) such that if \(\| \rho_0 \|_{L^1} > M^*\), at \(T^* < \infty\) we have then \(\rho_{T^*} = a \delta_0 + \cdots\) .
In \(\Rd\), [Herrero & Velázquez, 1996] showed \(M^* = 8 \pi\).
In more generality in [Dolbeault & Perthame, 2004] \[
\frac{\diff}{\diff t} \int_{\mathbb R^2} |x|^2 \rho(t,x) \diff x
=
4 M \Big( 1- \frac{M}{8\pi }\Big), \qquad \qquad M = \int_{\mathbb R^2} \rho_0 .
\] [Blanchet, Carrillo & Laurençot, 2009] studied \(d > 2\)
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].
Consider the free energy \[\begin{equation*} \tag{FE$^*$} \label{eq:free energy star} \mathcal F(\rho) =\int_\Omega U(\rho) + \int_\Omega V \rho + \frac 1 2 \iint_{\Omega \times \Omega } K(x,y) \rho(x) \rho(y) \diff x \diff y . \end{equation*}\] Its first variation in the pressure field we usually encounter \[\begin{equation} \label{eq:first variation} \frac{\delta \mathcal F}{\delta \rho} [\rho] \defeq U'(\rho) + V + \int_\Omega K(\cdot,y)\rho(y) \diff y. \end{equation}\]
We write ADE as \[\begin{equation*} \partial_t \rho = \diver\left( \mob(\rho) \nabla \frac{\delta \mathcal F}{\delta \rho} [\rho]\right) \end{equation*}\]
Then, the free energy is a Lyapunov functional \[\begin{equation*} \frac{d}{dt} \mathcal F[\rho_t] = \int \frac{\delta \mathcal F}{\delta \rho} [\rho] \partial_t \rho_t = - \int \mob(\rho_t) \left|\nabla \frac{\delta \mathcal F}{\delta \rho} [\rho] \right|^2. \end{equation*}\]
When \(\mob(\rho) = \rho\) there a nice gradient-flow theory:
Otto calculus (see, e.g., [Ambrosio, Brué & Semola, 2021]):
ADE is formally the \(2\)-Wasserstein gradient flow \[
\frac{\partial \rho}{\partial t} = - \nabla_{\Wass_2} \mathcal F[\rho_t].
\]
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 suitable Wasserstein distance
[Dolbeault, Nazaret & Savaré, 2009], [Carrillo, Lisini, Savaré & Slepčev, 2010]
This is only a distance if \(\mob\) is concave.
It is (almost) never displacement convex in our settings.
Under convexity, the asymptotic behaviour are usually the local minimisers.
If \(\calF(\widehat \rho) \le \calF(\rho)\) for all \(\rho \ge 0\) such that \(\int \rho = \int \widehat \rho\) we have that, for some \(C \in \mathbb R\)
\[\begin{equation*} \begin{dcases} \frac{\delta \calF}{\delta\rho} [\widehat \rho] = C & \text{if } \rho > 0, \\ \frac{\delta \calF}{\delta\rho} [\widehat \rho] \ge C & \text{if } \rho = 0. \\ \end{dcases} \end{equation*}\]
Recall \(\frac{\delta \calF}{\delta\rho} [\rho] = U'(\rho) + V + W*\rho\).
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 \(C\).
\[\begin{equation} \tag{NVE} \begin{dcases} \frac{\partial \rho}{\partial t} = \diver (\rho^\alpha \nabla V_t ) \\ -\Delta V_t = \rho. \end{dcases} \end{equation}\]
[Carrillo, Gómez-Castro & Vázquez, 2022a],[Carrillo, Gómez-Castro & Vázquez, 2022c]
We can look for solutions of the form \[ B(t,x)= t^{-\gamma} F(|x|\,t^{-\beta}). \]
Plugging this in the equation we recover the self-similar solution of mass 1 for \(\alpha \in (0,1)\) \[ \scriptsize B(t,x) = t^{-\frac 1 \alpha}\left( \alpha + \left( \frac{ \omega_d |x|^dt^{-\frac 1 {\alpha}} } { \alpha} \right)^{\frac {\alpha} {1-\alpha }} \right)^{-1/\alpha }. \]
The case \(\alpha = 1\) was already studied in [Serfaty & Vázquez, 2014].
The same algebra works for \(\alpha > 1\) but gives no finite-mass solutions
Let us define \[ m(t,x) = \int_{-\infty}^x \rho(t,y) \diff y \]
Integrating the equation for \(\rho\) in \(x\): \(\displaystyle\qquad \frac{\partial m}{\partial t} = \rho^\alpha \frac{\partial V_t}{\partial x}.\)
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 \[ m(t,v) = \int_{A_v} \rho(t,x) \diff x , \quad A_v = B(0,r) \text{ such that }|A_v| = v. \]
Mass conservation implies \(m(t,0) = 0, m(t,\infty) = m_0(\infty)\).
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 \[ G_M(\kappa)=\int \limits _{\omega_d |y|^d \le \kappa } F_M (|y|) \diff y \] 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 \(\varepsilon > 0\).
\[ \partial_t \rho = \Delta \rho^m + \nabla \cdot (\rho \nabla(V + W*\rho)). \]
\(W = 0\): [Carrillo, Gómez-Castro & Vázquez, 2022b]
\(W \ne 0\): [Carrillo, Fernández-Jiménez & Gómez-Castro, 2024]
The Euler-Lagrange equation when \(U'(\rho) = \frac{\rho^{m}}{m-1}\) for \(0 < m < 1\) becomes \[ \widehat \rho(x; C) = \left( \tfrac{1-m}{m} (C + V(x)) \right)^{-\frac 1 {1-m}}. \]
W.l.o.g. \(V(0) = 0\) so \(C \ge 0\).
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.
It may happen that \(\mathfrak M(0^+) < \infty\). Then
\[ \rho_t \overset{t \to \infty}{\longrightarrow} \begin{dcases} \widehat \rho(x; C) & \text{if } M = \mathfrak M(C), \\ \widehat \rho (x; 0) + (\mathfrak M(0) - M) \delta_0 & \text{if } M > \mathfrak M(0) . \end{dcases} \]
For technical convenience, we work in a ball \(B_R\), \(\nabla u \cdot n = \nabla V \cdot n = 0\) on \(\partial B_R\).
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.
\[ \partial_t \rho = \nabla \cdot (\mob(\rho) \nabla(U'(\rho) + W)). \]
where \(\mob(0) = \mob(\beta) = 0\) and \(0 \le \rho_0 \le \beta\).
\[ \partial_t \rho = \nabla \cdot (\mob(\rho) \nabla(U'(\rho) + V + W)). \]
Numerical results obtained in
[Bailo, Carrillo & Hu, 2023]
suggested the formation of free boundaries at levels \(0, \beta\).
Numerical results suggest that the steady states are the doubly-truncated Barenblatts \[ \widehat \rho = \min \Bigg\{ \beta, \max\Big\{ 0, (U')^{-1}(C - V - W*\widehat \rho) \Big\} \Bigg \} \]
We show existence and uniqueness for the approximating problems,
given by semigroup of \(L^1\)-contractions.
We construct a semigroup of \(L^1\)-contractions for the original problem.
We do not provide a notion of solution with uniqueness.1
We prove that the these approximating problems have a unique global attractor \[ \rho_{t}^{(\varepsilon)} \overset{t \to \infty }{\longrightarrow} \widehat \rho^{(\varepsilon)} = (U_\ee')^{-1} (C_\varepsilon - V). \] This is also the global minimiser of the free energy (for a given mass).
We show that \[ \widehat \rho^{(\varepsilon)} \overset{\varepsilon \to 0 }{\longrightarrow} \widehat \rho = \min \Bigg\{ \beta, \max\Big\{ 0, (U')^{-1}(C - V) \Big\} \Bigg \} \] and this is the unique global minimiser of the free energy (for a given mass).
We show that the semigroup solutions converge to steady states,
and that the limits depends continuously on the initial datum
The structure of the \(\omega\)-limit may be complicated, as in linear mobility (recall Figure 3).
Aggregation-diffusion equations with non-linear mobility. David Gómez-Castro (UAM)