In this section, we assume that a sequence of vanishing discretizations \(\boxplus \rightarrow 0\) is given, and we study the respective limit of the fully discrete solutions \((G_\boxplus ^n)_{n\ge 0}\) that are produced by the inductive minimization procedure (3.7). For the analysis of that limit trajectory, it is more natural to work with the induced densities and velocities,
$$\begin{aligned} \rho _\boxplus ^n:=(G_\boxplus ^n)_\#{\overline{\rho }}, \quad \mathbf {v}_\boxplus ^n:=\frac{{\mathrm {id}}-G_\boxplus ^{n-1}\circ (G_\boxplus ^n)^{-1}}{\tau }, \end{aligned}$$
instead of the Lagrangian maps \(G_\boxplus ^n\) themselves. Note that \(\mathbf {v}_\boxplus ^n\) is only well-defined on the support of \(\rho _\boxplus ^n\)—that is, on the image of \(G_\boxplus ^n\)—and can be assigned arbitrary values outside. Let us introduce the piecewise constant in time interpolations \({\widetilde{\rho }}_\boxplus :[0,T]\times {{\mathbb {R}}^d}\rightarrow {\mathbb {R}}_{\ge 0}\), and \({\widetilde{\mathbf {v}}}_\boxplus :[0,T]\times {{\mathbb {R}}^d}\rightarrow {\mathbb {R}}^d\) as usual,
$$\begin{aligned} {\widetilde{\rho }}_\boxplus (t) = \rho _\boxplus ^n, \quad {\widetilde{\mathbf {v}}}_\boxplus (t) = \mathbf {v}_\boxplus ^n \quad \text {with }n \text { such that }t\in ((n-1)\tau ,n\tau ]. \end{aligned}$$
Note that \({\widetilde{\rho }}(t,\cdot )\in {\mathcal {P}}_2^\text {ac}({\mathbb {R}}^d)\) and \({\widetilde{\mathbf {v}}}_\boxplus (t,\cdot )\in L^2({{\mathbb {R}}^d}\rightarrow {\mathbb {R}}^d;{\widetilde{\rho }}_\boxplus (t,\cdot ))\) at each \(t\ge 0\).
Energy Estimates
We start by proving the classical energy estimates on minimizing movements for our fully discrete scheme.
Lemma 4.1
For each discretization \(\boxplus \) and for any indices \(\overline{n}>\underline{n}\ge 0\), one has the a priori estimate
$$\begin{aligned} {\mathcal {E}}(\rho _\boxplus ^{\overline{n}}) +\frac{\tau }{2}\sum _{n=\underline{n}+1}^{\overline{n}}\left( \frac{\mathrm {W}_2(\rho _\boxplus ^n,\rho _\boxplus ^{n-1})}{\tau }\right) ^2 \le {\mathcal {E}}(\rho ^{\underline{n}}). \end{aligned}$$
(4.1)
Consequently:
-
(1)
\({\mathbf {E}}\) is monotonically decreasing, i.e., \({\mathcal {E}}({\widetilde{\rho }}_\boxplus (t))\le {\mathcal {E}}({\widetilde{\rho }}_\boxplus (s))\) for all \(t\ge s\ge 0\);
-
(2)
\({\widetilde{\rho }}_\boxplus \) is Hölder-\(\frac{1}{2}\)-continuous in \(\mathrm {W}_2\), up to an error \(\tau \),
$$\begin{aligned} \mathrm {W}_2\big ({\widetilde{\rho }}_\boxplus (t),{\widetilde{\rho }}_\boxplus (s)\big ) \le \sqrt{2{\mathcal {E}}(\rho _\boxplus ^0)}\big (|t-s|^{\frac{1}{2}}+\tau ^{\frac{1}{2}}\big ) \quad \text { for all }t\ge s\ge 0. \end{aligned}$$
(4.2)
-
(3)
\({\widetilde{\mathbf {v}}}_\boxplus \) is square integrable with respect to \({\widetilde{\rho }}_\boxplus \),
$$\begin{aligned} \int _0^T\int _{{\mathbb {R}}^d}\Vert {\widetilde{\mathbf {v}}}_\boxplus \Vert ^2{\widetilde{\rho }}_\boxplus \,\mathrm {d}x\,\mathrm {d}t \le 2{\mathcal {E}}(\rho _\boxplus ^0). \end{aligned}$$
(4.3)
Proof
By the definition of \(G_\boxplus ^n\) as a minimizer, we know that \({\mathbf {E}}_\boxplus (G_\boxplus ^n;G_\boxplus ^{n-1})\le {\mathbf {E}}_\boxplus (G;G_\boxplus ^{n-1})\) for any \(G\in {\mathcal {A}}_{\mathscr {T}}\), and in particular for the choice \(G:=G_\boxplus ^{n-1}\), which yields:
$$\begin{aligned} \frac{1}{2\tau }\int _K\Vert G_\boxplus ^n-G_\boxplus ^{n-1}\Vert ^2{\overline{\rho }}_{\mathscr {T}}\,\mathrm {d}\omega +{\mathbf {E}}(G_\boxplus ^n|{\overline{\rho }}_{\mathscr {T}}) \le {\mathbf {E}}(G_\boxplus ^{n-1}|{\overline{\rho }}_{\mathscr {T}}). \end{aligned}$$
(4.4)
Summing these inequalies for \(n=\underline{n}+1,\ldots ,\overline{n}\), recalling that \({\mathcal {E}}(\rho _\boxplus ^n)={\mathbf {E}}(G_\boxplus ^n|{\overline{\rho }}_{\mathscr {T}})\) by (1.9) and that \(\mathrm {W}_2(\rho _\boxplus ^n,\rho _\boxplus ^{n-1})^2\le \int _K|G_\boxplus ^n-G_\boxplus ^{n-1}|^2{\overline{\rho }}\,\mathrm {d}\omega \) by (2.3), produces (4.1).
Monotonicity of \({\mathcal {E}}\) in time is obvious.
To prove (4.2), choose \(\underline{n}\le \overline{n}\) such that \(s\in ((\underline{n}-1)\tau ,\underline{n}\tau ]\) and \(t\in ((\overline{n}-1)\tau ,\overline{n}\tau ]\). Notice that \(\tau (\overline{n}-\underline{n})\le t-s+\tau \). If \(\underline{n}=\overline{n}\), the claim (4.2) is obviously true; let \(\underline{n}<\overline{n}\) in the following. Combining the triangle inequality for the metric \(\mathrm {W}_2\), estimate (4.1) above and Hölder’s inequality for sums, we arrive at
$$\begin{aligned} \mathrm {W}_2\big ({\widetilde{\rho }}_\boxplus (t),{\widetilde{\rho }}_\boxplus (s)\big )&= \mathrm {W}_2(\rho _\boxplus ^{\overline{n}},\rho _\boxplus ^{\underline{n}}) \le \sum _{n=\underline{n}+1}^{\overline{n}}\mathrm {W}_2(\rho _\boxplus ^n,\rho _\boxplus ^{n-1}) \\&\le \left[ \sum _{n=\underline{n}+1}^{\overline{n}}\tau \right] ^{\frac{1}{2}} \left[ \sum _{n=\underline{n}+1}^{\overline{n}}\frac{\mathrm {W}_2(\rho _\boxplus ^n,\rho _\boxplus ^{n-1})^2}{\tau }\right] ^{\frac{1}{2}}\\&= \big [\tau (\overline{n}-\underline{n}) \big ]^{\frac{1}{2}} \left[ \tau \sum _{n=\underline{n}+1}^{\overline{n}}\left( \frac{\mathrm {W}_2(\rho _\boxplus ^n,\rho _\boxplus ^{n-1})}{\tau }\right) ^2\right] ^{\frac{1}{2}}\\&\le [t-s+\tau ]^{\frac{1}{2}} \left[ 2\big ({\mathcal {E}}(\rho _\boxplus ^{\underline{n}}) - {\mathcal {E}}(\rho _\boxplus ^{\overline{n}})\big )\right] ^{\frac{1}{2}} \\&\le \big [|t-s|^{\frac{1}{2}}+\tau ^{\frac{1}{2}}\big ]{\mathcal {E}}(\rho _\boxplus ^0)^{\frac{1}{2}}. \end{aligned}$$
Finally, changing variables using \(x=G_\boxplus ^n(\omega )\) in (4.4) yields
$$\begin{aligned} \frac{\tau }{2}\int _{{\mathbb {R}}^d}\Vert \mathbf {v}_\boxplus ^n\Vert ^2\rho _\boxplus ^n\,\mathrm {d}x + {\mathbf {E}}(G_\boxplus ^n) \le {\mathbf {E}}(G_\boxplus ^{n-1}), \end{aligned}$$
and summing these inequalities from \(n=1\) to \(n=N_\tau \) yields (4.3). \(\square \)
Compactness of the Trajectories and Weak Formulation
Our main result on the weak limit of \({\widetilde{\rho }}_\boxplus \) is the following.
Theorem 4.2
Along a suitable sequence \(\boxplus \rightarrow 0\), the curves \({\widetilde{\rho }}_\boxplus :{\mathbb {R}}_{\ge 0}\rightarrow {\mathcal {P}}_2^\text {ac}({\mathbb {R}}^d)\) convergence pointwise in time, i.e., \({\widetilde{\rho }}_\boxplus (t)\rightarrow \rho _*(t)\) narrowly for each \(t>0\), towards a Hölder-\(\frac{1}{2}\)-continuous limit trajectory \(\rho _* :{\mathbb {R}}_{\ge 0}\rightarrow {\mathcal {P}}_2^\text {ac}({\mathbb {R}}^d)\).
Moreover, the discrete velocities \({\widetilde{\mathbf {v}}}_\boxplus \) possess a limit \(\mathbf {v}_*\in L^2({\mathbb {R}}_{\ge 0}\times {{\mathbb {R}}^d};\rho _*)\) such that \({\widetilde{\mathbf {v}}}_\boxplus {\widetilde{\rho }}_\boxplus \overset{*}{\rightharpoonup }\mathbf {v}_*\rho _*\) in \(L^1({\mathbb {R}}_{\ge 0}\times {{\mathbb {R}}^d})\), and the continuity equation
$$\begin{aligned} \partial _t\rho _* + \nabla \cdot (\rho _*\mathbf {v}_*) = 0 \end{aligned}$$
(4.5)
holds in the sense of distributions.
Remark 4.3
The Hölder continuity of \(\rho _*\) implies that \(\rho _*\) satisfies the initial condition (1.1b) in the sense that \(\rho _*(t)\rightarrow \rho ^0\) narrowly as \(t\downarrow 0\).
Proof of Theorem 4.2
We closely follow an argument that is part of the general convergence proof for the minimizing movement scheme as given in Ambrosio et al. [1, Section 11.1.3]. Below, convergence is shown for some arbitrary but fixed time horizon \(T>0\); a standard diagonal argument implies convergence at arbitrary times.
First observe that by estimate (4.2)—applied with \(0=s\le t\le T\)—it follows that \(\mathrm {W}_2({\widetilde{\rho }}_\boxplus (t),\rho _\boxplus ^0)\) is bounded, uniformly in \(t\in [0,T]\) and in \(\boxplus \). Since further \(\rho ^0_\boxplus \) converges narrowly to \(\rho ^0\) by our hypotheses on the initial approximation, we conclude that all densities \({\widetilde{\rho }}_\boxplus (t)\) belong to a sequentially compact subset for the narrow convergence. The second observation is that the term on the right hand side of (4.2) simplifies to \((2\overline{{\mathcal {E}}})^\frac{1}{2}|t-s|^\frac{1}{2}\) in the limit \(\boxplus \rightarrow 0\). A straightforward application of the “refined version” of the Ascoli-Arzelà theorem (Proposition 3.3.1 in Ambrosio et al. [1]) yields the first part of the claim, namely the pointwise narrow convergence of \({\widetilde{\rho }}_\boxplus \) towards a Hölder continuous limit curve \(\rho _*\).
It remains to pass to the limit with the velocity \({\widetilde{\mathbf {v}}}_\boxplus \). Towards that end, we define a probability measure \(\widetilde{\gamma }_\boxplus \in {\mathcal {P}}(Z_T)\) on the set \(Z_T:=[0,T]\times {\mathbb {R}}^d\times {\mathbb {R}}^d\) as follows:
$$\begin{aligned} \int _{Z_T}\varphi (t,x,v)\,\mathrm {d}\widetilde{\gamma }_\boxplus (t,x,v) = \int _0^T\int _{{\mathbb {R}}^d} \varphi \big (t,x,{\widetilde{\mathbf {v}}}_\boxplus (t,x)\big )\,{\widetilde{\rho }}_\boxplus (t,x)\,\mathrm {d}x\frac{\mathrm {d}t}{T}, \end{aligned}$$
for every bounded and continuous function \(\varphi \in C^0_b(Z_T)\). For brevity, let \(\widetilde{M}_\boxplus \in {\mathcal {P}}([0,T]\times {\mathbb {R}}^d)\) be the (t, x)-marginals of \(\widetilde{\gamma }_\boxplus \), that have respective Lebesgue densities \(\frac{\rho _\boxplus (t,x)}{T}\) on \([0,T]\times {\mathbb {R}}^d\). Thanks to the result from the first part of the proof, \(\widetilde{M}_\boxplus \) converges narrowly to a limit \(M_*\), which has density \(\frac{\rho _*(t,x)}{T}\). On the other hand, the estimate (4.3) implies that
$$\begin{aligned} \int _{Z_T} |v|^2\,\mathrm {d}\widetilde{\gamma }_\boxplus (t,x,v) = \int _{[0,T]\times {\mathbb {R}}^d}|{\widetilde{\mathbf {v}}}_\boxplus (t,x)|^2\,\mathrm {d}\widetilde{M}_\boxplus (t,x) \le 2\overline{{\mathcal {E}}}. \end{aligned}$$
We are thus in the position to apply Theorem 5.4.4 in Ambrosio et al. [1], which yields the narrow convergence of \(\widetilde{\gamma }_\boxplus \) towards a limit \(\gamma _*\). Clearly, the (t, x)-marginal of \(\gamma _*\) is \(M_*\). Accordingly, we introduce the disintegration \(\gamma _{(t,x)}\) of \(\gamma _*\) with respect to \(M_*\), which is well-defined \(M_*\)-a.e.. Below, it will turn out that \(\gamma _*\)’s v-barycenter,
$$\begin{aligned} \mathbf {v}_*(t,x) := \int _{{\mathbb {R}}^d} v \,\mathrm {d}\gamma _{(t,x)}(v), \end{aligned}$$
(4.6)
is the sought-for weak limit of \({\widetilde{\mathbf {v}}}_\boxplus \). The convergence \({\widetilde{\mathbf {v}}}_\boxplus {\widetilde{\rho }}_\boxplus \overset{*}{\rightharpoonup }\mathbf {v}_*\rho _*\) and the inheritance of the uniform \(L^2\)-bound (4.3) to the limit \(\mathbf {v}_*\) are further direct consequences of Theorem 5.4.4 in Ambrosio et al. [1].
The key step to establish the continuity equation for the just-defined \(\mathbf {v}_*\) is to evaluate the limit as \(\boxplus \rightarrow 0\) of
$$\begin{aligned} J_\boxplus [\phi ] := \frac{1}{\tau }\left[ \int _0^T\int _{{\mathbb {R}}^d}\phi (t,x){\widetilde{\rho }}_\boxplus (t,x)\,\mathrm {d}x\,\mathrm {d}t - \int _0^T\int _{{\mathbb {R}}^d}\phi (t,x){\widetilde{\rho }}_\boxplus (t-\tau ,x)\,\mathrm {d}x\,\mathrm {d}t \right] \end{aligned}$$
for any given test function \(\phi \in C^\infty _c((0,T)\times {\mathbb {R}}^d)\) in two different ways. First, we change variables \(t\mapsto t+\tau \) in the second integral, which gives
$$\begin{aligned} J_\boxplus [\phi ]&= \int _0^T\int _{{\mathbb {R}}^d} \frac{\phi (t,x)-\phi (t+\tau ,x)}{\tau }{\widetilde{\rho }}_\boxplus (t,x)\,\mathrm {d}x\,\mathrm {d}t {\mathop {\longrightarrow }\limits ^{\boxplus \rightarrow 0}}\\&\quad -\int _0^T\int _{{\mathbb {R}}^d} \partial _t\phi (t,x)\,\rho _*(t,x)\,\mathrm {d}x\,\mathrm {d}t. \end{aligned}$$
For the second evaluation, we write
$$\begin{aligned} \rho _\boxplus ^{n-1} = \big (G_\boxplus ^{n-1}\circ (G_\boxplus ^n)^{-1}\big )_\#\rho _\boxplus ^n = \big ({\mathrm {id}}-\tau \mathbf {v}_\boxplus ^n\big )_\#\rho _\boxplus ^n, \end{aligned}$$
and substitute accordingly \(x\mapsto x-\tau {\widetilde{\mathbf {v}}}_\boxplus (t,x)\) in the second integral, leading to
$$\begin{aligned} J_\boxplus [\phi ]&= \int _0^T\int _{{\mathbb {R}}^d} \frac{\phi (t,x)-\phi \big (t,x-\tau {\widetilde{\mathbf {v}}}_\boxplus (t,x)\big )}{\tau }{\widetilde{\rho }}_\boxplus (t,x)\,\mathrm {d}x\,\mathrm {d}t \\&= \int _0^T\int _{{\mathbb {R}}^d} \nabla \phi (t,x)\cdot {\widetilde{\mathbf {v}}}_\boxplus (t,x){\widetilde{\rho }}_\boxplus (t,x)\,\mathrm {d}x\,\mathrm {d}t + \mathfrak {e}_\boxplus [\phi ]\\&= \int _{Z_T} \nabla \phi (t,x)\cdot v\,\mathrm {d}\widetilde{\gamma }_\boxplus (t,x,v) + \mathfrak {e}_\boxplus [\phi ] \\&\quad {\mathop {\longrightarrow }\limits ^{\boxplus \rightarrow 0}} \int _{Z_T} \nabla \phi (t,x)\cdot v\,\mathrm {d}\gamma _*(t,x,v) \\&= \int _{[0,T]\times {\mathbb {R}}^d}\nabla \phi (t,x)\cdot \left[ \int _{{\mathbb {R}}^d}v\,\mathrm {d}\gamma _{(t,x)}(v)\right] \,\mathrm {d}M_*(t,x) \\&= \int _0^T\int _{{\mathbb {R}}^d}\nabla \phi (t,x)\cdot \mathbf {v}_*(t,x)\rho _*(t,x)\,\mathrm {d}x\,\mathrm {d}t. \end{aligned}$$
The error term \(\mathfrak {e}_\boxplus [\phi ]\) above is controlled via Taylor expansion of \(\phi \) and by using (4.3),
$$\begin{aligned} \big |\mathfrak {e}_\boxplus [\phi ]\big | \le \int _0^T\int _{{\mathbb {R}}^d}\frac{\tau }{2}\Vert \phi \Vert _{C^2}\big \Vert {\widetilde{\mathbf {v}}}_\boxplus (t,x)\big \Vert ^2{\widetilde{\rho }}_\boxplus (t,x)\,\mathrm {d}x\,\mathrm {d}t \le \overline{{\mathcal {E}}}\Vert \phi \Vert _{C^2}T\;\tau . \end{aligned}$$
Equality of the limits for both evaluations of \(J_\boxplus [\phi ]\) for arbitrary test functions \(\phi \) shows the continuity Eq. (4.5). \(\square \)
Unfortunately, the convergence provided by Theorem 4.2 is generally not sufficient to conclude that \(\rho _*\) is a weak solution to (1.1), since we are not able to identify \(\mathbf {v}_*\) as \(\mathbf {v}[\rho _*]\) from (1.4b). The problem is two-fold: first, weak-\(\star \) convergence of \({\widetilde{\rho }}_\boxplus \) is insufficient to pass to the limit inside the nonlinear function P. Second, even if we would know that, for instance, \(P({\widetilde{\rho }}_\boxplus )\overset{*}{\rightharpoonup }P(\rho _*)\), we would still need a \(\boxplus \)-independent a priori control on the regularity (e.g., maximal diameter of triangles) of the meshes generated by the \(G_\boxplus ^n\) to justify the passage to limit in the weak formulation below.
The main difficulty in the weak formulation that we derive now is that we can only use “test functions” that are piecewise affine with respect to the changing meshes generated by the \(G_\boxplus ^n\). For definiteness, we introduce the space
$$\begin{aligned} \mathcal {D}({\mathscr {T}}):=\left\{ \Gamma :K\rightarrow {\mathbb {R}}^d\,;\,\Gamma \text { is globally continuous, and is piecewise affine w.r.t. }\Delta _m\right\} . \end{aligned}$$
Lemma 4.4
Assume \(S :{{\mathbb {R}}^d}\rightarrow {\mathbb {R}}^d\) is such that \(S\circ G_\boxplus ^n\in \mathcal {D}({\mathscr {T}})\). Then:
$$\begin{aligned} \int _{{{\mathbb {R}}^d}} P(\rho _\boxplus ^{n}) \, \nabla \cdot S \,\mathrm {d}x - \int _{{{\mathbb {R}}^d}} \nabla V \cdot S\, \rho _\boxplus ^{n} \,\mathrm {d}x = \int _{{{\mathbb {R}}^d}} S\cdot \mathbf {v}_\boxplus ^n \rho _\boxplus ^{n} \,\mathrm {d}x. \end{aligned}$$
(4.7)
Proof
For all sufficiently small \(\varepsilon >0\), let \(G_\varepsilon = ({\mathrm {id}}+S)\circ G_\boxplus ^n\). By definition of \(G_\boxplus ^n\) as a minimizer, we have that \({\mathbf {E}}_\boxplus (G_\varepsilon ;G_\boxplus ^{n-1})\ge {\mathbf {E}}_\boxplus (G_\boxplus ^n;G_\boxplus ^{n-1})\). This implies that
$$\begin{aligned} 0&\le \frac{1}{\varepsilon }\int _K\bigg (\frac{1}{2\tau }\big [\Vert G_\varepsilon -G_\boxplus ^{n-1}\Vert ^2-\Vert G_\boxplus ^n-G_\boxplus ^{n-1}\Vert ^2\big ] \nonumber \\&\qquad + \left[ \widetilde{h}\left( \frac{\det \mathrm {D}G_\varepsilon }{{\overline{\rho }}_{\mathscr {T}}}\right) -\widetilde{h}\left( \frac{\det \mathrm {D}G_\boxplus ^n}{{\overline{\rho }}_{\mathscr {T}}}\right) \right] + \big [V\circ G_\varepsilon -V\big ]\bigg ) {\overline{\rho }}_{\mathscr {T}}\,\mathrm {d}\omega . \end{aligned}$$
(4.8)
We discuss limits of the three terms under the integral for \(\varepsilon \searrow 0\). For the metric term:
$$\begin{aligned} \frac{1}{2\tau \varepsilon }\left[ \Vert G_\varepsilon -G_\boxplus ^{n-1}\Vert ^2-\Vert G_\boxplus ^n-G_\boxplus ^{n-1}\Vert ^2\right]&= \frac{G_\boxplus ^n-G_\boxplus ^{n-1}}{\tau }\cdot \frac{G_\varepsilon -G_\boxplus ^n}{\varepsilon }+ \frac{1}{2\tau \varepsilon }\Vert G_\varepsilon -G_\boxplus ^n\Vert ^2 \\&= \left[ \left( \frac{{\mathrm {id}}-T_\boxplus ^n}{\tau }\right) \cdot S\right] \circ G_\boxplus ^n + \frac{\varepsilon }{2\tau }\Vert S\Vert ^2\circ G_\boxplus ^n, \end{aligned}$$
and since S is bounded, the last term vanishes uniformly on K for \(\varepsilon \searrow 0\). For the internal energy, since \(\mathrm {D}G_\varepsilon =\mathrm {D}({\mathrm {id}}+\varepsilon S)\circ G_\boxplus ^n\cdot \mathrm {D}G_\boxplus ^n\), and recalling (3.8),
$$\begin{aligned}&\frac{1}{\varepsilon }\left[ \widetilde{h}\left( \frac{\det \mathrm {D}G_\varepsilon }{{\overline{\rho }}_{\mathscr {T}}}\right) -\widetilde{h}\left( \frac{\det \mathrm {D}G_\boxplus ^n}{{\overline{\rho }}_{\mathscr {T}}}\right) \right] \\&\quad = \frac{1}{\varepsilon }\left[ \widetilde{h}\left( \frac{\det \mathrm {D}G_\boxplus ^n}{{\overline{\rho }}_{\mathscr {T}}}\det ({\mathbb {1}}+\varepsilon \mathrm {D}S)\circ G_\boxplus ^n\right) -\widetilde{h}\left( \frac{\det \mathrm {D}G_\boxplus ^n}{{\overline{\rho }}_{\mathscr {T}}}\right) \right] \\&\quad {\mathop {\longrightarrow }\limits ^{\varepsilon \searrow 0}} \frac{\det \mathrm {D}G_\boxplus ^n}{{\overline{\rho }}_{\mathscr {T}}}\widetilde{h}'\left( \frac{\det \mathrm {D}G_\boxplus ^n}{{\overline{\rho }}_{\mathscr {T}}}\right) \left( \lim _{\varepsilon \searrow 0}\frac{\det ({\mathbb {1}}+\varepsilon \mathrm {D}S)}{\varepsilon }\right) \circ G_\boxplus ^n \\&\quad = -\frac{\det \mathrm {D}G_\boxplus ^n}{{\overline{\rho }}_{\mathscr {T}}} P\left( \frac{{\overline{\rho }}_{\mathscr {T}}}{\det \mathrm {D}G_\boxplus ^n}\right) {{\mathrm{tr}}}[\mathrm {D}S]\circ G_\boxplus ^n \\&\quad = - \frac{\det \mathrm {D}G_\boxplus ^n}{{\overline{\rho }}_{\mathscr {T}}}\big [P(\rho ^n)\,\nabla \cdot S\big ]\circ G_\boxplus ^n. \end{aligned}$$
Since the piecewise constant function \(\det \mathrm {D}G_\boxplus ^n\) has a positive lower bound, the convergence as \(\varepsilon \searrow 0\) is uniform on K. Finally, for the potential energy,
$$\begin{aligned} \frac{1}{\varepsilon }\big [V\circ ({\mathrm {id}}+\varepsilon S)\circ G_\boxplus ^n-V\circ G_\boxplus ^n\big ] {\mathop {\longrightarrow }\limits ^{\varepsilon \searrow 0}} \big [\nabla V\cdot S\big ]\circ G_\boxplus ^n. \end{aligned}$$
Again, the convergence is uniform on K. Passing to the limit in the integral (4.8) yields
$$\begin{aligned} 0&\le \int _K \left[ \left( \frac{{\mathrm {id}}-T_\boxplus ^n}{\tau }\right) \cdot S\right] \circ G_\boxplus ^n{\overline{\rho }}_{\mathscr {T}}\,\mathrm {d}\omega \\&\qquad - \int _K \big [P(\rho ^n)\,\nabla \cdot S\big ]\circ G_\boxplus ^n \det \mathrm {D}G_\boxplus ^n\,\mathrm {d}\omega + \int _K \big [\nabla V\cdot S\big ]\circ G_\boxplus ^n{\overline{\rho }}_{\mathscr {T}}\,\mathrm {d}\omega . \end{aligned}$$
The same inequality is true with \(-S\) in place of S, hence this inequality is actually an equality. Since \(\rho _\boxplus ^n=(G_\boxplus ^n)_\#{\overline{\rho }}_{\mathscr {T}}\), a change of variables \(x=S_\boxplus ^n(\omega )\) produces (4.7). \(\square \)
Corollary 4.5
In addition to the hypotheses of Theorem 4.2, assume that
-
(1)
\(P({\widetilde{\rho }}_\boxplus )\overset{*}{\rightharpoonup }p_*\) in \(L^1([0,T]\times \Omega )\);
-
(2)
each \(G_\boxplus ^n\) is injective;
-
(3)
as \(\boxplus \rightarrow 0\), all simplices in the images of \({\mathscr {T}}\) under \(G_\boxplus ^n\) have non-degenerate interior angles and tend to zero in diameter, uniformly w.r.t. n.
Then \(\rho _*\) satisfies the PDE
$$\begin{aligned} \partial _t\rho _* = \Delta p_* + \nabla \cdot (\rho _*\nabla V) \end{aligned}$$
(4.9)
in the sense of distributions.
Proof
Let a smooth test function \(\zeta \in C^\infty _c({\mathbb {R}}^d\rightarrow {\mathbb {R}}^d)\) be given. For each \(\boxplus \) and each n, a \(\zeta _\boxplus ^n :{\mathbb {R}}^d\rightarrow {\mathbb {R}}^d\) with \(\zeta _\boxplus ^n\circ G_\boxplus ^n\in \mathcal {D}({\mathscr {T}})\) can be constructed in such a way that
$$\begin{aligned} \zeta _\boxplus ^n\rightarrow \zeta , \quad \nabla \cdot \zeta _\boxplus ^n\rightarrow \nabla \cdot \zeta \end{aligned}$$
(4.10)
uniformly on \({\mathbb {R}}^d\), and uniformly in n as \(\boxplus \rightarrow 0\). This follows from our hypotheses on the \(\boxplus \)-uniform regularity of the Lagrangian meshes: inside the image of \(G_\boxplus ^n\), one can simply choose \(\zeta _\boxplus ^n\) as the affine interpolation of the values of \(\zeta \) at the points \(G_\boxplus ^n(\omega _\ell )\). Outside, one can take an arbitrary approximation of \(\zeta \) that is compatible with the piecewise-affine approximation on the boundary of \(G_\boxplus ^n\)’s image; one may even choose \(\zeta _\boxplus ^n\equiv \zeta \) at sufficient distance to that boundary. The uniform convergences (4.10) then follow by standard finite element analysis.
Further, let \(\eta \in C^\infty _c(0,T)\) be given. For each \(t\in ((n-1)\tau ,n\tau ]\), substitute \(S(t,x):=\eta (t)\zeta _\boxplus ^n(x)\) into (4.7). Integration of these equalities with respect to \(t\in (0,T)\) yields
$$\begin{aligned} \int _0^T\int _{{\mathbb {R}}^d} P({\widetilde{\rho }}_\boxplus )\nabla \cdot S\,\mathrm {d}x\,\mathrm {d}t - \int _0^T\int _{{\mathbb {R}}^d} \nabla V\cdot S\,\mathrm {d}x\,\mathrm {d}t = \int _0^T\int _{{\mathbb {R}}^d} S\cdot {\widetilde{\mathbf {v}}}_\boxplus {\widetilde{\rho }}_\boxplus \,\mathrm {d}x\,\mathrm {d}t. \end{aligned}$$
We pass to the limit \(\boxplus \rightarrow 0\) in these integrals. For the first, we use that \(P({\widetilde{\rho }})\overset{*}{\rightharpoonup }p_*\) by hypothesis, for the last, we use Theorem 4.2 above. Since any test function \(S\in C^\infty _c((0,T)\times \Omega )\) can be approximated in \(C^1\) by linear combinations of products \(\eta (t)\zeta (x)\) as above, we thus obtain the weak formulation of
$$\begin{aligned} \rho _*\mathbf {v}_* = \nabla p_* + \rho _*\nabla V. \end{aligned}$$
In combination with the continuity Eq. (4.5), we arrive at (4.9). \(\square \)
Remark 4.6
In principle, our discretization can also be applied to the linear Fokker–Planck equation with \(P(r)=r\) and \(h(r)=r\log r\). In that case, one automatically has \(P({\widetilde{\rho }})\overset{*}{\rightharpoonup }p_*\equiv P(\rho _*)\) thanks to Theorem 4.2. Corollary 4.5 above then provides an a posteriori criterion for convergence: if the Lagrangian mesh does not deform too wildly under the dynamics as the discretization is refined, then the discrete solutions converge to the genuine solution.