Numerical simulation of nonlinear continuity equations by ...

26
Numerical simulation of nonlinear continuity equations by evolving diffeomorphisms Jos´ e A. Carrillo a , Helene Ranetbauer b , Marie-Therese Wolfram c a Department of Mathematics, Imperial College London, London SW7 2AZ, UK; email: [email protected] b Radon Institute for Computational and Applied Mathematics, Austrian Academy of Sciences, Altenberger Strasse 69, 4040 Linz, Austria; email: [email protected] c Mathematics Institute, University of Warwick, Coventry CV4 7AL, UK and Radon Institute for Computational and Applied Mathematics, Austrian Academy of Sciences, Altenberger Strasse 69, 4040 Linz, Austria; email: [email protected] Abstract In this paper we present a numerical scheme for nonlinear continuity equations, which is based on the gradient flow formulation of an energy functional with respect to the quadratic transportation distance. It can be applied to a large class of nonlinear continuity equations, whose dynamics are driven by internal energies, given external potentials and/or interaction energies. The solver is based on its variational formulation as a gradient flow with respect to the Wasserstein distance. Positivity of solutions as well as energy decrease of the semi- discrete scheme are guaranteed by its construction. We illustrate this properties with various examples in spatial dimension one and two. Keywords: Lagrangian coordinates, variational scheme, optimal transport, implicit in time discretization. 1. Introduction In this work we propose a numerical method for solving nonlinear continuity equations of the form: t ρ = -∇ · [ρv] := ∇· [ρ(U 0 (ρ)+ V + W * ρ)] (1a) ρ(0, ·)= ρ 0 , (1b) where ρ = ρ(t, x) denotes the unknown time dependent probability density and ρ 0 = ρ 0 (x) a given initial probability density on Ω R d . The function U : R + R is an internal energy, V = V (x): R d R a given potential and W = W (x): R d R an interaction potential. Equation (1) can be interpreted as a gradient flow with respect to the Euclidean Preprint submitted to Journal of Computational Physics September 15, 2016

Transcript of Numerical simulation of nonlinear continuity equations by ...

Page 1: Numerical simulation of nonlinear continuity equations by ...

Numerical simulation of nonlinear continuity equations by

evolving diffeomorphisms

Jose A. Carrilloa, Helene Ranetbauerb, Marie-Therese Wolframc

aDepartment of Mathematics, Imperial College London, London SW7 2AZ, UK;email: [email protected]

bRadon Institute for Computational and Applied Mathematics, Austrian Academy of Sciences,Altenberger Strasse 69, 4040 Linz, Austria;email: [email protected]

cMathematics Institute, University of Warwick, Coventry CV4 7AL, UK and Radon Institute forComputational and Applied Mathematics, Austrian Academy of Sciences, Altenberger Strasse 69, 4040

Linz, Austria;email: [email protected]

Abstract

In this paper we present a numerical scheme for nonlinear continuity equations, which isbased on the gradient flow formulation of an energy functional with respect to the quadratictransportation distance. It can be applied to a large class of nonlinear continuity equations,whose dynamics are driven by internal energies, given external potentials and/or interactionenergies. The solver is based on its variational formulation as a gradient flow with respectto the Wasserstein distance. Positivity of solutions as well as energy decrease of the semi-discrete scheme are guaranteed by its construction. We illustrate this properties withvarious examples in spatial dimension one and two.

Keywords: Lagrangian coordinates, variational scheme, optimal transport, implicit intime discretization.

1. Introduction

In this work we propose a numerical method for solving nonlinear continuity equationsof the form:

∂tρ = −∇ · [ρv] := ∇ · [ρ∇ (U ′(ρ) + V +W ∗ ρ)] (1a)

ρ(0, ·) = ρ0, (1b)

where ρ = ρ(t, x) denotes the unknown time dependent probability density and ρ0 = ρ0(x)a given initial probability density on Ω ⊆ Rd. The function U : R+ → R is an internalenergy, V = V (x) : Rd → R a given potential and W = W (x) : Rd → R an interactionpotential. Equation (1) can be interpreted as a gradient flow with respect to the Euclidean

Preprint submitted to Journal of Computational Physics September 15, 2016

Page 2: Numerical simulation of nonlinear continuity equations by ...

Wasserstein distance of the free energy or entropy

E(ρ) =

∫RdU(ρ(x))dx+

∫RdV (x)ρ(x)dx+

1

2

∫Rd×Rd

W (x− y)ρ(x)ρ(y)dxdy.

Then the velocity field v = v(t, x) corresponds to v = −∇ δEδρ

, and the free energy is

dissipated along the trajectories of equation (1), i.e.

d

dtE(ρ)(t) = −D(ρ) ≡ −

∫Rd|v(t, x)|2ρ(t, x)dx.

Here D(ρ) denotes the so called entropy dissipation functional.The gradient flow formulation detailed above provides a natural framework to de-

scribe the evolution of densities and it has been successfully used to model transporta-tion processes in the life and social sciences. Examples of it include the heat equationwith U(s) = s ln s, V = W = 0 or the porous medium and fast diffusion equation withV = W = 0 and U(s) = sm/(m− 1) for m > 1 and 0 < m < 1 respectively. If the dynam-ics are in addition driven by a given potential V , we refer to (1) as a Fokker-Planck typeequation. Gradient flow techniques have also been applied to aggregation equations, whichcorrespond to (1) for a given interaction potential W and U = V = 0. These models appearnaturally to describe spatial shapes of collective dynamics in general and in particular foranimal swarms, for example the motion of bird flocks or fish schools, cf. [1, 2, 3, 4]. See alsothe reviews [5, 6] and the references therein. Applications in physics include the field ofgranular media [7, 8] or material sciences [9]. A common assumption in collective dynamicsis that W is a radial function, i.e. W = W (|x|). Hence interactions among individualsdepend on their distance only. The interaction dynamics are often driven by attractiveand repulsive forces, mimicking the tendency of individuals to stay close to the group butmaintain a minimal distance. A popular choice among attractive-repulsive potentials isthe Morse potential, that is

W (r) = −CAe−r/lA + CRe−r/lR ,

where CA and CR are the attractive and repulsive strength and lA, lR their respectivelength scales, see [10, 11, 12, 13, 14, 15]. Also power-law potentials of the form

W (r) =|r|a

a− |r|

b

b, a > b,

have been used thoroughly. We refer to the works of [14, 16, 17, 18, 19] for more details.Purely attractive potentials are often of the form W (r) = |r|a, a > 0. In this case the den-sity of particles collapses in finite or in infinite time and converges to a Delta Dirac locatedat the center of mass for certain range of values, see [20, 21, 22, 23]. Finally, aggregation-diffusion equations, in which both the linear or nonlinear diffusion term modelling repulsionand the aggregation term modeling attraction are present, are ubiquitous models in physicsand biology, including all different versions of the parabolic-elliptic Keller-Segel model forchemotaxis, see for instance [24, 25, 26, 27, 28, 29] and the references therein.

2

Page 3: Numerical simulation of nonlinear continuity equations by ...

The gradient flow structure of (1) provides a natural framework to construct solutionsusing variational schemes, usually called the JKO algorithm, as proposed in the seminalworks of Jordan, Kinderlehrer and Otto [30, 31]. The variational formulation providesan underlying structure for the construction of numerical methods, which have inherentadvantages such as build-in positivity and free-energy decrease. There has been an increas-ing interest in the development of such methods in the last years, for example variationalLagrangian schemes such as [32, 33, 34, 35, 27, 36, 37, 38] or finite volume schemes as in[39]. The common challenge of all methods is the high computational complexity, oftenrestricting them to spatial dimension one. They involve for example the solution of theMonge-Kantorovich transportation problem between measures, which require the compu-tation of the Wasserstein distance. In 1D the Wasserstein distance corresponds to theL2 distance between inverses of the distribution functions. In higher space dimensionsits computation involves the solution of an optimal control problem itself, which poses asignificant challenge for the development of numerical schemes. There are few results onthe numerical analysis of these schemes, for example Matthes and co-workers provided firstresults on the convergence in 1D in [36]. Recent works [40] are making use of recent de-velopments in the fast computation of optimal transportation maps to discretize the JKOsteps in the variational scheme.

Many of the Lagrangian methods are based on the “discretize-then-minimize” strategy,see for example [34, 35, 36, 40, 37, 38]. However, we follow closely the advantage of the 1Dformulation in terms of the pseudo-inverse function, see also [32, 33, 27], which correspondsto a “minimize-then-discretize” strategy. In higher dimensions, this strategy is based onthe variational formulation of (1) for a diffeomorphism mapping the uniform density tothe unknown density ρ. Evans et al. proved in [41] that this approach corresponds to aclass of L2-gradient flows of functionals on diffeomorphisms. This equivalent formulationserved as a basis for the numerical solver of Carrillo and Moll [42], which uses the vari-ational formulation and an explicit in time as well as a finite difference discretization inspace. We shall follow their approach but propose an implicit time stepping and a spatialdiscretization, which is based on finite differences in 1D and finite elements in 2D. Wewould like to mention that the our scheme coincides with the implicit in time finite differ-ence discretization of the pseudo-inverse function proposed by Blanchet et al. for the 1DPatlak-Keller-Segel model, see [27]. The finite element discretization allows us to considermore general computational domains as well as triangular meshes. This is advantageousin the case of radially symmetric solutions, which develop asymptotically for large timesfor many types of aggregation-diffusion equations. Due to the implicit in time steppingno CFL condition as in [42] is necessary. Moreover, if our dicretization were convergentfor the implicit semidiscretization of the PDE system satisfied by the diffeomorphisms,the numerical scheme would be convergent for an exact JKO step of the the variationalscheme.

This paper is organized as follows: we review the underlying ideas of the proposedvariational scheme in Section 2. In Section 3 we discuss the numerical algorithm; Section4 details the necessary pre- and postprocessing steps. Finally we present extensive numer-ical simulations for a large class of nonlinear aggregation-diffusion equations in Section 5

3

Page 4: Numerical simulation of nonlinear continuity equations by ...

including blow-up profiles and complicated asymptotic behaviors showing the flexibilityof the method to cope both with diffusive and aggregation behavior in the same model.Section 6 is devoted to discuss the main aspects of our approach together with futureperspectives.

2. Gradient flow formulation

We start by briefly reviewing the variational formulation of (1) as a gradient flow withrespect to a particular distance and energy. Let us consider the quadratic Wassersteindistance between two probability measures µ ∈ P(Rd) and ν ∈ P(Rd) given by

d2W (µ, ν) := inf

T :ν=T#µ

∫Rd|x− T (x)|2dµ(x) ,

for any µ absolutely continuous with respect to Lebesgue, see [43] for a general definitionand its properties. It is well known that solutions of (1) can be constructed via the so-calledJKO scheme, see [30], which corresponds to solving

ρn+1∆t ∈ arg min

ρ∈K

1

2∆td2W (ρn∆t, ρ) + E(ρ)

, (2)

for a fixed time step ∆t > 0 and K = ρ ∈ L1+(Rd) :

∫Rd ρ(x)dx = M, |x|2ρ ∈ L1(Rd).

Hence (2) can be understood as a time discretization of an abstract gradient flow equationin the space of probability measures. It has been proven that solutions of (2) converge tothe solutions of (1) first order in time, see [30, 31, 43, 44] for more detailed results on theanalysis.

Evans et al. [41] showed that there is a connection between the theory of steepestdescent schemes with respect to the Euclidean transport distance and the L2-gradient flowsof polyconvex functionals of diffeomorphisms. Since this formulation serves as a basis ofthe proposed numerical solver, we will review the main results in the following. Let Ω bea smooth, open, bounded and connected subset of Rd and Ω be an open subset of Rd. Let

D denote the set of diffeomorphisms from ¯Ω to Ω, mapping ∂Ω onto ∂Ω. Furthermore weconsider the energy functional:

I(Φ) =

∫Ω

Ψ(detDΦ)dx+

∫Ω

V (Φ(x))dx+1

2

∫Ω×Ω

W (Φ(x)− Φ(y))dxdy,

where Ψ(s) = sU(

1s

). Ambrosio et al [45] clarified even further this subtle relation found

in [41], by showing that the L2-gradient flow of I(Φ) with V = W = 0 given by

Φn+1∆t ∈ arg min

Φ∈D

1

2∆t‖Φn

∆t − Φ‖L2(Ω) + I(Φ)

(3)

is well defined and converges to the solutions of the nonlinear PDE system

∂Φ

∂t= ∇ ·

[Ψ′(detDΦ)(cof DΦ)T

],

4

Page 5: Numerical simulation of nonlinear continuity equations by ...

where DΦ is the Jacobian matrix of Φ and cof DΦ the corresponding cofactor matrix. Thisconnection was generalized subsequently in [42] to the case of interaction and potentialenergies, leading to the nonlinear PDE system

∂Φ

∂t= ∇ ·

[Ψ′(detDΦ)(cof DΦ)T

]−∇V Φ−

∫Ω

∇W (Φ(x)− Φ(y))dy, (4)

The diffeomorphism Φ maps a given reference density, for instance the uniform density ona reference domain Ω, to the unknown density ρ in Ω. Therefore the density ρ ∈ K canbe calculated via ρ = Φ#LN for every diffeomorphism Φ ∈ D, assuming that |Ω| = M ,where LN denotes the N -dimensional Lebesgue measure on the reference domain Ω. Anequivalent formulation in the case of a sufficiently smooth diffeomorphism Φ is given by

ρ(Φ(x)) det(DΦ(x)) = 1. (5)

Hence we can interpret equation (4) as the Lagrangian representation of the original non-linear continuity equation (1) in Eulerian coordinates.

3. Spatial and temporal discretization of the Lagrangian representation

In this section we present the details of the spatial and temporal discretization of equa-tion (4), focusing on the implicit in time scheme and the subsequent finite dimensionalapproximation of the nonlinear semi-discrete equations. The latter is discretized by a fi-nite difference scheme in 1D and a finite element method in 2D. In the following we shallonly present the finite element discretization of (4), since the 1D finite difference schemeis detailed in [27, 42] already.

Let ∆t denote the discrete time step, tn+1 = (n+1)∆t and Φn+1 the solution Φ = Φ(t, x)at time tn+1. Then the implicit in time discretization of (4) reads as

Φn+1 − Φn

∆t=∇ · [Ψ′(detDΦn+1)(cof DΦn+1)]

−∇V (Φn+1)−∫

Ω

∇W (Φn+1(x)− Φn+1(y))dy. (6)

Let us consider test functions ϕ = ϕ(x) ∈ H1(Ω). Then the nonlinear operator F definedvia (6) is given by:

F (Φ, ϕ) =1

∆t

∫Ω

(Φn+1 − Φn)ϕ(x)dx+

∫Ω

Ψ′(detDΦn+1)(cof DΦn+1)∇ϕ(x)dx

+

∫Ω

∇V (Φn+1)ϕ(x)dx+

∫Ω

[∫Ω

∇W (Φn+1(x)− Φn+1(y))dy

]ϕ(x)dx.

(7)

We use lowest order H1 conforming finite elements, also known as hat functions, for thediscretization of Φ, more precisely piece-wise linear functions. The discrete diffeomorphism

5

Page 6: Numerical simulation of nonlinear continuity equations by ...

Φh = Φh(x1, x2) can be written as

Φh(x1, x2) =∑j

(Φ1j

Φ2j

)ϕj(x1, x2),

where the index j corresponds to the nodal degrees of freedom. We solve the nonlinearoperator equation F (Φ, ϕ) = 0 using a Newton Raphson method in every time step (anddrop the subscript h to enhance readability in the following). To do so, we calculate theJacobian matrix DF of (7) and determine the Newton update Υn+1,k+1 via

DF (Φn+1,k, ϕ)Υn+1,k+1 = −F (Φn+1,k, ϕ),

for all test functions ϕ(x) ∈ H1(Ω). The index k corresponds to the Newton iteration andn to the temporal discretization. Note that the Jacobian matrix Υn+1,k+1 is a full matrixand has no sparse structure due to the convolution operator W . The Newton updates arecalculated via

Φn+1,k+1 = Φn+1,k + αΥn+1,k+1,

where α is a suitable damping parameter. The Newton iteration is terminated if

|F (Φn+1,k+1, ϕ)| ≤ ε1 or ‖Φn+1,k+1 − Φn+1,k‖ ≤ ε2,

with given error bounds ε1 and ε2. Note that due to the implicit in time discretization noCFL type condition for the time step ∆t, as in [42], is necessary.

The presented L2-gradient flow (3) is only valid if the total mass is conserved. Thiscan be implemented by adopting the image domain in 1D or by considering equation (1)with no flux boundary conditions in 2D, i.e.

v · n = 0 on ∂Ω,

where n denotes the outer unit normal vector on the boundary ∂Ω and v is given by (1).We shall only consider diffeomorphisms which map ∂Ω onto ∂Ω without rotations as in[42]. Then the corresponding natural boundary conditions for the equation in Lagrangianformulation are given by

nT (cof DΦ)T∂Φ

∂t= (cof DΦ)n · ∂Φ

∂t= 0. (8)

Note that the boundary conditions (8) have to be checked separately for every computa-tional domain, see for example [42] for the discussion of appropriate boundary conditionsin the case of the unit square. In the case of a circle with radius R we have:

1

R

(x1

x2

)( ∂Φ2

∂x2−∂Φ1

∂x2

−∂Φ2

∂x1

∂Φ1

∂x1

)(∂Φ1

∂t∂Φ2

∂t

)= 0.

6

Page 7: Numerical simulation of nonlinear continuity equations by ...

Using radial coordinates x1 = R cos θ and x2 = R sin θ, we obtain:

cos θ(∂Φ1

∂t

∂Φ2

∂x2

− ∂Φ2

∂t

∂Φ1

∂x2

) + sin θ(−∂Φ1

∂t

∂Φ2

∂x1

+∂Φ2

∂t

∂Φ1

∂x1

) = 0.

Since ∂Φ1

∂θ= − sin θ ∂Φ1

∂x= cos θ ∂Φ1

∂x2, we conclude that

sin θ∂Φ2

∂t

∂Φ1

∂x1

+ cos θ∂Φ1

∂t

∂Φ2

∂x2

= 0. (9)

Hence, in the case of a circle equation (9) is equivalent to (8).If we assume that the boundary of Ω is mapped onto the boundary of Ω, the correspondingdiffeomorphism is given by

Φ1(t, x) = Φ2(t, x) = Id,

which implies that ∂Φ1

∂t= ∂Φ2

∂t= 0 on the boundary. This choice ensures that the equivalent

boundary conditions given by (9) are satisfied.

4. Pre-processing and post-processing: calculating the initial diffeomorphismand the final density

The discretization of the L2-gradient flow formulation involves several pre- and post-processing steps. First the initial diffeomorphism Φ0 has to be computed for a given initialdensity ρ0. The postprocessing step corresponds to calculate the density ρT from the finaldiffeomorphism ΦT . We shall detail these two steps in the following.

4.1. Pre-processing:

Let ρ0(x) be a given smooth initial density with∫

Ωρ0(x)dx = M and denote by Φ0 =

Φ0(x) := Φ(0, x) the corresponding diffeomorphism. Then the initial diffeomorphism Φ0 :¯Ω→ Ω satisfies

ρ0(Φ0(x)) detDΦ0(x) = 1. (10)

Depending on the discretization of Ω different approaches can be used to determine Φ0(x)from a given ρ0(x). In the case of an equidistant mesh of squares or rectangles, onecan calculate the initial diffeomorphism by solving a one-dimensional Monge-Kantorovichproblem in x1 direction and subsequently a family of Monge-Kantorovich problems in thex2 direction, cf. [46, 42].

However this approach is not possible in case of general quadrilateral or triangularmeshes. In this case different strategies can be used: either based on the Monge Ampereequation (giving the optimal transportation plan in case of quadratic cost), Knote theoryor density equalizing maps. We shall follow the latter, which is based on an idea ofMoser [47] that was further studied by Avinyo and co-workers in [48]. It is based on the

7

Page 8: Numerical simulation of nonlinear continuity equations by ...

idea of constructing the initial diffeomorphism from solutions of the heat equation withhomogeneous Neumann boundary conditions. This approach was also used in cartographyto calculate density equalizing maps, cf. [49]. Its advantage is its flexibility - it can beused for general computational domains and their respective discretizations since it onlyrequires the efficient solution of the heat equation.

Consider the heat equation written as a continuity equation on a bounded domainΩ ⊂ R2 as

∂ρ

∂t+ div(ρv) = 0, with v = −∇ρ

ρ, (11)

and initial datum ρ(0, x) = ρ0(x) as well as homogeneous Neumann boundary conditions.Equation (11) corresponds to the time evolution of ρ0 transported by the velocity fieldv = −∇ρ

ρtowards the constant density ρ = 1

|Ω|

∫Ωρ0(x)dx as t → ∞. Hence the velocity

field can be calculated by solving the heat equation until equilibration. Then the cumulativedisplacement x(t) of any point at time t is determined by integrating the velocity field,which corresponds to solving

x(t) = x(0) +

∫ t

0

v(t′,x(t′)) dt′.

As t → ∞, the set of such displacements for all points x = x(t) in Ω, that is the gridpoints of the computational mesh, defines the new density-equalized domain. Note thatwe actually need to determine its inverse, since we need to find the map which maps theconstant density to the initial density ρ0. Hence we solve:

x′(t) = v(t′,x(t′)) = −∇ρ(t,x(t))

ρ(t,x(t))(12a)

x(T ) = x1, (12b)

for all mesh points x1 ∈ Ω. This corresponds to the solution of the integral equationx(0) = x(T )−

∫ T0v(t,x(t))dt. Altogether the pre-processing consists of two steps:

1. Solve the heat equation with Neumann boundary conditions for an initial datumρ0 = ρ0(x) until equilibration. The time of equilibration corresponds to the timewhere the L2−norm between two consecutive time steps is less than 10−6.

2. Starting with a given mesh at time t = T calculate the initial diffeomorphism bysolving (12) backward in time. Then the initial diffeomorphism is given by Φ0(x1) =x(0).

Remark 4.1. Note that the one could also use the initial density ρ0 as a reference measure,then the initial diffeomorphism φ0 is the identity map. In this case equation (5) would readas

ρ(Φ(x)) det(DΦ(x)) = ρ0(x).

8

Page 9: Numerical simulation of nonlinear continuity equations by ...

4.2. Post-processing:

The post-processing step corresponds to calculate ρT := ρ(t = T, x) given the finaldiffeomorphism ΦT := Φ(t = T, x). Since we expect numerical artifacts in case of compactlysupported and measure valued solutions we solve a regularized version of (5) (in 2D) givenby:

ε∆ρT (ΦT (x)) + ρT (ΦT (x)) =1

detDΦT (x)with 0 < ε 1. (13)

Note that equation (13) can be easily implemented using finite elements, since the assem-bling of the system matrix is based on the transformation of the triangle to the referencetriangle. In case of equation (13) we have to perform two successive transformations. Firstthe transformation to the displaced element given by Φ and second to the reference element.

5. Numerical simulations

In this section we illustrate the behavior of the numerical solver with simulations inspatial dimension one and two. The 1D simulations are based on a finite difference dis-cretization, in 2D we use finite elements. The respective solvers are based on Matlab in1D and on the finite element package Netgen/NgSolve in 2D. In the following we illustratethe flexibility of our approach with various simulations for a large class of PDEs. We startby summarizing all steps of the numerical solver in Algorithm 1.

5.1. Numerical simulations in spatial dimension one

If not stated otherwise we initiate all numerical simulations with a Gaussian of theform:

ρ0(x) =M√2πσ2

e−x22σ2 (14)

with M,σ > 0. Note that∫∞−∞ ρ0(x)dx = M . We perform our simulations on a bounded

domain Ω = [0, M ], where M is the approximate value of∫

Ωρ0(x)dx. Hence, the initial

diffeomorphism Φ0 is defined on Ω.

5.1.1. Nonlinear diffusion equations

In our first example we illustrate the behavior of our scheme for the Porous mediumequation, that is U(s) = 1

m−1sm,m > 1 and V = W = 0. Let Ω = [−1, 1] denote the

computational domain discretized into 501 intervals of size ∆x = 2501

. The initial densityis given by a Barenblatt Pattle profile (BPP) at time t0 = 10−3, i.e.

ρ0(x) =1

tα0

(c− αm− 1

2m

x2

t2α0

) 1m−1

+

,

where α = 1m+1

and c is chosen such that∫ 1

−1ρ0(x) dx = 2. The discrete time steps are

set to ∆t = 10−5. We consider the cases m = 2 and m = 4. Figures 1 and 2 show the

9

Page 10: Numerical simulation of nonlinear continuity equations by ...

Algorithm 1 Let ρ0 = ρ0(x) be a given initial density with∫

Ωρ0(x)dx = M and Ω = [a, b]

or Ω = [a, b]× [c, d].

1. Determine the initial diffeomorphism

(a) 1D: Define the initial diffeomorphism Φ0 : Ω = [0,M ] → [a, b] by: for everydiscretization point xi ∈ [0,M ] solve the one-dimensional Monge-Kantorovichproblem for Φi = Φ(xi): ∫ Φi

a

ρ0(y)dy = xi,

using Newton’s method.(b) 2D: Follow the two-step algorithm outlined in Section 4.1: Solve the heat equa-

tion ∂tρ = ∆ρ with homogeneous Neumann boundary conditions using an H1

conforming finite element method of order p = 6 until equilibration at timet = T . Determine the initial diffeomorphism by transporting each mesh pointof the computational mesh V (T ) = (xV1 (T ),xV2 (T )) ∈ V backward in time

x(0) = x(T )−∫ T

0

v(t,x(t))dt, v(t,x(t)) = −∇ρ(t,x(t))

ρ(t,x(t)).

2. At every time step tn+1 = (n+ 1)∆t in t ∈ (0, T ] solve

(a) 1D: the implicit equation (6) using Newton’s method and finite difference dis-cretization.

(b) 2D: the nonlinear operator equation F (Φ, ϕ) = 0, given by (7) using Newton’smethod and a spatial discretization of H1 conforming finite elements of orderp = 1.

3. Recover the final density ρ = ρ(T ) by

(a) 1D: calculating the final density at every discrete point Φi = Φ(xi) from therelation

ρ(Φ(x)) =1

detDΦ(x).

(b) 2D: solving the regularized equation

ε∆ρ(Φ(x)) + ρ(Φ(x)) =1

detDΦ(x)

on the transformed mesh N using H1 conforming elements of order p = 4.

10

Page 11: Numerical simulation of nonlinear continuity equations by ...

density profiles ρ at time t = 0.021 as well as the evolution of the free energy in time, whichdecays like t−α(m−1), cf. [50]. As seen from Figures 1b and 2b, these decays (indicated bythe green lines) are perfectly captured by the scheme validating the chosen discretization.Note that the boundary points of the support are calculated using an explicit one-sideddifference scheme of the velocity as proposed by Budd et al., see [51]. In particular, writingthe porous medium equation as

ρt +∇ · (ρv) = 0

with the velocity v = − mm−1

(ρm−1)x leads to the approximation Φn+1 = Φn + ∆tv at theboundary points. When determining ρx we assume that the density at the boundary ofthe support equals zero.

−2 −1.5 −1 −0.5 0 0.5 1 1.5 20

1

2

3

4

5

6

Computed solution

Initial data

BPP at t=0.021

(a) Reconstructed BP profile ρ = ρ(x, t) attime t = 0.021.

10−3 10−2 10−1 10010−1

100

101

102

Time t

E

(b) Evolution of the entropy and the ex-pected decay of t−α in time with a logarith-mic scale.

Figure 1: Solution of the PME for m = 2 after 2000 time steps.

5.1.2. Fokker-Planck type equations

Next we study the behavior of a nonlinear Fokker-Planck equation in the case of an internalenergy U(s) = ν s

m

m, m > 1 and a double-well confining potential V (x) = x4

4− x2

2. Let

Ω = [−2, 2] be the computational domain discretized into 501 intervals of size ∆x = 4501

.The initial density is given by (14) with σ = 0.2 and M = 1. The discrete time steps areset to ∆t = 10−3. In this case steady states are given by

ρ∞(x) =1

ν(C(x)− V (x))

1m−1

+ , (15)

where C = C(x) is a piecewise constant function possibly taking different values on eachconnected part of the support, cf. [39]. The density at time T and the entropy decay withrespect to ρ∞ are illustrated in Figure 3. We always choose T in Section 5.1 as the timewhere the L2−norm of the difference of two consecutive diffeomorphisms is less than 10−6

and refer to ρ at time T + t0 for t0 = 500∆t as the steady state ρ∞. We observe that the

11

Page 12: Numerical simulation of nonlinear continuity equations by ...

−2 −1.5 −1 −0.5 0 0.5 1 1.5 20

0.5

1

1.5

2

2.5

3

Computed solution

Initial data

BPP at t=0.021

(a) Reconstructed BP profile ρ = ρ(x, t) attime t = 0.021.

10−3

10−2

10−1

100

10−2

10−1

100

101

102

Time t

E

(b) Evolution of the entropy and the ex-pected decay of t−3α in time with a loga-rithmic scale.

Figure 2: Solution of the PME for m = 4 after 2000 time steps.

(a) Density ρ at time T = 9.718.

0 1 2 3 4 5 6 7 8 910

−8

10−7

10−6

10−5

10−4

10−3

10−2

10−1

100

E(ρ)−E(ρ∞)

Time t

(b) Evolution of the entropy in time with alogarithmic scale for the y-axis.

Figure 3: Solution of the nonlinear Fokker-Planck equation for m = 2 and ν = 0.05.

stationary state has two connected components. As seen in Figure 3b, the entropy decaysabruptly once the support separates into two pieces before final convergence towards thesteady state with the lowest free energy filling with equal mass each of the two wells of thepotential, i.e., C = C(x) is equal in each connected component.

If the confining potential is chosen as the harmonic potential V (x) = x2

2, the steady

state is also given by (15), cf. [52]. The convergence towards the steady state is exponen-tial, more precisely Carrillo et al. [52] showed that the distance towards equilibrium, i.e.d2W (ρ(t), ρ∞), converges like O(e−(m+1)t) since the initial data has zero center of mass. The

same convergence behavior can be observed in the numerical simulations, see Figure 4a

12

Page 13: Numerical simulation of nonlinear continuity equations by ...

and 4b. Note that the rate of convergence in relative energy in Figure 4b coincides withthe rate of convergence of d2

W (ρ(t), ρ∞).

−2 −1.5 −1 −0.5 0 0.5 1 1.5 20

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

2 Computed solutionInitial dataCorrect Solution ρ∞

(a) Density ρ at time T = 3.172.

1.4 1.6 1.8 2 2.2 2.4 2.6 2.810

−7

10−6

10−5

10−4

10−3

10−2

10−1

100

E(ρ)−E(ρ∞)

Time t

N=101N=301N=501line of slope −3

(b) Semi-log error plot at time t for different

values for N = |Ω|∆x with respect to the steady

state.

Figure 4: Solution of the nonlinear Fokker-Planck equation for m = 2 and ν = 2.

5.1.3. Aggregation equations

Next we consider aggregation potentials of the form

W (x) =|x|a

a− |x|

b

b, (16)

for a > b ≥ 0 using the convention that |x|0/0 = ln |x|. The set of stationary states ofthese equations can be quite complex, see [14, 16, 17], depending on the dimension. In onedimension with a = 2 and b = 1 the steady state profiles are constant on an interval, cf.[18, 19], for a = 2, b = 0 they correspond to the semicircle law, cf. [53, 54].Let a = 2, b = 1 and set Ω = [−2, 2] split into 201 intervals of size ∆x = 4

201. The initial

datum is given by (14) with σ = 0.35 and M = 4. The discrete time steps are set to∆t = 10−3. Figure 5 shows the computed solution at time T and the entropy decay withrespect to ρ∞. The numerical simulations confirm the theoretical results, i.e. the computedstationary profile corresponds to the constant density ρ∞(x) = 2 for all x ∈ [−1, 1]. Notethat we use the mid-point rule to calculate the convolution integral in (6) on the boundaryelements, as proposed in [39].

In the case a = 2, b = 0 we set Ω = [−6, 6] split into 501 intervals of size ∆x = 12501

.Then the unit-mass steady state is given by

ρ∞(x) =

√2− x2, |x| ≤

√2,

0 otherwise.

13

Page 14: Numerical simulation of nonlinear continuity equations by ...

−2 −1.5 −1 −0.5 0 0.5 1 1.5 20

0.5

1

1.5

2

2.5

3

3.5

4

4.5

5

Computed solution

Initial data

(a) Density ρ at time T = 2.412.

0 0.5 1 1.5 210

−8

10−6

10−4

10−2

100

102

Time t

E(ρ)−E(ρ∞)

(b) Evolution of the entropy and the decay ofe−7.5t in time with a logarithmic scale for they-axis.

Figure 5: Solution of the aggregation equation with potential (16) (a = 2, b = 1).

−3 −2 −1 0 1 2 30

0.5

1

1.5Computed solutionInitial dataCorrect Solution ρ∞

(a) Density ρ at time T = 4.891.

0 0.5 1 1.5 2 2.5 3 3.5 4 4.510

−10

10−8

10−6

10−4

10−2

100

102

Time t

E(ρ)−E(ρ∞)

(b) Evolution of the entropy and the decayof e−3t in time with a logarithmic scale forthe y-axis.

Figure 6: Solution of the aggregation equation with potential (16) (a = 2, b = 0).

This result is confirmed by the numerical simulations illustrated in Figure 6 for an initialdatum ρ0 given by (14) with σ = 1 and M = 1. It is known that the solutions decayin dW toward the corresponding stationary states ρ∞ exponentially fast. We observe thatbehavior at the level of the relative energy with numerical rates of decay given by 7.5 and3 approximately in Figures 5b and 6b.

5.1.4. Modified Keller-Segel model

In our final example we consider a modified Keller-Segel (KS) model. It is well known thatsolutions of the classical Keller-Segel may blow up in finite time in space dimensions d ≥ 2,

14

Page 15: Numerical simulation of nonlinear continuity equations by ...

see cf. [55, 27]. The same behavior can be observed in 1D by using the corresponding 2Dinteraction potential instead. Then the modified KS model reads as:

∂tρ = ∇ · (ρ(∇(ln ρ+W ∗ ρ))),

W =χ

dπln |x|,

ρ(x, 0) = ρ0 ≥ 0.

(17)

The blow-up behavior depends on the initial mass M0 =∫

Ωρ0(x)dx. If M0 < Mc, where

χMC = 2d2π denotes the critical mass, the system has a global in time solution. IfM0 > Mc

solutions blow up in finite time.Let Ω = [−6, 6] be the computational domain discretized into 501 intervals of size

∆x = 12501

. The initial density is given by (14) with σ = 1 and the discrete time steps areset to ∆t = 10−1. If the initial data has a smaller mass, in our case M0 = 2π − 0.1 thesolution diffuses to zero, see Figure 7. If the M0 > MC the solution blows up as illustratedin Figure 8 in the case of M0 = 2π + 0.1.

−400−200

0200

400

0

0.5

1

1.5

2

xTime t

solution

t=0t=2000t=4000t=8000t=100000

Figure 7: M = 2π − 0.1 < Mc. Evolution of the density ρ for χ = 1.

−10−5

05

10

0

10

20

30

40

50

60

xTime t

solution

t=0t=3t=6t=9t=12

−50

5

0

5

10

15

x 109

xTime t

solution

t=t*+33e−14t=t*+51e−14t=t*+57e−14t=t*+58e−14

Figure 8: M = 2π + 0.1 > Mc. Evolution of the density ρ for χ = 1, where t∗ = 16.818399684193.

15

Page 16: Numerical simulation of nonlinear continuity equations by ...

Next we study the behavior of (17) for two initial densities corresponding to the sumof two Gaussians, i.e.

ρ10(x) =

√600

(2π − 0.1√

2πe

−600(x−2)2

2 +2π − 0.5√

2πe

−600(x+2)2

2

),

and

ρ20(x) =

√600

(2π + 0.1√

2πe

−600(x−2)2

2 +2π − 0.5√

2πe

−600(x+2)2

2

),

on the computational domain Ω = [−5, 5] discretized into 301 intervals of size ∆x = 10301

with ∆t = 10−3 and ∆t = 10−4, respectively. The total mass is 4π − 0.6 in the firstsimulation and 4π − 0.4 in the second one, then they correspond to supercritical massesin which the blow-up will eventually happen. Figure 9 illustrates that if both peaks haveinitial masses smaller than the critical one, they initially diffuse while moving towards eachother until they accumulate enough mass to blow up at the center of mass. Note that theblow-up time is not included in the figure. But if one of them has an initial mass abovethe critical value, the blow up happens in the center of mass of the corresponding peakbefore getting closer to the center of mass, see Figure 10, again the final time does notcorrespond to the blow-up time.

−5−2

02

5

0

50

100

150

200

xTime t

solution

t=0t=0.3t=1.5t=1.8

−5 0 50

20

40

60

80

100

120

140

160

180

200

220Computed solutionInitial data

Figure 9: Evolution of the density ρ for χ = 1 with initial datum ρ0 = ρ10(x) up to time t = 1.8.

In case M > Mc, it is conjectured that the first blow-up should happen by the formationof a Dirac Delta with exactly the critical mass Mc = 2π. Moreover, a detailed asymptoticexpansion near the blow-up time for the blow-up profile was performed in [24]. The profileis not blowing-up selfsimilarly and its validity was numerically checked in [56] for twodimensions in a numerical tour-de-force. Checking the blowup profile in time for theKeller-Segel model is quite challenging for us too since the density is a derived quantityonly obtained from the diffeomorphism through (10)-(13) by numerical differentiation.Therefore, numerical errors are likely to happen when reconstructing the density profile.

However, we can now compare our computed diffeomorphism to the diffeomorphismrepresenting the zeroth order term in the expansion studied by [24]. This expansion does

16

Page 17: Numerical simulation of nonlinear continuity equations by ...

−5

0

5

0

200

400

600

800

xTime t

solution

t=0t=0.01t=0.02t=0.03

−5 0 50

100

200

300

400

500

600

700

800 Computed solutionInitial data

Figure 10: Evolution of the density ρ for χ = 1 with initial datum ρ0 = ρ20(x) up to time t = 0.03.

not carry over directly to our modified KS system (17). Nevertheless, it is easy to checkfollowing [24] that the zeroth order term should be given by a properly scaled stationarystate of the critical mass case. More precisely, the solution near the blow-up time shouldbehave like

ρ(t) =2γ(t)1/2

1 + γ(t)x2with γ(t) ' ρ(t, 0)2

in an interval of length 2L(t) centered at the point of blow-up with L(t)−1 ' ρ(t, 0) accord-ing to [24, 56]. In order to check this behavior, we try to avoid comparisons of the densitiesdue to the numerical errors mentioned above and we concentrate in estimating the errorbetween the diffeomorphisms Φ(t) and Ψ(t) representing ρ(t) and ρ(t) respectively. Thelast one is computed by calculating the diffeomorphism representing the steady solution tothe critical mass case 2(1 + x2)−1 and then multiplying by the dilation factor γ(t)−1/2.

We take as initial data a Gaussian already very picked at the origin, i.e. (14) withσ2 = 5 × 10−6 and M = 2π + 0.1 for which we have numerical blow-up quite quickly.Let Ω = [−0.1, 0.1] be the computational domain discretized into 1200 intervals of size∆x = 0.2

1200. The discrete time steps are set to ∆t = 10−9. Then, we compute the blow-

up time T = 3.328e − 6 as the one for which our Newton’s iterations do not converge.Once the blow-up time is approximated, we compute the errors in Wasserstein distancebetween ρ(t) and ρ(t) on the interval (−L(t), L(t)) and we can compare the correspondingdiffeomorphisms at the blow-up time. Figure (11)(a) shows the evolution in time of theL2-difference between Φ(t) and Ψ(t) over the mass interval corresponding to (−L(t), L(t))close to blow-up time. Figure (11)(b) shows the comparison between Φ(T ) and Ψ(T ) bothover the mass interval of length 2π and the zoom over the mass interval corresponding to(−L(T ), L(T )).

5.2. 2D simulations

Next we present several numerical simulations in 2D. Let Ω denote the unit circle, M =⋃V ∈V xV the computational mesh with vertexes xV , V = 1, . . . nV and N =

⋃V ∈V Φ(xV )

17

Page 18: Numerical simulation of nonlinear continuity equations by ...

2.6 2.7 2.8 2.9 3 3.1 3.2 3.3 3.4

x 10−6

0

1

2

3

4

5

6x 10−12

Time

L2er

roro

fΦ−Ψ

(a) Evolution of the error in dW between ρ(t)and ρ(t) near the blow-up time.

0.0532 3.1916 6.33Mass variable

Diff

eom

orph

ism

ΦΨ

2.2607 3.1916 4.1225

−L(T)

0

L(T)

Mass variable

Diff

eom

orph

ism

ΦΨ

(b) Comparison of the diffeomorphisms Φ and Ψ at the blow-up time T . Inset shows a zoom at the center of mass.

Figure 11: Comparison of the evolution of Φ(t) and Ψ(t).

the transformed mesh. We choose the following parameters if not stated otherwise:

• ∆t = 0.01

• nV = 1758 which corresponds to 3386 triangles of maximum size hmax = 0.05.

The initial datum is given by Gaussian centered around the origin

ρ0(x) =1

2πσ2e−

12σ2

(x2+y2) + c with σ = 0.3,

where c is chosen such that∫

Ωρ0(x)dx = 1. In the pre-processing step we solve the heat

equation on the time interval t ∈ (0, 2] using an implicit in time discretization and H1-conforming elements of order 6. The solution is computed at ti = 0.002i, i = 1, . . . , 1000and used for the computation of the initial diffeomorphism. The error bounds for theNewton schemes are set to ε1 = ε2 = 10−6, the regularization parameter ε in the post-processing step to ε = 10−2. All solutions are computed as detailed in Algorithm 1.

5.2.1. Attraction potentials

First we consider an attraction potential of harmonic type W (x) = 12|x|2. In the case of a

purely attractive potential we expect the formation of a Delta Dirac at the center of thedomain. Figure 12 illustrates the formation of the blow up. In the first picture we seethe transformed mesh, where all triangles are moved towards the center of the domain.The second and third picture show the reconstructed density profile ρ as well as the decayof the distance towards the Dirac Delta in time. The numerical simulations confirm thetheoretical results, indicated by the red line of slope −1, at the beginning of the simulation.

18

Page 19: Numerical simulation of nonlinear continuity equations by ...

The bad match towards the end results most likely from the formation of the Delta Diracand the inaccuracy caused by it.

(a) Transformed mesh (b) Density ρ (c) dW (ρ(t), δ0).

Figure 12: Simulation results in the case of a purely attractive potential W = 12 |x|

2.

5.2.2. Attraction-repulsion potentials

Next consider the attraction-repulsion potential (16). In the case a = 4 and b = 2 thesolution concentrates on a ring of radius r = 1

2, see [21, 22]. The formation of the Delta

(a) Transformed mesh (b) Density ρ (c) Relative entropy

Figure 13: Simulation results for an attractive-repulsive potential W = 14 |x|

4 − 12 |x|

2.

Dirac ring is clearly visible in Figure 13. The first picture corresponds to the transformedinitial mesh, the second and third show the reconstructed density and the decay of therelative entropy respectively. Note that we map the boundary of the computational domainonto itself, hence the boundary nodes do not move.

In the case of a logarithmic repulsion, that is a = 2 and b = 0 in (16) we expect theformation of a compactly supported steady state circle with radius r = 1

2. We see that

the vertexes of the mesh concentrate in a circle of radius r = 12, except for the boundary

nodes which are fixed, see Figure 14. Again the second and third plot correspond to thereconstructed density ρ at time t = 2 and the decay of the relative entropy functional.

19

Page 20: Numerical simulation of nonlinear continuity equations by ...

(a) Transformed mesh (b) Density ρ (c) Entropy

Figure 14: Simulation result for an attractive-repulsive potential W = 12 |x|

2 − ln(|x|).

Next we consider the agreggation equation with the logarithmic repulsive potential andharmonic confinement, a = 2 and b = 0 in (16), with an additional external potential ofthe form V (x) = −α

βln(|x|). This model has been proposed as a way to find the spatial

(a) Transformed mesh (b) Density ρ (c) Relative entropy

Figure 15: Simulation results for an attractive-repulsive potential W = 12 |x|

2 − ln(|x|) and an additionalpotential V = 1

4 ln(|x|).

shape of the milling profiles in microscopic models for the dynamics of bird flocks [57]. Weobtain this steady state also in the numerical simulations, see Figure 15. The inner andouter radius of this mill depend on the parameters α and β and are given

Ri =

√α

βand Ro =

√α

β+

1

2π.

see [39]. The transformed mesh clearly shows the formation of the steady state annuluswith α = 1 and β = 4. Note that the ’vacuum formation’ at the center distorts the meshat the center. Note that the distortion of the triangles in Figure 15 does not affect theperformance of the numerical method itself, since all computations are done using the

20

Page 21: Numerical simulation of nonlinear continuity equations by ...

original mesh on the reference domain Ω. It only affects the reconstruction of the finaldensity ρ, where we observe numerical instabilities close to the support of the steady stateannulus. Therefore we choose to solve a regularized version of (5).If we start with a not radially symmetric initial datum of the form

ρ0(x) =1

4π0.252e−

120.252

((x−0.15)2(y−0.25)2) +1

4π0.22e−

120.22

((x+0.3)2+(y+0.4)2)

and again consider a potential of the form (16) with a = 2, b = 0, we observe the formationof an off-centered compactly supported bump, see Figure 16.

(a) Transformed mesh (b) Density ρ (c) Relative entropy

Figure 16: Simulation results for an attractive-repulsive potential W = 12 |x|

2 − ln(|x|) in case of a notradially symmetric initial datum.

5.2.3. The Keller-Segel model

In our final example we consider the modified KS model (17) with an initial Gaussianof mass one and χ = 1.1× 8π. The time steps are set to ∆t = 5× 10−3 and we observe thefast formation of a Delta Dirac at the center of the domain, see Figure 17. Figure 17(c)indicates that the free energy decay is changing concavity as the free energy tries to decayfaster possibly tending to −∞ at the blow-up time.

6. Conclusion

In this paper we propose a numerical algorithm for nonlinear PDEs, which can be writtenas gradient flows with respect to the Wasserstein distance. The method is based on itsgradient flow formulation with respect to the Euclidean distance, corresponding to theLagrangian representation of the original nonlinear PDE. The construction of the solverguarantees the preservation of structural features, such as entropy decay or positivity ofsolutions. We presented extensive numerical simulations 1D, which illustrated the flexibilityof our approach and confirmed theoretical results concerning entropy decay and blow upbehavior. Even though these numerical simulations confirmed the predicted convergencebehavior towards the steady state, the numerical analysis is still an open problem (even in

21

Page 22: Numerical simulation of nonlinear continuity equations by ...

(a) Transformed mesh (b) Density ρ (c) Relative entropy

Figure 17: Simulation results for KS model with initial M = 1 and χ = 1.1× 8π.

1D). Hence we plan to investigate the numerical analysis of the proposed scheme in spatialdimension one and as a next step study the method for radially symmetric solutions.

The 2D algorithm involves several pre- and post-processing steps, which present addi-tional challenges with respect to numerical accuracy and computational complexity. Theproposed finite element discretization allows to consider more general domains and betterresolve features of radially symmetric solutions, which often arise in aggregation equations.However the computational complexity of the preprocessing step as well as the solver itselfremains a key limitation of the solver, which we plan to address in a future work.

Acknowledgments

JAC was partially supported by the Royal Society via a Wolfson Research Merit Award.HR and MTW acknowledge financial support from the Austrian Academy of Sciences OAWvia the New Frontiers Group NSP-001. The authors would like to thank the King AbdullahUniversity of Science and Technology for its hospitality and partial support while preparingthe manuscript.

References

[1] A. Mogilner, L. Edelstein-Keshet, A non-local model for a swarm, Journal of Mathe-matical Biology 38 (6) (1999) 534–570.

[2] C. M. Topaz, A. L. Bertozzi, Swarming patterns in a two-dimensional kinematic modelfor biological groups, SIAM J. Appl. Math. 65 (1) (2004) 152–174.

[3] C. M. Topaz, A. L. Bertozzi, M. A. Lewis, A nonlocal continuum model for biologicalaggregation, Bull. Math. Biol. 68 (7) (2006) 1601–1623.

[4] M. R. D’Orsogna, Y.-L. Chuang, A. L. Bertozzi, L. S. Chayes, Self-propelled particleswith soft-core interactions: patterns, stability, and collapse, Physical review letters96 (10) (2006) 104302.

22

Page 23: Numerical simulation of nonlinear continuity equations by ...

[5] T. Kolokolnikov, J. A. Carrillo, A. Bertozzi, R. Fetecau, M. Lewis, Emergent behaviourin multi-particle systems with non-local interactions [Editorial], Phys. D 260 (2013)1–4.

[6] J. A. Carrillo, M. Fornasier, G. Toscani, F. Vecil, Particle, kinetic, and hydrody-namic models of swarming, in: Mathematical modeling of collective behavior in socio-economic and life sciences, Model. Simul. Sci. Eng. Technol., Birkhauser Boston, Inc.,Boston, MA, 2010, pp. 297–336.

[7] J. A. Carrillo, R. J. McCann, C. Villani, et al., Kinetic equilibration rates for granularmedia and related equations: entropy dissipation and mass transportation estimates,Revista Matematica Iberoamericana 19 (3) (2003) 971–1018.

[8] G. Toscani, One-dimensional kinetic models of granular flows, M2AN Math. Model.Numer. Anal. 34 (6) (2000) 1277–1291.

[9] D. D. Holm, V. Putkaradze, Aggregation of finite-size particles with variable mobility,Phys. Rev. Lett. 95 (2005) 226106.

[10] A. J. Leverentz, C. M. Topaz, A. J. Bernoff, Asymptotic dynamics of attractive-repulsive swarms, SIAM J. Appl. Dyn. Syst. 8 (3) (2009) 880–908.

[11] A. J. Bernoff, C. M. Topaz, A primer of swarm equilibria, SIAM J. Appl. Dyn. Syst.10 (1) (2011) 212–250.

[12] J. A. Carrillo, S. Martin, V. Panferov, A new interaction potential for swarmingmodels, Phys. D 260 (2013) 112–126.

[13] J. A. Carrillo, Y. Huang, S. Martin, Explicit flock solutions for Quasi-Morse potentials,European J. Appl. Math. 25 (5) (2014) 553–578.

[14] T. Kolokolnikov, H. Sun, D. Uminsky, A. L. Bertozzi, Stability of ring patterns arisingfrom two-dimensional particle interactions, Phys. Rev. E 84 (2011) 015203.

[15] G. Albi, D. Balague, J. A. Carrillo, J. von Brecht, Stability analysis of flock andmill rings for second order models in swarming, SIAM J. Appl. Math. 74 (3) (2014)794–818.

[16] D. Balague, J. A. Carrillo, T. Laurent, G. Raoul, Nonlocal interactions by repulsive-attractive potentials: radial ins/stability, Phys. D 260 (2013) 5–25.

[17] D. Balague, J. A. Carrillo, T. Laurent, G. Raoul, Dimensionality of local minimizersof the interaction energy, Arch. Ration. Mech. Anal. 209 (3) (2013) 1055–1088.

[18] K. Fellner, G. Raoul, Stable stationary states of non-local interaction equations, Math-ematical Models and Methods in Applied Sciences 20 (12) (2010) 2267–2291.

23

Page 24: Numerical simulation of nonlinear continuity equations by ...

[19] K. Fellner, G. Raoul, Stability of stationary states of non-local equations with singularinteraction potentials, Mathematical and Computer Modelling 53 (7) (2011) 1436–1450.

[20] A. L. Bertozzi, J. A. Carrillo, T. Laurent, Blow-up in multidimensional aggregationequations with mildly singular interaction kernels, Nonlinearity 22 (3) (2009) 683–710.

[21] Y. Huang, A. L. Bertozzi, Self-similar blowup solutions to an aggregation equation inRn, SIAM J. Appl. Math. 70 (7) (2010) 2582–2603.

[22] Y. Huang, A. Bertozzi, Asymptotics of blowup solutions for the aggregation equation,Discrete Contin. Dyn. Syst. Ser. B 17 (4) (2012) 1309–1331.

[23] J. A. Carrillo, M. Di Francesco, A. Figalli, T. Laurent, D. Slepcev, Global-in-time weakmeasure solutions and finite-time aggregation for nonlocal interaction equations, DukeMathematical Journal 156 (2) (2011) 229–271.

[24] M. A. Herrero, J. J. L. Velazquez, Singularity patterns in a chemotaxis model, Math.Ann. 306 (3) (1996) 583–623.

[25] J. J. L. Velazquez, Point dynamics in a singular limit of the Keller-Segel model. I.Motion of the concentration regions, SIAM J. Appl. Math. 64 (4) (2004) 1198–1223.

[26] A. Blanchet, J. A. Carrillo, N. Masmoudi, Infinite time aggregation for the criticalPatlak-Keller-Segel model in R2, Communications on Pure and Applied Mathematics61 (10) (2008) 1449–1481.

[27] A. Blanchet, V. Calvez, J. A. Carrillo, Convergence of the mass-transport steepestdescent scheme for the subcritical Patlak-Keller-Segel model, SIAM Journal on Nu-merical Analysis 46 (2) (2008) 691–721.

[28] A. Blanchet, J. A. Carrillo, P. Laurencot, Critical mass for a Patlak-Keller-Segelmodel with degenerate diffusion in higher dimensions, Calc. Var. Partial DifferentialEquations 35 (2) (2009) 133–168.

[29] Y. Yao, A. L. Bertozzi, Blow-up dynamics for the aggregation equation with degen-erate diffusion, Physica D: Nonlinear Phenomena 260 (0) (2013) 77 – 89.

[30] R. Jordan, D. Kinderlehrer, F. Otto, The variational formulation of the Fokker–Planckequation, SIAM Journal on Mathematical Analysis 29 (1) (1998) 1–17.

[31] F. Otto, The geometry of dissipative evolution equations: the porous medium equa-tion, Comm. Partial Differential Equations 26 (1-2) (2001) 101–174.

[32] L. Gosse, G. Toscani, Lagrangian numerical approximations to one-dimensionalconvolution-diffusion equations, SIAM Journal on Scientific Computing 28 (4) (2006)1203–1227.

24

Page 25: Numerical simulation of nonlinear continuity equations by ...

[33] L. Gosse, G. Toscani, Identification of asymptotic decay to self-similarity for one-dimensional filtration equations, SIAM Journal on Numerical Analysis 43 (6) (2006)2590–2606.

[34] M. Westdickenberg, J. Wilkening, Variational particle schemes for the porous mediumequation and for the system of isentropic Euler equations, ESAIM: MathematicalModelling and Numerical Analysis 44 (01) (2010) 133–166.

[35] B. During, D. Matthes, J. P. Milisic, A gradient flow scheme for nonlinear fourth orderequations, Discrete Contin. Dyn. Syst. Ser. B 14 (3) (2010) 935–959.

[36] D. Matthes, H. Osberger, Convergence of a variational Lagrangian scheme for a nonlin-ear drift diffusion equation, ESAIM: Mathematical Modelling and Numerical Analysis48 (03) (2014) 697–726.

[37] J. A. Carrillo, Y. Huang, F. S. Patacchini, G. Wolansky, Numerical study of a particlemethod for gradient flows, arXiv preprint arXiv:1512.03029.

[38] O. Junge, D. Matthes, H. Osberger, A fully discrete variational scheme for solv-ing nonlinear Fokker-Planck equations in higher space dimensions, arXiv preprintarXiv:1509.07721.

[39] J. A. Carrillo, A. Chertock, Y. Huang, A finite-volume method for nonlinear nonlocalequations with a gradient flow structure, Communications in Computational Physics17 (01) (2015) 233–258.

[40] J.-D. Benamou, G. Carlier, Q. Merigot, E. Oudet, Discretization of functionals in-volving the Monge–Ampere operator, Numerische Mathematik (2014) 1–26.

[41] L. Evans, O. Savin, W. Gangbo, Diffeomorphisms and nonlinear heat flows, SIAMJournal on Mathematical Analysis 37 (3) (2005) 737–751.

[42] J. A. Carrillo, J. S. Moll, Numerical simulation of diffusive and aggregation phenom-ena in nonlinear continuity equations by evolving diffeomorphisms, SIAM Journal onScientific Computing 31 (6) (2009) 4305.

[43] C. Villani, Topics in Optimal Transportation, Vol. 58 of Graduate Studies in Mathe-matics, American Mathematical Society, Providence, RI, 2003.

[44] L. Ambrosio, N. Gigli, G. Savare, Gradient flows, Springer, 2005.

[45] L. Ambrosio, S. Lisini, G. Savare, Stability of flows associated to gradient vector fieldsand convergence of iterated transport maps, manuscripta mathematica 121 (1) (2006)1–50.

[46] S. Haker, A. Tannenbaum, Optimal mass transport and image registration, in: Varia-tional and Level Set Methods in Computer Vision, 2001. Proceedings. IEEE Workshopon, 2001, pp. 29 –36.

25

Page 26: Numerical simulation of nonlinear continuity equations by ...

[47] J. Moser, On the volume elements on a manifold, Transactions of the American Math-ematical Society (1965) 286–294.

[48] A. Avinyo, J. Sola-Morales, M. Valencia, On maps with given Jacobians involvingthe heat equation, Zeitschrift fur angewandte Mathematik und Physik ZAMP 54 (6)(2003) 919–936.

[49] M. T. Gastner, M. E. Newman, Diffusion-based method for producing density-equalizing maps, PNAS 101 (20) (2004) 7499–7504.

[50] J. L. Vazquez, Smoothing and decay estimates for nonlinear parabolic equations ofporous medium type, Oxford Lecture Notes in Maths and its Applications 33.

[51] C. Budd, G. Collins, W. Huang, R. Russell, Self–similar numerical solutions of theporous–medium equation using moving mesh methods, Phil. Trans. R. Soc. A: Math-ematical, Physical and Engineering Sciences 357 (1754) (1999) 1047–1077.

[52] J. A. Carrillo, M. Di Francesco, G. Toscani, Strict contractivity of the 2- Wasser-stein distance for the porous medium equation by mass-centering, Proceedings of theAmerican Mathematical Society 135 (2) (2007) 353–363.

[53] E. B. Saff, V. Totik, Logarithmic potentials with external fields, Grundlehren dermathematischen Wissenschaften, Springer, Berlin, New York.

[54] J. A. Carrillo, L. C. Ferreira, J. C. Precioso, A mass-transportation approach to a onedimensional fluid mechanics model with nonlocal velocity, Advances in Mathematics231 (1) (2012) 306–327.

[55] V. Calvez, B. Perthame, M. Sharifi Tabar, Modified Keller-Segel system and criticalmass for the log interaction kernel, Contemp. Math. 429 (2007) 45–62.

[56] C. J. Budd, R. Carretero-Gonzalez, R. D. Russell, Precise computations of chemotacticcollapse using moving mesh methods, J. Comput. Phys. 202 (2) (2005) 463–487.

[57] J. A. Carrillo, M. R. D’Orsogna, V. Panferov, Double milling in self-propelled swarmsfrom kinetic theory, Kinet. Relat. Models 2 (2) (2009) 363–378.

26