• No results found

A second-order numerical method for the aggregation equations

N/A
N/A
Protected

Academic year: 2022

Share "A second-order numerical method for the aggregation equations"

Copied!
32
0
0

Laster.... (Se fulltekst nå)

Fulltekst

(1)

EQUATIONS

JOS ´E A. CARRILLO, ULRIK S. FJORDHOLM, AND SUSANNE SOLEM

ABSTRACT. Inspired by so-called TVD limiter-based second-order schemes for hyper- bolic conservation laws, we develop a formally second-order accurate numerical method for multi-dimensional aggregation equations. The method allows for simulations to be continued after the first blow-up time of the solution. In the case of symmetric,λ-convex potentials with a possible Lipschitz singularity at the origin we prove that the method converges in the Monge–Kantorovich distance towards the unique gradient flow solution.

Several numerical experiments are presented to validate the second-order convergence rate and to explore the performance of the scheme.

1. INTRODUCTION

In this paper we derive and analyze a formally second-order accurate numerical method for the aggregation equation

(1.1) ∂tρ=∇ · ∇W ∗ρ

ρ

, ρ(0) =ρ0

whereρ = ρ(t) ∈ P(Rd)is a time-parametrized probability measure onRd andρ0 ∈ P(Rd)is given. The interaction potentialW : Rd →Ris assumed to satisfy some or all of the following conditions:

(A1) W is Lipschitz continuous,W(x) =W(−x)andW(0) = 0.

(A2) W ∈C1 Rd\ {0}

.

(A3) W isλ-convex for someλ>0, i.eW(x) +λ2|x|2is convex.

Potentials satisfying(A1)–(A3)with a Lipschitz singularity at the origin are the so-called pointy potentials. WhenWis a pointy potential, weak solutions to (1.1) might concentrate into Dirac measures in finite time. The finite time blow-up of solutions has attracted a lot of attention, see [44, 9, 5, 8, 33], wherein almost sharp conditions were given for finite time blow-up and typical blow-up profiles were studied. This finite time blow-up phenomenon explains the necessity of considering measure valued solutions of (1.1). By utilizing the gradient flow structure of (1.1), Carrillo et al. [15] proved existence and uniqueness of solutions to (1.1) whenW satisfies(A1)–(A3).

Aggregation equations of the form (1.1), being the continuum limits of particle systems described by

˙

xi=−X

i6=j

mj∇W(xi−xj), X

i

mi= 1, mi>0, (1.2)

wherexiare the particle positions andmithe weights, are ubiquitous in modelling con- centration in applied mathematics. They find applications in physical and biological sci- ences, to name a few: granular materials [4, 44, 12, 18], particle assembly [32], swarm- ing [49, 52, 47, 53, 39], bacterial chemotaxis [38, 27, 34], and opinion dynamics [48].

Furthermore, attraction-repulsion potentials have recently been proposed as very simple models of pattern formation due to the rich structure of the set of stationary solutions, see [50, 3, 55, 56, 2, 40, 6] for instance.

2010Mathematics Subject Classification. 35R09, 35D30, 35Q92, 65M12, 65M08.

Key words and phrases. Aggregation equations, numerical methods, weak measure solutions, measure reconstruction.

1

(2)

The numerical method proposed in this paper is motivated by the fact that the Burgers- type equation

(1.3) ∂tu+∂xf(u) = 0, f(u) =±(u−u2) and the one-dimensional aggregation equation

(1.4) ∂tρ=∂x W0∗ρ

ρ

withW(x) = ±|x|are equivalent, see [9, 10]. Indeed, defining the primitiveu(x, t) = Rx

−∞ρ(dy, t), we see that

W0∗ρ=±sgn∗ρ=±(2u−1).

Integrating (1.4) over(−∞, x]therefore gives (1.3). Thus, formally speaking, differenti- ating (1.3) in xyields (1.4). This intuition was made rigorous in [10] in which entropy solutions to (1.3) for nondecreasing initial data are shown to be equivalent to gradient flow solutions to (1.4) for measure valued initial data.

Our starting point is a formally second-order accurate finite volume method for solu- tions of Burgers’ equation (1.3). By “differentiating the method” inxwe obtain a numer- ical method for (1.4) with W(x) = ±|x|. This method is then extended to the class of potentials W satisfying (A1),(A2) and any dimensiond. The order of accuracy of the method is preserved when measured in the right metric, namely the Monge–Kantorovich distanced1. Indeed, the Monge–Kantorovich distanced1at the level of (1.4) corresponds to theL1norm at the level of (1.3) in one dimension due to the relation

d1(µ, ν) = sup

kϕkLip61

Z

R

ϕ(x)d(µ−ν)(x) = Z

R

(µ−ν)((−∞, x])

dx=ku−vkL1(R). The second-order accuracy of the numerical method for Burgers’ equation is obtained by reconstructing the numerical approximation into a piecewise linear function in every timestep (see e.g. [29, 43]). A reconstruction also takes place in the proposed scheme, but the result of the procedure is areconstructed measure. This measure consists of a combi- nation of constant values in the grid cells and Dirac deltas at the grid points. This mixed reconstruction, Diracs plus piecewise constants, is the main difference between our method compared to other methods (of lower order) to solve the aggregation equation with mea- sure valued initial data [35, 17, 22, 23]. Other numerical schemes based on finite volumes [13] or optimal transport strategies [30, 20, 21] have been proposed.

Above, and throughout the paper, we use the terms ‘formally second-order’ and ‘second- order’ in the sense of having a local truncation error of orderO(∆t∆x2). This nomencla- ture is standard in the literature on numerical methods for hyperbolic conservation laws [29, 41, 43]. Such truncation error estimates rely on Taylor expansions of the exact solu- tion and hence requires the existence of a smooth solution. There are very few rigorous convergence rate results available for such methods for general, non-smooth solutions (be- yond the suboptimalO(∆x1/2)estimate due to Kuznetsov [42]). We would expect that a rigorous convergence rate estimate for the methods presented here (beyond our local trun- cation estimate) would require a substantial amount of work, and only apply in a limited number of scenarios. We refer to [23] for a proof of anO(∆x1/2)convergence rate for a numerical method for (1.1).

We derive the method for (1.4) with W(x) = ±|x|before generalizing it in one di- mension to any potential satisfying(A1),(A2)in Section 3. We study its properties, and show the convergence of the scheme for measure valued solutions in the distance d1 in the main theorem. The scheme and the main theorem is generalized to any dimension in Section 4. Section 5 is devoted to validating the scheme in known particular cases to- gether with accuracy tests and numerical explorations for both potentials covered by the theory and attractive-repulsive potentials not covered. Section 2 deals with the necessary preliminaries about gradient flow solutions to the aggregation equation (1.1).

(3)

2. PRELIMINARIES ON GRADIENT FLOW SOLUTIONS

We define the space of probability measures with finitep-th order moment,16p <∞ as

Pp(Rd) =

µnonnegative Borel measure, µ(Rd) = 1, Z

Rd

|x|pµ(dx)<∞

. This space is endowed with the optimal transport distancedpdefined by

dp(µ, ν) = inf

γ∈Γ(µ,ν)

Z

|y−x|pγ(dx, dy) 1/p

whereΓ(µ, ν)is the set of measures onRd×Rdwith marginalsµandν(see e.g. [54, 1]).

The particular cases that will be useful in our present work are the Euclidean Wasserstein distanced2and the Monge–Kantorovich distanced1. Let

(2.1) W(ρ) =1

2 Z

Rd×Rd

W(x−y)ρ(dx)ρ(dy)

be the total potential energy associated to the aggregation equation (1.1). It is by now classical that the aggregation equation (1.1) can be written as

∂ρ

∂t =∇ ·

ρ∇δW δρ

,

with δWδρ = W ∗ρthe variational derivative of the functional W. This is the formal signature of thed2-gradient flow structure of evolutions equations [1, 54, 18, 19].

We say thatµ∈ACloc1/2([0,+∞);P2(Rd))ifµis locally H¨older continuous of exponent

1/2in time with respect to the distanced2inP2(Rd). A gradient flow solution associated to (2.1) is defined as follows, see [1, 15].

Definition 2.1(Gradient flow solutions). LetW satisfy the assumptions (A1)–(A3). We say that a map ρ ∈ ACloc1/2 [0,+∞);P2(Rd)

is a gradient flow solution of (1.1)asso- ciated with the functional (2.1), if there exists a Borel vector field v such that v(t) ∈ Tanρ(t)P2(Rd)for a.e.t >0, i.e.kv(t)kL2(ρ)∈L2loc(0,+∞), the continuity equation

(2.2) ∂tρ+∇ ·(vρ) = 0,

holds in the sense of distributions, andv(t) =−∂0W(ρ(t))for a.e.t >0. Here,∂0W(ρ) denotes the element of minimal norm in∂W(ρ), the subdifferential ofWat the pointρ.

In [15] it is shown that whenW satisfies(A1)–(A3)we have∂0W =∂0W∗ρ, where

0W(x) =∇W(x)forx6= 0and∂0W(0) = 0. Hence,

(2.3) ∂0W(x, t) =

Z

x6=y

∇W(x−y)ρ(dy, t) is the unique element of minimal norm whenW satisfies(A1)–(A3).

Theorem 2.2(Well-posedness of gradient flow solutions [15]). LetW satisfy assumptions (A1)–(A3). Givenρ0 ∈ P2(Rd)there exists a unique gradient flow solution of (1.1), i.e.

a curveρ∈ACloc1/2([0,+∞);P2(Rd))satisfying(2.2)inD0([0,∞)×Rd)withv(x, t) =

−∂0W∗ρandρ(0) =ρ0.

Let us connect this notion of solution to more classical concepts of weak solutions for PDEs.

(4)

Definition 2.3. A locally in time absolutely continuous in dp curve ρ : [0,+∞) → Pp(Rd),1 6 p < ∞is said to be adp-weak measure solutionto(1.1)with initial da- tumρ0∈ Pp(Rd)if and only if∂0W∗ρ∈L1loc((0,+∞);L2(ρ(t)))and

Z +∞

0

Z

Rd

∂ϕ

∂t(x, t)ρ(dx, t)dt+ Z

Rd

ϕ(x,0)ρ0(dx)

= Z +∞

0

Z

Rd

Z

Rd

∇ϕ(x, t)·∂0W(x−y)ρ(dy, t)ρ(dx, t)dt, (2.4)

for all test functionsϕ∈Cc([0,+∞)×Rd).

In [15] it is proven that the concept of gradient flow solutions to (1.1) under the assump- tions(A1)–(A3)is in fact equivalent to the concept ofd2-weak measure solutions. As a consequence, the uniqueness of gradient flow solutions imply the uniqueness ofd2-weak measure solutions, see [15, Section 2.3].

Observe thatd1-weak measure solutions to (1.1) are alsod2-weak measure solutions to (1.1). Indeed, sincekv(t)kL2(ρ)∈L2loc((0,+∞)), we can apply [1, Theorem 8.3.1] which implies the absolute continuity with respect tod2of the curve of probability measuresρ(t).

This fact will be the key to identifying the limit of the numerical schemes below.

The notion of gradient flow solutions has been proven to be equivalent to the notion of duality solutions in one dimension [36], the Fillipov flow solutions [17], and, as mentioned in the introduction, it is equivalent to the notion of entropy solutions of the one-dimensional Burgers’ equation in the particular case ofW(x) =±|x|, see [10].

Let us finally mention that global existence of measure valued solutions in one dimen- sion to (1.1) with∂xW∗ρreplaced bya(∂xW∗ρ), whereais aC1function, was obtained by James and Vauchelet [36] using the notion of duality solutions, introduced by Bouchut and James [11].

3. ANUMERICAL SCHEME FOR THE1DAGGREGATION EQUATION

Based on the relation

(3.1) u(x, t) =ρ((−∞, x], t)

between solutions to the one-dimensional aggregation equation (1.4) and Burgers’ equa- tion (1.3) in the case W(x) = ±|x|, we will derive a (formally) second-order accurate method for the aggregation model (1.4). Asρis assumed to be a probability measure in x, we can assume thatuwill be a nondecreasing function satisfyingu(−∞, t) = 0and u(+∞, t) = 1. We will later generalize the resulting method to general potentials and multiple dimensions.

3.1. Second-order schemes for Burgers’ equation. We discretize the space-time do- main R×R+ asxi−1/2 = (i−1/2)∆xandtn = n∆t for i ∈ Z andn ∈ N0, where

∆x,∆t > 0 are the discretization parameters. We define also the computational cell Ci:= [xi−1/2, xi+1/2). A finite volume approximation of (1.3) aims to approximate the cell averages

uni ≈ 1

∆x Z

Ci

u(x, tn)dx .

Such schemes are generally first-order accurate, and a popular method of increasing the order of accuracy is byreconstruction: Given cell averagesuni, compute a piecewise linear polynomial

Ru∆x(x, tn) =uniin(x−xi), x∈ Ci

(see e.g. [29, 43]). The slopesσni ∈Rare selected using e.g. the minmod limiter, which for increasing datauni 6uni+1is given byσin= ∆x1 min uni −uni−1, uni+1−uni

. Defining

(5)

the edge valuesun,±i+1/2=Ru∆x(xi+1/2±0, tn), a (formally) second-order accurate finite volume method for (1.3) is given by

un+1i =uni −β

F un,−i+1/2, un,+i+1/2

−F un,−i−1/2, un,+i−1/2

, β:= ∆t

∆x u0i = 1

∆x Z

Ci

u0(x)dx.

(3.2)

Here,Fis any monotone numerical flux function, such as the Lax–Friedrichs-type flux F(u, v) = f(u) +f(v)

2 − c

2(v−u), c= max

i |f0(uni)|

.

This numerical flux is chosen here for its simplicity, and is a Lax–Friedrichs-type flux where the usual constant1/βis replaced by the maximum velocityc.

3.2. Second-order schemes for the aggregation model. In this section we transfer the above approach to the one-dimensional aggregation equation (1.4), first for the Newtonian potential W(x) = ±|x|in Section 3.2.1 and then to more general potentials in Section 3.2.2.

3.2.1. Newtonian potential. Analogous to the relation (3.1), we defineρthrough the rela- tion

ρni+1/2=uni+1−uni

∆x ⇔ uni =X

j6i

∆xρnj−1/2, ρ−∞= 0.

As a simplifying assumption, let us assume that the initial data for the conservation law (1.3) has been sampled through point values,u0i =u0(xi). For the initial dataρ0i+1/2, this relation and the definitionu00((−∞, x])yield

ρ0i+1/2= 1

∆x u0(xi+1)−u0(xi)

= 1

∆xρ0((xi, xi+1]).

Taking the difference iniof (3.2) yields the following numerical method for (1.4) in the caseW(x) =±|x|:

ρn+1i+1/2ni+1/2+β 2

"

∆xX

j6=i

±sgn xi−xj

ρn,+j+1ρn,+i+1n,−j+1ρn,−i+1

−ρn,+j ρn,+i −ρn,−j ρn,−i

+cn ρn,+i+1−ρn,−i+1

−cn ρn,+i −ρn,−i

# , (3.3)

where

cn = ∆xmax

i

X

j6=i

±sgn(xi−xjn,−j ,

X

j6=i

±sgn(xi−xjn,+j

and

(3.4) ρn,+ini−1/2−1

2 σi+1n −σni

, ρn,−i+1ni+1/2+1

2 σi+1n −σin , with initial data given by

(3.5) ρ0i+1/2= 1

∆xρ0((xi, xi+1]).

Observe that

ρn,±i =un,±i+1/2−un,±i−1/2

∆x .

(6)

xi+1/2 xi+3/2

xi−1/2

ui+1

xi xi+1

ui

(a) Cell averages ofu∆x.

xi+1/2 xi+3/2

xi−1/2

ui+1+σi+1(x−xi+1)

xi xi+1

ui+σi(x−xi)

(b) The reconstructionRu∆x.

xi+1/2 xi+3/2

xi−1/2

∆xρi+1/2

xi xi+1

∆xρi+3/2

∆xρi−1/2

(c) Point masses ofρ∆x.

xi+1/2 xi+3/2

xi−1/2

∆x˜ρi+1/2

σi+1 σi

xi xi+1

∆x˜ρi+3/2

∆x˜ρi−1/2

(d) The reconstructed measure rof ρ∆x, where r([xi, xi+1)) = ∆xρi+1/2.

FIGURE1. The reconstruction ofu∆x(top) translated into a reconstruc- tion procedure forρ∆x(bottom). The solid, vertical lines represent Dirac measures centered at the midpointsxi+1/2.

Now, notice that whenW(x) =±|x|we haveW0(x) =±sgn(x)for allx6= 0, so that

±sgncan be replaced byW0. Thus, the method (3.3) can be written as follows, ρn+1i+1/2ni+1/2

2 h

an,+i+1ρn,+i+1+an,−i+1ρn,−i+1 −an,+i ρn,+i −an,−i ρn,−i +cn ρn,+i+1−ρn,−i+1

−cn ρn,+i −ρn,−i i , where

an,+i = ∆xX

j6=i

W0 xi−xj

ρn,+j , an,−i = ∆xX

j6=i

W0 xi−xj ρn,−j .

3.2.2. General potentials. Letρ0i+1/2be as in (3.5). We realize the numerical approxima- tionρni+1/2as the measure

ρ∆x(x, tn) = ∆xX

i

ρni+1/2δx

i+1/2.

This function is reconstructed by defining areconstructed measurernas (3.6) rn=X

i

h

∆x˜ρni+1/2δxi+1/2inL Ci

i

, ρ˜ni+1/2:=ρni+1/2−1

2(σinni+1) whereσni = min ρni−1/2, ρni+1/2

andL

Adenotes the Lebesgue measure restricted to the setA(cf. Figure 1). It is easy to check that the reconstruction preserves mass, in the sense

rn [xi, xi+1)

= ∆xρni+1/2,

and that rn is nonnegative. It follows thatrn ∈ P1(R) whenever ρ∆x(tn) ∈ P1(R).

Moreover, as the reconstruction procedure redistributes mass over a distance no greater than∆x, we have

(3.7) d1 ρ∆x(tn), rn

6∆x.

(7)

We can nowdefineρn,±i as taking information fromrnin the downwind or upwind direc- tion,

ρn,+i = 1

∆xrn (xi−1/2, xi+1/2]

ni+1/2−1

2(σi+1n −σni), ρn,−i = 1

∆xrn [xi−1/2, xi+1/2)

ni−1/2+1

2(σin−σni−1) (3.8)

(compare with (3.4)). Furthermore, if we set ρn,+∆x =X

i

∆xρn,+i δxi, ρn,−∆x =X

i

∆xρn,−i δxi then we find that

∆xX

j6=i

±sgn xi−xj

ρn,±j =∂x0W ∗ρn,±∆x(xi)

whenW(x) = ±|x|. We use the above expression to define the numerical velocities for general potentialsW as follows,

an,+i =∂x0W∗ρn,+∆x(xi) = ∆xX

j6=i

W0 xi−xj ρn,+j , an,−i =∂x0W∗ρn,−∆x(xi) = ∆xX

j6=i

W0 xi−xj ρn,−j . (3.9)

Moreover, we can replaceW0in (3.9) by a continuous and piecewise linear approximation W∆x0 satisfyingW∆x0 (k∆x) =W0(k∆x)for allk6= 0. This will be used in the upcoming convergence proof.

Summing up, with the reconstruction (3.6) and the velocities (3.9) we define the (for- mally) second-order accurate numerical scheme

(3.10a) ρn+1i+1/2ni+1/2+ ∆t

∆x Ji+1n −Jin , where the fluxes are given by the Lax–Friedrichs-type flux formula (3.10b) Jin =an,+i ρn,+i +an,−i ρn,−i

2 +cn

2 ρn,+i −ρn,−i

, cn= max

i |an,±i | andρ0i+1/2is defined in (3.5). We emphasize that this numerical scheme is well-defined for any potentialW satisfying assumptions(A1),(A2).

Notice that if we replaceJinin (3.10b) with the upwind flux (3.10c) Jin= max an,+i ,0

ρn,+i + min an,−i ,0 ρn,−i ,

then the scheme (3.10a), (3.10c) also defines a (formally) second-order accurate numerical scheme, and the upwinding flux formula (3.10c) is a perfect valid alternative to the Lax–

Friedrichs-type flux (3.10b).

From now on, we will refer to the numerical scheme (3.10) meaning that we discuss either the numerical scheme (3.10a), (3.10b) or (3.10a), (3.10c) indistinctively. We will only provide the proofs in the case of the Lax–Friedrichs-type flux (3.10b), but we empha- size that the upwind scheme shares the same stability and convergence properties as the Lax–Friedrichs method.

Remark 3.1. The numerical scheme(3.10)is only (formally) first-order accurate in time.

A higher-order integration in time, such as Heun’s method or another Runge–Kutta method, is needed to make the scheme second-order in both time and space. See e.g.[43, Section 19.4]for more details.

(8)

3.3. Properties of the scheme. The properties of the scheme (3.10) are similar to those of the first-order accurate schemes for (1.4) developed by James and Vauchelet [35, 37].

Define the linear time interpolation (3.11) ρ∆x(t) := tn+1−t

∆t ρ∆x(tn) +t−tn

∆t ρ∆x(tn+1), t∈[tn, tn+1) whereρ∆x(tn) = ∆xP

iρni+1/2δxi+1/2 andρni+1/2is computed with the numerical scheme (3.10).

Lemma 3.2. Assume thatρ0∈ P1(R)and thatW satsfies(A1)–(A2). Assume moreover thatβ := ∆t/∆xsatisfies theCFL condition

(3.12) β6 1

2kWkLip

. Then for allt>0andn∈N0:

(i) Positivity/mass preservation:ρ∆x(t)>0andR

Rρ∆x(dx, t) = 1, (ii) Finite speed of propagation:cn:= maxi

an,±i

6kWkLip, (iii) Bounded first order moment:

(3.13)

Z

R

|x|ρ∆x(dx, t)6 Z

R

|x|ρ0(dx) + 2tkWkLip, (iv) Uniform tightness:Letr>1andε >0. Then

Z

R\[−r,r]

|x|ρ0(dx)< ε =⇒ Z

R\[−R,R]

|x|ρ∆x(dx, t)< εC(t), whereR=r+t/βandC(t) = exp 32kWkLipt

. (v) Preservation of the center of mass:

Z

R

x ρ∆x(dx, t) = Z

R

x ρ∆x(dx,0)

(vi) Time continuity:The mapt7→ρ∆x(t)is uniformly Lipschitz, in the sense that (3.14) d1 ρ∆x(t), ρ∆x(s)

62kWkLip|t−s|

for allt, s>0, whered1denotes the Monge–Kantorovich–Rubinstein metric.

(vii) Bounded second order moment: If in addition ρ0 ∈ P2(R) then ρ∆x(t) has bounded second order moment:

Z

R

|x|2ρ∆x(dx, t)6 Z

R

|x|2ρ0(dx) + 6tkWkLip

Z

R

|x|ρ0(dx) + 12t2kWk2Lip, Proof. From the definition (3.11) ofρ∆x(t)it is clear that we only need to check each property at the discrete timest=tn.

(i) and (ii): The property(i)clearly holds forn = 0. Assume that(i)holds for some n∈N0. By the definition (3.8) we haveρn,±j >0for allj. It then follows that the velocity an,±i is bounded:

an,±i =

∆xX

j6=i

W0 xi−xj ρn,±j

6kWkLip∆xX

j

ρn,±j

=kWkLip∆xX

j

ρnj ±1

2(σnj+1−σjn)

=kWkLip∆xX

j

ρnj =kWkLip.

Using the fact thatρni+1/2= 12n,+in,−i+1), the scheme (3.10a) can be rewritten as ρn+1i+1/2= 1−β cn−an,−i+1

2 ρn,−i+1+1−β cn+an,+i 2 ρn,+i

(9)

2 cn+an,+i+1

ρn,+i+1

2 cn−an,−i ρn,−i . By the induction hypothesis, the definition (3.10b) ofcnand the CFL condition (3.12), we infer thatρn+1i+1/2 >0. Summing the conservative numerical method (3.10a) over alli∈Z and using the definition (3.5) of the initial data yields

X

i

∆xρn+1i+1/2=X

i

∆xρni+1/2=X

i

∆xρ0i+1/2= 1.

(iii): Assume thatρ∆x(tn)satisfies (3.13). From (3.10a) and summation by parts, the first order moment can be written as

∆xX

i

|xi+1/2n+1i+1/2= ∆xX

i

|xi+1/2ni+1/2

−∆xβ 2

X

i

(an,+i +cnn,+i |xi+1/2| − |xi−1/2| (3.15)

−∆xβ 2

X

i

(an,−i −cnn,−i |xi+1/2| − |xi−1/2| + lim

i→∞∆xβ

2|xi+1/2| an,+i ρn,+i +an,−i ρn,−i +cn ρn,+in,−i

− lim

i→−∞∆xβ

2|xi+1/2| an,+i ρn,+i +an,−i ρn,−i +cn ρn,+in,−i . (3.16)

The last two terms vanish becausean,±i satisfies (ii)andρ±∆x(tn) 6 32ρ∆x(tn), where ρ∆x(tn)∈ P1(R)by the induction hypothesis. From the bound

|xi+1/2|−|xi−1/2| 6∆x, (3.12),(ii)and the induction hypothesis, we get

∆xX

i

|xi+1/2n+1i+1/26∆xX

i

|xi+1/2ni+1/2

+ ∆xβ∆x 2

X

i

cn+an,+i

ρn,+i +

cn−an,−i ρn,−i

6∆xX

i

|xi+1/2ni+1/2+kWkLip∆t∆xX

i

ρn,+in,−i

= ∆xX

i

|xi+1/2ni+1/2+ 2∆tkWkLip 6∆xX

i

|xi+1/20i+1/2+ 2tn+1kWkLip.

(iv):We considerx > R. The casex <−Ris similar. Letk∈Zbe such thatR∈ Ck. Then, from a summation by parts,

Z

x>R

|x|ρ∆x(dx, tn+1)

= ∆xX

i>k

xi+1/2ρn+1i+1/2

= ∆xX

i>k

xi+1/2ρni+1/2−∆tX

i>k

xi+1/2−xi−1/2

Jin−∆txk−1/2Jkn

= ∆xX

i>k

xi+1/2ρni+1/2−∆t∆xX

i>k

Jin−∆txk−1/2Jkn 6∆xX

i>k

xi+1/2ρni+1/2−∆t∆xX

i>k

1

2 an,−i −cn

ρn,−i −∆txk−1/2

1

2 an,−k −cn ρn,−k

(10)

6∆xX

i>k

xi+1/2ρni+1/2+ ∆t∆x3

2kWkLip

X

i>k

ρni+1/2+ ∆xxk−1/2

3 4ρnk−1/2

6

1 +3

2kWkLip∆t

∆x X

i>k−1

xi+1/2ρni+1/2

6. . .6

1 + 3

2kWkLip∆t n

∆x X

i>k−n−1

xi+1/2ρ0i+1/2

where we have usedJin> 12(an,−i −cnn,−i in the first inequality, andρn,−i 63/2ρni−1/2, (3.12) and(ii)in the second. As long asr>1, the third inequality follows.

(v): The proof is based on the antisymmetry ofW0(x). Similar to (3.16), it is easy to check that(v)is equivalent to showing that

X

i

an,+i ρn,+i +an,−i ρn,−i +cn ρn,+i −ρn,−i

xi+1/2−xi−1/2

= 0 for alln∈N. Sincecndoes not depend oni,xi+1/2−xi−1/2= ∆x, and taking into account the formulas forρn,±i in (3.8), we deduce that the last statement is equivalent to

X

i

an,+i ρn,+i +an,−i ρn,−i

= 0 for alln∈N. Finally, we have due to the antisymmetry ofW0(x)that

X

i

an,±i ρn,±i =X

i6=j

W0(xi−xjn,±i ρn,±j

=X

i6=j

W0(xj−xin,±i ρn,±j =−X

i6=j

W0(xi−xjn,±i ρn,±j , leading to

X

i

an,±i ρn,±i = 0 for alln∈N.

(vi): The proof is similar to (iii): Multiplying (3.10a) byϕi+1/2 = ϕ(xi+1/2) for a Lipschitz continuous functionϕsatisfyingkϕkLip61gives

∆xX

i

ρn+1i+1/2−ρni+1/2

ϕi+1/2

= ∆xβ 2

X

i

c+an,+i

ρn,+i +

c−an,−i ρn,−i

ϕi+1/2−ϕi−1/2

6 β

22cn∆x2kϕkLip

X

i

ρn,+in,−i 62kϕkLipkWkLip∆t,

and taking the supremum over allϕwithkϕkLip 6 1yields (3.14) witht = tn+1 and s=tn. Iterating over timesteps yields (3.14) for any discrete timestn, tm,m, n∈N. The inequality (3.14) for anyt, s∈R+follows from (3.11).

(vii):The proof follows similarly to(iii).

Remark 3.3. ReplacingJin = 12 an,+i ρn,+i +an,−i ρn,−i +cnn,+i −ρn,−i )

in(3.10b) with the upwind flux(3.10c)in the above proof, we can easily deduce the same properties under the same CFL condition.

(11)

3.4. Convergence of the method. Using the properties derived in the previous section we can now prove convergence of the method using a standard compactness technique.

Theorem 3.4. Letρ0∈ P1(R), assume thatW satisfies properties(A1)and(A2), and that the CFL condition(3.12)is satisfied. Then for anyT >0, the numerical approximation (3.11)has a uniformly convergent subsequence,

(3.17) sup

t∈[0,T]

d1 ρ∆x0(t), ρ(t)

→0 as∆x0 →0, and the limitρis ad1-weak measure solution of(1.4),(2.3)which satisfies (3.18) d1 ρ(t), ρ(s)

62kWkLip(R)|t−s| ∀t, s∈R+.

IfW also fulfills(A3)andρ0 ∈ P2(R)then the whole sequenceρ∆xconverges, and the limitρis the unique gradient flow solution of (1.4).

Proof. Define the set

K:=

ρ∆x(t) : ∆x >0, t∈[0, T] ,

which by Lemma 3.2 (i)and(iii)is a subset ofP1(R)with uniformly bounded first mo- ment. Hence,Kis tight, so by Prohorov’s theoremKis sequentially precompact inP(R) with respect to the weak (or “narrow”) topology (cf. e.g. [1, Theorem 5.1.3]). We claim thatKis also sequentially precompact with respect tod1. By [54, Theorem 7.12], all we need to check is that the first moments are uniformly integrable with respect toK. Fix ε > 0and letr > 0be such thatR

R\[−r,r]|x|ρ0(dx) < ε. By Lemma 3.2(iv), we then have

sup

ρ∈K

Z

R\[−R,R]

|x|ρ(dx)< εC(T)

for some R > 0, which proves our claim. Using the2kWkLip-Lipschitz continuity of ρ∆x (Lemma 3.2(vi)), Ascoli’s theorem now implies the existence of a subsequence of ρ∆x(which we still denote asρ∆x) and some2kWkLip-Lipschitz continuousρ: [0, T]→ P1(R)such thatd1 ρ∆x(t), ρ(t)

→0uniformly fort∈[0, T].

We check that the limitρsatisfies (1.4) in the distributional sense. We multiplyρn+1∆x with a test functionϕ∈Cc2(R), use (3.10a) and perform a summation by parts,

(3.19) Z

R

ϕ(x)ρn+1∆x (dx) = ∆xX

i

ϕ(xi+1/2ni+1/2

−β∆x 2

X

i

an,+i ρn,+i +an,−i ρn,−i

ϕ(xi+1/2)−ϕ(xi−1/2)

−β∆x 2

X

i

cnn,+i −ρn,−i ) ϕ(xi+1/2)−ϕ(xi−1/2) .

By Taylor expanding the last term in (3.19) aroundxi+1/2, summing by parts, and taking into account (3.8), Lemma 3.2(ii), and that the mass is conserved, we find that

cnβ∆x 2

X

i

ρn,+i −ρn,−i

ϕ(xi+1/2)−ϕ(xi−1/2)

=cnβ∆x 2

X

i

ρni+1/2−ρni−1/2−1

2[σii+1−σi−1−σi]

ϕ(xi+1/2)−ϕ(xi−1/2)

= −cnβ∆x 2

X

i

ρni+1/2−σii+1 2

| {z }

∈[0,ρni+1/2]

ϕ(xi+3/2)−2ϕ(xi+1/2) +ϕ(xi−1/2)

| {z }

600kL∆x2

=O(∆x2).

(12)

We insert this into the expression (3.19), and by a new Taylor expansion aroundxi+1/2, we know that there existsyi∈[xi−1/2, xi+1/2]such that

Z

R

ϕ(x)ρn+1∆x (dx)

= ∆xX

i

ϕ(xi+1/2ni+1/2−β∆x 2

X

i

an,+i ρn,+i +an,−i ρn,−i

ϕ0(xi)∆x

−β∆x 2

X

i

an,+i ρn,+i +an,−i ρn,−i

ϕ00(yi)∆x2

2 +O(∆x2)

= Z

R

ϕ(x)ρ∆x(dx, tn)−∆t 2

Z

R

ϕ0(x)a+∆x(x, tn+∆x(dx, tn)

−∆t 2

Z

R

ϕ0(x)a∆x(x, tn∆x(dx, tn) +O(∆x2), wherea±∆x(x, t) =∂x0W∆x∗ρ±∆x(x, t)andρ±∆xis defined similar toρ∆x, cf. (3.11). Then for any test functionϕ∈Cc2(R×R+)we have

Z

R

ϕ(x, tn∆x(dx, tn+ ∆t)−ρ∆x(dx, tn)

∆t

=−1 2

Z

R

xϕ(x, tn)a+∆x(x, tn+∆x(dx, tn)

−1 2

Z

R

xϕ(x, tn)a∆x(x, tn∆x(dx, tn) +O(∆x).

The fact that ρ∆x → ρ, together with the stability property (3.7) of the reconstruction procedure, implies that alsoρ±∆x →ρ. Recall thatW∆x0 is everywhere continuous. Then the stability result [17, Lemma 3.1] implies thata±∆xρ±∆x* ∂0W∗ρ

ρ, where∂0W∗ρ is defined in (2.3). A standard argument of summation by parts in time implies thatρis a distributional solution of (1.4) in the sense of (2.4).

We have shown thatρis a distributional solution of a continuity equation of the form (2.2) where the velocity field is given byv(t) =−∂0W∗ρ(t)for a.e.t >0. Furthermore, kv(t)kL2(ρ) ∈L2loc(0,+∞)since|v(t, x)|6kWkLipfor a.e. t >0andx∈R. Finally, from (3.18) it follows that the continuous curve of probability measuresρ(t)is absolutely continuous in time with respect tod1, and we can thus conclude thatρis ad1-weak measure solution according to Definition 2.3.

Ifρ0∈ P2(R)thenρ∈ P2(R)follows from Lemma 3.2(vii). Under the additional as- sumption(A3),d2-weak measure solutions as defined in (2.3) are unique, see [15, Section 2.3], and they coincide with the unique gradient flow solutions of (1.4) given by Theorem 2.2. Thus, what remains to show to conclude thatρis the unique gradient flow solution, is that thed1-weak measure solutionρ(t)is locally in time absolutely continuous ind2. As pointed out in Section 2, sincekv(t)kL2(ρ) ∈ L2loc(0,+∞), we can apply the properties of continuity equations in [1, Theorem 8.3.1] which imply the absolute continuity with

respect tod2ofρ(t).

Remark 3.5. The repulsive potentialW(x) =−|x|does not satisfy(A3). However, due to the equivalence in[10]we can apply the proof of Theorem 3.4 to obtain the convergence of the numerical scheme also for this potential.

Remark 3.6. Also from the equivalence in [10], we can deduce from Theorem 3.4 the convergence of the minmod scheme(3.2)for Burgers’ equation(1.3)to the unique entropy solution whenever the initial data for Burgers’ equation is nondecreasing. See [41]for further results in this direction.

Referanser

RELATERTE DOKUMENTER

There had been an innovative report prepared by Lord Dawson in 1920 for the Minister of Health’s Consultative Council on Medical and Allied Services, in which he used his

Overall, the SAB considered 60 chemicals that included: (a) 14 declared as RCAs since entry into force of the Convention; (b) chemicals identied as potential RCAs from a list of

The SPH technique and the corpuscular technique are superior to the Eulerian technique and the Lagrangian technique (with erosion) when it is applied to materials that have fluid

In the analysis of flow around an acoustic antenna, various tensors appear, for example the strain rate tensor, structural tensors and tensorial expressions involved in the

We observe that the SP2 method yields numerical solutions that have over an order of magnitude less error for the same computational cost over the one second interval for the

Our approach is independent of the particular numerical method chosen for solving the shallow water equations, but here we implement first, second and third order accurate

The differences are so small that the Newman’s approximation can be suggested as a valid method to calculate second-order loads when considering the surge motion and the mooring

The graph plots the log10 of the absolute value of the Euler equation errors for the first order approximation (top panel) and the second order approximation (bottom panel) to