ON CONTINUITY EQUATIONS IN SPACE-TIME DOMAINS

arXiv:1701.06237v1 [math.AP] 23 Jan 2017

YUMING ZHANG

Abstract. In this paper we consider a class of continuity equations that are conditioned to stay in general space-time domains, which is formulated as a continuum limit of interacting particle systems. We study the well-posedness of the solutions and provide examples illustrating that the stability of solutions is strongly related to the decay of initial data at infinity. We also consider the vanishing viscosity approximation of the system, given with the co-normal boundary data. If the domain is spatially convex, the limit coincides with the solution of our original system, giving another interpretation to the equation.

1. Introduction Let ΩT be a given space-time domain in Rd × (0, T ), denoted by  ΩT := ∪0 c. Note at x ∈ Ω(t) in the interior, Pt is an identity map on v. We refer readers to [4,22] where they defined a similar projection operator on stationary domains. 1

2

Y ZHANG

Figure 1. the particle system and the Pt operator Set ΩT := ∪0 0: in ΩT , on ∂l ΩT ,

(1.5)

on Ω(0).

Here ∂l ΩT := ΩT \ΩT is the lateral boundary of ΩT . The co-normal boundary condition above guarantees the mass preservation. In this case the associated energy φa (µ) is given by φa (µ) = F a (µ) + V(µ) + W(µ) (1.6) Z Z 1 := a u log udx + V (x)dµ(x) + W (x − y)dµ(y)dµ(x). 2 R2d d d R R In the first term, u is the probability density function of µ if µ is absolutely continuous with respect to Euclidean measure. We set φa (µ) = ∞ if µ is not absolutely continuous. Z

With the convergence of a → 0 in mind, we show the existence of solutions by discrete-time gradientflow (JKO) solutions (see [16]). For this purpose, technically we further require V, W to be C 2 and bounded below. In time-dependent domain, the scheme is slightly different from the standard version that we minimize each movement among probability measures with support contained in ΩT (see (3.2)). To obtain the continuum time limit of the discrete-time solutions, we show the uniform boundedness of the second moment and the boundedness of φa along solutions from the discrete scheme. This is one part that the analysis for problems on stationary domains that cannot be directly carried over for the time-dependent domains. The problem is solved in Proposition 3.3, the proof of which is inspired by the work of Di Marino, Maury and Santambrogio [11] who encountered the same problem. Also let us mention that solutions obtained in this way inherit the gradient flow structure which will be important later.

4

Y ZHANG

After establishing the well-posedness of weak solutions, we send a → 0. We show that if the domain is bounded and spatially convex, equations (1.5) are indeed the vanishing viscosity approximation of the first order equation (1.2) in the first part. This convergence justifies the formulation of equation (1.2), in addition to the derivation via particle system. In [7], the vanishing viscosity limit problem in the whole domain was studied in the case when V = 0 and −W is the Newtonian potential. Their proof heavily relies on the specific choice of kernel W , and also the fact that the domain is Rn which eliminates the task of determining the limiting boundary condition. Now let us give the main theorem of the second part of this paper. Theorem B. Assume conditions (C1)-(C4)(O1)(O2) hold (see sections 2.1, 3.1 for details), µ0 is an absolutely continuous (with respect to Lebesgue measure) probability measure supported in Ω(0) with finite second moment, and T is any fixed positive number. Then (a) (Theorem 3.1) There exists a weak solution µ(·) to equation (1.5). (b) (Theorem 3.2) Suppose Ω(t) is bounded for each t ≥ 0 and the density of µ0 is in L2 (Ω(0)). Then there exists a unique weak solution µ(·) to equation (1.5) with density u(t) ∈ L2 (Ω(t)) a.e. t > 0. If µ1 (·), µ2 (·) are two solutions with initial data µ10 , µ20 , then there exits C such that their densities satisfy ku1 (t) − u2 (t)kL2 (Ω(t)) ≤ Cku10 − u20 kL2 (Ω(0) for a.e. t ∈ [0, T ].

(1.7)

(c) (Theorem 3.3) Suppose Ω(t) is bounded and convex for each t ≥ 0. Let µa (·) be the weak solutions to equations (1.5) and µ(·) be the weak solution to equation (1.2). Then for all t dW (µa (t), µ(t)) → 0 as a → 0. We conclude by outlining the strategy for proving the convergence as a → 0. Write  = a. Since the domain is convex, it can be shown that Z Z 1 d 2 (1.8) dW (µ, µ ) ≤ hξ  , tµµ − iidµ + hξ, tµµ − iidµ, 2 dt Ω(t) Ω(t) where tµµ12 is an optimal transport map from µ2 to µ1 defined in section 1.3 below, i is the identity map on Rd , ξ  and ξ belong to the Fr´echet subdifferential of φ , φ at µ , µ respectively. Next from the convexity conditions on V, W , we can show that the functionals φ, φ are uniformly ˜ ˜ By the variational inequalities in gradient flow theory (section 10.1.1 [2]), we have λ-convex for some λ. the following inequalities at each time t: R ˜ 2 λ   1 φ (ω1 ) − φ (µ ) ≥ Ω(t) hξ  , tω (1.9) µ − iidµ + 2 dW (ω1 , µ ), R ˜ λ 2 2 (1.10) φ(ω2 ) − φ(µ) ≥ Ω(t) hξ, tω µ − iidµ + 2 dW (ω2 , µ) where ω1 , ω2 can be any probability measures supported in Ω(t). Formally we take ω1 = µ, ω2 = µ and add up the above two inequalities which then, combining with equality (1.8), leads to 1 d 2 ˜ 2 (µ, µ ). d (µ, µ ) ≤ F  (µ ) + F  (µ) − λd (1.11) W 2 dt W Thus it remains to show that F  (µ ) and F  (µ) converge to 0. However we point out that it is possible F  (µ(t)) = +∞ for some t > 0, since in general even with smooth initial data, µ can concentrate mass in finite time as discussed in [3]. To overcome this problem we look for a µ ˜=u ˜dx to replace µ which is close to µ in some sense, and u ˜ is bounded point-wisely by −α for some 0 < α < 1. If such µ ˜ exists, using that the domain is bounded, we obtain F  (˜ µ) → 0. Then we encounter another problem when we take ω1 = u ˜ in the variational inequality (1.9). A key step is to show that µ, µ ˜ are close not only in 2-Wasserstein metric but also in Pseudo-Wasserstein distance with base µ (the definition will be given in 1.3). To do this we will use properties of generalized geodesics in probability measure space with Pseudo-Wasserstein metric as well as properties of optimal transport plan.

ON CONTINUITY EQUATIONS IN SPACE-TIME DOMAINS

5

1.3. Notations and Preliminaries. We write n(x, t)(or simply n) as the spatial normal vector and c(x, t)(or c) as the speed of the boundary at x ∈ ∂Ω(t). Throughout the paper, we fix a T > 0 which is assumed to be large. We say a constant is universal if it only depends on T and constants in conditions (O1)(O2)(C1)-(C4) (λ, rp , L and bounds about c, V, W ). We denote by C a constant which may depend on universal constants and µ0 , possibly changing from one estimate to another. A spatial ball in Rd centered at x with radius r is denoted by Br (x), and we may simply write Br if x is the origin. R Given a probability measure µ, we write m2 (µ) = x2 dµ as the second moment of µ. The set of all probability measures on Ω with finite second moment is denoted by P2 (Ω). The set of absolutely continuous (with respect to Lebesgue measure) probability measures with finite second moment is written as P2a . For µ ∈ P2a , we usually write µ = uLd where u is its density. For probability measures supported in Ω, we will think of them as measures in Rd , extended by 0 outside Ω. Now we give the definition of weak (measure) solutions to equations (1.2) and (1.5). Definition 1.1. A locally absolutely continuous (in Wasserstein metric) curve µ(·) ∈ P2 (Rd ) is a weak solution to (1.2) with initial value µ0 ∈ P2 (Ω(0)) if v(x, t) = Pt (−∇V − ∇W ∗ µ)(x, t) ∈ L1loc ([0, T ); L2 (µ(t)))  and for all ϕ(x, t) ∈ Cc∞ Rd × (0, T ) : Z Z ϕt dµdt + (∇ϕ · v)dµdt = 0, µ(0) = µ0 , Rd ×(0,T )

Rd ×(0,T )

and for each t ≥ 0, supp(µ(t)) ⊆ Ω(t). Definition 1.2. A locally absolutely continuous curve µ(·) ∈P2a (Rd ) is a weak solution to (1.5) with initial value µ0 ∈ P2a (Ω(0)) if for all ϕ(x, t) ∈ Cc∞ Rd × (0, T ) : Z (ϕt + ∆ϕ − ∇V · ∇ϕ − (∇W ∗ µ) · ∇ϕ) dµdt = 0, µ(0) = µ0 , Rd ×(0,T )

and for each t ≥ 0, supp(µ(t)) ⊆ Ω(t). Now we discuss Wasserstein metric, and we refer readers to [2] for details. Suppose X, Y are measurable subsets of Rd and µ1 ∈ P2 (X), µ2 ∈ P2 (Y ). A plan between µ1 , µ2 is any Borel measure γ on X × Y which has µ1 as its first marginal and µ2 as its second marginal. We write γ ∈ Γ(µ1 , µ2 ). From optimal transport theory, there exists an optimal transport plan γ ∈ Γ(µ1 , µ2 ) such that Z  Z 2 2 0 0 |x − y| dγ(x, y) = min |x − y| dγ (x, y), γ ∈ Γ(µ1 , µ2 ) . X×Y

X×Y

The above quantity is defined to be the 2-Wasserstein distance between µ1 , µ2 (the Kantorovich’s formulation). Throughout this paper we use this distance for probability measures with notation dW (·, ·) unless otherwise stated. And later by Wasserstein distance (metric) we always mean 2-Wasserstein distance (metric). We denote the set of optimal transport plans between µ1 , µ2 by Γ0 (µ1 , µ2 ). Let µ2 ∈ P2 (Y ), a measurable function t : Y → X transports µ2 onto µ1 ∈ P2 (X) if µ1 (B) = µ2 (t−1 (B)) for all measurable B ⊆ X, and we write µ1 = t# µ2 . If µ2 ∈ P2a (Y ), then for any µ1 ∈ P2 (X) there is an optimal transport map tµµ12 : Y → X such that tµµ12 # µ2 = µ1 (With reference to [18]). And we have, in Monge’s formulation, Z 2 dW (µ1 , µ2 ) = |tµµ12 (x) − x|2 dµ2 (x). Y

Given µ1 , µ2 ∈ P2 (X), µ ∈ P2a (X). Let tµµ1 , tµµ1 be an optimal transport maps from µ to µ1 and µ2 respectively. Then the Pseudo-Wasserstein distance with base µ is defined as Z d2µ (µ1 , µ2 ) = |tµµ1 − tµµ2 |2 dµ. X

6

Y ZHANG

By Proposition 1.15 [8], dµ is a metric on Pµ (X) := {µ0 ∈ probability measures on X, dW (µ, µ0 ) < +∞} . And we always have for any µ, dW (·, ·) ≤ dµ (·, ·). 2. Part One. Nonlocal First Order Equations 2.1. Settings and Assumptions. We study equation (1.2) in the first part of this paper. Recall S ⊂ Rn open, with C 1 boundary is said to be r-prox-regular if for any point x ∈ ∂S, y ∈ S we have |y − x|2 hn(x), y − xi ≤ r where n(x) is the unit normal at x (see [6] for more results). This is the same as: for any boundary point x, there is a ball of radius r that intersects S at exactly one point x. Also recall a C 1 function f (x) is called λ−convex in S ⊆ Rd if h∇f (x) − ∇f (y), x − yi ≥ λ|x − y|2 for all x, y ∈ S. We list the assumptions below. (O1) For each t, Ω(t) is a non-empty open subset of Rd which is always rp -prox-regular for some rp > 0. The lateral boundary ∪t (∂Ω(t) × {t}) is C 1 in both time and space direction, particularly we have the boundary speed c(x, t) is continuous if restricted to the boundary. In addition, we require |c(x, t)| ≤ C(1 + |x|). 1

d

(C1) V ∈ C (R ) and |∇V (x)| ≤ C(1 + |x|) for all x ∈ Rd . (C2) W ∈ C 1 (Rd ) with |∇W (x)| ≤ C(1 + |x|) for all x ∈ Rd , and W (x) = W (−x). (C3) There exists some λ that V, W are λ-convex in Rd . 2.2. Particle Approximations. As stated in the introduction, we use particle approximations. ConPN PN sider: µ0 = i=1 mi δxi,0 where N is a large integer and i=1 mi = 1. We look for a solution of the PN form µ(t) = i=1 mi δxi (t) . By the weak formulation of solutions the equation (1.2) becomes:  ( x˙i (t) = Pt − F (xi , µ) a.e. for t > 0 (2.1) xi (0) = xi,0 ∈ Ω(0) where F (y, µ(t)) := ∇V (y) +

X

 mj ∇W y − xj (t) .

(2.2)

j

To solve this ODE, we cite the following result from [13] about the existence of differential inclusions. N

Theorem 2.1. (Theorem 5.1 [13]) Assume S(t) = Ω(t)

⊂ RdN satisfies the following:

for each t ∈ [0, T ], S(t) is nonempty and r-prox-regular; the set S(t) varies absolutely continuously (see (H3) [13]).  Also assume that F : RdN × [0, T ] → nonempty convex compact subsets of RdN satisfies: F (y, x) is upper semicontinuous in y, x; there exists β(t) ≥ 0 ∈ L1 ([0, T ], R), such that |F (x, t)| ≤ β(t)(1 + |x|). N

Then for any x0 ∈ Ω(0) , the following sweeping process with perturbation (  N − x(t) ˙ ∈ N Ω(t) , x(t) + F (x(t), t) a.e. t ∈ I, x(0) = x0

ON CONTINUITY EQUATIONS IN SPACE-TIME DOMAINS

7

 N has at least one absolutely continuous solution x(t). Here N Ω(t) , x(t) denotes the normal cone at x(t) if x(t) is on the boundary, otherwise it is an empty set. In our case of C 1 boundary, the normal cone simply means the collection of all outer normal vectors. Now we prove the following proposition: Proposition 2.1. Assume conditions (O1)(C1)(C2) hold. Let µ0 = Then the ODE system (2.1) has a locally absolutely continuous solution.

P

i

mi δxi,0 be as stated above.

Proof. In order to use Theorem 2.1, we need to verify all the conditions required. Notice that Proposition 2.5 [4] showed that if Ω(t) is r-prox-regular then so is Ω(t)N . The absolute continuity of Ω(t)N follows from condition (O1). Also the upper semicontinuity follows from the definition of F (x, µ) (2.2). Then let us check boundedness condition.  X   ≤ C(1 + |xi (t)|) F xi (t), µ(t) = ∇V xi (t) + m ∇W x (t) − x (t) j i j j   (2.3) X X 1 mj |xj (t)|2 ) 2  . (mj |xi (t)| + mj |xj (t)|) ≤ C 1 + |xi (t)| + ( +C j

j

So |(F (x1 , µ), ..., F (xN , µ))| ≤ C(N )(1 + |(x1 , ..., xN )|). P Actually it can be seen later by estimate (2.6), the bound of j mj |xj (t)|2 is independent of N . In all we verified all the assumptions made in Theorem 2.1, and thus there exits x(t) = (x1 , ..., xN )(t) absolutely continuous such that    N  − x(t) ˙ ∈ N Ω(t) , x(t) + F (x1 , µ), ..., F (xN , µ) , a.e. t ≥ 0, (2.4)  x(t) ∈ Ω(t)N , x(0) = (x , x ...x ). 1,0 2,0 N,0 N

Write −x˙i (t) = hi (x, t) + F (xi , µ). Note for x(t) ∈ Ω(t) : N

h(x, t) = (h1 , ..., hN )(x, t) ∈ N (Ω(t) , x(t)) ⇔ hi (x, t) ∈ N (Ω(t), xi (t)). So we can write ki (x, t)n(xi , t) = hi (x, t) such that − x˙i (t) = F (xi , µ) + ki (x, t)n(xi , t).

(2.5)

We may set ki (x, t) = 0 if xi (t) ∈ / ∂Ω(t). We claim  −F (xi , µ) − ki (x, t)n(xi , t) = Pt − F (xi , µ) a.e. in time. Since xi (·) is absolutely continuous in R, we only consider all the t ∈ [0, T ] where xi (t) are differentiable. Also note that for xi (t) not on the boundary, the claim is trivial. Hence we consider the case when xi (t) ∈ ∂Ω(t). Because xi (t) is supported in Ω(t), x˙i (t) · n(xi , t) ≤ c(xi , t). If x˙i · n = c, using the fact ki ≥ 0, we have −F (xi , µ) · n(xi , t) ≥ c(xi , t). By definition of Pt , we only need to check the normal direction:   Pt − F (xi , µ) · n(xi , t) = c(xi , t) = − F (xi , µ) − ki (x, t)n(xi , t) · n(xi , t). i (t0 ) i (t0 ) If x˙i (t0 ) · n(xi , t0 ) < c(xi , t0 ), we know limt→t+ xi (t)−x = limt→t− xi (t)−x = x˙i < c(xi , t0 ). In t−t0 t−t0 0 0 view of the continuity of c(x, t) on the boundary, we find that x(t) is in the interior of Ω(t) a.e. for t close to t0 . Then we also have at t = t0  Pt0 − F (xi , µ) = −F (xi , µ) = −F (xi , µ) − ki (x, t0 )n(xi , t0 ).



8

Y ZHANG

2.2.1. Some Estimates. Since

d dt xj (t)

 = Pt −F xj (t), µ(t) a.e. for t > 0, we have

d xj (t)| ≤ |F (xj , µ)| + |c(xj , t)|. dt P We know that m2 (µ0 ) = i mi x2i,0 is bounded. By (2.3), X X d (m2 (µ(t)) ≤ 2 mi |xi ||x˙ i | ≤ 2 mi |xi |(|F (xi , µ)| + |c(xi , t)|) dt i i   X X 1 mi (|xi | + |xi |2 ) + ≤C mi |xi |m2 (µ(t)) 2  ≤ C(1 + m2 (µ(t))). |

i

(2.6)

j

 This provides us a uniform bound for m2 µ(t) which only depends on T, m2 (µ0 ). Then we give bound to |F (x, µ)|: 1 (2.7) |F (xi , µ)| ≤ C(1 + |xi | + m2 µ(t) 2 ) ≤ C(1 + |xi |). Also we have d |xi |2 ≤ C|xi |(|F (xi , µ, t)| + |c(xi , t)|) ≤ C(1 + |xi |2 ). dt This shows the linear growth of xi : |xi (t)|2 ≤ C(1 + |xi,0 |2 )

for t ≤ T.

(2.8)

Recall ki which satisfies (2.5), we deduce |ki (x, t)| ≤ |F (xi , µ)| + |c(x, t)| ≤ C(1 + |xi |).

(2.9)

From the above d2W (µ(t), µ(s)) ≤

X

mi |xi (t) − xi (s)|2 ≤ C

i

X i

Z

t

(1 + |xi (r)|)dr ≤ C|t − s|

mi s

which gives the 21 -holder continuity of µ(·). Now let us recall the metric derivative of an absolutely continuous curve in P2 (Rd ):  dW µ(t + h), µ(t) 0 . |µ |(t) := lim h→0 h We give the following proposition regarding the well-posedness of solutions for the projected discrete systems. PN Proposition 2.2. Assume (O1)(C1)(C2) and µ0 = i=1 mi δxi,0 ∈ P(Ω(0)), then equation (1.2) has  a weak solution µ(t) ∈ P2 Ω(t) with µ(0) = µ0 . And v(x, t) := Pt − ∇V (x) − ∇W ∗ µ(x) satisfies that Z |µ0 |2 (t) = |v|2 dµ(x, t) for a.e. t > 0. Rd

Cc∞

d



Proof. Let ζ(x, t) ∈ R × (0, T ) be a test function. By Proposition 2.1 Z X ∂ζ  d 0= ζ(x, t)dµ(t) = mi (xi (t), t) + mi ∇x ζ(xi (t), t) · − F (xi , µ) − hi (xt , t) dt Rd ∂t Z Z  ∂ζ = (x, t)dµ(t) + ∇ζ(x, t) · Pt − ∇V (x) − ∇W ∗ µ(x) dµ. Rd ∂t Rd Since  X 2 d2W µ(t + h), µ(t) ≤ mi xi (t + h) − xi (t)

(2.10)

i

and xi (t) are absolutely continuous, µ(t) is an absolutely continuous curve with respect to Wasserstein metric. Then it is not hard to see that µ(t) is a weak solution.

ON CONTINUITY EQUATIONS IN SPACE-TIME DOMAINS

9

Next by direct computation (similarly as in Theorem 3.4 [4] Theorem 8.31 [2]) Z  0 2 |µ | (t) = |Pt − ∇V (x) − ∇W ∗ µ(x) |2 dµ(x, t) for a.e. t > 0. Rd

And by (2.10) and the prior estimates we have v(x, t) ∈ L1 ([0, T ); L2 (µ)).



2.2.2. Stability of Discrete Solutions. The following proposition gives the stability result of solutions in the discrete case. The proof is similar to the one in Proposition 5.1 [4]. The only difference is the movement of the boundary which can be controlled by condition (O1). Proposition 2.3. Assume (O1)(C1)-(C3). Suppose µ1 , µ2 are solutions with discrete type initial measures µ10 , µ20 as in Proposition 2.2. Then there exists a constant C depending on the support of initial data, the conditions and T , such that:   dW µ1 (t), µ2 (t) ≤ CdW µ10 , µ20 for all 0 < t ≤ T. Proof. Let µi (t), v i (x, t), F i (x, µi ) be defined as in Proposition 2.1 for i = 1, 2. Let k i (x, t) be such that v i (x, t) = −F i (x, µi ) − k i (x, t)n(x, t). By rp -prox regularity for each y ∈ Ω(t) hk i (x, t)n(x, t), y − xi ≤

|k i (x, t)| |y − x|2 . rp i

≤ C in the Estimate (2.8) shows that, for t ≤ T , µi (t) are compactly supported. By (2.9), |k (x,t)| rp i 1 2 support of µ (t). Let γt be an optimal transport plan between µ (t) and µ (t), then Z  − hk 1 (x, t)n(x, t) − k 2 (y, t)n(y, t), x − yidγt (x, y) ≤ Cd2W µ1 (t), µ2 (t) . R2d

By Theorem 8.4.7 from [2]: Z  1 d 2 dW µ1 (t), µ2 (t) = hv 1 (x, t) − v 2 (y, t), x − yidγt (x, y) 2 dt R2d Z  ≤− hF (x, µ1 ) − F (y, µ2 ), x − yidγt (x, y) + Cd2W µ1 (t), µ2 (t) . R2d

Since F (x, µi ) = (∇V + ∇W ∗ µi )(x), by condition (C2), Z − h∇V (x) − ∇V (y), x − yidγt (x, y) ≤ −λd2W (µ1 (t), µ2 (t)). Rd

Since W is λ-convex and even, we get Z − h∇W ∗ µ1 (x) − ∇W ∗ µ2 (y), x − yidγt (x, y) R2d Z 1 =− h∇W (x − x0 ) − ∇W (y − y 0 ), x − x0 − y + y 0 idγt (x0 , y 0 )dγt (x, y) 2 R2d Z  λ ≤− |x − x0 − y + y 0 |2 dγt (x0 , y 0 )dγt (x, y) ≤ C(λ)d2W µ1 (t), µ2 (t) . 2 R2d   d 2 Here C(λ) = min{0, 2λ}. So in all dt dW µ1 (t), µ2 (t) ≤ Cd2W µ1 (t), µ2 (t) which gives   d2W µ1 (t), µ2 (t) ≤ Cd2W µ1 (0), µ2 (0) for all 0 ≤ t ≤ T.  Remark 2.4. If we were more careful on the dependence of the constant on T and the support of the initial data (suppose supp(µi0 ) ⊂ BR ), we find out that the constant in above proposition can be bounded by C exp(CRT exp(CT )) where C is a universal constant.

10

Y ZHANG

2.3. Compactly Supported Initial Data. Suppose µ0 ∈ P2 (Ω(0)). Let µn0 be a sequence of delta masses converging to µ0 in Wasserstein metric and at the same time they are all compactly supported in the same compact set. Suppose µn (t) is a solution to (1.2) given by Proposition 2.2 with initial value µn0 . Proposition 2.3 shows that for each t, µn (t) is a Cauchy sequence once µn0 is Cauchy. So we can write the limit as µ(t) which is compactly supported. Now we need to show that the limit µ(t) is indeed a solution of (1.2) which is the hard part. We do not expect showing that  v(t) := Pt − F (x, µ) = Pt (−∇V − ∇W ∗ µ) (2.11) is the tangent velocity field of µ(t) by simply letting n goes to infinity due to the discontinuity of Pt operator, as explained in Remark 3.2 [4]. To overcome this problem, we use gradient flow method. Let us start with the following definition. Definition 2.1. Let µ(·) be an absolutely continuous curve in P2 (Rd ) with compact support in ΩT . We say that µ(·) is a curve of maximal slope with respect to φ in a time-dependent domain, if for all 0 ≤ s < t < T: Z Z Z    1 t 02 1 t Pr − F (x, µ) 2 dµ(r)dr |µ | (r)dr + φ µ(s) ≥ φ µ(t) + 2 s 2 s Rd (2.12) Z tZ   − F (x, µ) + Pr (−F (x, µ)) Pr − F (x, µ) dµ(r)dr. s

Rd

As mentioned in the introduction, since we are considering a time-dependent domain, the last term (will be denoted by E(µ)) appears. This integration is only possibly non-zero on the boundary ∂l ΩT where c(x, t) 6= 0. Now we give the main theorem which generalizes Theorem 1.10 in [4]: Theorem 2.2. Assume conditions (O1)(C1)-(C3) hold. Suppose µ0 is a probability measure compactly supported in Ω(0). µn0 , µn (·) are as stated above. Then µn (·) converges to µ(·) which is an absolutely continuous curve in P2 (Rn ) of maximal slope with respect to φ (see definition 2.12). And µ(·) satisfies the equation (1.2) in the weak sense. Let i = 1, 2. Suppose we have two such solutions µi (·) with initial data µi0 , and µi (·) are compactly supported for t ∈ [0, T ]. Then there exists a constant C that  dW µ1 (t), µ2 (t) ≤ CdW (µ10 , µ20 ) for all 0 ≤ t ≤ T. Here C depends on universal constants and the support of µi0 .  Proof. First let us show that t → φ µn (t) is absolutely continuous.     X   X φ µn (t) − φ µn (s) ≤ mi mj W xi (t) − W xj (t) mj V xj (t) − V xj (s) + j



X

i0}

Rd ∩{c(x,t)>0}

Z =

(ai + bi w)dµ. Rd ∩{c(x,t)>0}

Then take sup over i ∈ N and integrate in time. We see Z tZ Z tZ n n n n lim inf (w − Pr w )Pr w dµ dr ≥ n

s

Rd ∩{c(x,t)>0}

s

(w − Pr w)Pr wdµdr.

(2.17)

Rd ∩{c(x,r)>0}

 For the part c(x, t) < 0, we can show that w − Pt (w) Pt (w) is concave. Then for each t, there exist two families of countable many bounded and continuous functions a0i (x), b0i (x) such that (wn − Pt wn )Pt wn = inf {a0i (x) + b0i (x)wn } . i∈N

12

Y ZHANG

By conditions (C1)(C2), wn (·, t) are equi-continuous on compact sets and converge to w(·, t) pointwise. Fix any small  > 0, since a0i (x), b0i (x) are bounded, there exits N > 0 such that for all n > N, i ∈ N and x in the compact support of µn , µ we have a0i (x) + b0i (x)wn ≥ a0i (x) + b0i (x)w − . Also there are integers n0 (x) > N and i0 that lim inf (wn − Pt wn )Pt wn (x) ≥ lim inf {a0i (x) + b0i (x)wn0 (x)} −  n

i



a0i0 (x)

+

b0i0 (x)wn0 (x)

− 2 ≥ a0i0 (x) + b0i0 (x)w(x) − 3

≥ (w − Pt w)Pt w(x) − 3. Because  can be any small and µ (t), µ(t) are always supported in a compact set for t ≤ T , we conclude Z tZ (wn − Pr wn )Pr wn dµn dr lim inf n

n

Rd ∩{c(x,r) 0.

14

Y ZHANG

Remark 2.6. From the theorem we have the uniqueness of solutions to (1.2) with compact support for any finite time. But it is not clear to us whether solutions can spread out to infinity far of the domain within a finite time, even with compactly supported initial data. It is unknown to us about the general uniqueness result. 2.4. Non-Compactly Supported Data. In this section we consider equation (1.2) with non-compactly supported initial data. Let µ ∈ P 2 (Rd ), define Z δ(R, µ) := { x2 dµ} C BR

where

C BR

denotes the complement of the ball BR . We say µ satisfies condition (R) if

(R) There exists some constant cr > 0 such that δ(R, µ)exp(cr R) → 0 as R → +∞. Measures satisfying the condition decay exponentially fast at infinity which is slightly more generous than compact supported ones. We say a curve of measures {µ(t), t ≥ 0} satisfy condition (R) locally uniformly if for each t ∈ [0, T ], µ(t) satisfies the condition for some cr (T ). Theorem 2.3. Suppose conditions (O1)(C1)-(C3) hold and µ0 satisfies (R). Then there exists a weak solution µ(·) to equation (1.2). And it is an absolutely continuous curve in P2 (Rd ) which satisfies condition (R) locally uniformly. Let i = 1, 2. Suppose we have two such solutions µi (·) with initial data µi0 , and for 0 ≤ t ≤ T µi (·) satisfy condition (R) uniformly with constant cr . Then for any 0 < p < 1, there is tp > 0 (depending on p and universal constants) that for all 0 ≤ t ≤ tp , if dW (µ10 , µ20 ) is small enough (depending on p, cr and universal constants), we have  dW µ1 (t), µ2 (t) ≤ 2dW (µ10 , µ20 )p . P Proof. For existence, we use the particle approximation method as before: let µn (·) = δxi (·) be solutions to equation (1.2) with discrete initial data µn0 → µ0 . Then we need to show that µn (·) converges and also the limit satisfies the equation. First let us assume the convergence and we follow the proofs given in Theorem 2.2 to show the rest. Expressions or estimates (2.13) (2.14) (2.15) (2.17) still hold. But we need to be careful on (2.18) and (2.24) seeing that here {xi (t)} are no longer uniformly bounded. If (2.18) (2.24) are valid, we deduce (2.20) (2.26) and from which we draw the conclusion. To show (2.18), we use truncation method. Recall estimate (2.8), within time T , we get R n ,µ ) δ(R, µn (t)) ≤ δ( CT 0 which converges 0 exponentially fast as R → ∞. By (2.7)(2.8) and definition of Pt |k n (x, t)| ≤ C(1 + |x|),

|Pt wn (x, t)| ≤ |wn (x, t)| + |c(x, t)| ≤ C(1 + |x|).

And similar linear bounds also hold for k(x, t), Pt w(x, t). So for any small  > 0, we can choose R large enough such that Z Z Z t n n n n (w − P w )P w dµ dr ≤ C |x|2 dµn dr <  s {c(x,r)> 1, let C 0 be the constant as given in (2.8) and L0 := C 0 L. Then the above Z Z |k n | + |k m | BL0 B 2 ≤ |x − y| dγBL0 + (|k n | + |k m |)|x − y|d(γBLL00 )C . r 2d 2d p R R By (2.9), |k n | + |k m | B |x − y|2 dγBLL00 ≤ CLd2W (µn , µm ), r 2d p R Z B B n m (|k | + |k |)|x − y|d(γBLL00 )C ≤ C (|x|2 + |y|2 )d(γBLL00 )C Z

Z R2d

Z

2

R2d

n

Z

2

|x| dµ +

≤C |x|≥L0

m

m

Z

|y| dµ + m2 (µ ) |y|≥L0

n

n

Z

1dµ + m2 (µ ) |x|≥L0

! m

1dµ

(2.28)

|y|≥L0

By (2.8), for all 0 ≤ t ≤ T , if xi (t) ∈ BLC0 , then xi (0) ∈ BLC and |xi (t)| ≤ C 0 (|xi (0)| + 1). This, combining with (2.6), gives Z Z  (2.28) ≤ CC 02 |x|2 dµn (0) + |y|2 dµm (0) ≤ CC 02 δ(L, µ0 ). |x|≥L

|y|≥L

0

Note C is a universal constant. In all, finally we have d 2 d (µn , µm ) ≤ CLd2W (µn , µm ) + Cδ(L, µ0 ). dt W This shows  exp(CLt) − 1 + d2W (µn0 , µm (2.29) d2W µn (t), µm (t) ≤ Cδ(L, µ0 ) 0 ) exp(CLt). L cr Select t0 = C where cr comes from (R). By the condition, δ(L, µ0 ) exp(cr L) can be any small if L is n large. Recall that {µn0 } is Cauchy and we further take dW (µn0 , µm 0 ) to be small, thus {µ (t)} is also Cauchy for 0 ≤ t ≤ t0 . Then we can consider each time interval: [0, t0 ], [t0 , 2t0 ]... inductively and we proved that µn (t) → µ(t) for all t ≤ T . −2 log  For any 0 < p < 1, write  := dW (µn0 , µm 0 ) and choose L = Ct0 , t0 as above. For t ≤ t0 (1 − p),  2 n m dW µ (t), µ (t) ≤ Cδ(L, µ0 ) exp(cr L − cr pL) + 2p .

Let  be small enough and L is then large enough. By (2.29) and (R),  n m d2W µn (t), µm (t) ≤ 4d2p W (µ0 , µ0 ). Notice if µ(·) solves the equation (1.2), m2 (µ(t)) ≤ C for t ∈ [0, T ]. The second claim about the stability of solutions satisfying condition (R) follows from the above argument for the discrete type solutions. This shows that solutions satisfying condition (R) are unique.  Remark 2.7. We comment on several situations where the condition (R) can (or possibly) be dropped. (i) Similarly as in Theorem 1.9 [4], if Ω(t) is convex for all t and µi (t)(i = 1, 2) are solutions with general initial data µi0 ∈ P2 (Ω(t)), then we actually have for some universal constant C,  dW µ1 (t), µ2 (t) ≤ CdW (µ10 , µ20 ). Here we do not need any assumptions on the decay of solutions. The proof follows from the observation that in (2.27) hk(x, t)n(x, t), x − yi ≥ 0 by the convexity of the domain. And as a corollary we have the uniqueness result. (ii) If ∇V, ∇W and c(x, t) are uniformly bounded, we can conclude the same as in (i).

16

Y ZHANG

(iii) Note ∇W ∗ µ is non-local. If W = 0, the problem is not hard. We guess for W compactly supported, there is a better stability result. 2.5. Examples and Stability of Solutions. In this section, we will show that the stability result in Theorem 2.3 cannot be improved to dW µ1 (t), µ2 (t) ≤ CdW (µ10 , µ20 ) as long as the domain is unbounded and non-convex. Moreover we give examples showing that without condition (R), the stability of solutions is even weaker than that in Theorem 2.3. All these suggest that the stability of solutions to (1.2) is strongly related to the decay of initial data at infinity. Theorem 2.4. There exists V, W, Ω satisfy conditions (C1)-(C3)(O1) such that the following holds for any t0 > 0. (i) There is µ0 ∈ P2 (Ω) and µn0 → µ0 (write µn (t) as a solution to equation (1.2) with initial data µn0 ) that µ(t), µn (t) satisfy condition (R) locally uniformly and  lim inf dW µ(t0 ), µn (t0 ) /dW (µ0 , µn0 ) = +∞. n

(ii) For any p >

1 2,

there is µ0 ∈ P2 (Ω) and µn0 → µ0 that  lim inf dW µ(t0 ), µn (t0 ) /dpW (µ0 , µn0 ) = +∞. n

Proof. Consider the following stationary domain in R2 Ω := {(x, y)| y ≥ cos(2πx)} and the equation ∂ µ + ∇ · (µP (−∇V )) = 0 with V = xy. (2.30) ∂t Here P is the projection operator defined in (1.1) with zero boundary speed. Let us start by solving the equation with initial data µ0 = δ(x0 ,cos(2πx0 )) where x0 is close to some large integer. Suppose δ(xt ,yt ) is a solution, then by simple calculations (x0t , yt0 ) = P (−∇V (xt , yt ) = P (−yt , −xt ). It is not hard to see that if xt is large enough, the delta mass moves along the boundary. So P (−yt , −xt ) = P (− cos(2πxt ), −xt ) and vt := (x0t , yt0 ) = (wt , −2π sin(2πxt )wt ) where wt =

−(cos(2πxt ) − 2πxt sin(2πxt )) . 1 + 4π 2 sin2 (2πxt )

(2.31)

Notice the equation x0t = wt is an autonomous system. For each large integer N , there are two equilibrium points in [N − 12 , N + 21 ] namely the solutions of 2πx = cot(2πx). And one is close to N − 12 which is stable and the other is close to N which is unstable. We write the stable one as (N )1∗ and the other as (N )2∗ . We can show (N )1∗ − (N − 21 ) ≈ (N )2∗ − N ≈ N1 . Now we give the initial data ∞ X µ0 = mj δ(j,cos(2πj)) . j=1

Also we construct a family of

µn0

which converge to µ0 . Denote jα := j + (j)−α with 0 < α < 1 and let

µn0 =

n−1 X

mj δ(j,cos(2πj)) +

j=1

As usual, we have the solutions µn (t) = µn0 , µ(t) = µ0 . So for any fix t0 > 0:

∞ X

mj δ(jα ,cos(2πjα )) .

j=n

P

j

δ(xnj (t),yjn (t)) and µ(t) =

P

j

δ(xj (t),yj (t)) with µn (t) =

 d2W µ(t0 ), µn (t0 ) /d2W (µ0 , µn0 ) & X  2 2  X  mj xj (t0 ) − xnj (t0 ) + yj (t0 ) − yjn (t0 ) / mj (3j)−2α . j≥n

j≥n

ON CONTINUITY EQUATIONS IN SPACE-TIME DOMAINS

17

Because α < 1 and jα > (j)1∗ , for j large enough we have (xnj (t), yjn (t)) starts at (jα , cos(2πjα )) and moves towards ((j +1)1∗ , cos(2π(j +1)1∗ )). However the mass starting from (j, cos(2πj)) will move towards d n ((j)1∗ , cos(2π(j)1∗ )). Note by (2.31) we get if xnj (t) − j ≤ 41 , dt xj (t) ≥ Cj 1−α . Hence for any fix t0 if n is large enough, xnj (t0 ) − 3j ≥ 14 . While (xj (t), yj (t)) goes to the opposite direction, so the distance between them is larger than some constant say 41 . And it cannot be too large since their limits are ((j + 1)1∗ , ...), ((j)1∗ , ...)). P At last we choose mj = Cm e−j where Cm is a constant satisfying Cm j e−j = 1. We have X X  mj j −2α ) ≥ lim Cn2α = +∞. lim inf d2W µ(t0 ), µn (t0 ) /d2W (µ0 , µn0 ) & lim inf ( Cmj )/( n

n

j≥n

n

j≥n

It is straight forward to check that the condition (R) is satisfied in the above construction. We claim that this is the example promised which shows that dW (µ(t), µn (t)) cannot be bounded by CdW (µ0 , µn0 ) no matter how small the time is and how close µ0 and µn0 are. 0 −β 0 Next we select mj = C with β > 3 where Cm is some constant making µ0 a probability measure. Pm j −β 2 −β+1 In this case δ(R, µ0 ) ≈ j≥R j j ≈ R which fails condition (R). On the other hand X n d2p j −β j −2α )p ≈ n(−β−2α+1)p and W (µ0 , µ0 ) ≈ ( j≥n

X  d2W µ(t0 ), µn (t0 ) & ( j −β ) ≈ n−β+1 . j≥n

Thus for any p >

β−1 β+2α−1 ,

we deduce

 n lim inf d2W µ(t0 ), µn (t0 ) /d2p W (µ0 , µ0 ) = +∞ n

for all t0 > 0.

We can take α = 1 − , β = 3 −  and then p can be any close to 12 . This shows that in Theorem 2.3 if without condition (R), p cannot be greater than 12 . So at least we can claim that the stability is weaker.  3. Part Two. Second Order Equations In the second part of this paper we show the well-posedness of the second order continuity equation (1.5) and then we send the diffusion term to 0 (a → 0). If Ω(t) is bounded and convex for each t, we show that (1.5) is indeed the vanishing viscosity approximation of (1.2). 3.1. Assumptions and JKO Scheme. Let us assume the following conditions for a moment. We will make a remark about several generations later. (O2) The lateral boundary of ΩT is uniformly C 1 in space. And there exists L > 0 that d(Ω(t), Ω(s)) ≤ L|t − s| for 0 ≤ s, t ≤ T. For V, W , we assume (C1)-(C3) hold and furthermore we assume: (C4) V, W ∈ C 2 (Rd ) are bounded below. Without lose of generality, we take a = 1 while proving the well-posedness. The equation becomes  ∂  in ΩT ,   ∂t µ − ∇ · (∇µ + ∇V µ + (∇W ∗ µ)µ) = 0  (3.1) ∇µ + ∇V µ + (∇W ∗ µ)µ + cµ · n = 0 on ∂l ΩT ,    µ = µ0 on Ω(0). Suppose µ0 ∈ P2a (Ω(0)) and conditions (C1)-(C4)(O1)(O2) hold. Recall that the associated energy functional φ1 are as defined in (1.3). We use the following variation of JKO scheme: fix a small time

18

Y ZHANG

  step τ > 0, define Jτ,t : P2a Ω(t) → P2a Ω(t + τ ) by  Jτ,t (µ) ∈ argmin



v∈P2 Ω(t+τ )

 1 2 1 d (µ, v) + φ (v) . 2τ W

(3.2)

First we show the existence of such minimizers. With the assumptions (C1)-(C4) on V, W , we have φ1 is lower semi-continuous, coercive, compact. Then   1 2 1 lim inf dW (µ, v) + φ (v) v∈P2 (Ω(t+τ )) 2τ is bounded below. And we can find a sequence of measures whose energy goes to the infimum value and they all belong to P2a due to the internal energy. Lower semi-continuous and compactness guarantee the existence of the converging limit. Details can be found in section 2.1 in [2] or Lemma 4.2 of [22]. Actually if {Ω(t), t ≥ 0} are always convex, we have the uniqueness of the minimizer. However, here we only need the existence result. 3.2. Priori Estimates. First we prove the following technical lemma which is enlightened by Corollary 2.6 in [11]. It will be used to compare µ(t) and Jτ,t (µ(t)) whose support are different. Proposition 3.1. Suppose the domain satisfies condition (O2). Then for t ∈ [0, T − τ ] and µ ∈  P2a Ω(t) , there exists a C 1 map t : Ω(t) → Ω(t + τ ) such that kt − ikL2 (Ω(t),µ) ≤ CLτ,

(3.3)

det(Dt) ≥ 1 − CLτ for all x ∈ Ω(t).

(3.4)

Here C is some constant independent of t, τ which only depends on the geometry of ΩT . Proof. By (O2), ∂Ω(t)C is rp -prox regular for each t ∈ [0, T ] for some rp > 0. So there is a Lipschitz map Qt1 : ∂Ω(t) → Ω(t) such that  rp Qt1 (y) = z with d(z, y) = d z, ∂Ω(t) = . 2 Let Qt : Ω(t) → Ω(t) that (   rp Qt1 (y) s.t. y ∈ ∂Ω(t), d(x, y) = d x, ∂Ω(t) , for d x, ∂Ω(t) ≤ , t 2 Q (x) = x, otherwise. Observe B(Qt (x),

rp 3 )

∈ Ω(t). We define t(x) = (1 −

3Lτ 3Lτ t )x + Q (x). rp rp

Without lose of generality we can modify the Qt to be C 1 with |DQt | uniformly bounded. And this shows estimate (3.4). We then show that t : Ω(t) → Ω(t + τ ). Otherwise if there exists x ∈ Ω(t) such that t(x) 6= Ω(t + τ ), r by (O2) for any z ∈ ∂B(0, 3p ), the line segment connecting x and Qt (x) + z lies in Ω(t). Then t(x) +

3Lτ z ∈ Ω(t). rp r

In view of the fact that z is an arbitrary vector with length 3p , we end up with a contradiction to the Lipschitz variation of Ω(·). Next for (3.3), Z Z Lτ |t(x) − x|2 dµ = ( )2 | − x + Qt (x)|2 dµ ≤ CL2 τ 2 . r p Ω(t) Ω(t)  As a corollary, we have Lemma 3.2 below. Notice that the hypotheses (3.5) is weaker than (3.3) which is made to allow possible weaker assumptions (than (O1)(O2)) on the domain. See Remark 3.5.

ON CONTINUITY EQUATIONS IN SPACE-TIME DOMAINS

19

Lemma 3.2. Assume conditions (C1)(C2) hold. Suppose for for all τ > 0 small enough and t > 0, Jτ,t is well-defined, and there is a map t : Ω(t) → Ω(t + τ ) and universal constants C, q < 12 such that estimate (3.4) holds and kt − ikL2 (Ω(t),µ) ≤ CLτ (1 + m2 (µ)q ). (3.5)  0 0 a Then for some universal constants q < 1, C and any µ ∈ P2 Ω(t) , we have   0 1 2 d µ, Jτ,t (µ) + φ1 Jτ,t (µ) ≤ φ1 (µ) + C 0 τ (1 + m2 (µ)q ). 2τ W Proof. Write v = t#µ and by the assumption Z d2W (µ, v) ≤ (t(x) − x)2 dµ ≤ Cτ 2 (1 + m2 (µ)q )2 . Ω(t) 1

1

Now let us estimate φ (v) − φ (µ). By simple calculation µ . t#µ ◦ t = det(Dt) Write F (µ) = u log u and then Z Z F (v)dx − Ω(t+τ )

Z

Z F (µ)dx =

Ω(t)

Ω(t+τ )

F (t#µ) dt#µ − t#u

Z Ω(t)

F (µ) dµ u

 F (t#µ ◦ t) F (µ) µ − )dµ = F( ) det(Dt) − F (µ) dx t#u ◦ t u det(Dt) Ω(t) Z ≤ −u log det Dtdx ≤ Cτ. Z

=

( Ω(t)

Ω(t)∩{det(Dt)≤1}

Next we calculate the difference between V(v), V(µ). From (C1), |∇V (x)|2 ≤ C(1 + x2 ). Then Z Z Z  V dv − V dµ = V t(x) − V (x)dµ Ω(t+τ )

Z

Ω(t)

Ω(t)

Z ∇V (ϑ)(t(x) − x)dµ ≤ C(

= Ω(t)

1

(1 + x2 + t(x)2 )dµ) 2 (

Ω(t) 1 2

Z

1

(t(x) − x)2 dµ) 2

Ω(t)

q

0

q

≤ Cτ (1 + m2 (µ) + τ m2 (µ) )(1 + m2 (µ) ) ≤ Cτ (1 + m2 (µ)q ). By mean-value theorem, ϑ above lies in the segment connecting x, t(x). And q 0 = 21 + q < 1. The last but two inequality holds because: t(x)2 = 2x2 + 2(t(x) − x)2 and kt − ikL2 (Ω(t),u) can be bounded by Cτ (1 + m2 (µ)q ) by the assumption. The last one is holder inequality. Then similar computation yields Z Z W (x − y)dv(y)dv(x) − W (x − y)dµ(y)dµ(x) Ω(t+τ )×Ω(t+τ )

Z ≤ C(

Ω(t)×Ω(t) 1

(1 + x2 + t(x)2 + y 2 + t(y)2 )dµ2 ) 2 (

Ω(t)

Z

1

(t(x) − x)2 + (t(y) − y)2 dµ2 ) 2

Ω(t) 0

≤ Cτ (1 + m2 (µ)q ). In all we proved 0

φ1 (v) ≤ φ1 (µ) + Cτ (1 + m2 (µ)q ). Then the optimality of Jτ,t (µ) gives:   1 2 1 2 dW µ, Jτ,t (µ) + φ1 Jτ,t (µ) ≤ d (µ, v) + φ1 (v) 2τ 2τ W 0

≤ φ1 (µ) + Cτ (1 + m2 (µ)q ) + Cτ (1 + m2 (µ)q )2 . Note 2q < q 0 , we finished the proof.



20

Y ZHANG

Define µkτ := Jτ,(k−1)τ ◦ ... ◦ Jτ,0 (µ0 ) ∈ P(Ω((kτ ))). Let tk be the transport plan as in Lemma 3.1 with t = kτ . Apply this lemma by replacing µ by µkτ .  Denote vk := tk #µkτ which is then supported in Ω (k + 1)τ . We get Proposition 3.3. Suppose conditions (O1)(C1)-(C4) hold and under the assumption of Lemma 3.2. For fixed µ0 , T , if τ is small enough and nτ < T , then there exists C > 0 independent of τ, k, n such that n−1 X

d2W (µkτ , µk+1 ) ≤ Cτ, τ

φ1 (µnτ ) ≤ C.

k=0

Proof. By Proposition 3.1 and Lemma 3.2, 0 1 2 d (µk , µk+1 ) + φ1 (µk+1 ) ≤ φ1 (µkτ ) + Cτ (1 + m2 (µkτ )q ). τ 2τ W τ τ By iteration n−1 0 0 1 X 2 dW (µkτ , µk+1 ) + φ1 (µnτ ) ≤ φ1 (µ0 ) + Cτ n + Cτ m2 (u0τ )q + ... + Cτ m2 (µn−1 )q . τ τ 2τ k=0 Pn−1 0 n By dW (µτ , µτ ) ≤ k=0 dW (µkτ , µk+1 ) and the above, we obtain τ

(3.6)

n−1 1 2 1 X 2 dW (µkτ , µk+1 ) dW (µ0τ , µnτ ) ≤ τ 2nτ 2τ k=0

≤ φ1 (µ0 ) − φ1 (µnτ ) + Cτ

n−1 X

0

m2 (µkτ )q + C.

(3.7)

k=1

To give a bound to m2 (µ), we use the trick as in proposition 4.1 [16]. Since m2 (µnτ ) ≤ 2m2 (µ0τ ) + 2d2W (µ0τ , µnτ ) Combining with (3.7) and the lower bound of V, W , m2 (µnτ ) ≤ C + Cτ

n−1 X

0

m2 (µkτ )q .

k=0 0

Here q < 1. Considering that C is independent of τ, n (only depends on T, µ0 , ΩT ) and n can be any positive integer such that nτ < T , the above shows m2 (µnτ ) is uniformly bounded. According to (3.7) n−1 X d2W (µkτ , µk+1 ) ≤ Cτ. τ k=0

And in view of (3.6), φ

1

(µkτ )

are uniformly bounded.



3.3. Convergence of Discrete Solutions. We define a discrete solution with time step τ as µτ (t) := Jτ,(k−1)τ ◦ ... ◦ Jτ,0 (µ0 ) = µkτ

if t ∈ ((k − 1)τ, kτ ].

(3.8)

As proved in [16], 

µ ∈ P2 (Rd ) : φ1 (µ) ≤ C, m2 (µ) ≤ C for some t ≤ T



is compact in P2a (Rd ). Then according to Proposition 3.3, {µτ (t), t ≤ T, τ > 0} belongs to a compact subset in P2a (Rd ). We connect every pair of consecutive discrete values (µk−1 , µkτ ) with a constant speed geodesic parametrized τ in each interval [(k − 1)τ, kτ ] by   µ ˆτ (k − 1)τ + s := (1 − s)i + st # µk−1 , s ∈ [0, 1]. τ

ON CONTINUITY EQUATIONS IN SPACE-TIME DOMAINS

21

Here t is an optimal transport map from µk−1 to µkτ . Again by Proposition 3.3, µ ˆτ are holder τ  continuous curves. Ascoli-Arzela Theorem yields the relative compactness of µ ˆτ in C 0 [0, T ]; P2 (Rd ) . Then there is a subsequence of µ ˆτ that converge to some µ(t) for 0 ≤ t ≤ T . Obviously µ(t) is supported in Ω(t) and µτ (·) → µ(·) along the subsequence τ → 0. From a by-now standard computation presented in [16, 20], the Euler-Lagrange equation associated with (3.8) is the following: for any ψ, a smooth vector field compactly supported in Ω(kτ ), Z (y − x) · ψ(y)dγτk + Ω((k−1)τ )×Ω(kτ ) Z (3.9) k k τ (−∇ · ψ + ∇V · ψ + ∇W ∗ µτ · ψ)dµτ = 0. Ω(kτ )

γτk

Here is an optimal transport plan from µkτ to µk−1 . By holder inequality and conditions (C1)(C2) τ Z Z 1 k k−1 k , µkτ ), | (y − x) · ψ(y)dγτ | ≤ dW (µτ , µτ )( |ψ|2 dµkτ ) 2 ≤ kψkL∞ dW (µk−1 τ Ω((k−1)τ )×Ω(kτ ) Ω(kτ ) Z Z 1 | ∇V · ψukτ dx| ≤ kψkL∞ |∇V |dµkτ ≤ CkψkL∞ (m22 (µkτ ) + 1), Ω(kτ ) Ω(kτ ) Z Z 1 | ∇W ∗ µkτ · ψukτ dx| ≤ CkψkL∞ (1 + |x| + |y|)dµkτ (y)dµkτ (x) ≤ CkψkL∞ (m22 (µkτ ) + 1). Ω2 (kτ )

Ω(kτ )

By the uniform bound of m2 (µτ ), we find Z 1 | ukτ ∇ · ψdx| ≤ dW (µkτ , µk−1 )kψkL∞ + CkψkL∞ . τ τ Ω(kτ ) Then we have C dW (µk−1 , µkτ ) + C τ τ which shows that ∇ukτ exists in L1 (dx). Then by approximation, we deduce that equality (3.9) remains valid for ψ ∈ C0∞ (Rn ; Rn ). Actually if applying Proposition 3.3, we get ∇uτ exists in (L1 (dxdt)). . Then Write tkτ as an optimal transport map from µkτ to µk−1 τ Z Z  (y − x) · ψ(y)dγτk = (tkτ − i)ψ tkτ (x) dµkτ . (3.10) k∇ukτ kL1 (Ω(kτ ),dx) ≤

Ω((k−1)τ )×Ω(kτ )

Cc∞

Ω(kτ ) d



Also for every test function ϕ ∈ R × (0, T ) Z  1 | ϕ tkτ (x) − ϕ(x) − h∇ϕ(tkτ (x)), tkτ − iidµkτ | ≤ k∇2 ϕkL∞ d2W (µk−1 , µkτ ). τ 2 Ω(kτ )

(3.11)

Take ψ = ∇ϕ in above. By (3.9),(3.10),(3.11) and Proposition 3.3, we get Z Z Z  1X ∂t ϕdµ = lim ∂t ϕdµτ = lim ϕ(x) − ϕ tkτ (x) dµkτ τ →0 τ →0 τ ΩT ΩT Ω(kτ ) k Z X = lim (−∆ϕ + ∇V · ∇ϕ + ∇W ∗ µkτ · ∇ϕ)dµkτ τ →0

k

Ω(kτ )

Z (−∆ϕ + ∇V · ∇ϕ + ∇W ∗ µ · ∇ϕ)dµdt.

= ΩT

Till now we proved that µ is a weak solution to equation (3.1). We conclude with the following theorem. Theorem 3.1. Suppose conditions (C1)-(C4)(O1)(O2) hold. Then for µ0 ∈ P2a (Ω(0)), there exists an absolutely continuous curve µ(·) in Pa2 (Rd ) which solves equation (3.1) weakly in ΩT . And we have ∇u ∈ L1 (ΩT , dxdt).

22

Y ZHANG

Proof. From the above discussion, along a subsequence of τ → 0, µτ (·) converges narrowly to µ(·) uniformly for all t ∈ [0, T ]. The limit µ is holder continuous in Pa2 (Rd ) and it is a weak solution to equation (3.1). We are left to show that ∇u exits. Since ∇uτ exists in L1 (dxdt) with uniform bound, then uτ are uniformly bounded in BVloc (ΩT ). We can extract a further subsequence that uτ converges in L1loc (ΩT ). And it is not hard to see that the limit is just u which is also a function of bounded variation and µ = udx. Therefore we deduce that ∇uτ → ∇u in the Rsense of distributions for some ∇u. From the above we know ∇uτ are uniformly bounded in L1 (ΩT ) and Rd ×[0,T ] uτ log uτ dxdt are uniformly bounded. Therefore we conclude u ∈ W 1,1 (ΩT ).  Remark 3.4. Proposition 10.4.13 [2] gave another proof for ∇u ∈ L1 (ΩT ) using the finite slope of φ1 . Actually it can be shown that there exists w ∈ L2 (dµdt; Rd ) such that uw = ∇u a.e. dxdt. Similar argument will be used in our Lemma 3.9. Remark 3.5. Here we required conditions (O1)(O2) on the domain. However the condition (O2) is only used in Proposition 3.1 and so that we can apply Lemma 3.2. But according to the assumptions made in the lemma, (O2) is more than what is needed. For example, bound (3.5) can still be achieved if there is a wedge on the boundary. It is technical to construct such maps t depending on the geometry of the time-dependent domain. 3.4. Uniqueness Result. In this section, we prove a uniqueness result of equation (3.1) on bounded non-convex domains. For this purpose, we study the L2 norm of the solutions and we will assume that the density of the initial data u0 is bounded in L2 . Theorem 3.2. Suppose conditions (C1)-(C4)(O1)(O2) hold, ΩT is bounded in Rd+1 . Also assume that µ0 ∈ P2a (Ω(0)) and its density u0 ∈ L2 (Ω(0)). Then there exists a unique weak solution µ to equation (3.1) with density u ∈ L2 (ΩT ). If µi (t)(i = 1, 2) are two solutions with initial data µi0 ∈ P2a (Ω(0)) with their densities ui0 ∈ L2 (Ω(0)), then there is C depending on the domain and universal constants such that for a.e. t ≤ T ku1 (t) − u2 (t)kL2 (Ω(t)) ≤ Cku10 − u20 kL2 (Ω(0)) . Proof. Let µ be a solution to equation (3.1) with initial data µ0 . First we show that u(t) is bounded in L2 norm. Set uι = ηι ∗ u in Rd where ηι (x) is a space modifier which is assumed to be non-negative, 1 d ι supported in a ball of radius ι 0 according to t. Since ρι (·, t) → ρ(·, t) in L2 (Rd ) a.e. dt and ρ(0) = ρι (0) = 0, we see that ρ ≡ 0 in L2 (ΩT ). The second statement can be proved similarly.  Remark 3.6. In [12], equations without interaction terms (W = 0) are considered. Remark 2.2 [12] shows that under the condition that the domain is always convex while moving, we have the uniqueness of solutions. They used gradient flow method. Remark 3.7. In [19], Otto established the L1 -contraction prinple and uniqueness  of solutions for general quasilinear parabolic equations which include equation ∂t u − ∇ · a(∇u, u) = 0. And in our case a(p, u) = p + ∇V u + (∇W ∗ µ)µ. But weak solutions are assumed to be in class Lr (0, T ), W 1,r (Ω(·)) with r > 1 which is unknown here. 3.5. Approximation to First Order Equation. In this section, we consider equations (1.2) and (1.5) in bounded, convex domain. As proved before solutions exist and are unique (see Remark 3.6). Let µ be the weak solution to (1.5) with a there replaced by  and µ be the weak solution to (1.2). We want to show that µ converges to µ in Wasserstein metric. Recall definition (1.6), let us write φ(µ), φ (µ ) as the potentials. Denote the internal energy as Z F  (µ) :=  u log udx where µ = uLd . Rd

24

Y ZHANG

The proper domain of functional φ is defined as n o D(φ , t) := µ ∈ P2 (Ω(t)), φ (µ) < ∞ . The metric slope of functional φ for µ ∈ D(φ , t) at time t is 

|∂φ (t)|(µ) :=

lim sup

+ φ (µ) − φ (w) .  dW (w, µ)

w→µ,w∈P2a Ω(t)

We give the following lemma which will be one of the two ingredients for proving the approximation result. Lemma 3.8. Suppose the domain is bounded and satisfies conditions (O1)(O2), and also µ0 ∈ D(φ1 , 0). R 0 Then for any 0 < T ≤ T , 0≤t≤T 0 F  (µ (t))dt → 0 as  → 0. To prove this lemma, we need Lemma 3.9. Settings are as above. For any 0 <  < 1, Z ∇u 2 |  (x, t)|2 u (x, t)dx Ω(t) u are uniformly bounded in L1 (dt; [0, T ]). As a convention,

∇u u

is defined as  ∇u   u ∇u := 0  u  +∞

if u 6= 0, if ∇u = 0, if ∇u 6= 0, u = 0.

Proof. Recall JKO scheme (3.8), let µτ be defined as the discrete solution with time step τ . From the above we know that for t ≤ T , µτ converges to µ uniformly in Wasserstein metric. For abbreviation, = µτ (kτ ). By exactly the same proof as in Lemma 3.1.3 (Slope estimate) [2] we can show write µ,k τ has finite metric slope at time kτ and the metric slope satisfies that µ,k τ ,k−1 ) dW (µ,k τ , µτ . τ For the convenience of readers we write the proof below. Let v ∈ D(φ , kτ ), then  1 2  dW (v, µτ,k−1 ) − d2W (µτ,k−1 , µ,k φ (µ,k τ ) τ ) − φ (v) ≤ 2τ  1 ,k−1 ≤ dW (v, µ,k ) + dW µ,k−1 , µ,k τ )(dW (v, µτ τ τ ) . 2τ This proves the claim, because +   dW (µτ,k−1 , µ,k φ (µ,k 1 τ ) − φ (v) τ ) ,k−1 ,k−1 ,k lim sup ≤ lim sup (d (v, µ ) + d µ , µ ) = . W W τ τ τ ,k 2τ τ dW (v, µτ ) v→µ,k v→µ,k τ

|∂φ (kτ )|(µ,k τ )≤

Next by Proposition 3.3, X τ 1≤k≤T /τ

|∂φ,k (kτ )|(µ,k τ )≤

X 1≤k≤T /τ

,k−1  1 d2W (µ,k ) 2 τ , µτ ≤ C. τ

(3.14)

It can be checked that the constant C above is independent of τ, . d ,k Let t ∈ L2 (dµ,k τ , R ) be a transport map which is µτ −a.e. differentiable and compactly supported in Ω(kτ ). Write ts = (1 − s)i + st. For s small,  ,k φ (µ,k τ ) − φ (ts #µτ )

ON CONTINUITY EQUATIONS IN SPACE-TIME DOMAINS

Z

−u,k τ log(det Dts )dx +

= Ω(kτ )

Z

25

,k h∇V + ∇W ∗ µ,k τ , ts − iidµτ

Ω(kτ ) 2

2

+C kD V kL∞ + kD W kL∞



!

Z

2

|ts − i|

dµ,k τ

.

Ω(kτ )

Due to the expansion det Dts = 1 + str∇(t − i) + o(s) as well as the compact support of t − i in Ω(kτ ), we find the above Z  = hvτ , ts − iidµτ + o dW (µτ , ts #µτ ) . Ω(kτ ) ∇u,k τ  u,k τ

,k ,k ,k ,k + ∇V + ∇W ∗ µ,k Here := τ . Note dW (µτ , ts #µτ ) = sdW (µτ , tµτ ). We divide the above by s and let s → 0 to obtain Z hvτ,k , tidµτ ≤ |∂φ (kτ )|(u,k τ )ktkL2 (dµ,k τ )

vτ,k

Ω(kτ )

Since t can be any compactly supported vector field, we have ! Z ,k 2 ,k |vτ | dµτ ≤ C|∂φ (kτ )|(u,k τ ). Ω(kτ )

Let us define vτ (t) := vτ,k if t ∈ ((k − 1)τ, kτ ]. Then by (3.8), (3.14), vτ ∈ L2 (uτ dxdt, Rd ) with a bound independent of both τ and . Recall µτ → µ in Wasserstein metric uniformly for all t ∈ [0, T ], then by Theorem 5.4.4 [2] vτ converges weakly to some w ∈ L2 (u dxdt, Rd ). From the previous discussion and the equation, we know that vτ converges weakly    to v  =  ∇u u + ∇V + ∇W ∗ µ . It is not hard to see that such limit is unique, so we have v = w a.e.  dµ dt. Then considering the regularity of V, W , we get Z Z ∇u 2 |  |2 u dxdt ≤ 2 |v  |2 dµ dt + C ≤ C u ΩT ΩT where C does not depends on .



Proof. (of Lemma 3.8) By the Euclidean Logarithmic Sobolev inequality (see [15] [10]) and the fact that u (t) is supported in Ω(t), Z Z 1 |∇u |2 d dx). u log u dx ≤ log( 2 2πde Ω(t) u Rd Write F  (t) := F  (µ (t)). Then Z T Z  exp −1 F  (t) dt ≤ C 0

ΩT

|∇u |2 dxdt ≤ C−2 . u

We used Lemma 3.9 in the last inequality. Now for  small enough, assume 2 eN ≥ N holds for all 1 N ≥ − 2 . Thus Z T Z T Z 1 −1 F  (t)dt ≤ − 2 dt + −1 F  (t)dt 1 0

0

1

≤ C− 2 +

Z

T

F  (t)≥ 2

 1 2 exp −1 F  (t) dt ≤ C− 2

0

which finishes the proof.



Remark 3.10. If we have the uniform convergence of µτ → µ for all , then Lemma 3.8 can be proved fairly easier from JKO scheme. Specifically, we expect some kind of exponential formula. The formula for JKO scheme with Wasserstein metric in Rd is proved in [2, 8]. But we cannot prove it in the case that the domain is particularly time-dependent. This can be an interesting direction for future research.

26

Y ZHANG

Figure 2. map (1 − s)tµv + stev



The following lemma is another important ingredient to the proof of convergence. Note it is possible that µ ∈ / the proper domain of φ1 (or equivalently φ ), the plan is to regularize it and replace it by a 2 µ ˜ ∈ Pa . As explained in the introduction, we look for a µ ˜ with density function uniformly bounded by −α for some 0 < α < 1. Additionally we need dµ (µ, µ ˜) to be small where dµ (·, ·) is the PseudoWasserstein metric with base µ . As a remark, this is stronger than requiring dW (µ, µ ˜) to be small. Lemma 3.11. Given any µ ∈ P2 (Ω), v ∈ P2a (Ω) where Ω is a bounded, convex subset of Rd . For any s > 0 small enough, there exists µs ∈ Pa2 (Ω) such that dv (µ, µs ) ≤ Cs and max {µs (x), x ∈ Ω} ≤ s−d . The constant C only depends on the diameter and the volume of Ω. Proof. Without lose of generality, suppose Ω has volume 1 in Euclidean measure. Let e be the Euclidean measure restricted in Ω and then e ∈ Pa2 (Ω). Since v is absolutely continuous, tev and tµv exist and tev is one to one on Ω outside a v zero measure subset. Let    µs := (1 − s)tµv + stev # v be the generalized geodesic joining µ, e with base v, which is defined the same as in Definition 9.2.2 [2]. By the convexity of the domain, µs ∈ P2 (Ω). Moreover by Proposition 2.6.4 [8], the generalized geodesic is of constant speed in the sense that dv (µ, µs ) = sdv (µ, e). Since the domain is bounded, dv (µ, e) is uniformly bounded for all probability measures v, µ, e. We deduce that dv (µ, µs ) ≤ Cs. Now we show the L∞ bound of µs . Let ϕ = χBr (x) which is 1 in the ball and 0 outside. Thus Z Z  ϕdµs = ϕ (1 − s)tµv + stev dv Ω

Z =



n o  −1 ϕ (1 − s)tµv + stev dv = v (1 − s)tµv + stev Br (x)

(3.15)



−1  Write S := (1 − s)tµv + stev Br (x). We claim that, for all s < 1 the map (1 − s)tµv + stev is one to one on some S\Ss such that v(Ss ) = 0. From the results of optimal transport theory (see Theorem 5.10 Kantorovich duality [21]), (i × tµv )# dv and (i × tev )# dv are concentrated on c-cyclically monotone sets. Here in our case the cost function c(x, y) = |x − y|2 . Then for any x, y ∈ Ω outside a v-measure 0 set, by cyclical monotonicity     c x, tµv (x) + c y, tµv (y) ≤ c x, tµv (y) + c y, tµv (x)

ON CONTINUITY EQUATIONS IN SPACE-TIME DOMAINS

27

which simplifies to  (x − y) tµv (y) − tµv (x) ≤ 0.  Similarly (x − y) tev (y) − tev (x) ≤ 0. These two inequalities forbid   (1 − s)tµv + stev (x) = (1 − s)tµv + stev (y), which is equivalent to   (1 − s) tµv (x) − tµv (y) = s tev (y) − tev (x) . (3.16)   If otherwise, (3.16) shows that tµv (x) − tµv (y) , tev (x) − tev (y) are parallel and pointing to opposite directions. This then, as well as the two inequalities above, leads to x = y. We proved the claim. Notice e(tev S) = vol {tev S} = v(S). P By the claim of the one-to-one map, the following holds for µ = i δxi and thus by approximation it holds for all µ. That is   vol {Br (x)} = vol (1 − s)tµv + stev S ≥ vol {stev S} = sd v(S). And so v(S) ≤ vol {B0 (1)} ( rs )d . Recall (3.15), for any such ϕ = χBr (x) we find out Z 1 ϕdµs ≤ s−d . vol {Br (x)} Ω This shows that us is an L∞ function in Ω with L∞ bound s−d .



Now we give our main theorem in the second part of this paper. Theorem 3.3. Assume conditions (C1)-(C4)(O1)(O2) hold and µ0 ∈ D(φ1 , 0). Suppose ΩT is bounded and Ω(t) is convex for all t. Let µ be the unique weak solutions to equations (1.5) and µ be the unique weak solution to equation (1.2). Then µ → µ in Wasserstein metric.     1 Proof. For any ω1 ∈ P2a Ω(t) , let µs := (stω µ + (1 − s)i)# µ with µ = µ (t). Since the domain is assumed to be convex, by direct computation as did in Lemma 3.9 Z φ (µs ) − φ (µ )  1 lim inf ≥ hξ  , tω µ − iidµ s→0 s Ω(t) where ξ  ∈ L2 (µ ; Rd ) belongs to the Fr´echet subdifferential of φ at µ (see section 10 [2]). By (C3) ˜ ˜ = min{λ, 3λ}. So and the definition of λ-convexity, it is not hard to see that φ , φ are λ−convex for λ by the Characterization by Variational inequalities and monotonicity in 10.1.1 [2], ˜ φ (µs ) − φ (µ ) λ ≤ φ (ω1 ) − φ (µ ) − (1 − s)d2W (ω1 , µ ). s 2 Then we take s → 0 and find Z ˜ λ 2   1 φ (ω1 ) − φ (µ ) ≥ hξ  , tω (3.17) µ − iidµ + dW (ω1 , µ ). 2 Ω(t) By direct computation, we can let ξ := ∇V + (∇W ∗ µ) to be the Fr´echet subdifferential of φ at µ and then for any ω2 ∈ P2 (Ω(t)) Z ˜ λ 2 2 φ(ω2 ) − φ(µ) ≥ hξ, tω (3.18) µ − iidµ + dW (ω2 , µ). 2 Ω(t) We want to take ω1 = µ. However it is possible that µ ∈ / D(φ , t), we use a modification of µ by 1  d+2 Lemma 3.11 by taking v = µ , s =  . Thus let µ ˜ = µs ∈ P2a (ΩT ) and µ ˜=u ˜Ld , then for all 0 ≤ t ≤ T d

max {˜ u(x, t)} ≤ − d+2 ,

1

dµ (˜ µ, µ) ≤ C d+2 .

28

Y ZHANG

Plug in ω1 = µ ˜, Z

  φ µ ˜(t) − φ µ (t) ≥

hξ  , tµµ˜ − iidµ +

Ω(t)

Z

hξ  , tµµ − iidµ +



Z

Ω(t)

hξ  , tµµ˜ − tµµ idµ +

Ω(t)

˜ λ d2 (˜ µ, µ ) 2 W

˜ 1 λ (dW (µ, µ ) + C d+2 )2 . 2



Let γ be an optimal transport plan between µ, µ . The above Z Z 2 ≥ hξ  (y), x − yidγ  + hξ  , tµµ˜ − tµµ idµ − Cd2W (µ, µ ) − C d+2 . Ω(t)2

(3.19)

Ω(t)

Similarly   φ µ (t) − φ µ(t) ≥

Z

hξ(x), y − xidγ  +

Ω(t)

˜ λ d2 (µ, µ ). 2 W

(3.20)

As already proved in Lemma 3.9, for some constant C independent of , t Z Z  12 Z  12 µ ˜ µ 2 ˜ µ  2    µ  |ξ | dµ dt |tµ − tµ | dµ dt . hξ , tµ − tµ idµ dt ≤ ΩT

ΩT

Notice

R ΩT

 2

ΩT



|ξ | dµ dt is uniformly bounded and ! 21

Z

|tµµ˜



tµµ |2 dµ

= dµ (µ, µ ˜)

Ω(t)

is the Pseudo-Wasserstein distance induced by µ ∈ P2a . So Z Z T 1 ˜ µ  µ  dµ (µ, µ ˜)dt ≤ C d+2 . hξ , tµ − tµ idµ dt ≤ C 0

ΩT

Now combining the above and inequalities (3.19) (3.20), we get Z Z Z h−ξ(x) + ξ  (y), x − yidγ  dt ≤  (˜ u log u ˜ − u log u )dxdt + C ΩT

2

ΩT

T

1

d2W (µ, µ )dt + C d+2 . (3.21)

0

By Theorem 8.4.7 and Lemma 4.3.4 from [2], Z d 2  d (µ, µ ) = 2 hv(x) − v  (y), x − yidγ  2 dt W Ω(t) where v and v  are µ and µ ’s tangent vector fields respectively. By the JKO scheme, µ is a gradient flow solution and we can choose ξ  = −v  . Recall (2.26), v = Pt (−ξ). Since the domain is convex, the following holds for all x ∈ ∂Ω(t), y ∈ Ω. h−Pt (−ξ)(x) − ξ(x), x − yi ≤ 0. Also we know that (−Pt (−ξ) − ξ) is supported on the boundary. This shows Z d 2  d (µ, µ ) ≤ 2 h−ξ(x) + ξ  (y), x − yidγ  . (3.22) 2 dt W Ω(t)  Then by Lemma 3.8, F  (µ ) converges to 0 in L1 (0, T ), dt as  → 0. Also F  (˜ µ) converges to 0 d − d+2 uniformly, because u ˜≤ and the domain is bounded. Then by (3.21), (3.22) and d2W (µ, µ )(0) = 0, we deduce that Z T0 2  0 dW (µ, µ )(T ) ≤ C d2W (µ, µ )dt + δ() 0 0

1

where 0 < T ≤ T and δ() → 0 as  → 0. From the above, δ() can be chosen to be  d+2 for small . Finally Gronwall’s inequality finishes the proof. 

ON CONTINUITY EQUATIONS IN SPACE-TIME DOMAINS

29

Acknowledgements. The author would like to thank Inwon Kim for suggesting the problem as well as all the stimulating guidance and discussions. The author would like to thank Katy Craig, Alp´ ar Rich´ ard M´esz´ aros and Wilfrid Gangbo for fruitful discussions. Also the author thank Katy Craig for mentioning the work [7]. References [1] L Ambrosio and G Savar´ e. Gradient flows of probability measures, handbook of differential equations, evolutionary equations 3, ed. by cm dafermos and e. feireisl, 2007. [2] Luigi Ambrosio, Nicola Gigli, and Giuseppe Savar´ e. Gradient flows: in metric spaces and in the space of probability measures. Springer Science & Business Media, 2008. [3] JA Carrillo, M Difrancesco, A Figalli, T Laurent, and D Slepcev. Global-in-time weak measure solutions, finite-time aggregation and confinement for nolocal interaction equations. 2009. [4] Jos´ e A Carrillo, Dejan Slepˇ cev, and Lijiang Wu. Nonlocal-interaction equations on uniformly prox-regular sets. arXiv preprint arXiv:1405.1111, 2014. [5] Jos´ e Antonio Carrillo, Stefano Lisini, and Edoardo Mainini. Gradient flows for non-smooth interaction potentials. Nonlinear Analysis: Theory, Methods & Applications, 100:122–147, 2014. [6] Francis H Clarke, RJ Stern, and PR Wolenski. Proximal smoothness and the lower-c2 property. J. Convex Anal, 2(1/2):117–144, 1995. [7] Elaine Cozzi, Gung-Min Gie, and James P Kelliher. The aggregation equation with newtonian potential. arXiv preprint arXiv:1608.01348, 2016. [8] Katy Craig. The exponential formula for the wasserstein metric. ESAIM: Control, Optimisation and Calculus of Variations, 22(1):169–187, 2016. [9] Guido De Philippis, Alp´ ar Rich´ ard M´ esz´ aros, Filippo Santambrogio, and Bozhidar Velichkov. Bv estimates in optimal transportation and applications. Archive for Rational Mechanics and Analysis, 219(2):829–860, 2016. [10] Manuel Del Pino and Jean Dolbeault. The optimal euclidean l p-sobolev logarithmic inequality. Journal of Functional Analysis, 197(1):151–161, 2003. [11] Simone Di Marino, Bertrand Maury, and Filippo Santambrogio. Measure sweeping processes. 2015. [12] Simone Di Marino and Alp´ ar Rich´ ard M´ esz´ aros. Uniqueness issues for evolutive equations with density constraints. arXiv preprint arXiv:1507.02900, 2015. [13] Jean Fenel Edmond and Lionel Thibault. Bv solutions of nonconvex sweeping process differential inclusion with perturbation. Journal of Differential Equations, 226(1):135–179, 2006. [14] Irene Fonseca and Giovanni Leoni. Modern Methods in the Calculus of Variations: Lˆ p Spaces. Springer Science & Business Media, 2007. [15] Leonard Gross. Logarithmic sobolev inequalities. American Journal of Mathematics, 97(4):1061–1083, 1975. [16] Richard Jordan, David Kinderlehrer, and Felix Otto. The variational formulation of the fokker–planck equation. SIAM journal on mathematical analysis, 29(1):1–17, 1998. [17] Hans G Kellerer. Duality theorems for marginal problems. Zeitschrift f¨ ur Wahrscheinlichkeitstheorie und verwandte Gebiete, 67(4):399–432, 1984. [18] Robert J McCann et al. Existence and uniqueness of monotone measure-preserving maps. Duke Mathematical Journal, 80(2):309–324, 1995. [19] Felix Otto. L 1-contraction and uniqueness for quasilinear elliptic–parabolic equations. Journal of differential equations, 131(1):20–38, 1996. [20] Luca Petrelli and Adrian Tudorascu. Variational principle for general diffusion problems. Applied Mathematics and Optimization, 50(3):229–257, 2004. [21] C´ edric Villani. Optimal transport: old and new, volume 338. Springer Science & Business Media, 2008. [22] Lijiang Wu and Dejan Slepˇ cev. Nonlocal interaction equations in environments with heterogeneities and boundaries. Communications in Partial Differential Equations, 40(7):1241–1281, 2015. Department of Mathematics, University of California, Los Angeles, USA E-mail address: [email protected]