12.07.2015 Views

Lp Theory for the Multidimensional Aggregation Equation

Lp Theory for the Multidimensional Aggregation Equation

Lp Theory for the Multidimensional Aggregation Equation

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

(ii) If α ≥ 2 <strong>the</strong>n <strong>the</strong> aggregation equation is globally well posed in P 2 (R d ) ∩ L p (R d ) <strong>for</strong> every p > 1.As a consequence we have existence and uniqueness <strong>for</strong> all potentials which are less singular than <strong>the</strong>Newtonian potential K(x) = |x| 2−d at <strong>the</strong> origin. In two dimensions this includes potentials with cuspsuch as K(x) = |x| 1/2 . In three dimensions this includes potentials that blow up such as K(x) = |x| −1/2 .From [5, 16] we know that <strong>the</strong> support of compactly supported solutions shrinks to a point in finite time,proving <strong>the</strong> second assertion in point (i) above. The first part of (i) and statement (ii) are direct corollaryof Theorem 1, Definition 2 and <strong>the</strong> fact that α ≥ 2 implies ∆K bounded.In <strong>the</strong> case where α = 1, i.e. K(x) ∼ |x| as |x| → 0, <strong>the</strong> previous Theorem gives local well posedness inP 2 (R d ) ∩ L p (R d ) <strong>for</strong> all p > p s =dd−1. The next Theorem shows that it is not possible to obtain local wellposedness in P 2 (R d ) ∩ L p (R d ) <strong>for</strong> p < p s =dd−1 .Theorem 4 (Critical p-exponent to generate instantaneous mass concentration). Suppose K(x) = |x| in aneighborhood of <strong>the</strong> origin, and suppose ∇K is compactly supported (or decays exponentially fast at infinity).Then, <strong>for</strong> any p < p s =dd−1 , <strong>the</strong>re exists initial data in P 2(R d ) ∩ L p (R d ) <strong>for</strong> which a delta Dirac appearsinstantaneously in <strong>the</strong> measure solution.In order to make sense of <strong>the</strong> statement of <strong>the</strong> previous Theorem, we need a concept of measure solution.The potentials K(x) = |x| is semi-convex, i.e. <strong>the</strong>re exist λ ∈ R such that K(x) − λ 2 |x|2 is convex. In[16], Carrillo et al. prove global well-posedness in P 2 (R d ) of <strong>the</strong> aggregation equation with semi-convexpotentials. The solutions in [16] are weak measure solutions - <strong>the</strong>y are not necessarily absolutely continuouswith respect to <strong>the</strong> Lebesgue measure. Theorems 3 and 4 give a sharp condition on <strong>the</strong> initial data in order<strong>for</strong> <strong>the</strong> solution to stay absolutely continuous with respect to <strong>the</strong> Lebesgue measure <strong>for</strong> short time. Theorem4 is proven in Section 4.Finally, in section 5 we consider a class of potential that will be referred to as <strong>the</strong> class of naturalpotentials. A potential is said to be natural if it satisfies thata) it is a radially symmetric potential, i.e.: K(x) = k(|x|),b) it is smooth away from <strong>the</strong> origin and it’s singularity at <strong>the</strong> origin is not worse than Lipschitz,c) it doesn’t exhibit pathological oscillation at <strong>the</strong> origin,d) its derivatives decay fast enough at infinity.All <strong>the</strong>se conditions will be more rigorously stated later. It will be shown that <strong>the</strong> gradient of naturalpotentials automatically belongs to W 1,q <strong>for</strong> q < d, <strong>the</strong>re<strong>for</strong>e, using <strong>the</strong> results from <strong>the</strong> sections 2 and 3,we have local existence and uniqueness in P 2 (R d ) ∩ L p (R d ), p >dd−1 .A natural potential is said to be repulsive in <strong>the</strong> short range if it has a local maximum at <strong>the</strong> originand it is said to be attractive in <strong>the</strong> short range if it has a local minimum at <strong>the</strong> origin. If <strong>the</strong> maximum(respectively minimum) is strict, <strong>the</strong> natural potential is said to be strictly repulsive (respectively strictlyattractive) at <strong>the</strong> origin. The main <strong>the</strong>orem of section 5 is <strong>the</strong> following:Theorem 5 (Osgood condition <strong>for</strong> global well posedness). Suppose K is a natural potential.(i) If K is repulsive in <strong>the</strong> short range, <strong>the</strong>n <strong>the</strong> aggregation equation is globally well posed in P 2 (R d ) ∩L p (R d ), p > d/(d − 1).(ii) If K is strictly attractive in <strong>the</strong> short range, <strong>the</strong> aggregation equation is globally well posed in P 2 (R d ) ∩L p (R d ), p > d/(d − 1), if and only ifr ↦→ 1k ′ (r)is not integrable at 0. (1.9)3


By globally well posed in P 2 (R d ) ∩ L p (R d ), we mean that <strong>for</strong> any initial data in P 2 (R d ) ∩ L p (R d ) <strong>the</strong>unique solution of <strong>the</strong> aggregation equation will exist <strong>for</strong> all time and will stay in P 2 (R d ) ∩ L p (R d ) <strong>for</strong> alltime.Notice that <strong>the</strong> exponent d/(d − 1) is not sharp in this <strong>the</strong>orem.Condition (1.9) will be refered as <strong>the</strong> Osgood condition. It is easy to understand why <strong>the</strong> Osgoodcondition is relevant while studying blowup: <strong>the</strong> quantityT (d) =∫ d0drk ′ (r)can be thought as <strong>the</strong> amount of time it takes <strong>for</strong> a particle obeying <strong>the</strong> ODE Ẋ = −∇K(X) to reach <strong>the</strong>origin if it starts at a distance d from it. For a potential satisfying <strong>the</strong> Osgood condition, T (d) = +∞, whichmeans that <strong>the</strong> particle can not reach <strong>the</strong> origin in finite time. The Osgood condition was already shownin [5] to be necessary and sufficient <strong>for</strong> global well posedness of L ∞ -solutions. Extension to L p requires L pestimates ra<strong>the</strong>r than L ∞ estimates. See also [43] <strong>for</strong> an example of <strong>the</strong> use of <strong>the</strong> Osgood condition in <strong>the</strong>context of <strong>the</strong> Euler equations <strong>for</strong> incompressible fluid.The “only if” part of statement (ii) was proven in [5] and [16]. In <strong>the</strong>se two works it was shown thatif (1.9) is not satisfied, <strong>the</strong>n compactly supported solutions will collapse into a point mass – and <strong>the</strong>re<strong>for</strong>eleave L p – in finite time. In section 5 we prove statement (i) and <strong>the</strong> ’if’ part of statement (ii).2 Existence of L p -solutionsIn this section we show that if <strong>the</strong> interaction potential satisfies∇K ∈ W 1,q (R d ), 1 < q < +∞, (2.10)and if <strong>the</strong> initial data is nonnegative and belongs to L p (R d ) (p and q are Hölder conjugates) <strong>the</strong>n <strong>the</strong>re existsa solution to <strong>the</strong> aggregation equation. Moreover, ei<strong>the</strong>r this solution exists <strong>for</strong> all times, or its L p -normblows up in finite time. The duality between L p and L q guarantees enough smoothness in <strong>the</strong> velocity fieldv = −∇K ∗ u to define characteristics. We use <strong>the</strong> characteristics to construct a solution. The argumentis inspired by <strong>the</strong> existence of L ∞ solutions of <strong>the</strong> incompressible 2D Euler equations by Yudovich [44] andof L ∞ solutions of <strong>the</strong> aggregation equation [4]. Section 3 proves uniqueness provided u ∈ P 2 . We prove inTheorem 18 of <strong>the</strong> present section that if u 0 ∈ P 2 <strong>the</strong>n <strong>the</strong> solution stays in P 2 . Finally we prove that if inaddition to (2.10), we haveess sup ∆K < +∞, (2.11)<strong>the</strong>n <strong>the</strong> solution constructed exists <strong>for</strong> all time.Most of <strong>the</strong> section is devoted to <strong>the</strong> proof of <strong>the</strong> following <strong>the</strong>orem:Theorem 6 (Local existence). Consider 1 < q < ∞ and p its Hölder conjugate. Suppose ∇K ∈ W 1,q (R d )and suppose u 0 ∈ L p (R d ) is nonnegative. Then <strong>the</strong>re exists a time T ∗ > 0 and a nonnegative functionu ∈ C([0, T ∗ ], L p (R d )) ∩ C 1 ([0, T ∗ ], W −1,p (R d )) such thatu ′ (t) + div ( u(t) v(t) ) = 0 ∀t ∈ [0, T ∗ ], (2.12)Moreover <strong>the</strong> function t → ‖u(t)‖ p Lis differentiable and satisfiesp∫ddt {‖u(t)‖p L p} = −(p − 1)The choice of <strong>the</strong> spacev(t) = −u(t) ∗ ∇K ∀t ∈ [0, T ∗ ], (2.13)u(0) = u 0 . (2.14)R d u(t, x) p div v(t, x) dx ∀t ∈ [0, T ∗ ]. (2.15)Y p := C([0, T ∗ ], L p (R d )) ∩ C 1 ([0, T ∗ ], W −1,p (R d ))is motivated by <strong>the</strong> fact that, if u ∈ Y p and ∇K ∈ W 1,q , <strong>the</strong>n <strong>the</strong> velocity field is automatically C 1 in spaceand time:4


Lemma 7. Consider 1 < q < ∞ and p its Hölder conjugate. If ∇K ∈ W 1,q (R d ) and u ∈ Y p <strong>the</strong>nu ∗ ∇K ∈ C 1 ( [0, T ∗ ] × R d) and ‖u ∗ ∇K‖ C1 ([0,T ∗ ]×R d ) ≤ ‖∇K‖ W 1,q (R d ) ‖u‖ Y p(2.16)where <strong>the</strong> norm ‖·‖ C 1 ([0,T ∗ ]×R d ) and ‖·‖ Y pare defined by∣ ∣ ∣ ∣∣∣‖v‖ C1 ([0,T ∗ ]×R d ) = sup∂vd∑ ∣∣∣|v| + sup[0,T ∗ ]×R d [0,T ∗ ]×R ∂t ∣ + ∂v ∣∣∣sup , (2.17)d [0,T ∗ ]×R ∂x d i‖u‖ Yp= sup ‖u(t)‖ <strong>Lp</strong> (R d ) +t∈[0,T ∗ ]supt∈[0,T ∗ ]i=1‖u ′ (t)‖ W −1,p (R d ) . (2.18)Proof. Recall that <strong>the</strong> convolution between a L p -function and a L q -function is continuous and sup x∈R d |f ∗g(x)| ≤ ‖f‖ L p ‖g‖ L q There<strong>for</strong>e, since ∇K and ∇K xi are in L q , <strong>the</strong> mappingf ↦→ ∇K ∗ fis a bounded linear trans<strong>for</strong>mation from L p (R d ) to C 1 (R d ), where C 1 (R d ) is endowed with <strong>the</strong> norm‖f‖ C 1= supx∈R d |f(x)| +d∑sup | ∂f (x)|.x∈R ∂x d iSince u ∈ C([0, T ∗ ], L p ) it is <strong>the</strong>n clear that u ∗ ∇K ∈ C([0, T ∗ ], C 1 ). In particular w(t, x) = (u(t) ∗ ∇K) (x)and ∂w∂x i(t, x) are continuous on [0, T ∗ ] × R d . Let us now show that ∂w∂t(t, x) exists and is continuous on[0, T ∗ ] × R d . Since u ′ (t) ∈ C([0, T ∗ ], W −1,p ) and ∇K ∈ W 1,q , we havei=1∂w∂t (t, x) = − (u′ (t) ∗ ∇K) (x) = −〈u ′ (t), τ x ∇K〉where 〈 , 〉 denote <strong>the</strong> pairing between <strong>the</strong> two dual spaces W −1,p (R d ) and W 1,q (R d ), and τ x denote <strong>the</strong>translation by x. Since x ↦→ τ x ∇K is a continuous mapping from R d to W 1,q it is clear that ∂w∂t(t, x) iscontinuous with respect to space. The continuity with respect to time come from <strong>the</strong> continuity of u ′ (t) withrespect to time. Inequality (2.16) is easily obtained.Remark 8. Let us point out that (2.12) indeed makes sense, when understood as an equality in W −1,p .Since v ∈ C([0, T ∗ ], C 1 (R d )) one can easily check that uv ∈ C([0, T ∗ ], L p (R d )). Also recall that <strong>the</strong> injectioni : L p (R d ) → W −1,p (R d ) and <strong>the</strong> differentiation ∂ xi : L p (R d ) → W −1,p (R d ) are bounded linear operators.There<strong>for</strong>e it is clear that both u and div(uv) belong to C([0, T ∗ ], W −1,p (R d )). <strong>Equation</strong> (2.12) has to beunderstand as an equality in W −1,p .The rest of this section is organized as follows. First we give <strong>the</strong> basic a priori estimates in subsection 2.1.Then, in subsection 2.2, we consider a mollified and cutted-off version of <strong>the</strong> aggregation equation <strong>for</strong> whichwe have global existence of smooth and compactly supported solutions. In subsection 2.3 we show that <strong>the</strong>characteristics of this approximate problem are uni<strong>for</strong>mly Lipschitz continuous on [0, T ∗ ]×R d , where T ∗ > 0is some finite time depending on ‖u 0 ‖ L p. In subsection 2.4 we pass to <strong>the</strong> limit in C([0, T ∗ ], L p ). To do thiswe need <strong>the</strong> uni<strong>for</strong>m Lipschitz bound on <strong>the</strong> characteristics toge<strong>the</strong>r with <strong>the</strong> fact that <strong>the</strong> translation byx, x ↦→ τ x u 0 , is a continuous mapping from R d to L p (R d ). In subsection 2.5 we prove three <strong>the</strong>orems. Wefirst prove continuation of solutions. We <strong>the</strong>n prove that L p -solutions which start in P 2 stay in P 2 as longas <strong>the</strong>y exist. And finally we prove global existence in <strong>the</strong> case where ∆K is bounded from above.2.1 A priori estimatesSuppose u ∈ Cc 1 ((0, T ) × R d ) is a nonnegative function which satisfies (2.12)-(2.13) in <strong>the</strong> classical sense.Suppose also that K ∈ Cc∞ (R d ). Integrating by part, we obtain that <strong>for</strong> any p ∈ (1, +∞):∫∫du(t, x) p dx = −(p − 1) u(t, x) p div v(t, x)dx ∀t ∈ (0, T ). (2.19)dt R d R d5


As a consequence we have:ddt ‖u(t)‖p L ≤ (p − 1)‖div v(t)‖ p L ∞‖u(t)‖p <strong>Lp</strong> ∀t ∈ (0, T ), (2.20)and by Hölder’s inequality:ddt ‖u(t)‖p L ≤ (p − 1)‖∆K‖ pL q‖u(t)‖p+1 L . (2.21)pWe now derive L ∞ estimates <strong>for</strong> <strong>the</strong> velocity field v = −∇K ∗ u and its derivatives. Hölder’s inequalityeasily givesSince ∂v∂t∂u= −∇K ∗∂t ∣|v(t, x)| ≤ ‖u(t)‖ L p‖∇K‖ L q ∀(t, x) ∈ (0, T ) × R d , (2.22)∥ ∂v j∣ (t, x)∂x i∣ ≤ ‖u(t)‖ L p∂ 2 K ∥∥∥L∥∀(t, x) ∈ (0, T ) × R d . (2.23)∂x i dx j q= ∇K ∗ div(uv) = ∆K ∗ uv we have∣∂v ∣∣∣∂t (t, x) ≤ ‖u(t)v(t)‖ L p‖∆K‖ L q ≤ ‖u(t)‖ L p‖v(t)‖ L ∞‖∆K‖ L q,which in light of (2.22) gives∣ ∂v ∣∣∣∣ ∂t (t, x) ≤ ‖u(t)‖ 2 L p‖∇K‖ L q‖∆K‖ L q ∀(t, x) ∈ (0, T ) × Rd . (2.24)2.2 Approximate smooth compactly supported solutionsIn this section we deal with a smooth version of equation (2.12)-(2.14). Suppose u 0 ∈ L p (R d ), 1 < p < +∞,and ∇K ∈ W 1,q (R d ). Consider <strong>the</strong> approximate problemu t + div(uv) = 0 in (0, +∞) × R d , (2.25)v = −∇K ɛ ∗ u in (0, +∞) × R d , (2.26)u(0) = u ɛ 0, (2.27)where K ɛ = J ɛ K, u ɛ 0 = J ɛ u 0 and J ɛ is an operator which mollifies and cuts-off, J ɛ f = (fM Rɛ ) ∗ η ɛ whereη ɛ (x) is a standard mollifier:η ɛ (x) = 1 ( x)∫ɛ d η , η ∈ Cc ∞ (R d ), η ≥ 0, η(x)dx = 1,ɛR dand M Rɛ (x) is a standard cut-off function: M Rɛ (x) = M( x R ɛ), R ɛ → ∞ as ɛ → 0,M ∈ C ∞ c (R d ),Let τ x denote <strong>the</strong> translation by x, i.e.:⎧⎨⎩M(x) = 1 if |x| ≤ 1,0 < M(x) < 1 if 1 < |x| < 2,M = 0 if 2 ≤ |x|.τ x f(y) := f(y − x).It is well known that given a fixed f ∈ L r (R d ), 1 < r < +∞, <strong>the</strong> mapping x ↦→ τ x f from R d to L r (R d )is uni<strong>for</strong>mly continuous. In (iv) of <strong>the</strong> next lemma we show a slightly stronger result which will be neededlater.Lemma 9 (Properties of J ɛ ). Suppose f ∈ L r (R d ), 1 < r < +∞, <strong>the</strong>n6


(i) J ɛ f ∈ C ∞ c (R d ),(ii) ‖J ɛ f‖ L r ≤ ‖f‖ L r,(iii) lim ɛ→0 ‖J ɛ f − f‖ L r = 0,(iv) The family of mappings x ↦→ τ x J ɛ f from R d to L r (R d ) is equicontinuous, i.e.: <strong>for</strong> each δ > 0, <strong>the</strong>re isa η > 0 independent of ɛ such that ‖τ x J ɛ f − τ y J ɛ f‖ L r ≤ δ if |x − y| ≤ η.Proof. Statements (i) and (ii) are obvious. If f is compactly supported, one can easily prove (iii) by notingthat fM Rɛ = f <strong>for</strong> ɛ small enough. If f is not compactly supported, (iii) is obtained by approximating f bya compactly supported function and by using (ii). Let us now turn to <strong>the</strong> proof of (iv). Using (ii) we obtain‖τ x J ɛ f − J ɛ f‖ L r ≤ ‖(τ x M Rɛ )(τ x f) − M Rɛ f)‖ L r≤ ‖τ x M Rɛ − M Rɛ ‖ L ∞‖τ x f‖ L r + ‖M Rɛ ‖ L ∞‖τ x f − f‖ L r.Because x ↦→ τ x f is continuous, <strong>the</strong> second term can be made as small as we want by choosing |x| smallenough. Since ‖τ x M Rɛ − M Rɛ ‖ L ∞ ≤ ‖∇M Rɛ ‖ L ∞|x| ≤ 1 R ɛ‖∇M‖ L ∞|x|, <strong>the</strong> first term can be made as smallas we want by choosing |x| small enough and independently of ɛ.Proposition 10 (Global existence of smooth compactly-supported approximates). Given ɛ, T > 0, <strong>the</strong>reexists a nonnegative function u ∈ C 1 c ((0, T ) × R d ) which satisfy (2.27) in <strong>the</strong> classical sense.Proof. Since u ɛ 0 and K ɛ belong to Cc∞function u ɛ satisfying(R d ), we can use <strong>the</strong>orem 3 p. 1961 of [28] to get <strong>the</strong> existence of au ɛ ∈ L ∞ (0, T ; H k ), u ɛ t ∈ L ∞ , (0, T ; H k−1 ) <strong>for</strong> all k, (2.28)u ɛ t + div (u ɛ (−∇K ɛ ∗ u ɛ )) = 0 in (0, T ) × R d , (2.29)u ɛ (0) = u ɛ 0, (2.30)u ɛ (t, x) ≥ 0 <strong>for</strong> a.e. (t, x) ∈ (0, T ), ×R d . (2.31)Statement (2.28) implies that u ɛ ∈ C((0, T ); H k−1 ). Using <strong>the</strong> continuous embedding H k−1 (R d ) ⊂ C 1 (R d )<strong>for</strong> k large enough we find that u ɛ and u ɛ x i, 1 ≤ i ≤ d, are continuous on (0, T ) × R d . Finally, (2.29) showsthat u ɛ t is also continuous on (0, T ) × R d . We have proven that u ɛ ∈ C 1 ((0, T ) × R d ). It is <strong>the</strong>n obvious thatv ɛ = −∇K ɛ ∗ u ɛ ∈ C 1 ((0, T ) × R d ). Note moreover that|v ɛ (x, t)| ≤ ‖u ɛ ‖ L∞ (0,T ;L 2 )‖∇K ɛ ‖ L 2<strong>for</strong> all (t, x) ∈ (0, T ) × R d . This combined with <strong>the</strong> fact that v ɛ is in C 1 shows that <strong>the</strong> characteristics arewell defined and propagate with finite speed. This proves that u ɛ is compactly supported in (0, T ) × R d(because u ɛ 0 is compactly supported in R d ).2.3 Study of <strong>the</strong> velocity field and <strong>the</strong> induced flow mapNote that K ɛ and u ɛ are in <strong>the</strong> right function spaces so that we can apply to <strong>the</strong>m to <strong>the</strong> a priori estimatesderived in section 2.1. In particular we have:ddt ‖uɛ (t)‖ L p ≤ (p − 1)‖∆K‖ L q‖u ɛ (t)‖ p+1L p ,‖u ɛ (0)‖ L p ≤ ‖u 0 ‖ L p.Using Gronwall inequality and <strong>the</strong> estimate on <strong>the</strong> supremum norm of <strong>the</strong> derivatives derived in section 2.1we obtain:7


Lemma 11 (uni<strong>for</strong>m bound <strong>for</strong> <strong>the</strong> smooth approximates). There exists a time T ∗ > 0 and a constantC > 0, both independent of ɛ, such that‖u ɛ (t)‖ L p ≤ C ∀t ∈ [0, T ∗ ], (2.32)|v ɛ t, x)|, |v ɛ x i(t, x)|, |v ɛ t(t, x)| ≤ C ∀(t, x) ∈ [0, T ∗ ] × R d . (2.33)From (2.33) it is clear that <strong>the</strong> family {v ɛ } is uni<strong>for</strong>mly Lipschitz on [0, T ∗ ] × R d , with Lipschitz constantC. We can <strong>the</strong>re<strong>for</strong>e use <strong>the</strong> Arzela-Ascoli Theorem to obtain <strong>the</strong> existence of a continuous function v(t, x)such thatv ɛ → v uni<strong>for</strong>mly on compact subset of [0, T ∗ ] × R d . (2.34)It is easy to check that this function v is also Lipschitz continuous with Lipschitz constant C. The Lipschitzand bounded vector field v ɛ generates a flow map X ɛ (t, α), t ∈ [0, T ∗ ], α ∈ R d :∂X ɛ (t, α)= v ɛ (X ɛ (t, α), t),∂tX ɛ (0, α) = α,where we denote by X t ɛ : R d → R d <strong>the</strong> mapping α ↦→ X ɛ (t, α) and by X −tɛ <strong>the</strong> inverse of X t ɛ.The uni<strong>for</strong>m Lipschitz bound on <strong>the</strong> vector field implies uni<strong>for</strong>m Lipschitz bound on <strong>the</strong> flow map andits inverse (see <strong>for</strong> example [4] <strong>for</strong> a proof of this statement) we <strong>the</strong>re<strong>for</strong>e have:Lemma 12 (uni<strong>for</strong>m Lipschitz bound on Xɛ t and Xɛ−t ). There exists a constant C > 0 independent of ɛsuch that:(i) <strong>for</strong> all t ∈ [0, T ∗ ] and <strong>for</strong> all x 1 , x 2 ∈ R d|Xɛ(x t 1 ) − Xɛ(x t 2 )| ≤ C|x 1 − x 2 | and |Xɛ−t (x 1 ) − x −tɛ (x 2 )| ≤ C|x 1 − x 2 |,(ii) <strong>for</strong> all t 1 , t 2 ∈ [0, T ∗ ] and <strong>for</strong> all x ∈ R d|X t1ɛ (x) − X t2ɛ (x)| ≤ C|t 1 , t 2 | and |Xɛ−t1(x) − X −t2ɛ (x)| ≤ C|t 1 − t 2 |.The Arzela-Ascoli Theorem <strong>the</strong>n implies that <strong>the</strong>re exists mapping X t and X −t such thatX t ɛ k(x) → X t (x) uni<strong>for</strong>mly on compact subset of [0, T ∗ ] × R d ,X −tɛ k(x) → X −t (x) uni<strong>for</strong>mly on compact subset of [0, T ∗ ] × R d .Moreover it is easy to check that X t and X −t inherit <strong>the</strong> Lipschitz bounds of X t ɛ and X −tɛ .Since <strong>the</strong> mapping X t : R d → R d is Lipschitz continuous, by Rademecher’s Theorem it is differentiablealmost everywhere. There<strong>for</strong>e it makes sense to consider its Jacobian matrix DX t (α). Because of Lemma12-(i) we know that <strong>the</strong>re exists a constant C independent of t and ε such thatsupαɛR d |det DX t (α)| ≤ C and supα∈R d |det DX t ɛ(α)| ≤ C.By <strong>the</strong> change of variable we <strong>the</strong>n easily obtain <strong>the</strong> following Lemma:Lemma 13. The mappings f ↦→ f ◦ X −t and f ↦→ f ◦ Xɛ−t , t ∈ [0, T ∗ ], ɛ > 0, are bounded linear operatorsfrom L p (R d ) to L p (R d ). Moreover <strong>the</strong>re exists a constant C ∗ independent of t and ɛ such that‖f ◦ X −t ‖ L p ≤ C ∗ ‖f‖ L p and ‖f ◦ X −tɛ ‖ L p ≤ C ∗ ‖f‖ L p <strong>for</strong> all f ∈ L p (R d ).8


Note that Lemma 12-(ii) impliesand <strong>the</strong>re<strong>for</strong>eThis gives us <strong>the</strong> following lemma:Lemma 14. let Ω be a compact subset of R d , <strong>the</strong>nwhere <strong>the</strong> compact set Ω + Ct is defined by2.4 Convergence in C([0, T ∗ ), L p )|X t ɛ(α) − α| ≤ Ct <strong>for</strong> all (t, α) ∈ [0, T ∗ ) × R d ,|X t (α) − α| ≤ Ct <strong>for</strong> all (t, α) ∈ [0, T ∗ ) × R d .X t ɛ(Ω) ⊂ Ω + Ct and X t (Ω) ⊂ Ω + Ct,Ω + Ct := {x ∈ R d : dist (x, Ω) ≤ Ct}.Since u ɛ and v ɛ = u ɛ ∗ ∇K are C 1 functions which satisfyu ɛ t + v ɛ · ∇u ɛ = −(div v ɛ )u ɛ and u ɛ (0) = u ɛ 0 (2.35)we have <strong>the</strong> simple representation <strong>for</strong>mula <strong>for</strong> u ɛ (t, x), t ∈ [0, T ∗ ), x ∈ R d :u ɛ (t, x) = u ɛ 0(Xɛ−t (x))e − R tdiv 0 vɛ (s,X −(t−s)ɛ(x))ds = u ɛ 0(X −t (x)) a ɛ (t, x).Lemma 15. There exists a function a(t, x) ∈ C 1 ([0, T ∗ ) × R d ) and a sequence ɛ k → 0 such thata ɛ k(t, x) → a(t, x) uni<strong>for</strong>mly on compact subset of [0, T ∗ ] × R d . (2.36)Proof. By <strong>the</strong> Arzela-Ascoli Theorem, it is enough to show that <strong>the</strong> familyb ɛ (t, x) :=∫ t0div v ɛ (s, Xɛ−(t−s) (x))dxis equicontinuous and uni<strong>for</strong>mly bounded. The uni<strong>for</strong>m boundedness simply come from <strong>the</strong> fact that|div v ɛ | = |u ɛ ∗ ∆K ɛ | ≤ ‖u ɛ ‖ L p‖∆K‖ L q.Let us now prove equicontinuity in space, i.e., we want to prove that <strong>for</strong> each δ > 0, <strong>the</strong>re is η > 0 independentof ɛ and t such that|b ɛ (t, x 1 ) − b ɛ (t, x 2 )| ≤ δ if |x 1 − x 2 | ≤ η.First, note that by Hölder’s inequality we haveɛ|b ɛ (t, x 1 ) − b ɛ (t, x 2 )| ≤∫ t0‖u ɛ (s)‖ L p‖τ ξ J ɛ ∆K − τ ζ J ɛ ∆K‖ L qdswhere ξ stands <strong>for</strong> Xɛ−(t−s) (x 1 ) and ζ <strong>for</strong> Xɛ−(t−s) (x 2 ). Then equicontinuity in space is a consequence ofLemma 9 (iv) toge<strong>the</strong>r with <strong>the</strong> fact thatwhere C is independent of t, s and ɛ.|Xɛ−(t−s) (x 1 ) − Xɛ −(t−s) (x 2 )| ≤ C|x 1 − x 2 |9


Let us finally prove equicontinuity in time. First not that, assuming that t 1 < t 2 ,b ɛ (t 1 , x) − b ɛ (t 2 , x) =−∫ t10∫ t2t 1div v ɛ (s, X −(t1−s) (x)) − div v ɛ (s, X −(t2−s) (x))dsdiv v ɛ (s, Xɛ−(t2−s) (x))ds.Since div v ɛ is uni<strong>for</strong>mly bounded we clearly have∫ t2∣ div v ɛ (s, Xɛ−(t2−s) (x))ds∣ ≤ C|t 1 − t 2 |.t 1The o<strong>the</strong>r term can be treated exactly as be<strong>for</strong>e, when we proved equicontinuity in space.Recall that <strong>the</strong> functionu ɛ (t, x) = u ɛ 0(Xɛ −t (x)) a ɛ (t, x)satisfies <strong>the</strong> ɛ-problem (2.27). We also have <strong>the</strong> following convergences:Define <strong>the</strong> functionu ɛ 0 → u 0 in L p (R d ), (2.37)Xɛ −tk(x) → X −t (x) unif. on compact subset of [0, T ∗ ] × R d , (2.38)a ɛ k(t, x) → a(t, x) unif. on compact subset of [0, T ∗ ] × R d , (2.39)v ɛ k(t, x) → v(t, x) unif. on compact subset of [0, T ∗ ] × R d . (2.40)u(t, x) := u 0 (X −t (x)) a(t, x). (2.41)Convergence (2.37)-(2.40) toge<strong>the</strong>r with Lemma 13 and 14 allow us to prove <strong>the</strong> following proposition.Proposition 16. : u, u ɛ , uv and u ɛ v ɛ all belong to <strong>the</strong> space C([0, T ∗ ), L p (R d )). Moreover we have:Proof. Straight <strong>for</strong>ward, see appendix at <strong>the</strong> end of <strong>the</strong> paper.We now turn to <strong>the</strong> proof of <strong>the</strong> main <strong>the</strong>orem of this section.u ɛ k→ u in C([0, T ∗ ), L p ), (2.42)u ɛ kv ɛ k→ uv in C([0, T ∗ ), L p ). (2.43)Proof of Theorem 6. Let φ ∈ C ∞ c(0, T ∗ ) be a scalar test function. It is obvious that u ɛ and v ɛ satisfy:∫ T∗∫ T∗− u ɛ (t) φ ′ (t) dt + div ( u ɛ (t) v ɛ (t) ) φ(t) dt = 0, (2.44)00v ɛ (t, x) = (u ɛ (t) ∗ ∇K ɛ )(x) <strong>for</strong> all (t, x) ∈ [0, T ∗ ] × R d , (2.45)u ɛ (0) = u ɛ 0, (2.46)where <strong>the</strong> integrals in (2.44) are <strong>the</strong> integral of a continuous function from [0, T ∗ ] to <strong>the</strong> Banach spaceW −1,p (R d ). Recall that <strong>the</strong> injection i : L p (R d ) → W −1,p (R d ) and <strong>the</strong> differentiation ∂ xi : L p (R d ) →W −1,p (R d ) are bounded linear operator. There<strong>for</strong>e (2.42) and (2.43) implyu ɛ k→ u in C([0, T ∗ ], W −1,p (R d )),div [u ɛ kv ɛ k] → div [uv] in C([0, T ∗ ], W −1,p (R d )), (2.47)hich is more than enough to pass to <strong>the</strong> limit in relation (2.44). To pass to <strong>the</strong> limit in (2.45), it is enoughto note that <strong>for</strong> all (t, x) ∈ [0, T ∗ ] × R d we have|(u ɛ (t) ∗ ∇K ɛ ) − (u(t) ∗ ∇K)(x)| ≤ ‖u ɛ (t) − u(t)‖ L p‖∇K ɛ ‖ L q + ‖u(t)‖ L p‖∇K ɛ − ∇K‖ L q, (2.48)10


and finally it is trivial to pass to <strong>the</strong> limit in relation (2.46).<strong>Equation</strong> (2.44) means that <strong>the</strong> continuous function u(t) (continuous function with values in W −1,p (R d ))satisfies (2.12) in <strong>the</strong> distributional sense. But (2.12) implies that <strong>the</strong> distributional derivative u ′ (t) is itselfa continuous function with value in W −1,p (R d ). There<strong>for</strong>e u(t) is differentiable in <strong>the</strong> classical sense, i.e, itbelongs to C 1 ([0, T ∗ ], W −1,p (R d )), and (2.12) is satisfied in <strong>the</strong> classical sense.We now turn to <strong>the</strong> proof of (2.15). The u ɛ ’s satisfies (2.19). Integrating over [0, t], t < T ∗ , we get∫ t ∫‖u ɛ (t)‖ p L = p ‖uɛ 0‖ p L − (p − 1) u ɛ (s, x) p div v ɛ (s, x) dxdt. (2.49)p0 R dProposition 16 toge<strong>the</strong>r with <strong>the</strong> general inequality()‖|f| p − |g| p ‖ L 1 ≤ 2p ‖f‖ p−1L + p ‖g‖p−1 L‖f − g‖ p L p (2.50)implies that(u ɛ ) p , u p ∈ C([0, T ∗ ], L 1 (R d )), (2.51)(u ɛ k) p → u p ∈ C([0, T ∗ ], L 1 (R d )). (2.52)On <strong>the</strong> o<strong>the</strong>r hand, replacing ∇K by ∆K in (2.48) we see right away thatCombining (2.51)-(2.54) we obtainSo we can pass to <strong>the</strong> limit in (2.49) to obtaindiv v, div v ɛ ∈ C([0, T ∗ ], L ∞ (R d )), (2.53)div v ɛ k→ div v in C([0, T ∗ ], L ∞ (R d )). (2.54)u p div v, (u ɛ ) p div v ɛ ∈ C([0, T ∗ ], L 1 (R d )), (2.55)(u ɛ k) p div v ɛ k→ u p div v in C([0, T ∗ ], L 1 (R d )). (2.56)∫ t ∫‖u(t)‖ p L = ‖u 0‖ p p L − (p − 1) u(s, x) p div v(s, x) dxdt.p0 R dBut (2.55) implies that <strong>the</strong> function t → ∫ u(t, x) p div v(t, x) dx is continuous, <strong>the</strong>re<strong>for</strong>e <strong>the</strong> functionR dt → ‖u(t)‖ p <strong>Lp</strong> is differentiable and satisfies (2.15).2.5 Continuation and conserved propertiesTheorem 17 (Continuation of solutions). The solution provided by Theorem 6 can be continued up to atime T max ∈ (0, +∞]. If T max < +∞, <strong>the</strong>n lim t→Tmax sup τ∈[0,t] ‖u(τ)‖ L p = +∞Proof. The proof is standard. One just needs to use <strong>the</strong> continuity of <strong>the</strong> solution with respect to time.Theorem 18 (Conservation of mass/ second moment). (i) Under <strong>the</strong> assumption of Theorem 6, and if weassume moreover that u 0 ∈ L 1 (R d ), <strong>the</strong>n <strong>the</strong> solution u belongs to C([0, T ∗ ], L 1 (R d )) and satisfies ‖u(t)‖ L 1 =‖u 0 ‖ L 1 <strong>for</strong> all t ∈ [0, T ∗ ].(ii) Under <strong>the</strong> assumption of Theorem 6, and if we assume moreover that u 0 has bounded second moment,<strong>the</strong>n <strong>the</strong> second moment of u(t) stays bounded <strong>for</strong> all t ∈ [0, T ∗ ].Proof. We just need to revisit <strong>the</strong> proof of Proposition 16. Since u 0 ∈ L 1 ∩ L p it is clear thatu ɛ 0 = J ɛ u 0 → u 0 in L 1 (R d ) ∩ L p (R d ). (2.57)11


Using convergences (2.57), (2.38), (2.39) and (2.40) we prove thatu ɛ k→ u in C([0, T ∗ ], L 1 ∩ L p ). (2.58)The proof is exactly <strong>the</strong> same than <strong>the</strong> one of Proposition 16 . Since <strong>the</strong> aggregation equation is a conservationlaw, it is obvious that <strong>the</strong> smooth approximates satisfy ‖u ɛ (t)‖ L 1 = ‖u ɛ 0‖ L 1. Using (2.58) we obtain‖u(t)‖ L 1 = ‖u 0 ‖ L 1.We now turn to <strong>the</strong> proof of (ii). Since <strong>the</strong> smooth approximates u ɛ have compact support, <strong>the</strong>ir secondmoment is clearly finite, and <strong>the</strong> following manipulation are justified:∫∫(∫) 1/2 (∫) 1/2d|x| 2 u ɛ (t, x)dx = 2 ⃗x · v ɛ du ɛ (x) ≤ 2 |x| 2 u ɛ (t, x)dx |v ɛ | 2 u ɛ (t, x)dxdt R d R d R d R(∫) d 1/2≤ C |x| 2 u ɛ (t, x)dx . (2.59)R dAssume now that <strong>the</strong> second moment of u 0 is bounded. A simple computation shows that if η ɛ is radiallysymmetric, <strong>the</strong>n |x| 2 ∗ η ɛ = |x| 2 + second moment of η ɛ . There<strong>for</strong>e∫∫|x| 2 u ɛ 0(x)dx ≤ |x| 2 ∗ η ɛ (x) u 0 (x)dxR d R∫d ∫≤ |x| 2 u 0 (x)dx + |x| 2 η ɛ (x)dxR d R∫d≤ |x| 2 u 0 (x)dx + 1 <strong>for</strong> ɛ small enough. (2.60)R dInequality (2.60) come from <strong>the</strong> fact that <strong>the</strong> second moment of η ɛ goes to 0 as ɛ goes to 0. Estimate (2.59)toge<strong>the</strong>r with (2.60) provide us with a uni<strong>for</strong>m bound of <strong>the</strong> second moment of <strong>the</strong> u ɛ (t) which only depends<strong>the</strong> second moment of u 0 . Since u ɛ converges to u in L 1 , we obviously have, <strong>for</strong> a given R and t:∫∫|x|≤R|x| 2 u(t, x)dx = limɛ→0∫|x|≤R|x| 2 u ɛ (t, x)dx ≤ lim supɛ→0R d |x| 2 u ɛ (t, x)dx.Since R is arbitrary, this show that <strong>the</strong> second moment of u(t, ·) is bounded <strong>for</strong> all t <strong>for</strong> which <strong>the</strong> solutionexists.Combining Theorem 17 and 18 toge<strong>the</strong>r with equality (2.15) we get:Theorem 19 (Global existence when ∆K is bounded from above). Under <strong>the</strong> assumption of Theorem 6,and if we assume moreover that u 0 ∈ L 1 (R d ) and ess sup ∆K < +∞, <strong>the</strong>n <strong>the</strong> solution u exists <strong>for</strong> all times(i.e.: T max = +∞).Proof. Equality (2.15) can be written∫ddt {‖u(t)‖p L p} = (p − 1)R d u(t, x) p (u(t) ∗ ∆K)(x) dx. (2.61)Since ∆K is bounded from above we have(u(s) ∗ ∆K)(x) ≤ (ess sup ∆K)∫u(s, x)dx = (ess sup ∆K) ‖u 0 ‖ L 1 .R d (2.62)Combining (2.61), (2.62) and Gronwall inequality gives‖u(t)‖ p L p ≤ ‖u 0‖ p L p e(p−1)(ess sup ∆K)‖u0‖ L 1 t ,so <strong>the</strong> L p -norm can not blow-up in finite time which, because of Theorem 17, implies global existence.12


3 Uniqueness of solutions in P 2 (R d ) ∩ L p (R d )In this section we use an optimal transport argument to prove uniqueness of solutions in P 2 (R d ) ∩ L p (R d ),when ∇K ∈ W 1,q <strong>for</strong> 1 < q < ∞ and p its Hölder conjugate. To do that, we shall follow <strong>the</strong> steps in [19],where <strong>the</strong> authors extend <strong>the</strong> work of Loeper [34] to prove, among o<strong>the</strong>r, uniqueness of P 2 ∩ L ∞ -solutionsof <strong>the</strong> aggregation equation when <strong>the</strong> interaction potential K has a Lipschitz singularity at <strong>the</strong> origin.One can easily check that solutions of <strong>the</strong> aggregation equation constructed in previous sections aredistribution solutions, i.e. <strong>the</strong>y satisfy∫ T ∫ ( )∫∂ϕ0 R ∂t (t, x) + v(t, x) · ∇ϕ(t, x) u(t, x) dx dt = ϕ(0, x)u 0 (x) dx (3.63)N R N<strong>for</strong> all ϕ ∈ C0 ∞ ([0, T ∗ ) × R N )). A function u(t, x) satisfying (3.63) is said to be a distribution solution to <strong>the</strong>continuity equation (2.12) with <strong>the</strong> given velocity field v(t, x) and initial data u 0 (x). In fact, it is uniquelycharacterized by∫∫u(t, x) dx = u 0 (x) dxBX −t (B)<strong>for</strong> all measurable set B ⊂ R d , see [1]. Here X t : R d → R d is <strong>the</strong> flow map associated with <strong>the</strong> velocityfield v(t, x) and X −t is its inverse. In <strong>the</strong> optimal transport terminology this is equivalent to say that X ttransports <strong>the</strong> measure u 0 onto u(t) (u(t) = X t #u 0 ).We recall, <strong>for</strong> <strong>the</strong> sake of completeness, [19, Theorem 2.4], where several results of [2, 35, 22, 1] are puttoge<strong>the</strong>r.Theorem 20 ([19]). Let ρ 1 and ρ 2 be two probability measures on R N , such that <strong>the</strong>y are absolutely continuouswith respect to <strong>the</strong> Lebesgue measure and W 2 (ρ 1 , ρ 2 ) < ∞, and let ρ θ be an interpolation measurebetween ρ 1 and ρ 2 , defined as in [34] byρ θ = ((θ − 1)T + (2 − θ)I R N ) # ρ 1 (3.64)<strong>for</strong> θ ∈ [1, 2], where T is <strong>the</strong> optimal transport map between ρ 1 and ρ 2 due to Brenier’s <strong>the</strong>orem [12] and I R Nis <strong>the</strong> identity map. Then <strong>the</strong>re exists a vector field ν θ ∈ L 2 (R N , ρ θ dx) such thatdi.dθ ρ θ + div(ρ θ ν θ ) = 0 <strong>for</strong> all θ ∈ [1, 2].∫ii. ρ θ |ν θ | 2 dx = W2 2 (ρ 1 , ρ 2 ) <strong>for</strong> all θ ∈ [1, 2].R Niii. We have <strong>the</strong> L p -interpolation estimate‖ρ θ ‖ L p (R N ) ≤ max { }‖ρ 1 ‖ L p (R N ), ‖ρ 2 ‖ L p (R N )<strong>for</strong> all θ ∈ [1, 2].Here, W 2 (f, g) is <strong>the</strong> Euclidean Wasserstein distance between two probability measures f, g ∈ P(R n ),{∫∫} 1/2W 2 (f, g) = inf|v − x| 2 dΠ(v, x) , (3.65)Π∈Γ R n ×R nwhere Π runs over <strong>the</strong> set of joint probability measures on R n × R n with marginals f and g.Now we are ready to prove <strong>the</strong> uniqueness of solutions to <strong>the</strong> aggregation equation.Theorem 21 (Uniqueness). Let u 1 , u 2 be two bounded solutions of equation (2.12) in <strong>the</strong> interval [0, T ∗ ]with initial data u 0 ∈ P 2 (R d ) ∩ L p (R d ), 1 < p < ∞ and assume that v is given by v = −∇K ∗ u, with Ksuch that ∇K ∈ W 1,q (R N ), p and q conjugates. Then u 1 (t) = u 2 (t) <strong>for</strong> all 0 ≤ t ≤ T ∗ .13


Here ∂ 0 K is <strong>the</strong> unique element of minimal norm in <strong>the</strong> subdifferential of K. Simply speaking, sinceK(x) = |x| is smooth away from <strong>the</strong> origin and radially symmetric, we have ∂ 0 K(x) = x|x|<strong>for</strong> x ≠ 0 and∂ 0 K(0) = 0, and thus:∫(∂ 0 x − yK ∗ µ)(x) =dµ(y). (4.77)|x − y|Note that, µ t being a measure, it is important <strong>for</strong> ∂ 0 K to be defined <strong>for</strong> every x ∈ R d so that (4.75) makessense. <strong>Equation</strong> (4.74) means that∫ +∞0∫y≠xR d ( dψdt (x, t) + ∇ψ(x, t) · v t(x))dµ t (x) dt = 0, (4.78)<strong>for</strong> all ψ ∈ C ∞ 0 (R d × (0, +∞)). From (4.77) it is clear that |v t (x)| ≤ 1 <strong>for</strong> all x and t, <strong>the</strong>re<strong>for</strong>e <strong>the</strong> aboveintegral makes sense.The main Theorem of this section is <strong>the</strong> following:Theorem 24 (Instantaneous mass concentration). Consider <strong>the</strong> initial datau 0 (x) ={L|x| d−1+ɛ if |x| < 1,0 o<strong>the</strong>rwise,(4.79)where ɛ ∈ (0, 1) and L :=( ∫|x| 0 we haveµ t ({0}) > 0,i.e., mass is concentrated at <strong>the</strong> origin instantaneously and <strong>the</strong> solution is no longer continuous with respectto <strong>the</strong> Lebesgue measure.Theorem 24 is a consequence of <strong>the</strong> following estimate on <strong>the</strong> velocity field:Proposition 25. Let (µ t ) t∈(0,+∞) be <strong>the</strong> unique measure solution of <strong>the</strong> aggregation equation with interactionpotential K(x) = |x| and with initial data (4.79). Then, <strong>for</strong> all t ∈ [0, +∞) <strong>the</strong> velocity field v t = −∂ 0 K ∗ µ tis focussing and <strong>the</strong>re exists a constant C > 0 such that|v t (x)| ≥ C|x| 1−ɛ <strong>for</strong> all t ∈ [0, +∞) and x ∈ B(0, 1). (4.80)By focussing, we mean that <strong>the</strong> velocity field points inward, i.e.λ t : [0, +∞) → [0, +∞) such that v t (x) = −λ t (|x|) ⃗x x .<strong>the</strong>re exists a nonnegative function4.1 Representation <strong>for</strong>mula <strong>for</strong> radially symmetric measure solutionsIn this section, we show that <strong>for</strong> radially symmetric measure solutions, <strong>the</strong> characteristics are well defined.As a consequence, <strong>the</strong> solution to (2.12) can be expressed as <strong>the</strong> push <strong>for</strong>ward of <strong>the</strong> initial data by <strong>the</strong> flowmap associated with <strong>the</strong> ODE defining <strong>the</strong> characteristics.In <strong>the</strong> following <strong>the</strong> unit sphere {x ∈ R d , |x| = 1} is denoted by S d and its surface area by ω d .Definition 26. If µ ∈ P(R d ) is a radially symmetric probability measure, <strong>the</strong>n we define ˆµ ∈ P([0, +∞)) by<strong>for</strong> all I ∈ B([0, +∞)).ˆµ(I) = µ({x ∈ R d : |x| ∈ I})16


Remark 27. If a measure µ is radially symmetric, <strong>the</strong>n µ({x}) = 0 <strong>for</strong> all x ≠ 0, and <strong>the</strong>re<strong>for</strong>e∫∇K(x − y) dµ(y) =R d \{x}∫∇K(x − y) dµ(y)R d <strong>for</strong> all x ≠ 0.In o<strong>the</strong>r words, <strong>for</strong> x ≠ 0, (∇K ∗ µ)(x) is well defined despite <strong>the</strong> fact that ∇K is not defined at x = 0. Asa consequence (∂ 0 K ∗ µ)(x) = (∇K ∗ µ)(x) if x ≠ 0 and (∂ 0 K ∗ µ)(0) = 0.Remark 28. If <strong>the</strong> radially symmetric measure µ is continuous with respect to <strong>the</strong> Lebesgue measure andhas radially symmetric density u(x) = ũ(|x|), <strong>the</strong>n ˆµ is also continuous with respect to <strong>the</strong> Lebesgue measureand has density û, whereû(r) = ω d r d−1 ũ(r). (4.81)Lemma 29 (Polar coordinate <strong>for</strong>mula <strong>for</strong> <strong>the</strong> convolution). Suppose µ ∈ P(R d ) is radially symmetric. LetK(x) = |x|, <strong>the</strong>n <strong>for</strong> all x ≠ 0 we have:(∫ +∞( ) ) |x|x(µ ∗ ∇K) (x) = φ dˆµ(ρ)(4.82)ρ|x|0where <strong>the</strong> function φ : [0, +∞) → [−1, 1] is defined byφ(r) = 1 ∫re 1 − yω d |re 1 − y| · e 1dσ(y). (4.83)Proof. This come from simple algebraic manipulations. These manipulations are shown in [5].In <strong>the</strong> next Lemma we state properties of <strong>the</strong> function φ defined in (4.83).Lemma 30 (Properties of <strong>the</strong> function φ).S d(i) φ is continuous and non-decreasing on [0, +∞). Moreover φ(0) = 0, and lim r→∞ φ(r) = 1.(ii) φ(r) is O(r) as r → 0. To be more precise:Proof. Consider <strong>the</strong> function F : [0, +∞) × S d → [−1, 1] defined byφ(r)lim = 1 − 1 (y · e 1 )r→0 r ω dr>0∫S 2 dσ(y). (4.84)d(r, y) ↦→ re 1 − y|re 1 − y| · e 1. (4.85)Since F is bounded, we have thatφ(r) = 1 ∫F (r, y)dσ(y) = 1 F (r, y)dσ(y).ω d S dω d∫S d \{e 1}If y ∈ S d \{e 1 } <strong>the</strong>n <strong>the</strong> function r ↦→ F (r, y) is continuous on [0, +∞) and C ∞ on (0, +∞). An explicitcomputation shows <strong>the</strong>n that∂F 1 − F (r, y)2(r, y) = ≥ 0, (4.86)∂r |re 1 − y|thus φ is non-decreasing and, by <strong>the</strong> Lebesgue dominated convergence, it is easy to see that φ is continuous,φ(0) = 0 and lim r→∞ φ(r) = 1, which prove (i).To prove (ii), note that <strong>the</strong> function ∂F∂r(r, y) can be extended by continuity on [0, +∞). There<strong>for</strong>e <strong>the</strong>right derivative with respect to r of F (r, y) is well defined:F (r, y) − F (0, y) 1 − F (0, y)2lim= = 1 − (y · e 1 ) 2 .r→0 r|y|r>017


and since ∂F/∂r is bounded on (0, +∞) × S d \{e 1 }, we can now use <strong>the</strong> Lebesgue dominated convergence<strong>the</strong>orem to conclude:φ(r)limr→0 rr>0φ(r) − φ(0)= limr→0 rr>0= limr→0r>0∫1ω dF (r, y) − F (0, y)S dr(Remark 31. Note that <strong>for</strong> r > 0 <strong>the</strong> function ρ ↦→ φrρdσ(y) = 1 ω d∫S d1 − (y · e 1 ) 2 dσ(y).)is non increasing and continuous. Indeed, it isequal to 1 when ρ = 0 and it decreases to 0 as ρ → ∞. In particular, <strong>the</strong> integral in (4.82) is well defined<strong>for</strong> any probability measure ˆµ ∈ P([0, +∞)).Remark 32. In dimension two, it is easy to check that lim r→1 φ ′ (r) = +∞ which implies that <strong>the</strong> derivativeof <strong>the</strong> function φ has a singularity at r = 1 and thus, that <strong>the</strong> function φ is not C 1 .Proposition 33 (Characteristic ODE). Let K(x) = |x| and let (µ t ) t∈[0,+∞) be a weakly continuous familyof radially symmetric probability measures. Then <strong>the</strong> velocity field{−(∇K ∗ µ t )(x) if x ≠ 0v(x, t) =(4.87)0 if x = 0is continuous on R d \{0} × [0 + ∞). Moreover, <strong>for</strong> every x ∈ R d , <strong>the</strong>re exists an absolutely continuousfunction t → X t (x), t ∈ [0, +∞), which satisfiesddt X t(x) = v(X t (x), t) <strong>for</strong> a.e. t ∈ (0, +∞), (4.88)X 0 (x) = x. (4.89)Proof. From <strong>for</strong>mula (4.82), Remark 31, and <strong>the</strong> weak continuity of <strong>the</strong> family (µ t ) t∈[0,+∞) , we obtaincontinuity in time. The continuity in space simply comes from <strong>the</strong> continuity and boundedness of <strong>the</strong>function φ toge<strong>the</strong>r with <strong>the</strong> Lebesgue dominated convergence <strong>the</strong>orem.Since v is continuous on R d \{0} × [0 + ∞) we know from <strong>the</strong> Peano <strong>the</strong>orem that given x ∈ R d \{0}, <strong>the</strong>initial value problem (4.88)-(4.89) has a C 1 solution at least <strong>for</strong> short time. We want to see that it is defined<strong>for</strong> all time. For that, note that by a continuation argument, <strong>the</strong> interval given by Peano <strong>the</strong>orem can beextended as long as <strong>the</strong> solution stays in R d \{0}. Then, if we denote by T x <strong>the</strong> maximum time so that <strong>the</strong>solution exists in [0, T x ) we have that ei<strong>the</strong>r T x = ∞ and we are done, or T x < +∞, in which case clearlylim t→Tx X t (x) = 0, and we can extend <strong>the</strong> function X t (x) on [0, +∞) by setting X t (x) := 0 <strong>for</strong> t ≥ T x .The function t → X t (x) that we have just constructed is continuous on [0, +∞), C 1 on [0, +∞)\{T x }and satisfies (4.88) on [0, +∞)\{T x }. If x = 0, we obviously let X t (x) = 0 <strong>for</strong> all t ≥ 0.Finally, we present <strong>the</strong> representation <strong>for</strong>mula, by which we express <strong>the</strong> solution to (2.12) as a push<strong>for</strong>wardof <strong>the</strong> initial data. See [1] or [41] <strong>for</strong> a definition of <strong>the</strong> push-<strong>for</strong>ward of a measure by a map.Proposition 34 (Representation <strong>for</strong>mula). Let (µ t ) t∈[0,+∞) be a radially symmetric measure solution of <strong>the</strong>aggregation equation with interaction potential K(x) = |x|, and let X t : R d → R d be defined by (4.87), (4.88)and (4.89). Then <strong>for</strong> all t ≥ 0,µ t = X t #µ 0 .Proof. In this proof, we follow arguments from [1]. Since <strong>for</strong> a given x <strong>the</strong> function t ↦→ X t (x) is continuous,one can easily prove, using <strong>the</strong> Lebesgue dominated convergence <strong>the</strong>orem, that t ↦→ X t #µ 0 is weaklycontinuous. Let us now prove that µ t := X t #µ 0 satisfies (4.78) <strong>for</strong> all ψ ∈ C ∞ 0 (R d × (0, ∞)). Giventhat <strong>the</strong> test function ψ is compactly supported, <strong>the</strong>re exist T > 0 such that ψ(x, t) = 0 <strong>for</strong> all t ≥ T . We<strong>the</strong>re<strong>for</strong>e have:∫∫()0 = ψ(x, T )dµ T (x) − ψ(x, 0)dµ 0 (x) = ψ(X T (x), T ) − ψ(x, 0) dµ 0 (x). (4.90)R d R∫R d d18


If we now take into account that from Proposition 33 <strong>the</strong> mapping t → φ(X t (x), t) is absolutely continuous,we can rewrite (4.90) as0 ====∫R d ∫ T0∫ T ∫R d∫ T0∫ T00( ddt ψ(X t(x), t))dt dµ 0 (x) (4.91)(∇ψ(X t (x), t) · v(X t (x), t) + dψ∫R d (∇ψ(X t (x), t) · v(X t (x), t) + dψdt (X t(x), t)dt (X t(x), t)(∇ψ(x, t) · v(x, t) +∫R dψ )dt (x, t) dµ t (x) dt.d)dt dµ 0 (x) (4.92))dµ 0 (x) dt (4.93)The step from (4.92) to (4.93) holds because of <strong>the</strong> fact that |v(x, t)| ≤ 1, which justifies <strong>the</strong> use of <strong>the</strong>Fubini Theorem.Remark 35 (Representation <strong>for</strong>mula in polar coordinates). Let µ t and X t be as in <strong>the</strong> previous proposition.Let R t : [0, +∞) → [0, +∞) be <strong>the</strong> function such that |X t (x)| = R t (|x|). Thenˆµ t = R t #ˆµ 0 . (4.94)Remark 36. Since φ is nonnegative (Lemma 30), from (4.82), (4.87) and (4.88) we see that <strong>the</strong> functiont ↦→ |X t (x)| = R t (|x|) is non increasing.4.2 Proof of Proposition 25 and Theorem 24We are now ready to prove <strong>the</strong> estimate on <strong>the</strong> velocity field and <strong>the</strong> instantaneous concentration result. Westart by giving a frozen in time estimate of <strong>the</strong> velocity field.Lemma 37. Let K(x) = |x|, and let u 0 (x) be defined by (4.79) <strong>for</strong> some ɛ ∈ (0, 1). Then <strong>the</strong>re exist aconstant C > 0 such that|(u 0 ∗ ∇K) (x)| ≥ C|x| 1−ɛ (4.95)<strong>for</strong> all x ∈ B(0, 1)\{0}.Proof. Note that if we do <strong>the</strong> change of variable s = |x|ρ|(u 0 ∗ ∇K) (x)| = |x|in equation (4.82), we find that∫ +∞0φ(s) û 0 ( |x|s )ds s 2 . (4.96)On <strong>the</strong> o<strong>the</strong>r hand, using (4.81) we see that <strong>the</strong> û 0 (r) corresponding to <strong>the</strong> u 0 (x) defined by (4.79) isû 0 (r) ={ωdr ɛ if r < 1,0 o<strong>the</strong>rwise.(4.97)Then, plugging (4.97) in (4.96) we obtain that <strong>for</strong> all x ≠ 0|(u 0 ∗ ∇K) (x)| = ω d |x| 1−ɛ ∫ +∞|x|φ(ρ)dρ. (4.98)ρ2−ɛ In light of statement (ii) of Lemma 30, we see that <strong>the</strong> previous integral converges as |x| → 0.|(u 0 ∗ ∇K) (x)| is O(|x| 1−ɛ ) as |x| → 0 and (4.95) follows.HenceFinally, <strong>the</strong> last piece we need in order to prove Proposition 25 from <strong>the</strong> previous lemma, is <strong>the</strong> followingcomparison principle:19


Lemma 38 (Temporal monotonicity of <strong>the</strong> velocity). Let (µ t ) t∈(0,+∞) be a radially symmetric measuresolution of <strong>the</strong> aggregation equation with interaction potential K(x) = |x| . Then, <strong>for</strong> every x ∈ R d \{0} <strong>the</strong>functiont ↦→ |(∇K ∗ µ t )(x)|is non decreasing.Proof. Combining (4.82) and (4.94) we see that|(µ t ∗ ∇K)(x)| =(∫ +∞0)φ( |x|R t (ρ) ) dˆµ 0(ρ) . (4.99)Now, by Lemma 30, φ is non decreasing and due to Remark 35, t ↦→ R t (ρ) is non increasing. Hence<strong>for</strong>th itis clear that (4.99) is itself non decreasing.At this point, Proposition 25 follows as a simple consequence of <strong>the</strong> frozen in time estimate (4.95) toge<strong>the</strong>rwith Lemma 38, and we can give an easy proof <strong>for</strong> <strong>the</strong> main result we introduced at <strong>the</strong> beginning of <strong>the</strong>section.Proof of Theorem 24. Using <strong>the</strong> representation <strong>for</strong>mula (Proposition 34) and <strong>the</strong> definition of <strong>the</strong> push <strong>for</strong>wardwe getµ t ({0}) = (X t #µ 0 )({0}) = µ 0 (Xt −1 ({0})).Then, note that <strong>the</strong> solution of <strong>the</strong> ODE ṙ = −r 1−ɛ reaches zero in finite time. There<strong>for</strong>e, from Proposition25 and 33 we obtain that <strong>for</strong> all t > 0, <strong>the</strong>re exists δ > 0 such thatX t (x) = 0 <strong>for</strong> all |x| < δ.In o<strong>the</strong>r words, <strong>for</strong> all t > 0, <strong>the</strong>re exists δ > 0 such that B(0, δ) ⊂ Xt−1 ({0}). Clearly, given our choice ofinitial condition, we have that µ 0 (B(0, δ)) > 0 if δ > 0, and <strong>the</strong>re<strong>for</strong>e µ t ({0}) > 0 if t > 0.4.3 Remark about <strong>the</strong> initial data 1/ |x| d−1An interesting open problem is we<strong>the</strong>r or not <strong>the</strong>re is instantaneous mass concentration when <strong>the</strong> initial datais defined by (4.79) with ɛ = 0. We right now can not answer this question, but below are some interestingremarks about this case.Lemma 39. Let u 0 be defined by (4.79) with ɛ = 0. Then <strong>the</strong>re exists constants C > 0 and β > 1 such that|(∇K ∗ u 0 )(x)| ≤ C|x||x|∣log β ∣ <strong>for</strong> all x ∈ B(0, 1). (4.100)Proof. From Lemma 30 it is clear that <strong>the</strong>re exists a constant α > 0 such that φ(r)/r ≤ α <strong>for</strong> all r ∈ (0, 1).We can <strong>the</strong>n use (4.98) with ɛ = 0 to obtain that, <strong>for</strong> |x| < 1,( ∫ 1 ∫ )φ(ρ) dρ ∞|(u 0 ∗ ∇K)(x)| = ω d |x|ρ ρ + φ(ρ)1 ρ 2 dρ≤ ω d |x|(α|x|∫ 1|x|dρρ + cst )= ω d |x| (−α log |x| + cst) = −C|x| log |x|β ,<strong>for</strong> some constants C > 0 and β > 1.Consider <strong>the</strong> push<strong>for</strong>ward S t (r) defined by <strong>the</strong> ODEddt S t(r) = −CS t (r)∣ log S t(r)β ∣ ,S 0 (r) = r.20


This ODE has an explicit solution: <strong>for</strong> r < 1 we have( ) eCtrS t (r) = ββThe push <strong>for</strong>ward of u 0 by <strong>the</strong> map S t can also be explicitly computed:(S t #u 0 )(r) ={ L(t)r α(t)if r < β 1−exp(Ct)0 o<strong>the</strong>rwisewhere α(t) = (d − 1) + ∫ eCt − 1β1−exp(Ct)e Ct , and L(t) =r −α(t) dr.0In dimension 2 and greater, α(t) ≤ d − 1 <strong>for</strong> t > 0, so S t #u 0 is less singular than u 0 .5 Osgood condition <strong>for</strong> global well-posednessThis section considers <strong>the</strong> global well-posedness of <strong>the</strong> aggregation equation in P 2 (R d ) ∩ L p (R d ), dependingon <strong>the</strong> potential K. We start by giving a precise definitions of “natural potential”, “repulsive in <strong>the</strong> shortrange” and “strictly attractive in <strong>the</strong> short range”, and <strong>the</strong>n we prove Theorem 5.Definition 40. A natural potential is a radially symmetric potential K(x) = k(|x|), where k : (0, +∞) → Ris a smooth function which satisfies <strong>the</strong> following conditions:(C1)sup |k ′ (r)| < +∞,r∈(0,∞)(C2) ∃ α > d such that k ′ (r) and k ′′ (r) are O(1/r α ) as r → +∞,(MN1) ∃ δ 1 > 0 such that k ′′ (r) is monotonic (ei<strong>the</strong>r increasing or decreasing) in (0, δ 1 ),(MN2) ∃ δ 2 > 0 such that rk ′′ (r) is monotonic (ei<strong>the</strong>r increasing or decreasing) in (0, δ 2 ).Remark 41. Note that monotonicity condition (MN1) implies that k ′ (r) and k(r) are also monotonic insome (different) neighborhood of <strong>the</strong> origin (0, δ). Also, note that (C1) and (MN1) imply(C3)limr→0 k′ (r) exists and is finite.+Remark 42. The far field condition (C2) can be dropped when <strong>the</strong> data has compact support.Definition 43. A natural potential is said to be repulsive in <strong>the</strong> short range if <strong>the</strong>re exists an interval (0, δ)on which k(r) is decreasing. A natural potential is said to be strictly attractive in <strong>the</strong> short range if <strong>the</strong>reexists an interval (0, δ) on which k(r) is strictly increasing.We would like to remark that <strong>the</strong> two monotonicity conditions are not very restrictive as, in order toviolate <strong>the</strong>m, a potential would have to exhibit some pathological behavior around <strong>the</strong> origin, like oscillatingfaster and faster as r → 0.5.1 Properties of natural potentialsAs a last step be<strong>for</strong>e proving Theorem 5, let us point out some properties of natural potentials, which show<strong>the</strong> reason behind <strong>the</strong> choice of this kind of potentials to work with.Lemma 44. If K(x) = k(|x|) is a natural potential, <strong>the</strong>n k ′′ (r) = o(1/r) as r → 0.21


Proof. First, note that since k(r) is smooth away from 0 we havek ′ (1) − k ′ (ɛ) =∫ 1ɛk ′′ (r)dr.Now, because of (MN1) we know that <strong>the</strong>re exists a neighborhood of zero in which k ′′ doesn’t change sign.There<strong>for</strong>e, letting ɛ → 0 and using (C3) we conclude that k ′′ is integrable around <strong>the</strong> origin. A simpleintegration by part, toge<strong>the</strong>r with (C3) gives <strong>the</strong>n that∫ r0k ′′ (s)s ds = −Dividing both sides by r and letting r → 0 we obtain1limr→0 r∫ r0∫ rwhich, combined with (MN2) implies lim r→0 k ′′ (r)r = 0.0k ′ (s)dr + k ′ (r)r.k ′′ (s)s ds = 0.The next lemma shows that <strong>the</strong> existence <strong>the</strong>ory developed in <strong>the</strong> previous section applies to this classof potentials.Lemma 45. If K is a natural potential <strong>the</strong>n ∇K ∈ W 1,q (R d ) <strong>for</strong> all 1 ≤ q < d. As a consequence, <strong>the</strong>critical exponents p s and q s associated to a natural potential satisfyProof. Recall that∇K(x) = k ′ (|x|) x|x|andq s ≥ d and p s ≤ dd − 1 .∂ 2 ()K(x) = k ′′ (|x|) − k′ (|x|) xi x j∂x i ∂x j |x| |x| 2 + δ ijk ′ (|x|),|x|where δ ij is <strong>the</strong> Kronecker delta symbol. In order to prove <strong>the</strong> lemma, it is enough to show thatk ′ (|x|), k ′′ (|x|) and k′ (|x|)|x|belong to L q (R d ) <strong>for</strong> all 1 ≤ q < d. To do that, observe that <strong>the</strong> decay condition(C2) implies that <strong>the</strong>y belong to L q (B(0, 1) c ) <strong>for</strong> all q ≥ 1. Then, we take into account that (C3)implies that k′ (r)r= O(1/r) as r → 0 and that we have seen in <strong>the</strong> previous Lemma that k ′′ (r) = o(1/r) asr → 0. This is enough to conclude, since <strong>the</strong> function x ↦→ 1/|x| is in L q (B(0, 1)) <strong>for</strong> all 1 ≤ q < d.The following Lemma toge<strong>the</strong>r with Theorem 19 gives global existence of solutions <strong>for</strong> natural potentialswhich are repulsive in <strong>the</strong> short range, whence part (i) of Theorem 5 follows.Lemma 46. Suppose K is a natural potential which is repulsive in <strong>the</strong> short range. Then ∆K is boundedfrom above.Proof. We will prove that <strong>the</strong>re is a neighborhood of zero on which ∆K ≤ 0. This combined with <strong>the</strong> decaycondition (C2) give <strong>the</strong> desired result. First, recall that ∆K(x) = k ′′ (|x|) + (d − 1)k ′ (|x|)|x| −1 . Then, sincek is repulsive in <strong>the</strong> short range, <strong>the</strong>re exists a neighborhood of zero in which k ′ ≤ 0. Now, we have twopossibilities: on one hand, if lim r→0 + k ′ (r) = 0, <strong>the</strong>n given r ∈ (0, +∞) <strong>the</strong>re exists s ∈ (0, r) such thatk ′ (r)r= k ′′ (s). Toge<strong>the</strong>r with (MN1), this implies that k ′′ is also non-positive in some neighborhood of zero.On <strong>the</strong> o<strong>the</strong>r hand if lim r→0 + k ′ (r) < 0, <strong>the</strong>n <strong>the</strong> fact that k ′′ (r) = o(1/r) implies that rk′′ (r)+(d−1)k ′ (r)risnegative <strong>for</strong> r small enough.Finally, <strong>the</strong> next Lemma will be needed to prove global existence <strong>for</strong> natural potentials which are strictlyattractive in <strong>the</strong> short range and satisfy <strong>the</strong> Osgood criteria.22


Lemma 47. Suppose that K is a natural potential which is strictly attractive in <strong>the</strong> short range and satisfies<strong>the</strong> Osgood criteria (1.9). If moreover sup x≠0 ∆K(x) = +∞ <strong>the</strong>n <strong>the</strong> following holds(Z1) lim r→0 + k ′′ (r) = +∞ and lim r→0 +k′ (r)r= +∞,(Z2) ∃δ 1 > 0 such that k ′′ (r) and k′ (r)rare decreasing <strong>for</strong> r ∈ (0, δ 1 ),(Z3) ∃δ 2 > 0 such that k ′′ (r) ≤ k′ (r)r<strong>for</strong> r ∈ (0, δ 2 ).Proof. Let us start by proving by contradiction thatIf we suppose thatlimsupr→0 + p∈(0,r)k ′ (r)rk ′ (r)r= +∞. (5.101)< C ∀r ∈ (0, 1], (5.102)<strong>the</strong>n given a sequence r n → 0 + <strong>the</strong>re will exist ano<strong>the</strong>r sequence s n → 0 + , 0 < s n < r n , such thatk ′ (r n )r n= k ′′ (s n ) < C.Since k ′′ (r) is monotonic around zero, this implies that k ′′ is bounded from above, and combining thiswith (5.102) we see that ∆K must also be bounded from above, which contradicts our assumption. Now,statements (Z1), (Z2), (Z3) follow easily: First note that if lim r→0 + k ′ (r) > 0, <strong>the</strong>n clearly <strong>the</strong> Osgoodcondition (1.9) is not satisfied, whence lim r→0 + k ′ (r) = 0. This implies that <strong>for</strong> all r > 0 <strong>the</strong>re existss ∈ (0, r) such thatk ′ (r)r= k ′′ (s). (5.103)Combining (5.103), (5.101) and <strong>the</strong> monotonicity of k ′′ we get that lim r→0 + k ′′ (r) = +∞ and k ′′ is decreasingon some interval (0, δ), which corresponds with <strong>the</strong> first part of (Z1) and (Z2). Now, going back to (5.103)we see that if 0 < s < r < δ <strong>the</strong>n k′ (r)rddr= k ′′ (s) ≥ k ′′ (r) which proves (Z3). This implies{ k ′ }(r)= 1 r r()k ′′ (r) − k′ (r)≤ 0rand <strong>the</strong>re<strong>for</strong>e k ′ (r)/r decreases on (0, δ). Thus, (5.101) implies lim r→0 + k′ (r)rcomplete.= +∞ and <strong>the</strong> proof is5.2 Global bound of <strong>the</strong> L p -norm using Osgood criteriaWe have already proven global existence when <strong>the</strong> potential is bounded from above. In this section we prove<strong>the</strong> following proposition, which allows us to prove global existence <strong>for</strong> potentials which are attractive in <strong>the</strong>short range, satisfy <strong>the</strong> Osgood criteria, and whose Laplacian is not bounded from above. From it, secondpart of Theorem 5 follows readily. This extends prior work on L ∞ -solutions [5] to <strong>the</strong> L p case.Proposition 48. Suppose that K is a natural potential which is strictly attractive in <strong>the</strong> short range, satisfies<strong>the</strong> Osgood criteria (1.9) and whose Laplacian is not bounded from above (i.e. sup x≠0 ∆K(x) = +∞). Letu(t) be <strong>the</strong> unique solution of <strong>the</strong> aggregation equation starting with initial data u 0 ∈ P 2 (R d ) ∩ L p (R d ), wherep > d/(d − 1). Define <strong>the</strong> length scale( ) q/d ‖u(t)‖L 1R(t) =.‖u(t)‖ L p23


Then <strong>the</strong>re exists positive constants δ and C such that <strong>the</strong> inequalityholds <strong>for</strong> all time t <strong>for</strong> which R(t) ≤ δ.dRdt ≥ −C k′ (R) (5.104)Remark 49. R(t) is <strong>the</strong> natural length scale associated with blow up of L p norm. Given that mass isconserved, R(t) > 0 means that <strong>the</strong> L p norm is bounded. The differential inequality (5.104) tell us that R(t)decays slower than <strong>the</strong> solutions of <strong>the</strong> ODE ẏ = −Ck ′ (y). Since k ′ (r) satisfies <strong>the</strong> Osgood criteria (1.9),solutions of this ODE do not go to zero in finite time. R(t) <strong>the</strong>re<strong>for</strong>e stays away from zero <strong>for</strong> all time whichprovides us with a global upper bound <strong>for</strong> ‖u(t)‖ L p.Proof. Equality (2.15) can be writtenUsing <strong>the</strong> chain rule we getddt∫ddt {‖u(t)‖p L p} = (p − 1){ }‖u(t)‖ − q dL= − p(p − 1)qpd≥ −R d u(t, x) p (u(t) ∗ ∆K)(x) dx.‖u(t)‖ − q d −pL p ∫R d u(t, x) p (u(t) ∗ ∆K)(x) dx (5.105)(p − 1)q‖u(t)‖ − q d<strong>Lp</strong>dsup {(u(t) ∗ ∆K)(x)} . (5.106)px∈R dTo obtain (5.104), we now need to carefully estimate sup x∈R d {(u(t) ∗ ∆K)(x)}:Lemma 50 (potential <strong>the</strong>ory estimate). Suppose that K(x) = k(|x|) satisfies (C2), (Z1), (Z2) and (Z3).Suppose also that p > . Then <strong>the</strong>re exists positive constants δ and C such that inequalitydd−1k ′ (R)sup (u ∗ ∆K) (x) ≤ C‖u‖ L 1x∈R Rd( ) q/d ‖u‖L 1where R =(5.107)‖u‖ L pholds <strong>for</strong> all nonnegative u ∈ L 1 (R d ) ∩ L p (R d ) satisfying R ≤ δ.Remark: By Lemma 47, potentials satisfying <strong>the</strong> conditions of Proposition 48 automatically satisfy (C2),(Z1), (Z2) and (Z3).Proof of Lemma 50. Recall that ∆K(x) = k ′′ (|x|) + (d − 1) k′ (|x|)|x|so that (Z3) implies that ∆K(x)


To go from (5.108) to (5.109) we have used <strong>the</strong> fact that k ′ is nonnegative in a neighborhood of zero (becauselim r→0 + k ′ (r)/r = +∞ ) and that u is also nonnegative. To go from (5.111) to (5.112) we have used <strong>the</strong>fact that k ′ is increasing in a neighborhood of zero (because lim r→0 + k ′′ (r) = +∞). Finally, to go from(5.112) to (5.113) we have used <strong>the</strong> fact that, since q < d, ∫ ɛ∫0 rd−1−q dr = ɛ d−q /(d − q). Let us now estimate|y|≥ɛ u(x − y)∆K(y)dy. On one hand k′′ (r) and k ′ (r)/r go to 0 as r → +∞. On <strong>the</strong> o<strong>the</strong>r hand k ′′ (r) andk ′ (r)/r go monotonically to +∞ as r → 0 + . There<strong>for</strong>e <strong>for</strong> ɛ small enough we have{}sup ∆K(x) = sup k ′′ (r) + (d − 1) k′ (r)|x|≥ɛr≥ɛrwhich gives, since u is nonnegative,∫|y|≥ɛu(x − y)∆K(y)dy ≤ d ‖u‖ L 1≤ k ′′ (ɛ) + (d − 1) k′ (ɛ)ɛCombining (5.113) and (5.114) we see that <strong>for</strong> ɛ small enough we have≤ d k′ (ɛ)ɛk ′ (ɛ). (5.114)ɛsup (u ∗ ∆K) (x) ≤ c k′ (ɛ)(ɛ d/q ‖u‖ L p + ‖u‖ L 1) (5.115)x∈R ɛd(‖u‖L 1) q/d.where c is a positive constant depending on d and q. To conclude <strong>the</strong> proof, choose ɛ = R =‖u‖ L pCombining (5.106) and (5.107) allows us to get (5.104), which concludes <strong>the</strong> proof of Proposition 48.6 ConclusionsThis paper develops refined analysis <strong>for</strong> well-posedness of <strong>the</strong> multidimensional aggregation equation <strong>for</strong>initial data in L p . An additional assumption of bounded second moment is needed as a decay condition <strong>for</strong>uniqueness; <strong>for</strong>tunately this condition is preserved by <strong>the</strong> dynamics of <strong>the</strong> equation. The results connectrecent <strong>the</strong>ory developed <strong>for</strong> L ∞ initial data [4, 5, 6] to recent <strong>the</strong>ory <strong>for</strong> measure solutions [16]. It turnsout that L p spaces provide a good understanding of <strong>the</strong> transition from a regular (bounded) solution to ameasure solution, which was proved to occur in [5] whenever <strong>the</strong> potential violates <strong>the</strong> Osgood condition.In [5], it is shown that <strong>for</strong> <strong>the</strong> special case of K(x) = |x|, no ‘first kind’ similarity solutions exist to describe<strong>the</strong> blowup to a mass concentration <strong>for</strong> odd space dimensions larger than one. A subsequent numerical studyof blowup [26], <strong>for</strong> K = |x|, illustrates that <strong>for</strong> dimensions larger than two, <strong>the</strong>re is a ‘second kind’ self-similarblowup in which no mass concentration occurs at <strong>the</strong> initial blowup time. Ra<strong>the</strong>r, <strong>the</strong> solution is in L p <strong>for</strong>some p < p s = d/(d − 1). The solution has asymptotic structure like <strong>the</strong> example constructed in Section 4of this paper, thus we expect after <strong>the</strong> initial blowup time, that it concentrates mass in a delta. We remarkthat <strong>the</strong>se results provide an interesting connection to classical results <strong>for</strong> Burgers equation. Our equation inone space dimension, with K = |x|, reduces to a <strong>for</strong>m of Burgers equation by defining w(x) = ∫ xu(y)dy [10].0Thus, an initial blowup <strong>for</strong> <strong>the</strong> aggregation problem is <strong>the</strong> same as a a singularity in <strong>the</strong> slope <strong>for</strong> Burgersequation. Generically, Burgers singularities <strong>for</strong>m by creating a |x| 1/3 power singularity in w, which gives a−2/3 power blowup in u. This corresponds to a blowup in L p <strong>for</strong> p > 3/2, but does not result in an initialmass concentration. However, as we well know, <strong>the</strong> Burgers solution <strong>for</strong>ms a shock immediately afterwards,resulting in a delta concentration in u. Thus <strong>the</strong> scenario described above is a multidimensional analogueof <strong>the</strong> well-known behavior of how singularities initially <strong>for</strong>m in Burgers equation. The delta-concentrationsare analogues of shock <strong>for</strong>mation in scalar conservation laws.Section 4 constructs an example of a measure solution with initial data in L p , p < p s , that instantaneouslyconcentrates mass, <strong>for</strong> <strong>the</strong> special kernel K(x) = |x| (near <strong>the</strong> origin). We conjecture that such solutions exist<strong>for</strong> more general power-law kernels K(x) = |x| α , 2 − d < α < 2. The proof of instantaneous concentrationuses some monotonicity properties of <strong>the</strong> convolution operator ⃗x/|x| which would need to be proved <strong>for</strong> <strong>the</strong>more general case.25


Several interesting open problems remain, in addition to proving sharpness of <strong>the</strong> exponent p s <strong>for</strong> moregeneral kernels. For initial data in L ps local well-posedness is not known and we do not have intuition <strong>for</strong>this.7 Appendix – Proof of Proposition 16Since all <strong>the</strong> convergences (2.38)-(2.40) take place on compact sets, one of <strong>the</strong> key ideas of this proof will beto approximate u 0 ∈ L p by a function with compact support and to use <strong>the</strong> fact that X t maps compact setsto compact sets (Lemma 14).PART I: We will prove that u(t, x) = u 0 (X −t (x))a(t, x) belongs to C([0, T ∗ ), L p (R d )). Assume first thatu 0 ∈ C c (R d ). The function u 0 is <strong>the</strong>n uni<strong>for</strong>mly continuous and, sinceit is clear that <strong>the</strong> quantitysup |X −t (x) − X −s (x)| ≤ C|t − s|,x∈R d‖u 0 (X −t (x)) − u 0 (X s (x))‖ L pcan be made as small as we want by choosing |t−s| small enough. We have <strong>the</strong>re<strong>for</strong>e proven that u 0 (X −t (x)) ∈C([0, T ∗ ), L p (R d ). Assume now that u 0 ∈ L p (R d ). Approximate it by a function g ∈ C c (R d ) and write:‖u 0 (X −t (x)) − u 0 (X −s (x))‖ L p ≤ ‖u 0 (X −t (x)) − g(X −t (x))‖ L p + ‖g(X −t (x)) − g(X −s (x))‖ L p= I + II + III.+‖g(X −s (x)) − u 0 (X −s (x))‖ L pAs we have seen above, <strong>the</strong> second term can be made as small as we want by choosing |t − s| small enough.Using Lemma 13 we getI, III ≤ C ∗ ‖u 0 − g‖ L p,which can be made as small as we want since C c (R d ) is dense in L p (R d ). We have <strong>the</strong>re<strong>for</strong>e proven that, ifu 0 ∈ L p (R d ), <strong>the</strong>n(t, x) ↦→ u 0 (X −t (x)) ∈ C([0, T ∗ ), L p (R d )). (7.116)Let us now consider <strong>the</strong> function u 0 (X −t (x))a(t, x).[0, T ∗ ] × R d . Write‖u 0 (X −t (x))a(t, x) − u 0 (X −s (x))a(s, x)‖ L p ≤‖u 0 (X −t (x)){a(t, x) − a(s, x)}‖ L pRecall that a(t, x) is continuous and bounded on+‖{u 0 (X −t (x)) − u 0 (X −s (x))}a(s, x)‖ L p= I + II.Since a(s, x) is bounded, (7.116) implies that II can be made as small as we want by choosing |t − s| smallenough. Let us now take care of I. Approximate u 0 by a function g ∈ C c (R d ) and writeI≤ ‖{u 0 (X −t (x)) − g(X −t (x))}{a(t, x) − a(s, x)}‖ L p+‖g(X −t (x)){a(t, x) − a(s, x)}‖ L p= A + B.From Lemma 13 we haveA ≤ 2 C ∗ ‖u 0 − g‖ L psup |a(t, x)|,(t,x)∈[0,T ∗ ]×R d26


and since C c (R d ) is dense in L p , A can be made as small as we want. Let Ω denote <strong>the</strong> compact support ofg. Using Lemmas 13 and 14 we obtain:(∫B =(≤Ω+Ct∣∣g(X −t (x)){a(t, x) − a(s, x)} ∣ ) 1/pp dxsup |a(t, x) − a(s, x)|x∈Ω+ct)C ∗ ‖g‖ L p.Since a(t, x) is uni<strong>for</strong>mly continuous on compact subset of [0, T ∗ ] × R d , it is clear that by choosing |t − s|small enough we can make B as small as we want. We have proven thatu ∈ C([0, T ∗ ], L p (R d ).PART II: In Part I we have proven that u ∈ C([0, T ∗ ], L p ). The same proof work to show that u ɛ , uvand u ɛ v ɛ belong to C([0, T ∗ ], L p ) (v(t, x) is continuous and bounded, so we can handle it exactly like a(t, x).)PART III: Let us now prove that u ɛ (t, x) = u ɛ 0(Xɛ−t (x))a ɛ (t, x) converges to u(t, x) = u 0 (X −t (x))a(t, x).For convenience we write ɛ instead of ɛ k . To do this, we will successively prove:u 0 (Xɛ −t (x)) → u 0 (X −t (x)) in C([0, T ∗ ), L p ), (7.117)u ɛ 0(Xɛ −t (x)) → u 0 (X −t (x)) in C([0, T ∗ ), L p ), (7.118)u ɛ 0(Xɛ −t (x))a ɛ (t, x) → u 0 (X −t (x))a(t, x) in C([0, T ∗ ), L p ). (7.119)To prove (7.117), approximate u 0 ∈ L p (R d ) by a function g ∈ C c (R d ) and write:sup t∈[0,T ∗ ) ‖u 0 (Xɛ −t (x) − u 0 (X −t (x))‖ L p ≤ sup t∈[0,T ∗ ) ‖u 0 (Xɛ−t (x)) − g(Xɛ −t (x))‖ L p +sup t∈[0,T ∗ ) ‖g(Xɛ−t (x)) − g(X −t (x))‖ L p + sup t∈[0,T ∗ ) ‖g(X −t (x)) − u 0 (X −t (x))‖ L p= I + II + III.From Lemma 13 it is clear that I and III can be made as small as we want by choosing an appropriatefunction g. Using Lemma 14, we see that, if Ω is <strong>the</strong> support of gII =sup ∣ g(X−tɛ (x)) − g(Xt∈[0,T(∫Ω+Ct−t (x)) ∣ 1/p dx) p .∗Using <strong>the</strong> uni<strong>for</strong>m continuity of g toge<strong>the</strong>r with <strong>the</strong> fact that Xɛ−t (x) converges uni<strong>for</strong>mly to X −t (x) on[0, T ∗ ] × Ω + CT ∗ , we can make II as small as we want by choosing ɛ small enough.This concludes <strong>the</strong> proof of (7.117). To prove (7.118), writesup ‖u ɛ 0(Xɛ −t (x)) − u 0 (X −t (x))‖ L p ≤ sup ‖u ɛ 0(Xɛ−t (x)) − u 0 (Xɛ−t (x)‖ L pt∈[0,T ∗ )t∈[0,T ∗ )= I + II.+ sup ‖u 0 (Xɛ−t (x)) − u 0 (X −t (x))‖ L pt∈[0,T ∗ )From (7.117) we know that II can be made as small as we want by choosing ɛ small enough. Using Lemma13 we obtainI ≤ C ∗ ‖u ɛ 0 − u 0 ‖ L p → 0 as ɛ → 0.27


Let us now prove (7.119). Writesup ‖u ɛ 0(Xɛ−t (x))a ɛ (t, x) − u 0 (X −t (x))a(t, x)‖ L pt∈[0,T ∗ ]≤supt∈[0,T ∗ ]‖{u ɛ 0(Xɛ−t (x)) − u 0 (X −t (x))}a ɛ (t, x)‖ L p+ sup ‖u 0 (X −t (y)){a ɛ (t, x) − a(t, x)}‖ L pt∈[0,T ∗ ]= I + II.Since a ɛ (t, x) is uni<strong>for</strong>mly bounded on [0, T ∗ ] × R d , it is clear from (7.118) that I → 0 as ɛ → 0. If u 0 is inC c (R d ), <strong>the</strong>n it is easy to prove that II → 0 as ɛ → 0. If u 0 ∈ L p (R d ), <strong>the</strong>n approximate it by g ∈ C c (R d )and proceed as be<strong>for</strong>e.Part IV: To prove that u ɛ kv ɛ kconverges to uv in C([0, T ∗ ], L p ), proceed exactly as in <strong>the</strong> proof of(7.119).AcknowledgmentsWe thank José Antonio Carrillo and Dejan Slepčev <strong>for</strong> useful comments. The authors were supportedby NSF grant DMS-0907931. JR acknowledge support from <strong>the</strong> project MTM2008-06349-C03-03 DGI-MCI (Spain) and 2009-SGR-345 from AGAUR-Generalitat de Catalunya. We thank <strong>the</strong> Centre de RecercaMatemàtica (Barcelona) <strong>for</strong> providing an excellent atmosphere <strong>for</strong> research.References[1] L.A. Ambrosio, N. Gigli, G. Savaré, Gradient flows in metric spaces and in <strong>the</strong> space of probabilitymeasures, Lectures in Ma<strong>the</strong>matics, Birkhäuser, 2005.[2] J.D. Benamou, Y. Brenier, A computational Fluid Mechanics solution to <strong>the</strong> Monge-Kantorovichmass transfer problem, Numer. Math., 84 (2000), pp. 375–393.[3] D. Benedetto, E. Caglioti, M. Pulvirenti, A kinetic equation <strong>for</strong> granular media, RAIRO Modél.Math. Anal. Numér., 31, (1997), 615–641.[4] A. Bertozzi, J. Brandman, Finite-time blow-up of L ∞ -weak solutions of an aggregation equation, toappear in Comm. Math. Sci., special issue in honor of Andrew Majda’s 60th birthday.[5] A. Bertozzi, J.A. Carrillo, T. Laurent, Blowup in multidimensional aggregation equations withmildly singular interaction kernels, Nonlinearity, 22, (2009), pp. 683-710.[6] A. Bertozzi, T. Laurent, Finite-time blow-up of solutions of an aggregation equation in R n , Comm.Math. Phys., 274 (2007), pp. 717–735.[7] P. Biler and A. Woyczyński Global and Expoding Solutions <strong>for</strong> Nonlocal Quadratic Evolution Problems,SIAM J. Appl. Math., 59 (1998), pp. 845-869.[8] A. Blanchet, J. A. Carrillo, and N. Masmoudi, Infinite Time <strong>Aggregation</strong> <strong>for</strong> <strong>the</strong> Critical Patlak-Keller-Segel model in R 2 , to appear in Comm. Pure Appl. Math.[9] A. Blanchet, J. Dolbeault, and B. Perthame, Two-dimensional Keller-Segel model: optimalcritical mass and qualitative properties of <strong>the</strong> solutions, Electron. J. Differential <strong>Equation</strong>s, (2006), No.44, 32 pp. (electronic).[10] M. Bodnar and J.J.L. Velázquez, An integro-differential equation arising as a limit of individualcell-based models, J. Differential <strong>Equation</strong>s 222, (2006), pp. 341–380.28


[11] S. Boi, V. Capasso and D. Morale, Modeling <strong>the</strong> aggregative behavior of ants of <strong>the</strong> species Polyergusrufescens, Spatial heterogeneity in ecological models (Alcalá de Henares, 1998), Nonlinear Anal. RealWorld Appl., 1 (2000), pp. 163–176.[12] Y. Brenier, Polar factorization and monotone rearrangement of vector-valued functions, Comm. PureAppl. Math. 44, (1991), pp. 375–417.[13] M.P. Brenner, P. Constantin, L.P. Kadanoff, A. Schenkel and S.C. Venkataramani,Diffusion, attraction and collapse, Nonlinearity 12, (1999), pp. 1071–1098.[14] M. Burger, V. Capasso and D. Morale, On an aggregation model with long and short rangeinteractions, Nonlinear Analysis. Real World Applications. An International Multidisciplinary Journal,8 (2007), pp. 939–958.[15] M. Burger and M. Di Francesco, Large time behavior of nonlocal aggregation models with nonlineardiffusion, Networks and Heterogenous Media, 3(4), (2008), pp. 749-785.[16] J.A. Carrillo, M. Difrancesco, A. Figalli, T. Laurent and D. Slepčev, Global-in-time weakmeasure solutions, finite-time aggregation and confinement <strong>for</strong> nonlocal interaction equations, preprint.[17] J.A. Carrillo, R.J. McCann, and C. Villani, Kinetic equilibration rates <strong>for</strong> granular media andrelated equations: entropy dissipation and mass transportation estimates, Rev. Matemática Iberoamericana,19 (2003), pp. 1-48.[18] J.A. Carrillo, R.J. McCann, C. Villani, Contractions in <strong>the</strong> 2-Wasserstein length space and<strong>the</strong>rmalization of granular media, Arch. Rat. Mech. Anal., 179 (2006), pp. 217–263.[19] J.A. Carrillo, J. Rosado Uniqueness of bounded solutions to <strong>Aggregation</strong> equations by optimaltransport methods, preprint.[20] J. Dolbeault and B. Perthame, Optimal critical mass in <strong>the</strong> two-dimensional Keller-Segel modelin R 2 , C. R. Math. Acad. Sci. Paris, 339 (2004), pp. 611–616.[21] Qiang Du, Ping Zhang, Existence of weak solutions to some vortex density models, Siam J. Math.Anal., Vol. 34, No. 6, pp. 1279–1299.[22] W. Gangbo, R.J. McCann, The geometry of optimal transportation, Acta Math., 177 (1996),pp. 113161.[23] V. Gazi and K. Passino, Stability analysis of swarms, IEEE Trans. Auto. Control, 48 (2003), pp. 692-697.[24] D. Holm and V. Putkaradze, Formation of clumps and patches in self-aggregation of finite-sizeparticles , Physica D 220, (2006), pp. 183–196.[25] D. Holm and V. Putkaradze, <strong>Aggregation</strong> of finite size particles with variable mobility, Phys. Rev.Lett., (2005) 95:226106.[26] Y. Huang and A. L. Bertozzi, Self-similar blow-up solutions to an aggregation equation, preprint2009.[27] E.F. Keller and L.A. Segel, Initiation of slide mold aggregation viewed as an instability, J. Theor.Biol., 26 (1970).[28] T. Laurent, Local and Global Existence <strong>for</strong> an <strong>Aggregation</strong> <strong>Equation</strong>, Communications in PartialDifferential <strong>Equation</strong>s, 32 (2007), pp. 1941-1964.29


[29] D. Li and J. Rodrigo, Finite-time singularities of an aggregation equation in R n with fractionaldissipation, Comm. Math. Phys., 287(2), (2009), pp. 687-703.[30] D. Li and J. Rodrigo, Refined blowup criteria and nonsymmetric blowup of an aggregation equation,Advances in Ma<strong>the</strong>matics, 220(1), (2009), pp. 1717-1738.[31] H. Li and G. Toscani, Long-time asymptotics of kinetic models of granular flows, Arch. Ration. Mech.Anal. 172, (2004), pp. 407–428.[32] D. Li and X. Zhang, On a nonlocal aggregation model with nonlinear diffusion, (2008) preprint.[33] F. Lin and P. Zhang, On <strong>the</strong> hydrodynamic limit of Ginzburg-Landau vortices, Discrete Contin.Dynam. Systems, 6 (2000), pp. 121–142.[34] G. Loeper, Uniqueness of <strong>the</strong> solution to <strong>the</strong> Vlasov-Poisson system with bounded density, J. Math.Pures Appl. 86 (2006), pp. 68–79.[35] R. J. McCann, A convexity principle <strong>for</strong> interacting gases, Adv. Math., 128 (1997), pp. 153–179.[36] A. Mogilner and L. Edelstein-Keshet, A non-local model <strong>for</strong> a swarm, J. Math. Bio. 38, (1999),pp. 534–570.[37] D. Morale, V. Capasso and K. Oelschläger, An interacting particle system modelling aggregationbehavior: from individuals to populations, J. Math. Biol., 50 (2005), pp. 49–66.[38] C.M. Topaz and A.L. Bertozzi, Swarming patterns in a two-dimensional kinematic model <strong>for</strong> biologicalgroups, SIAM J. Appl. Math., 65 (2004), pp. 152–174.[39] C.M. Topaz, A.L. Bertozzi, and M.A. Lewis, A nonlocal continuum model <strong>for</strong> biological aggregation,Bulletin of Ma<strong>the</strong>matical Biology, 68(7), pp. 1601-1623, 2006.[40] G. Toscani, One-dimensional kinetic models of granular flows, RAIRO Modél. Math. Anal. Numér.,34, 6 (2000), pp. 1277–1291.[41] C. Villani, Topics in optimal transportation, volume 58 of Graduate Studies in Ma<strong>the</strong>matics, AmericanMa<strong>the</strong>matical Society, Providence, RI, 2003.[42] C. Villani, Optimal transport, old and new, Lecture Notes <strong>for</strong> <strong>the</strong> 2005 Saint-Flour summer school, toappear, Springer 2008.[43] M. Vishik, Incompressible flows of an ideal fluid with vorticity in borderline spaces of Besov type, Ann.scient. Éc. Norm. Sup., 4e série, t. 32 (1999), pp. 769–812.[44] V. I. Yudovich, Non-stationary flow of an incompressible liquid., Zh. Vychisl. Mat. Mat. Fiz, 3 (1963),pp. 1032-1066.30

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!