ebook img

Critical mass for a Patlak-Keller-Segel model with degenerate diffusion in higher dimensions PDF

0.39 MB·English
Save to my drive
Quick download
Download
Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.

Preview Critical mass for a Patlak-Keller-Segel model with degenerate diffusion in higher dimensions

Critical mass for a Patlak-Keller-Segel model with degenerate diffusion in higher dimensions Adrien Blanchet∗, Jos´e A. Carrillo† & Philippe Laurenc¸ot ‡ 8 0 0 February 2, 2008 2 n a J 5 Abstract 1 This paper is devoted to the analysis of non-negative solutions for a generalisation of the ] classical parabolic-elliptic Patlak-Keller-Segel system with d ≥ 3 and porous medium-like P non-lineardiffusion. Here,thenon-lineardiffusionischoseninsuchawaythatitsscalingand A the one of the Poisson term coincide. We exhibit that the qualitative behaviour of solutions . h is decided by the initial mass of the system. Actually, there is a sharp critical mass Mc such t that if M ∈ (0,Mc] solutions exist globally in time, whereas there are blowing-up solutions a m otherwise. We also show the existence of self-similar solutions for M ∈ (0,Mc). While characterising the eventual infinite time blowing-up profile for M =Mc, we observe that the [ long time asymptotics are much more complicated than in the classical Patlak-Keller-Segel 1 system in dimension two. v 0 1 1 Introduction 3 2 . In this work, we analyse qualitative properties of non-negative solutions for the Patlak-Keller- 1 0 Segel system in dimension d ≥ 3 with homogeneous non-linear diffusion given by 8 0 ∂u : (t,x) = div[∇um(t,x)−u(t,x)∇φ(t,x)] t > 0, x∈ Rd , v ∂t i  X  −∆φ(t,x) = u(t,x) , t > 0, x∈ Rd , (1.1)   r  a u(0,x) = u (x) x∈ Rd. 0     Initial data will be assumed throughout this paper to verify u ∈ L1(Rd;(1+|x|2) dx) ∩ L∞(Rd), ∇um ∈ L2(Rd) and u ≥0 . (1.2) 0 0 0 A fundamental property of the solutions to (1.1) is the formal conservation of the total mass of the system M := u (x) dx= u(t,x) dx for t ≥ 0 . 0 Rd Rd Z Z ∗EPI SIMPAF – INRIA Futurs, Parc Scientifique de la haute Borne, F–59650 Villeneuve d’Ascq, France & Laboratoire Paul Painlev´e – Universit´e de Lille 1, F–59655 Villeneuve d’Ascq C´edex, France. E-mail: [email protected], Internet: http://www.ceremade.dauphine.fr/∼blanchet/ †ICREA (Instituci´o Catalana de Recerca i Estudis Avan¸cats) and Departament de Matem`atiques, Uni- versitat Aut`onoma de Barcelona, E–08193 Bellaterra, Spain. E-mail: [email protected], Internet: http://kinetic.mat.uab.es/∼carrillo/ ‡Institut de Math´ematiques de Toulouse, CNRS UMR 5219 & Universit´e de Toulouse, 118 route de Nar- bonne, F–31062 Toulouse C´edex 9, France. E-mail: [email protected], Internet: http://www.mip.ups- tlse.fr/∼laurenco/ 1 As the solution to the Poisson equation −∆φ= u is given up to an harmonic function, we choose the one given by φ= K∗u with 1 1 K(x) = c and c := d |x|d−2 d (d−2)σ d where σ := 2πd/2/Γ(d/2) is the surface area of the sphere Sd−1 in Rd. This system has been d proposed as a model for chemotaxis-driven cell movement or in the study of large ensemble of gravitationally interacting particles, see [17, 12, 4] and the literature therein. We will concentrate on a particular choice of the non-linear diffusion exponent m in any di- mensioncharacterisedforproducinganexactbalanceinthescalingofdiffusionandpotentialdrift inequation (1.1). Tothis endweusetheby-now classical scaling leadingto thenonlinearFokker- Planck equation for porous media as in [10], that is, let us define ρ by ρ(s,y) := edtu β(t),etx and c:= K∗ρ with β strictly increasing to be chosen. Then, it is straightforward to check that (cid:0) (cid:1) ∂ρ (s,y) = div yρ(s,y)+β′(t) e−λt∇ρm(s,y)−e−dtρ(s,y)∇c(s,y) s >0, y ∈ Rd , ∂s   −∆c(s,y) = ρ(s,(cid:2)y), (cid:8) (cid:9)(cid:3) s >0, y ∈ Rd ,    ρ(0,y) = u (y) ≥ 0 y ∈ Rd, 0     with λ = d(m−1)+2. From this scaling, theonly possiblechoice of m leading to a compensation effect between diffusion and concentration is given by λ = d or equivalently 2(d−1) m := . d d In that case, β′(t)= edt determines the change of variables and the final scaled equation reads: ∂ρ (s,y) = div[yρ(s,y)+∇ρmd(s,y)−ρ(s,y)∇c(s,y)] s > 0, y ∈ Rd , ∂s   −∆c(s,y) = ρ(s,y) , s > 0, y ∈ Rd , (1.3)    ρ(0,y) = u (y) ≥ 0 y ∈ Rd, 0     Notethat the case d = 2 and m = 1 corresponds to the Patlak-Keller-Segel system or to the 2 classical Smoluchowski-Poisson system in two dimensions with linear diffusion [29, 19]. In this case, a simple dichotomy result have been shown in [14, 4] improving over previous results in [18,26],namely,thebehaviourofthesolutionsisjustdeterminedbytheinitialmassofthesystem. More precisely, there exists a critical value of the mass M := 8π such that if 0 < M < M (sub- c c critical case) the solutions exist globally and if M > M (super-critical case) the solutions blow c up in finite time. Moreover, in the sub-critical case solutions behave self-similarly as t → ∞ [2, 4]. Finally, the critical case M = M was studied in [3] showing that solutions exist globally c and blow up as a Dirac mass at the centre of mass as t → ∞. Solutions have to be understood as free energy solutions, concept that we will specify below. In this work, we will show that a similar situation to the classical PKS system in d = 2, although with some important differences, happens for the critical variant of the PKS model in any dimension d ≥ 3 reading as: ∂u (t,x) = div[∇umd(t,x)−u(t,x)∇(K∗u)(t,x)] t > 0, x ∈ Rd , ∂t (1.4)   u(0,x) = u0(x) ≥ 0 x∈ Rd.   2 We will simply denote by m the critical exponent 2(d−1) m := m = ∈ (1,2), d d as long as d ≥ 3, in the rest of the paper for notational convenience. The main tool for the analysis of this equation is the following free energy functional: um(t,x) 1 t 7→ F[u(t)] := − K(x−y)u(t,x)u(t,y) dxdy ZRd m−1 2ZZRd×Rd um(t,x) c 1 d = − u(t,x)u(t,y) dxdy ZRd m−1 2 ZZRd×Rd |x−y|d−2 which is related to its time derivative, the Fisher information, in the following way: given a smooth positive fast-decaying solution to (1.4), then 2 d m F[u(t)] = − u(t,x) ∇ um−1(t,x)−φ(t,x) dx . (1.5) dt ZRd (cid:12) (cid:18)m−1 (cid:19)(cid:12) (cid:12) (cid:12) (cid:12) (cid:12) We will give a precise sense to this entrop(cid:12)y/entropy-dissipation relation b(cid:12)elow. The system (1.4) can formally be considered a particular instance of the general family of PDEs studied in [8, 1, 9]. The free energy functional F structurally belongs to the general class of free energies for interacting particles introduced in [25, 8, 9]. The functionals treated in those references are of the general form: 1 E[n]:= U[n(x)] dx+ n(x)V(x) dx+ W(x−y)n(x)n(y) dxdy ZRd ZRd 2 ZZRd×Rd under the basic assumptions U : R+ → R is a density of internal energy, V : Rd → R is a convex smooth confinement potential and W :Rd → R is a symmetric convex smooth interaction potential. The internal energy U should satisfy the following dilation condition, introduced in McCann [25] λ 7−→ λdU(λ−d) is convex non-increasing on R+. In our case, the interaction potential is singular and the key tool of displacement convexity of the functional fails, making the theory in the previous references not useful for our purposes. Nevertheless, the free energy functional plays a central role for this problem as we shall see below. Before proceeding further, let us state the notion of solutions we will deal with in the rest: Definition 1.1 (Weak and free energy solution) Let u be an initial condition satisfying (1.2) 0 and T ∈ (0,∞]. (i) A weak solution to (1.4) on [0,T) with initial condition u is a non-negative function 0 u ∈ C([0,T);L1(Rd)) such that u ∈ L∞((0,t) × Rd)), um ∈ L2(0,t;H1(Rd)) for each t ∈ [0,T) and T u (x)ψ(0,x) dx= [∇um(t,x)−u(t,x)∇φ(t,x)]·∇ψ(t,x) dxdt 0 ZRd Z0 ZRd T − u(t,x)∂ ψ(t,x) dxdt (1.6) t Z0 ZRd for any test function ψ ∈D([0,T)×Rd) with φ = K∗u. 3 (ii) A free energy solution to (1.4) on [0,T) with initial condition u is a weak solution to (1.4) 0 on [0,T) with initial condition u satisfying additionally: u(2m−1)/2 ∈ L2(0,t;H1(Rd)) and 0 t 2m 2 F[u(t)]+ ∇u(2m−1)/2(s,x)−u1/2(s,x)∇φ(s,x) dxds≤ F[u ] (1.7) 0 Z0ZRd(cid:12)(cid:18)2m−1 (cid:19)(cid:12) (cid:12) (cid:12) (cid:12) (cid:12) for all t ∈(0,T) w(cid:12)ith φ= K∗u. (cid:12) In (1.7), we cannot write the Fisher information factorised by u as in (1.5) because of the lack of regularity of u. We notethat both(1.6)and(1.7)aremeaningful. Indeed,theregularity required for u implies that the solution φ = K∗u to the Poisson equation satisfies φ ∈ L∞(0,t;H1(Rd)) for all t ∈ (0,T). In addition, it follows from (1.6) by classical approximation arguments that ku(t)k = u(t,x) dx= u (x) dx= ku k = M for t ∈[0,T) . (1.8) 1 0 0 1 Rd Rd Z Z Letuspointoutthattheexistenceoffreeenergysolutionsforarelatedproblemwasessentially obtained in [30, 31, 27] where the Poisson equation is replaced by −∆φ = u− φ. There, the authors also show that the mass is the suitable quantity for (1.4) allowing for a dichotomy. Precisely, the author shows that there exist two masses 0 < M < M such that if 0 < M < M 1 2 1 the solutions exist globally in time, while for M > M there are solutions blowing up in finite 2 time. The values of these masses, are related to the sharp constants of the Sobolev inequality. Here, we will make a fundamental use of a variant to the Hardy-Littlewood-Sobolev (VHLS) inequality, see Lemma 3.2: for all h ∈ L1(Rd)∩Lm(Rd), there exists an optimal constant C such ∗ that h(x)h(y) dxdy ≤ C khkmkhk2/d . (cid:12)ZZRd×Rd |x−y|d−2 (cid:12) ∗ m 1 (cid:12) (cid:12) This inequality will play(cid:12)the same role as the loga(cid:12)rithmic HLS inequality proved in [6] for the (cid:12) (cid:12) classical PKS system in d = 2 [14, 4, 3]. The VHLS inequality and the identification of the equality cases allow us to give the first main result of this work, namely, the following sharp critical mass d/2 2 M := c (m−1)C c (cid:20) ∗ d(cid:21) for equation (1.4). More precisely, we will show that free energy solutions exist globally for M ∈ (0,M ] while there are finite time blowing-up solutions otherwise. However, the long time c asymptotics of the solutions is much more complicated compared to the classical PKS system in two dimensions. The main results of this work and the open problems related to large times asymptotics can be summarised as follows: • Sub-critical case: 0 < M < M , solutions exist globally in time and there exists a radially c symmetriccompactly supportedself-similar solution, although we arenotabletoshow that it attracts all global solutions. See Proposition 4.3, Theorem 5.2 and Corollary 5.7. • Critical case: M = M , solutions exist globally in time, see Proposition 4.6. There are c infinitely many compactly supportedstationary solutions. The second moment of solutions is non-decreasing in time, with two possibilities we cannot exclude: either is uniformly bounded in time or diverges. Moreover, the Lm-norm of the solution could be divergent as t → ∞ or a diverging sequence of times could exist with bounded Lm-norm. However, we show a striking difference with respect to the classical PKS system in two dimensions [3], namely, the existence of global in time solutions not blowing-up in infinite time. We will comment further on these issues in Section 4.2.3. 4 • Super-critical case: M > M , we prove that there exist solutions, corresponding to initial c data with negative free energy, blowing up in finite time, see Proposition 4.2. However, we cannot exclude the possibility that solutions with positive free energy may be global in time. Theresultsareorganisedasfollows. Section2showsakeymaximaltimeofexistencecriterion forfreeenergysolutionsofequation (1.4). Thiscriterion improvesovertheresultsin[30,31]since it is only based on the boundedness or unboundedness in time of the Lm-norm of the solutions and it has to be compared to a similar criterion based on the logarithmic entropy in the classical PKS system in two dimensions obtained in [3]. Section 3 is devoted to the variational study of the minimisation of the free energy functional over the set of densities with a fixed mass. With that aim the proof of the VHLS inequality and the identification of the equality cases are performed. Section 4 uses this variational information to show the above main results concerning the dichotomy, the global existence for M < M and the characterisation by concentration- c compactness techniques of the nature of the possible blow-up in the critical case leading to the global existence for this critical value. Finally, the last section is devoted to the study of the free energy functional in self-similar variables and the proof of the existence of self-similar solutions in the sub-critical case. 2 Existence criterion As in [30, 31], we consider the regularised problem ∂u ε(t,x) = div[∇(f ◦u )(t,x)−u (t,x)∇φ (t,x)] t > 0, x∈ Rd , ε ε ε ε ∂t   φε(t,x) = K∗uε(t,x) , t > 0, x∈ Rd , (2.1)   u (0,x) = uε ≥ 0 x∈ Rd, ε 0    where f : [0,∞) −→ R is given by f (u) := (u+ε)m −εm. Here, uε is the convolution of u ε ε 0 0 with a sequence of mollifiers and kuεk = ku k = M in particular. This regularised problem has 0 1 0 1 global in time smooth solutions. This approximation has been proved to be convergent. More precisely, the result in [31, Section 4] asserts that if we assume that sup ku (t)k ≤κ (2.2) ε ∞ 0<t<T where κ is independent of ε > 0, then there exists a sub-sequence ε → 0, such that n u → u strongly in C([0,T],Lp (Rd)) and a.e. in (0,T)×Rd, (2.3) εn loc ∇um ⇀ ∇um weakly-* in L∞(0,T;L2(Rd)), (2.4) εn φ (t) → φ(t) strongly in Lr (Rd) a.e. in (0,T), (2.5) εn loc ∇φ (t) → ∇φ(t) strongly in Lr (Rd) a.e. in (0,T), (2.6) εn loc for any p ∈ (1,∞) and r ∈ (1,∞], and u is a weak solution to (1.4) on [0,T) with φ = K∗u. Moreover, the free energy [30, Proposition 6.1] satisfies F[u(t)] ≤ F[u ] for a.e. t ≥ 0. However, 0 a detailed analysis of the proof in [30, Proposition 6.1] shows that the weak solution is in fact a free energy solution. Proposition 2.1 (Existence of free energy solutions) Under assumption (1.2) on the ini- tial data and (2.2) on the approximation sequence, there exists a free energy solution to (1.4) in [0,T). 5 Proof. The only remaining points not covered by the results in [30, 31] are the lower semi- continuity of the free energy dissipation and the fact that u(2m−1)/2 belongs to L2(0,t;H1(Rd)) for t ∈ [0,T). The latter will actually be shown in the proof of Lemma 2.3, see (2.9) below. Concerning the former, a careful reading of the proof of [30, Proposition 6.1] gives that 3 t m 2 F [u (t)]+ [u (s,x)+ε] ∇ [u (s,x)+ε]m−1−φ (s,x) ψ (x) dxdt ≤ F [u ] ε,l ε ε ε ε l ε,l 0 4 Z0 ZRd (cid:12) (cid:18)m−1 (cid:19)(cid:12) (cid:12) (cid:12) (cid:12) (cid:12) for a.e. t ∈ (0,T) where ψ is a st(cid:12)andard cut-off function in Rd for any l(cid:12)∈ N and l [u (t,x)+ε]m 1 ε F [u (t)] = ψ (x) dx− K(x−y)u (t,x)u (t,y) dxdy . ε,l ε l ε ε ZRd m−1 2 ZZRd×Rd In this regularised setting, we can write that 2 2 m 2m (u +ε) ∇ (u +ε)m−1−φ = ∇(u +ε)(2m−1)/2 −(u +ε)1/2∇φ . ε ε ε ε ε ε m−1 2m−1 (cid:12) (cid:20) (cid:21)(cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) (cid:12) As proved i(cid:12)n [30], we have F [u (t)](cid:12)→ F(cid:12) [u(t)] as ε → 0 and l → ∞. In addition(cid:12), it is εn,l εn n straightforward from the convergence properties (2.3)-(2.6) above to pass to the limit as ε → 0 n in the free energy dissipation functional with the help of a lower semi-continuity argument. We leave the details to the reader, see e.g. [28] or [7, Lemma 10]. Hence, passing to the limit as l → ∞, then u is a free energy solution as it satisfies the free energy inequality (1.7). (cid:3) Remark 2.2 The free energy inequality (1.7) can be obtained with constant 3/4 multiplying the entropy dissipation directly from [30, Proposition 6.1] and the procedure above. This is a technical issue that can be improved to constant 1 by redoing the proof in [30,Proposition6.1] treating more carefully the free energy dissipation term. In fact, the proof in [30, Proposition 6.1] shows that you can choose the constant as close to 1 as you want. We are now ready to characterise the maximal time of existence by showing the local in time boundedness of the Lm-norm independently of the approximation parameter ε >0 and how this estimate implies the local in time L∞-estimate (2.2). Lemma 2.3 (From uniform integrability to L∞-bounds) For any η > 0 there exists τ > η 0 depending only on d, M, and η such that, if sup ku (t∗)k ≤ η ε m ε∈(0,1) for some t∗ ∈ [0,∞), then (i) the family (u ) is bounded in L∞(t∗,t∗+τ ;Lm(Rd)). ε ε η (ii) Moreover, if (u (t∗)) is also bounded in Lp(Rd) for some p ∈ (m,∞], then (u ) is bounded ε ε ε ε in L∞(t∗,t∗+τ ;Lp(Rd)). η Proof. To prove this result we need to refine the argument already used in the two-dimensional situation d = 2 with linear diffusion m = 1 in [4, 3]. We follow a procedure analogous to the ones in [20, 5, 30, 13]. 6 Step 1 - Lm-estimates: By (2.1) we have d ku km = −m(m−1) um−2∇u · m(u +ε)m−1∇u −u ∇φ dx dt ε m Rd ε ε ε ε ε ε Z 4m2(m−1) (cid:0)2 (cid:1) ≤ − ∇u(2m−1)/2 −(m−1) um∆φ dx (2m−1)2 ε 2 Rd ε ε (cid:13) (cid:13) Z 4m2(m−1) (cid:13) (cid:13)2 = − (cid:13)∇u(2m−1)/2(cid:13) +(m−1)ku km+1 . (2m−1)2 ε 2 ε m+1 (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) As 2m 2(m+1) 2d 1 < < < , 2m−1 2m−1 d−2 we have the following Gagliardo-Nirenberg-Sobolev inequality: there exists a positive constant C such that [(2m−1)d]/[(m+1)(2m+d−2)] 2m2/[(m+1)(2m+d−2)] kwk ≤ Ck∇wk kwk 2(m+1)/(2m−1) 2 2m/(2m−1) (2m−1)/2 which we apply with w = u to obtain ε ku k(2m−1)/2 ≤ C ∇u(2m−1)/2 [(2m−1)d]/[(m+1)(2m+d−2)]ku km2(2m−1)/[(m+1)(2m+d−2)] . ε m+1 ε ε m 2 (cid:13) (cid:13) (cid:13) (cid:13) It leads to (cid:13) (cid:13) ku km+1 ≤C ∇u(2m−1)/2 2d/(2m+d−2) ku k2m2/(2m+d−2) ε m+1 ε ε m 2 ≤ (2m(cid:13)(cid:13)(cid:13)2m−21)2 ∇u(cid:13)(cid:13)(cid:13)(ε2m−1)/2 22+Ckuεkmm2/(m−1) . (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) We thus end up with d ku km+ 2m2(m−1) ∇u(2m−1)/2 2 ≤ (m−1)Cku km2/(m−1) . (2.7) dt ε m (2m−1)2 ε 2 ε m (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) (cid:13) In particular, for any t ≥ t ≥ 0 2 1 −(m−1) ku (t )km ≤ ku (t )k−m/(m−1) −C(t −t ) (2.8) ε 2 m ε 1 m 2 1 h i Taking t = t∗ we deduce from (2.8) that 1 −(m−1) ku (t)km ≤ η−m/(m−1) −C(t−t∗) for t ∈ [t∗,t∗+2τ ) ε m η (cid:16) (cid:17) with τ = 1/ 2Cηm/(m−1) . Consequently, ku (t)km ≤ (Cτ )−(m−1) for t ∈ [t∗,t∗ +τ ] and the η ε m η η proof of the first assertion of Lemma 2.3 is complete. In addition, coming back to (2.7), we (cid:0) (cid:1) further deduce that t∗+τη 2 ∇u(2m−1)/2 ≤ C(t∗,η) . (2.9) ε Zt∗ (cid:13) (cid:13)2 (cid:13) (cid:13) (cid:13) (cid:13) 7 Step 2 - Lp-estimates, p ∈ (m,∞): For t ∈ [t∗,t∗+τ ], K ≥ 1, and p > m, we infer from (2.1) η that d k(u (t)−K) kp ≤ −mp(p−1) (u (t)−K)p−2(u +ε)m−1|∇u |2 dx dt ε + p Rd ε + ε ε Z p−1 p−2 +p(p−1) (u (t)−K) +K(u (t)−K) ∇u ·∇φ dx ε + ε + ε ε Rd Z h i ≤ −mp(p−1) (u (t)−K)m+p−3|∇u |2 dx ε + ε Rd Z p p−1 − (p−1)(u (t)−K) +pK(u (t)−K) ∆φ dx ε + ε + ε Rd Z h i 4mp(p−1) 2 (m+p−1)/2 ≤ − ∇ (u (t)−K) +(I) (m+p−1)2 ε + 2 (cid:13) h i(cid:13) (cid:13) (cid:13) with (cid:13) (cid:13) (I) := pK2k(u (t)−K) kp−1+(2p−1)Kk(u (t)−K) kp+(p−1)k(u (t)−K) kp+1 . ε + p−1 ε + p ε + p+1 We now use the following interpolation inequality 2 kwkp+1 ≤ C(p) ∇ w(m+p−1)/2 kwk2/d p+1 1 2 (cid:13) (cid:16) (cid:17)(cid:13) (cid:13) (cid:13) which is aconsequence of the Gagliardo-Nirenberg-Sobolev and H¨older inequalities (see, e.g., [31, (cid:13) (cid:13) Lemma 3.2]) to obtain (I) ≤ (p−1)K2k(u (t)−K) kp +K2 |{x : u (t,x) ≥ K}|+(2p−1)Kk(u (t)−K) kp ε + p ε ε + p 2 (m+p−1)/2 2/d + C(p) ∇ (u (t)−K) k(u (t)−K) k . ε + ε + 1 2 (cid:13) h i(cid:13) Noting that (cid:13) (cid:13) (cid:13) (cid:13) (m−1)/m ku (t)k ε 1 k(u (t)−K) k ≤ ku (t)k ε + 1 ε m K (cid:18) (cid:19) and recalling that ku (t)k = M we conclude that ε 1 2/d ku (t)k 2 ε m (m+p−1)/2 (I) ≤ C(p) ∇ (u (t)−K) K2(m−1)/md ε + 2 + K[2p−1+(p−1)(cid:13)K]hk(u (t)−K) kp+KiM(cid:13) . (cid:13) ε + p (cid:13) (cid:13) (cid:13) By Step 1, we may choose K = K large enough such that ∗ 2/d ku (t)k 4mp(p−1) ε m C ≤ K2(m−1)/md (m+p−1)2 ∗ for all t ∈ [t∗,t∗+τ ] and ε ∈ (0,1), hence η 4mp(p−1) 2 (I)≤ ∇ (u (t)−K )(m+p−1)/2 +C(p,t∗,η) 1+k(u (t)−K ) kp . (m+p−1)2 ε ∗ + 2 ε ∗ + p (cid:13) h i(cid:13) (cid:13) (cid:13) (cid:2) (cid:3) Therefore (cid:13) (cid:13) d k(u (t)−K ) kp ≤ C(p,t∗,η) 1+k(u (t)−K ) kp , dt ε ∗ + p ε ∗ + p so that (cid:2) (cid:3) k(u (t)−K ) kp ≤ C(p,t∗,η) for t ∈ [t∗,t∗+τ ] and ε ∈ (0,1) . ε ∗ + p η 8 As ku (t)kp ≤C(p) Kp−mku (t)km +k(u (t)−K ) kp , ε p ∗ ε m ε ∗ + p (cid:16) (cid:17) the previous inequality and Step 1 warrant that ku (t)k ≤ C(p,t∗,η) for t ∈ [t∗,t∗+τ ] and ε ∈ (0,1) . ε p η Step 3 - L∞-estimates: As a direct consequence of Step 2 with p = d + 1 and Morrey’s embedding theorem (∇φ ) is bounded in L∞((t∗,t∗ + τ ) × Rd;Rd). This property in turn ε ε η implies that (u ) is bounded in L∞((t∗,t∗ +τ )×Rd) and we refer to [5, Lemma 3.2] and [20] ε ε η for a proof (see also [31, Section 5] and [30, Theorem 1.2] for alternative arguments). (cid:3) As a consequence of the previous lemma, we are able to construct a free energy solution defined on a maximal existence time. Theorem 2.4 (Maximal free energy solution) Under assumption (1.2) on the initial con- dition there are T ∈ (0,∞] and a free energy solution u to (1.4) on [0,T ) with the following ω ω alternative: Either T = ∞ or T < ∞ and ku(t)k → ∞ as t ր T . Furthermore there exists ω ω m ω a positive constant C depending only on d such that u satisfies 0 −(m−1) ku(t )km ≤ ku(t )k−m/(m−1) −C (t −t ) (2.10) 2 m 1 m 0 2 1 (cid:16) (cid:17) for t ∈ [0,T ) and t ∈(t ,T ). 1 ω 2 1 ω Proof. We put ξ (t) = sup ku (t)k ∈ (0,∞] for t ≥ 0 and p ∈ [m,∞] and p ε∈(0,1) ε p T = sup{T > 0 : ξ ∈ L∞(0,T)} . 1 m Clearly the definition of the sequence (uε) and (1.2) ensure that ξ (0) is finite for all p ∈ [m,∞]. 0 ε p By Lemma 2.3 there exists t > 0 such that ξ is boundedon [0,t ] for all p ∈ [m,∞]. Then (2.2) 1 p 1 is fulfilled for T = t and there is a free energy solution to (1.4) on [0,t ) by Proposition 2.1 1 1 and (2.9). This ensures in particular that T ≥ t > 0. 1 1 We next claim that ξ ∈ L∞(0,T) for any T ∈[0,T ) . (2.11) ∞ 1 Indeed, consider T∞ = sup{T ∈ (0,T ) : ξ ∈ L∞(0,T)} and assume for contradiction that 1 1 ∞ T∞ < T . Then ξ belongs to L∞(0,T∞) and we put η = kξ k and t∗ = T∞−(τ /2), 1 1 m 1 m L∞(0,T1∞) 1 η τ being defined in Lemma 2.3. As ξ (t∗) ≤ η and ξ (t∗) is finite we may apply Lemma 2.3 η m ∞ to deduce that both ξ and ξ belong to L∞(t∗,t∗ +τ ), the latter property contradicting the m ∞ η definition of T∞ as t∗+τ = T∞+(τ /2). 1 η 1 η Now, thanks to (2.11), (2.2) is fulfilled for any T ∈ [0,T ) and the existence of a free energy 1 solution u to (1.4) on [0,T ) follows from Proposition 2.1 and (2.9). Moreover, either T = ∞ or 1 1 T < ∞ and ku(t)k → ∞ as t ր T , and the proof of Theorem 2.4 is complete with T = T . 1 m 1 ω 1 Or T < ∞ and 1 liminfku(t)k < ∞ . m t→T1 In that case, there are η > 0 and an increasing sequence of positive real numbers (s ) such j j≥1 that s → T as j → ∞ and ku(s )k ≤ η. Fix j ≥ 1 such that s ≥ T − (τ /2) with τ j 1 j m 0 j0 1 η η defined in Lemma 2.3 and put u˜ = u(s ); According to Definition 1.1 and (2.4) u˜ fulfils (1.2) 0 j0 0 and we may proceed as above to obtain a free energy solution u˜ to (1.4) on [0,T ) for some 2 T ≥ τ . Setting u¯(t) = u(t) for t ∈ [0,s ] and u¯(t) = u˜(t − s ) for t ∈ [s ,s + T ) we 2 η j0 j0 j0 j0 2 first note that u¯ is a free energy solution to (1.4) on [0,s +T ) and a true extension of u as j0 2 9 s + T ≥ T − (τ /2) + τ ≥ T + (τ /2). We then iterate this construction as long as the j0 2 1 η η 1 η alternative stated in Theorem 2.4 is not fulfilled to complete the proof. Thanks to the regularity of weak solutions we may next proceed as in the proof of (2.8) to deduce (2.10). (cid:3) Corollary 2.5 (Lower bound on the blow-up time) Let u be a free energy solution to (1.4) on [0,T ) with an initial condition u satisfying (1.2). If T is finite, then ω 0 ω ku(t)k ≥ [C (T −t)]−(m−1)/m , m 0 ω where C is defined in Theorem 2.4. 0 Proof. Let t ∈ (0,T ) and t ∈ (t,T ). By (2.10), we have ω 2 ω ku(t )k−m/(m−1) ≥ ku(t)k−m/(m−1) −C (t −t). 2 m m 0 2 Letting t going to T gives 2 ω 0 ≥ ku(t)k−m/(m−1) −C (T −t), m 0 ω hence the expected result. (cid:3) 3 The free energy functional F As we have just seen in the existence proof, the existence time of a free energy solution to (1.4) heavily depends on the behaviour of its Lm-norm. As the free energy F involves the Lm-norm, the information given F will beof paramount importance. Let us then proceed to a deeper study of this functional. Lemma 3.1 (Scaling properties of the free energy) Given h ∈ L1(Rd) ∩ Lm(Rd), let us define h (x) := λdh(λx), then λ F[h ]= λd−2F[h] for all λ ∈ (0,∞) . λ Proof. We have 1 c h(λx)h(λy) F[h ]= λ2d−2h(λx)m dx− d λ2d dxdy λ m−1 ZRd 2 ZZRd×Rd |x−y|d−2 λd−2 c λd−2 h(x)h(y) = h(x)m dx− d dxdy m−1 ZRd 2 ZZRd×Rd |x−y|d−2 = λd−2F[h] , giving the announced scaling property. (cid:3) We next establish a variant to the Hardy-Littlewood-Sobolev (VHLS) inequality: Lemma 3.2 (VHLS inequality) For h ∈ L1(Rd)∩Lm(Rd) we put h(x)h(y) W(h) := dxdy. ZZRd×Rd |x−y|d−2 Then W(h) C := sup , h∈ L1(Rd)∩Lm(Rd) < ∞ . (3.1) ∗ h6=0(khk2/dkhkm ) 1 m 10

See more

The list of books you might like

Most books are stored in the elastic cloud where traffic is expensive. For this reason, we have a limit on daily download.