We are concerned with the uniqueness of mild solutions in the critical Lebesgue space Ln2(Rn) for the parabolic-elliptic Keller-Segel system, n≥4. For that, we prove the bicontinuity of the bilinear term of the mild formulation in the critical weak-Ln2 space, without using Kato time-weighted norms, time-spatial mixed Lebesgue norms (i.e., Lq((0,T);Lp)-norms with q≠∞), and any other auxiliary norms. Our proofs are based on Yamazaki's estimate, duality and Hölder's inequality, as well as an adapted Meyer-type argument. Since they are different from those of Kozono, Sugiyama and Yahagi [J. Diff. Eq. 253 (2012)] and it is not clear whether mild solutions are weak solutions in the critical C([0,T);Ln2), our results complement theirs in a twofold way. Moreover, the bilinear estimate together heat semigroup estimates yield a well-posedness result whose dependence with respect to the decay rate γ of the chemoattractant is also analyzed.
Citation: Lucas C. F. Ferreira. On the uniqueness of mild solutions for the parabolic-elliptic Keller-Segel system in the critical Lp-space[J]. Mathematics in Engineering, 2022, 4(6): 1-14. doi: 10.3934/mine.2022048
[1] | Lucio Boccardo . A "nonlinear duality" approach to W1,10 solutions in elliptic systems related to the Keller-Segel model. Mathematics in Engineering, 2023, 5(5): 1-11. doi: 10.3934/mine.2023085 |
[2] | Kentaro Fujie, Jie Jiang . A note on construction of nonnegative initial data inducing unbounded solutions to some two-dimensional Keller–Segel systems. Mathematics in Engineering, 2022, 4(6): 1-12. doi: 10.3934/mine.2022045 |
[3] | Kyungkeun Kang, Dongkwang Kim . Existence of generalized solutions for Keller-Segel-Navier-Stokes equations with degradation in dimension three. Mathematics in Engineering, 2022, 4(5): 1-25. doi: 10.3934/mine.2022041 |
[4] | Mikyoung Lee, Jihoon Ok . Local Calderón-Zygmund estimates for parabolic equations in weighted Lebesgue spaces. Mathematics in Engineering, 2023, 5(3): 1-20. doi: 10.3934/mine.2023062 |
[5] | François Murat, Alessio Porretta . The ergodic limit for weak solutions of elliptic equations with Neumann boundary condition. Mathematics in Engineering, 2021, 3(4): 1-20. doi: 10.3934/mine.2021031 |
[6] | Nicolai Krylov . On parabolic Adams's, the Chiarenza-Frasca theorems, and some other results related to parabolic Morrey spaces. Mathematics in Engineering, 2023, 5(2): 1-20. doi: 10.3934/mine.2023038 |
[7] | Márcio Batista, Giovanni Molica Bisci, Henrique de Lima . Spacelike translating solitons of the mean curvature flow in Lorentzian product spaces with density. Mathematics in Engineering, 2023, 5(3): 1-18. doi: 10.3934/mine.2023054 |
[8] | Patrizia Pucci, Letizia Temperini . On the concentration–compactness principle for Folland–Stein spaces and for fractional horizontal Sobolev spaces. Mathematics in Engineering, 2023, 5(1): 1-21. doi: 10.3934/mine.2023007 |
[9] | María Ángeles García-Ferrero, Angkana Rüland . Strong unique continuation for the higher order fractional Laplacian. Mathematics in Engineering, 2019, 1(4): 715-774. doi: 10.3934/mine.2019.4.715 |
[10] | Patrizia Di Gironimo, Salvatore Leonardi, Francesco Leonetti, Marta Macrì, Pier Vincenzo Petricca . Existence of solutions to some quasilinear degenerate elliptic systems with right hand side in a Marcinkiewicz space. Mathematics in Engineering, 2023, 5(3): 1-23. doi: 10.3934/mine.2023055 |
We are concerned with the uniqueness of mild solutions in the critical Lebesgue space Ln2(Rn) for the parabolic-elliptic Keller-Segel system, n≥4. For that, we prove the bicontinuity of the bilinear term of the mild formulation in the critical weak-Ln2 space, without using Kato time-weighted norms, time-spatial mixed Lebesgue norms (i.e., Lq((0,T);Lp)-norms with q≠∞), and any other auxiliary norms. Our proofs are based on Yamazaki's estimate, duality and Hölder's inequality, as well as an adapted Meyer-type argument. Since they are different from those of Kozono, Sugiyama and Yahagi [J. Diff. Eq. 253 (2012)] and it is not clear whether mild solutions are weak solutions in the critical C([0,T);Ln2), our results complement theirs in a twofold way. Moreover, the bilinear estimate together heat semigroup estimates yield a well-posedness result whose dependence with respect to the decay rate γ of the chemoattractant is also analyzed.
We are concerned with the parabolic-elliptic Keller-Segel (or Patlak-Keller-Segel) system
{∂tu=∇⋅(∇u−u∇v),inx∈Rn, t∈(0,T),−△v+γv=κu,inx∈Rn, t∈(0,T),u|t=0=u0≥0,inx∈Rn, | (1.1) |
where 0<T≤∞, u(x,t)≥0 represents the density of cells and v(x,t)≥0 is the concentration of the chemoattractant. The parameters γ≥0 and κ>0 denote the decay and production rate of the chemoattractant, respectively. The model works for n≥2 but, as will be seen later, we restrict themselves to n≥4 due to technical issues.
System (1.1) is a chemotaxis model with a mathematical structure of parabolic-elliptic type. It is used in the study of aggregation of biological species, describing the behavior of organisms (e.g., bacteria) that move towards high concentration of a chemical secreted by themselves.
In view of (1.1)2, we can express v=κ(−Δ+γI)−1u and, according to Duhamel's principle, the Cauchy problem (1.1) can be formally converted to the integral equation
u(t)=G(t)u0+B(u,u)(t), | (1.2) |
where G(t)=eΔt stands for the heat semigroup and the bilinear term B is given by
B(u,w)(t)=−κ∫t0∇x⋅G(t−s)[u∇x(−Δ+γI)−1w](s)ds. | (1.3) |
Functions u(x,t) satisfying (1.2) are called mild solutions for (1.1). Here the mild formulation (1.2) is considered in a suitable dual sense, see Section 3 for details.
For γ=0, Eqs. (1.1)1–(1.1)2 has the scaling
u(x,t)→λ2u(λx,λ2t), | (1.4) |
which, for the initial data, induces
u0(x)→λ2u0(λx). | (1.5) |
Spaces invariant under the scaling (1.5), namely ‖u0‖X≈‖λ2u0(λx)‖X for all λ>0, are called critical spaces for (1.1).
In the dimension n=2, it is well-known that there exists a threshold value for the initial mass M=∫u0dx that decides if solutions exist globally (M<8π/κ) or blow up in a finite time (M>8π/κ) (see, e.g., [6,7]). Note that the space L1(R2) is critical for (1.1). For n≥3, one might wonder if some critical space could play a similar role as the L1-space in n=2 (for example, Ln2(Rn)), however, it is still an open problem to know whether there exists such a suitable space. In connection with that, in dimensions n≥3, there is a huge literature about existence of mild solutions for (1.1) and its parabolic-parabolic version with smallness conditions on the initial data in critical spaces. Without making a complete list, we mention the results in L1∩Ln2 [10], Ln2 [22] (weak solutions), Marcinkiewicz Ln2,∞ (weak−Lp spaces) [23], PMn−2 [3], Besov ˙Bnq−2q,∞ [18], Triebel-Lizorkin ˙F−2∞,2 [19], Morrey Mq,n−2q [2], Fourier-Besov F˙B−21,r [18], Besov-Morrey Nn−μq−2q,μ,∞ [13], and Fourier-Besov-Morrey spaces FNn−2−n−μqq,μ,∞[9], among others. It is worth noting that most of the above existence results of small mild solutions in critical spaces were inspired by those for Navier-Stokes equations, see, e.g., [1,11,15,20,21,24,25,30], and their references.
On the other hand, the uniqueness in critical spaces X is more subtle and needs some care. For n≥3, most of the above existence results are proved by constructing a fixed point argument in time-dependent spaces with norms composed of two ou more parts. One is the norm of the persistence space L∞((0,∞);X) and the others are auxiliary norms such as Kato time-weighted type norms, time-spatial mixed Lebesgue norms (i.e., Lq((0,T);Lp)-norms or, more generally, Lq((0,T);Y) with q≠∞) and Chemin-Lerner type norms, which are used to control the bilinear term B(u,w). Also, solutions are continuous at t>0 but only time-weakly continuous at t=0+, since the heat semigroup {etΔ}t≥0 is not strongly continuous at t=0+. This lack of continuity can be overcome by considering either the maximal subspace ˜X in which {etΔ}t≥0 is continuous or the closure of C∞0(Rn) in X, and then solutions belong to C([0,T);˜X) with large initial data u0∈˜X and small T>0. The estimates involving auxiliary norms in the proof of existence results, in principle, provide only a conditional uniqueness result, that is, uniqueness in a space more restricted than the natural one C([0,T);˜X). For the sake of completeness, in the case n=2 we would like to mention the uniqueness results of weak/mild solutions for (1.1) (and its parabolic-parabolic version) with finite mass, finite second moment and finite entropy (see, e.g., [8,12,27] and their references).
A way to obtain unconditional uniqueness in the critical class C([0,T);˜X) (or uniqueness of small solutions in L∞((0,∞);X)) is to prove the bilinear estimate
‖B(u,w)‖L∞((0,T);X)≤C‖u‖L∞((0,T);X)‖w‖L∞((0,T);X), | (1.6) |
where C>0 is a constant. This approach has already been employed in the context of Navier-Stokes equations. For example, see [11,14,25,26,28,30] to results in the framework of critical Lebesgue, Marcinkiewicz, Morrey and weak-Morrey spaces.
Next, let us discuss in more detail the works [22] and [23], which are more directly related to our results. In [23], Kozono-Sugiyama proved local well-posedness of mild solutions for (1.1) with small data u0∈Ln2,∞ and n≥3, where the existence and uniqueness are obtained in the class
u∈BC((0,T);Ln2,∞) and tβu∈BC((0,T);Lq) with n2<q<n, | (1.7) |
where u is time-weakly continuous at t=0+ and β=1−n2q. Also, u∈BC([0,T);Ln2) provided that u0∈Ln2↪Ln2,∞. They employed a point-fixed argument by using Kato's approach (see [20]) and their bilinear estimates are performed with the help of the auxiliary time-weighted norm supt∈(0,T)tα‖u(⋅,t)‖Lq. The solution u is global (T=∞) if ‖u0‖Ln2,∞ is small enough. For the uniqueness, besides assuming (1.7), it is required a smallness condition in the auxiliary norm, that is, the uniqueness is obtained in a space more restricted than BC((0,T);Ln2,∞). Moreover, assuming additional regularity on the initial data, they obtained the existence of local (or small global) strong Lp-solutions.
In [22], Kozono-Sugiyama-Yahagi proved the existence of local weak solutions u with respect to a suitable set of test functions. More precisely, for initial data u0∈Ln2(Rn), they obtained local-in-time solutions in
BC([0,T);Ln2(Rn))∩Lq((0,T);Lp(Rn)), |
where T>0 is small enough, n≥3, 2/q+n/p=2, 2<q<∞, n/2<p<n, and p≤n2/2(n−2). These weak solutions can be extended globally (T=∞) if ‖u0‖Ln2 is small enough. They constructed solutions via an approximation scheme of strong solutions whose existence was obtained in [23]. The uniqueness was obtained in the class Lq((0,T);Lp(Rn)) with n≥3, 2/q+n/p=2 and n/2<p<n. For n≥4, the uniqueness in the natural persistence space C([0,T);Ln2(Rn)) was obtained without any further condition, while the uniqueness in L∞((0,T);Ln2(Rn)) required small conditions on u and u0. For that, Kozono-Sugiyama-Yahagi converted the uniqueness problem to the one of obtaining global strong solutions for the associated adjoint equation, where coefficients depend on weak solutions, by using maximal regularity of the heat equation and suitable estimates involving Lq((0,T);Lp(Rn))-norms with q≠∞.
As pointed out by the authors of [22], it is not clear whether mild solutions satisfy their weak formulation. So, a natural question is to know if we have unconditional uniqueness of mild solutions for (1.1) in C([0,T);Ln2(Rn)), as well as the uniqueness in L∞((0,T);Ln2(Rn)) with supt∈(0,T)‖u(⋅,t)‖Ln2 small enough. Another question is to know if it is possible to obtain these uniqueness properties (and also existence and continuous dependence on initial data) without using auxiliary norms such as, for example, those in [22] and [23].
In this paper we obtain positive answers for the above questions for n≥4. First, inspired by [30], we prove estimate (1.6) with X=Ln2,∞(Rn) by means of careful estimates on the predual space of Ln2,∞(Rn), that is, the Lorentz space Lnn−2,1(Rn) (see Section 2 for the definition). So, adapting arguments found in [28], we obtain the unconditional uniqueness of mild solutions in the class C([0,T);Ln2(Rn)) with 0<T≤∞ and initial data u0∈Ln2(Rn) (see Theorem 3.1), because Ln2(Rn) is contained in ˜X (maximal subspace of Ln2,∞(Rn) where {etΔ}t≥0 is continuous). Moreover, using heat semigroup estimates and (1.6), we have the well-posedness of small solutions in L∞((0,T);Ln2,∞(Rn)) with 0<T≤∞. Since we have the continuous inclusion Ln2↪ Ln2,∞, the uniqueness of mild solutions in L∞((0,T);Ln2(Rn)) holds true provided that supt∈(0,T)‖u(⋅,t)‖Ln2 is small enough. Furthermore, we analyze the dependence of the well-posedness result with respect to the decay rate γ of the chemoattractant (see Remark 3.2).
Finally, we observe that our results work well by considering non-negative u0 and u in (1.1) as well as without any sign restrictions on them. However, we have considered the former for physical reasons associated to the model.
This paper is organized as follows. In Section 2, we give some preliminaries about Lorentz spaces and, in particular, weak-Lp spaces. Section 3 is dedicated to the statements of our results and some further remarks. The proofs of results are performed in Section 4.
This section is devoted to recalling some basic definitions and properties about Lorentz spaces.
Let |⋅| stands for the Lebesgue measure and let λf(s)=|{x∈Rn:|f(x)|>s}| be the distribution function of a measurable function f:Rn→R. The rearrangement of f is defined as
f∗(t)=inf{s>0:λf(s)≤t}, for t>0. | (2.1) |
The Lorentz space is the complete quasi-normed space
Lp,d(Rn)={f:Rn→R measurable; ‖f‖∗p,d<∞}, | (2.2) |
where the quasi-norm ‖f‖∗p,d is given by
‖f‖∗p,d={[∫∞0(t1p[f∗(t)])ddtt]1d,0<p<∞,1≤d<∞supt>0t1p[f∗(t)],0<p≤∞,d=∞. | (2.3) |
Taking d=p and d=∞ in (2.2), we obtain the Lebesgue space Lp,p(Rn)=Lp(Rn) and the Marcinkiewicz space Lp,∞, also called weak-Lp, respectively. The case p=∞ and 1≤d<∞ was removed from (2.3) because L∞,d is the trivial space.
In general, the quantity (2.3) is not a norm on Lp,d. Consider the double-rearrangement
f∗∗(t)=1t∫t0f∗(s)ds, |
and define the quantity ‖⋅‖p,d exactly as in (2.3) but replacing f∗ with f∗∗. For 1<p≤∞, the quantities ‖⋅‖p,d and ‖⋅‖∗p,d are topologically equivalent on Lp,d, since we have
‖⋅‖∗Lp,d≤‖⋅‖Lp,d≤pp−1‖⋅‖∗Lp,d. |
The pair (Lp,r,‖⋅‖p,d) is a Banach space and, unless mentioned otherwise, we consider it when 1<p≤∞. For 0<p≤1, Lp,d is endowed with ‖⋅‖∗p,d. In the case p=d=1 we have ‖⋅‖∗1,1 coincides with the standard L1-norm and L1,1=L1.
Lorentz spaces have the scaling property
‖f(λx)‖p,d=λ−np‖f(x)‖p,d. | (2.4) |
For 1≤d1≤p≤d2≤∞ and 1≤p≤∞, we have the continuous strict inclusions
Lp,1⊂Lp,d1⊂Lp⊂Lp,d2⊂Lp,∞ | (2.5) |
and then Lp,d becomes larger as the second index d goes from 1 to ∞.
Next we recall the Hölder inequality in Lorentz spaces (see [17,29]). Let 1<p1,p2,p3≤∞ and 1≤d1,d2,d3≤∞ be such that 1/p3=1/p1+1/p2 and 1/d1+1/d2≥1/d3. Then, there exists a constant C>0 (independent of f and g) such that
‖fg‖p3,d3≤C‖f‖p1,d1‖g‖p2,d2. | (2.6) |
For 1≤p,d<∞ (with d=1 when p=1), we have that the dual space of Lp,d is Lp′,d′(see [16]). In particular, the dual of Lp,1 is Lp′,∞ for 1≤p<∞. The space of compactly-supported smooth functions C∞0(Rn) is dense in Lp,d(Rn) for 1≤p,d<∞.
Young inequality works well in the framework of Lorentz spaces. In fact, if 1<p1,p2,p3≤∞ and 1≤d1,d2,d3≤∞ with 1/p3=1/p1+1/p2−1 and 1/d1+1/d2≥1/d3, then (see [29])
‖f∗g‖p3,d3≤C‖f‖p1,d1‖g‖p2,d2, | (2.7) |
where C>0 is a universal constant. Moreover, for p1=1 and 1<p=p2=p3≤∞, we have the inequality (see [4])
‖f∗g‖p,∞≤pp+1p(p−1)−1‖f‖L1‖g‖p,∞. | (2.8) |
Lorentz spaces has nice interpolation properties. For that, recall that the functor (⋅,⋅)θ,r constructed via the Kθ,q-method is exact of exponent θ on the categories of quasi-normed and normed spaces. Let 0<p1<p2≤∞, 0<θ<1, 1p=1−θp1+θp2 and 1≤d1,d2,d≤∞. Employing (⋅,⋅)θ,r in Lorentz spaces, we obtain (see [5,Chapter 5])
(Lp1,d1,Lp2,d2)θ,d=Lp,d. | (2.9) |
For 1<p≤∞ and 1≤d≤∞, by interpolating (2.8), we arrive at the inequality
‖f∗g‖p,d≤C‖f‖L1‖g‖p,d, | (2.10) |
where C>0 is a universal constant.
In this section we state the bilinear estimate (1.6) in weak- Ln2 and uniqueness result. Also, we make some comments about global existence and its dependence on the decay rate γ of the chemoattractant, non-negativity and mass conservation.
Before proceeding, we point out that the mild formulation (1.2) and its bilinear term should be meant in a suitable dual sense in the Ln2,∞-setting by using its predual space Lnn−2,1 and the duality pairing ⟨f,g⟩=∫Rnfgdx (see [30]). More precisely, for u,w∈L∞((0,T);Ln2,∞) we define B(u,w) by
⟨B(u,w),ϕ⟩=−κ∫t0⟨(u∇x(−Δ+γI)−1w),∇xG(t−s)ϕ⟩ds, | (3.1) |
for all ϕ∈Lnn−2,1(Rn) and a.e. t∈(0,T). Note also that, for u0∈Ln2,∞(Rn), the convolution G(t)u0 is well-defined and
⟨G(t)u0,ϕ⟩=⟨u0,G(t)ϕ⟩, for all ϕ∈Lnn−2,1(Rn). | (3.2) |
Thus, the formulation (1.2) should be meant as
⟨u(⋅,t),ϕ⟩=⟨u0,G(t)ϕ⟩−κ∫t0⟨[u∇x(−Δ+γI)−1u](s),∇xG(t−s)ϕ⟩ds, |
for all ϕ∈Lnn−2,1(Rn) and a.e. t∈(0,T). In other words, taking into account (3.2), u verifies (1.2) with B(u,u) given by (3.1).
Our results read as follows.
Theorem 3.1. Let n≥4,γ≥0, κ>0 and 0<T≤∞.
(i) (Bilinear estimate) Let B(⋅,⋅) be the bilinear form (1.3). There exists a constant K>0 (independent of γ) such that
‖B(u,w)‖L∞((0,T);Ln2,∞)≤κg(γ)K‖u‖L∞((0,T);Ln2,∞)‖w‖L∞((0,T);Ln2,∞), | (3.3) |
for all u,v∈L∞((0,T);Ln2,∞(Rn)), where g(γ)=1 and g(γ)=γ−(n−1) if γ=0 and γ>0, respectively.
(ii) (Uniqueness) Let u and w be mild solutions of (1.1) in the class C([0,T);Ln2(Rn)) with the same initial data u0. Then u(⋅,t)=w(⋅,t) in Ln2(Rn), for all t∈[0,T).
Remark 3.2. (Further comments)
(i) (The restriction n≥4) Due to the duality arguments in our proofs, we need to employ Lemma 4.3 with p=n2 and r=n3>1, and then n≥4. In turn, we need r>1 for Hölder's inequality (2.6) and the quantity ‖⋅‖r,∞ in Lr,∞ to be a norm. An interesting question would be to know whether, with some suitable relaxed integrability conditions, those arguments could be adapted to obtain a similar uniqueness result in lower dimensions.
(ii) (Well-posedness) Let 0<T≤∞, γ≥0, n≥4 and u0∈ Ln2,∞(Rn). Under a small assumption on ‖u0‖n2,∞ and a fixed-point argument, Theorem 3.1 together with heat semigroup estimates (see (4.1)) imply the well-posedness of small mild solutions in the class L∞((0,T);Ln2,∞(Rn)). In particular, we obtain the uniqueness of sufficiently small mild solutions in L∞((0,T);Ln2,∞(Rn)) or, using the continuous inclusion Ln2(Rn)↪Ln2,∞(Rn), in L∞((0,T);Ln2(Rn)). Moreover, the solution u belongs to BC((0,T);Ln2,∞(Rn)) with time-weak continuity at t=0+. For u0∈Ln2(Rn), one obtains u∈BC([0,T);Ln2(Rn)) and the smallness condition on u0 can be replaced with a smallness one on the existence-time T>0, and u∈ C([0,T);Ln2(Rn)) can be large. For γ=0 and T=∞, the obtained solution is self-similar provided that u0 is homogeneous of degree −2.
(iii) (Non-negativity and mass conservation) Due to the fixed-point argument, the solution u in item (ii) of this remark can be obtained as the limit of the Picard sequence u(k+1)=u(1)+B(u(k),u(k)), k∈N, and u(1)=G(t)u0. Let u0∈Ln2,∞(Rn) be non-negative. Using the parabolic regularization of the heat semigroup (see, e.g., (4.1)), an induction procedure, and the divergence structure of B(⋅,⋅), one can show that u(k) is smooth and non-negative, for each k. Since the convergence u(k)→u in L∞((0,T);Ln2,∞(Rn)) preserves non-negativity, it follows that u is non-negative. Furthermore, for u0∈L1+(Rn)∩Ln2,∞(Rn), using (4.1) and reducing the size of ‖u0‖n2,∞ (if necessary), one can show further integrability properties and polynomial time decay of Lp-norms for u(⋅,t) and B(⋅,t), for t>0, and then obtain u(⋅,t)∈L1+ and B(u,u)(⋅,t)∈L1. After, using the divergence form of B(⋅,⋅), one can obtain the mass conservation of the solution, that is, ∫Rnu(⋅,t)dx=∫Rnu0dx, for t>0.
(iv) (Large decay rate of the chemoattractant) Let 0<T≤∞. Considering γ>0 large enough, we can make κg(γ)K small enough and then obtain the well-posedness of mild solutions for (1.1) in L∞((0,T);Ln2,∞(Rn)), without smallness conditions on the existence-time T and initial data u0. More precisely, in order to employ a fixed-point argument, we need 4‖u0‖n2,∞(κg(γ)K)<1 which leads us to
γ>(4κK‖u0‖n2,∞)1n−1. |
We start with a lemma that will be useful to handle the coupling operator in (1.1) in Lorentz spaces.
Lemma 4.1. Let n≥2, 1<p<n, 1≤d≤∞ and 1q=1p−1n. The operator Lj=∂j(−Δ+γI)−1 is continuous from Lp,d(Rn) to Lq,d(Rn), for each j=1,2,...,n. Moreover, for γ>0, there exists a constant C>0 (independent of f and γ) such that
‖Ljf‖q,d≤Cγ−(n−1)‖f‖p,d. |
Proof. We can write the multiplier operator Lj as
Ljf=Kj,γ∗f, where ˆKj,γ(ξ)=−iξj|ξ|2+γ. |
Taking γ=0 and γ=1, we have that
Kj,0=(−iξj|ξ|2)∨∈Lnn−1,∞ and Kj,1=(−iξj|ξ|2+1)∨∈Lnn−1,∞. |
For γ=0, by using Young's inequality in Lorentz space (2.7) with p3=q, p1=nn−1, and p2=p, we obtain that
‖Ljf‖q,d=‖Kj,0∗f‖q,d≤C‖Kj,0‖p1,∞‖f‖p,d=C‖f‖p,d. |
Next we deal with the case γ>0. By a scaling argument, note that ˆKj,γ(ξ)=γ−1/2ˆKj,1(γ−1/2ξ) and then
Kj,γ(x)=γ−1/2γn/2Kj,1(γ1/2x)=γ(n−1)/2Kj,1(γ1/2x). |
Thus, again using (2.7) with the same indexes above, it follows that
‖Ljf‖q,d=‖Kj,γ∗f‖q,d≤C‖γ(n−1)/2Kj,1(γ1/2x)‖p1,∞‖f‖p,d=Cγ−n−12γ−n2p1‖Kj,1‖p1,∞‖f‖p,d=Cγ−(n−1)‖f‖p,d. |
In the lemma below, we recall some known estimates in Lorentz spaces for the heat semigroup (see [30]).
Lemma 4.2. (i) Let m∈{0}∪N, 1<r≤p≤∞, and 1≤d1,d2≤∞. Then, there exists a constant C>0 such that
‖∇mxG(t)φ‖p,d2≤Ct−m2−n2(1r−1p)‖φ‖r,d1,forallφ∈Lr,d1. | (4.1) |
(ii) (Yamazaki's estimate) Let 1<r<p<∞ . There is a constant C>0 such that
∫∞0sn2(1r−1p)−12‖∇x⋅G(s)ϕ‖p,1ds≤C‖ϕ‖r,1,forallϕ∈Lr,1. | (4.2) |
In the next lemma, by means of a duality argument and (4.2), we provide estimates for the linear operator
Q(f)(x)=∫∞0∇x⋅G(s)f(⋅,s)ds, | (4.3) |
which is linked to "Duhamel structure" of (1.3). Just like (1.3), the operator (4.3) is understood in the sense of duality, as explained in Section 3. Note that the lemma is valid for n≥2.
Lemma 4.3. Let n≥2 and 1<r<p<∞ be such that nr−np=1. There exists a constant C>0 such that
‖Q(f)‖p,∞≤Csupt>0‖f(⋅,t)‖r,∞, | (4.4) |
for all f∈L∞((0,∞), Lr,∞), where the supremum over t>0 is taken in the essential sense.
Proof. First, denoting the heat kernel by ς(x,t), we have that G(t)ϕ= ς(⋅,t)∗ϕ(⋅). A duality argument and Hölder's inequality (2.6) allow us to estimate the Lp,∞-norm of Q(f) as
‖Q(f)‖p,∞≤Csup‖ϕ‖Lp′,1=1|∫RnQ(f)ϕ(x)dx|≤Csup‖ϕ‖Lp′,1=1∫Rn∫∞0|((∇⋅g)(x,s)∗f(x,s))ϕ(x)|dsdx≤Csup‖ϕ‖Lp′,1=1∫∞0∫Rn|(∇⋅g(−x,s)∗ϕ(x))f(x,s)|dxds≤Csup‖ϕ‖Lp′,1=1∫∞0‖f(⋅,s)‖r,∞‖∇xg(x,s)∗ϕ‖r′,∞ds. | (4.5) |
Now note that 1<p′<r′<∞ and that the condition nr−np=1 implies
n2(1p′−1r′)−12=n2(1r−1p)−12=0. | (4.6) |
Using (4.6) and (4.2) in order to estimate the integral in the right-hand side of (4.5), we arrive at
‖Q(f)‖p,∞≤Csup‖ϕ‖Lp′,1=1∫∞0‖f(⋅,s)‖r,∞ sn2(1p′−1r′)−12‖g(x,s)∗ϕ‖r′,∞ds≤Csupt>0‖f(⋅,t)‖r,∞(Csup‖ϕ‖Lp′,1=1‖ϕ‖p′,1)=Csupt>0‖f(⋅,t)‖r,∞, |
as required.
Proof of item (i). Let 0<T≤∞. For each t∈(0,T), consider ft(x,s) given by
ft(⋅,s)=−κ(u∇(−Δ+γI)−1w)(⋅,t−s), for a.e. s∈(0,t), ft(⋅,s)=0, for s∈(t,∞), |
and note that (1.3) can be rewritten as
B(u,w)(t)=−κ∫t0∇⋅G(t−s)(s)[u∇(−Δ+γI)−1w]ds=Q(ft). |
Taking p=n2, l=n and r=n3, observe that 1r=1p+1l and 1l=1p−1n. Then, we can employ Hölder's inequality and Lemma 4.1 to estimate
sup0<s<T‖ft(⋅,s)‖r,∞ =κsup0<s<t<T‖(u∇(−Δ+γI)−1w)(⋅,t−s)‖r,∞≤Cκsup0<s<T‖u(⋅,s)‖p,∞sup0<s<T‖∇(−Δ+γI)−1w(⋅,s)‖l,∞≤Cκg(γ)sup0<s<T‖u(⋅,s)‖p,∞sup0<s<T‖w(⋅,s)‖p,∞, | (4.7) |
where g(γ)=1 and g(γ)=γ−(n−1) if γ=0 and γ>0, respectively. It follows that ft∈L∞((0,∞),Lr,∞(Rn)), for all t∈(0,T).
Now, noting that
nr−np=3−2=1, |
using Lemma 4.3 and afterwards (4.7), we arrive at
sup0<t<T‖B(u,w)‖p,∞=sup0<t<T‖Q(ft)‖p,∞≤Cκg(γ)sup0<t<T(sup0<s<T‖ft(⋅,s)‖r,∞)≤κg(γ)Ksup0<t<T‖u(⋅,t)‖p,∞sup0<t<T‖w(⋅,t)‖p,∞. |
With the bilinear estimate (3.3) in hands, the uniqueness of solutions in C([0,T);Ln2,∞) follows by adapting an argument due to Meyer [28] for our mild formulation (see also [11,25]).
Let u and w be mild solutions of (1.1) in C([0,T);Ln2(Rn)) such that u(0)=w(0)=u0∈Ln2(Rn). We claim that there exists 0<T1<T such that u(⋅,t)=w(⋅,t) in Ln2(Rn), for all t∈[0,T1). Considering h=u−w, h1=G(t)u0−u and h2=G(t)u0−w, we can rewrite the difference of the quadratic terms inside (1.3) as follows
u∇(−Δ+γI)−1u)−w∇(−Δ+γI)−1w)=h∇(−Δ+γI)−1u)+w∇(−Δ+γI)−1h)=h∇(−Δ+γI)−1G(t)u0)+G(t)u0∇(−Δ+γI)−1h)−h∇(−Δ+γI)−1h1)−h2∇(−Δ+γI)−1h). |
Thus, we can estimate h as
‖h(⋅,t)‖n2,∞=‖∫t0∇x⋅G(t−s)[u∇(−Δ+γI)−1u)−w∇(−Δ+γI)−1w)]ds‖n2,∞≤‖∫t0∇x⋅G(t−s)[h∇(−Δ+γI)−1h1)+h2∇(−Δ+γI)−1h)]ds‖n2,∞+‖∫t0∇x⋅G(t−s)[h∇(−Δ+γI)−1G(t)u0)+G(t)u0∇(−Δ+γI)−1h)]ds‖n2,∞:=J1(t)+J2(t). | (4.8) |
Using (3.3), the parcel J1(t) can be estimated as
J1(t)≤Kκg(γ)sup0<t<T‖h‖n2,∞(sup0<t<T‖h1‖n2,∞+sup0<t<T‖h2‖n2,∞). | (4.9) |
Next we turn to I2. Take 1<l<n2<β<n and b>β satisfying 1l=2n+1b, 1b=1β−1n and let ηβ=1−n2β. Also, note that 1l=1β+1n and 1n=2n−1n. Then, using (4.1), Hölder's inequality and afterwards (4.1), we arrive at
J2(t)≤C∫t0(t−s)−12(nl−nn/2)−12‖h(⋅,s)‖n2,∞‖∇(−Δ+γI)−1G(t)u0‖b,∞ds+C∫t0(t−s)−12(nl−nn/2)−12‖G(t)u0‖β,∞‖∇(−Δ+γI)−1h(⋅,s)‖n,∞ds≤C∫t0(t−s)−12(nl−nn/2)−12‖h(⋅,s)‖n2,∞‖G(t)u0‖β,∞ds+C∫t0(t−s)−12(nl−nn/2)−12‖G(t)u0‖β,∞‖h(⋅,s)‖n2,∞ds≤Csup0<t<T‖h(⋅,s)‖n2,∞(sup0<t<Ttηβ‖G(t)u0‖β,∞)∫t0(t−s)−n2βs−ηβds≤Csup0<t<T‖h(⋅,s)‖n2,∞(sup0<t<Ttηβ‖G(t)u0‖β,∞), | (4.10) |
for all t∈(0,T), where we used above that −n2β−ηβ+1=0 and
∫t0(t−s)−n2βs−ηβds=t−n2β−ηβ+1∫10(1−s)−n2βs−ηβds=C<∞. |
Inserting (4.9)–(4.10) into (4.8) yields
sup0<t<T‖h(⋅,t)‖n2,∞≤CM(T)sup0<t<T‖h(⋅,t)‖n2,∞, | (4.11) |
where
M(T)=(sup0<t<T‖h1(⋅,t)‖n2,∞+sup0<t<T‖h2(⋅,t)‖n2,∞+sup0<t<Ttηβ‖G(t)u0‖β,∞). | (4.12) |
Using that G(t)u0→u0 and u,v→u0 in Ln2(Rn), as t→0+, and Ln2(Rn)↪Ln2,∞(Rn), we obtain
lim supt→0+(‖h1(⋅,t)‖n2,∞+‖h2(⋅,t)‖n2,∞)=0. | (4.13) |
Moreover, since u0∈Ln2(Rn), there exists a sequence {u0,k}k⊂Ln2(Rn)∩Lβ,∞(Rn) such that u0,k→u0 in Ln2(Rn)↪Ln2,∞(Rn). In fact, it is sufficient to take u0,k=G(1k)u0 and use (4.1). Then, we can estimate
lim supt→0+tηβ‖G(t)u0‖β,∞≤sup0<t<∞tηβ‖G(t)(u0−u0,k)‖β,∞+lim supt→0+tηβ‖G(t)u0,k‖β,∞≤C‖u0−u0,k‖n2,∞+C‖u0,k‖β,∞lim supt→0+tηβ≤C‖u0−u0,k‖n2→0, as k→∞. | (4.14) |
In view of (4.12), (4.13) and (4.14), we can take T=T1>0 small enough such that CA(T1)<1. Now, estimate (4.11) implies that h(⋅,t)=0 for all t∈ [0,T1),which gives the desired claim.
Finally, we are going to show that the smallness condition on T1 can be removed. For that, consider
T∗=sup{˜T:0<˜T<T, u(⋅,t)=w(⋅,t) in Ln2 for all t∈[0,˜T)}. |
If T∗=T, then u=w in [0,T), as desired. If T∗<T≤∞, we have that u(⋅,t)=w(⋅,t) for all t∈[0,T∗). By time-continuity, we obtain that u(⋅,T∗)=w(⋅,T∗) and then, by the first part, there exists a sufficiently small ρ>0 such that u(⋅,t)=w(⋅,t) for all t∈[T∗,T∗+ρ). So, u(⋅,t)=w(⋅,t) in [0,T∗+ρ), which contradics the maximality of T∗<∞.
L. C. F. Ferreira was supported by FAPESP and CNPQ, Brazil.
The author declares no conflict of interest.
[1] | M. F. de Almeida, L. C. F. Ferreira, L. S. M. Lima, Uniform global well-posedness of the Navier-Stokes-Coriolis system in a new critical space, Math. Z., 287 (2017), 735–750. |
[2] | P. Biler, The Cauchy problem and self-similar solutions for a nonlinear parabolic equation, Studia Math., 114 (1995), 181–205. |
[3] | P. Biler, M. Cannone, I. A. Guerra, G. Karch, Global regular and singular solutions for a model of gravitating particles, Math. Ann., 330 (2004), 693–708. |
[4] | P. Benilan, H. Brezis, M. Crandall, A semilinear equation in L1(Rn), Ann. Scuola Norm. Sup. Pisa, Ser. 4 (1975), 523–555. |
[5] | J. Berg, J. Lofstrom, Interpolation spaces, Berlin-Heidelberg-New York: Springer, 1976. |
[6] | A. Blanchet, J. Dolbeault, B. Perthame, Two-dimensional Keller-Segel model: optimal critical mass and qualitative properties of the solutions, Electron. J. Differ. Eq., 2006 (2006), 1–33. |
[7] | V. Calvez, L. Corrias, The parabolic-parabolic Keller-Segel model in R2, Commun. Math. Sci., 6 (2008), 417–447. |
[8] | K. Carrapatoso, S. Mischler, Uniqueness and long time asymptotics for the parabolic-parabolic Keller-Segel equation, Commun. Part. Diff. Eq., 42 (2017), 291–345. |
[9] | X. Chen, Well-posedness of the Keller-Segel system in Fourier-Besov-Morrey spaces, Z. Anal. Anwend., 37 (2018), 417–433. |
[10] | L. Corrias, B. Perthame, H. Zaag, Global solutions of some chemotaxis and angiogenesis systems in high space dimensions, Milan J. Math., 72 (2004), 1–28. |
[11] | M. Cannone, Harmonic analysis tools for solving the incompressible Navier-Stokes equations, In: Handbook of mathematical fluid dynamics, Amsterdam: North-Holland, 3 (2004), 161–244. |
[12] | G. E. Fernández, S. Mischler, Uniqueness and long time asymptotic for the Keller-Segel equation: the parabolic-elliptic case, Arch. Rational Mech. Anal., 220 (2016), 1159–1194. |
[13] | L. C. F. Ferreira, J. C. Precioso, Existence and asymptotic behaviour for the parabolic-parabolic Keller-Segel system with singular data, Nonlinearity, 24 (2011), 1433–1449. |
[14] | L. C. F. Ferreira, On a bilinear estimate in weak-Morrey spaces and uniqueness for Navier-Stokes equations, J. Math. Pure. Appl., 105 (2016), 228–247. |
[15] | Y. Giga, T. Miyakawa, Navier-Stokes flow in R3 with measures as initial vorticity and Morrey spaces, Commun Part. Diff. Eq., 14 (1989), 577–618. |
[16] | L. Grafakos, Classical and modern Fourier analysis, Upper Saddle River, NJ: Pearson Education, Inc., 2004. |
[17] | R. Hunt, On L(p,q) spaces, L 'Enseignement Mathématique, 12 (1966), 249–276. |
[18] | T. Iwabuchi, Global well-posedness for Keller-Segel system in Besov type spaces, J. Math. Anal. Appl., 379 (2011), 930–948. |
[19] | T. Iwabuchi, M. Nakamura, Small solutions for nonlinear heat equations, the Navier-Stokes equation, and the Keller-Segel system in Besov and Triebel-Lizorkin spaces, Adv. Differential Equ., 18 (2013), 687–736. |
[20] | T. Kato, Strong Lp-solutions of the Navier-Stokes equation in Rm, with applications to weak solutions, Math. Z., 187 (1984), 471–480. |
[21] | T. Kato, Strong solutions of the Navier-Stokes equations in Morrey spaces, Bol. Soc. Brasil Mat., 22 (1992), 127–155. |
[22] | H. Kozono, Y. Sugiyama, Y. Yahagi, Existence and uniqueness theorem on weak solutions to the parabolic-elliptic Keller-Segel system, J. Differ. Equations, 253 (2012), 2295–2313. |
[23] | H. Kozono, Y. Sugiyama, Strong solutions to the Keller-Segel system with the weak−Ln2 initial data and its application to the blow-up rate, Math. Nachr., 283 (2010), 732–751. |
[24] | H. Kozono, M. Yamazaki, Semilinear heat equations and the Navier-Stokes equation with distributions in new function spaces as initial data, Commun. Part. Diff. Eq., 19 (1994), 959–1014. |
[25] | P. G. Lemarie-Rieusset, Recent developments in the Navier-Stokes equations, Chapman and Hall, 2002. |
[26] | P. G. Lemarié-Rieusset, On some classes of time-periodic solutions for the Navier-Stokes equations in the whole space, SIAM J. Math. Anal., 47 (2015), 1022–1043. |
[27] | J.-G. Liu, J. Wang, Refined hyper-contractivity and uniqueness for the Keller-Segel equations, Appl. Math. Lett., 52 (2016), 212–219. |
[28] | Y. Meyer, Wavelets, paraproducts and Navier-Stokes equations, In: Current developments in mathematics 1996, Cambridge: International Press, 1999,105–212. |
[29] | R. O'Neil, Convolution operators and L(p,q) spaces, Duke Math. J., 30 (1963), 129–142. |
[30] | M. Yamazaki, The Navier-Stokes equations in the weak-Ln space with time-dependent external force, Math. Ann., 317 (2000), 635–675. |
1. | Achraf Azanzal, Chakir Allalou, Said Melliani, Gevrey class regularity and stability for the Debye-H¨uckel system in critical Fourier-Besov-Morrey spaces, 2022, 41, 2175-1188, 1, 10.5269/bspm.62517 | |
2. | Shouming Zhou, Li Zhang, On the Cauchy problem for Keller‐Segel model with nonlinear chemotactic sensitivity and signal secretion in Besov spaces, 2023, 0170-4214, 10.1002/mma.9104 | |
3. | Nguyen Thi Loan, Van Anh Nguyen Thi, Tran Van Thuy, Pham Truong Xuan, Periodic solutions of the parabolic–elliptic Keller–Segel system on whole spaces, 2024, 297, 0025-584X, 3003, 10.1002/mana.202300311 | |
4. | Nguyen Thi Loan, Pham Truong Xuan, Almost periodic solutions of the parabolic-elliptic Keller–Segel system on the whole space, 2024, 123, 0003-889X, 431, 10.1007/s00013-024-02023-8 | |
5. | Tran Van Thuy, Nguyen Thi Van, Pham Truong Xuan, On asymptotically almost periodic solutions of the parabolic-elliptic Keller-Segel system on real hyperbolic manifolds, 2024, 1468-9367, 1, 10.1080/14689367.2024.2360204 |