We construct generalized solutions for the Keller-Segel system with a degradation source coupled to Navier Stokes equations in three dimensions, in case that the power of degradation is smaller than quadratic. Furthermore, if the logistic type source is purely damping with no growing effect, we prove that solutions converge to zero in some norms and provide upper bounds of convergence rates in time.
Citation: Kyungkeun Kang, Dongkwang Kim. Existence of generalized solutions for Keller-Segel-Navier-Stokes equations with degradation in dimension three[J]. Mathematics in Engineering, 2022, 4(5): 1-25. doi: 10.3934/mine.2022041
[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] | Lars Eric Hientzsch . On the low Mach number limit for 2D Navier–Stokes–Korteweg systems. Mathematics in Engineering, 2023, 5(2): 1-26. doi: 10.3934/mine.2023023 |
[3] | 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 |
[4] | Ivan Fumagalli . Discontinuous Galerkin method for a three-dimensional coupled fluid-poroelastic model with applications to brain fluid mechanics. Mathematics in Engineering, 2025, 7(2): 130-161. doi: 10.3934/mine.2025006 |
[5] | Gennaro Ciampa, Gianluca Crippa, Stefano Spirito . Propagation of logarithmic regularity and inviscid limit for the 2D Euler equations. Mathematics in Engineering, 2024, 6(4): 494-509. doi: 10.3934/mine.2024020 |
[6] | Andrea Manzoni, Alfio Quarteroni, Sandro Salsa . A saddle point approach to an optimal boundary control problem for steady Navier-Stokes equations. Mathematics in Engineering, 2019, 1(2): 252-280. doi: 10.3934/mine.2019.2.252 |
[7] | Lucas C. F. Ferreira . On the uniqueness of mild solutions for the parabolic-elliptic Keller-Segel system in the critical Lp-space. Mathematics in Engineering, 2022, 4(6): 1-14. doi: 10.3934/mine.2022048 |
[8] | Yangyang Qiao, Qing Li, Steinar Evje . On the numerical discretization of a tumor progression model driven by competing migration mechanisms. Mathematics in Engineering, 2022, 4(6): 1-24. doi: 10.3934/mine.2022046 |
[9] | Giuseppe Procopio, Massimiliano Giona . Bitensorial formulation of the singularity method for Stokes flows. Mathematics in Engineering, 2023, 5(2): 1-34. doi: 10.3934/mine.2023046 |
[10] | Daniele Del Santo, Francesco Fanelli, Gabriele Sbaiz, Aneta Wróblewska-Kamińska . On the influence of gravity in the dynamics of geophysical flows. Mathematics in Engineering, 2023, 5(1): 1-33. doi: 10.3934/mine.2023008 |
We construct generalized solutions for the Keller-Segel system with a degradation source coupled to Navier Stokes equations in three dimensions, in case that the power of degradation is smaller than quadratic. Furthermore, if the logistic type source is purely damping with no growing effect, we prove that solutions converge to zero in some norms and provide upper bounds of convergence rates in time.
We consider a mathematical model to decribe the dynamics of biological organism influenced by chemical signal and living in fluid. The original Keller-Segel system was proposed to write the motion of biological individuals sensing gradient of a chemical substance and moving toward its higher concentration (see [9]). Such biological organisms often live in fluid, and thus their behaviors are influenced by motions of viscous fluid flows as well. There are, for example, the bacteria living in fluid such as Bacillus subtilus ([1,2,7,11,18,24]) or Escherichia coli ([12,22]) or phenomena of coral fertilization in sea resulting from the chemotatic behavior of sperm ([4,6,10,24]).
In this note, we study the following Keller-Segel system with degradation coupled to the Navier-Stokes equations in a bounded domain in three dimensions:
nt+u⋅∇n=Δn−∇⋅(n∇c)+ρn−μnq, | (1.1) |
ct+u⋅∇c=Δc−c+n, | (1.2) |
ut+(u⋅∇)u=Δu+∇P+n∇ϕ,∇⋅u=0 | (1.3) |
in Ω×(0,T), where Ω⊂R3 is a bounded domain with smooth boundary and T>0. Here n, c, u, and P are the population density of the chemotactic organisms, the concentration of signal substances, the fluid velocity, and the associated pressure, respectively. No flux condition is assigned for n and c at the boundary, and u has no slip boundary condition there, namely
∂n∂ν=∂c∂ν=0,u=0 on ∂Ω. | (1.4) |
We assume that initial data (n0,c0,u0) satisfies
{0⩽n0∈C0(¯Ω) with n0≢0,0⩽c0∈W1,∞(Ω),u0∈W2,2(Ω)∩W1,20(Ω) with ∇⋅u0=0. | (1.5) |
In case that the Eq (1.1) has the logistic degradation, i.e., q=2, Tao and Winkler [16] proved global existence and large time behavior of classical solutions to the system (1.1)–(1.3) in two dimensions. Such result was extended to the case of three dimensions, provided that the fluid equation is given by the Stokes system, instead of the Navier-Stokes equations, and μ is sufficiently large (see [15]).
For the chemotaxis-Navier-Stokes system (1.1)–(1.3) with q=2, the existence of generalized solutions was proved by Winkler [22].
To the best of our knowledge, if q<2, it is not known whether or not classical solutions exist globally in time for general data and parameters. Instead of classical solutions, recently it was shown in [8] that generalized solutions to the chemotaxis-Stokes system exists globally in time for q∈(2−1d,2), where d is dimensions two or three, i.e., d=2,3. (the notion of generalized solutions is reminded in Definition 2). In the absence of fluid, i.e., u=0, one can refer to [19,20,23] for generalized solutions.
The main objective of this note is to establish the existence of generalized solutions globally in time, in case that the degradation power q is less than 2, and the Navier-Stokes equations are coupled for the fluid equations in three dimensions.
To begin with, we recall the notion of generalized solution of (1.1)–(1.3). Firstly, we remind the γ−entrophy super(or sub) solution of the Eq (1.1).
Definition 1. Let γ∈(0,1). Assume that a pair of functions (n,c) and a vector field u satisfy the following:
∇n and ∇c are measurable in Ω×(0,∞),nγ,nγ−2|∇n|2,nγ−1∇n⋅∇c,nq+γ−1∈L1loc(¯Ω×[0,∞)),nγ∇c,nγu∈L1loc(¯Ω×[0,∞);R3),∇⋅u=0inD′(Ω×(0,∞)). |
Then such (n,c,u) is called a weak γ-entropy super-solution(resp., sub-) of the first equation in (1.1)–(1.3) if
−∞∫0∫Ωnγφt−∫Ωnγ0φ(⋅,0)⩾(⩽)γ(1−γ)∞∫0∫Ωnγ−2|∇n|2φ+∞∫0∫ΩnγΔφ+(1−γ)∞∫0∫ΩnγΔcφ+∞∫0∫Ωnγ∇c⋅∇φ+ργ∞∫0∫Ωnγφ−μγ∞∫0∫Ωnq+γ−1φ+∞∫0∫Ωnγu⋅∇φ, |
for all nonnegative φ∈C∞0(¯Ω×[0,∞)).
Next, we define the notion of the generalized solutions of (1.1)–(1.3).
Definition 2. A triple of two functions and a vector field
n∈L1loc(¯Ω×[0,∞)),c∈L1loc([0,∞);W1,1(Ω)),u∈L1loc([0,∞);W1,10(Ω,R3)) |
satisfying
cu∈L1loc(¯Ω×[0,∞)),u⊗u∈L1loc(¯Ω×[0,∞);R3×R3) |
is called a generalized solution of (1.1)–(1.3), if
−∞∫0∫Ωcφt−∫Ωc0φ(⋅,0)=−∞∫0∫Ω∇c⋅∇φ−∞∫0∫Ωcφ+∞∫0∫Ωnφ+∞∫0∫Ωcu⋅∇φ | (1.6) |
for all φ∈C∞0(¯Ω×[0,∞)) and, if ∇⋅u=0 in D′(Ω×(0,∞)) and
−∞∫0∫Ωuφt−∫Ωu0φ(⋅,0)=−∞∫0∫Ω∇u⋅∇φ+∞∫0∫Ω(u⊗u)⋅∇φ+∞∫0∫Ωn∇ϕ⋅φ | (1.7) |
for all φ∈C∞0(Ω×[0,∞);R3) with ∇⋅φ≡0, and if there exist γ1,γ2∈(0,1) such that (n,c,u) is a weak γ1-entropy super-solution and a weak γ2-entropy sub-solution of the first equations in (1.1)–(1.3).
For logistic coefficients ρ,μ and the potential function ϕ, we assume that
ρ∈R,μ>0 and ϕ∈C1(Ω). | (1.8) |
We are now ready to state our main result.
Theorem 1.1. Let q∈(2011,2). Then the Eqs (1.1)–(1.5) with (1.8) admit at least one generalized solution in the sense of Definition 2.
Remark 1. The result Theorem 1.1 is an improvemnt of that of [22], which showed the existence of the generalized solution in case that q=2. Furthermore, it is also an extension to the result of [8], since the Navier-Stokes equations are considered instead the Stokes system. In such case, the range of q is, however, restrictive, compared to the case that q∈(53,2) in [8]. This is mainly due to the fact that the control of u is more difficult for the Navier-Stokes equations, which causes lower regularity of u⋅∇c and, in turn, ∇c (see Lemma 3.6 for the details). Therefore, passing to the limit for regularized solutions, convergence to n∇c is well understood only for q∈(2011,2).
Next, in case that ρ⩽0, we can show that generalized solutions converge to zero in an appropriate sense, passing time to the limit. More precisely, we obtain the following:
Theorem 1.2. Let (n,c,u) be the generalized solution established in Theorem 1.1. If ρ=0, then (n,c,u) vanishes in L1(Ω)×Ll(Ω)×L2(Ω) as time tends to infinity. Furthermore, (n,c,u) satisfies
∫Ωn(⋅,t)dx⩽C(1+t)−1q−1,∫Ω|u(⋅,t)|2dx⩽C(1+t)−3q(4−q)−103(q−1)2 |
and∫Ω(c(⋅,t))ldx⩽{C(1+t)−2lq+q−3l3l(q−1)2,if1⩽l⩽3q−2,C(1+t)−3q−(5−2q)ll(3q−5)(q−1),if3q−2<l⩽3q5−2q. |
Morerover, if ρ<0, then (n,c,u) satisfies
∫Ωn(⋅,t)dx⩽Ceρt,∫Ω|u(⋅,t)|2dx⩽Ce−δ∗t |
and∫Ω(c(⋅,t))ldx⩽Ce−3q−(5−2q)l5(q−1)lρ∗t,if1⩽l⩽3q5−2q. |
where ρ∗=min, \delta_* = \frac{1}{2}\min\left\{ {\frac{C_p}{2}, -\rho\frac{5q-6}{3(q-1)}} \right\} and C_p is the Poincaré constant for \Omega.
Remark 2. The result of Theorem 1.2 can be extended to the case q = 2 and \rho = 0. In such case, in particular, estimates of c read as follows:
\int\limits_{\Omega} \left( {c(\cdot, t)} \right)^l {\,\mathrm{d}} x \leqslant \left\{ \begin{array}{ll} C(1+t)^{-\frac{l+2}{3l} }, & \mbox{ if }\quad1 \leqslant l \leqslant 4,\\ C(1+t)^{-\frac{6-l}{l}}, & \mbox{ if }\quad 4 < l \leqslant 6. \end{array} \right. |
This estimate of decay for c is slightly better, compared to those of [22,Section 8]. On the other hand, in case that q = 2 and \rho > 0 , it was also shown in [22] that if \mu > \chi\sqrt{\rho}/4 , then
\limsup\limits_{t\rightarrow \infty}\left\Vert {n(\cdot,t)-\frac{\rho}{\mu}} \right\Vert_{1}+\left\Vert {c(\cdot,t)-\frac{\rho}{\mu}} \right\Vert_{p}+\left\Vert {u(\cdot,t)} \right\Vert_{2} = 0,\qquad 1 \leqslant p < 6. |
This convergence is based on stabilization of a certain energy functional (see [22,Section 8]). Although similar results are expected, such a method doesn't seem to be valid unless q = 2 . Therefore, we leave the asymptotic behaviors as an open question in case that \rho > 0 and q < 2 .
This paper is organized as follows: In Section 2, we introduce an approximated system and recall some useful lemma for our purpose. Section 3 is devoted to obtaining estimates, independent of a regularizing parameter, of the approximated system. We then discuss the convergence of approximated solutions to a generalized solution in Section 4. Finally, in Section 5, asymptotic estimates are provided.
Throughout this paper, we shall abbreviate \left\Vert {f} \right\Vert_{L^p({\Omega})} as \left\Vert {f} \right\Vert_p for simplicity. Further, we denote by C>0 generic constants which may be different from line to line.
In the following proposition we define an appropriate approximated system to (1.1)–(1.3), for which global classical solutions can be verified. The approximated system is given by
\begin{equation} \begin{cases} {\partial}_tn_ {\epsilon}+u_ {\epsilon}\cdot {\nabla } n_ {\epsilon} = \Delta n_ {\epsilon} - {{\nabla }\cdot}(n_ {\epsilon} {\nabla } c_ {\epsilon})+\rho n_ {\epsilon}-\mu n_ {\epsilon}^q- {\epsilon} n_ {\epsilon}^ {\kappa},\\ {\partial}_tc_ {\epsilon}+u_ {\epsilon}\cdot {\nabla } c_ {\epsilon} = \Delta c_ {\epsilon} -c_ {\epsilon}+n_ {\epsilon} ,\\ {\partial}_tu_ {\epsilon}+(Y_ {\epsilon} u_ {\epsilon}\cdot {\nabla }) u_ {\epsilon} = \Delta u_ {\epsilon} + {\nabla } P_ {\epsilon}+n_ {\epsilon} {\nabla }\phi ,\\ {{\nabla }\cdot} u_ {\epsilon} = 0 ,\\ \frac{\partial {n_ {\epsilon}}}{\partial {\nu}} = \frac{\partial {c_ {\epsilon}}}{\partial {\nu}} = u_ {\epsilon} = 0, \\ n_ {\epsilon}(x,0) = n_0,\quad c_ {\epsilon}(x,0) = n_0,\quad u_ {\epsilon}(x,0) = u_0. \end{cases} \end{equation} | (2.1) |
Here {\epsilon}\in(0, 1), {\kappa}>2 and Y_ {\epsilon} is the Yosida approximation defined by
Y_ {\epsilon} f: = (I+ {\epsilon} A)^{-1}f,\quad f\in L^2_\sigma( {\Omega}), |
where A is the realization of the stokes operator in D(A) = W^{2, 2}({\Omega})\cap W^{1, 2}_{0, \sigma}({\Omega})\subset L^2_\sigma({\Omega}).
Following method of proofs developped in [8] and [22], one can prove the existence of classical solution of the approximated system (2.1). Since its verification is similar to thoes of [8] and [22], we skip its proof.
Proposition 1. For each {\epsilon}\in(0, 1), there exist functions
\begin{equation*} \begin{cases} n_ {\epsilon}\in C^0(\overline{ {\Omega}}\times [0,\infty)\cap C^{2,1}(\overline{ {\Omega}}\times (0\,\infty)), \\ c_ {\epsilon}\in C^0(\overline{ {\Omega}}\times [0,\infty)\cap C^{2,1}(\overline{ {\Omega}}\times (0\,\infty)), \\ u_ {\epsilon}\in C^0(\overline{ {\Omega}}\times [0,\infty)\cap C^{2,1}(\overline{ {\Omega}}\times (0\,\infty)), \\ P_ {\epsilon}\in C^{1,0}(\overline{ {\Omega}}\times (0,\infty)) \end{cases} \end{equation*} |
such that (n_ {\epsilon}, c_ {\epsilon}, u_ {\epsilon}, P_ {\epsilon}) solves (2.1) classically in \overline{ {\Omega}}\times (0, \infty).
We recall an effective inequality in Sobolev spaces called the Gagliardo-Nirenberg interpolation inequality. Here we only consider a version of bounded Lipschitz domain \Omega in \mathbb{R}^3. The proof can be found in [3,Theorem 1.5.2] and [13].
Lemma 2.1. Let 1 \leqslant p, r \leqslant \infty and 0 \leqslant n < m\in\mathbb{N}. Then there exist constants C_1 and C_2>0 such that
\begin{equation} \left\Vert {D^n f} \right\Vert_q \leqslant C\left\Vert {D^mf} \right\Vert_p^\theta\left\Vert {f} \right\Vert_r^{1-\theta}+C_2\left\Vert {f} \right\Vert_s ,\quad f\in\mathcal{D}'(\Omega) \end{equation} | (2.2) |
where \frac{1}{q}-\frac{n}{3} = \left({\frac{1}{p}-\frac{m}{3}} \right)\theta+\frac{1}{r}(1-\theta), \theta\in [\frac{n}{m}, 1], and s>0 is arbitrary.
The following two Lemmas named maximal estimates are crutial to obtain a regularity of approximated solutions (see [5,8,14]).
Lemma 2.2. Let T>0, v_0\in W^{1, p}(\Omega) and h\in L^p(0, T;L^p(\Omega; \mathbb{R}^3)) for 1 < p < \infty. Then there exists a unique solution v\in L^p(0, T;W^{1, p}(\Omega)) solving
\begin{equation*} \begin{cases} v_t-\Delta v = {{\nabla }\cdot} h,\quad\quad&(x,t)\in\Omega\times (0,T),\\ v(x,0) = v_0(x), &x\in\Omega,\\ \frac{\partial {v}}{\partial {\nu}} = 0, &(x,t)\in {\partial}\Omega\times (0,T). \end{cases} \end{equation*} |
Furthermore, v attains the following estimate.
\begin{equation} \int\limits_{0}^{T}\left\Vert {v(s)} \right\Vert_p^p {\,\mathrm{d}}{s}+\int\limits_{0}^{T}\left\Vert { {\nabla } v(s)} \right\Vert^p_p {\,\mathrm{d}} s \leqslant C_T\left( {\int\limits_{0}^{T}\left\Vert {h(s)} \right\Vert^p_p {\,\mathrm{d}} s+\left\Vert {v_0} \right\Vert^p_{W^{1,p}(\Omega)}} \right). \end{equation} | (2.3) |
Lemma 2.3. Let T>0 and p\in(1, 2]. Then for every v_0\in W^{1, \infty}({\Omega}) and h\in L^p\left({ {\Omega}\times(0, T)} \right), the following heat equation with Neumann boundary condition
\begin{equation} \begin{cases} v_t-\Delta v = h,\quad\quad\quad&(x,t)\in\Omega\times (0,T),\\ v(x,0) = v_0(x), &x\in\Omega,\\ \frac{\partial {v}}{\partial {\nu}} = 0, &(x,t)\in {\partial}\Omega\times (0,T) \end{cases} \end{equation} | (2.4) |
has a unique solution v\in W^{1, p}((0, T);L^p({\Omega}))\cap L^p((0, T);W^{2, p}({\Omega})) satisfying
\begin{equation} \left\Vert {v_t} \right\Vert_{L^p( {\Omega}\times(0,T))}+\left\Vert {v} \right\Vert_{L^p(0,T;W^{2,p}( {\Omega}))} \leqslant C_T\left( {\left\Vert {h} \right\Vert_{L^p( {\Omega}\times(0,T))}+1} \right) \end{equation} | (2.5) |
with some C_T>0.
Proof. Set X = L^p({\Omega}) and X_1 = W^{2, p}_\nu({\Omega}): = \{f\in W^{2, p}({\Omega}) : \frac{\partial {f}}{\partial {\nu}} = 0 on {\partial} {\Omega}\}. From [14] and [19,Proposition 2] we have
\begin{equation*} \left\Vert {v_t} \right\Vert_{L^p\left( { {\Omega}\times\left( {0,T} \right)} \right)}+\left\Vert {v} \right\Vert_{L^p\left( {0,T;W^{2,p}( {\Omega})} \right)} \leqslant C_T\left( {\left\Vert {v_0} \right\Vert_{1-\frac{1}{p},p}+\left\Vert {h} \right\Vert_{L^p\left( { {\Omega}\times({0,T})} \right)}} \right), \end{equation*} |
where \left\Vert {\cdot} \right\Vert_{1-\frac{1}{p}, p} stands for the norm in the real interpolation space \left({X, X_1} \right)_{1-\frac{1}{p}, p}. Now (2.5) is achieved from the embedding [21,Lemma 2.1.(ii)]
\begin{equation*} W^{1,\infty}( {\Omega})\hookrightarrow W^{1,p}( {\Omega})\hookrightarrow W^{2(1-\frac{1}{p}),p}( {\Omega})\cong\left( {X,X_1} \right)_{1-\frac{1}{p},p}, \end{equation*} |
for any p\in(1, 2].
Remark 3. For the purpose of our analysis, we consider only the case p\in(1, 2] in Lemma 2.3. One can refer to [21] for more general cases, in particular p \geqslant3, where the interpolation space \left({X, X_1} \right)_{1-\frac{1}{p}, p} is not equaivalent to W^{2(1-\frac{1}{p}), p}({\Omega}).
Next, we present a compactness theorem called Aubin-Lions Lemma [17,Theorem 2.1] that will be used to give convergence results for the approximated solution (n_ {\epsilon}, c_ {\epsilon}, u_ {\epsilon}).
Lemma 2.4. Let T>0, 1 \leqslant {\alpha}_0, {\alpha}_1 < \infty and X_0, X, X_1 be Banach spaces with X_0\subset X\subset X_1. Suppose further that the embedding X_0\hookrightarrow X is compact and the embedding X\hookrightarrow X_1 is continuous. Let
W = \left\{ {v\in L^{ {\alpha}_0}(0,T ; X_0)\,\, | \,\, {\partial}_t v\in L^{ {\alpha}_1}(0,T ; X_1)} \right\}. |
Then the embedding W\hookrightarrow L^{ {\alpha}_0}(0, T; X) is compact.
The following basic properties of these solutions are well-known.
Lemma 3.1. Let T>0. For each {\epsilon}\in(0, 1), the solution of (2.1) fulfills
\begin{equation} \int\limits_{ {\Omega}}n_ {\epsilon}(x,t) {\,\mathrm{d}} x \leqslant m\quad for\; all \,\, t < T \end{equation} | (3.1) |
and
\begin{equation} \mu\int\limits_{0}^{T}\int\limits_{ {\Omega}}n_ {\epsilon}^q(x,s) {\,\mathrm{d}} x {\,\mathrm{d}} s+ {\epsilon}\int\limits_{0}^{T}\int\limits_{ {\Omega}}n^ {\kappa}_{ {\epsilon}}(x,s) {\,\mathrm{d}} x {\,\mathrm{d}} s \leqslant ( \rho_{+}T+1)m , \end{equation} | (3.2) |
where m = \max\left\{ {{\int\limits_{ {\Omega}}n_0, \left({\frac{\rho_{+}\left| { {\Omega}} \right|}{\mu}} \right)^\frac{1}{q-1}}} \right\} and \rho_+ = \max\left\{ {\rho, 0} \right\}.
Proof. Integrating the first equation in (2.1) over {\Omega}, employing the divergence theorem, and using the Hölder inequality yield that, for all t>0,
\begin{equation} \frac{d{}}{d{t}}\int\limits_{ {\Omega}}n_ {\epsilon} = \rho \int\limits_{\Omega} n_ {\epsilon}-\mu \int\limits_{\Omega} n^q_ {\epsilon} - {\epsilon} \int\limits_{\Omega} n^ {\kappa}_ {\epsilon} \leqslant \rho_{+} \int\limits_{\Omega} n_ {\epsilon}-\frac{\mu}{\left| { {\Omega}} \right|}\left( { \int\limits_{\Omega} n_ {\epsilon}} \right)^q. \end{equation} | (3.3) |
An ODE comparison implies (3.1). Integrating (3.3) with respect to time and then using (3.1), we have
\begin{align*} \mu\int\limits_{0}^{T} \int\limits_{\Omega} n^q_ {\epsilon}+ {\epsilon}\int\limits_{0}^{T} \int\limits_{\Omega} n^ {\kappa}_ {\epsilon} \leqslant\rho_{+}\int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}+ \int\limits_{\Omega} n_0(x) {\,\mathrm{d}} x- \int\limits_{\Omega} n_ {\epsilon}(x, T) {\,\mathrm{d}} x \leqslant (\rho_{+}T+1)m, \end{align*} |
which implies (3.2).
The following estimate is easily obtained by (3.1).
Lemma 3.2. For each {\epsilon}\in(0, 1), we have
\begin{equation} \int\limits_{\Omega} c_ {\epsilon}(x,t) {\,\mathrm{d}} x \leqslant\max\left\{ { \int\limits_{\Omega} c_0,m} \right\} {\quad { for\; all }\; t > 0 }. \end{equation} | (3.4) |
Proof. Integrating the equation for c_ {\epsilon} in (2.1) and using (3.1), we have
\begin{equation*} \frac{d{}}{d{t}} \int\limits_{\Omega} c_ {\epsilon}+ \int\limits_{\Omega} c_ {\epsilon} = \int\limits_{\Omega} n_ {\epsilon} \leqslant m\quad{\mbox{for all}}\;t < T, \end{equation*} |
which yields (3.4) by the ODE comparison.
We recall a useful result shown in [22,Lemma 3.4].
Lemma 3.3. Let T\in(0, \infty], \tau\in(0, T), a>0 and b>0. Suppose that a nonnegative function h\in L^1_{loc}(\mathbb{R}) be such that
\begin{equation*} \int\limits_{t}^{t+\tau}h(s) {\,\mathrm{d}} s \leqslant b\tau \quad\mathit{\text{for all}}\; t\in[0,T-\tau). \end{equation*} |
If a nonnegative function y\in C^0[0, T)\cap C^1(0, T) satisfies
\begin{equation*} y'(t)+ay(t) \leqslant h(t), \end{equation*} |
then
\begin{equation*} y(t) \leqslant y(0)+\frac{b\tau}{1-e^{-a\tau}} {\quad { for\; all }\; t > 0 }. \end{equation*} |
The following lemma is a variant of the result with q = 2 in [22,Lemma 3.6].
Lemma 3.4. Let T>0 and q\in(\frac{5}{3}, 2). Then there exists C>0 such that for any {\epsilon}\in(0, 1) we obtain
\begin{equation} \int\limits_{ {\Omega}}\left| {c_ {\epsilon}(x,t)} \right|^r {\,\mathrm{d}} x \leqslant C\quad for\; all\; t > 0. \end{equation} | (3.5) |
Moreover,
\begin{equation} \int\limits_{0}^{T}\Big( { \int\limits_{\Omega}\left| {c_ {\epsilon}(x,s)} \right|^{3r}} \Big)^\frac{1}{3} {\,\mathrm{d}}{x} {\,\mathrm{d}}{s} \leqslant C(T+1), \end{equation} | (3.6) |
where r = \frac{3q}{5-2q}.
Proof. Multiplying the equation for c_ {\epsilon} in (2.1) by c_ {\epsilon}^{r-1} and integrating over \Omega, we have for all t>0,
\begin{equation} \frac{d{}}{d{t}}\frac{1}{r} \int\limits_{\Omega} c_ {\epsilon}^r+ \frac{4(r-1)}{r^2} \int\limits_{\Omega}\left| { {\nabla } c_ {\epsilon}^{\frac{r}{2}}} \right|^2+ \int\limits_{\Omega} c_ {\epsilon}^r = \int\limits_{\Omega} n_ {\epsilon} c_ {\epsilon}^{r-1} \leqslant\left\Vert {n_ {\epsilon}} \right\Vert_q\left\Vert {c_ {\epsilon}^{r-1}} \right\Vert_\frac{q}{q-1} , \end{equation} | (3.7) |
where the Hölder inequality is used. Using the Gagliardo-Nirenberg inequality and (3.4), we note that
\begin{align*} \left\Vert {c_ {\epsilon}^{r-1}} \right\Vert_\frac{q}{q-1} = \left\Vert {c_ {\epsilon}^\frac{r}{2}} \right\Vert^{\frac{2(r-1)}{r}}_{\frac{2(r-1)}{r}\frac{q}{q-1}}& \leqslant C\left( {\left\Vert { {\nabla } c_ {\epsilon}^\frac{r}{2}} \right\Vert^{\frac{2(r-1)}{r}\theta}_2\left\Vert {c_ {\epsilon}^\frac{r}{2}} \right\Vert^{\frac{2(r-1)}{r}(1-\theta)}_2+\left\Vert {c_ {\epsilon}^\frac{r}{2}} \right\Vert^\frac{2(r-1)}{r}_\frac{2}{r}} \right)\\ & \leqslant C\left\Vert { {\nabla } c_ {\epsilon}^\frac{r}{2}} \right\Vert^{\frac{2(r-1)}{r}\theta}_2\left\Vert {c_ {\epsilon}^\frac{r}{2}} \right\Vert^{\frac{2(r-1)}{r}(1-\theta)}_2+C {\quad {\rm{ for \;all }}\; t > 0 }, \end{align*} |
where \theta = \frac{3}{2}(1-\frac{r}{r-1}\frac{g-1}{q})\in(0, 1) since r = \frac{3q}{5-2q}. Employing Young's inequality, we have
\begin{align} \left\Vert {n_ {\epsilon}} \right\Vert_q\left\Vert {c_ {\epsilon}^{r-1}} \right\Vert_\frac{q}{q-1} \leqslant& \frac{2(r-1)}{r^2}\left\Vert { {\nabla } c_ {\epsilon}^\frac{r}{2}} \right\Vert^{2}_2+C\left\Vert {n_ {\epsilon}} \right\Vert^q_q\left\Vert {c_ {\epsilon}^\frac{r}{2}} \right\Vert^\frac{4(q-1)}{3}_2+\left\Vert {n_ {\epsilon}} \right\Vert^q_q+C\\ \leqslant&\frac{2(r-1)}{r^2}\left\Vert { {\nabla } c_ {\epsilon}^\frac{r}{2}} \right\Vert^{2}_2+C\left\Vert {n_ {\epsilon}} \right\Vert^q_q\left( {\left\Vert {c_ {\epsilon}^\frac{r}{2}} \right\Vert^2_2+1} \right)+\left\Vert {n_ {\epsilon}} \right\Vert^q_q+C. \end{align} | (3.8) |
Combining (3.7) with (3.8) implies that there exist C_5>0 such that
\begin{equation} \frac{d{}}{d{t}}\frac{1}{r} \int\limits_{\Omega} c_ {\epsilon}^r+ \frac{2(r-1)}{r^2} \int\limits_{\Omega}\left| { {\nabla } c_ {\epsilon}^{\frac{r}{2}}} \right|^2+ \int\limits_{\Omega} c_ {\epsilon}^r+1 \leqslant C\left( {\left\Vert {n_ {\epsilon}} \right\Vert^q_q+1} \right)\left( {\left\Vert {c_ {\epsilon}} \right\Vert^r_r+1} \right). \end{equation} | (3.9) |
Let y(t): = \left\Vert {c_ {\epsilon}(t)} \right\Vert^r_r+1 and h(t): = \left\Vert {n_ {\epsilon}(t)} \right\Vert^q_q+1, which is in L^1 locally in time. Then, dividing (3.9) by y(t) yields that
\begin{equation} \frac{d{}}{d{t}}\ln{y}+\frac{2(r-1)}{r}\frac{1}{y}\left\Vert { {\nabla } c_ {\epsilon}^\frac{r}{2}} \right\Vert^2_2+1 \leqslant C h. \end{equation} | (3.10) |
We use again the Gagliardo-Nirenberg inequality to obtain that for all t>0
\begin{align*} y(t) \leqslant C\left\Vert { {\nabla } c_ {\epsilon}^\frac{r}{2}} \right\Vert^{\frac{6(r-1)}{3r-1}}_2\left\Vert {c_ {\epsilon}^\frac{r}{2}} \right\Vert^\frac{4}{3r-1}_\frac{2}{r}+C\left\Vert {c_ {\epsilon}^\frac{r}{2}} \right\Vert^2_\frac{2}{r}+1 \leqslant C \left( {\left\Vert { {\nabla } c_ {\epsilon}^\frac{r}{2}} \right\Vert^{\frac{6(r-1)}{3r-1}}_2+1} \right), \end{align*} |
which leads that \left\Vert { {\nabla } c_ {\epsilon}^\frac{r}{2}} \right\Vert_2^2 \geqslant\left({\frac{1}{C}y(t)-1} \right)^\frac{3r-1}{3(r-1)} \geqslant Cy^\frac{3r-1}{3(r-1)}-1. Hence, it follows that
\begin{equation} \frac{1}{y}\left\Vert { {\nabla } c_ {\epsilon}^\frac{r}{2}} \right\Vert^2_2 \geqslant Cy^\frac{2}{3(r-1)}-\frac{1}{y} \geqslant C\ln y-1 {\quad {\rm{ for\; all }}\; t > 0 }, \end{equation} | (3.11) |
where we use the trivial inequality \ln{y}\lesssim y^k for k>0. Putting the above inequality (3.11) into (3.10), we have
\begin{equation*} \frac{d{}}{d{t}}\ln{y}+C\ln{y} \leqslant h. \end{equation*} |
By Lemma 3.3, we can conclude that there exists C>0 satisfying y(t) \leqslant C for all t>0 which proves (3.5) as required. Integrating (3.10) with respect to time and exploiting the boundedness of y(t), guaranteed by (3.5), yield that
\begin{equation*} \int\limits_{0}^{T} \left\Vert { {\nabla } c_ {\epsilon}^\frac{r}{2}} \right\Vert_2^2 \leqslant C(1+T) \end{equation*} |
for some C>0. Using (2.2) and (3.4), we finally have (3.6).
We adopt well-known energy estimate for the Navier-Stokes system to gain a bound for u_ {\epsilon} in energy class.
Lemma 3.5. Let T>0 and q\in\left({\frac{5}{3}, 2} \right). Then there exists C>0 such that for each {\epsilon}\in(0, 1), we have
\begin{equation} \int\limits_{ {\Omega}}\left| {u_ {\epsilon}(x,t)} \right|^2 {\,\mathrm{d}} x \leqslant C\quad for \; all \;t > 0 \end{equation} | (3.12) |
and
\begin{equation} \int\limits_{0}^{T}\int\limits_{ {\Omega}}\left| { {\nabla } u_ {\epsilon}(x,s)} \right|^2 {\,\mathrm{d}} x {\,\mathrm{d}} s \leqslant C(1+T). \end{equation} | (3.13) |
Proof. We test the fluid equation in (2.1) by u_ {\epsilon} to find the following L^2 estimate
\begin{equation} \frac{d{}}{d{t}} \int\limits_{\Omega} u_ {\epsilon}^2+ \int\limits_{\Omega} \left| { {\nabla } u_ {\epsilon}} \right|^2 = \int\limits_{\Omega} n_ {\epsilon} u_ {\epsilon} {\nabla } \phi \end{equation} | (3.14) |
We can estimate the right hand side of (3.14) using the H \ddot{\rm o} lder inequality, the Sobolev embedding W^{1, 2}_{0, \sigma}\hookrightarrow L^6, and the interpolation inequality for n_ {\epsilon} that
\begin{align} \int\limits_{\Omega} n_ {\epsilon} u_ {\epsilon} {\nabla } \phi& \leqslant C\left\Vert {n_ {\epsilon}} \right\Vert_\frac{6}{5}\left\Vert {u_ {\epsilon}} \right\Vert_6\\ & \leqslant C\left\Vert {n_ {\epsilon}} \right\Vert_\frac{6}{5}^2+\frac{1}{2}\left\Vert { {\nabla } u_ {\epsilon}} \right\Vert_2^2\\ & \leqslant C\left\Vert {n_ {\epsilon}} \right\Vert^{\frac{q}{3(q-1)}}_q\left\Vert {n_ {\epsilon}} \right\Vert^\frac{5q-6}{3(q-1)}_1+\frac{1}{2}\left\Vert { {\nabla } u_ {\epsilon}} \right\Vert^2_2\\ & \leqslant C\left( {\left\Vert {n_ {\epsilon}} \right\Vert^q_q+1} \right)+\frac{1}{2}\left\Vert { {\nabla } u_ {\epsilon}} \right\Vert^2_2 {\quad {\rm{ for \; all }}\; t > 0 } , \end{align} | (3.15) |
where we used that \frac{q}{3(q-1)} \leqslant q.
Thus, with the aid of (3.15) and the Poincaré inequality, we have for some C
\begin{equation*} \frac{d{}}{d{t}} \int\limits_{\Omega} u_ {\epsilon}^2+C \int\limits_{\Omega} u_ {\epsilon}^2 \leqslant C \left\Vert {n_ {\epsilon}} \right\Vert^q_q+1. \end{equation*} |
(3.12) is proved if we use (3.2) and Lemma 3.3, and then (3.13) can be calculated by integrating (3.14) with respect to time and using (3.15).
A direct consequence of Lemma 3.5 is the following.
Corollary 1. Let T>0 and \frac{3}{ {\alpha}}+\frac{2}{ {\beta}} = \frac{3}{2}, 2 \leqslant {\alpha} \leqslant6. Then
\begin{equation} \int\limits_{0}^{T}\Big( { \int\limits_{\Omega}\left| {u_ {\epsilon}(x,s)} \right|^ {\alpha}} \Big)^\frac{ {\beta}}{ {\alpha}} {\,\mathrm{d}}{x} {\,\mathrm{d}}{s} \leqslant C(1+T), \end{equation} | (3.16) |
in particular, if {\alpha} = {\beta} = \frac{10}{3}, then
\begin{equation} \int\limits_{0}^{T}\int\limits_{ {\Omega}}\left| {u_ {\epsilon}(x,s)} \right|^\frac{10}{3} {\,\mathrm{d}} x {\,\mathrm{d}} s \leqslant C(1+T). \end{equation} | (3.17) |
Proof. In view of Lemma 3.5, (3.16), in particular (3.17), is derived from the Gagliardo-Nirenberg inequality (2.2).
Since u_{\epsilon} only belong to energy class, we have lower regularity of {\nabla } c_{ {\epsilon}} , due to difficulties of controlling convective term u\cdot {\nabla } c, than the case that the Stokes sysem is coupled. Nevertheless, using the divergence free condtion, we obtain a certain integrability of {\nabla } c_{ {\epsilon}} by the following decompsition, which makes computations easier. More precisely, let w_ {\epsilon} be a solution satisfying
\begin{equation*} \begin{cases} \partial_t w_ {\epsilon}-\Delta w_ {\epsilon} = -c_ {\epsilon}+n_ {\epsilon},\quad (x,t)\in\Omega\times [0,t),\\ w_ {\epsilon}(x,0) = c_0, \quad\quad\quad\quad\quad\,\,\,\, x\in\Omega. \end{cases} \end{equation*} |
Now we set \tilde{w}_ {\epsilon}: = c_ {\epsilon}-w_ {\epsilon}. Then, due to the divergence free condition for u_ {\epsilon}, it follows that \tilde{w}_ {\epsilon} solves
\begin{equation*} \begin{cases} \partial_t \tilde{w}_ {\epsilon}-\Delta \tilde{w}_ {\epsilon} = - {{\nabla }\cdot}(u_ {\epsilon} c_ {\epsilon}),\quad (x,t)\in\Omega\times [0,t),\\ \tilde{w}_ {\epsilon}(x,0) = 0, \quad\quad\quad\quad\quad\quad\,\,\,\, x\in\Omega. \end{cases} \end{equation*} |
In next lemma, estimating each solutions of the decompsition, we show that \nabla c_{\epsilon}\in L^{10q/(10-q)}(\Omega\times (0, T)) .
Lemma 3.6. Let T>0 and q\in\left({\frac{5}{3}, 2} \right). Then given {\epsilon}\in(0, 1), there exists C = C(T)>0 such that
\begin{equation} \int\limits_{0}^{T} \int\limits_{\Omega} \left| { {\nabla } c_ {\epsilon}(x,s)} \right|^m {\,\mathrm{d}}{x} {\,\mathrm{d}}{s} \leqslant C, \end{equation} | (3.18) |
where m = \frac{10q}{10-q}.
Proof. We first observe reularity of w_ {\epsilon}. On account of (2.5), we can find a constant C = C(T)>0 satisfying
\begin{align} \int\limits_{0}^{T} \left\Vert {\Delta w_ {\epsilon}} \right\Vert^q_q& \leqslant C\int\limits_{0}^{T}\left( {\left\Vert {n_ {\epsilon}} \right\Vert^q_q+\left\Vert {c_ {\epsilon}} \right\Vert^q_q+1} \right)\\ & \leqslant C\left( {\left( {\sup\limits_{t > 0}\left\Vert {c_ {\epsilon}} \right\Vert_r} \right)^q+\int\limits_{0}^{T}\left\Vert {n_ {\epsilon}} \right\Vert^q_q+1} \right). \end{align} | (3.19) |
Then the Gagliardo-Nirenberg interpolation inequality (2.2) and (3.5) yield that
\begin{align} \int\limits_{0}^{T}\left\Vert { {\nabla } w_ {\epsilon}} \right\Vert^\frac{5q}{5-q}_\frac{5q}{5-q}& \leqslant C\int\limits_{0}^{T}\left( {\left\Vert {\Delta w_ {\epsilon}} \right\Vert^{\frac{5q}{5-q}\left( {1-\frac{q}{5}} \right)}_q\left\Vert {w_ {\epsilon}} \right\Vert^{\frac{5q}{5-q}\cdot\frac{q}{5}}_\frac{3q}{5-2q}+\left\Vert {w_ {\epsilon}} \right\Vert^\frac{5q}{5-q}_\frac{3q}{5-2q}} \right)\\ & \leqslant C\left( {\int\limits_{0}^{T}\left\Vert {\Delta w_ {\epsilon}} \right\Vert^q_q+1} \right). \end{align} | (3.20) |
Thus, from (3.19) and (3.20) we see that for some C = C(T)>0
\begin{equation*} \int\limits_{0}^{T}\left\Vert { {\nabla } w_ {\epsilon}} \right\Vert^\frac{5q}{5-q}_\frac{5q}{5-q} \leqslant C\left( {\int\limits_{0}^{T}\left\Vert {n_ {\epsilon}} \right\Vert^q_q+\left( {\sup\limits_{t > 0}\left\Vert {c_ {\epsilon}} \right\Vert_r} \right)^q+1} \right). \end{equation*} |
The last term is finite because of (3.2), (3.5) and the fact that q \leqslant r = \frac{3q}{5-2q}. Next, let {\alpha} and {\beta} be in Lemma 3.5 with {\alpha} = \frac{90q}{11q+40} and {\beta} = \frac{30q}{17q-20}. It can be easily checked that 2 < {\alpha} < 6 and 2 < {\beta} because q\in\left({\frac{5}{3}, 2} \right). Then we can see via the maximal estimate (2.3) and the Hölder inequality that
\begin{align} \int\limits_{0}^{T}\left\Vert { {\nabla } \tilde{w}} \right\Vert^m_m \leqslant C_T\int\limits_{0}^{T}\left\Vert {u_ {\epsilon} c_ {\epsilon}} \right\Vert^m_m \leqslant C\left( {\int\limits_{0}^{T}\left\Vert {u_ {\epsilon}} \right\Vert^ {\beta}_ {\alpha}} \right)^\frac{m}{ {\beta}}\left( {\int\limits_{0}^{T}\left\Vert {c_ {\epsilon}} \right\Vert_{3r}^r} \right)^\frac{m}{r} \end{align} | (3.21) |
which is valid since \frac{1}{m} = \frac{1}{ {\alpha}}+\frac{1}{3r} = \frac{1}{ {\beta}}+\frac{1}{r}, where r = \frac{3q}{5-2q}. The last term in (3.21) is finite due to (3.16) and (3.6). Hence, we have
\begin{equation*} \int\limits_{0}^{T}\left\Vert { {\nabla } c_ {\epsilon}} \right\Vert_m^m \leqslant \int\limits_{0}^{T}\left\Vert { {\nabla } w_ {\epsilon}} \right\Vert^m_m+\int\limits_{0}^{T}\left\Vert { {\nabla } \tilde{w}_ {\epsilon}} \right\Vert^m_m, \end{equation*} |
which is finite since m < \frac{5q}{5-q} and (3.21). Then (3.18) is proved.
Taking advantage of Lemma 3.6, we can obtain the maximal estimate for c_ {\epsilon}.
Lemma 3.7. Let T>0 and q\in\left({\frac{5}{3}, 2} \right). Then there exists C = C(T)>0 such that for any {\epsilon}>0,
\begin{equation} \int\limits_{0}^{T}\left\Vert { {\partial}_t c_ {\epsilon}} \right\Vert_\frac{5q}{5+q}^\frac{5q}{5+q}+\int\limits_{0}^{T}\left\Vert { {\Delta} c_ {\epsilon}} \right\Vert_\frac{5q}{5+q}^\frac{5q}{5+q} \leqslant C. \end{equation} | (3.22) |
Proof. Applying (2.5), we obtain
\begin{equation*} \int\limits_{0}^{T}\left\Vert { {\partial}_t c_ {\epsilon}} \right\Vert_\frac{5q}{5+q}^\frac{5q}{5+q}+\int\limits_{0}^{T}\left\Vert { {\Delta} c_ {\epsilon}} \right\Vert_\frac{5q}{5+q}^\frac{5q}{5+q} \leqslant C\left( {\int\limits_{0}^{T}\left\Vert {c_ {\epsilon}} \right\Vert_\frac{5q}{5+q}^\frac{5q}{5+q}+\int\limits_{0}^{T}\left\Vert {n_ {\epsilon}} \right\Vert_\frac{5q}{5+q}^\frac{5q}{5+q}+\int\limits_{0}^{T}\left\Vert {u_ {\epsilon} {\nabla } c_ {\epsilon}} \right\Vert_\frac{5q}{5+q}^\frac{5q}{5+q}+1} \right) \end{equation*} |
\begin{equation*} \leqslant C\left( {\left( {\sup\limits_{t > 0}\left\Vert {c_ {\epsilon}} \right\Vert_r} \right)^\frac{5q}{5+q}+\int\limits_{0}^{T} \left\Vert {n_ {\epsilon}} \right\Vert_q^q+\int\limits_{0}^{T}\left\Vert {u_ {\epsilon}} \right\Vert_\frac{10}{3}^\frac{5q}{5+q}\left\Vert { {\nabla } c_ {\epsilon}} \right\Vert_m^\frac{5q}{5+q}+1} \right) \end{equation*} |
\begin{equation*} \leqslant C\left( {\left( {\sup\limits_{t > 0}\left\Vert {c_ {\epsilon}} \right\Vert_r} \right)^\frac{5q}{5+q}+\int\limits_{0}^{T} \left\Vert {n_ {\epsilon}} \right\Vert_q^q+\int\limits_{0}^{T}\left\Vert {u_ {\epsilon}} \right\Vert_\frac{10}{3}^\frac{10}{3}+\int\limits_{0}^{T}\left\Vert { {\nabla } c_ {\epsilon}} \right\Vert_m^m+1} \right) < C, \end{equation*} |
due to (3.2), (3.5), (3.17) and (3.18). This proves (3.22).
The following two lemmas are crucial to achieving the convergence property for n_ {\epsilon}.
Lemma 3.8. Let T>0 and q\in\left({\frac{5}{3}, 2} \right). Then for any {\gamma}\in(0, 1) with {\gamma} \leqslant\frac{4q-5}{5}, there exists C = C(T)>0 satisfying
\begin{equation} \int\limits_{0}^{T} \int\limits_{\Omega} \left| { {\nabla } (n_ {\epsilon}+1)^\frac{ {\gamma}}{2}(x,s)} \right|^2 {\,\mathrm{d}}{x} {\,\mathrm{d}}{s} \leqslant C. \end{equation} | (3.23) |
Proof. Testing the first equation in (2.1) by {\gamma} n_ {\epsilon}^{ {\gamma}-1} and using integration by parts, we obtain
\begin{align} \frac{4(1- {\gamma})}{ {\gamma}}\int\limits_{0}^{T} \int\limits_{\Omega} \left| { {\nabla } n_ {\epsilon}^\frac{ {\gamma}}{2}} \right|^2 = & \int\limits_{\Omega} n_ {\epsilon}^ {\gamma}(\cdot,T)- \int\limits_{\Omega} n_0^ {\gamma}-(1- {\gamma})\int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^ {\gamma} {\Delta} c_ {\epsilon}\\ &-\rho {\gamma}\int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^ {\gamma}+\mu {\gamma}\int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^{ {\gamma}+q-1}+ {\epsilon} {\gamma}\int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^{ {\kappa}+ {\gamma}-1}. \end{align} | (3.24) |
Using Young's inequality and (3.2), we have
\begin{equation*} \int\limits_{\Omega} n_ {\epsilon}^ {\gamma}(\cdot,T)- \int\limits_{\Omega} n_0^ {\gamma} \leqslant C\left( { \int\limits_{\Omega} n_ {\epsilon}+1 } \right) < C, \end{equation*} |
and
\begin{align} -\rho {\gamma}\int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^ {\gamma}+\mu {\gamma}\int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^{ {\gamma}+q-1}+ {\epsilon} {\gamma}\int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^{ {\kappa}+ {\gamma}-1}\\ \leqslant C\left( {\mu\int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^q+ {\epsilon}\int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^ {\kappa}+1} \right) < C. \end{align} | (3.25) |
Since 0 < {\gamma} \leqslant\frac{4q-5}{5}, we see that \frac{5+q}{5q}+\frac{ {\gamma}}{q} \leqslant1. This leads
\begin{align} (1- {\gamma})\int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^ {\gamma} {\Delta} c_ {\epsilon} \leqslant \int\limits_{0}^{T}\left\Vert {n_ {\epsilon}} \right\Vert_q^ {\gamma}\left\Vert { {\Delta} c_ {\epsilon}} \right\Vert_\frac{5q}{5+q}\\ \leqslant C\left( {\int\limits_{0}^{T}\left\Vert {n_ {\epsilon}} \right\Vert^q_q+\int\limits_{0}^{T}\left\Vert { {\Delta} c_ {\epsilon}} \right\Vert_\frac{5q}{5+q}^ \frac{5q}{5+q}+1} \right) < C . \end{align} | (3.26) |
Collecting (3.24), (3.25) and (3.26), we obtain
\begin{equation} \int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^{ {\gamma}-2}\left| { {\nabla } n_ {\epsilon}} \right|^2 = \frac{4}{ {\gamma}^2}\int\limits_{0}^{T} \int\limits_{\Omega}\left| { {\nabla } n_ {\epsilon}^\frac{ {\gamma}}{2}} \right|^2 \leqslant C. \end{equation} | (3.27) |
Since {\gamma}-2 < 0, we get (n_ {\epsilon}+1)^{ {\gamma}-2} \leqslant n_ {\epsilon}^{ {\gamma}-2}, hence (3.23).
In the following lemma, we mean by (W^{k, 2}_0)^* the dual space of W^{k, 2}_0.
Lemma 3.9. Let T>0 and q\in\left({\frac{5}{3}, 2} \right). Then for any {\gamma}\in(0, 1) with {\gamma} \leqslant\frac{4q-5}{5}, there exists k\in\mathbb{N} and C = C(T)>0, independent of {\epsilon}, satisfying
\begin{equation*} \left\Vert { {\partial}_t(1+n_ {\epsilon})^\frac{ {\gamma}}{2}} \right\Vert_{L^1(0,T;(W^{k,2}_0(\Omega))^*)} \leqslant C. \end{equation*} |
Proof. Fix k\in\mathbb{N} to be choosen later and let {\varphi}\in W^{k, 2}_0(\Omega) be a test function. We observe that
\begin{align*} \frac{2}{ {\gamma}} \int\limits_{\Omega} & {\partial}_t(n_ {\epsilon}+1)^\frac{ {\gamma}}{2} {\varphi} = \int\limits_{\Omega} (1+n_ {\epsilon})^{\frac{ {\gamma}}{2}-1} {\partial}_t n_ {\epsilon} {\varphi}\\ & = \int\limits_{\Omega} (1+n_ {\epsilon})^{\frac{ {\gamma}}{2}-1}\left( {\Delta n_ {\epsilon}-u_ {\epsilon}\cdot {\nabla } n_ {\epsilon}- {{\nabla }\cdot}(n_ {\epsilon} {\nabla } c_ {\epsilon})+\rho n_ {\epsilon}-\mu n_ {\epsilon}^q- {\epsilon} n_ {\epsilon}^ {\kappa}} \right) {\varphi} = :\sum\limits_{i = 1}^6 J_i. \end{align*} |
First, employing integration by parts and H \ddot{\rm o} lder inequality, we can estimate J_1 as follows:
\begin{align} \left| {J_1} \right|& \leqslant C \int\limits_{\Omega} (1+n_ {\epsilon})^{\frac{ {\gamma}}{2}-2}\left| { {\nabla } n_ {\epsilon}} \right|^2\left| { {\varphi}} \right|+ C \int\limits_{\Omega} (1+n_ {\epsilon})^{\frac{ {\gamma}}{2}-1}\left| { {\nabla } n_ {\epsilon}} \right|\left| { {\nabla } {\varphi}} \right|\\ & \leqslant C\left\Vert { {\varphi}} \right\Vert_\infty\left\Vert { {\nabla } n_ {\epsilon}^\frac{ {\gamma}}{2}} \right\Vert^2_2 +C\left\Vert { {\nabla } {\varphi}} \right\Vert_2\left( {1+\left\Vert { {\nabla } n_ {\epsilon}^\frac{ {\gamma}}{2}} \right\Vert^2_2} \right), \end{align} | (3.28) |
where we used the fact that (1+n_ {\epsilon})^{\frac{ {\gamma}}{2}-2} \leqslant (1+n_ {\epsilon})^{ {\gamma}-2} \leqslant n_ {\epsilon}^{ {\gamma}-2}. Similarly, the second and third terms are controlled as follows:
\begin{align} &\left| {J_2} \right| \leqslant C \int\limits_{\Omega}(1+n_ {\epsilon})^{\frac{ {\gamma}}{2}-2} n_ {\epsilon}^{2-\frac{ {\gamma}}{2}} \left| { {\nabla }{n_ {\epsilon}^\frac{ {\gamma}}{2}}} \right|\left| {u_ {\epsilon}} \right|\left| { {\varphi}} \right|+C \int\limits_{\Omega}(1+n_ {\epsilon})^{\frac{ {\gamma}}{2}-1}\left| {n_ {\epsilon}} \right|\left| {u_ {\epsilon}} \right|\left| { {\nabla } {\varphi}} \right| \\ & \leqslant C\left\Vert { {\nabla } n_ {\epsilon}^\frac{ {\gamma}}{2}} \right\Vert_2\left\Vert {u_ {\epsilon}} \right\Vert_\frac{10}{3}\left\Vert { {\varphi}} \right\Vert_5+C\left\Vert {1+n_ {\epsilon}} \right\Vert_q^\frac{ {\gamma}}{2}\left\Vert {u_ {\epsilon}} \right\Vert_\frac{10}{3}\left\Vert { {\nabla } {\varphi}} \right\Vert_\frac{10q}{7q-5 {\gamma}} \\ & \leqslant C\left( {\left\Vert { {\nabla } n_ {\epsilon}^\frac{ {\gamma}}{2}} \right\Vert_2^\frac{10}{7}+C\left\Vert {u_ {\epsilon}} \right\Vert_\frac{10}{3}^\frac{10}{3}} \right)\left\Vert { {\varphi}} \right\Vert_5+C\left( {\left\Vert {1+n_ {\epsilon}} \right\Vert_q^\frac{5 {\gamma}}{7}+\left\Vert {u_ {\epsilon}} \right\Vert_\frac{10}{3}^\frac{10}{3}} \right)\left\Vert { {\nabla } {\varphi}} \right\Vert_\frac{10q}{3q+5}\\ & \leqslant C\left( {\left\Vert { {\nabla } n_ {\epsilon}^\frac{ {\gamma}}{2}} \right\Vert_2^2+C\left\Vert {u_ {\epsilon}} \right\Vert_\frac{10}{3}^\frac{10}{3}+1} \right)\left\Vert { {\varphi}} \right\Vert_5+C\left( {\left\Vert {n_ {\epsilon}} \right\Vert_q^q+\left\Vert {u_ {\epsilon}} \right\Vert_\frac{10}{3}^\frac{10}{3}+1} \right)\left\Vert { {\nabla } {\varphi}} \right\Vert_\frac{10q}{3q+5} \end{align} | (3.29) |
because {\gamma} < 1 < \frac{7q}{5} and \frac{10q}{7q-5 {\gamma}} \leqslant\frac{10q}{3q+5}.
\begin{align} \left| {J_3} \right|& \leqslant C \int\limits_{\Omega}(1+n_ {\epsilon})^{\frac{ {\gamma}}{2}-2} n_ {\epsilon}^{2-\frac{ {\gamma}}{2}} \left| { {\nabla }{n_ {\epsilon}^\frac{ {\gamma}}{2}}} \right|\left| { {\nabla } c_ {\epsilon}} \right|\left| { {\varphi}} \right|+C \int\limits_{\Omega}(1+n_ {\epsilon})^{\frac{ {\gamma}}{2}-1}\left| {n_ {\epsilon}} \right|\left| { {\nabla } c_ {\epsilon}} \right|\left| { {\nabla } {\varphi}} \right|\\ & \leqslant C\left\Vert { {\nabla } n_ {\epsilon}^\frac{ {\gamma}}{2}} \right\Vert_2\left\Vert { {\nabla } c_ {\epsilon}} \right\Vert_q\left\Vert { {\varphi}} \right\Vert_\frac{2q}{2-q}+C\left\Vert {1+n_ {\epsilon}} \right\Vert^\frac{ {\gamma}}{2}_q\left\Vert { {\nabla } c_ {\epsilon}} \right\Vert_q\left\Vert { {\nabla } {\varphi}} \right\Vert_\frac{2q}{2q-2- {\gamma}}\\ & \leqslant C\left( {\left\Vert { {\nabla } n_ {\epsilon}^\frac{ {\gamma}}{2}} \right\Vert_2^2+\left\Vert { {\nabla } c_ {\epsilon}} \right\Vert_m^m+1} \right)\left\Vert { {\varphi}} \right\Vert_\frac{2q}{2-q}\\&\quad+C\left( {\left\Vert {n_ {\epsilon}} \right\Vert^q_q+\left\Vert { {\nabla } c_ {\epsilon}} \right\Vert_m^m+1} \right)\left\Vert { {\nabla } {\varphi}} \right\Vert_\frac{2q}{2q-2- {\gamma}}, \end{align} | (3.30) |
where we used the fact that q < m and {\gamma} \leqslant\frac{4q-5}{5} < 2q-2. Estimates for J_4, J_5 and J_6 can be easily obtained by the following calculation
\begin{align} \left| {J_4} \right|& \leqslant \int\limits_{\Omega}(1+n_ {\epsilon})^\frac{ {\gamma}}{2}\left| { {\varphi}} \right| \leqslant C\left( {\left\Vert {n_ {\epsilon}} \right\Vert_q^q+1} \right)\left\Vert { {\varphi}} \right\Vert_\infty, \end{align} | (3.31) |
\begin{align} \left| {J_5} \right|& \leqslant \int\limits_{\Omega}(1+n_ {\epsilon})^{\frac{ {\gamma}}{2}+q-1}\left| { {\varphi}} \right| \leqslant C\left( {\left\Vert {n_ {\epsilon}} \right\Vert^q_q+1} \right)\left\Vert { {\varphi}} \right\Vert_\infty , \end{align} | (3.32) |
\begin{align} \left| {J_6} \right|& \leqslant {\epsilon} \int\limits_{\Omega}(1+n_ {\epsilon})^{\frac{ {\gamma}}{2}+ {\kappa}-1}\left| { {\varphi}} \right| \leqslant C\left( { {\epsilon}\left\Vert {n_ {\epsilon}} \right\Vert^ {\kappa}_ {\kappa}+1} \right)\left\Vert { {\varphi}} \right\Vert_\infty . \end{align} | (3.33) |
Collecting all of estimates (3.28)-(3.33) and applying the Sobolev embedding theorem, we have
\begin{align} \left| { \int\limits_{\Omega} {\partial}_t(1+n_ {\epsilon})^\frac{ {\gamma}}{2} {\varphi}} \right| \leqslant C&\left( {\left\Vert { {\nabla } n_ {\epsilon}^\frac{ {\gamma}}{2}} \right\Vert_2^2+\left\Vert {u_ {\epsilon}} \right\Vert^\frac{10}{3}_\frac{10}{3}+\left\Vert { {\nabla } c_ {\epsilon}} \right\Vert^m_m+\left\Vert {n_ {\epsilon}} \right\Vert^q_q+ {\epsilon}\left\Vert {n_ {\epsilon}} \right\Vert^ {\kappa}_ {\kappa}+1} \right)\\ &\times\left\Vert { {\varphi}} \right\Vert_{W^{1,\infty}_0(\Omega)}. \end{align} | (3.34) |
Choose k sufficiently large that k>\frac{5}{2}. Then W^{k, 2}_0(\Omega) is embedded into W^{1, \infty}(\Omega) by Sobolev embedding. Finally, integration of (3.34) over (0, T) leads, with the help of (3.1), (3.2), (3.18), (3.16) and (3.23), that
\begin{equation*} \left\Vert { {\partial}_t(1+n_ {\epsilon})^\frac{ {\gamma}}{2}} \right\Vert_{L^1(0,T;(W^{k,2}_0(\Omega))^*)} \leqslant C, \end{equation*} |
as desired.
The estimate for the time derivative of u_ {\epsilon} is obtained by the simple calculation.
Lemma 3.10. Let T>0. Then there exists C>0 such that for any {\epsilon}>0,
\begin{equation} \left\Vert { {\partial}_tu_ {\epsilon}} \right\Vert_{L^1(0,T;(W^{1,5}_{0,\sigma}(\Omega))^*)} \leqslant C(1+T). \end{equation} | (3.35) |
Proof. Given {\varphi}\in C^\infty_0({\Omega}\times [0, \infty); \mathbb{R}^3) with {{\nabla }\cdot} {\varphi} = 0, we compute
\begin{align} \left| { \int\limits_{\Omega} {\partial}_t u_ {\epsilon} {\varphi}} \right|& = \left| {- \int\limits_{\Omega} {\nabla } u_ {\epsilon}\cdot {\nabla } {\varphi}- \int\limits_{\Omega} (Y_ {\epsilon} u_ {\epsilon}\otimes u_ {\epsilon}) {\nabla } {\varphi}+ \int\limits_{\Omega} n_ {\epsilon} {\nabla }\phi {\varphi}} \right|\\ & \leqslant\left\Vert { {\nabla } u_ {\epsilon}} \right\Vert_2\left\Vert { {\nabla } {\varphi}} \right\Vert_2+\left\Vert {Y_ {\epsilon} u_ {\epsilon}\otimes u_ {\epsilon}} \right\Vert_\frac{5}{4}\left\Vert { {\nabla } {\varphi}} \right\Vert_5+\left\Vert {n_ {\epsilon}} \right\Vert_q\left\Vert { {\varphi}} \right\Vert_\frac{q}{q-1}\left\Vert {\nabla\phi} \right\Vert_{\infty}\\ & \leqslant\left( {\left\Vert { {\nabla } u_ {\epsilon}} \right\Vert_2^2+1} \right)\left\Vert { {\nabla } {\varphi}} \right\Vert_2+C\left( {\left\Vert {Y_ {\epsilon} u_ {\epsilon}} \right\Vert_2^2+\left\Vert {u_ {\epsilon}} \right\Vert_\frac{10}{3}^\frac{10}{3}+1} \right)\left\Vert { {\nabla } {\varphi}} \right\Vert_5\\ &\quad+C\left( {\left\Vert {n_ {\epsilon}} \right\Vert_q^q+1} \right)\left\Vert { {\varphi}} \right\Vert_\infty\\ & \leqslant C\left( {\left\Vert { {\nabla } u_ {\epsilon}} \right\Vert^2_2+\left\Vert {u_ {\epsilon}} \right\Vert^\frac{10}{3}_\frac{10}{3}+\left\Vert {n_ {\epsilon}} \right\Vert^q_q+1} \right)\left\Vert { {\varphi}} \right\Vert_{W^{1,5}_0( {\Omega})} . \end{align} | (3.36) |
Here we used the well-known inequality \left\Vert {Y_ {\epsilon} u_ {\epsilon}} \right\Vert_2^2 \leqslant C\left\Vert {u_ {\epsilon}} \right\Vert_2^2. Thus, integrating (3.36) over (0, T) yields (3.35).
We are now ready to prove the convergence property for (n_ {\epsilon}, c_ {\epsilon}, u_ {\epsilon}).
Lemma 4.1. Let q\in(\frac{5}{3}, 2), {\gamma}\in(0, 1) with {\gamma} \leqslant\frac{4q-5}{5} and p\in(1, q). A number m is given in Lemma 3.6. Then the classical solution (n_ {\epsilon}, c_ {\epsilon}, u_ {\epsilon}) of (2.1) satisfies the following convergence property.
\begin{gather} n_ {\epsilon}\rightarrow n\quad {\text{a.e. in}}\;\Omega\times\left( {0,\infty} \right), \end{gather} | (4.1) |
\begin{gather} n_ {\epsilon}\rightharpoonup n\quad {\text{in}}\;L^q_{loc}(\overline{\Omega}\times[0,\infty)), \end{gather} | (4.2) |
\begin{gather} n_ {\epsilon}\rightarrow n\quad {\text{in}}\;L^p_{loc}(\overline{\Omega}\times[0,\infty)), \end{gather} | (4.3) |
\begin{gather} n_ {\epsilon}^\frac{ {\gamma}}{2}\rightharpoonup n^\frac{ {\gamma}}{2} \quad {\text{in}}\; L^2_{loc}([0,\infty);W^{1,2}(\Omega)), \end{gather} | (4.4) |
\begin{gather} c_ {\epsilon}\rightarrow c\quad {\text{a.e. in}}\;\Omega\times\left( {0,\infty} \right), \end{gather} | (4.5) |
\begin{gather} c_ {\epsilon}\rightharpoonup c\quad {\text{in}}\; L^m_{loc}([0,\infty);W^{1,m}(\Omega)), \end{gather} | (4.6) |
\begin{gather} {\Delta} c_ {\epsilon}\rightharpoonup {\Delta} c\quad {\text{in}}\; L^\frac{5q}{5+q}_{loc}(\overline{\Omega}\times[0,\infty)), \end{gather} | (4.7) |
\begin{gather} u_ {\epsilon}\rightarrow u\quad {\text{a.e. in}}\; {\Omega}\times\left( {0,\infty} \right), \end{gather} | (4.8) |
\begin{gather} u_ {\epsilon}\rightarrow u\quad {\text{in}}\; L^2_{loc}(\overline{\Omega}\times[0,\infty)), \end{gather} | (4.9) |
\begin{gather} u_ {\epsilon}\rightharpoonup u \quad {\text{in}}\; L^\frac{10}{3}_{loc}(\overline{\Omega}\times[0,\infty)), \end{gather} | (4.10) |
\begin{gather} {\nabla } u_ {\epsilon}\rightharpoonup {\nabla } u \quad {\text{in}}\; L^2_{loc}(\overline{\Omega}\times[0,\infty)). \end{gather} | (4.11) |
Proof. For convenience, we denote a subsequence ({\epsilon}_j)_{j\in\mathbb{N}} of {\epsilon} by {\epsilon} itself. First, Lemma 2.4 gives the pointwise convergence of c_ {\epsilon} in (4.5):
\begin{equation*} c_ {\epsilon}\rightarrow c\quad \text{ a.e. in }\Omega\times\left( {0,\infty} \right). \end{equation*} |
Indeed, using Lemma 2.4, bounds for c_ {\epsilon} in L^m_{loc}([0, \infty); W^{1, m}(\Omega)) and {\partial}_t c_ {\epsilon} in L^\frac{5q}{5+q}_{loc}(\overline{\Omega}\times[0, \infty)), asserted in Lemma 3.6 and Lemma 3.7, yield the strong convergence of c_ {\epsilon} in L^m_{loc}(\overline{\Omega}\times[0, \infty)) which in particular implies (4.5). Similarly, by Lemma 3.8 and 3.9, we see that \left({1+n_ {\epsilon}} \right)^\frac{ {\gamma}}{2}_{ {\epsilon}\in(0, 1)} is relatively compact in L^2_{loc}(\overline{\Omega}\times[0, \infty)) with respect to the strong topology by Lemma 2.4. we can thus see that
\begin{equation*} n_ {\epsilon}\rightarrow n\quad \text{ a.e. in }\Omega\times\left( {0,\infty} \right), \end{equation*} |
which proves (4.1), as well as (4.4) holds. Likewise, exploiting boundedness of u_ {\epsilon} and of its time derivative, as proved in Lemma 3.5 and Lemma 3.10, and using Lemma 2.4 again, we have (4.8) and (4.9). The convergence properites (4.2), (4.6), (4.7), (4.10) and (4.11) is a direct consequence of (3.2), (3.18), (3.22), (3.17) and (3.13), respectively. In order to prove (4.3), we use (3.2) again, which implies that \int_{0}^{T}\left\Vert {n_ {\epsilon}^p} \right\Vert_\frac{q}{p} \leqslant C for all t>0. Hence we have
\begin{equation*} n_ {\epsilon}^p\rightharpoonup n^p\quad \text{ in }L^\frac{q}{p}_{loc}(\overline{\Omega}\times[0,\infty)) \end{equation*} |
as {\epsilon}\searrow 0. By this weak convergence we have
\begin{equation*} \int\limits_{0}^{T} \int\limits_{\Omega} n_ {\epsilon}^p \rightarrow \int\limits_{0}^{T} \int\limits_{\Omega} n^p {\quad {\rm{ for \;all }}\; t > 0 }, \end{equation*} |
which asserts that n_ {\epsilon}\rightarrow n \text{ in }L^p_{loc}(\overline{\Omega}\times[0, \infty)) due to uniform convexity of L^p-space for p>1. This proves (4.3).
We shall prove the limit (n, c, u) in Lemma 4.1 is a solution of our main system (1.1)–(1.3) in the sense of Definition 2. We first focus on c and u which satisfy (1.1) and (1.2) in the standard weak sence. In addition, we show that n is a weak sub-solution in the sense of Definition 1.
Lemma 4.2. Let (n, c, u) be the limit function and vector field in Lemma 4.1. Then (1.6) and (1.7) hold.
Proof. We multiply the second equation in (2.1) by the test function {\varphi}\in C^\infty_0(\overline{\Omega}\times[0, \infty)) to get, for all {\epsilon}\in(0, 1),
\begin{align*} - \int\limits_{0}^\infty\int\limits_{\Omega} c_ {\epsilon} {\varphi}_t-\int\limits_ {\Omega} c_0 {\varphi}(\cdot,0) = &-\int\limits^\infty_0\int\limits_ {\Omega} {\nabla } c_ {\epsilon}\cdot {\nabla } {\varphi}-\int\limits^\infty_0\int\limits_ {\Omega} c_ {\epsilon} {\varphi}\\ &+\int\limits^\infty_0\int\limits_ {\Omega} n_ {\epsilon} {\varphi}+\int\limits^\infty_0\int\limits_{ {\Omega}} c_ {\epsilon} u_ {\epsilon}\cdot {\nabla } {\varphi}. \end{align*} |
Applying (4.6) and (4.2), we easily obtain
\begin{equation} \int\limits_{0}^\infty\int\limits_{\Omega} c_ {\epsilon} {\varphi}_t\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} c {\varphi}_t,\qquad \int\limits_{0}^\infty\int\limits_{\Omega} c_ {\epsilon} {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} c {\varphi}, \end{equation} | (4.12) |
\begin{equation} \int\limits_{0}^\infty\int\limits_{\Omega} {\nabla } c_ {\epsilon}\cdot {\nabla } {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} {\nabla } c\cdot {\nabla } {\varphi},\qquad \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon} {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} n {\varphi} \end{equation} | (4.13) |
as {\epsilon} = {\epsilon}_j\searrow 0. On the other hand, combining (4.3) and (4.10) infers that c_ {\epsilon} u_ {\epsilon}\rightharpoonup cu in L^s_{loc} for s: = \frac{10+3p}{10p} \geqslant1, which proves
\begin{equation} \int\limits_{0}^\infty\int\limits_{\Omega} c_ {\epsilon} u_ {\epsilon}\cdot {\nabla } {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} cu\cdot {\nabla } {\varphi} \end{equation} | (4.14) |
as {\epsilon}\searrow 0. Next we multiply the third equation in (2.1) by {\varphi}\in C^\infty_0({\Omega}\times [0, \infty); \mathbb{R}^3) with {{\nabla }\cdot} {\varphi} = 0 that gives
\begin{equation*} -\int\limits_{0}^\infty\int\limits_{ {\Omega}}u_ {\epsilon}\cdot {\varphi}_t-\int\limits_{ {\Omega}} u_0\cdot {\varphi}(\cdot,0) = - \int\limits_{0}^\infty\int\limits_{\Omega} {\nabla } u_ {\epsilon}\cdot {\nabla } {\varphi} + \int\limits_{0}^\infty\int\limits_{ {\Omega}}(Y_ {\epsilon} u_ {\epsilon}\otimes u_ {\epsilon})\cdot {\nabla } {\varphi} +\int\limits_{0}^\infty\int\limits_{ {\Omega}} n_ {\epsilon} {\nabla }\phi\cdot {\varphi} \end{equation*} |
for all {\epsilon}\in(0, 1). Similar to the above, (4.10), (4.11), (4.2) and the condition on {\nabla }\phi, as assumed in (1.8), imply that
\begin{equation} \int\limits_{0}^\infty\int\limits_{\Omega} u_ {\epsilon}\cdot {\varphi}_t\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} u\cdot {\varphi}_t,\qquad \int\limits_{0}^\infty\int\limits_{\Omega} {\nabla } u_ {\epsilon}\cdot {\nabla } {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} {\nabla } u\cdot {\nabla } {\varphi}, \end{equation} | (4.15) |
\begin{equation} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon} {\nabla }\phi\cdot {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} n {\nabla }\phi\cdot {\varphi} \end{equation} | (4.16) |
as {\epsilon}\searrow 0. Since it is well known that Y_ {\epsilon} u_ {\epsilon}\, \rightarrow u in L^2_{loc}({\Omega}\times(0, \infty)), with the aid of (4.9), we obtain Y_ {\epsilon} u_ {\epsilon}\otimes u_ {\epsilon}\rightarrow u\otimes u in L^1_{loc}({\Omega}\times(0, \infty)). This proves
\begin{equation} \int\limits_{0}^\infty\int\limits_{ {\Omega}}(Y_ {\epsilon} u_ {\epsilon}\otimes u_ {\epsilon})\cdot {\nabla } {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} (u\otimes u)\cdot {\nabla } {\varphi} \end{equation} | (4.17) |
as {\epsilon}\searrow 0. We collect (4.12)–(4.17) to conclude the proof.
So far, we used that q>\frac{5}{3}. In the next Lemma, however, it is necessary to assume that q>\frac{20}{11}, which is crucial to show convergence of n_ {\epsilon} {\nabla } c_ {\epsilon} (see the estimate (4.21) below).
Lemma 4.3. Let q\in\left({\frac{20}{11}, 2} \right) and (n, c, u) be the limit function and vector field in Lemma 4.1. Then n is a {\gamma}-entropy sub-solution of (1.1)–(1.3) with {\gamma} = 1, that is, n satisfies the following integral inequality
\begin{align*} -\int\limits_{0}^\infty\int\limits_{ {\Omega}}n {\varphi}_t-\int\limits_{ {\Omega}}n_0 {\varphi}(\cdot,0)& \leqslant\int\limits_{0}^\infty\int\limits_{ {\Omega}}n\Delta {\varphi}+ \int\limits_{0}^\infty\int\limits_{\Omega} n {\nabla } c\cdot {\nabla } {\varphi}\nonumber\\ &+\rho \int\limits_{0}^\infty\int\limits_{\Omega} n {\varphi}-\mu \int\limits_{0}^\infty\int\limits_{\Omega} n^q {\varphi}+ \int\limits_{0}^\infty\int\limits_{\Omega} nu\cdot {\nabla } {\varphi} \end{align*} |
for all nonnegative {\varphi}\in C^\infty_0(\overline{ {\Omega}}\times[0, \infty)).
Proof. We multiply the first equation in (2.1) by a nonnegative test function {\varphi}\in C^\infty_0 (\overline{ {\Omega}}\times[0, \infty)) and integrate over {\Omega}\times(0, \infty). By suitable integration by parts,
\begin{equation*} \begin{split} -\int\limits_{0}^\infty\int\limits_{ {\Omega}}n_ {\epsilon} {\varphi}_t-\int\limits_{ {\Omega}}n_0 {\varphi}(\cdot,0)& = \int\limits_{0}^\infty\int\limits_{ {\Omega}}n_ {\epsilon}\Delta {\varphi}+ \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon} {\nabla } c_ {\epsilon}\cdot {\nabla } {\varphi}+\rho \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon} {\varphi}\nonumber\\ &-\mu \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^q {\varphi}- {\epsilon} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\kappa} {\varphi}+ \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon} u_ {\epsilon}\cdot {\nabla } {\varphi} \end{split} \end{equation*} |
for all {\epsilon}\in(0, 1). Using (4.2), we see that
\begin{align} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon} {\varphi}_t\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} n {\varphi}_t,\qquad \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}\Delta {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} n\Delta {\varphi}, \end{align} | (4.18) |
\begin{align} \mbox{ and }\quad\rho \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon} {\varphi}\, {\rightarrow }\,\rho \int\limits_{0}^\infty\int\limits_{\Omega} n {\varphi} \end{align} | (4.19) |
as {\epsilon}\searrow 0. Funthermore, applying strong convergence of (n_ {\epsilon})_{ {\epsilon}\in(0, 1)}, (u_ {\epsilon})_{ {\epsilon}\in(0, 1)} as asserted in Lemma 4.1, we have
\begin{equation} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon} u_ {\epsilon}\cdot {\nabla } {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} nu\cdot {\nabla } {\varphi} \end{equation} | (4.20) |
as {\epsilon}\searrow 0. Since q\in(\frac{20}{11}, 2), we can take p < q close to q satisfying \frac{1}{p}+\frac{1}{m} < 1. Then, by (4.3) and (4.6) we see that
\begin{equation} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon} {\nabla } c_ {\epsilon}\cdot {\nabla } {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} n {\nabla } c\cdot {\nabla } {\varphi} \end{equation} | (4.21) |
as {\epsilon}\searrow 0. Besides, the nonnegativity of n_ {\epsilon} and {\varphi} leads that
\begin{equation} - {\epsilon} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\kappa} {\varphi} \leqslant 0 \end{equation} | (4.22) |
for all {\epsilon}\in(0, 1). Lastly, we observe that by Fatou's lemma
\begin{equation} \mu \int\limits_{0}^\infty\int\limits_{\Omega} n^q {\varphi} \leqslant \liminf\limits_{ {\epsilon}\searrow0}\left\{ {\mu \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^q {\varphi}} \right\}. \end{equation} | (4.23) |
Hence, combining (4.18)–(4.23), we conclude that n is a {\gamma}-entropy sub-solution with {\gamma} = 1.
Now we shall prove that (n, c, u) as in Lemma 4.1 is a {\gamma}-entropy super-solution.
Lemma 4.4. Let q\in\left({\frac{5}{3}, 2} \right) and (n, c, u) be the limit functions and vector field in Lemma 4.1. Then for any fixed {\gamma}\in\left({0, \frac{4q-5}{5}} \right), n is a {\gamma}-entropy supersolution of (1.1)–(1.3).
Proof. Let 0 \leqslant {\varphi}\in C^\infty_0(\overline{ {\Omega}}\times[0, \infty)) be arbitralily. Testing the first equation in (2.1) by {\gamma} n_ {\epsilon}^{ {\gamma}-1} {\varphi} and integrating by parts, we have
\begin{align*} - \int\limits_{0}^\infty\int\limits_{\Omega}&n_ {\epsilon}^ {\gamma} {\varphi}_t-\int\limits_{ {\Omega}}n_0^ {\gamma} {\varphi}(\cdot,0) = {\gamma}(1- {\gamma})\int\limits_{0}^\infty\int\limits_{ {\Omega}}n_ {\epsilon}^{ {\gamma}-2}\left| { {\nabla } n_ {\epsilon}} \right|^2 {\varphi}+ \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\gamma}\Delta {\varphi}\nonumber\\ &+(1- {\gamma}) \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\gamma} {\Delta} c_ {\epsilon} {\varphi}+ \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\gamma} {\nabla } c_ {\epsilon}\cdot {\nabla } {\varphi}\\ &+\rho {\gamma} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\gamma} {\varphi}-\mu {\gamma} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^{q+ {\gamma}-1} {\varphi}- {\epsilon} {\gamma} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^{ {\kappa}+ {\gamma}-1} {\varphi}+ \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\gamma} u_ {\epsilon}\cdot {\nabla } {\varphi}\nonumber \end{align*} |
for all {\epsilon}\in(0, 1). Since {\gamma}\in(0, 1), we obtaing the strong convergence n_ {\epsilon}^ {\gamma} {\rightarrow }\, n^ {\gamma} in L^p_{loc}({\Omega}\times(0, \infty)) for p\in(1, q) due to (4.3) which follows
\begin{equation} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\gamma} {\varphi}_t\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} n^ {\gamma} {\varphi}_t,\quad \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\gamma}\Delta {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} n^ {\gamma}\Delta {\varphi},\quad \rho \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\gamma} {\varphi}\, {\rightarrow }\,\rho \int\limits_{0}^\infty\int\limits_{\Omega} n^ {\gamma} {\varphi} \end{equation} | (4.24) |
as {\epsilon}\searrow 0. Furthermore, referring to (4.20) and (4.21) we have
\begin{equation} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\gamma} {\nabla } c_ {\epsilon}\cdot {\nabla } {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} n^ {\gamma} {\nabla } c\cdot {\nabla } {\varphi}\quad\mbox{ and }\quad \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\gamma} u_ {\epsilon}\cdot {\nabla } {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} n^ {\gamma} u\cdot {\nabla } {\varphi} \end{equation} | (4.25) |
as {\epsilon}\searrow 0. As n_ {\epsilon}^{q+ {\gamma}-1} is bounded in L^k_{loc}({\Omega}\times(0, \infty)) for k = \frac{q}{q+ {\gamma}-1} >1, uniformly in {\epsilon}, the weak convergence n_ {\epsilon}^{q+ {\gamma}-1}\rightharpoonup n^{q+ {\gamma}-1} in L^k_{loc}({\Omega}\times(0, \infty)) holds. Thus, we have
\begin{equation} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^{q+ {\gamma}-1} {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} n^{q+ {\gamma}-1} {\varphi} \end{equation} | (4.26) |
as {\epsilon}\searrow0. Since \frac{5+q}{5q}+\frac{ {\gamma}}{q} < 1, it follows from (4.3) and (4.7) that
\begin{equation} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^ {\gamma} {\Delta} c_ {\epsilon} {\varphi}\, {\rightarrow } \int\limits_{0}^\infty\int\limits_{\Omega} n^ {\gamma} {\Delta} c\, {\varphi} \end{equation} | (4.27) |
as {\epsilon}\searrow0. For the regularizing term, we note that from Hölder inequality and (3.2)
\begin{equation*} \left| {- {\gamma} {\epsilon} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^{ {\kappa}+ {\gamma}-1} {\varphi}} \right| \leqslant C_1 {\gamma} {\epsilon}^\frac{1- {\gamma}}{ {\kappa}}\left\Vert { {\varphi}} \right\Vert_\infty\left( { {\epsilon} \int\limits_{0}^\infty\int\limits_{\Omega}{n_ {\epsilon}}^ {\kappa}} \right)^\frac{ {\kappa}+ {\gamma}-1}{ {\kappa}} \leqslant C_2 {\epsilon}^\frac{1- {\gamma}}{ {\kappa}} \end{equation*} |
for all {\epsilon}\in(0, 1). Hence, we have
\begin{equation} - {\gamma} {\epsilon} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^{ {\kappa}+ {\gamma}-1} {\varphi}\, {\rightarrow } \,0 \end{equation} | (4.28) |
as {\epsilon}\searrow0. Finally, from (4.4) and the lower semicontinuity of the seminorm \left\Vert {\cdot} \right\Vert defined by \left\Vert {f} \right\Vert: = (\int\limits_{0}^\infty\int\limits_{\Omega} f^2 {\varphi})^\frac{1}{2} with respect to weak convergence, we obtain
\begin{equation} {\gamma}(1- {\gamma}) \int\limits_{0}^\infty\int\limits_{\Omega} n^{ {\gamma}-2}\left| { {\nabla } n} \right|^2 {\varphi} \leqslant {\gamma}(1- {\gamma})\liminf\limits_{ {\epsilon}\searrow0} \int\limits_{0}^\infty\int\limits_{\Omega} n_ {\epsilon}^{ {\gamma}-2}\left| { {\nabla } n_ {\epsilon}} \right|^2 {\varphi}. \end{equation} | (4.29) |
Therefore, collecting (4.24)–(4.29) proves that n is a {\gamma}-entropy super-solution of (1.1)–(1.3).
Proof of Theorem 1.1. This is the combination of Lemma 4.2, Lemma 4.3 and Lemma 4.4.
The following Lemma is elementary, but for clarity, we give its detail.
Lemma 5.1. Let a > 1 and f\in L^1([0, \infty)) . Suppose there is t_0 > 0 such that f(t) \leqslant Nt^{-a} for sufficiently large t \geqslant t_0 . Assume further that a non-negative measurable function y(t) satisfies
y'(t)+y(t) \leqslant f(t). |
Then, y(t) \leqslant Ct^{-a} for sufficiently large t .
Proof. Firstly we note that y(t) is bounded uniformly in time. Then, using the integrating factor, we have for t \geqslant t_0
e^{2t}y(2t)-e^{t}y(t) \leqslant \int\limits_{t}^{2t} e^{\tau}f(\tau) {\,\mathrm{d}}\tau, |
which yields, using integration by parts,
\begin{align*} y(2t)& \leqslant e^{-t}y(t)+Ne^{-2t}\int\limits_{t}^{2t}e^\tau\tau^{-a} {\,\mathrm{d}}\tau\\ & \leqslant Ce^{-t}+Ne^{-2t}\left[ {e^{2t}\left( {2t} \right)^{- {\alpha}}-e^t t^{- {\alpha}}+ {\alpha}\int\limits_{t}^{2t}e^\tau\tau^{- {\alpha}-1} {\,\mathrm{d}}\tau} \right]\\ & \leqslant C(2t)^{- {\alpha}}. \end{align*} |
Proof of Theorem 1.2. \bullet (The case \rho = 0 ) Noting that \rho = 0, we integrate the equation for n_ {\epsilon} in (2.1) over \Omega to get
\frac{d}{dt} \int\limits_{\Omega} n_ {\epsilon}(\cdot, t) {\,\mathrm{d}} x \leqslant -\frac{\mu}{\left| {\Omega} \right|^{q-1}}\left( { \int\limits_{\Omega} n_ {\epsilon}(\cdot, t) {\,\mathrm{d}} x} \right)^q. |
A standard argument of ODE implies that
\int\limits_{\Omega} n_ {\epsilon}(\cdot, t) {\,\mathrm{d}} x \leqslant C(1+t)^{-\frac{1}{q-1}} {\quad {\rm{ for \; all }}\; t > 0 }. |
Next, integrating the equation of c_ {\epsilon} , it follows that for all t>0,
\frac{d}{dt} \int\limits_{\Omega} c_ {\epsilon}(\cdot, t) {\,\mathrm{d}} x+ \int\limits_{\Omega} c_ {\epsilon}(\cdot, t) {\,\mathrm{d}} x = \int\limits_{\Omega} n_ {\epsilon}(\cdot, t) {\,\mathrm{d}} x. |
Let g(t) = \int\limits_{\Omega} n_ {\epsilon}(\cdot, t) {\, \mathrm{d}} x . Then, since \frac{1}{q-1} > 1 , we observe that g\in L^{1}([0, \infty)) , and thus, via Lemma 5.1, it follows that
\int\limits_{\Omega} c_ {\epsilon}(\cdot, t) {\,\mathrm{d}} x \leqslant C(1+t)^{-\frac{1}{q-1}} {\quad {\rm{ for \; all }}\; t > 0 }. |
On the other hand, putting m = 3q-2 and testing the equation for c_ {\epsilon} in (2.1) by c^{m-1} , we get
\frac{1}{m}\frac{d}{dt} \int\limits_{\Omega} c_ {\epsilon}^m(\cdot, t) {\,\mathrm{d}} x+ \int\limits_{\Omega} \left| {\nabla c_ {\epsilon}^{\frac{m}{2}}} \right|^2 {\,\mathrm{d}} x+ \int\limits_{\Omega} c_ {\epsilon}^m {\,\mathrm{d}} x = \int\limits_{\Omega} n_ {\epsilon} c_ {\epsilon}^{m-1} {\,\mathrm{d}} x |
\leqslant\left\Vert {n_ {\epsilon}} \right\Vert_{\frac{3m}{2m+1}}\left\Vert {c_ {\epsilon}^{m-1}} \right\Vert_{\frac{3m}{m-1}} = \left\Vert {n_ {\epsilon}} \right\Vert_{\frac{3m}{2m+1}}\left\Vert {c_ {\epsilon}^{\frac{m}{2}}} \right\Vert^{\frac{2(m-1)}{m}}_{6} |
\leqslant C\left\Vert {n_ {\epsilon}} \right\Vert_{\frac{3m}{2m+1}}\left( {\left\Vert {\nabla c_ {\epsilon}^{\frac{m}{2}}} \right\Vert^{\frac{2(m-1)}{m}}_{2}+1} \right) \leqslant C\left\Vert {n_ {\epsilon}} \right\Vert^m_{\frac{3m}{2m+1}}+\frac{1}{2}\left\Vert {\nabla c_ {\epsilon}^{\frac{m}{2}}} \right\Vert^{2}_{2}. |
Since m = 3q-2 , we observe that
\left\Vert {n_ {\epsilon}} \right\Vert^m_{\frac{3m}{2m+1}} = \left\Vert {n_ {\epsilon}} \right\Vert^{3q-2}_{\frac{3q-2}{2q-1}} \leqslant \left\Vert {n_ {\epsilon}} \right\Vert_{1}^{2(q-1)}\left\Vert {n_ {\epsilon}} \right\Vert_{q}^q \leqslant C(1+t)^{-2}\left\Vert {n_ {\epsilon}(t)} \right\Vert_{q}^q. |
Let h(t) = (1+t)^{-2}\left\Vert {n_ {\epsilon}(t)} \right\Vert_{q}^q . Then, it is direct that h\in L^1((0, \infty)) . Setting Z(t) = \int\limits_{\Omega} c_ {\epsilon}^m(\cdot, t) {\, \mathrm{d}} x , we have Z'(t)+Z(t) \leqslant h(t) , which yields
e^{2t}Z(2t)-e^{t}Z(t) = \int\limits_{t}^{2t}e^{\tau} h(\tau)d\tau, |
which implies that
Z(2t) \leqslant e^{-t}Z(t)+C(1+t)^{-2}\int\limits_{t}^{2t} \left\Vert {n_ {\epsilon}(\tau)} \right\Vert_{q}^qd\tau \leqslant C(1+t)^{-2}. |
Noting that Z(t) \leqslant C for all t>0, we have
\left\Vert {c_ {\epsilon}(t)} \right\Vert_{3q-2} \leqslant C(1+t)^{-\frac{2}{3q-2}}. |
Hence, interpolation gives
\left\Vert {c_ {\epsilon}(t)} \right\Vert_l \leqslant \left\Vert {c_ {\epsilon}(t)} \right\Vert_1^{1-\theta}\left\Vert {c_ {\epsilon}(t)} \right\Vert_{3q-2}^\theta\\ \leqslant C(1+t)^{-\frac{2lq+q-3l}{3l(q-1)^2}}, |
where 1 \leqslant l \leqslant 3q-2 and \theta = \frac{(l-1)(3q-2)}{3l(q-1)}. On the other hand, in case that 3q-2 \leqslant l \leqslant \frac{3q}{5-2q}, interpolation gives
\left\Vert {c_ {\epsilon}(t)} \right\Vert_k \leqslant\left\Vert {c_ {\epsilon}(t)} \right\Vert_{3q-2}^{\theta_1}\left\Vert {c_ {\epsilon}(t)} \right\Vert_\frac{3q}{5-2q}^{1-\theta_1} \leqslant C(1+t)^{-\frac{3q-(5-2q)k}{k(3q-5)(q-1)}}, |
where \theta_1 = \frac{(3q-(5-2q)k)(3q-2)}{2k(3q-5)(q-1)}. Finally, recalling (3.14) and (3.15), we have
\begin{align*} \frac{d}{dt} \int\limits_{\Omega} \left| {u_ {\epsilon}(\cdot, t)} \right|^2 {\,\mathrm{d}} x+\frac{1}{2} \int\limits_{\Omega} \left| {\nabla u_ {\epsilon}(\cdot, t)} \right|^2 {\,\mathrm{d}} x& \leqslant C\left( { \int\limits_{\Omega} \left| {n_ {\epsilon}(\cdot,t)} \right|^\frac{6}{5}} \right)^\frac{5}{3}\\ & \leqslant C\left\Vert {n_ {\epsilon}(t)} \right\Vert^{\frac{5q-6}{3(q-1)}}_{1}\left\Vert {n_ {\epsilon}(t)} \right\Vert^{\frac{q}{3(q-1)}}_{q}, \end{align*} |
where we used
\left\Vert {n_ {\epsilon}} \right\Vert_{\frac{6}{5}} \leqslant \left\Vert {n_ {\epsilon}} \right\Vert^{\theta}_{1}\left\Vert {n_ {\epsilon}} \right\Vert^{1-\theta}_{q},\qquad \theta = \frac{5q-6}{6(q-1)}. |
We set h(t) = \left\Vert {n_ {\epsilon}(t)} \right\Vert^{\frac{5q-6}{3(q-1)}}_{1}\left\Vert {n_ {\epsilon}(t)} \right\Vert^{\frac{q}{3(q-1)}}_{q} \leqslant C(1+t)^{-\frac{5q-6}{3(q-1)^2}}\left\Vert {n_ {\epsilon}(t)} \right\Vert^{\frac{q}{3(q-1)}}_{q} . We note that h\in L^1((0, \infty)) , since n_ {\epsilon}\in L^q({\Omega}\times(0, \infty)) and
\int\limits_{0}^{\infty} h(t) {\,\mathrm{d}} t \leqslant \left( {\int\limits_{0}^{\infty} (1+t)^{-\frac{5q-6}{(3q-4)(q-1)}} {\,\mathrm{d}} t} \right)^{\frac{3q-4}{3(q-1)}}\left( {\int\limits_{0}^{\infty}\left\Vert {n_ {\epsilon}(t)} \right\Vert_{q}^q {\,\mathrm{d}} t} \right)^{\frac{1}{3(q-1)}} < C. |
Using the Poincaré inequality, it follows that
\begin{equation} \frac{d}{dt} \int\limits_{\Omega} \left| {u_{ {\epsilon}}(\cdot, t)} \right|^2 {\,\mathrm{d}} x+\frac{C_p}{2} \int\limits_{\Omega} \left| {u_{ {\epsilon}}(\cdot, t)} \right|^2 {\,\mathrm{d}} x \leqslant h(t). \end{equation} | (5.1) |
Since h is in L^1 , we have \left\Vert {u_ {\epsilon}(\cdot, t)} \right\Vert_{2} \leqslant C for all t . In addition, we obtain, for sufficiently large t ,
\left\Vert {u_ {\epsilon}(t)} \right\Vert_{2} \leqslant C(1+t)^{-\frac{-3q^2+12q-10}{3(q-1)^2}}. |
Indeed, setting z(t): = \left\Vert {u_ {\epsilon}(t)} \right\Vert^2_{2} , it leads that
\begin{align*} z(2t) \leqslant& e^{-t}z(t)+e^{-2t}\int\limits_{t}^{2t} e^{\tau}h(\tau)d\tau \leqslant e^{-t}z(t)+\int\limits_{t}^{2t} h(\tau)d\tau\\ \leqslant& Ce^{-t}+C\left( {\int\limits_{t}^{2t} (1+t)^{-\frac{5q-6}{(3q-4)(q-1)}}} \right)^{\frac{3q-4}{3(q-1)}}\\ \leqslant& Ce^{-t}+C(1+t)^{\frac{3q^2-12q+10}{3(q-1)^2}} \leqslant C(1+t)^{-\frac{-3q^2+12q-10}{3(q-1)^2}}. \end{align*} |
\bullet (The case \rho < 0 ) Firstly, we integrate the equation for n_ {\epsilon} over \Omega to get
\begin{equation*} \frac{d{}}{d{t}} \int\limits_{\Omega} n_ {\epsilon} -\rho \int\limits_{\Omega} n_ {\epsilon} \leqslant-\mu \int\limits_{\Omega} n_ {\epsilon}^ {\kappa} \leqslant 0, \end{equation*} |
which directly yields
\begin{equation} \int\limits_{\Omega} n_ {\epsilon}(\cdot,t) {\,\mathrm{d}} x \leqslant me^{\rho t} {\quad {\rm{ for \; all }}\; t > 0 }, \end{equation} | (5.2) |
where m is as in Lemma 3.1. Next, again integrating the equation for c_ {\epsilon} over \Omega and letting z(t): = \int_\Omega c_ {\epsilon}(\cdot, t) {\, \mathrm{d}} x, it follows that
\begin{equation*} z'(t)+z(t) \leqslant me^{\rho t}, \end{equation*} |
which leads that for all t>0,
\begin{equation*} z(t) \leqslant e^{-t}z_0+me^{-t}\int\limits_{0}^{t}e^{(1+\rho)\tau} {\,\mathrm{d}}\tau \leqslant C\left( {e^{-t}+\frac{1}{1+\rho}\left( {e^{\rho t}-e^{-t}} \right)} \right), \end{equation*} |
where C = \max\left\{ {m, \int_\Omega c_0} \right\}. Thus, we have
\begin{equation} \int\limits_{\Omega} c_ {\epsilon}(\cdot,t) {\,\mathrm{d}} x \leqslant Ce^{-\rho_*t} {\quad {\rm{ for \; all }}\; t > 0 }, \end{equation} | (5.3) |
where \rho_* = \min\left\{ {-\rho, 1} \right\}>0. Using the interpolation inequality, (3.5) and (5.3), we obtain for 1 \leqslant l \leqslant \frac{3q}{5-2q},
\begin{align*} \left\Vert {c_ {\epsilon}(t)} \right\Vert_l \leqslant\left\Vert {c_ {\epsilon}(t)} \right\Vert_1^\frac{3q-(5-2q)l}{5(q-1)l}\left\Vert {c_ {\epsilon}(t)} \right\Vert_\frac{3q}{5-2q}^\frac{3q(l-1)}{5(q-1)l} \leqslant Ce^{-\frac{3q-(5-2q)l}{5(q-1)l}\rho_* t} {\quad {\rm{ for \; all }}\; t > 0 }. \end{align*} |
Lastly, we recall the inequality (5.1):
\begin{equation*} \frac{d}{dt} \int\limits_{\Omega} \left| {u_ {\epsilon}(\cdot, t)} \right|^2 {\,\mathrm{d}} x+{C_*} \int\limits_{\Omega} \left| {u_ {\epsilon}(\cdot, t)} \right|^2 {\,\mathrm{d}} x \leqslant h(t). \end{equation*} |
Here h(t) = \left\Vert {n_ {\epsilon}} \right\Vert_1^\frac{5q-6}{3(q-1)}\left\Vert {n_ {\epsilon}} \right\Vert_q^\frac{q}{3(q-1)} \leqslant C_3e^{-\delta t}\left\Vert {n_ {\epsilon}(t)} \right\Vert_q^\frac{q}{3(q-1)} with \delta = -\frac{5q-6}{3(q-1)}\rho>0 and C_* = \frac{C_p}{2}>0, where C_p is the constant appeared in the Poincaré inequality. Letting z(t): = \left\Vert {u_ {\epsilon}(t)} \right\Vert_2^2, we have
\begin{align*} z(t) \leqslant& e^{-C_*t}z(0)+e^{-C_*t}\int\limits_{0}^{t} e^{C_*\tau}h(\tau)d\tau\\ \leqslant& e^{-C_*t}z(0)+C_3e^{-C_*t}\int\limits_{0}^{t}e^{(C_*-\delta)\tau}\left\Vert {n_ {\epsilon}(\tau)} \right\Vert_q^\frac{q}{3(q-1)} d\tau\\ \leqslant& e^{-C_*t}z(0)+C_3e^{-C_*t}e^{(C_*-\delta)_+t}t^\frac{3q-4}{3(q-1)}\left( {\int\limits_{0}^{t}\left\Vert {n_ {\epsilon}(\tau)} \right\Vert_q^q d\tau} \right)^\frac{1}{3(q-1)}\\ \leqslant&C_4\left( {e^{-C_*t}+e^{-\min\left\{ {C_*,\delta} \right\}\frac{t}{2}}} \right)\\ \leqslant&C_5e^{-\delta_*t}, \end{align*} |
where \delta_* = \frac{1}{2}\min\left\{ {C_*, \delta} \right\}. In both cases \rho = 0 and \rho < 0, we finally get the estimates for (n, c, u) in Theorem 1.2 by passing {\epsilon} to the limit via the Fatou's Lemma which is guaranteed by (4.1), (4.5) and (4.8).
K. Kang is partially supported by NRF-2019R1A2C1084685 and NRF-2015R1A5A1009350. D. Kim is supported by NRF-2019R1A2C1084685.
The authors declare no conflict of interest.
[1] |
M. Chae, K. Kang, J. Lee, Existence of smooth solutions to coupled chemotaxis-fluid equations, DCDS, 33 (2013), 2271–2297. doi: 10.3934/dcds.2013.33.2271
![]() |
[2] |
M. Chae, K. Kang, J. Lee, Global existence and temporal decay in keller-segel models coupled to fluid equations, Commun. Part. Diff. Eq., 39 (2014), 1205–1235. doi: 10.1080/03605302.2013.852224
![]() |
[3] | P. Cherrier, A. Milani, Linear and quasi-linear evolution equations in Hilbert spaces, Providence, RI: American Mathematical Society, 2012. |
[4] | J. C. Coll, B. F. Bowden, G. V. Meehan, G. M. Konig, A. R. Carroll, D. M. Tapiolas, et al., Chemical aspects of mass spawning in corals. I. sperm-attractant molecules in the eggs of the scleractinian coral montipora digitata, Marine Biology, 118, (1994), 177–182. |
[5] |
R. Denk, M. Hieber, J. Prüss, Optimal L^p-L^q-estimates for parabolic boundary value problems with inhomogeneous data, Math. Z., 257 (2007), 193–224. doi: 10.1007/s00209-007-0120-9
![]() |
[6] |
E. Espejo, T. Suzuki, Reaction terms avoiding aggregation in slow fluids, Nonlinear Anal. Real, 21 (2015), 110–126. doi: 10.1016/j.nonrwa.2014.07.001
![]() |
[7] |
S. Ishida, Global existence and boundedness for chemotaxis-navier-stokes systems with position-dependent sensitivity in 2d bounded domains, DCDS, 35 (2015), 3463–3482. doi: 10.3934/dcds.2015.35.3463
![]() |
[8] |
K. Kang, K. Kim, C. Yoon, Existence of weak and regular solutions for keller-segel system with degradation coupled to fluid equations, J. Math. Anal. Appl., 485 (2020), 123750. doi: 10.1016/j.jmaa.2019.123750
![]() |
[9] | E. F. Keller, L. A. Segel, Model for chemotaxis, J. Theor. Biol., 30 (1971), 225–234. |
[10] |
A. Kiselev, L. Ryzhik, Biomixing by chemotaxis and enhancement of biological reactions, Commun. Part. Diff. Eq., 37 (2012), 298–318. doi: 10.1080/03605302.2011.589879
![]() |
[11] |
J. Lankeit, Long-term behaviour in a chemotaxis-fluid system with logistic source, Math. Mod. Meth. Appl. S, 26 (2016), 2071–2109. doi: 10.1142/S021820251640008X
![]() |
[12] |
N. Mittal, E. O. Budrene, M. P. Brenner, A. Van Oudenaarden, Motility of escherichia coli cells in clusters formed by chemotactic aggregation, PNAS, 100 (2003), 13259–13263. doi: 10.1073/pnas.2233626100
![]() |
[13] | L. Nirenberg, An extended interpolation inequality, Ann. Scuola Norm. Sup. Pisa Cl. Sci., 20 (1966), 733–737. |
[14] |
J. Prüss, R. Schnaubelt, Solvability and maximal regularity of parabolic evolution equations with coefficients continuous in time, J. Math. Anal. Appl., 256 (2001), 405–430. doi: 10.1006/jmaa.2000.7247
![]() |
[15] |
Y. Tao, M. Winkler, Boundedness and decay enforced by quadratic degradation in a three-dimensional chemotaxis-fluid system, Z. Angew. Math. Phys., 66 (2015), 2555–2573. doi: 10.1007/s00033-015-0541-y
![]() |
[16] |
Y. Tao, M. Winkler, Blow-up prevention by quadratic degradation in a two-dimensional keller-segel-navier-stokes system, Z. Angew. Math. Phys., 67 (2016), 138. doi: 10.1007/s00033-016-0732-1
![]() |
[17] | R. Temam, Navier-Stokes equations. Theory and numerical analysis, Amsterdam-New York-Oxford: North-Holland Publishing Co., 1977. |
[18] |
I. Tuval, L. Cisneros, C. Dombrowski, C. W. Wolgemuth, J. O. Kessler, R. E. Goldstein, Bacterial swimming and oxygen transport near contact lines, PNAS, 102 (2005), 2277–2282. doi: 10.1073/pnas.0406724102
![]() |
[19] |
G. Viglialoro, Very weak global solutions to a parabolic-parabolic chemotaxis-system with logistic source, J. Math. Anal. Appl., 439 (2016), 197–212. doi: 10.1016/j.jmaa.2016.02.069
![]() |
[20] |
G. Viglialoro, Boundedness properties of very weak solutions to a fully parabolic chemotaxis-system with logistic source, Nonlinear Anal. Real, 34 (2017), 520–535. doi: 10.1016/j.nonrwa.2016.10.001
![]() |
[21] |
W. Wang, A quasilinear fully parabolic chemotaxis system with indirect signal production and logistic source, J. Math. Anal. Appl., 477 (2019), 488–522. doi: 10.1016/j.jmaa.2019.04.043
![]() |
[22] |
M. Winkler, A three-dimensional keller-segel-navier-stokes system with logistic source: global weak solutions and asymptotic stabilization, J. Funct. Anal., 276 (2019), 1339–1401. doi: 10.1016/j.jfa.2018.12.009
![]() |
[23] |
M. Winkler, Chemotaxis with logistic source: very weak global solutions and their boundedness properties, J. Math. Anal. Appl., 348 (2008), 708–729. doi: 10.1016/j.jmaa.2008.07.071
![]() |
[24] |
M. Winkler, Stabilization in a two-dimensional chemotaxis-navier-stokes system, Arch. Ration. Mech. Anal., 211 (2014), 455–487. doi: 10.1007/s00205-013-0678-9
![]() |
1. | Kyungkeun Kang, Dongkwang Kim, Soo-Oh Yang, Existence of generalized solutions to an attraction-repulsion Keller-Segel system with degradation, 2022, 511, 0022247X, 126069, 10.1016/j.jmaa.2022.126069 | |
2. | Yu Tian, Zhaoyin Xiang, A three-dimensional Keller-Segel-Navier–Stokes system involving subquadratic logistic degradation: global generalized solutions and eventual smoothness, 2025, 64, 0944-2669, 10.1007/s00526-024-02891-6 |