Processing math: 83%
Research article Special Issues

Regularization effect of the mixed-type damping in a higher-dimensional logarithmic Keller-Segel system related to crime modeling


  • We study a logarithmic Keller-Segel system proposed by Rodríguez for crime modeling as follows:

    {ut=Δuχ(ulnv)κuv+h1,vt=Δvv+u+h2,

    in a bounded and smooth spatial domain ΩRn with n3, with the parameters χ>0 and κ>0, and with the nonnegative functions h1 and h2. For the case that κ=0, h10 and h20, recent results showed that the corresponding initial-boundary value problem admits a global generalized solution provided that χ<χ0 with some χ0>0.

    In the present work, our first result shows that for the case of κ>0 such problem possesses global generalized solutions provided that χ<χ1 with some χ1>χ0, which seems to confirm that the mixed-type damping κuv has a regularization effect on solutions. Besides the existence result for generalized solutions, a statement on the large-time behavior of such solutions is derived as well.

    Citation: Bin Li, Zhi Wang, Li Xie. Regularization effect of the mixed-type damping in a higher-dimensional logarithmic Keller-Segel system related to crime modeling[J]. Mathematical Biosciences and Engineering, 2023, 20(3): 4532-4559. doi: 10.3934/mbe.2023210

    Related Papers:

    [1] Thierry Colin, Marie-Christine Durrieu, Julie Joie, Yifeng Lei, Youcef Mammeri, Clair Poignard, Olivier Saut . Modeling of the migration of endothelial cells on bioactive micropatterned polymers. Mathematical Biosciences and Engineering, 2013, 10(4): 997-1015. doi: 10.3934/mbe.2013.10.997
    [2] Shangbing Ai, Zhian Wang . Traveling bands for the Keller-Segel model with population growth. Mathematical Biosciences and Engineering, 2015, 12(4): 717-737. doi: 10.3934/mbe.2015.12.717
    [3] Tong Li, Zhi-An Wang . Traveling wave solutions of a singular Keller-Segel system with logistic source. Mathematical Biosciences and Engineering, 2022, 19(8): 8107-8131. doi: 10.3934/mbe.2022379
    [4] Sunwoo Hwang, Seongwon Lee, Hyung Ju Hwang . Neural network approach to data-driven estimation of chemotactic sensitivity in the Keller-Segel model. Mathematical Biosciences and Engineering, 2021, 18(6): 8524-8534. doi: 10.3934/mbe.2021421
    [5] Lin Zhang, Yongbin Ge, Zhi Wang . Positivity-preserving high-order compact difference method for the Keller-Segel chemotaxis model. Mathematical Biosciences and Engineering, 2022, 19(7): 6764-6794. doi: 10.3934/mbe.2022319
    [6] Lin Zhang, Yongbin Ge, Xiaojia Yang . High-accuracy positivity-preserving numerical method for Keller-Segel model. Mathematical Biosciences and Engineering, 2023, 20(5): 8601-8631. doi: 10.3934/mbe.2023378
    [7] Fadoua El Moustaid, Amina Eladdadi, Lafras Uys . Modeling bacterial attachment to surfaces as an early stage of biofilm development. Mathematical Biosciences and Engineering, 2013, 10(3): 821-842. doi: 10.3934/mbe.2013.10.821
    [8] Chunxue Wu, Fang Yang, Yan Wu, Ren Han . Prediction of crime tendency of high-risk personnel using C5.0 decision tree empowered by particle swarm optimization. Mathematical Biosciences and Engineering, 2019, 16(5): 4135-4150. doi: 10.3934/mbe.2019206
    [9] Hany Bauomy . Safety action over oscillations of a beam excited by moving load via a new active vibration controller. Mathematical Biosciences and Engineering, 2023, 20(3): 5135-5158. doi: 10.3934/mbe.2023238
    [10] Hany Bauomy, Ashraf Taha . Nonlinear saturation controller simulation for reducing the high vibrations of a dynamical system. Mathematical Biosciences and Engineering, 2022, 19(4): 3487-3508. doi: 10.3934/mbe.2022161
  • We study a logarithmic Keller-Segel system proposed by Rodríguez for crime modeling as follows:

    {ut=Δuχ(ulnv)κuv+h1,vt=Δvv+u+h2,

    in a bounded and smooth spatial domain ΩRn with n3, with the parameters χ>0 and κ>0, and with the nonnegative functions h1 and h2. For the case that κ=0, h10 and h20, recent results showed that the corresponding initial-boundary value problem admits a global generalized solution provided that χ<χ0 with some χ0>0.

    In the present work, our first result shows that for the case of κ>0 such problem possesses global generalized solutions provided that χ<χ1 with some χ1>χ0, which seems to confirm that the mixed-type damping κuv has a regularization effect on solutions. Besides the existence result for generalized solutions, a statement on the large-time behavior of such solutions is derived as well.



    Let u(x,t) denote the density of criminals, and let v(x,t) represent the abstract so-called attractiveness value. A class of logarithmic Keller-Segel models of the following form

    {ut=Δuχ(ulnv)κuv+h1, xΩ, t>0,vt=Δvv+u+h2, xΩ, t>0, (1.1)

    with the parameters χ>0 and κ>0, was introduced in [1] to model the propagation of criminal activities, where ΩRn are bounded and smooth spatial domains. In the model (1.1), the given functions h1(x,t) and h2(x,t) describe the density of additional criminals and the source of attractiveness, respectively.

    When +u in the second equation in (1.1) is replaced by +uv, it arrives at the original Short et al. crime model [2,3], which is rewritten as

    {ut=Δuχ(ulnv)κuv+h1, xΩ, t>0,vt=Δvv+uv+h2, xΩ, t>0, (1.2)

    with the particular value χ=2. Note that results on related stationary problems, as in [4,5,6,7,8,9,10,11,12], strongly support that the model (1.2) is adequate to describe the formation of crime hotspots encountered in reality. As for the corresponding initial-boundary value problems, { the understanding of them are incomplete. The local-in-time classical solution established in [13] is global provided that either n=1 [14,15] or n2 and χ<2n [16,17] or both the initial data and the given functions h1 and h2 are appropriately small [18,19].} For larger ranges of χ, global existence results, without imposing smallness on the initial data and on the given functions, are only available for either certain types of weak solutions or certain modified versions which contain additional regularizing ingredients: the globally radial renormalized solution was obtained for n=2 and any χ>0 [20], which was extended to n=3 with restriction that χ(0,3) [21]; the global weak solution was established in [22] for n=2 and χ>0 by nonlinear diffusion enhancement (i.e., Δu is replaced by Δum with m>32); the global generalized solution was structured in [23] for n=2 and χ>0 by incorporating the logistic source (i.e., aubu2), which was extended to the case without incorporating the logistic source in [24]. Moreover, to suppress the formation of crime hotspots, the effects of law enforcement agents can be incorporated into (1.2) [3,11,25,26,27], and we also refer to [28,29,30,31] for the existence and stability of the related steady states.

    Note that, whenever κ=0, h10 and h20, the model (1.1) becomes the celebrated logarithmic Keller-Segel model [32]:

    {ut=Δuχ(ulnv), xΩ, t>0,vt=Δvv+u, xΩ, t>0, (1.3)

    in which u and v respectively represent the density of chemotactic cells and the chemoattractant concentration. To motivate our study, we also recall some results on (1.3). As to the global solvability of (1.3), various thresholds of χ have been introduced. Namely, the initial-boundary value problem possesses a global bounded classical solution for suitably regular initial data (u0,v0), provided that either χ<2n [33,34], or n=2 and χ<ˆχ with some ˆχ(1,2) [35], or χ4n [18]. Beyond this, the restrictions on χ have been relaxed within suitably generalized solution frameworks, for instance, χ<n+23n4 in the weak sense [34], χ<nn2 in radially symmetric setting [36], χ<χ0 with χ0= for n=2 and

    χ0={8,n=3,nn2,n4, (1.4)

    in the integrable sense [37], and χ>0 in the measure-valued sense [38]. In the case that ut in the first equation in (1.3) is replaced by εut with appropriately small ε, there exists an unbounded solution for large initial data, provided that χ>nn2 with n3 [39].{ As to the asymptotic stability of constant steady states, for a variant of (1.3) in more general non-normalized parameter settings it was established in [40] under the smallness of the domain size |Ω|, and later on, this restriction was removed out in [41] by assuming χ12 and the convexity of Ω.} {In addition, when the second equation in (1.3) is replaced by vt=Δvuv, the corresponding model is known as the logarithmic Keller-Segel model with signal absorption, which has also been studied in a series of papers, see for instance [42,43,44,45,46,47,48,49].}

    Concerning the mathematical analysis, the model (1.1) is expected to have better solution properties than that of the models (1.2) and (1.3). However, to the best of our knowledge, the analysis results on model (1.1) are very sparse: Rodríguez in [1] presented that the corresponding initial-boundary value problem admits a global classical solution for the case that κ=1, χ=1 and n=2, which was extended to the case that χ4n, κ0 and n2 in [18]; very recently, we showed in [50] that such problem possesses globally generalized solutions in the two-dimensional setting for any χ>0, and investigated the eventual smoothness of these generalized solutions. Compared these to aforementioned results related to (1.3), an appealing problem naturally appears: Does the mixed-type damping term κuv possess some regularization effect that contributes to enlarging the range of the parameter χ within which the higher-dimensional initial-boundary value problem of (1.1) admits global solvability at least within some generalized framework?

    To reveal it, the purpose of the present work is to explore the regularization effect of the quadratic absorption term κuv with κ>0 in the following initial-boundary value problem related to (1.1):

    {ut=Δuχ(ulnv)κuv+h1, xΩ, t>0,vt=Δvv+u+h2, xΩ, t>0,uν=vν=0, xΩ, t>0,u(x,0)=u0(x), v(x,0)=v0(x), xΩ, (1.5)

    with the parameters χ>0 and κ>0, where ΩRn(n3) are bounded and smooth domains, and ν denotes the exterior normal vector to the boundary Ω.

    To specify the setup for our analysis, we assume throughout the sequel that the initial data (u0,v0) fulfill that

    {u0C0(¯Ω)withu00andu00,v0W1,(¯Ω)withinfx¯Ωv0>0, (1.6)

    and the given functions h1 and h2 satisfy that

    0hiC1(¯Ω×[0,))L(Ω×(0,)),i=1,2. (1.7)

    The first attempt is to show that the initial-boundary value problem (1.5) admits some global generalized solutions for general initial data and arbitrary χ<χ1 with some χ1>χ0, where χ0 is given in (1.4). For any given (u0,v0,h1,h2) obeying (1.6) and (1.7), the global generalized solution of the problem (1.5) can be defined as follows:

    Definition 1.1. A pair (u,v) is called a global generalized solution to the initial-boundary value problem (1.5) if for any T>0,

    (1) it holds that for some r>1, p(0,1) and q(0,1)

    {uLr(Ω×(0,T)),ln(1+u)L2(Ω×(0,T)),uvL1(Ω×(0,T)),up+1vq1L1(Ω×(0,T)),vL(0,T;L2(Ω))L2(0,T;H1(Ω)),lnvL2(Ω×(0,T))u(x,t)0,v(x,t)>0,a.e.inΩ×[0,T],Ωu(,t)dx+κt0ΩuvdxdsΩu0dx+t0Ωh1dxds,a.e.in[0,T]; (1.8)

    (2) it holds that for each nonnegative φ(x,t)C0(¯Ω×[0,T))

    T0Ω(ln(u+1)φt+ln(1+u)φφ|ln(1+u)|2χu1+uφlnv+χuφ1+uln(1+u)lnv+κuv1+uφh11+uφ)dxdtΩln(u0+1)φ|t=0dx; (1.9)

    (3) it holds that for any φ(x,t)L(Ω×(0,T))L2(0,T;H1(Ω)) having compact support in ¯Ω×[0,T) with φtL2(Ω×(0,T))

    T0Ω(vφt+vφ+vφuφh2φ)dxdt=Ωv0φ|t=0dx. (1.10)

    We would like to remark that such concept of generalized solutions resembles those for the (logarithmic) Keller-Segel system with signal absorption used in [49,51], but is different from that proposed in [37] by Lankeit and Winkler for the model (1.3). The first result on the global existence of such generalized solutions can be stated as follows.

    Theorem 1.2. Let (1.6) and (1.7) hold, and let χ>0 fulfill that

    χ<χ1:={23,n=3,21+4n,n4. (1.11)

    Then the initial-boundary value problem (1.5) possesses at least one global generalized solution (u,v) in the sense of Definition 1.1.

    Remark 1.1. Simple computation shows that χ1 defined in (1.11) is larger than χ0 provided by (1.4). This reveals in some sense that κuv with κ>0 indeed has some regularization effect on solutions.

    With the global existence statement at hand, it is natural to focus on the large-time behavior of generalized solutions. To achieve it, we need the following additional assumptions on h1 and h2:

    inft>0Ωh2(x,t)dx>0, (1.12)
    t+1tΩh1(,t)dxds0,ast, (1.13)
    t+1tΩ|h2(,t)h2,()|2dxds0,ast, (1.14)

    with some 0h2,C1(¯Ω). The corresponding result can be stated as follows.

    Theorem 1.3. Let Ω be convex and the extra assumptions (1.12)–(1.14) hold. Then, for the global generalized solution of the initial-boundary value problem (1.5) from Theorem 1.2, there exists a null set N(0,) such that

    u(,t)L1+v(,t)v()L20,as(0,)Nt, (1.15)

    where v denotes the solution of the boundary value problem

    {0=Δvv+h2,,xΩ,vν=0,xΩ. (1.16)

    Let's state of the art and strategy of our proofs:

    {The main objective of this paper is to present that κuv with κ>0 has a regularization effect on the solution of the problem (1.5) in the n-dimensional settings with n3. Precisely, we prove that the initial-boundary value problem (1.5) possesses a global generalized solution for any χ<χ1 (given in (1.11)), where χ1 is greater than χ0 (given in (1.4)).} Note that the condition that χ<χ0 is required in [37] to guarantee the global existence of generalized solutions to the initial-boundary value problem of (1.3). Usually, to get the generalized solvability, one should seek an appropriate generalized framework, and thereby obtain the global existence of generalized solutions by an appropriate approximation procedure. Here, our novelty of analysis consists of further developing the generalized framework given in [49] by Winkler for the logarithmic Keller-Segel system with signal absorption, and using the coupled quantity upεvqε with some p,q(0,1) introduced in [37] by Lankeit and Winkler for the model (1.3) to derive the uniform in ε bound of

    T0Ωup+1εvq1εdxdt,T>0,

    see Lemma 3.1, in which, in contrast with [37], we must deal with the additional term

    2pκT0Ωupεvq+1εdxdt.

    To this end, we need some additional assumption on p to obtain some now uniform in ε estimates by using the benefit of κuεvε, see Lemma 2.2. After this, by taking advantage of Lemma 2.2 again, we address ourselvers to the uniform in ε bound of uεLr(Ω×(0,T)) with some r>1, see Lemma 3.3. We would like to remark that, in contrast with [37], the additional condition that q<4n will be required in our situation when n5, which results in that [37,Lemma 5.1] is no longer in force. Here, our novelty of analysis contains establishing some fragile estimates, by which we can get the core of our requirement (1.11) on χ. Based on the above processes, we can employ the result that uεu a.e. in Ω×(0,T) and the Vitali convergence theorem to get the strong convergence of {uε} in L1(Ω×(0,T)), see Lemma 3.5. Moreover, by establishing a series of uniform a-priori estimates as desired, we can get the global existence of generalized solutions to the initial-boundary value problem (1.5) via passing to limit, and subsequently complete the proof of Theorem 1.2 in Section 3.

    The second objective of this paper is to show the large-time behavior of such generalized solutions, under the additional assumptions on h1 and h2. Here, our novelty of analysis consists of tracking the time evolution of the combinational functional of the form

    Eε(t):=Ω|vεv|2+μuεdx,t>0,

    with some μ>0, where v is a classical solution of the boundary value problem (1.16). A fragile calculation yields that

    Ω|vεv|2(,t)+uε(,t)dx0astuniformlyinε,

    see Lemma 4.5 for details. This, together with the Fubini-Tonelli theorem and Fatou's lemma, ensures the desired results in Theorem 1.3, see Section 4 for details.

    To structure the generalized solution of the initial-boundary value problem (1.5) by an approximation procedure, for any ε(0,1) we shall consider the following approximate problem

    {uεt=Δuεχ(uε1+εuεlnvε)κuεvε+h1,xΩ,t>0,vεt=Δvεvε+uε+h2,xΩ,t>0,uεν=vεν=0,xΩ,t>0,uε(x,0)=u0(x),vε(x,0)=v0(x),xΩ, (2.1)

    and then first obtain the following.

    Lemma 2.1. Assume that the assumptions (1.6) and (1.7) hold. For each ε(0,1) and any χ>0, there exists a unique pair (uε,vε) of positive functions with the properties that for any T>0

    {uεC0(¯Ω×[0,T])C2,1(¯Ω×(0,T)),vεr>nC0(0,T;W1,r(¯Ω))C2,1(¯Ω×(0,T)),

    such that (uε,vε) solves the approximate problem (2.1) classically in Ω×[0,T). Moreover, the following two statements are true:

    vε(,t)etinfxΩv0(x),t>0, (2.2)

    and

    uε(,t)L1+t0Ωuεvε(,s)dxdsC(1+t),t>0, (2.3)

    for some C>0, independent of ε.

    Proof. An application of the well-known strategy, as in [34,52], implies that there exist time Tmax,ε(0,] and a unique pair (uε,vε) of positive functions with the properties that

    {uεC0(¯Ω×[0,Tmax,ε))C2,1(¯Ω×(0,Tmax,ε)),vεr>nC0([0,Tmax,ε);W1,r(¯Ω))C2,1(¯Ω×(0,Tmax,ε)),

    such that (uε,vε) solves the approximate problem (1.5) classically in Ω×[0,Tmax,ε). Moreover, if Tmax,ε<, then for any q>n

    lim suptTmax,ε(uε(,t)L+vε(,t)Lq+v1ε(,t)L)=. (2.4)

    To show that Tmax,ε=, let us start with the pointwise lower bound for the solution component vε and the bound of uεL1. Indeed, according to the variation-of-constants formula for vε

    vε(,t)=et(Δ1)v0+t0e(ts)(Δ1)uε(,s)ds+t0e(ts)(Δ1)h2(,s)ds, (2.5)

    the comparison principle for the Neumann problem associated with the heat equation and the facts that h20 and uε>0, we have

    vε(,t)et(Δ1)v0etinfxΩv0(x),t(0,Tmax,ε). (2.6)

    To get the bound of uεL1, we integrate the first equation in (2.1) over Ω to obtain

    ddtΩuεdx+κΩuεvεdx=Ωh1dx, (2.7)

    which, integrating over [0,t], implies that

    uε(,t)L1+κt0Ωuεvεdxdsu0L1+h1L(Ω×(0,))t,t(0,Tmax,ε). (2.8)

    We now estimate uεL. Indeed, according to the variation-of-constants formula for uε, we can infer from the maximum principle and the nonnegativity of κuεvε that

    uε(,t)et(Δ1)u0+t0e(ts)(Δ1){χ(uε1+εuεlnvε)+uε+h1}dset(Δ1)u0L+t0e(ts)(Δ1){χ(uε1+εuεlnvε)+uε+h1}Lds,

    which, combining with the properties of the Neumann heat semigroup (see [53,54]), (2.6) and the nonnegativity of uε, leads to that for any r>n

    uε(,t)Lu0L+Cεett0(1+(ts)12n2r+(ts)n2r)e(ts)(vεLr+uε+h1Lr)ds. (2.9)

    By means of the interpolation inequality and (2.8), we obtain

    uεLruε1rL1uε11rLC(1+t1r)uε11rL,t(0,Tmax,ε). (2.10)

    In addition, the application of the properties of the Neumann heat semigroup to (2.5) entails that for any r>n

    vε(,t)Lret(Δ1)v0Lr+t0e(ts)(Δ1)(uε+h2)LrdsCv0Lr+Ct0(1+(ts)12)e(ts)uε+h2Lrds.

    Using (2.10) and letting K(T):=supt(0,T)uε(,t)L for any T(0,Tmax,ε), we arrive at

    supt(0,T)vε(,t)LrC+C(1+T1r)K11r(T). (2.11)

    Substituting this and (2.10) into (2.9), we conclude that

    K(T)C+CεeT(1+(1+T1r)K11r(T)).

    Since 0<11r<1, an application of Young's inequality implies that K(T)Cε(T). Hence for any T(0,Tmax,ε) we infer that

    uε(,t)LCε(T),t(0,T).

    This, together with (2.6) and (2.11), establishes a contradiction to (2.4) and thereby ensures that actually we must have Tmax,ε=.

    Moreover, using (2.6) and (2.8) with Tmax,ε=, the assertions (2.2) and (2.3) hold as desired.

    The estimate (2.3) turns out to be sufficient for the derivation of the bound of vεL2.

    Lemma 2.2. Let r[1,nn2). Assume that (uε,vε) is taken from Lemma 2.1. Then there exists C=C(r)>0, with the property that

    vε(,t)LrC(1+t),t>0andε(0,1). (2.12)

    Moreover, there is C>0 such that

    Ωv2ε(,t)dxC(1+t),t>0andε(0,1), (2.13)

    and

    t0vε(,s)2L2dsC(1+t),t>0andε(0,1). (2.14)

    Proof. Recalling (2.5) and invoking the properties of the Neumann heat semigroup (see [53,54]), for any r[1,nn2) and t>0 we have

    vε(,t)LrC1v0Lr+C1t0(1+(ts)n2(11r))e(ts)(uε(,s)L1+h2(,s)L1)ds,

    which, using (1.7) and (2.3), leads to (2.12) as desired.

    Next, we test the second equation in (2.1) by vε and use the integration by part to get

    12ddtΩv2εdx+Ω|vε|2dx+Ωv2εdx=Ωuεvεdx+Ωh2vεdx,t>0.

    By means of (1.7), the applications of Young's inequality and Hölder's inequality yield

    Ωh2vεdx12vε2L2+12h22L212vε2L2+C3.

    Invoking this we arrive at

    ddtΩv2εdx+2Ω|vε|2dx+Ωv2εdx2Ωuεvεdx+2C3,t>0. (2.15)

    Integrating (2.15) over [0,t] we obtain that

    Ωv2ε(,t)dx+2t0Ω|vε|2dxds+t0Ωv2εdxdsΩv20dx+2t0Ωuεvεdxds+2C3t,

    which, combined with (2.3), ensures (2.13) and (2.14).

    In comparison to Lemma 2.2, deriving the bound for t0Ω|uε|2dxds seems to be more delicate, due to the presence of the taxis-type term in the first equation in (2.1). Motivated by [49,51], we resort to estimating ln(uε+1) instead.

    Lemma 2.3. Let (uε,vε) be given in Lemma 2.1. There exists C>0, with the property that

    t0ln(uε(,s)+1)2L2dsC(1+t)et,t>0andε(0,1). (2.16)

    Proof. Multiplying the first equation in (2.1) by 11+uε and using the integration by parts, we see that

    ddtΩln(1+uε)dx=Ω11+uε{Δuεχ(uε1+εuεlnvε)κuεvε+h1}=Ω|uε|2(uε+1)2dxχΩuε(uε+1)2(1+εuε)uεlnvεdxΩκuεvε1+uεdx+Ωh11+uεdx,t>0.

    Invoking this, an application of Young's inequality yields that

    12Ω|uε|2(uε+1)2dxddtΩln(1+uε)dx+χ22Ωu2ε(uε+1)2(εuε+1)2|lnvε|2dx+Ωκuεvε1+uεdxΩh11+uεdxddtΩln(1+uε)dx+χ22Ω|lnvε|2dx+κΩvεdx,t>0,

    which, using (2.2) and integrating in time, leads to

    12t0Ω|uε|2(uε+1)2dxdsΩln(1+uε(,t))dx+Cett0Ω|vε|2dxds+κt0Ωvεdxds,t>0.

    This, together with (2.12) and (2.14), ensures there exists C>0 such that

    12t0Ω|uε|2(uε+1)2dxdsΩln(1+uε(,t))dx+C(1+t)et,t>0.

    Since ζln(1+ζ)0 for any ζ0, we obtain that

    12t0Ω|uε|2(uε+1)2dxdsΩuε(,t)dx+C(1+t)et,t>0,

    which, in view of (2.3), entails (2.16).

    To obtain the desired integrability for u in Definition 1.1, a crucial step in our analysis will consist of deriving the uniform spatio-temporal integrability of uε. To this end, we further develop the framework presented in [37,Lemma 5.1] to get some essential a-priori estimates for (2.1).

    Lemma 3.1. Let n3, p(0,1) satisfying p<1χ2 and p<4n, and q(q(p),q+(p)) with

    q±(p):=1p2(1±1pχ2). (3.1)

    Assume that (uε,vε) is given in Lemma 2.1. Then, there exists C>0, with the property that for any t>0 and ε(0,1)

    t0Ωupεvq2ε|vε|2dxds+t0Ωup2εvqε|uε|2dxds+t0Ωup+1εvq1εdxdsC(1+t)4. (3.2)

    Proof. Using the facts that uε>0 and vε>0, we have

    ddtΩupεvqεdx=pΩup1εvqεtuεdx+qΩupεvq1εtvεdx=pΩup1εvqε{Δuεχ(uε1+εuεlnvε)κuεvε+h1}dx+qΩupεvq1ε(Δvεvε+uε+h2)dx,t>0,

    which, using the integration by parts, leads to

    ddtΩupεvqεdx=p(1p)Ωup2εvqε|uε|2dxΩ(2pq+p(1p)χ1+εuε)up1εvq1εuεvεdx+Ω(pqχ1+εuε+q(1q))upεvq2ε|vε|2dxpκΩupεvq+1εdx+pΩup1εvqεh1dxqΩupεvqεdx+qΩup+1εvq1εdx+qΩupεvq1εh2dx,t>0.

    Thanks to the nonnegativity of h1, h2, uε and vε, we arrive at

    ddtΩupεvqεdxp(1p)Ωup2εvqε|uε|2dxΩ(2pq+p(1p)χ1+εuε)up1εvq1uεvεdx+Ω(pqχ1+εuε+q(1q))upεvq2ε|vε|2dx+qΩup+1εvq1εdxpκΩupεvq+1εdxqΩupεvqεdx=:P1+P2+P3+P4+P5+P6,t>0.

    A straightforward rearrangement in the first three integrands on the right entails

    P1+P2+P3=p(1p)Ω|up21εvq2εuε2q+(1p)χ1+εuε2(1p)up2εvq21εvε|2dx+Ω{q(pχ1+εuε+1q)p(2q+(1p)χ1+εuε)24(1p)}upεvq2ε|vε|2dx=Ωq(pχ1+εuε+1q)|up2εvq21εvεp(2q+(1p)χ1+εuε)2q(pχ1+εuε+1q)vp21vq2εuε|dx+Ω{p(1p)p2(2q+(1p)χ1+εuε)24q(pχ1+εuε+1q)}up2εvqε|uε|2dx.

    Invoking this, we obtain

    2ddtΩupεvqεdx+2pκΩupεvq+1εdx+2qΩupεvqεdxΩc1(x,t)upεvq2ε|vε|2dx+Ωc2(x,t)up2εvqε|uε|2dx+2qΩup+1εvq1εdx,

    where

    c1(x,t):=q(pχ1+εuε+1q)p(2q+(1p)χ1+εuε)24(1p),c2(x,t):=p(1p)p2(2q+(1p)χ1+εuε)24q(pχ1+εuε+1q).

    We also note that the assumption (3.1) on q warrants that q<1p and

    4(1p)q4q2p(1p)2χ2=4(qq+(p))(qq(p))>0,

    which, due to p(0,1), ensures

    4(1p)c1(x,t)=4q(1p)4q2p(1p)2χ2(1+εuε)24(1p)q4q2p(1p)2χ2>0.

    Similarly, we have

    4q(pχ+1q)p1c2(x,t)4q(pχ1+εuε+1q)p1c2(x,t)4(1p)q4q2p(1p)2χ2>0.

    Collecting these, there exist two positive constants ^c1 and ^c2, denoted by

    ^c1:=q(pχ+1q)p(2q+(1p)χ)24(1p),^c2:=p(1p)p2(2q+(1p)χ)24q(pχ+1q),

    such that

    2ddtΩupεvqεdx+2pκΩupεvq+1εdx+2qΩupεvqεdx^c1Ωupεvq2ε|vε|2dx+^c2Ωup2εvqε|uε|2dx+2qΩup+1εvq1εdx.

    Hence, an integration in time shows

    ^c1t0Ωupεvq2ε|vε|2dxds+^c2t0Ωup2εvqε|uε|2dxds+2qt0Ωup+1εvq1εdxds2Ωupεvqε(x,t)dx+2pκt0Ωupεvq+1εdxds+2qt0Ωupεvqεdxds. (3.3)

    Using Hölder's inequality and the fact that q<1p again, we have

    ΩupεvqεdxupεL1pvqεL11p=uεpL1vεqLq1pCuεpL1vεqL2,

    which, together with (2.3) and (2.13), implies that there exists C>0, independent of ε, such that

    2Ωupεvqε(x,t)dx+2qt0ΩupεvqεdxdsC(1+t)p+q2+1C(1+t)2,t>0. (3.4)

    Similarly, an application of Hölder's inequality and Young's inequality yields that

    2pκt0Ωupεvq+1εdxds2pκt0(Ωup+1εvq1εdx)pp+1(Ωv2p+q+1εdx)1p+1dsqt0Ωup+1εvq1εdxds+Ct0Ωv2p+q+1εdxds. (3.5)

    Since q<1p and 0<p<1, we have 2p+q+1<2+p<3. In the case that n=3, thanks to nn2=3, it follows from (2.12) that there exists C>0 such that

    t0Ωv2p+q+1εdxdsCt0(1+s)2p+q+1dsC(1+t)4,t>0. (3.6)

    In the case that n4, if 2p+q+12, according to Young's inequality and (2.13) there exists C>0 such that

    t0Ωv2p+q+1εdxdst0Ωv2ε+1dxdsC(1+t)2,t>0. (3.7)

    In addition, if 2<2p+q+12nn2, then we can infer from the Gagliardo-Nirenberg inequality that

    vεL2p+q+1C(vε1θL2vεθL2+vεL2),θ:=n(2p+q1)2(2p+q+1),

    which, combined with (2.13), leads to

    vε(,t)2p+q+1L2p+q+1C(1+t)2p+q+1(vε(,t)n2(2p+q1)L2+1),t>0.

    Recalling the fact that 2<2p+q+1, we have 0<n2(2p+q1)<n2p<2 due to p+q<1 and p<4n, and thereby infer from Young's inequality and (2.14) that

    t0Ωv2p+q+1εdxdsC(1+t)p+q2+12t0(vε(,s)2L2+1)dsC(1+t)3,t>0. (3.8)

    Collecting (3.6)–(3.8), it follows from (3.5) that, whenever n3,

    2pκt0Ωupεvq+1εdxdsqt0Ωup+1εvq1εdxds+C(1+t)4,t>0. (3.9)

    Substituting (3.4) and (3.9) into (3.3), we have (3.2) as desired, due to the facts that ^c1,^c2>0 and q<1.

    Indeed, using Lemma 3.1, we can get the bound of uε in some reflexive Lr spaces. To achieve it, we need to identify the minimal possible choice of an integrability exponent arising in (3.19) below, which will form the core of our requirement (1.11) on χ.

    Lemma 3.2. Let χ>0, and for p(0,min{1,4n,1χ2}) let q±(p) be defined in (3.1). If n{3,4}, then

    infp(0,1),p<1χ2q(q(p),q+(p))1qp={1ifχ1,χifχ(1,2),1+χ24ifχ2, (3.10)

    and if n5, then

    infp(0,min{4n,1χ2})q(q(p),q+(p))1qp={n8+12(n812)14nχ2ifχ2nn+4,χifχ(2nn+4,2),1+χ24ifχ2. (3.11)

    Proof. If n{3,4}, then p(0,min{1,1χ2}). In this case, the assertion (3.11) directly follows from [37,Lemma 5.1].

    If n5, then we have p(0,min{4n,1χ2}). A straightforward calculation shows that

    I(χ):=infp(0,min{4n,1χ2})q(q(p),q+(p))1qp=infp(0,min{4n,1χ2})1+p(1p)1pχ22p,

    and that I(χ)1 for any χ>0. Setting ξ:=1pχ2, we get that

    ξ{(14nχ2,1), if χn2,(0,1) if χ>n2. (3.12)

    Note that p=1ξ2χ2, simple computation shows that

    1+p(1p)1pχ22p=12(χ21+ξ+1+ξ)=:g(ξ).

    Accordingly, we have

    I(χ)={infξ[14nχ2,1]g(ξ), if χn2,infξ[0,1]g(ξ), if χ>n2. (3.13)

    As

    g(ξ):=12(χ2(1+ξ)2+1),

    which implies that g(ξ) is strictly monotonely decreasing in the interval [0,χ1] and strictly monotonely increasing in the intrerval [χ1,+), correspondingly, we have

    I(χ)={g(χ1), if χ(1+14nχ2,2](0,n2],g(14nχ2), if χ(0,1+14nχ2](0,n2],g(1), if χ(2,n2],n>16,g(χ1), if χ[1,2](n2,+)g(1), if χ(2,+)(n2,+). (3.14)

    Direct calculation shows that

    χ(1+14nχ2,2](0,n2]χ(2nn+4,min{2,n2}],

    and

    χ(0,1+14nχ2](0,n2]χ(0,2nn+4],

    moreover,

    (2,+)(n2,+)=(2,+)ifn16

    and

    (2,n2]((2,+)(n2,+))=(2,+)ifn>16,

    as well as

    (2nn+4,min{2,n2}]([1,2](n2,+))=(2nn+4,2].

    Thus, it follows from (3.14) that

    I(χ)={g(14nχ2), if χ(0,2nn+4],g(χ1), if χ(2nn+4,2],g(1), if χ(2,+) (3.15)

    Note that

    g(14nχ2)=12(1+n4+(1n4)14nχ2),
    g(χ1)=χ,

    and

    g(1)=1+14χ2,

    these together with (3.15) gives us the desired (3.11).

    Now under the assumptions on χ in Theorem 1.2, the interpolation argument, invoking Lemma 3.2, indeed bears fruit of the desired flavour.

    Lemma 3.3. Let p and q be taken from Lemma 3.1, and χ satisfy (1.11). Then, for (uε,vε) given in Lemma 2.1, there exist r>1 and C>0, with the property that

    t0Ωurε(,s)dxdsC(1+t)4,t>0andε(0,1). (3.16)

    Proof. According to (3.10), we infer that if n=3, then we have, as long as χ<23,

    infp(0,1),p<1χ2q(q(p),q+(p))1qp<4,

    with q(p) and q+(p) given by Lemma 3.1, which ensures that we can find p(0,min{1,1χ2} and q(q(p),q+(p)) such that

    (1q)p<4. (3.17)

    Similarly, we can deduce from (3.10) and (3.11) that

    infp(0,min{4n,1χ2})q(q(p),q+(p))1qp<2+4n,

    for any n4, as long as χ<21+4n, which also guarantees that we can choose p(0,min{4n,1χ2} and q(q(p),q+(p)) such that

    (1q)p<2+4n. (3.18)

    Fix p and q in (3.17) and (3.18) respectively, utilizing a continuity argument we can further pick r(1,1+p) sufficiently close to 1 such that

    β:=(1q)rp+1r{<4,n=3,<2+4n,n4.

    For such r, an application of Young's inequality yields that

    t0Ωurεdxdst0Ωup+1εvq1εdxds+t0Ωv(1q)r1+prεdxds,t>0,

    which, united (3.2), ensures that

    t0ΩurεdxdsC(1+t)4+t0Ωvβεdxds,t>0, (3.19)

    with β=(1q)r1+pr and C>0 independent of ϵ.

    In the case n=3, if β<nn2, i.e, β<3, then it follows from (2.12) that

    t0ΩvβεdxdsC(1+t)β+1C(1+t)4,t>0,

    thus we have from (3.19) that

    t0ΩurεdxdsC(1+t)4,t>0,

    namely, (3.16) is valid. Meanwhile, if β[3,4), then γ:=3β62[32,3) and 6(βγ)6γ=2, the Gagliardo-Nirenberg inequality implies that

    vεβLβC(vεβ2Lγvε2L2+vεβL2),t>0,

    which, together with (2.12)–(2.14), gives us

    t0ΩvβεdxdsCt0(1+s)β2vε(,s)2L2+(1+s)β2dsC{(1+t)β1+(1+t)β2+1},t>0.

    This, combined with (3.19), also entails (3.16) due to β<4.

    In the case n4, if β2, then the Young inequality and (2.13) entail that

    t0Ωvβεdxdst0Ωv2ε+1dxdsC(1+t)2,t>0.

    This, together with (3.19), also guarantees the validity of (3.16). If β(2,2+4n), then n(β2)2<2. Applications of the Gagliardo-Nirenberg inequality and the Young inequality imply that

    vεβLβC(vεβn(β2)2L2vεn(β2)2L2+vεβL2)C(vεβn(β2)2L2vε2L2+vεβL2+1),t>0.

    In this case, we can infer from (2.13) and (2.14) that

    t0ΩvβεdxdsC(1+t)β2+1,t>0,

    which, combined (3.19), implies that (3.16) is also valid due to β<2+4n.

    As a final preparation for our limit procedure, we establish some regularity features of the time derivatives in (2.1).

    Lemma 3.4. Let (uε,vε) be established in Lemma 2.1. For any T>0, there exists C(T)>0, with the property that for r>n

    T0vεs(,s)2(W1,r)dsC(T)foranyε(0,1), (3.20)
    T0sln(uε(,s)+1)(W1,r)dsC(T)foranyε(0,1). (3.21)

    Proof. On the basis of the second equation in (2.1), we obtain from the integration by parts and Hölder's inequality that for any φC(¯Ω) and t>0

    |vεt,φ|vεL2φL2+vεL2φL2+uεL1φL+h2LφL1,

    which, combined with the Sobolev embedding theorem, entails that for any r>n there exists C>0 independent of ε such that for any t>0

    |vεt,φ|C(vεH1+uεL1+h2L)φW1,r.

    This, in view of (2.3), (2.13), (2.14) and (1.7), in turn ensures (3.20).

    Next, multiplying the first equation in (2.1) by φuε+1 for any φC(¯Ω) we have for any t>0

    Ωtln(1+uε)φdx=Ω|uε|2φ(uε+1)2dxχΩuε(uεlnvε)φ(uε+1)2(1+εuε)dxΩuεφ1+uεdx+χΩuεlnvεφ(1+uε)(1+εuε)dxκΩuεvεφ1+uεdx+Ωh1φ1+uεdx,

    which, by using Hölder's inequality, Young's inequality and Sobolev's inequality, entails that for any r>n there exists C>0 such that for any t>0

    |tln(1+uε),φ|φLΩ|uε|2(uε+1)2dx+χ(Ω|uε|2(uε+1)2dx)12lnvεL2φL+(Ω|uε|2(uε+1)2dx)12φL2+χlnvεL2φL2+κvεL2φL2+h1LφL1CφW1,r(Ω|uε|2(uε+1)2dx+lnvε2L2+vεL2+h1L+1).

    After an integration in time, we infer from (1.7), (2.13), (2.14), (2.16) and (2.2) that (3.21) holds as desired.

    On the basis of the standard compactness arguments, we can find a candidate (u,v) for a generalized solution.

    Lemma 3.5. Let (uε,vε) be taken from Lemma 2.1. Then, for any T>0 there exist functions u0 and v>0 defined on Ω×(0,T) and a sequence {εj}j=1(0,1) such that εj0 as j, with the properties that as ε=εj0,

    ln(1+uε)ln(1+u)inL2(Ω×(0,T)), (3.22)
    ln(1+uε)ln(1+u)inL2(0,T;H1(Ω)), (3.23)
    uεua.e.inΩ×(0,T), (3.24)
    uεuinL1(Ω×(0,T)), (3.25)
    uεuinLr(Ω×(0,T)), (3.26)
    vεvinL2(0,T;Lq(Ω)), (3.27)
    v1εv1inL2(0,T;Lq(Ω)), (3.28)
    lnvεlnvinL2(0,T;Lq(Ω)), (3.29)
    vεvinL(0,T;L2(Ω)), (3.30)
    vεvinL2(0,T;H1(Ω)), (3.31)

    where q<2nn2.

    Proof. Since 12ln2(1+ζ)ζ for any ζ0, we infer from the bounds (2.16) and (2.3) that

    t0ln(1+uε)2H1dsCt0(uε(,s)L1+ln(1+uε)(,s)2L2)dsC(T),t(0,T).

    Based on this, the bound (3.21), combined with the Aubin-Lions compactness theorem [55], implies that there exist a subsequence of {εj}j=1 (still expressed as {εj}j=1) and a function wL2(0,T;H1(Ω)), fulfilling that as ε=εj0,

    ln(uε+1)w,ln(uε+1)w inL2(Ω×(0,T),

    and thereby

    ln(uε+1)wanduεew1 a.e. inΩ×(0,T).

    Hence, denoting u=ew1 and using the bound (2.3) again, we obtain the assertions (3.22)–(3.24). Due to (3.24), according to the uniform integrability property implied by Lemma 3.3 we may apply the Vitali convergence theorem to get that in fact (3.25) also holds. Meanwhile, using (3.24) and invoking Lemma 3.3, we arrive at (3.26).

    According to the bounds (2.13), (2.14) and (3.20), and the Sobolev embedding theorem, a standard subsequence extraction procedure resorting to the Aubin-Lions compactness theorem (see [55]) entails model (3.27) immediately. Due to (2.2), we have

    v1ε(,t)L2C(T),t(0,T),

    and also infer from (2.14) that

    t0v1ε(,s)2L2dsC(T),t(0,T).

    Since (v1ε)t=v2ε(Δvεvε+uε+h2), similar to (3.20), using (2.2) again we get

    T0v1εs(,s)2(W1,r)dsC(T),r>n.

    Hence, invoking the Aubin-Lions compactness theorem ([55]), there exists a subsequence of {εj}j=1 (still expressed as {εj}j=1) such that (3.28) holds as desired, as ε=εj0. Similarly, (3.29) also holds. On the other hand, using the bounds (2.13) and (2.14) again yields the last two assertions in lemma.

    Up to now, our knowledge on approximation of (u,v) by (uε,vε) is enough to pass to the limit ε=εj0 in the weak formulation of the second equation in the approximate problem (2.1), which also show that v is indeed a weak solution of the respective sub-problem of (1.5) in the sense of Definition 1.1.

    Lemma 3.6. Let u and v be given in Lemma 3.5. For any T>0, the identity (1.10) in Definition 1.1 is valid for any φ(x,t)L(Ω×(0,T))L2(0,T;H1(Ω)) having compact support in ¯Ω×[0,T) with φtL2(Ω×(0,T)).

    Proof. For each φ from the class indicated in (1.10), it follows from (3.25) and the Lebesgue dominated convergence theorem that there exists a subsequence of {εj}j=1 (still expressed as \{\varepsilon_j\}_{j = 1}^\infty ) such that for any T > 0 , as \varepsilon = \varepsilon_j\rightarrow0 ,

    \begin{align*} \int_{0}^T \int_{\Omega} u_{\varepsilon} \varphi dx ds\rightarrow \int_{0}^T \int_{\Omega} u \varphi dxds. \end{align*}

    Hence, we can take the limit \varepsilon = \varepsilon_j\rightarrow0 on the second equation in (2.1) in the weak sense by employing Lemma 3.5. Moreover, the functions u and v obtained in Lemma 3.5 satisfy the identity (1.10) in Definition 1.1.

    To take the limit also in the first equation in the approximate problem (2.1) in an appropriate manner, we shall obtain the strongly convergence of \nabla\ln v_{\varepsilon} in L^2(\Omega\times(0, T)) for any T > 0 .

    Lemma 3.7. Let (u_\varepsilon, v_\varepsilon) be described in Lemma 2.1, and let u and v be established in Lemma 3.5. Then there exists a subsequence of \{\varepsilon_j\}_{j = 1}^\infty (still expressed as \{\varepsilon_j\}_{j = 1}^\infty ) such that for any T > 0 , as \varepsilon = \varepsilon_j\rightarrow0 ,

    \begin{equation} \nabla\ln v_\varepsilon\rightarrow \nabla\ln v \quad {\mathrm{in}}\quad L^2\big(\Omega\times(0,T)\big). \end{equation} (3.32)

    Proof. We can adopt a strategy similar to [49,Lemma 2.10] to get (3.32) as desired.

    Invoking Lemma 3.7, we can present the validity of (1.9) in Definition 1.1.

    Lemma 3.8. Let u and v be given in Lemma 3.5. For any T > 0 , the inequality (1.9) in Definition 1.1 is valid for each nonnegative \varphi(x, t)\in \mathcal{C}_0^\infty(\overline{\Omega}\times[0, T)) .

    Proof. Testing the first equation in (2.1) by \frac{\varphi}{1+u_\varepsilon} with 0\leq\varphi\in C_0^\infty(\Omega\times [0, T)) , we have

    \begin{align*} & \int_{0}^T \int_{\Omega}\left|\nabla \ln \left(u_{\varepsilon}+1\right)\right|^{2} \varphi dxdt\\ & = -\int_{0}^T \int_{\Omega} \ln \left(u_{\varepsilon}+1\right) \varphi_{t} dxdt-\int_{\Omega} \ln \left(u_{0}+1\right) \varphi(\cdot, 0)dx+\int_{0}^T \int_{\Omega} \nabla \ln \left(u_{\varepsilon}+1\right) \cdot \nabla \varphi dxdt\\ &\quad+\chi\int_{0}^T \int_{\Omega} \frac{u_{\varepsilon} }{(u_{\varepsilon}+1)(1+\varepsilon u_\varepsilon)}\left(\nabla \ln \left(u_{\varepsilon}+1\right) \cdot \nabla \ln v_{\varepsilon}\right) \varphi dxdt\\ &\quad-\chi\int_{0}^T \int_{\Omega} \frac{u_{\varepsilon} }{(u_{\varepsilon}+1)(1+\varepsilon u_\varepsilon)} \nabla \ln v_{\varepsilon} \cdot \nabla \varphi dxdt\\ &\quad+\kappa\int_0^T\int_\Omega\frac{ u_\varepsilon v_\varepsilon}{1+u_\varepsilon}\varphi dxdt-\int_0^T\int_\Omega\frac{h_1}{1+u_\varepsilon}\varphi dxdt. \end{align*}

    We conclude from (3.23) that as \varepsilon = \varepsilon_j\rightarrow0 ,

    \begin{align*} \int_{0}^T \int_{\Omega} \ln \left(u_{\varepsilon}+1\right) \varphi_{t} dxdt \rightarrow \int_{0}^T \int_{\Omega} \ln (u+1) \varphi_{t}dxdt,\\ \int_{0}^T \int_{\Omega} \nabla \ln \left(u_{\varepsilon}+1\right) \cdot \nabla \varphi dxdt \rightarrow \int_{0}^T \int_{\Omega} \nabla \ln (u+1) \cdot \nabla \varphi dxdt. \end{align*}

    Since \frac{u_{\varepsilon} }{(u_{\varepsilon}+1)(1+\varepsilon u_\varepsilon)} \rightarrow \frac{u}{u+1} a.e. in \Omega\times(0, T) as \varepsilon = \varepsilon_j\rightarrow0 , we infer from (3.32) and [51,Lemma A.4] that, as \varepsilon = \varepsilon_j\rightarrow0 ,

    \begin{align*} \frac{u_{\varepsilon}}{(u_{\varepsilon}+1)(1+\varepsilon u_\varepsilon)} \nabla \ln v_{\varepsilon} \rightarrow \frac{u}{u+1} \nabla \ln v \quad {\text { in }} L^{2}(\Omega \times(0, T)), \end{align*}

    which, combined with (3.23), further implies that, as \varepsilon = \varepsilon_j\rightarrow0 ,

    \begin{align*} \int_{0}^T \int_{\Omega} \frac{u_{\varepsilon} }{(u_{\varepsilon}+1)(1+\varepsilon u_\varepsilon)}\left(\nabla \ln \left(u_{\varepsilon}+1\right) \cdot \nabla \ln v_{\varepsilon}\right) \varphi dxdt\\ \rightarrow \int_{0}^T \int_{\Omega} \frac{u}{u+1}\left(\nabla \ln \left(u+1\right) \cdot \nabla \ln v\right) \varphi dxdt \end{align*}

    and

    \begin{align*} \int_{0}^T \int_{\Omega} \frac{u_{\varepsilon}}{(u_{\varepsilon}+1)(1+\varepsilon u_\varepsilon)} \nabla \ln v_{\varepsilon} \cdot \nabla \varphi dxdt\rightarrow \int_{0}^T \int_{\Omega} \frac{u}{u+1} \nabla \ln v \cdot \nabla \varphi dxdt. \end{align*}

    Similarly, we obtain that \varepsilon = \varepsilon_j\rightarrow0 ,

    \begin{align*} \kappa\int_0^T\int_\Omega\frac{u_\varepsilon v_\varepsilon}{1+u_\varepsilon}\varphi dxdt\rightarrow \kappa\int_0^T\int_\Omega\frac{uv}{1+u}\varphi dxdt. \end{align*}

    By using the Lebesgue dominated convergence theorem, we have

    \begin{align*} \int_0^T\int_\Omega\frac{h_1}{1+u_\varepsilon}\varphi dxdt\rightarrow \int_0^T\int_\Omega\frac{h_1}{1+u}\varphi dxdt. \end{align*}

    Invoking (3.23), an application of the weak lower semicontinuity of the norm implies

    \begin{equation*} \int_{0}^T \int_{\Omega}|\nabla \ln (u+1)|^{2} \varphi dxdt\leq \liminf \limits_{\varepsilon = \varepsilon_{j} \searrow 0} \int_{0}^T \int_{\Omega}\left|\nabla \ln \left(u_{\varepsilon}+1\right)\right|^{2} \varphi dxdt. \end{equation*}

    Hence, collecting these, (1.9) holds as desired.

    We are now in the position to prove Theorem 1.2.

    Proof of Theorem 1.2. In fact, we only need to combine Lemma 3.6 with Lemma 3.8.

    In this section, we will investigate the large-time behavior of the generalized solution (u, v) determined in Theorem 1.2, under the additional assumptions (1.12)–(1.14). To achieve this, we begin with the following pointwise lower bound for the solution component v_\varepsilon , which will play a key role in the sequel.

    Lemma 4.1. Let (u_\varepsilon, v_\varepsilon) come from Lemma 2.1, and let (1.12) be in force. Under the additional assumption that \Omega is convex, then there exists c_1 > 0 , independent of t and \varepsilon , fulfilling that

    \begin{align} v_\varepsilon(x,t)\geq c_1, \quad x\in\Omega,\,\, t > 0. \end{align} (4.1)

    Proof. It immediately follows from [50,Corollary 3.1].

    Let us state a straightforward consequence of Lemma 4.1.

    Lemma 4.2. Let all the assumptions in Lemma 4.1 be fulfilled. Then there exists a positive constant c_2 , with the property that

    \begin{align} \int_\Omega (u_\varepsilon+ v_\varepsilon^2)(\cdot,t)dx+\int_t^{t+1}\int_\Omega(u_\varepsilon v_\varepsilon +|\nabla v_\varepsilon|^2)(\cdot,s)dxds\leq c_2,\quad t > 0\,\,{\mathrm{and}}\,\, \varepsilon\in(0,1). \end{align} (4.2)

    Proof. Invoking (2.7) and (4.1), we arrive at

    \begin{align} \frac{d}{dt}\int_\Omega u_\varepsilon dx+\frac12\kappa \int_\Omega u_\varepsilon v_\varepsilon dx+\frac12\kappa c_1 \int_\Omega u_\varepsilon dx\leq \int_\Omega h_1dx\leq\|h_1\|_{L^\infty(\Omega\times(0,\infty))}|\Omega|,\,\,t > 0, \end{align} (4.3)

    where c_1 is given in (4.1). By taking \lambda: = \frac6\kappa , this, combined with (2.15), leads to

    \begin{align} \frac{d}{dt}\int_\Omega\lambda u_\varepsilon+ v_\varepsilon^2 dx+\int_\Omega 3 c_1 u_\varepsilon+v_\varepsilon^2 dx+\int_\Omega u_\varepsilon v_\varepsilon +2|\nabla v_\varepsilon|^2 dx\leq C_1,\,\,t > 0. \end{align} (4.4)

    Setting y(t): = \int_\Omega\lambda u_\varepsilon+ v_\varepsilon^2 dx , we get

    y'(t)+\min\{3c_1\lambda^{-1},1\}y(t)\leq C_1,\quad t > 0,

    which, employing a standard ODE argument, warrants that

    \begin{align} \int_\Omega(\lambda u_\varepsilon+ v_\varepsilon^2)(\cdot,t) dx\leq C_2,\quad t > 0. \end{align} (4.5)

    Using this and integrating (4.4) over [t, t+1] , it follows that for any t > 0

    \begin{align*} \int_\Omega (\lambda u_\varepsilon+ v_\varepsilon^2)(\cdot,t+1)dx+\int_t^{t+1}\int_\Omega(u_\varepsilon v_\varepsilon +2|\nabla v_\varepsilon|^2)(\cdot,s)dxds\leq \int_\Omega (\lambda u_\varepsilon+ v_\varepsilon^2)(\cdot,t)dx +C_1, \end{align*}

    which, combined with (4.5), evidently ensures (4.2).

    To prove the long-time behavior in Theorem 1.3, we shall consider the Helmholtz problem (1.16).

    Lemma 4.3. For given 0\not\equiv h_{2, \infty}\in \mathcal{C}^1(\overline{\Omega}) , the problem (1.16) possesses a unique classical solution v_\infty with the property that v_\infty\in \mathcal{C}^{2+\theta}(\overline{\Omega}) for some \theta\in(0, 1) .

    Proof. The assertion directly follows from [56].

    We are also concerned with the decay in a linear differential inequality (see [50,Lemma 2.5]).

    Lemma 4.4. For \varepsilon\in(0, 1) , let y_\varepsilon\in \mathcal{C}^1([0, \infty)) be non-negative functions. If y_\varepsilon(0) is dependent of \varepsilon , and there exist a > 0 and the nonnegative function g(t)\in C([0, \infty))\cap L^\infty([0, \infty)) which satisfies

    \begin{align} \lim\limits_{t\rightarrow \infty}\int_t^{t+1}g(s)ds = 0 \end{align}

    such that

    \begin{equation*} y_\varepsilon'(t)+ay_\varepsilon(t)\leq g(t)\quad for\; all\quad t > 0\,\,{\mathrm{and}}\,\, \varepsilon\in(0,1), \end{equation*}

    then

    \begin{equation*} y_\varepsilon(t)\rightarrow0 \quad {\rm{as}}\quad t\rightarrow \infty\quad uniformly \;in\quad \varepsilon. \end{equation*}

    As a consequence, under the additional assumptions (1.13)–(1.14), a stronger result than Lemma 4.2 can be shown as follows.

    Lemma 4.5. Let all the assumptions in Lemma 4.2 hold, and let (1.13)–(1.14) be in force. Then we have

    \begin{align} \int_\Omega|v_\varepsilon-v_\infty|^2(\cdot,t)+ u_\varepsilon(\cdot,t) dx\rightarrow0\quad{\mathrm{as}}\,\,\, t\rightarrow \infty\quad {\mathrm{uniformly\,\, in}}\,\,\, \varepsilon, \end{align} (4.6)
    \begin{align} \int_t^{t+1}\int_\Omega|\nabla(v_\varepsilon-v_\infty)|^2(\cdot,s) dxds \rightarrow0\quad{\mathrm{as}}\,\,\, t\rightarrow \infty\quad {\mathrm{uniformly\,\, in}}\,\,\, \varepsilon, \end{align} (4.7)

    where v_\infty is a unique classical solution of (1.16).

    Proof. Set \widehat{v_\varepsilon}: = v_\varepsilon-v_\infty for convenience. Lemmas 2.1 and 4.3 imply that for fixed u_\varepsilon from Lemma 2.1, the initial-boundary value problem

    \begin{equation} \left\{ \begin{split} &\widehat{v_\varepsilon}_t = \Delta\widehat{v_\varepsilon}-\widehat{v_\varepsilon}+u_\varepsilon+h_2-h_{2,\infty},&x\in\Omega,\,\,t > 0,\\ &\nabla \widehat{v_\varepsilon}\cdot\nu = 0,&x\in\partial\Omega,\,\,t > 0,\\ &\widehat{v_\varepsilon}(x,0) = v_0(x)-v_\infty(x),&x\in\Omega, \end{split} \right. \end{equation} (4.8)

    admits a unique classical solution \widehat{v_\varepsilon} . We multiply the first equation in (4.8) by \widehat{v_\varepsilon} to get

    \begin{align*} \frac12\frac{d}{dt}\int_\Omega\widehat{v_\varepsilon}^2 dx+\int_\Omega|\nabla\widehat{v_\varepsilon}|^2 dx+\int_\Omega\widehat{v_\varepsilon}^2 dx\leq\int_\Omega u_\varepsilon v_\varepsilon-\int_\Omega u_\varepsilon v_\infty dx+\int_\Omega \widehat{v_\varepsilon}(h_2-h_{2,\infty})dx, \end{align*}

    and thereby obtain from Young's inequality that

    \begin{align*} \frac{d}{dt}\int_\Omega\widehat{v_\varepsilon}^2 dx+2\int_\Omega|\nabla\widehat{v_\varepsilon}|^2 dx+\int_\Omega\widehat{v_\varepsilon}^2 dx\leq2\int_\Omega u_\varepsilon v_\varepsilon+2\|v_\infty\|_{L^\infty}\int_\Omega u_\varepsilon dx+\int_\Omega (h_2-h_{2,\infty})^2dx. \end{align*}

    By taking \lambda\geq\max\left\{\frac4\kappa, \frac{4\|v_\infty\|_{L^\infty}+2}{c_1\kappa}\right\} , this, combined with (4.3), ensures

    \begin{align*} \frac{d}{dt}\int_\Omega\widehat{v_\varepsilon}^2+\lambda u_\varepsilon dx+2\int_\Omega|\nabla\widehat{v_\varepsilon}|^2 dx+\int_\Omega\widehat{v_\varepsilon}^2 +u_\varepsilon dx\leq\int_\Omega (h_2-h_{2,\infty})^2dx+\lambda\int_\Omega h_1dx. \end{align*}

    Setting g(t): = \int_\Omega (h_2-h_{2, \infty})^2dx+\lambda\int_\Omega h_1dx and y_\varepsilon(t): = \int_\Omega\widehat{v_\varepsilon}^2+\mu u_\varepsilon dx , we have

    \begin{align} y_\varepsilon'(t)+\min\{\lambda^{-1},1\}y_\varepsilon(t)+2\int_\Omega|\nabla\widehat{v_\varepsilon}|^2 dx\leq g(t). \end{align} (4.9)

    By means of (1.13)–(1.14) and Lemma 4.4, the desired (4.6) holds. We now integrate (4.9) over [t, t + 1] to get

    \begin{align*} 2\int_t^{t+1}\int_\Omega|\nabla\widehat{v_\varepsilon}|^2 dxds\leq \int_t^{t+1}g(s)ds+y_\varepsilon(t). \end{align*}

    This, in view of (1.13), (1.14) and (4.6) again, ensures that (4.7) holds.

    Our second result on the large-time behavior of generalized solutions featured in Theorem 1.3 is in fact a by-product of our previous analysis.

    Proof of Theorem 1.3. In fact, Lemma 3.5, combining with the Fubini-Tonelli theorem, provides (\varepsilon_j)_{j\in\mathbb{N}}\subset(0, 1) and a null set \mathcal{N}\subset(0, \infty) such that \varepsilon_j\rightarrow0 as j\rightarrow \infty and

    u_\varepsilon(\cdot,t)\rightarrow u(\cdot,t) \,\,\,{\mathrm{and}}\,\,\, v_\varepsilon(\cdot,t)\rightarrow (\cdot,t)\,\,\, a.e.\,\, {\mathrm{in}}\,\,\Omega\,\,{\mathrm{for\,\,all}}\,\,t\in(0,\infty)\setminus\mathcal{N},

    as \varepsilon = \varepsilon_j\rightarrow0 . Based on this, Lemma 4.5, together with Fatou's lemma, presents the desired large-time behavior of the generalized solution in Theorem 1.3.

    The authors are sincerely grateful to the anonymous reviewers for the detailed comments and valuable suggestions which really helped us to make the paper more readable and meaningful. The research of ZW is supported by the National Natural Science Foundation of China (No. 11701304) and Natural Science Foundation of Ningbo Municipality (No. 2019A610041, No. 2021J143). Wang also gratefully acknowledges the support of KC Wong Education Foundation. The research of LX is partially supported by the Chongqing Science and Technology Commission Project (No. sctc2020jcyj-msxmX0560, No. csts2020jcyj-jqX0022), and the Science Technology Research Program of Chongqing Municipal Education Commission (No. KJZD-M202000502, No. CXQT21014).

    The authors declare there is no conflict of interest.



    [1] N. Rodríguez, On the global well-posedness theory for a class of PDE models for criminal activity, Phys. D Nonlinear Phenom., 260 (2013), 191–200. https://doi.org/10.1016/j.physd.2012.08.003 doi: 10.1016/j.physd.2012.08.003
    [2] M. Short, M. D'Orsogna, V. Pasour, G. Tita, P. Brantingham, A. Bertozzi, et al., A statistical model of criminal behavior, Math. Mod. Meth. Appl. Sci., 18 (2008), 1249–1267. https://doi.org/10.1142/S0218202508003029 doi: 10.1142/S0218202508003029
    [3] M. Short, A. Bertozzi, P. Brantingham, G. Tita, Dissipation and displacement of hotspots in reaction-diffusion model of crime, Proc. Natl. Acad. Sci. USA, 107 (2010), 3961–3965. https://doi.org/0.1073/pnas.0910921107 doi: 10.1073/pnas.0910921107
    [4] H. Berestycki, J. Wei, M. Winter, Existence of symmetric and asymmetric spikes for a crime hotspot model, SIAM J. Math. Anal., 46 (2014), 691–719. https://doi.org/10.1137/130922744 doi: 10.1137/130922744
    [5] R. Cantrell, C. Cosner, R. Manásevich, Global bifurcation of solutions for crime modeling equations, SIAM J. Math. Anal., 44 (2012), 1340–1358. https://doi.org/10.1137/110843356 doi: 10.1137/110843356
    [6] Y. Gu, Q. Wang, G. Yi, Stationary patterns and their selection mechanism of urban crime models with heterogeneous near-repeat victimization effect, Eur. J. Appl. Math., 28 (2017), 141–178. https://doi.org/10.1017/S0956792516000206 doi: 10.1017/S0956792516000206
    [7] T. Kolokolnikov, M. Ward, J. Wei, The stability of steady-state hot-spot patterns for a reaction-diffusion model of urban crime, Discrete Contin. Dyn. Syst. Ser. B, 19 (2014), 1373–1410. https://doi.org/10.3934/dcdsb.2014.19.1373 doi: 10.3934/dcdsb.2014.19.1373
    [8] D. Lloyd, H. O'Farrell, On localised hotspots of an urban crime model, Phys. D Nonlinear Phenom., 253 (2013), 23–39. https://doi.org/10.1016/j.physd.2013.02.005 doi: 10.1016/j.physd.2013.02.005
    [9] D. Lloyd, N. Santitissadeekorn, M. Short, Exploring data assimilation and forecasting issues for an urban crime model, Eur. J. Appl. Math., 27 (2016), 451–478. https://doi.org/10.1017/S0956792515000625 doi: 10.1017/S0956792515000625
    [10] L. Mei, J. Wei, The existence and stability of spike solutions for a chemotax is system modeling crime pattern formation, Math. Models Methods Appl. Sci., 30 (2020), 1727–1764. https://doi.org/10.1142/S0218202520500359 doi: 10.1142/S0218202520500359
    [11] M. Short, A. Bertozzi, P. Brantingham, Nonlinear patterns in urban crime: hotspots, bifurcations, and suppression, SIAM J. Appl. Dyn. Syst., 9 (2010), 462–483. https://doi.org/10.1137/090759069 doi: 10.1137/090759069
    [12] W. Tse, M. Ward, Hotspot formation and dynamics for a continuum model of urban crime, Eur. J. Appl. Math., 27 (2016), 583–624. https://doi.org/10.1017/S0956792515000376 doi: 10.1017/S0956792515000376
    [13] N. Rodríguez, A. Bertozzi, Local existence and uniqueness of solutions to a PDE model for criminal behavior, Math. Models Methods Appl. Sci., 20 (2010), 1425–1457. https://doi.org/10.1142/S0218202510004696 doi: 10.1142/S0218202510004696
    [14] N. Rodríguez, M. Winkler, On the global existence and qualitative behavior of one-dimensional solutions to a model for urban crime, Eur. J. Appl. Math., 33 (2022), 919–959. https://doi.org/10.1017/S0956792521000279 doi: 10.1017/S0956792521000279
    [15] Q. Wang, D. Wang, Y. Feng, Global well-posedness and uniform boundedness of urban crime models: One-dimensional case, J. Differ. Equations, 269 (2020), 6216–6235. https://doi.org/10.1016/j.jde.2020.04.035 doi: 10.1016/j.jde.2020.04.035
    [16] M. Freitag, Global solutions to a higher-dimensional system related to crime modeling, Math. Meth. Appl. Sci., 41 (2018), 6326–6335. https://doi.org/10.1002/mma.5141 doi: 10.1002/mma.5141
    [17] J. Shen, B. Li, Mathematical analysis of a continuous version of statistical models for criminal behavior, Math. Meth. Appl. Sci., 43 (2020), 409–426. https://doi.org/10.1002/mma.5898 doi: 10.1002/mma.5898
    [18] J. Ahn, K. Kang, J. Lee, Global well-posedness of logarithmic Keller-Segel type systems, J. Differ. Equations, 287 (2021), 185–211. https://doi.org/10.1016/j.jde.2021.03.053 doi: 10.1016/j.jde.2021.03.053
    [19] Y. Tao, M. Winkler, Global smooth solutions in a two-dimensional cross-diffusion system modeling propagation of urban crime, Commun. Math. Sci., 19 (2021), 829–849. https://doi.org/10.4310/CMS.2021.v19.n3.a12 doi: 10.4310/CMS.2021.v19.n3.a12
    [20] M. Winkler, Global solvability and stabilization in a two-dimensional cross-diffusion system modeling urban crime propagation, Ann. Inst. Henri Poincaré, Anal. Non Linéaire, 36 (2019), 1747–1790. https://doi.org/10.1016/j.anihpc.2019.02.004 doi: 10.1016/j.anihpc.2019.02.004
    [21] Y. Jiang, L. Yang, Global solvability and stabilization in a three-dimensional cross-diffusion system modeling urban crime propagation, Acta Appl. Math., 178 (2022). https://doi.org/10.1007/s10440-022-00484-z doi: 10.1007/s10440-022-00484-z
    [22] N. Rodríguez, M. Winkler, Relaxation by nonlinear diffusion enhancement in a two-dimensional cross-diffusion model for urban crime propagation, Math. Models Methods Appl. Sci., 30 (2020), 2105–2137. https://doi.org/10.1142/S0218202520500396 doi: 10.1142/S0218202520500396
    [23] F. Heihoff, Generalized solutions for a system of partial differential equations arising from urban crime modeling with a logistic source term, Z. Für Angew. Math. Phys., 71 (2020). https://doi.org/10.1007/s00033-020-01304-w doi: 10.1007/s00033-020-01304-w
    [24] B. Li, L. Xie, Generalized solution to a 2D parabolic-parabolic chemotaxis system for urban crime: Global existence and large time behavior, submitted for publication, 2022.
    [25] P. Jones, P. Brantingham, L. Chayes, Statistical models of criminal behavior: The effects of law enforcement actions, Math. Models Methods Appl. Sci., 20 (2010), 1397–1423. https://doi.org/10.1142/S0218202510004647 doi: 10.1142/S0218202510004647
    [26] A. Pitcher, Adding police to a mathematical model of burglary, Eur. J. Appl. Math., 21 (2010), 401–419. https://doi.org/10.1017/S0956792510000112 doi: 10.1017/S0956792510000112
    [27] J. Zipkin, M. Short, A. Bertozzi, Cops on the dots in a mathematical model of urban crime and police response, Discrete Contin. Dyn. Syst. Ser. B, 19 (2014), no. 5, 1479–1506. https://doi.org/10.3934/dcdsb.2014.19.1479 doi: 10.3934/dcdsb.2014.19.1479
    [28] W. Tse, M. Ward, Asynchronous instabilities of crime hotspots for a 1-D reaction-diffusion model of urban crime with focused police patrol, SIAM J. Appl. Dyn. Syst., 17 (2018), 2018–2075. https://doi.org/10.1137/17M1162585 doi: 10.1137/17M1162585
    [29] A. Buttenschoen, T. Kolokolnikov, M. Ward, J. Wei, Cops-on-the-dots: the linear stability of crime hotspots for a 1-D reaction-diffusion model of urban crime, Eur. J. Appl. Math., 31 (2020), 871–917. https://doi.org/10.1017/S0956792519000305 doi: 10.1017/S0956792519000305
    [30] B. Li, L. Xie, Global large-data generalized solutions to a two-dimensional chemotaxis system stemming from crime modelling, Discrete Contin. Dyn. Syst. Ser. B, 2022 (2022). https://doi.org/10.3934/dcdsb.2022167 doi: 10.3934/dcdsb.2022167
    [31] N. Rodríguez, Q. Wang, L. Zhang, Understanding the effects of on- and off-hotspot policing: Evidence of hotspot, oscillating, and chaotic activities, SIAM J. Appl. Dyn. Syst., 20 (2021), 1882–1916. https://doi.org/10.1137/20M1359572 doi: 10.1137/20M1359572
    [32] E. Keller, L. Segel, Initiation of slime mold aggregation viewed as an instability, J. Theor. Biol., 26 (1970), 399–415. https://doi.org/10.1016/0022-5193(70)90092-5 doi: 10.1016/0022-5193(70)90092-5
    [33] K. Fujie, Boundedness in a fully parabolic chemotaxis system with strongly singular sensitivity, J. Math. Anal. Appl., 424 (2015), 675–684. https://doi.org/10.1016/j.jmaa.2014.11.045 doi: 10.1016/j.jmaa.2014.11.045
    [34] M. Winkler, Global solutions in a fully parabolic chemotaxis system with singular sensitivity, Math. Methods Appl. Sci., 34 (2011), 176–190. https://doi.org/10.1002/mma.1346 doi: 10.1002/mma.1346
    [35] J. Lankeit, A new approach toward boundedness in a two-dimensional parabolic chemotaxis system with singular sensitivity, Math. Meth. Appl. Sci., 39 (2016), 394–404. https://doi.org/10.1002/mma.3489 doi: 10.1002/mma.3489
    [36] C. Stinner, M. Winkler, Global weak solutions in a chemotaxis system with large singular sensitivity, Nonliear Anal. Real Word Appl., 12 (2011), 3727–3740. https://doi.org/10.1016/j.nonrwa.2011.07.006 doi: 10.1016/j.nonrwa.2011.07.006
    [37] J. Lankeit, M. Winkler, A generalized solution concept for the Keller-Segel system with logarithmic sensitivity: global solvability for large nonradial data, NoDEA-Nonlinear Differ. Equ. Appl., 24 (2017). https://doi.org/10.1007/s00030-017-0472-8 doi: 10.1007/s00030-017-0472-8
    [38] A. Zhigun, Generalised supersolutions with mass control for the Keller-Segel system with logarithmic sensitivity, J. Math. Anal. Appl., 467 (2018), 1270–1286. https://doi.org/10.1016/j.jmaa.2018.08.001 doi: 10.1016/j.jmaa.2018.08.001
    [39] M. Winkler, Unlimited growth in logarithmic Keller-Segel systems, J. Differ. Equations, 309 (2022), 74–97. https://doi.org/10.1016/j.jde.2021.11.026 doi: 10.1016/j.jde.2021.11.026
    [40] M. Winkler, T. Yokota, Stabilization in the logarithmic Keller-Segel system, Nonlinear Anal., 170 (2018), 123–141. https://doi.org/10.1016/j.na.2018.01.002 doi: 10.1016/j.na.2018.01.002
    [41] J. Ahn, Global well-posedness and asymptotic stabilization for chemotaxis system with signal-dependent sensitivity, J. Differ. Equations, 266 (2019), 6866–6904. https://doi.org/10.1016/j.jde.2018.11.015 doi: 10.1016/j.jde.2018.11.015
    [42] Q. Hou, Z. Wang, K. Zhao, Boundary layer problem on a hyperbolic system arising from chemotaxis, J. Differ. Equations, 261 (2016), 5035–5070. https://doi.org/10.1016/j.jde.2016.07.018 doi: 10.1016/j.jde.2016.07.018
    [43] H. Jin, J. Li, Z. Wang, Asymptotic stability of traveling waves of a chemotaxis model with singular sensitivity, J. Differ. Equations, 255 (2013), 193–219. doilinkhttps://doi.org/10.1016/j.jde.2013.04.002 doi: 10.1016/j.jde.2013.04.002
    [44] H. Li, K. Zhao, Initial-boundary value problems for a system of hyperbolic balance laws arising from chemotaxis, J. Differ. Equations, 258 (2015), 302–308. https://doi.org/10.1016/j.jde.2014.09.014 doi: 10.1016/j.jde.2014.09.014
    [45] J. Li, T. Li, Z. Wang, Stability of traveling waves of the Keller-Segel system with logarithmic sensitivity, Math. Models Methods Appl. Sci., 24 (2014), 2819–2849. https://doi.org/10.1142/S0218202514500389 doi: 10.1142/S0218202514500389
    [46] Y. Tao, L. Wang, Z. Wang, Large-time behavior of a parabolic-parabolic chemotaxis model with logarithmic sensitivity in one dimension, Discrete Contin. Dyn. Syst. Ser. B, 18 (2013), 821–845. https://doi.org/10.3934/dcdsb.2013.18.821 doi: 10.3934/dcdsb.2013.18.821
    [47] Z. Wang, Z. Xiang, P. Yu, Asymptotic dynamics in a singular chemotaxis system modeling onset of tumor angiogenesis, J. Differ. Equations, 260 (2016), 2225–2258. https://doi.org/10.1016/j.jde.2015.09.063 doi: 10.1016/j.jde.2015.09.063
    [48] M. Winkler, Renormalized radial large-data solutions to the higher-dimensional Keller-Segel system with singular sensitivity and signal absorption, J. Differ. Equations, 264 (2018), 2310–2350. https://doi.org/10.1016/j.jde.2017.10.029 doi: 10.1016/j.jde.2017.10.029
    [49] M. Winkler, The two-dimensional Keller-Segel system with singular sensitivity and signal absorption: global large-data solutions and their relaxation properties, Math. Models Methods Appl. Sci., 26 (2016), 987–1024. https://doi.org/10.1142/S0218202516500238 doi: 10.1142/S0218202516500238
    [50] B. Li, L. Xie, Generalized solution and its eventual smoothness to a logarithmic Keller-Segel system for criminal activities, submitted for publication, 2022.
    [51] M. Winkler, Large-data global generalized solutions in a chemotaxis system with tensor-valued sensitivities, SIAM J. Math. Anal., 47 (2015), 3092–3115. https://doi.org/dx.doi.org/10.1137/140979708 doi: 10.1137/140979708
    [52] M. Aida, K. Osaka, T. Tsujikawa, M. Mimura, Chemotaxis and growth system with sigular sensitivity function, Nonliear Anal. Real Word Appl., 6 (2005), 323–336. https://doi.org/10.1016/j.nonrwa.2004.08.011 doi: 10.1016/j.nonrwa.2004.08.011
    [53] X. Cao, Global bounded solutions of the higher-dimensional Keller-Segel system under smallness conditions in optimal spaces, Discrete Contin. Dyn. Syst., 35 (2015), 1891–1904. https://doi.org/10.3934/dcds.2015.35.1891 doi: 10.3934/dcds.2015.35.1891
    [54] M. Winkler, Aggregation vs. global diffusive behavior in the higher-dimensional Keller-Segel model, J. Differ. Equations, 248 (2010), 2889–2905. https://doi.org/10.1016/j.jde.2010.02.008 doi: 10.1016/j.jde.2010.02.008
    [55] J. Simon, Compact sets in the space L^p(0, T;B), Ann. Mat. Pura Appl., 146 (1996), 65–96. https://doi.org/10.1007/BF01762360 doi: 10.1007/BF01762360
    [56] O. Ladyzhenskaya, N. Ural'tseva, Linear and Quasilinear Elliptic Equations, Translated from the Russian by Scripta Technica, Inc. Translation editor: Leon Ehrenpreis, Academic Press, New York-London, 1968.
  • This article has been cited by:

    1. Zixuan Qiu, Bin Li, Eventual smoothness of generalized solutions to a singular chemotaxis system for urban crime in space dimension 2, 2023, 31, 2688-1594, 3218, 10.3934/era.2023163
    2. Bin Li, Li Xie, Smoothness effects of a quadratic damping term of mixed type on a chemotaxis-type system modeling propagation of urban crime, 2023, 73, 14681218, 103912, 10.1016/j.nonrwa.2023.103912
    3. Yan Zhao, Li Xie, Global classical solvability and asymptotic behaviors of a parabolic-elliptic Chemotaxis-type system modeling crime activities, 2024, 532, 0022247X, 127909, 10.1016/j.jmaa.2023.127909
    4. Bin Li, Li Xie, Generalized solution and eventual smoothness in a logarithmic Keller–Segel system for criminal activities, 2023, 33, 0218-2025, 1281, 10.1142/S0218202523500306
    5. Bin Li, Li Xie, Can Dirac-type singularities in Keller-Segel systems be ruled out by power-type singular sensitivities?, 2024, 379, 00220396, 413, 10.1016/j.jde.2023.10.013
    6. Wenjing Jiang, Qi Wang, Global well-posedness and uniform boundedness of 2D urban crime models with nonlinear advection enhancement, 2024, 0956-7925, 1, 10.1017/S0956792524000792
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1728) PDF downloads(97) Cited by(6)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog