Skip to main content
Log in

On a Two-Species Attraction–Repulsion Chemotaxis System with Nonlocal Terms

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

Abstract

This paper deals with a two-species attraction–repulsion chemotaxis system

$$\begin{aligned} \left\{ \begin{aligned}{}&u_t=d_{1}\Delta u-\xi _{1}\nabla \cdot (u\nabla v)+\chi _{1}\nabla \cdot (u\nabla z)+g_{1}(u,w),&(x,t)\in \Omega \times (0,\infty ), \\&\tau v_{t}=d_{2}\Delta v+w-v,&(x,t)\in \Omega \times (0,\infty ),\\&w_t=d_{3}\Delta w-\xi _{2}\nabla \cdot (w\nabla z)+\chi _{2}\nabla \cdot (w\nabla v)+g_{2}(u,w),&(x,t)\in \Omega \times (0,\infty ), \\&\tau z_{t}=d_{4}\Delta z+u-z,&(x,t)\in \Omega \times (0,\infty ) \end{aligned} \right. \end{aligned}$$

under homogeneous Neumann boundary conditions in a smoothly bounded domain \(\Omega \subset {\mathbb {R}}^{n}\) for \(n\ge 1\), where \(\tau \in \{0,1\}\), the parameters \(d_{i}(i=1,2,3,4),\xi _{j},\chi _{j}(j=1,2)\) are positive and the kinetic terms \(g_{1}(u,w),g_{2}(u,w)\) satisfy

$$\begin{aligned} \left\{ \begin{aligned}{}&g_{1}(u,w)=u\bigg (a_{0}-a_{1}u-a_{2}w-a_{3}\int _{\Omega }u{\text {d}}x-a_{4}\int _{\Omega }w{\text {d}}x\bigg ),\\&g_{2}(u,w)=w\bigg (b_{0}-b_{1}u-b_{2}w-b_{3}\int _{\Omega }u{\text {d}}x-b_{4}\int _{\Omega }w{\text {d}}x\bigg )\\ \end{aligned} \right. \end{aligned}$$

with \(a_{0},a_{1},b_{0},b_{2}>0,a_{2},a_{3},a_{4},b_{1},b_{3},b_{4}\in {\mathbb {R}}\). It is shown that under some suitable parameter conditions, the above system possesses a unique global and uniformly bounded solution in any spatial dimension. Moreover, we investigate the asymptotic stability of solutions under the locally intraspecific competition and globally interspecific cooperation. Finally, we present some numerical simulations, which not only support our analytically theoretical results, but also find some new and interesting phenomena.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12

Similar content being viewed by others

Data availability

Data sharing is not applicable to this article as no new data were created or analyzed in this study.

References

  • Agmon, S., Douglis, A., Nirenberg, L.: Estimates near the boundary for solutions of elliptic partial differential equations satisfying general boundary conditions. I. Commun. Pure Appl. Math. 12, 623–727 (1959)

    MathSciNet  MATH  Google Scholar 

  • Agmon, S., Douglis, A., Nirenberg, L.: Estimates near the boundary for solutions of elliptic partial diffferential equations satisfying general boundary conditions. II. Commun. Pure Appl. Math. 17, 35–92 (1964)

    MATH  Google Scholar 

  • Alikakos, N.: \( L^{p} \)-bounds of solutions of reaction-diffusion equations. Commun. Partial Differ. Equ. 4, 827–868 (1979)

    MATH  Google Scholar 

  • Armstrong, N., Painter, K., Sherratt, J.: A continuum approach to modelling cell-cell adhesion. J. Theoret. Biol. 243, 98–113 (2006)

    MathSciNet  MATH  Google Scholar 

  • Bai, X., Winkler, M.: Equilibration in a fully parabolic two-species chemotaxis system with competitive kinetics. Indiana Univ. Math. J. 65, 553–583 (2016)

    MathSciNet  MATH  Google Scholar 

  • Bellomo, N., Bellouquid, A., Tao, Y., Winkler, M.: Toward a mathematical theory of Keller–Segel models of pattern formation in biological tissues. Math. Models Methods Appl. Sci. 25, 1663–1763 (2015)

    MathSciNet  MATH  Google Scholar 

  • Black, T., Lankeit, J., Mizukami, M.: On the weakly competitive case in a two-species chemotaxis model. IMA J. Appl. Math. 81, 860–876 (2016)

    MathSciNet  MATH  Google Scholar 

  • Black, T.: Global existence and asymptotic stability in a competitive two-species chemotaxis system with two signals. Discrete Contin. Dyn. Syst. Ser. B 22, 1253–1272 (2017)

    MathSciNet  MATH  Google Scholar 

  • Budrene, E., Berg, H.: Complex patterns formed by motile cells of Escherichia coli. Nature 349, 630–633 (1991)

    Google Scholar 

  • Burger, M., Francesco, M., Dolak, Y.: The Keller–Segel model for chemotaxis with prevention of overcrowding: linear vs. nonlinear diffusion. SIAM J. Math. Anal. 38, 1288–1315 (2006)

    MathSciNet  MATH  Google Scholar 

  • Chaplain, M., Logas, G.: Mathematical modelling of cancer cell invasion of tissue: the role of the urokinase plasminogen activation system. Math. Models Methods Appl. Sci. 15, 1685–1734 (2005)

    MathSciNet  MATH  Google Scholar 

  • Chiyo, Y., Yokota, T.: Boundedness in a fully parabolic attraction-repulsion chemotaxis system with nonlinear diffusion and signal-dependent sensitivit. Nonlinear Anal. Real World Appl. 66, 103533 (2022)

    MATH  Google Scholar 

  • Chiyo, Y., Yokota, T.: Boundedness and finite-time blow-up in a quasilinear parabolic-elliptic-elliptic attraction-repulsion chemotaxis system. Z. Angew. Math. Phys. 73, 1–27 (2022)

    MathSciNet  MATH  Google Scholar 

  • Coville, J., Dávila, J., Martínez, S.: Nonlocal anisotropic dispersal with monostable nonlinearity. J. Differ. Equ. 244, 3080–3118 (2008)

    MathSciNet  MATH  Google Scholar 

  • Espejo, E., Suzuki, T.: Global existence and blow-up for a system describing the aggregation of microglia. Appl. Math. Lett. 35, 29–34 (2014)

    MathSciNet  MATH  Google Scholar 

  • Evje, S., Winkler, M.: Mathematical analysis of two competing cancer cell migration mechanisms driven by interstitial fluid flow. J. Nonlinear Sci. 30, 1809–1847 (2020)

    MathSciNet  MATH  Google Scholar 

  • Freitag, M.: Global existence and boundedness in a chemorepulsion system with superlinear diffusion. Discrete Contin. Dyn. Syst. Ser. A 38, 5943–5961 (2018)

    MathSciNet  MATH  Google Scholar 

  • Friedman, A.: Partial Differential Equations. Holt, Rinehart and Winston, New York (1969)

    MATH  Google Scholar 

  • Gajewski, H., Zacharias, K.: Global behavior of a reaction-diffusion system modeling chemotaxis. Math. Nachr. 195, 77–114 (1998)

    MathSciNet  MATH  Google Scholar 

  • Gerisch, A., Chaplain, M.: Mathematical modelling of cancer cell invasion of tissue: local and nonlocal models and the effect of adhesion. J. Theoret. Biol. 250, 684–704 (2008)

    MathSciNet  MATH  Google Scholar 

  • Hazelbauer, G.: Taxis and Behavior: Elementary Sensory Systems in Biology, vol. 3, pp. 185–186. Chapman and Hall, London (1979)

    Google Scholar 

  • Heihoff, F.: On the existence of global smooth solutions to the parabolic-elliptic Keller–Segel system with irregular initial data. J. Dyn. Differ. Equ. 9, 1–25 (2021)

    MathSciNet  Google Scholar 

  • Henry, D.: Geometric Theory of Semilinear Parabolic Equations. Lecture Notes in Mathematics, vol. 840. Springer, New York (1981)

    Google Scholar 

  • Hillen, T., Painter, K.: A user’s guide to PDE models for chemotaxis. J. Math. Biol. 58, 183–217 (2009)

    MathSciNet  MATH  Google Scholar 

  • Horstmann, D., Winkler, M.: Boundedness vs. blow-up in a chemotaxis system. J. Differ. Equ. 215, 52–107 (2005)

    MathSciNet  MATH  Google Scholar 

  • Hsieh, C., Yu, Y.: Boundedness of solutions to an attraction-repulsion chemotaxis model in \({\mathbb{R} }^{2}\). J. Differ. Equ. 317, 422–438 (2022)

    MATH  Google Scholar 

  • Hu, R., Zheng, P.: On a quasilinear fully parabolic attraction or repulsion chemotaxis system with nonlinear signal production. Discrete Contin. Dyn. Syst. Ser. B 12, 7227–7244 (2022)

    MathSciNet  MATH  Google Scholar 

  • Hu, R., Zheng, P., Gao, Z.: Boundedness of solutions in a quasilinear chemo-repulsion system with nonlinear signal production. Evol. Equ. Control Theory 11, 2209–2219 (2022)

    MathSciNet  MATH  Google Scholar 

  • Hu, R., Zheng, P.: Global stability in a two-species attraction-repulsion system with competitive and nonlocal kinetics. J. Dyn. Differ. Equ. (2022). https://doi.org/10.1007/s10884-022-10215-5

    Article  Google Scholar 

  • Ishida, S., Seki, K., Yokota, T.: Boundedness in quasilinear Keller–Segel systems of parabolic-parabolic type on non-convex bounded domains. J. Differ. Equ. 256, 2993–3010 (2014)

    MathSciNet  MATH  Google Scholar 

  • Issa, T., Salako, R.: Asymptotic dynamics in a two-species chemotaxis model with nonlocal terms. Discrete Contin. Dyn. Syst. Ser. B 22, 3839–3874 (2017)

    MathSciNet  MATH  Google Scholar 

  • Jin, H.: Boundedness of the attraction-repulsion Keller–Segel system. J. Math. Anal. Appl. 422, 1463–1478 (2015)

    MathSciNet  MATH  Google Scholar 

  • Kao, C., Lou, Y., Shen, W.: Random dispersal vs nonlocal dispersal. Discrete Contin. Dyn. Syst. Ser. A 26, 551–596 (2010)

    MATH  Google Scholar 

  • Keller, E., Segel, L.: Initiation of slime mold aggregation viewed as an instability. J. Theoret. Biol. 26, 399–415 (1970)

    MathSciNet  MATH  Google Scholar 

  • Kurt, H., Shen, W.: Finite-time blow-up prevention by logistic source in parabolic-elliptic chemotaxis models with singular sensitivity in any dimensional setting. SIAM J. Math. Anal. 53, 973–1003 (2021)

    MathSciNet  MATH  Google Scholar 

  • Li, S., Muneoka, K.: Cell migration and chick limb development: chemotactic action of FGF-4 and the AER. Dev. Cell 211, 335–347 (1999)

    Google Scholar 

  • Li, Y., Li, Y.: Blow-up of nonradial solutions to attraction-repulsion chemotaxis system in two dimensions. Nonlinear Anal. Real World Appl. 30, 170–183 (2016)

    MathSciNet  MATH  Google Scholar 

  • Li, X., Wang, Y.: Boundedness in a two-species chemotaxis parabolic system with two chemicals. Discrete Contin. Dyn. Syst. Ser. B 22, 2717–2729 (2017)

    MathSciNet  MATH  Google Scholar 

  • Lin, K., Mu, C., Wang, L.: Large time behavior for an attraction-repulsion chemotaxis system. J. Math. Anal. Appl. 426, 105–124 (2015)

    MathSciNet  MATH  Google Scholar 

  • Lin, K., Xiang, T.: Strong damping effect of chemo-repulsion prevents blow-up. J. Math. Phys. 62, 041508 (2021)

    MathSciNet  MATH  Google Scholar 

  • Liu, A., Dai, B.: Blow-up vs boundedness in a two-species attraction-repulsion chemotaxis system with two chemicals. J. Math. Phys. 62, 111508 (2021)

    MathSciNet  MATH  Google Scholar 

  • Liu, A., Dai, B.: Boundedness and stabilization in a two-species chemotaxis system with two chemicals. J. Math. Anal. Appl. 506, 125609 (2022)

    MathSciNet  MATH  Google Scholar 

  • Liu, A., Dai, B., Chen, Y.: Boundedness in a two-species attraction-repulsion chemotaxis system with two chemicals. Discrete Contin. Dyn. Syst. Ser. B 27, 6037–6062 (2022)

    MathSciNet  MATH  Google Scholar 

  • Liu, D., Tao, Y.: Boundedness in a chemotaxis system with nonlinear signal production. Appl. Math. J. Chinese Univ. Ser. B 31, 379–388 (2016)

    MathSciNet  MATH  Google Scholar 

  • Liu, J., Wang, Z.: Classical solutions and steady states of an attraction-repulsion chemotaxis in one dimension. J. Biol. Dynam. 6, 31–41 (2012)

    MathSciNet  MATH  Google Scholar 

  • Luca, M., Chavez-Ross, A., Edelstein-Keshet, L., Mogilner, A.: Chemotactic signalling, microglia, and Alzheimer’s disease senile plagues: Is there a connection? Bull. Math. Biol. 65, 693–730 (2003)

    MATH  Google Scholar 

  • Mimura, M., Tsujikawa, T.: Aggregating pattern dynamics in a chemotaxis model including growth. Phys. A 230, 449–543 (1996)

    Google Scholar 

  • Mizukami, M., Yokota, T.: Global existence and asymptotic stability of solutions to a two-species chemotaxis system with any chemical diffusion. J. Differ. Equ. 261, 2650–2669 (2016)

    MathSciNet  MATH  Google Scholar 

  • Nagai, T.: Blow-up of radially symmetric solutions of a chemotaxis system. Adv. Math. Sci. Appl. 5, 581–601 (1995)

    MathSciNet  MATH  Google Scholar 

  • Negreanu, M., Tello, J.: On a competitive system under chemotactic effects with nonlocal terms. Nonlinearity 26, 1083–1103 (2013)

    MathSciNet  MATH  Google Scholar 

  • Negreanu, M., Tello, J.: On a two species chemotaxis model with slow chemical diffusion. SIAM J. Math. Anal. 46, 3761–3781 (2014)

    MathSciNet  MATH  Google Scholar 

  • Painter, K.: Continuous models for cell migration in tissues and applications to cell sorting via differential chemotaxis. Bull. Math. Biol. 71, 1117–1147 (2009)

    MathSciNet  MATH  Google Scholar 

  • Painter, K., Hillen, T.: Volume-filling and quorum-sensing in models for chemosensitive movement. Canad. Appl. Math. Quart. 10, 501–543 (2002)

    MathSciNet  MATH  Google Scholar 

  • Petter, G., Byrne, H., Mcelwain, D., Norbury, J.: A model of wound healing and angiogenesis in soft tissue. Math. Biosci. 136, 35–63 (2003)

    MATH  Google Scholar 

  • Shen, W., Zhang, A.: Stabilization solutions and spreading speeds of nonlocal monostable equations in space periodic habitats. Proc. Amer. Math. Soc. 140, 1681–1696 (2012)

    MathSciNet  MATH  Google Scholar 

  • Sherratt, J., Gourley, S., Armstrong, N., Painter, K.: Boundedness of solutions of a nonlocal reaction-diffusion model for adhesion in cell aggregation and cancer invasion. European J. Appl. Math. 20, 123–144 (2009)

    MathSciNet  MATH  Google Scholar 

  • Stinner, C., Surulescu, C., Winkler, M.: Global weak solutions in a PDE-ODE system modeling multiscale cancer cell invasion. SIAM J. Math. Anal. 46, 1969–2007 (2014)

    MathSciNet  MATH  Google Scholar 

  • Stinner, C., Tello, J., Winkler, M.: Competitive exclusion in a two-species chemotaxis model. J. Math. Biol. 68, 1607–1626 (2014)

    MathSciNet  MATH  Google Scholar 

  • Tao, Y.: Global dynamics in a higher-dimensional repulsion chemotaxis model with nonlinear sensitivity. Discrete Contin. Dyn. Syst. Ser. B 18, 2705–2722 (2013)

    MathSciNet  MATH  Google Scholar 

  • Tao, Y., Wang, Z.: Competing effects of attraction vs. repulsion in chemotaxis. Math. Models Methods Appl. Sci. 23, 1–36 (2013)

    MathSciNet  MATH  Google Scholar 

  • Tao, Y., Winkler, M.: Boundedness in a quasilinear parabolic-parabolic Keller–Segel system with subcritical sensitivity. J. Differ. Equ. 252, 692–715 (2012)

    MathSciNet  MATH  Google Scholar 

  • Tao, Y., Winkler, M.: Boundedness vs. blow-up in a two-species chemotaxis system with two chemicals. Discrete Contin. Dyn. Syst. Ser. B 20, 3165–3183 (2015)

    MathSciNet  MATH  Google Scholar 

  • Tello, J., Winkler, M.: A chemotaxis system with logistic source. Commun. Partial Differ. Equ. 32, 849–877 (2007)

    MathSciNet  MATH  Google Scholar 

  • Tello, J., Winkler, M.: Stabilization in a two-species chemotaxis system with a logistic source. Nonlinearity 25, 1413–1425 (2012)

    MathSciNet  MATH  Google Scholar 

  • Temam, R.: Infinite-Dimensional Dynamical Systemsin Mechanics and Physics. Appl. Math. Sci., vol. 68, 2nd edn. Springer, New York (1997)

    Google Scholar 

  • Tu, X., Mu, C., Zheng, P., Lin, K.: Global dynamics in a two-species chemotaxis-competition system with two signals. Discrete Contin. Dyn. Syst. Ser. A 38, 3617–3636 (2018)

    MathSciNet  MATH  Google Scholar 

  • Wang, L., Mu, C.: A new result for boundedness and stabilization in a two-species chemotaxis system with two chemicals. Discrete Contin. Dyn. Syst. Ser. B 25, 4585–4601 (2020)

    MathSciNet  MATH  Google Scholar 

  • Wang, W., Zhuang, M., Zheng, S.: Positive effects of repulsion on boundedness in a fully parabolic attraction-repulsion chemotaxis system with logistic source. J. Differ. Equ. 264, 2011–2027 (2018)

    MathSciNet  MATH  Google Scholar 

  • Weinberger, H.: Long-time behavior of a class of biology models. SIAM J. Math. Anal. 13, 353–396 (1982)

    MathSciNet  MATH  Google Scholar 

  • Weinberger, H.: On spreading speeds and traveling waves for growth and migration models in a periodic habitat. J. Math. Biol. 45, 511–548 (2002)

    MathSciNet  MATH  Google Scholar 

  • Winkler, M.: Boundedness in the higher-dimensional parabolic-parabolic chemotaxis system with logistic source. Commun. Partial Differ. Equ. 35, 1516–1537 (2010)

    MathSciNet  MATH  Google Scholar 

  • Winkler, M.: Aggregation vs. global diffusive behavior in the higher-dimensional Keller–Segel model. J. Differ. Equ. 248, 2889–2905 (2010)

    MathSciNet  MATH  Google Scholar 

  • Winkler, M.: Global asymptotic stability of constant equilibria in a fully parabolic chemotaxis system with strong logistic dampening. J. Differ. Equ. 257, 1056–1077 (2014)

    MathSciNet  MATH  Google Scholar 

  • Winkler, M.: Large-data global generalized solutions in a chemotaxis system with tensor-valued sensitivities. SIAM J. Math. Anal. 47, 3092–3115 (2015)

    MathSciNet  MATH  Google Scholar 

  • Winkler, M.: How far can chemotactic cross-diffusion enforce exceeding carrying capacities? J. Nonlinear Sci. 24, 809–855 (2014)

    MathSciNet  MATH  Google Scholar 

  • Xu, G.: Boundedness and asymptotically stability to chemotaxis system with competitive kinetics and nonlocal terms. Preprint

  • Yu, H., Guo, Q., Zheng, S.: Finite time blow-up of nonradial solutions in an attraction-repulsion chemotaxis system. Nonlinear Anal. Real World Appl. 34, 335–342 (2017)

    MathSciNet  MATH  Google Scholar 

  • Zhang, Q., Liu, X., Yang, X.: Global existence and asymptotic behavior of solutions to a two-species chemotaxis system with two chemicals. J. Math. Phys. 58, 111504 (2017)

    MathSciNet  MATH  Google Scholar 

  • Zhang, Q.: Competitive exclusion for a two-species chemotaxis system with two chemicals. Appl. Math. Lett. 83, 27–32 (2018)

    MathSciNet  MATH  Google Scholar 

  • Zheng, J.: Boundedness in a two-species quasilinear chemotaxis system with two chemicals. Topol. Methods Nonlinear Anal. 49, 463–480 (2017)

    MathSciNet  MATH  Google Scholar 

  • Zheng, P.: Asymptotic stability in a chemotaxis-competition system with indirect signal production. Discrete Contin. Dyn. Syst. Ser. A 41, 1207–1223 (2021)

    MathSciNet  MATH  Google Scholar 

  • Zheng, P., Hu, R.: Boundedness and stabilization in a two-species attraction-repulsion chemotaxis-competition system. Preprint

  • Zheng, P., Mu, C.: Global boundedness in a two-competing-species chemotaxis system with two chemicals. Acta Appl. Math. 148, 157–177 (2017)

    MathSciNet  MATH  Google Scholar 

  • Zheng, P., Mu, C., Mi, Y.: Global stability in a two-competing-species chemotaxis system with two chemicals. Differ. Integral Equ. 31, 547–558 (2018)

    MathSciNet  MATH  Google Scholar 

  • Zheng, P., Xiang, Y., Xing, J.: On a two-species chemotaxis system with indirect signal production and general competition terms. Math. Models Methods Appl. Sci. 32, 1385–1430 (2022)

    MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors would like to deeply thank the editor and anonymous reviewers for their insightful and constructive comments. Pan Zheng is deeply grateful to Professor Renjun Duan for his help and support at The Chinese University of Hong Kong. The work is partially supported by the National Natural Science Foundation of China (Grant Nos.: 11601053, 12271064), Natural Science Foundation of Chongqing (Grant No. cstc2019jcyj-msxmX0082), the Science and Technology Research Program of Chongqing Municipal Education Commission (Grant No. KJZD-K202200602), China-South Africa Young Scientist Exchange Project in 2020, The Hong Kong Scholars Program (Grant Nos.: XJ2021042, 2021-005), Young Hundred Talents Program of CQUPT in 2022-2024 and Chongqing Postgraduate Research and Innovation Project in 2022 (Grant No.: CYS22451).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pan Zheng.

Ethics declarations

Conflict of interest

The authors declare that this work does not have any conflicts of interests.

Additional information

Communicated by Anthony Blochv.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendix A. Proof of Lemma 2.1

Appendix A. Proof of Lemma 2.1

Proof

The ideas of proof are similar to (Winkler 2010a, Lemma 1.1) and (Stinner et al. 2014b, Lemma 2.1). For reader’s convenience, we give the sketch of the proof.

(i) Existence. Under the assumptions of Lemma 2.1, we claim that for all \(L>0\) there exists \(T=T(L)>0\) such that \(\Vert u_{0}\Vert _{L^{\infty }(\Omega )}\le L,\Vert w_{0}\Vert _{L^{\infty }(\Omega )}\le L,\Vert v_{0}\Vert _{W^{1,q}(\Omega )}\le L\) and \(\Vert z_{0}\Vert _{W^{1,q}(\Omega )}\le L\), then system (1.1) is classically solvable in \(\Omega \times (0,T)\). As a consequence of a standard extension argument, this will imply the existence of a maximal existence time \(T_{\max }\) satisfying (2.1).

Now, we prove the local existence of solutions for system (1.1) when \(\tau =1\) and \(\tau =0\), respectively.

When \(\tau =1\), according to the well-known Neumann heat semigroup \(\big (e^{t\Delta }\big )_{t\ge 0}\) in (Winkler 2010b, Lemma 3.1), we can pick \(K>0\) such that \(\Vert e^{t\Delta }v\Vert _{W^{1,q}(\Omega )}\le K\Vert v\Vert _{W^{1,q}(\Omega )}\) and \(\Vert e^{t\Delta }z\Vert _{W^{1,q}(\Omega )}\le K\Vert z\Vert _{W^{1,q}(\Omega )}\) for all \(v,z\in W^{1,q}(\Omega )\). For small \(T\in (0, 1)\) to be fixed below, we introduce the Banach space

$$\begin{aligned} X:= & {} C^{0}\bigg ([0,T];C^{0}({\overline{\Omega }})\bigg )\times C^{0}\bigg ([0,T];W^{1,q}(\Omega )\bigg )\times C^{0}\bigg ([0,T];C^{0}({\overline{\Omega }})\bigg )\\{} & {} \times C^{0}\bigg ([0,T];W^{1,q}(\Omega )\bigg ), \end{aligned}$$

and the close subset

$$\begin{aligned} \begin{aligned} F:=\big \{&(u,v,w,z)\in X\mid \Vert u\Vert _{L^{\infty }((0,T);L^{\infty }(\Omega ))}\le L+1,\Vert v\Vert _{L^{\infty }((0,T);W^{1,q}(\Omega ))}\le KL+1,\\&\Vert w\Vert _{L^{\infty }((0,T);L^{\infty }(\Omega ))}\le L+1,\Vert z\Vert _{L^{\infty }((0,T);W^{1,q}(\Omega ))}\le KL+1\big \}. \end{aligned} \end{aligned}$$

For \((u,v,w,z)\in F\) and \(t\in (0,T)\), we define the mapping

$$\begin{aligned} \begin{gathered} \Psi _{1}(u,v,w,z)(t):=\left( \begin{array}{c} \Psi _{11}(u,v, w,z)(t) \\ \Psi _{12}(u,v,w,z)(t) \\ \Psi _{13}(u,v,w,z)(t) \\ \Psi _{14}(u,v,w,z)(t) \end{array}\right) :=\\ \left( \begin{array}{c} e^{td_{1}\Delta }u_{0}-\xi _{1}\int _{0}^{t} e^{(t-s)d_{1}\Delta }\nabla \cdot \big (u(s)\nabla v(s)\big ) ds+\chi _{1}\int _{0}^{t}e^{(t-s)d_{1}\Delta }\nabla \cdot \big (u(s)\nabla z(s)\big )ds+\int _{0}^{t} e^{(t-s)d_{1}\Delta }g_{1}(u,w)ds\\ e^{t(d_{2}\Delta -1)}v_{0}+\int _{0}^{t} e^{(t-s)(d_{2}\Delta -1)}w(s)ds\\ e^{td_{3}\Delta }w_{0}-\xi _{2}\int _{0}^{t} e^{(t-s)d_{3}\Delta }\nabla \cdot \big (w(s)\nabla z(s)\big ) ds+\chi _{2}\int _{0}^{t}e^{(t-s)d_{3}\Delta }\nabla \cdot \big (w(s)\nabla v(s)\big )ds+\int _{0}^{t} e^{(t-s)d_{3}\Delta }g_{2}(u,w)ds\\ e^{t(d_{4}\Delta -1)}z_{0}+\int _{0}^{t} e^{(t-s)(d_{4}\Delta -1)}u(s)ds \end{array}\right) . \end{gathered} \end{aligned}$$

Then, we have

$$\begin{aligned} \begin{aligned} \Vert \Psi _{11}(u,v,w,z)(t)\Vert _{L^{\infty }(\Omega )}\le&\Vert e^{t d_{1}\Delta } u_{0}\Vert _{L^{\infty }(\Omega )}+\xi _{1} \int _{0}^{t}\bigg \Vert e^{(t-s)d_{1}\Delta }\nabla \cdot \bigg (u(s)\nabla v(s)\bigg )\bigg \Vert _{L^{\infty }(\Omega )}ds\\&+\chi _{1} \int _{0}^{t}\bigg \Vert e^{(t-s)d_{1}\Delta }\nabla \cdot \bigg (u(s) \nabla z(s)\bigg )\bigg \Vert _{L^{\infty }(\Omega )}ds\\&+\int _{0}^{t}\bigg \Vert e^{(t-s)d_{1}\Delta }g_{1}(u,w)\bigg \Vert _{L^{\infty }(\Omega )}ds\\ \le&\xi _{1} \int _{0}^{t}\bigg \Vert e^{(t-s)d_{1}\Delta }\nabla \cdot \bigg (u(s) \nabla v(s)\bigg )\bigg \Vert _{L^{\infty }(\Omega )}ds\\&+\chi _{1} \int _{0}^{t}\bigg \Vert e^{(t-s)d_{1}\Delta }\nabla \cdot \bigg (u(s) \nabla z(s)\bigg )\bigg \Vert _{L^{\infty }(\Omega )}ds\\&+L+T\cdot \Vert g_{1}(u,w)\Vert _{L^{\infty }((-R-1,R+1))}, \end{aligned} \end{aligned}$$
(7.1)

where by the maximum principle

$$\begin{aligned} \begin{aligned} \Vert e^{td_{1}\Delta }u_{0}\Vert _{L^{\infty }(\Omega )} \le \Vert u_{0}\Vert _{L^{\infty }(\Omega )}\le L \end{aligned} \end{aligned}$$
(7.2)

and

$$\begin{aligned} \begin{aligned} \int _{0}^{t}\Vert e^{td_{1}\Delta }g_{1}(u,w)\Vert _{L^{\infty }(\Omega )}ds&\le \int _{0}^{t}\Vert g_{1}(u,w)\Vert _{L^{\infty }(\Omega )}ds\\&\le T\cdot \Vert g_{1}(u,w)\Vert _{L^{\infty }((-R-1,R+1))}, \end{aligned} \end{aligned}$$
(7.3)

for all \(t\in (0,T)\). Furthermore, by picking any \(p>\frac{nq}{q-n}\) and then \(\alpha \in \big (\frac{n}{p}, \frac{1}{2}-\frac{n}{2}(\frac{1}{q}-\frac{1}{p})\big )\), we obtain \(p\alpha >n\) and the fractional power \(A^{\alpha }\) of the sectorial operator \(A:=-d_{1}\Delta +1\) with Neumann data in \(L^{p}(\Omega )\) satisfies \(\Vert \phi \Vert _{L^{\infty }(\Omega )}\le C\Vert A^{\alpha } \phi \Vert _{L^{p}(\Omega )}\) as well as \(\Vert A^{\alpha } e^{\rho d_{1}\Delta }\phi \Vert _{L^{p}(\Omega )}\le C \rho ^{-\alpha }\Vert \phi \Vert _{L^{p}(\Omega )}\) for all \(\phi \in C^{\infty }_{0}(\Omega )\) (cf. Henry (1981)). Here and below, \(C_{i}(i=1,2,\cdots ,23)\) denote generic positive constants. Therefore, by \(T<1,\alpha \in \bigg (\frac{n}{p}, \frac{1}{2}-\frac{n}{2}(\frac{1}{q}-\frac{1}{p})\bigg )\) and \(\Vert e^{\rho d_{1}\Delta }\nabla \cdot \psi \Vert _{L^{p}(\Omega )}\le C \rho ^{-\frac{1}{2}-\frac{n}{2}(\frac{1}{q}-\frac{1}{p})}\Vert \psi \Vert _{L^{q}(\Omega )}\) for \(\rho <1\) and all \({\mathbb {R}}-\)valued \(\psi \in C^{\infty }_{0}(\Omega )\) (cf. Weinberger (1982)), we have

$$\begin{aligned} \begin{aligned}&\xi _{1} \int _{0}^{t}\bigg \Vert e^{(t-s)d_{1}\Delta }\nabla \cdot \bigg (u(s) \nabla v(s)\bigg )\bigg \Vert _{L^{\infty }(\Omega )}ds\\&\le C_{1}\int _{0}^{t}\bigg \Vert A^{\alpha }e^{(t-s)d_{1}\Delta }\nabla \cdot \bigg (u(s)\nabla v(s)\bigg )\bigg \Vert _{L^{p}(\Omega )}ds\\&\le C_{2}\int _{0}^{t}(t-s)^{-\alpha }\bigg \Vert e^{\frac{t-s}{2}d_{1}\Delta }\nabla \cdot \bigg (u(s)\nabla v(s)\bigg )\bigg \Vert _{L^{p}(\Omega )}ds\\&\le C_{3}\int _{0}^{t}(t-s)^{-\alpha } \cdot (t-s)^{-\frac{1}{2}-\frac{n}{2}(\frac{1}{q}-\frac{1}{p})}\bigg \Vert u(s)\nabla v(s)\bigg \Vert _{L^{q}(\Omega )}ds\\&\le C_{4}T^{\frac{1}{2}-\alpha -\frac{n}{2}(\frac{1}{q}-\frac{1}{p})}\cdot (L+1)\cdot (KL+1) \end{aligned} \end{aligned}$$
(7.4)

for all \(t\in (0,T)\). Similarly, we obtain

$$\begin{aligned} \begin{aligned}&\chi _{1} \int _{0}^{t}\bigg \Vert e^{(t-s)d_{1}\Delta }\nabla \cdot \bigg (u(s) \nabla z(s)\bigg )\bigg \Vert _{L^{\infty }(\Omega )}ds\\&\quad \le C_{5}T^{\frac{1}{2}-\alpha -\frac{n}{2}(\frac{1}{q}-\frac{1}{p})}\cdot (L+1)\cdot (KL+1) \end{aligned} \end{aligned}$$
(7.5)

for all \(t\in (0,T)\). For the term \(\Vert \Psi _{13}(u,v,w,z)(t)\Vert _{L^{\infty }(\Omega )}\), we can use the similar way.

For \(\Vert \Psi _{12}(u,v,w,z)(t)\Vert _{L^{\infty }(\Omega )}\), we have

$$\begin{aligned} \begin{aligned} \Vert \Psi _{12}(u,v,w,z)(t)\Vert _{W^{1,q}(\Omega )}&\le e^{-t}\Vert e^{td_{1}\Delta }v_{0}\Vert _{W^{1,q}(\Omega )}+C_{6}\int _{0}^{t}(t-s)^{-\frac{1}{2}}\Vert w(s)\Vert _{L^{q}(\Omega )}ds\\&\le K\Vert v_{0}\Vert _{W^{1,q}(\Omega )}+C_{7}\int _{0}^{t}(t-s)^{-\frac{1}{2}}\Vert w(s)\Vert _{L^{\infty }(\Omega )}ds\\&\le KL+C_{8}T^{\frac{1}{2}}\cdot (L+1) \end{aligned} \end{aligned}$$
(7.6)

for all \(t\in (0,T)\). Similarly, we can estimate the term \(\Vert \Psi _{14}(u,v,w,z)(t)\Vert _{L^{\infty }(\Omega )}\). Then, it follows from (7.1)–(7.6) that if we fix \(T_{0} \in (0, 1)\) small enough such that \(T\in (0, T_{0})\), then \(\Psi _{1}\) maps F into itself.

Moreover, using the same ideas with (7.4), for \((u,v,w,z)\in F\) and \(({\bar{u}},{\bar{v}},{\bar{w}},{\bar{z}})\in F\), we get

$$\begin{aligned}{} & {} \left\| \Psi _{11}(u, v, w, z)(t)-\Psi _{11}({\bar{u}}, {\bar{v}}, {\bar{w}}, {\bar{z}})(t)\right\| _{L^{\infty }(\Omega )}\nonumber \\{} & {} \le C_{9} \int _{0}^{t}\left\| A^{\alpha } e^{(t-l)d_{1}\Delta }\nabla \cdot \bigg (u(s)\nabla v(s)-{\bar{u}}(s)\nabla {\bar{v}}(s)\bigg )\right\| _{L^{p}(\Omega )}ds\nonumber \\{} & {} +C_{9}\int _{0}^{t}\left\| A^{\alpha } e^{(t-s)d_{1}\Delta }\nabla \cdot \bigg (u(s)\nabla z(s)-{\bar{u}}(s)\nabla {\bar{z}}(s)\bigg )\right\| _{L^{p}(\Omega )}ds\nonumber \\{} & {} +\int _{0}^{t}\left\| e^{(t-s) d_{1}\Delta }\bigg (g_{1}(u, w)-g_{1}({\bar{u}}, {\bar{w}})\bigg )\right\| _{L^{\infty }(\Omega )}ds\nonumber \\{} & {} \le C_{10} \int _{0}^{t}(t-s)^{-\alpha -\frac{1}{2}-\frac{n}{2}\left( \frac{1}{q}-\frac{1}{p}\right) }\Vert u(s)\nabla v(s)-{\bar{u}}(s)\nabla {\bar{v}}(s)\Vert _{L^{q}(\Omega )}ds\nonumber \\{} & {} +C_{10} \int _{0}^{t}(t-s)^{-\alpha -\frac{1}{2}-\frac{n}{2}\left( \frac{1}{q}-\frac{1}{p}\right) }\Vert u(s)\nabla z(s)-{\bar{u}}(s)\nabla {\bar{z}}(s)\Vert _{L^{q}(\Omega )}ds\nonumber \\{} & {} +\int _{0}^{t} \Vert g_{1}(u,w)-g_{1}({\bar{u}}, {\bar{w}})\Vert _{L^{\infty }(\Omega )}ds\nonumber \\{} & {} \le C_{11} T^{\frac{1}{2}-\alpha -\frac{n}{2}\left( \frac{1}{q}-\frac{1}{p}\right) }\bigg ((L+1)+(KL+1)\bigg )\cdot \Vert (u, v, w, z)-({\bar{u}}, {\bar{v}}, {\bar{w}}, {\bar{z}})\Vert _{X} \nonumber \\{} & {} +T\cdot \left\| g_{1}^{\prime }\right\| _{L^{\infty }((-L-1, L+1))}\cdot \Vert (u, v, w, z)-({\bar{u}}, {\bar{v}}, {\bar{w}}, {\bar{z}})\Vert _{X}. \end{aligned}$$
(7.7)

Similarly, we have

$$\begin{aligned} \begin{aligned}&\left\| \Psi _{13}(u, v, w, z)(t)-\Psi _{13}({\bar{u}}, {\bar{v}}, {\bar{w}}, {\bar{z}})(t)\right\| _{L^{\infty }(\Omega )}\\ \le&C_{12} T^{\frac{1}{2}-\alpha -\frac{n}{2}\left( \frac{1}{q}-\frac{1}{p}\right) }\bigg ((L+1)+(KL+1)\bigg )\cdot \Vert (u, v, w, z)-({\bar{u}}, {\bar{v}}, {\bar{w}}, {\bar{z}})\Vert _{X} \\&+T\cdot \left\| g_{2}^{\prime }\right\| _{L^{\infty }((-L-1, L+1))}\cdot \Vert (u, v, w, z)-({\bar{u}}, {\bar{v}}, {\bar{w}}, {\bar{z}})\Vert _{X} \end{aligned} \end{aligned}$$
(7.8)

and

$$\begin{aligned} \begin{aligned}&\left\| \Psi _{12}(u, v, w, z)(t)-\Psi _{12}({\bar{u}}, {\bar{v}}, {\bar{w}}, {\bar{z}})(t)\right\| _{L^{\infty }(\Omega )}\\&\le C_{13} \int _{0}^{t}(t-s)^{-\frac{1}{2}}\Vert w(s)-{\bar{w}}(s)\Vert _{L^{q}(\Omega )}ds \\&\le C_{14} T^{\frac{1}{2}}\cdot \Vert (u, v, w, z)-({\bar{u}},{\bar{v}},{\bar{w}},{\bar{z}})\Vert _{X} \end{aligned} \end{aligned}$$
(7.9)

as well as

$$\begin{aligned} \begin{aligned}&\left\| \Psi _{14}(u,v,w,z)(t)-\Psi _{14}({\bar{u}}, {\bar{v}}, {\bar{w}}, {\bar{z}})(t)\right\| _{L^{\infty }(\Omega )} \\ {}&\le C_{15} T^{\frac{1}{2}}\cdot \Vert (u, v, w, z)-({\bar{u}},{\bar{v}},{\bar{w}},{\bar{z}})\Vert _{X} \end{aligned} \end{aligned}$$
(7.10)

for all \(t\in (0,T)\), which shows that \(\Psi \) is a contraction mapping if \(T\in (0, T_{0})\) is small enough. Then, by using the Banach fixed point theorem, we know that the existence of some \((u,v,w,z)\in F\) such that \(\Psi _{1}(u,v,w,z)=(u,v,w,z)\). Once again using standard arguments involving semigroup estimates, it can easily be checked that in fact (uvwz) lies in the asserted regularity class and is a classical solution of (1.1) in \(\Omega \times (0,T)\). Since \(g_{1}(0,0)\ge 0\) and \(g_{2}(0,0)\ge 0\) hold, the maximum principle moreover ensures that uwvz are nonnegative.

When \(\tau =0\), we introduce the Banach space

$$\begin{aligned} {\bar{X}}:=C^{0}\big ([0,T];C^{0}({\overline{\Omega }})\big )\times C^{0}\big ([0,T];C^{0}({\overline{\Omega }})\big ), \end{aligned}$$

and consider the close subset

$$\begin{aligned} \begin{aligned} {\bar{F}}:=\bigg \{(u,w)\in X\mid \Vert u\Vert _{L^{\infty }((0,T);L^{\infty }(\Omega ))}\le L+1,\Vert w\Vert _{L^{\infty }((0,T);L^{\infty }(\Omega ))}\le L+1\bigg \}, \end{aligned} \end{aligned}$$

where \(T\in (0,1)\) is small. Similarly, we define the mapping

$$\begin{aligned} \begin{gathered} \Psi _{2}(u,v,w,z)(t):=\left( \begin{array}{c} \Psi _{21}(u,v, w,z)(t) \\ \Psi _{22}(u,v,w,z)(t) \\ \end{array}\right) \\:=\left( \begin{array}{c} e^{td_{1}\Delta }u_{0}+\int _{0}^{t} e^{(t-s)d_{1}\Delta }\nabla \cdot \bigg (\chi _{1}u(s)\nabla z(s)-\xi _{1}u(s)\nabla v(s)\bigg )ds+\int _{0}^{t} e^{(t-s)d_{1}\Delta }g_{1}(u,w)ds\\ e^{td_{3}\Delta }w_{0}+\int _{0}^{t} e^{(t-s)d_{3}\Delta }\nabla \cdot \bigg (\chi _{2}w(s)\nabla v(s)-\xi _{2}w(l)\nabla z(s)\bigg )ds+\int _{0}^{t} e^{(t-s)d_{3}\Delta }g_{2}(u,w)ds \end{array}\right) . \end{gathered} \end{aligned}$$

for \((u,w)\in {\bar{F}}\) and \(t\in (0,T)\), where \(\big (e^{td_{i}\Delta }\big )_{t\ge 0}\) denotes the Neumann heat semigroup. From the second and fourth equation in (1.1), we have \(-d_{1}\Delta v+v=w\) and \(-d_{3}\Delta z+z=u\) under homogeneous Neumann boundary conditions. According to the same methods in case of \(\tau =1\), we get that \(\Psi _{2}\) is a contraction mapping on \({\bar{F}}\) if \(T\in (0, T_{0})\) is sufficiently small. Hence, the Banach fixed point theorem implies the existence of some \((u,w)\in {\bar{F}}\) such that \(\Psi _{2}(u,w)=(u,w)\). Moreover, by applying the similar arguments and the strong maximum principle, we deduce that (uw) is nonnegative. And by the strong elliptic maximum principle applied to the second and fourth equation in (1.1), we also obtain the nonnegativity of (vz).

(ii) Uniqueness. Proceeding as in Gajewski and Zacharias (1998), for given \(T>0\) and two solutions \((u,v,w,z),({\bar{u}},{\bar{v}},{\bar{w}},{\bar{z}})\) in \(\Omega \times (0,T)\), we fix \(T_{1}\in (0,T)\) and set \(U:=u-{\bar{u}},V:=v-{\bar{v}},W:=w-{\bar{w}},Z:=z-{\bar{z}}\). By applying straightforward testing procedures to (1.1), we have

$$\begin{aligned} \begin{aligned} \frac{{\text {d}}}{{\text {d}}t}\int _{\Omega }U^{2}+2d_{1}\int _{\Omega }|\nabla U|^{2}&=2\xi _{1}\int _{\Omega }U\nabla v\cdot \nabla U+2\xi _{1}\int _{\Omega }{\bar{u}}\nabla U\cdot \nabla V\\&\quad -2\chi _{1}\int _{\Omega }U\nabla z\cdot \nabla U -2\chi _{1}\int _{\Omega }{\bar{u}}\nabla U\cdot \nabla Z \\&\quad +2\int _{\Omega }\bigg (g_{1}(u,w)-g_{1}({\bar{u}}, {\bar{w}})\bigg )U \end{aligned} \end{aligned}$$
(7.11)

and

$$\begin{aligned} \begin{aligned} \int _{\Omega }W^{2}+2d_{3}\int _{\Omega }|\nabla W|^{2}&=2\xi _{2}\int _{\Omega }W\nabla z\cdot \nabla W+2\xi _{2}\int _{\Omega }{\bar{w}}\nabla W\cdot \nabla Z\\&\quad -2\chi _{2}\int _{\Omega }W\nabla v\cdot \nabla W -2\chi _{2}\int _{\Omega }{\bar{w}}\nabla W\cdot \nabla V \\&\quad +2\int _{\Omega }\bigg (g_{2}(u,w)-g_{2}({\bar{u}},{\bar{w}})\bigg )W \end{aligned} \end{aligned}$$
(7.12)

for all \(t\in (0,T_{1})\).

When \(\tau =1\), by the second and fourth equations in (1.1), we obtain

$$\begin{aligned} \frac{{\text {d}}}{{\text {d}}t}\int _{\Omega }|\nabla V|^{2}+2d_{2}\int _{\Omega }|\Delta V|^{2}+2\int _{\Omega }|\nabla V|^{2}=-2\int _{\Omega }W\Delta V \end{aligned}$$
(7.13)

and

$$\begin{aligned} \frac{{\text {d}}}{{\text {d}}t}\int _{\Omega }|\nabla Z|^{2}+2d_{4}\int _{\Omega }|\Delta Z|^{2}+2\int _{\Omega }|\nabla Z|^{2}=-2\int _{\Omega }U\Delta Z \end{aligned}$$
(7.14)

for all \(t\in (0,T_{1})\). By the Hölder, Young and Gagliardo–Nirenberg inequalities, we get

$$\begin{aligned} \begin{aligned} 2\xi _{1}\int _{\Omega }U\nabla v\cdot \nabla U&\le 2\xi _{1}\bigg (\int _{\Omega }|\nabla U|^{2}\bigg )^{\frac{1}{2}}\cdot \bigg (\int _{\Omega }|\nabla v|^{q}\bigg )^{\frac{1}{q}}\cdot \bigg (\int _{\Omega } U^{\frac{2 q}{q-2}}\bigg )^{\frac{q-2}{2 q}}\\&\le C_{22}\bigg (\int _{\Omega }|\nabla U|^{2}\bigg )^{\frac{1}{2}+\frac{n}{2 q}}\cdot \bigg (\int _{\Omega }|\nabla v|^{q}\bigg )^{\frac{1}{q}}\cdot \bigg (\int _{\Omega } U^{2}\bigg )^{\frac{q-n}{2q}}\\&\le \frac{d_{1}}{2}\int _{\Omega }|\nabla U|^{2}+C_{16}\int _{\Omega } U^{2}, \end{aligned} \end{aligned}$$
(7.15)

where we have used the fact that \(\int _{\Omega }U=0\) by a simple integration of (1.1), and \(\Vert \nabla v\Vert _{L^{q}(\Omega )}\le C_{17}\) for \(t\in (0,T_{1})\) as well as \(q>n\ge 2\). By using the same method with (7.15), we have

$$\begin{aligned} \begin{aligned} -2\chi _{1}\int _{\Omega }U\nabla z\cdot \nabla U\le \frac{d_{1}}{2}\int _{\Omega }|\nabla U|^{2}+C_{18}\int _{\Omega }U^{2}. \end{aligned} \end{aligned}$$
(7.16)

Furthermore, we have

$$\begin{aligned} \begin{aligned} 2\xi _{1}\int _{\Omega }{\bar{u}}\nabla U\cdot \nabla V\le \frac{d_{1}}{2}\int _{\Omega }|\nabla U|^{2}+C_{17}\int _{\Omega }|\nabla V|^{2} \end{aligned} \end{aligned}$$
(7.17)

and

$$\begin{aligned} \begin{aligned} -2\chi _{1}\int _{\Omega }{\bar{u}}\nabla U\cdot \nabla Z\le \frac{d_{1}}{2}\int _{\Omega }|\nabla U|^{2}+C_{19}\int _{\Omega }|\nabla Z|^{2} \end{aligned} \end{aligned}$$
(7.18)

as well as

$$\begin{aligned} \begin{aligned} 2\int _{\Omega }\bigg (g_{1}(u,w)-g_{1}({\bar{u}}, {\bar{w}})\bigg )U\le C_{20}\int _{\Omega }U^{2}, \end{aligned} \end{aligned}$$
(7.19)

in view of the boundedness of u and \({\bar{u}}\) in \(\Omega \times (0,T_{1})\) and the local Lipschitz continuity of \(g_{1}\). Then, by substituting (7.15)–(7.19) into (7.11), we derive

$$\begin{aligned} \frac{{\text {d}}}{{\text {d}}t}\int _{\Omega }U^{2}\le (C_{16}+C_{18}+C_{20})\int _{\Omega }U^{2}+C_{17}\int _{\Omega }|\nabla V|^{2}+C_{19}\int _{\Omega }|\nabla Z|^{2}. \end{aligned}$$
(7.20)

By using the same method to (7.12), we have

$$\begin{aligned} \frac{{\text {d}}}{{\text {d}}t}\int _{\Omega }W^{2}\le C_{21}\int _{\Omega }W^{2}+C_{21}\int _{\Omega }|\nabla V|^{2}+C_{21}\int _{\Omega }|\nabla Z|^{2}. \end{aligned}$$
(7.21)

By Young’s inequality, we obtain from (7.13) and (7.14) that

$$\begin{aligned} \frac{{\text {d}}}{{\text {d}}t}\int _{\Omega }|\nabla V|^{2}+2\int _{\Omega }|\nabla V|^{2}\le \frac{1}{2d_{2}}\int _{\Omega }W^{2} \end{aligned}$$
(7.22)

and

$$\begin{aligned} \frac{{\text {d}}}{{\text {d}}t}\int _{\Omega }|\nabla Z|^{2}+2\int _{\Omega }|\nabla Z|^{2}\le \frac{1}{2d_{4}}\int _{\Omega }U^{2}. \end{aligned}$$
(7.23)

By combining (7.20)–(7.23), one can find a positive constant \(C_{22}\) such that

$$\begin{aligned} \begin{aligned}&\frac{{\text {d}}}{{\text {d}}t}\bigg (\int _{\Omega }U^{2}+\int _{\Omega }W^{2}+\int _{\Omega }|\nabla V|^{2}+\int _{\Omega }|\nabla Z|^{2}\bigg ) \\&\le C_{22}\bigg (\int _{\Omega }U^{2}+\int _{\Omega } W^{2}+\int _{\Omega }|\nabla V|^{2}+\int _{\Omega }|\nabla Z|^{2}\bigg ). \end{aligned} \end{aligned}$$
(7.24)

When \(\tau =0\), by a straightforward computation, we deduce

$$\begin{aligned} \frac{{\text {d}}}{{\text {d}}t}\int _{\Omega }U^{2}\le C_{23}\bigg (\int _{\Omega } U^{2}+\int _{\Omega }W^{2}+\int _{\Omega }|\nabla V|^{2}+\int _{\Omega }|\nabla Z|^{2}\bigg ) \end{aligned}$$
(7.25)

and

$$\begin{aligned} \frac{{\text {d}}}{{\text {d}}t}\int _{\Omega }W^{2}\le C_{23}\bigg (\int _{\Omega } U^{2}+\int _{\Omega }W^{2}+\int _{\Omega }|\nabla V|^{2}+\int _{\Omega }|\nabla Z|^{2}\bigg ) \end{aligned}$$
(7.26)

as well as

$$\begin{aligned} \int _{\Omega }|\nabla V|^{2}\le \int _{\Omega }W^{2} \end{aligned}$$
(7.27)

and

$$\begin{aligned} \int _{\Omega }|\nabla Z|^{2}\le \int _{\Omega }U^{2}. \end{aligned}$$
(7.28)

Then, by combining (7.25)–(7.28), we have

$$\begin{aligned} \begin{aligned} \frac{{\text {d}}}{{\text {d}}t}\bigg (\int _{\Omega }U^{2}+\int _{\Omega }W^{2}\bigg )\le C_{23}\bigg (\int _{\Omega }U^{2}+\int _{\Omega }W^{2}\bigg ). \end{aligned} \end{aligned}$$
(7.29)

Now with the aid of Grönwall’s lemma, we obtain that \(U\equiv 0,V\equiv 0,W\equiv 0,Z\equiv 0\) in \(\Omega \times (0,T_{1})\). Hence, we obtain \((u,v,w,z)\equiv ({\bar{u}},{\bar{v}},{\bar{w}},{\bar{z}})\) in \(\Omega \times (0,T)\), because \(T_{1}\in (0,T)\) is arbitrary. The proof of Lemma 2.1 is complete. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zheng, P., Hu, R. & Shan, W. On a Two-Species Attraction–Repulsion Chemotaxis System with Nonlocal Terms. J Nonlinear Sci 33, 57 (2023). https://doi.org/10.1007/s00332-023-09912-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00332-023-09912-2

Keywords

Mathematics Subject Classification

Navigation