Skip to main content

Advertisement

Log in

A diffusive virus model with a fixed intracellular delay and combined drug treatments

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

The method of administration of an effective drug treatment to eradicate viruses within a host is an important issue in studying viral dynamics. Overuse of a drug can lead to deleterious side effects, and overestimating the efficacy of a drug can result in failure to treat infection. In this study, we proposed a reaction-diffusion within-host virus model which incorporated information regarding spatial heterogeneity, drug treatment, and the intracellular delay to produce productively infected cells and viruses. We also calculated the basic reproduction number \({\mathcal {R}}_0\) under the assumptions of spatial heterogeneity. We have shown that the infection-free periodic solution is globally asymptotically stable when \({\mathcal {R}}_0<1\), whereas when \({\mathcal {R}}_0>1\) there is an infected periodic solution and the infection is uniformly persistent. By conducting numerical simulations, we also revealed the effects of various parameters on the value of \({\mathcal {R}}_0\). First, we showed that the dependence of \({\mathcal {R}}_0\) on the intracellular delay could be monotone or non-monotone, depending on the death rate of infected cells in the immature stage. Second, we found that the mobility of infected cells or virions could facilitate the virus clearance. Third, we found that the spatial fragmentation of the virus environment enhanced viral infection. Finally, we showed that the combination of spatial heterogeneity and different sets of diffusion rates resulted in complicated viral dynamic outcomes.

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

Similar content being viewed by others

References

  • Bai Z, Peng R, Zhao XQ (2018) A reaction-diffusion malaria model with seasonality and incubation period. J Math Biol 77:201–228

    Article  MathSciNet  MATH  Google Scholar 

  • Browne CJ, Cheng CY (2020) Age-structured viral dynamics in a host with multiple compartments. Math Biosci Eng 17:538–574

    Article  MathSciNet  Google Scholar 

  • Browne CJ, Shu H, Pan X, Wang XS, (2020) Resonance of periodic combination antiviral therapy and intracellular delays in virus model. Bull Math Biol 82:1–29

    Article  MathSciNet  MATH  Google Scholar 

  • Chekroun A, Kuniya T (2020a) An infection age-space structured SIR epidemic model with Neumann boundary condition. Appl Anal 99:1972–1985

    Article  MathSciNet  MATH  Google Scholar 

  • Chekroun A, Kuniya T (2020b) Global threshold dynamics of an infection age-structured SIR epidemic model with diffusion under the Dirichlet boundary condition. J Differ Equ 269:117–148

    Article  MathSciNet  MATH  Google Scholar 

  • Chen SS, Cheng CY, Rong L (2019) Within-host virus models with periodic antiviral therapy. Bull Math Biol 81:4271–4308

    Article  MathSciNet  Google Scholar 

  • Cheynier R, Henrichwark S, Hadida F, Pelletier E, Oksenhendler E, Autran B, Wain-Hobson S (1994) HIV and T cell expansion in splenic white pulps is accompanied by infiltrationof HIV-specific cytotoxic T lymphocytes. Cell 78:373–387

    Article  Google Scholar 

  • Culshaw RV, Ruan S, Webb G (2003) A mathematical model of cell-to-cell spread of HIV-1 that includes a time delay. J Math Biol 46:425–444

    Article  MathSciNet  MATH  Google Scholar 

  • Daners D, Koch Medina P (1992) Abstract Evolution Equations, Periodic Problems and Applications Pitman Res Notes Math Ser, Vol. 279. Longman, Harlow, UK

    MATH  Google Scholar 

  • De Boer RJ, Ribeiro RM, Perelson AS (2010) Current estimates for HIV-1 production imply rapid viral clearance in lymphoid tissues. PLoS Comput Biol 6:1000906

    Article  MathSciNet  Google Scholar 

  • De Leenheer P (2009) Within-host virus models with periodic antiviral therapy. Bull Math Biol 71:189–210

    Article  MathSciNet  MATH  Google Scholar 

  • Dixit NM, Perelson AS (2004) Complex patterns of viral load decay under antiretroviral therapy: influence of pharmacokinetics and intracellular delay. J Theor Biol 226:95–109

    Article  MathSciNet  MATH  Google Scholar 

  • Edgar RS, Stangherlin A, Nagy AD, Nicoll MP, Efstathiou S, O’Neill JS, Reddy AB (2016) Cell autonomous regulation of herpes and influenza virus infection by the circadian clock. Proc Natl Acad Sci USA 113:10085–10090

    Article  Google Scholar 

  • Friedman A (1964) Partial Differential Equations of Parabolic Type. Prentice-Hall, Englewood Cliffs, NJ

    MATH  Google Scholar 

  • Fung HB, Stone EA, Piacenti FJ (2002) Tenofovir disoproxil fumarate: a nucleotide reverse transcriptase inhibitor for the treatment of HIV infection. Clin Ther 24:1515–1548

    Article  Google Scholar 

  • Graw F, Perelson AS (2013) Spatial aspects of HIV infection. In: Schättler H, Friedman A, Kashdan E, Ledzewicz U (eds) Mathematical methods and models in biomedicine, Springer, New York, pp 3–31

    Chapter  Google Scholar 

  • Guo Z, Wang FB, Zou X (2012) Threshold dynamics of an infective disease model with a fixed latent period and non-local infections. J Math Biol 65:1387–1410

    Article  MathSciNet  MATH  Google Scholar 

  • Hess P (1991) Periodic-Parabolic Boundary Value Problems and Positivity. Longman Scientific and Technical, Harlow, UK

    MATH  Google Scholar 

  • Hirsch WM, Smith HL, Zhao XQ (2001) Chain transitivity, attractivity, and strong repellers for semidynamical systems. J Dyn Differ Equ 13:107–131

    Article  MATH  Google Scholar 

  • Hübner W, McNerney GP, Chen P, Dale BM, Gordan RE, Chuang FYS, Li XD, Asmuth DM, Huser T, Chen BK (2009) Quantitative 3D video microscopy of HIV transfer across T cell virological synapses. Science 323:1743–1747

    Article  Google Scholar 

  • Kepler TB, Perelson AS (1998) Drug concentration heterogeneity facilitates the evolution of drug resistance. Proc Natl Acad Sci USA, 95:11514–11519

    Article  MATH  Google Scholar 

  • Lai X, Zou X (2014) Repulsion effect on superinfecting virions by infected cells. Bull Math Biol 76:2806–2833

    Article  MathSciNet  MATH  Google Scholar 

  • Lenbury Y, Ouncharoen R, Tumrasvin N (2000) Higher-dimensional separation principle for the analysis of relaxation oscillations in nonlinear systems: application to a model of HIV infection. IMA J Math Appl Med Biol 17:243–261

    Article  MATH  Google Scholar 

  • Li MY, Shu H (2010) Impact of intracellular delays and target-cell dynamics on in vivo viral infections. SIAM J Appl Math 70:2434–2448

    Article  MathSciNet  MATH  Google Scholar 

  • Liang X, Zhang L, Zhao XQ (2019) Basic reproduction ratios for periodic abstract functional differential equations (with application to a spatial model for Lyme disease). J Dyn Differ Equ 31:1247–1278

    Article  MathSciNet  MATH  Google Scholar 

  • Lou Y, Zhao XQ (2011) A reaction-diffusion malaria model with incubation period in the vector population. J Math Biol 62:543–568

    Article  MathSciNet  MATH  Google Scholar 

  • Magal P, Ruan S (2018) Theory and Applications of Abstract Semilinear Cauchy Problems. Springer International Publishing, Applied Mathematical Sciences

    Book  MATH  Google Scholar 

  • Magal P, Zhao XQ (2005) Global attractors and steady states for uniformly persistent dynamical systems. SIAM J Math Anal 37:251–275

    Article  MathSciNet  MATH  Google Scholar 

  • Martin RH, Smith HL (1990) Abstract functional differential equations and reaction-diffusion systems. Trans Amer Math Soc 321:1–44

    MathSciNet  MATH  Google Scholar 

  • McFarren A, Lopez L, Williams DW, Veenstra M, Bryan RA, Goldsmith A, Morgenstern A, Bruchertseifer F, Zolla-Pazner S, Gorny MK, Eugenin EA, Berman JW, Dadachova E (2016) A fully human antibody to gp41 selectively eliminates HIV-infected cells that transmigrated across a model human blood brain barrier. AIDS 30:563–572

    Article  Google Scholar 

  • Miller M, Wei S, Parker I, Cahalan M (2002) Two-photon imaging of lymphocyte motility and antigen response in intact lymph node. Science 296:1869–1873

    Article  Google Scholar 

  • Neagu IA, Olejarz J, Freeman M, Rosenbloom DI, Nowak MA, Hill AL (2018) Life cycle synchronization is a viral drug resistance mechanism. PLoS computational biology 14:1005947

    Article  Google Scholar 

  • Nowak MA, May RM (2000) Virus Dynamics. Oxford University Press, New York

    MATH  Google Scholar 

  • Pankavich S, Parkinson C (2016) Mathematical analysis of an in-host model of viral dynamics with spatial heterogeneity. Discrete Contin Dyn Syst B 21:1237–1257

    Article  MathSciNet  MATH  Google Scholar 

  • Perelson AS, Nelson PW (1999) Mathematical analysis of HIV-1 dynamics in vivo. SIAM Rev 41:3–44

    Article  MathSciNet  MATH  Google Scholar 

  • Ren X, Tian Y, Liu L, Liu X (2018) A reaction-diffusion within-host HIV model with cell-to-cell transmission. J Math Biol 76:1831–1872

    Article  MathSciNet  MATH  Google Scholar 

  • Rong L, Feng Z, Perelson AS (2007) Emergence of HIV-1 drug resistance during antiretroviral treatment. Bull Math Biol 69:2027–2060

    Article  MathSciNet  MATH  Google Scholar 

  • Scheiermann C, Gibbs J, Ince L, Loudon A (2018) Clocking in to immunity. Nature reviews Immunology 18:423–437

    Article  Google Scholar 

  • Shen L, Peterson S, Sedaghat AR, McMahon MA, Callender M, Zhang H, Zhou Y, Pitt E, Anderson KS, Acosta EP, Siliciano RF (2008) Dose-response curve slope sets class-specific limits on inhibitory potential of anti-HIV drugs. Nature Med 14:762–766

    Article  Google Scholar 

  • Smith HL (1995) Monotonic Dynamical Systems: An Introduction to the Theory of Competitive and Cooperative systems, Math Surveys Monogr, Vol. 41. American Mathematical Society, Providence

    Google Scholar 

  • Song XY, Neumann AU (2007) Global stability and periodic solution of the viral dynamics. J Math Anal Appl 329:281–297

    Article  MathSciNet  MATH  Google Scholar 

  • Strain MC, Richman DD, Wong JK, Levine H (2002) Spatiotemporal dynamics of HIV propagation. J Theor Biol 218:85–96

    Article  MathSciNet  Google Scholar 

  • Thieme HR (2009) Spectral bound and reproduction number for infinite-dimensional population structure and time heterogeneity. SIAM J Appl Math 70:188–211

    Article  MathSciNet  MATH  Google Scholar 

  • Thieme HR, Zhao XQ (2001) A non-local delayed and diffusive predator-prey model. Nonlinear Anal RWA 2:145–160

    Article  MathSciNet  MATH  Google Scholar 

  • Vaidya NK, Rong L (2017) Modeling pharmacodynamics on HIV latent infection: choice of drugs is key to successful cure via early therapy. SIAM J Appl Math 77:1781–1804

    Article  MathSciNet  MATH  Google Scholar 

  • Wang W, Zhao XQ (2008) Threshold dynamics for compartmental epidemic models in periodic environments. J Dyn Differ Equ 20:699–717

    Article  MathSciNet  MATH  Google Scholar 

  • Wu J (1996) Theory and applications of partial functional differential equations. Springer, New York

    Book  MATH  Google Scholar 

  • Zhang L, Wang Z, Zhao XQ (2015) Threshold dynamics of a time periodic reaction-diffusion epidemic model with latent period. J Differ Equ 258:3011–3036

    Article  MathSciNet  MATH  Google Scholar 

  • Zhao XQ (2017) Basic reproduction ratios for periodic compartmental models with time delay. J Dyn Differ Equ 29:67–82

    Article  MathSciNet  MATH  Google Scholar 

  • Zhao XQ (2017) Dynamical Systems in Population Biology, 2nd ed. Springer, New York

    Book  MATH  Google Scholar 

  • Zhang L, Dailey PJ, Gettie A, Blanchard J, Ho DD (2002) The liver is a major organ for clearing simian immunodeficiency virus in rhesus monkeys. J Virol 76:5271–5273

    Article  Google Scholar 

  • Zhang F, Zhao XQ (2007) A periodic epidemic model in a patchy environment. J Math Anal Appl 325:496–516

    Article  MathSciNet  MATH  Google Scholar 

  • Zhang C, Zhou S, Groppelli E, Pellegrino P, Williams I, Borrow P, Chain BM, Jolly C (2015) Hybrid spreading mechanisms and T cell activation shape the dynamics of HIV-1 infection. PLoS Comput Biol 11:1004179

    Article  Google Scholar 

Download references

Acknowledgements

Research of F.-B. Wang was supported in part by Ministry of Science and Technology, Taiwan; and National Center for Theoretical Sciences, National Taiwan University; and Chang Gung Memorial Hospital (BMRPD18, NMRPD5J0202 and CLRPG2L0051). C.-Y. Cheng was partially supported by the Ministry of Science and Technology of Taiwan, R.O.C. (MOST 109-2115-M-153-003). We sincerely thank Prof. Lei Zhang for helpful discussions on the numerical computation of the basic reproduction number. We are grateful to the Associate Editor, Prof. Yuan Lou, and two anonymous referees for their careful reading and helpful suggestions which led to significant improvements of our original manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chang-Yuan Cheng.

Additional information

Publisher's Note

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

Appendix

Appendix

In this appendix, we will finish the proofs of Lemma 2.1, Lemma 2.2, Lemma 2.3, and Theorem 3.5.

Proof of Lemma 2.1

In view of system (2.16), we see that (2.13)–(2.14) admits a mild solution in the following integral form:

$$\begin{aligned} {\left\{ \begin{array}{ll} u(\cdot ,t) = {\mathbb {T}}(t,0)\phi (0) + \int ^t_0{\mathbb {T}}(t,s)F(s,u_s)ds,~\phi \in {\mathbb {C}}_{\tau }^+,~t>0, \\ u_0(\theta )=\phi (\theta )~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~~,~\theta \in [-\tau ,0]. \end{array}\right. } \end{aligned}$$

Clearly, F is locally Lipschitz continuous. Set \(\bar{a}=\max _{x\in {\bar{\Omega }},t\in [0,\omega ]} a(x,t)\), \({\bar{\beta }}_j=\max _{x\in {\bar{\Omega }},t\in [0,\omega ]} \beta _j(x,t)\) for \(j=V,~I\), and \({\underline{K}}=\min _{x\in {\bar{\Omega }},t\in [0,\omega ]} K(x,t)\). For any \((t,\phi )\in {\mathbb {R}}_+\times {\mathbb {C}}_{\tau }^+\) and \(\xi >0\), we observe that

$$\begin{aligned}&\phi (x,0) + \xi F(t,\phi )(x) \nonumber \\&~~~=\left( \begin{array}{l} \phi _1(x,0) + \xi \left[ \Lambda (x,t) + a(x,t)\phi _1(x,0)\left( 1 - \frac{\phi _1(x,0)}{K(x,t)} \right) \right. \\ ~~~~~~~~~~~~~~~~~~~ - \left. \frac{\beta _V(x,t)\phi _1(x,0)\phi _3(x,0)}{1+\alpha _V(x,t)\phi _3(x,0)} - \frac{\beta _I(x,t)\phi _1(x,0) \phi _2(x,0)}{1+\alpha _I(x,t)\phi _2(x,0)}\right] \\ \phi _2(x,0) + \xi \int _{\Omega } {\mathcal {G}}(t,t-\tau ,x,y) \\ ~~~~~~~~~~~~~~~~~~~ \times \left[ \frac{\beta _V(y,t-\tau )\phi _1(y,-\tau )\phi _3(y,-\tau )}{1+\alpha _V(y,t)\phi _3(y,-\tau )} + \frac{\beta _I(y,t-\tau )\phi _1(y,-\tau )\phi _2(y,-\tau )}{1+\alpha _I(y,t)\phi _2(y,-\tau )}\right] dy \\ \phi _3(x,0) + \xi \gamma (x,t)\phi _2(x,0) \end{array} \right) \\&~~~\ge \left( \begin{array}{l} \phi _1(x,0) \left[ 1 - \xi \left( \frac{\bar{a}}{{\underline{K}}} \phi _1(x,0) + {\bar{\beta }}_V\phi _3(x,0) + {\bar{\beta }}_I\phi _2(x,0)\right) \right] \\ \phi _2(x,0) \\ \phi _3(x,0) \end{array} \right) , \end{aligned}$$

which implies that

$$\begin{aligned} \lim _{\xi \rightarrow 0^+} \frac{1}{\xi }\mathrm{dist}\left( \phi (x,0) + \xi F(t,\phi )(x),~{\mathbb {X}}^+ \right) = 0,~\forall ~(t,\phi )\in {\mathbb {R}}_+\times {\mathbb {C}}_{\tau }^+. \end{aligned}$$

In addition, from (Hess 1991), the operator \( {\mathbb {T}}(t,s):{\mathbb {X}}^+\rightarrow {\mathbb {X}}^+\) is positive, for all \(t\ge s\). Consequently, Corollary 4 in (Martin and Smith 1990) implies that system (2.13)–(2.14) admits a unique non-continuable mild solution \(u(\cdot ,t;\phi )\) on its maximal existence interval \([0,t_{\phi })\) with \(u_0=\phi \), where \(t_{\phi }\le +\infty \), and \(u(\cdot ,t;\phi )\in {\mathbb {C}}_{\tau }^+\) for \(t\in [0,t_{\phi })\). Furthermore, by the analyticity of \( {\mathbb {T}}(t,s)\) with respective to \((t,s)\in {\mathbb {R}}^2\) with \(t>s\), we see that \(u(\cdot ,t;\phi )\) is a classical solution for \(t>\tau \).

Proof of Lemma 2.2

The proof is similar to those in (Zhang et al. 2015, Lemma 2.1), and we only sketch our arguments. For any \({\hat{\varphi }}\in {\mathbb {Y}}^+\), we can show that system (2.18) admits a unique solution \(w(\cdot ,t;{\hat{\varphi }})\) on its maximal existence interval \([0,t_{{\hat{\varphi }}})\), where \(t_{{\hat{\varphi }}}\le +\infty \), and \(w(\cdot ,t;{\hat{\varphi }})\in {\mathbb {Y}}^+\) for \(t\in [0,t_{{\hat{\varphi }}})\). Let

$$\begin{aligned} \bar{p}=\max _{x\in {\bar{\Omega }},t\in [0,\omega ]}p(x,t),\ \forall \ p=\Lambda ,\ a,\ K\ ; {\underline{p}}=\min _{x\in {\bar{\Omega }},t\in [0,\omega ]}p(x,t),\ \forall \ p=n_T,\ a. \end{aligned}$$

Then it follows from (2.18) that

$$\begin{aligned}&\frac{\partial w(x,t)}{\partial t}\le \bigtriangledown \cdot \left[ d_1(x,t)\bigtriangledown w(x,t)\right] + {\bar{\Lambda }}-{\underline{n}}_Tw(x,t)+\bar{a} w(x,t)-{\underline{a}} \frac{(w(x,t))^2}{\bar{K}} \nonumber \\&\quad = \bigtriangledown \cdot \left[ d_1(x,t)\bigtriangledown w(x,t)\right] + {\bar{\Lambda }}-{\underline{n}}_Tw(x,t)+\bar{a} w(x,t)\left( 1 - \frac{w(x,t)}{\hat{K}} \right) , \end{aligned}$$
(6.1)

where \(\hat{K}=\frac{\bar{a}\bar{K}}{{\underline{a}}}\). In view of (6.1) and the comparison principle, we see that

$$\begin{aligned} \limsup _{t\rightarrow \infty }w(x,t)\le \zeta _0,\ \forall \ x\in {\bar{\Omega }}, \end{aligned}$$

where \( \zeta _0>0\) is the unique positive root of the following quadratic equation

$$\begin{aligned} {\bar{\Lambda }}-{\underline{n}}_T\zeta +\bar{a} \zeta \left( 1 - \frac{\zeta }{\hat{K}}\right) =0. \end{aligned}$$

Thus, solutions of system (2.18) are ultimately bounded, and hence, \(t_{{\hat{\varphi }}}= +\infty \).

Therefore, we may define the solution maps \(S(t):{\mathbb {Y}}^+\rightarrow {\mathbb {Y}}^+\) associated with system (2.18). Since system (2.18) is scalar, we see that S(t) is strongly monotone for \(t>0\), see, e. g., pages 56 and 133 in (Smith 1995). Observing that the reaction term in (2.18)

$$\begin{aligned} f(x,t,w):=\Lambda (x,t)-n_T(x,t)w + a(x,t)w\left( 1 - \frac{w}{K(x,t)} \right) \end{aligned}$$

is strictly subhomogeneous in the sense that \(f(x,t,\vartheta w)>\vartheta f(x,t,w)\) for \(\vartheta \in (0,1)\) and \(w>0\). Then the rest of the proofs are same to those in (Zhang et al. 2015, Lemma 2.1), and we omit the details.

Remark 6.1

In this remark, we further discuss the global dynamics of system (2.18) on \(C([-\tau ,0],{\mathbb {Y}}^+)\). For any \(\varphi \in C([-\tau ,0],{\mathbb {Y}}^+)\), we let \(w(\cdot ,t;\varphi (\cdot ,0))\) be the solution of system (2.18) with initial value \(w(\cdot ,0;\varphi (\cdot ,0))=\varphi (\cdot ,0)\in {\mathbb {Y}}^+\). Define

$$\begin{aligned} w_t(x,\theta ;\varphi )= \left\{ \begin{array}{llll} w\left( x,t+\theta ;\varphi (\cdot ,0)\right) ,\quad \text {if}\ t+\theta>0,\ t>0,\ x\in {\bar{\Omega }},\ \theta \in [-\tau ,0],\\ \varphi (\cdot ,t+\theta ),\quad ~~~~~~~~~~~\text {if} \ t+\theta \le 0,\ t>0,\ x\in {\bar{\Omega }},\ \theta \in [-\tau ,0]. \end{array} \right. \!\!\!\nonumber \\ \end{aligned}$$
(6.2)

Then \(w_t\) defines the solution map of system (2.18) on \(C([-\tau ,0],{\mathbb {Y}}^+)\). Let \({\bar{P}}(\varphi )=w_\omega (\varphi )\) be the Poincaré map associated with system (2.18) on \(C([-\tau ,0],{\mathbb {Y}}^+)\). By Lemma 2.2, it follows that for any \(\varphi \in C([-\tau ,0],{\mathbb {Y}}^+)\), we see that the omega limit set of \(\varphi \) for \({\bar{P}}\) is \(\{u^*_{1,0}\}\), where \(u^*_{1,0}\in C([-\tau ,0],{\mathbb {Y}}^+)\) is defined by \(u^*_{1,0}(\cdot ,\theta )=u^*_1(\cdot ,0+\theta )=u^*_1(\cdot ,\theta )\), for \(\theta \in [-\tau ,0]\).

Proof of Lemma 2.3

In view of Lemma 2.1, we see that for any \(\phi \in {\mathbb {C}}_{\tau }^+\), system (2.13)–(2.14) has a unique solution \(u(\cdot ,t;\phi )\) on its maximal existence interval \([0,t_{\phi })\) with \(u_0=\phi \), where \(t_{\phi }\le +\infty \). In view of Lemma 2.2, we see that system (2.18) admits a unique positive \(\omega \)-periodic solution \(u_1^*(x,t)\) which is globally attractive in \({\mathbb {Y}}^+\). We further observe that the \(u_1\)-equation in system (2.13) is dominated by the equation (2.18). Thus, the comparison principle implies that there is a constant \(B_1>0\) such that for any \(\phi \in {\mathbb {C}}_{\tau }^+\), there exists a \(t_1(\phi )\) such that

$$\begin{aligned} u_1(x,t)\le B_1,\ \forall \ x\in {\bar{\Omega }},\ t\ge t_1(\phi ). \end{aligned}$$
(6.3)

We are ready to adopt the similar ideas in (Thieme and Zhao 2001), see also (Guo et al. 2012; Zhang et al. 2015), to establish the ultimate boundedness of \(u_i(x,t)\) for \(i=2,3\). Set \(U_i(t)=\int _{\Omega }u_i(x,t)dx\) for \(i=1,2,3\). Integrating the first equation in (2.13) over \(x\in \Omega \), and using the boundary condition, we see that

$$\begin{aligned}&\frac{dU_1(t)}{dt} \le \int _{\Omega }\Lambda (x,t)dx- \int _{\Omega }n_T(x,t)u_1(x,t)dx+\int _{\Omega }a(x,t)u_1(x,t)dx\nonumber \\&\quad - \int _{\Omega } \left[ \frac{\beta _V(x,t)u_1(x,t)u_3(x,t)}{1+\alpha _V(x,t)u_3(x,t)} + \frac{\beta _I(x,t)u_1(x,t) u_2(x,t)}{1+\alpha _I(x,t)u_2(x,t)} \right] dx,\ \forall \ t>0.\nonumber \\ \end{aligned}$$
(6.4)

Let

$$\begin{aligned} \Lambda _1:= & {} \max _{t\in [0,\omega ]}\int _{\Omega }\Lambda (x,t)dx+B_1\max _{t\in [0,\omega ]}\int _{\Omega }a(x,t)dx>0,\\ {\underline{p}}= & {} \min _{x\in {\bar{\Omega }},t\in [0,\omega ]}p(x,t),\ \forall \ p=n_T,\ b,\ c, \end{aligned}$$

and

$$\begin{aligned} {\bar{\gamma }}=\max _{x\in {\bar{\Omega }},t\in [0,\omega ]}\gamma (x,t). \end{aligned}$$

Then (6.3) and (6.4) imply that

$$\begin{aligned}&\int _{\Omega } \left[ \frac{\beta _V(x,t)u_1(x,t)u_3(x,t)}{1+\alpha _V(x,t)u_3(x,t)} + \frac{\beta _I(x,t)u_1(x,t) u_2(x,t)}{1+\alpha _I(x,t)u_2(x,t)} \right] dx\nonumber \\&\quad \le \Lambda _1 -{\underline{n}}_T U_1(t) - \frac{dU_1(t)}{dt},\ \forall \ t\ge t_1(\phi ). \end{aligned}$$
(6.5)

Integrating the \(u_2\)-equation in (2.13), and using the facts that \({\mathcal {G}}(t,t-\tau ,x,y)\) is bounded and (6.5), we obtain

$$\begin{aligned} \frac{d}{dt}U_2(t) \le -{\underline{b}}U_2(t) + k_1 - k_2U_1(t-\tau ) -k_3\frac{d}{dt}U_1(t-\tau ),\ \forall \ t\ge t_1(\phi )+\tau ,\nonumber \\ \end{aligned}$$
(6.6)

where \(k_i\), \(i=1,2,3\), are some positive values independent of \(\phi \). Similarly, integrating the \(u_3\)-equation in (2.13) over \(\Omega \) yields

$$\begin{aligned} \frac{d}{dt}U_3(t) \le {\bar{\gamma }}U_2(t) - {\underline{c}}U_3(t),\ \forall \ t>0. \end{aligned}$$
(6.7)

Let \(m:=\frac{2{\bar{\gamma }}}{{\underline{b}}}\) and \(k_4:=\min \{\frac{k_2}{k_3},\frac{1}{2}{\underline{b}},{\underline{c}}\}>0\). Then it follows from (6.6) and (6.7) that

$$\begin{aligned}&\frac{d}{dt}\left[ mk_3U_1(t-\tau )+mU_2(t)+U_3(t) \right] \nonumber \\&\quad \le mk_3\frac{d}{dt}U_1(t-\tau ) - 2{\bar{\gamma }}U_2(t)+mk_1-mk_2U_1(t-\tau )-mk_3\frac{d}{dt}U_1(t-\tau )\nonumber \\&\qquad +{\bar{\gamma }}U_2(t)-{\underline{c}}U_3(t) \nonumber \\&\quad \le mk_1-k_4\left[ mk_3U_1(t-\tau )+mU_2(t)+U_3(t) \right] ,\ \forall \ t\ge t_1(\phi )+\tau . \end{aligned}$$

This implies that we can find a \(t_2(\phi )\) with \(t_2(\phi )\ge t_1(\phi )\) such that

$$\begin{aligned} mk_3U_1(t-\tau )+mU_2(t)+U_3(t) \le \frac{mk_1}{k_4}+1,\ \forall \ t\ge t_2(\phi )+\tau . \end{aligned}$$
(6.8)

Using the positivity of solutions, (6.3), (6.8) and the facts that \({\mathcal {G}}(t,t-\tau ,x,y)\), \(\beta _V(x,t)\), \(\beta _I(x,t)\) are bounded, it derives from the second equation of (2.13) that there exist positive constants \(k_5\) and \(k_6\) such that

$$\begin{aligned} \frac{\partial u_2(x,t)}{\partial t}\le & {} \bigtriangledown \cdot [d_2(x,t)\bigtriangledown u_2(x,t)] - {\underline{b}}u_2(x,t) + \int _{\Omega } {\mathcal {G}}(t,t-\tau ,x,y) \times \nonumber \\&\left[ \beta _V(y,t-\tau )u_1(y,t-\tau )u_3(y,t-\tau )\right. \nonumber \\&\left. + \beta _I(y,t-\tau )u_1(y,t-\tau )u_2(y,t-\tau )\right] dy \nonumber \\\le & {} \bigtriangledown \cdot \left[ d_2(x,t)\bigtriangledown u_2(x,t)\right] \nonumber \\&- {\underline{b}}u_2(x,t) + k_5 U_3(t-\tau )+k_6U_2(t-\tau ) \nonumber \\\le & {} \bigtriangledown \cdot \left[ d_2(x,t)\bigtriangledown u_2(x,t)\right] \nonumber \\&- {\underline{b}}u_2(x,t) + \left( k_5+\frac{k_6}{m}\right) \left( \frac{mk_1}{k_4}+1\right) ,\ \forall \ t\ge t_2(\phi )+2\tau . \end{aligned}$$
(6.9)

From (6.9) and the comparison principle, it follows that

$$\begin{aligned} \limsup _{t\rightarrow \infty }u_2(x,t)\le \frac{mk_5+k_6}{m{\underline{b}}}\left( \frac{mk_1}{k_4}+1\right) ,\ \forall \ x\in {\bar{\Omega }}. \end{aligned}$$

Then there exists \(t_3(\phi )\ge t_2(\phi )+2\tau \) and \(B_2>\frac{mk_5+k_6}{m{\underline{b}}}\left( \frac{mk_1}{k_4}+1\right) \) such that

$$\begin{aligned} u_2(x,t)\le B_2,\ \forall \ x\in {\bar{\Omega }},\ t\ge t_3(\phi ). \end{aligned}$$
(6.10)

In view of the third equation in (2.13), and (6.10), we see that

$$\begin{aligned} \frac{\partial u_3(x,t)}{\partial t} \le \bigtriangledown \cdot \left[ d_3(x,t)\bigtriangledown u_3(x,t)\right] + B_2{\bar{\gamma }}- {\underline{c}}u_3(x,t),\ \forall \ x\in \Omega ,\ t\ge t_3(\phi ).\nonumber \\ \end{aligned}$$
(6.11)

From (6.11) and the comparison principle, it follows that

$$\begin{aligned} \limsup _{t\rightarrow \infty }u_3(x,t)\le \frac{B_2{\bar{\gamma }}}{{\underline{c}}},\ \forall \ x\in {\bar{\Omega }}. \end{aligned}$$
(6.12)

In view of (6.3), (6.10) and (6.12), we see that solutions of system (2.13)–(2.14) are ultimately bounded. Thus, \(t_{\phi }= +\infty \) for any \(\phi \in {\mathbb {C}}_{\tau }^+\).

Define the solution maps \(\Phi (t):{\mathbb {C}}_{\tau }^+\rightarrow {\mathbb {C}}_{\tau }^+\) associated with system (2.13)–(2.14) by

$$\begin{aligned} \Phi (t)(\phi )(x,\theta )\!=\!u_t(\phi )(x,\theta )\!=\!u(x,t\!+\!\theta ;\phi ),\ \forall \ \phi \in {\mathbb {C}}_{\tau }^+ ,\ t\ge 0, \ \theta \in [\!-\!\tau ,0],\ x\in \Omega . \end{aligned}$$

We can further show that \(\{\Phi (t)\}_{t\ge 0}\) is an \(\omega \)-periodic semiflow on \({\mathbb {C}}_{\tau }^+\) in the sense defined in (Zhao 2017a, Sect. 3.1). From the ultimate boundedness of \(u_i(x,t)\), for \(i=1,2,3\), it follows that \(\Phi (t):{\mathbb {C}}_{\tau }^+\rightarrow {\mathbb {C}}_{\tau }^+\) is point dissipative. In view of (Wu 1996, Theorem 2.1.8), we see that \(\Phi (t):{\mathbb {C}}_{\tau }^+\rightarrow {\mathbb {C}}_{\tau }^+\) is compact for each \(t>\tau \). Then it follows from (Magal and Zhao 2005, Theorem 2.9) that the Poincaré map \(\Phi (\omega )\) admits a global compact attractor in \({\mathbb {C}}_{\tau }^+\).

Theorem 3.5 states a threshold-type result on the global dynamics of system (2.13)–(2.14) in terms of \({\mathcal {R}}_0\). In order to prove Theorem 3.5, we first give a primary estimation for the solution.

Lemma 6.2

Let \(u(\cdot ,t;\phi )\) be the solution of system (2.13)–(2.14) on \([0,\infty )\) with \(u_0=\phi \in {\mathbb {C}}_{\tau }^+\). Then the following statements are valid:

  1. (i)

    If there exists some \(t_0\ge 0\) such that \(u_i(\cdot ,t_0;\phi )\not \equiv 0\) for some \(i\in \{2,3\}\), then \(u_i(x,t;\phi )>0\), \(\forall \ t>t_0\), \(x\in {\bar{\Omega }}\).

  2. (ii)

    For any \(\phi \in {\mathbb {C}}_{\tau }^+\), we have

    $$\begin{aligned} u_1(x,t;\phi )>0,\ \forall t>0,\ x\in {\bar{\Omega }}, \end{aligned}$$
    (6.13)

    and

    $$\begin{aligned} \liminf \limits _{t\rightarrow \infty }u_1(x,t;\phi )\ge {\hat{\varrho }}, \ \text{ uniformly } \text{ for }\ x\in {\bar{\Omega }}, \end{aligned}$$
    (6.14)

    where \( {\hat{\varrho }}\) is a positive constant independent of \(\phi \).

Proof

Part (i). From the second and third equations of (2.13) and the positivity of solutions, one can easily see that for a given \(\phi \in {\mathbb {C}}_{\tau }^+\), \(u_2(x,t;\phi )\) and \(u_3(x,t;\phi )\), respectively, satisfy

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\partial u_2(x,t)}{\partial t} \ge \bigtriangledown \cdot \left[ d_2(x,t)\bigtriangledown u_2(x,t)\right] - b(x,t)u_2(x,t),\ (x,t)\in \Omega \times [0,\infty ),\\ \left[ d_2(x,t)\bigtriangledown u_2(x,t)\right] \cdot \nu =0,\ x\in \partial \Omega ,\ t>0, \end{array}\right. }\nonumber \\ \end{aligned}$$
(6.15)

and

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\partial u_3(x,t)}{\partial t} \ge \bigtriangledown \cdot \left[ d_3(x,t)\bigtriangledown u_3(x,t)\right] - c(x,t) u_3(x,t),\ (x,t)\in \Omega \times [0,\infty ),\\ \left[ d_3(x,t)\bigtriangledown u_3(x,t)\right] \cdot \nu =0,\ x\in \partial \Omega ,\ t>0. \end{array}\right. }\nonumber \\ \end{aligned}$$
(6.16)

By the parabolic maximum principle, see, e.g. (Hess 1991), it follows that the results in Part (i) hold.

Part (ii). By the positivity of solutions (see Lemma 2.1), it follows that

$$\begin{aligned} u_1(x,t;\phi )\ge 0,\ \forall t>0,\ x\in {\bar{\Omega }}. \end{aligned}$$

Suppose that (6.13) was not true, that is, there exists \(x_0\in {\bar{\Omega }}\) and \(t_0\in (0,\infty )\) such that \(u_1(x_0,t_0;\phi )=0\). Let \({\mathcal {T}}_0>0\) be such that \(t_0<{\mathcal {T}}_0\). Then \((x_0, t_0)\in {\bar{\Omega }} \times [0,{\mathcal {T}}_0]\) and \(u_1\) attains its minimum on \({\bar{\Omega }} \times [0,{\mathcal {T}}_0]\) at the point \((x_0, t_0)\). In view of the first equation of (2.13), we have

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\partial u_1(x,t)}{\partial t}\ge \bigtriangledown \cdot \left[ d_1(x,t)\bigtriangledown u_1(x,t)\right] - \left[ n_T(x,t)+\frac{ a(x,t)u_1(x,t)}{K(x,t)}\right. \\ \ \ \ \ \ \ \ \ \ \ \ \ \quad \left. +\beta _V(x,t)u_3(x,t)\! \quad +\!\beta _I(x,t)u_2(x,t)\right] u_1(x,t),\ x\in \Omega ,\ t\in (0,{\mathcal {T}}_0],\\ \left[ d_1(x,t)\bigtriangledown u_1(x,t)\right] \cdot \nu =0,\ x\in \partial \Omega ,\ t\in (0,{\mathcal {T}}_0]. \end{array}\right. }\!\!\!\!\nonumber \\ \end{aligned}$$
(6.17)

In case \(x_0\in \partial \Omega \), we apply the Hopf boundary lemma, see e.g. (Hess 1991, Sect. II.13), and get \(\frac{\partial u_1}{\partial \nu }(x_0, t_0)<0\), which contradicts the boundary condition of \(u_1\). In case \(x_0\in \Omega \), we apply the strong maximum principle, see e.g. (Hess 1991, Section II.13), to obtain that \(u_1(x,t)\equiv u_1(x_0, t_0)=0,\ \forall \ (x,t)\in {\bar{\Omega }}\times [0,{\mathcal {T}}_0]\). Inserting this result into the first equation of (2.13), it leads to \(\Lambda (x,t)\equiv 0,\ \forall \ (x, t)\in {\bar{\Omega }} \times [0,{\mathcal {T}}_0]\), a contradiction. Thus, (6.13) is true.

Next, we establish the result in (6.14). By Lemma 2.3, it follows that solutions of (2.13)–(2.14) are ultimately bounded. Thus, there is a constant \(\hat{B}>0\) such that for any \(\phi \in {\mathbb {C}}_{\tau }^+\), there exists a \(t_1>0\) such that

$$\begin{aligned} u_i(x,t)\le \hat{B},\ \forall \ x\in {\bar{\Omega }},\ t\ge t_1,\ i=1,2,3. \end{aligned}$$

For \(x\in \Omega ,\ t\ge t_1\), it follows from the first equation of (2.13) that

$$\begin{aligned} \frac{\partial u_1(x,t)}{\partial t}\ge & {} \bigtriangledown \cdot \left[ d_1(x,t)\bigtriangledown u_1(x,t)\right] + \Lambda (x,t) \nonumber \\&- \left[ n_T(x,t) +a(x,t)\frac{u_1(x,t)}{K(x,t)}+\beta _V(x,t)u_3(x,t)+\beta _I(x,t)u_2(x,t)\right] u_1(x,t) \nonumber \\\ge & {} \bigtriangledown \cdot \left[ d_1(x,t)\bigtriangledown u_1(x,t)\right] + \Lambda (x,t) \nonumber \\&- \left[ n_T(x,t) +a(x,t)\frac{\hat{B}}{K(x,t)}+\beta _V(x,t)\hat{B}+\beta _I(x,t)\hat{B}\right] u_1(x,t)\nonumber \\\ge & {} \bigtriangledown \cdot \left[ d_1(x,t)\bigtriangledown u_1(x,t)\right] + {\underline{\Lambda }}-C_0u_1(x,t), \end{aligned}$$
(6.18)

where

$$\begin{aligned} C_0=\max _{x\in {\bar{\Omega }},t\in [0,\omega ]}\left[ n_T(x,t) +a(x,t)\frac{\hat{B}}{K(x,t)}+\beta _V(x,t)\hat{B}+\beta _I(x,t)\hat{B}\right] , \end{aligned}$$

and \({\underline{\Lambda }}=\min _{x\in {\bar{\Omega }},t\in [0,\omega ]}\Lambda (x,t)\). By (6.18) and the comparison principle, we see that

$$\begin{aligned} \liminf \limits _{t\rightarrow \infty }u_1(x,t;\phi )\ge \frac{{\underline{\Lambda }}}{C_0}, \ \text{ uniformly } \text{ for }\ x\in {\bar{\Omega }}, \end{aligned}$$

which establishes (6.14). \(\square \)

Proof of Theorem 3.5

Recall that \(\Phi (t)(\cdot ):=u_t(\cdot ):{\mathbb {C}}_{\tau }^+\rightarrow {\mathbb {C}}_{\tau }^+\) is an \(\omega \)-periodic semiflow of system (2.13)–(2.14) (see Lemma 2.3), and let \(Q:=\Phi (\omega )\) be the associated Poincaré map on \({\mathbb {C}}_{\tau }^+\).

Part (i). In the case where \({\mathcal {R}}_0<1\), Lemma 3.1 implies that \(r(P(\omega ))<1\), and hence \(\mu =\frac{\ln [r(P(\omega ))]}{\omega }<0\). Consider the following system with a parameter \(\varepsilon >0\):

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\partial v_1(x,t)}{\partial t} = \bigtriangledown \cdot [d_2(x,t)\bigtriangledown v_1(x,t)] - b(x,t)v_1(x,t) \\ \ \ \ \ \ \ \ \ \ \ + \int _{\Omega } {\mathcal {G}}(t,t-\tau ,x,y)(u_1^*(y,t-\tau )+\varepsilon ) \left[ \beta _V(y,t-\tau )v_2(y,t-\tau ) + \beta _I(y,t-\tau )v_1(y,t-\tau ) \right] dy, \\ \frac{\partial v_2(x,t)}{\partial t} = \bigtriangledown \cdot [d_3(x,t)\bigtriangledown v_2(x,t)] + \gamma (x,t)v_1(x,t) - c(x,t) v_2(x,t),\ \ (x,t)\in \Omega \times [0,\infty ), \\ \left[ d_2(x,t)\bigtriangledown v_1(x,t)\right] \cdot \nu =\left[ d_3(x,t)\bigtriangledown v_2(x,t)\right] \cdot \nu =0,\ \ x\in \partial \Omega ,\ t>0. \end{array}\right. }\nonumber \\ \end{aligned}$$
(6.19)

For any \(\psi \in C([-\tau ,0],{\mathbb {E}})\), let \(v^{\varepsilon }(x,t;\psi )=(v^{\varepsilon }_1(x,t;\psi ),v^{\varepsilon }_2(x,t;\psi ))\) be the unique solution of system (6.19) with \(v^{\varepsilon }_0(\psi )(x,\theta )=\psi (x,\theta )\) for all \(\theta \in [-\tau ,0],\ x\in \Omega \), where

$$\begin{aligned} v_t^{\varepsilon }(\psi )(x,\theta )=(v_1^{\varepsilon }(x,t+\theta ;\psi ),v_2^{\varepsilon }(x,t+\theta ;\psi )), \ \theta \in [-\tau ,0], \end{aligned}$$

for any \(t\ge 0\) and \(x\in \Omega \). Let \(P_{\varepsilon }(\omega ):C([-\tau ,0],{\mathbb {E}})\rightarrow C([-\tau ,0],{\mathbb {E}})\) be the Poincaré map of system (6.19), i.e., \(P_{\varepsilon }(\omega )\psi =v_{\omega }^{\varepsilon }(\psi )\), \(\forall \psi \in C([-\tau ,0],{\mathbb {E}})\), and let \(r(P_{\varepsilon }(\omega ))\) be the spectral radius of \(P_{\varepsilon }(\omega )\). Since \(\lim \limits _{\varepsilon \rightarrow 0}r(P_{\varepsilon }(\omega ))=r(P(\omega ))<1\), we can fix a sufficiently small number \(\varepsilon _0>0\) such that \(r(P_{\varepsilon _0}(\omega ))<1\). By Lemma 3.2, there is a positive \(\omega \)-periodic function \(v^*_{\varepsilon _0}(x,t)\) such that \(v^{\varepsilon _0}(x,t)=e^{\mu _{\varepsilon _0} t}v^*_{\varepsilon _0}(x,t)\) is a solution of system (6.19), where \(\mu _{\varepsilon _0}=\frac{\ln [r(P_{\varepsilon _0}(\omega ))]}{\omega }<0\). In view of the first equation in (2.13), we see that

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\partial u_1(x,t)}{\partial t}\le \bigtriangledown \cdot \left[ d_1(x,t)\bigtriangledown u_1(x,t)\right] + \Lambda (x,t)-n_T(x,t)u_1(x,t)\\ \ \ \ \ \ \ \ \ \ \ \ \ \ \quad + a(x,t)u_1(x,t)\left( 1 - \frac{u_1(x,t)}{K(x,t)} \right) ,\ (x,t)\in \Omega \times [0,\infty ),\\ \left[ d_1(x,t)\bigtriangledown u_1(x,t)\right] \cdot \nu =0,\ x\in \partial \Omega ,\ t>0. \end{array}\right. } \end{aligned}$$

By the global attractivity of \(u_1^*(x,t)\) for system (2.18) (see Lemma 2.2) and the comparison principle, there exists a sufficiently large integer \(n_1>0\) such that \(n_1\omega >\tau \) and

$$\begin{aligned} u_1(x,t)\le u_1^*(x,t)+\varepsilon _0, \ \forall \ t\ge n_1\omega -\tau ,\ x\in {{\bar{\Omega }}}. \end{aligned}$$

From the second and third equations in (2.13), it follows that for any \(t\ge n_1\omega \), \(x\in \Omega \), we have

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\partial u_2(x,t)}{\partial t}\le \bigtriangledown \cdot [d_2(x,t)\bigtriangledown u_2(x,t)] - b(x,t)u_2(x,t) \\ \ \ \ \ \ \ \ \ \ \ + \int _{\Omega } {\mathcal {G}}(t,t-\tau ,x,y) (u_1^*(y,t-\tau )+\varepsilon _0)\left[ \beta _V(y,t-\tau )u_3(y,t-\tau )+\beta _I(y,t-\tau )u_2(y,t-\tau )\right] dy, \\ \\ \frac{\partial u_3(x,t)}{\partial t} = \bigtriangledown \cdot [d_3(x,t)\bigtriangledown u_3(x,t)] + \gamma (x,t)u_2(x,t) - c(x,t) u_3(x,t), \end{array}\right. }\!\!\!\!\nonumber \\ \end{aligned}$$
(6.20)

with the no flux boundary condition. For any \(\phi \in {\mathbb {C}}_{\tau }^+\), we may find a \(m_1>0\) such that

$$\begin{aligned} \left( u_2(x,t;\phi ),u_3(x,t;\phi )\right) \le m_1e^{\mu _{\varepsilon _0} t}v^*_{\varepsilon _0}(x,t), \ \forall \ t\in \left[ n_1\omega -\tau ,n_1\omega \right] ,\ x\in {\bar{\Omega }}. \end{aligned}$$

By the comparison theorem for abstract functional differential equation, see e.g. (Martin and Smith 1990, Proposition 1), it follows that

$$\begin{aligned} \left( u_2(x,t;\phi ),u_3(x,t;\phi )\right) \le m_1e^{\mu _{\varepsilon _0} t}v^*_{\varepsilon _0}(x,t), \ \forall t\ge n_1\omega , \ x\in {\bar{\Omega }}. \end{aligned}$$

Since \(\mu _{\varepsilon _0}<0\), we see that \(\lim \limits _{t\rightarrow \infty }(u_2(x,t;\phi ),u_3(x,t;\phi ))=(0,0)\) uniformly for \(x\in {\bar{\Omega }}\). Then the \(u_1\)-equation in (2.13) is asymptotic to system (2.18).

Next, we are ready to use the theory of internally chain transitive sets, see e.g. (Hirsch et al. 2001; Zhao 2017b), to prove the following result

$$\begin{aligned} \lim \limits _{t\rightarrow \infty }\left( u_1(x,t;\phi )-u_1^*(x,t)\right) =0,\ \text{ uniformly } \text{ for }\ x\in {\bar{\Omega }}, \end{aligned}$$

where \(u_1^*(x,t)\) is a globally attractive solution of system (2.18). Let \({\mathcal {J}}\) be the omega limit set of \(\phi \in {\mathbb {C}}_{\tau }^+\) for \(Q:=\Phi (\omega )\). Since \(\lim \limits _{t\rightarrow \infty }u_i(x,t;\phi )=0 \ (i=2,3)\) uniformly for \(x\in {\bar{\Omega }}\), we have \({\mathcal {J}}={\mathcal {J}}_1\times \{\hat{0}\}\times \{\hat{0}\}\), where \({\hat{0}}(\cdot ,\theta )=0\), \(\forall \ \theta \in [-\tau ,0]\). By (Hirsch et al. 2001, Lemma 2.1), see also (Zhao 2017b, Lemma 1.2.1), we note that \({\mathcal {J}}\) is an internally chain transitive set for Q, and we can further show that \({\mathcal {J}}_1\) is an internally transitive chain set for \({\bar{P}}\), where \({\bar{P}}\) is the Poincaré map associated with system (2.18) on \(C([-\tau ,0],{\mathbb {Y}}^+)\) (see Remark 6.1). By (6.14) in Lemma 6.2, we see that \({\mathcal {J}}_1\ne \{\hat{0}\}\). Since \(u^*_{1,0}\) defined in Remark 6.1 is globally attractive in \(C([-\tau ,0],{\mathbb {Y}}^+)\) for \({\bar{P}}\), we have \({\mathcal {J}}_1\cap W^s(\{u^*_{1,0}\})\ne \emptyset \), where \(W^s(\{u^*_{1,0}\})\) is the stable set of \(\{u^*_{1,0}\}\). By (Hirsch et al. 2001, Theorem 3.1), see also (Zhao 2017b, Theorem 1.2.1), we get \({\mathcal {J}}_1\subseteq \{u^*_{1,0}\}\), that is, \({\mathcal {J}}_1=\{u^*_{1,0}\}\). This proves

$$\begin{aligned} {\mathcal {J}}={\mathcal {J}}_1\times \{\hat{0}\}\times \{\hat{0}\}=\left\{ \left( u^*_{1,0},\hat{0},\hat{0}\right) \right\} . \end{aligned}$$

Thus, \((u_1^*(x,t),\hat{0},\hat{0})\) is globally attractive for system (2.13)–(2.14) in \({\mathbb {C}}_{\tau }^+\).

Part (ii). In the case where \({\mathcal {R}}_0>1\), Lemma 3.1 implies that \(r(P(\omega ))>1\), and hence \(\mu =\frac{\ln [r(P(\omega ))]}{\omega }>0\).

Let

$$\begin{aligned} {\mathbb {W}}_0:=\left\{ \phi \in {\mathbb {C}}_{\tau }^+ : \phi _2(\cdot ,0)\not \equiv 0 \ \text {and} \ \phi _3(\cdot ,0)\not \equiv 0 \right\} , \end{aligned}$$

and

$$\begin{aligned} \partial {\mathbb {W}}_0:={\mathbb {C}}_{\tau }^+\backslash {\mathbb {W}}_0=\left\{ \phi \in {\mathbb {C}}_{\tau }^+ : \phi _2(\cdot ,0)\equiv 0 \ \text {or} \ \phi _3(\cdot ,0)\equiv 0 \right\} . \end{aligned}$$

Note that for any \(\phi \in {\mathbb {W}}_0\), Lemma 6.2 implies that \(u_i(x,t;\phi )>0, \forall \ t>0,\ x\in {{\bar{\Omega }}} \ (i=1,2,3)\). It follows that \(Q^n{\mathbb {W}}_0\subset {\mathbb {W}}_0\), \(\forall n\in {\mathbb {N}}\). In view of Lemma 2.3, we know that Q admits a global attractor in \({\mathbb {C}}_{\tau }^+ \).

Let

$$\begin{aligned} {\mathbb {M}}_{\partial }:=\{\phi \in \partial {\mathbb {W}}_0 : Q^n\phi \in \partial {\mathbb {W}}_0,\forall n\in {\mathbb {N}}\}, \end{aligned}$$

and \(\omega (\phi )\) be the omega limit set of the orbit \(\gamma ^+=\{Q^n\phi : \forall n\in {\mathbb {N}}\}\). Set \(M=(u^*_{1,0},{\hat{0}},{\hat{0}})\), where \(u^*_{1,0}\in C([-\tau ,0],{\mathbb {Y}}^+)\) is defined by \(u^*_{1,0}(\cdot ,\theta )=u^*_1(\cdot ,0+\theta )=u^*_1(\cdot ,\theta )\), for \(\theta \in [-\tau ,0]\) (see Remark 6.1). Then the following claim indicates that \(\{M\}\) cannot form a cycle for Q in \(\partial {\mathbb {W}}_0\).

Claim 1

For any \({\hat{\phi }}\in {\mathbb {M}}_{\partial }\), the omega limit set \(\omega ({\hat{\phi }})=\{M\}\).

For any given \({\hat{\phi }}\in {\mathbb {M}}_{\partial }\), \(Q^n({\hat{\phi }})\in \partial {\mathbb {W}}_0\), \(\forall \ n\in {\mathbb {N}}\). Thus, for each \(n\in {\mathbb {N}}\), either \(u_2(\cdot ,n\omega ;{\hat{\phi }})\equiv 0\) or \(u_3(\cdot ,n\omega ;{\hat{\phi }})\equiv 0\). It then follows that for each \(t\ge 0\), \(u_2(\cdot ,t;{\hat{\phi }})\equiv 0\) or \(u_3(\cdot ,t;{\hat{\phi }})\equiv 0\). Otherwise, it contradicts Lemma 6.2. In case where \(u_3(\cdot ,t;{\hat{\phi }})\equiv 0\) for \(t\ge 0\). Then it follows from the third equation of (2.13) that \(u_2(\cdot ,t;{\hat{\phi }})\equiv 0\) for \(t\ge 0\). Thus, \(u_1(\cdot ,t;{\hat{\phi }})\) satisfies system (2.18) for \(t\ge 0\), and hence, Lemma 2.2 implies that \(\lim \limits _{t\rightarrow \infty }(u_1(x,t;{\hat{\phi }})-u_1^*(x,t))=0\) uniformly for \(x\in {\bar{\Omega }}\). In case where \(u_3(\cdot ,t_0;{\hat{\phi }})\not \equiv 0\) for some \(t_0\ge 0\). Then it follows from Lemma 6.2 that \(u_3(\cdot ,t;{\hat{\phi }})> 0 ,\ \forall \ t >t_0\). This implies that \(u_2(\cdot ,t;{\hat{\phi }})\equiv 0,\ \forall \ t >t_0\). Inserting this equality into the second equation of (2.13), we derive

$$\begin{aligned} u_1\left( \cdot ,t-\tau ;{\hat{\phi }}\right) u_3\left( \cdot ,t-\tau ;{\hat{\phi }}\right) \equiv 0,\ \forall \ t >t_0, \end{aligned}$$

which is a contradiction. From the above discussions, Claim 1 holds.

Consider the following system with a parameter \(\delta >0\):

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\partial v_1(x,t)}{\partial t} = \bigtriangledown \cdot [d_2(x,t)\bigtriangledown v_1(x,t)] - b(x,t)v_1(x,t) \\ \ \ \ \ \ \ \ \ \ \ + \int _{\Omega } {\mathcal {G}}(t,t-\tau ,x,y)(u_1^*(y,t-\tau )-\delta ) \left[ \frac{ \beta _V(y,t-\tau )v_2(y,t-\tau )}{1+\alpha _V(y,t)\delta } + \frac{\beta _I(y,t-\tau )v_1(y,t-\tau )}{1+\alpha _I(y,t)\delta } \right] dy, \\ \frac{\partial v_2(x,t)}{\partial t} = \bigtriangledown \cdot [d_3(x,t)\bigtriangledown v_2(x,t)] + \gamma (x,t)v_1(x,t) - c(x,t) v_2(x,t),\ \ (x,t)\in \Omega \times [0,\infty ), \\ \left[ d_2(x,t)\bigtriangledown v_1(x,t)\right] \cdot \nu =[d_3(x,t)\bigtriangledown v_2(x,t)]\cdot \nu =0,\ \ x\in \partial \Omega ,\ t>0. \end{array}\right. } \end{aligned}$$
(6.21)

For any \(\psi \in C([-\tau ,0],{\mathbb {E}})\), let \(v^{\delta }(x,t;\psi )=(v^{\delta }_1(x,t;\psi ),v^{\delta }_2(x,t;\psi ))\) be the unique solution of system (6.21) with \(v^{\delta }_0(\psi )(x,\theta )=\psi (x,\theta )\) for all \(\theta \in [-\tau ,0],\ x\in \Omega \), where

$$\begin{aligned} v_t^{\delta }(\psi )(x,\theta )=(v_1^{\delta }(x,t+\theta ;\psi ),v_2^{\delta }(x,t+\theta ;\psi )), \ \theta \in [-\tau ,0], \end{aligned}$$

for any \(t\ge 0\) and \(x\in \Omega \). Let \(P_{\delta }(\omega ):C([-\tau ,0],{\mathbb {E}})\rightarrow C([-\tau ,0],{\mathbb {E}})\) be the Poincáre map of system (6.21) , i.e., \(P_{\delta }(\omega )\psi =v_{\omega }^{\delta }(\psi )\), \(\forall \psi \in C([-\tau ,0],{\mathbb {E}})\), and let \(r(P_{\delta }(\omega ))\) be the spectral radius of \(P_{\delta }(\omega )\). Since \(\lim \limits _{\delta \rightarrow 0}r(P_{\delta }(\omega ))=r(P(\omega ))>1\), we can fix a sufficiently small number \(\delta _0>0\) such that

$$\begin{aligned} \delta _0<\min \limits _{x\in {\bar{\Omega }},t\in [0,\omega ]}u_1^*(x,t) \ \text {and} \ r(P_{\delta _0}(\omega ))>1. \end{aligned}$$

From the continuous dependence of solutions on the initial values, for the fixed \(\delta _0 >0\), there exists a \(\delta ^*>0\) such that for any \(\phi \in {\mathbb {C}}_{\tau }^+\) with \(\Vert \phi -M\Vert \le \delta ^*\), we have \(\Vert \Phi (t)\phi -\Phi (t)M\Vert <\delta _0\) for all \(t\in [0,\omega ]\). We now prove the following claim.

Claim 2

\(\lim \sup _{n\rightarrow \infty }\Vert Q^n(\phi )-M\Vert \ge \delta ^*,\ \forall \ \phi \in {\mathbb {W}}_0\).

Suppose, by contradiction, that \(\lim \sup _{n\rightarrow \infty }\Vert Q^n(\phi _0)-M\Vert < \delta ^*\) for some \(\phi _0\in {\mathbb {W}}_0\). Then there exists \(n_2\ge 1\) such that \(\Vert Q^n(\phi _0)-M\Vert < \delta ^*\) for \(n\ge n_2\). For any \(t\ge n_2\omega \), letting \(t=n\omega +t'\) with \(n=[t/\omega ]\) and \(t'\in [0,\omega )\), we have

$$\begin{aligned} \Vert \Phi (t)\phi _0-\Phi (t)M\Vert =\Vert \Phi (t')(\Phi (\omega )^n(\phi _0))-\Phi (t')M\Vert <\delta _0. \end{aligned}$$
(6.22)

We also note that

$$\begin{aligned} \Phi (t)\phi _0=u_t(\phi _0)=u\left( x,t+\theta ;\phi _0\right) ,\ \Phi (t)M=\left( u^*_{1,t},0,0\right) =\left( u^*_1(x,t+\theta ),0,0\right) ,\ \forall \ \theta \in [-\tau ,0].\nonumber \\ \end{aligned}$$
(6.23)

In view of (6.22), (6.23) and Lemma 6.2, it follows that

$$\begin{aligned} u_1\left( x,t;\phi _0\right) >u_1^*(x,t)-\delta _0 \ \text {and} \ 0<u_j\left( x,t;\phi _0\right) <\delta _0 \ (j=2,3), \end{aligned}$$
(6.24)

for any \(t\ge n_2\omega -\tau \), and \(x\in {\bar{\Omega }}\). By the second and third equations in (2.13) and using (6.24), for \(t\ge n_2\omega \), we see that \(u_2(x,t;\phi _0)\) and \(u_3(x,t;\phi _0)\) satisfy

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{\partial u_2(x,t)}{\partial t} \ge \bigtriangledown \cdot \left[ d_2(x,t)\bigtriangledown u_2(x,t)\right] - b(x,t)u_2(x,t) \\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \quad + \int _{\Omega } {\mathcal {G}}(t,t-\tau ,x,y) \left( u_1^*(y,t-\tau )-\delta _0\right) \left[ \frac{\beta _V(y,t-\tau )u_3(y,t-\tau )}{1+\alpha _V(y,t)\delta _0} + \frac{\beta _I(y,t-\tau )u_2(y,t-\tau )}{1+\alpha _I(y,t)\delta _0}\right] dy, \\ \frac{\partial u_3(x,t)}{\partial t} = \bigtriangledown \cdot [d_3(x,t)\bigtriangledown u_3(x,t)] + \gamma (x,t)u_2(x,t) - c(x,t) u_3(x,t),\\ \left[ d_i(x,t)\bigtriangledown u_i(x,t)\right] \cdot \nu =0,\ x\in \partial \Omega ,\ t>0,\ i=2,3. \end{array}\right. }\quad \end{aligned}$$
(6.25)

By Lemma 3.2, there is a positive \(\omega \)-periodic function \(v^*_{\delta _0}(x,t)\) such that \(v^{\delta _0}(x,t)=e^{\mu _{0} t}v^*_{\delta _0}(x,t)\) is a solution of system (6.21), where \(\mu _{0}=\frac{\ln [r(P_{\delta _0}(\omega ))]}{\omega }>0\). From Lemma 6.2, we see that \(u(x,t;\phi _0)\gg 0\) for all \(t\ge 0\) and \(x\in {\bar{\Omega }}\). Then there exists an \(m_2>0\) such that

$$\begin{aligned} \left( u_2\left( x,t;\phi _0\right) ,u_3\left( x,t;\phi _0\right) \right) \ge m_2e^{\mu _{0} t}v^*_{\delta _0}(x,t), \ \forall t\in \left[ n_2\omega -\tau ,n_2\omega \right] , \ x\in {\bar{\Omega }}.\nonumber \\ \end{aligned}$$
(6.26)

From (6.25), (6.26), and the comparison principle, it follows that

$$\begin{aligned} \left( u_2\left( x,t;\phi _0\right) ,u_3\left( x,t;\phi _0\right) \right) \ge m_2e^{\mu _{0} t}v^*_{\delta _0}(x,t), \ \forall t\ge n_2\omega , \ x\in {\bar{\Omega }}. \end{aligned}$$

Since \(\mu _{0}>0\), it is easy to see that \(u_i(\cdot ,t;\phi _0)\rightarrow +\infty \ (i=2,3)\) as \(t\rightarrow +\infty \). This contradicts (6.24).

The above claim implies that M is an isolated invariant set for Q in \({\mathbb {C}}_{\tau }^+\), and \(W^s(M)\bigcap {\mathbb {W}}_0=\emptyset \), where \(W^s(M)\) is the stable set of M for Q. By (Magal and Zhao 2005, Theorem 3.7), as applied to Q, we know that Q admits a global attractor \(A_0\) in \({\mathbb {W}}_0\). It then follows from (Zhao 2017b, Theorem 1.3.1) that Q is uniformly persistent with respect to \(({\mathbb {W}}_0,\partial {\mathbb {W}}_0)\) in the sense that there exists \({\tilde{\varrho }}>0\) such that

$$\begin{aligned} \liminf \limits _{n\rightarrow \infty }\text {d}(Q^n(\phi ),\partial {\mathbb {W}}_0)\ge {\tilde{\varrho }}, \ \forall \phi \in {\mathbb {W}}_0. \end{aligned}$$
(6.27)

Next, we establish the practical persistence. Since \(A_0=QA_0=\Phi (\omega )A_0\), we have that \(\phi _2(\cdot ,0)>0\) and \(\phi _3(\cdot ,0)>0\) for all \(\phi \in A_0\). Let \(B_0:=\bigcup \limits _{t\in [0,\omega ]}\Phi (t)A_0\). Then \(B_0\subset {\mathbb {W}}_0\) and \(\lim \limits _{t\rightarrow \infty }\text {d}(\Phi (t)\phi ,B_0)=0\), \(\forall \phi \in {\mathbb {W}}_0\). Define a continuous function \(p:{\mathbb {C}}_{\tau }^+\rightarrow {\mathbb {R}}_+\) by

$$\begin{aligned} p(\phi )=\min \left\{ \min \limits _{x\in {\bar{\Omega }}}\phi _2(x,0),\min \limits _{x\in {\bar{\Omega }}}\phi _3(x,0)\right\} , \ \forall \phi =\left( \phi _1,\phi _2,\phi _3\right) \in {\mathbb {C}}_{\tau }^+. \end{aligned}$$

Since \(B_0\) is a compact subset of \({\mathbb {W}}_0\), it follows that \(\inf \limits _{\phi \in B_0}p(\phi )=\min \limits _{\phi \in B_0}p(\phi )>0\). Consequently, there exists a \(\varrho ^*>0\) such that

$$\begin{aligned} \liminf \limits _{t\rightarrow \infty }p(\Phi (t)\phi )\!=\!\liminf \limits _{t\rightarrow \infty }\min \left\{ \min \limits _{x\in {\bar{\Omega }}}u_2(x,t;\phi ),\min \limits _{x\in {\bar{\Omega }}}u_3(x,t;\phi )\right\} \ge \varrho ^*, \ \forall \phi \in {\mathbb {W}}_0.\!\!\!\!\nonumber \\ \end{aligned}$$
(6.28)

Let \(\varrho :=\min \{{\hat{\varrho }},\ \varrho ^*\}\), where \({\hat{\varrho }}\) is given in Lemma 6.2 (ii). Then it follows from Lemma 6.2 and (6.28) that

$$\begin{aligned} \liminf \limits _{t\rightarrow \infty }\min \limits _{x\in {\bar{\Omega }}}u_i(x,t;\phi )\ge \varrho , \ \forall \phi \in {\mathbb {W}}_0 \ (1\le i\le 3). \end{aligned}$$

It remains to prove the existence of a positive periodic solution. By (Zhao 2017b, Theorem 3.5.1 and Remark 3.5.1), it follows that for each \(t>0\), the solution map \(\Phi (t):{\mathbb {C}}_{\tau }^+\rightarrow {\mathbb {C}}_{\tau }^+\) is an \(\alpha \)-contraction with respect to an equivalent norm on \({\mathbb {C}}_{\tau }^+\). Applying (Magal and Zhao 2005, Theorem 4.5) to \(Q=\Phi (\omega )\), it reveals that Q has a fixed point in \(A_0\). Thus, system (2.13)–(2.14) has an \(\omega \)-periodic solution \((\bar{u}_1(\cdot ,t),\bar{u}_2(\cdot ,t),\) \(\bar{u}_3(\cdot ,t))\) with \((\bar{u}_{1t},\bar{u}_{2t},\bar{u}_{3t})\in {\mathbb {W}}_0\). By Lemma 6.2, we see that \((\bar{u}_1(\cdot ,t),\bar{u}_2(\cdot ,t),\) \(\bar{u}_3(\cdot ,t))\) is a (componentwise) positive solution of system (2.13)–(2.14). The proof is finished.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Wang, FB., Cheng, CY. A diffusive virus model with a fixed intracellular delay and combined drug treatments. J. Math. Biol. 83, 19 (2021). https://doi.org/10.1007/s00285-021-01646-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00285-021-01646-7

Keywords

Mathematical Subject Classification:

Navigation