Skip to main content
Log in

Repulsion Effect on Superinfecting Virions by Infected Cells

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

In this paper, the repulsion effect of superinfecting virion by infected cells is studied by a reaction diffusion equation model for virus infection dynamics. In this model, the diffusion of virus depends not only on its concentration gradient but also on the concentration of infected cells. The basic reproduction number, linear stability of steady states, spreading speed and existence of traveling wave solutions for the model are discussed. It is shown that viruses spread more rapidly with the repulsion effect of infected cells on superinfecting virions, than with random diffusion only. For our model, the spreading speed of free virus is not consistent with the minimal traveling wave speed. With our general model, numerical computations of the spreading speed show that the repulsion of superinfecting virion promotes the spread of virus, which confirms, not only qualitatively but also quantitatively, the experimental result of Doceul et al. (Science 327:873–876, 2010).

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

References

  • Adams RA (1975) Sobolev spaces. Academic Press, New York

    MATH  Google Scholar 

  • Amann H (1989) Dynamical theory of quasilinear parabolic equations III: global existence. Math Z 202:219–250

    Article  MathSciNet  MATH  Google Scholar 

  • Amann H (1993) Nonhomogeneous linear and quasilinear elliptic and parabolic boundary value problems. In: Schmeisser HJ, Triebel H (eds) Function spaces, differential operators and nonlinear analysis (Friedrichroda, 1992), vol 133. Teubner-Texte zur Mathematik. Teubner, Stuttgart, pp 9–126

  • Beyn WJ (1990) The numerical computation of connecting orbits in dynamical systems. IMA J Numer Anal 9:379–405

    Article  MathSciNet  Google Scholar 

  • Bonhoeffer S, May RM, Shaw GM, Nowak MA (1997) Virus dynamics and drug therapy. Proc Natl Acad Sci USA 94:6971

    Article  Google Scholar 

  • Chatelin F (1981) The spectral approximation of linear operators with application to the computation of eigenelements of differential and integral operators. SIAM Rev 23:495–522

    Article  MathSciNet  MATH  Google Scholar 

  • Condit RC (2010) Surf and turf: mechanism of enhanced virus spread during poxvirus infection. Viruses 2:1050–1054

    Article  Google Scholar 

  • Doceul V, Hollinshead M, van der Linden L, Smith GL (2010) Repulsion of superinfecting virions: a mechanism for rapid virus spread. Science 327:873–876

    Article  Google Scholar 

  • Gan Q, Xu R, Yang P (2010) Travelling waves of a hepatitis B virus infection model with spatial diffusion and time delay. J Appl Math 75:392–417

    MathSciNet  MATH  Google Scholar 

  • Guo Z, Wang F-B, 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 

  • Haberman R (1998) Elementary applied partial differential equations: with Fourier series and boundary value problems. Prentice Hall, New Jersey

    MATH  Google Scholar 

  • Heffernan JM, Smith RJ, Wahl LM (2005) Perspectives on the basic reproductive ratio. J R Soc Interface 2(4):281–293

    Article  Google Scholar 

  • Ikebe Y (1972) The Galerkin method for the numerical solution of Fredholm integral equations of the 2nd kind. SIAM Rev 14:465–491

    Article  MathSciNet  MATH  Google Scholar 

  • Komarova NL (2007) Viral reproductive strategies: how can lytic viruses be evolutionarily competitive? J Theor Biol 249:766–784

    Article  MathSciNet  Google Scholar 

  • Korobeinikov A (2004) Global properties of basic virus dynamics models. Bull Math Biol 66:879–883

    Article  MathSciNet  Google Scholar 

  • Li B, Weinberger HF, Lewis MA (2005) Spreading speeds as slowest wave speeds for cooperative systems. Math Biosci 196:82–98

    Article  MathSciNet  MATH  Google Scholar 

  • Lewis MA, Li B, Weinberger HF (2002) Spreading speeds and linear determinacy for two-species competition models. J Math Biol 45:219–233

    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 

  • Mothes W, Sherer NM, Jin J, Zhong P (2010) Virus cell-to-cell transmission. J Virol 84:8360–8368

    Article  Google Scholar 

  • Murray JD (2000) Mathematical biology II: spatial models and biomedical applications. Springer, New York

    Google Scholar 

  • Neubert MG, Parker IM (2004) Projecting rates of spread for invasive species. Risk Anal 24:817–831

    Article  Google Scholar 

  • Sattentau Q (2008) Avoiding the void: cell-to-cell spread of human viruses. Nat Rev Microbiol 6:815–826

    Article  Google Scholar 

  • Sattentau Q (2011) The direct passage of animal viruses between cells. Curr Opin Virol 1:396–402

    Article  Google Scholar 

  • Smith HL (1995) Monotone dynamic systems: an introduction to the theory of competitive and cooperative systems. Math Surveys Monogr, vol 41. American Mathematical Society, Providence, RI

  • 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 

  • Vaidya NK, Wang F-B, Zou X (2012) Avian influenza dynamics in wild birds with bird mobility and spatial heterogeneous environment. Discrete Cont Dyn Syst B 17:2829–2848

    Article  MathSciNet  MATH  Google Scholar 

  • Wang K, Wang W (2007) Propagation of HBV with spatial dependence. Math Biosci 210:78–95

    Article  MathSciNet  MATH  Google Scholar 

  • Wang W, Zhao XQ (2011) A nonlocal and time-delayed reaction–diffusion model of dengue transmission. SIAM J Appl Math 71:147–168

    Article  MathSciNet  MATH  Google Scholar 

  • Wang W, Zhao XQ (2012) Basic reproduction numbers for reaction diffusion epidemic models. SIAM J Appl Dyn Sys 11:1652–1673

    Article  MATH  Google Scholar 

  • Weinberger HF, Lewis MA, Li B (2002) Analysis of linear determinacy for spread in cooperative models. J Math Biol 45:183–218

    Article  MathSciNet  MATH  Google Scholar 

  • Weinberger HF, Lewis MA, Li B (2007) Anomalous spreading speeds of cooperative recursion systems. J Math Biol 55:207–222

    Article  MathSciNet  MATH  Google Scholar 

  • Xu R, Ma Z (2009) An HBV model with diffusion and time delay. J Theor Biol 257:499–509

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgments

The authors would like to thank the two anonymous referees for their helpful comments which have led to an improvement to the presentation of the paper.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Xingfu Zou.

Additional information

Research supported by CSC Overseas Doctoral Scholarship (China) and NSERC (Canada).

Appendices

Appendix 1

Proofs of Theorems 2.1 and 2.3

Proof of Theorem 2.1

Note that the system (5) is normally elliptic and triangular (in fact diagonal). According to Theorem 1 Amann (1989) or Theorems 14.4 and 14.6 in Amann (1993), (5)–(4) has a unique classical solution \((T,I,V)\) defined on \([0,\tau _0)\times \Omega \) such that

$$\begin{aligned} (T,I,V)\in {\mathbf C}([0,\tau _0),{\mathbf X}_+)\cap {\mathbf C}^{2,1}((0,\tau _0)\times \bar{\Omega },{{\mathbb {R}}}^3), \end{aligned}$$

where \(\tau _0>0 \) is the maximal value for interval of existence of the solution. The nonnegativity of the solution follows from Theorem 15.1 by Amann (1993). In order to show that \(\tau _0= \infty \), by Theorem 5.2 in Amann (1989) and the nonnegativeness of the solution confirmed above, it suffices to prove that the solution \((T,I,V)\) is bounded above by some positive values.

From the \(T\) and \(I\) equations in (2), we see that

$$\begin{aligned} \frac{\partial }{\partial t}(T+I)&= D_T\Delta (T+I)+h(x)-d_T T-d_{I}I\\&\le D_T\Delta (T+I)+\bar{h}-d_m (T+I), \end{aligned}$$

where \(\bar{h}=\max _{x \in \Omega }h(x)\) and \(d_m=\min \{d_T,d_{I}\}\). By Lemma 1 in Lou and Zhao (2011), \(\bar{h}/d_m\) is the globally attractive steady state for the scalar parabolic equations

$$\begin{aligned} \frac{\partial w(t,x)}{\partial t}&= D_T\Delta w(t,x) +\bar{h}-d_m w(t,x),\quad x\in \Omega , \,t>0, \\ \frac{\partial w(t,x)}{\partial \nu }&= 0, \quad x\in \partial \Omega ,\, t>0. \end{aligned}$$

The parabolic comparison theorem (Smith 1995, Theorem 7.3.4) implies that \(T+I\) is bounded. This together with the nonnegativity of \(T\) and \(I\) further implies that both \(T(t,x)\) and \(I(t,x)\) are bounded. We assume \(0\le T(t,x)\le T_M\), \(0 \le I(t,x) \le I_M\).

Let \(\bar{\gamma }=\max _{x\in \Omega }\{\gamma (x)\}\), and \(V_M= \bar{\gamma } I_M/d_V\). For any given \(I\), define the operator \({\mathcal {P}}\) by

$$\begin{aligned} {\mathcal {P}} V=V_t- \nabla \cdot (D_V(I) \nabla V) - \gamma (x) I + d_V V. \end{aligned}$$

For any solution \((T,I,V)\) of the system (2)–(3)–(4), we have \({\mathcal {P}}V=0\). On the other hand,

$$\begin{aligned} {\mathcal {P}} V_M= d_V V_M- \gamma (x) I \ge d_V V_M-\bar{\gamma }I_M=0 =PV. \end{aligned}$$

On the boundary \(\partial \Omega \), we have \(\frac{\partial V_M}{\partial \nu }=0\). Thus, \(V=V_M\) is an upper solution of the \(V\) equation in the system (2)–(3). By the comparison principle, we obtain that \(V(t,x)\le V_M\). Therefore, the solution \((T,I,V)\) is bounded, and hence, it exists globally.

Proof of Theorem 2.3

Linearizing (2) at \(\bar{E}=(\bar{T}, \bar{I},\bar{V})\) gives

$$\begin{aligned} \frac{\partial }{\partial t}u(t,x)= (\bar{\mathrm{D}} \Delta + \bar{\mathrm{A}})u(t,x), \end{aligned}$$

where

$$\begin{aligned} \bar{\mathrm{D}} =\left( \begin{array}{ccc} D_T &{} 0 &{} 0 \\ 0 &{} D_T &{} 0 \\ 0 &{} 0 &{} D_V(\bar{I}) \end{array} \right) ,\, \bar{\mathrm{A}}= \left( \begin{array}{ccc} -d_T-\beta \bar{V} &{} 0 &{} -\beta \bar{T} \\ \beta \bar{V} &{} -d_{I} &{} \beta \bar{T} \\ 0 &{} \gamma &{} -d_V \end{array} \right) , \quad u=\left( \begin{array}{c} u_1 \\ u_2 \\ u_3 \end{array} \right) . \end{aligned}$$

The corresponding characteristic polynomial of this linearized system is

$$\begin{aligned} \mid \lambda I + \bar{\mathrm{D}} k^2-\bar{\mathrm{A}} \mid =0, \end{aligned}$$
(27)

where \(k\) is the wavenumber, \(\lambda \) is the eigenvalue which determines temporal growth (Murray 2000). The positive steady state \(\bar{E}\) is linearly stable if all eigenvalues have negative real parts.

Substituting the two matrices \(\bar{\mathrm{A}}\) and \(\bar{\mathrm{D}}\) into (27), we obtain

$$\begin{aligned} \left| \begin{array}{ccc} \lambda + D_T k^2+d_T +\beta \bar{V}&{} 0 &{} \beta \bar{T} \\ -\beta \bar{V} &{} \lambda + D_T k^2+d_{I} &{} - \beta \bar{T} \\ 0 &{} -\gamma &{} \lambda + D_V(\bar{I}) k^2+d_V \end{array} \right| =0, \end{aligned}$$

that is,

$$\begin{aligned} \lambda ^3 +b_1(k^2)\lambda ^2+b_2(k^2)\lambda + b_3(k^2)=0, \end{aligned}$$
(28)

where

$$\begin{aligned} b_1(k^2)&= (2D_T+D_V(\bar{I}))k^2+(d_T+d_{I}+d_V)>0,\\ b_2(k^2)&= [D_T^2 +2D_T D_V(\bar{I})]k^4\\&+ \,[D_T d_{I} +D_T (d_T +\beta \bar{V})+D_T d_V+ D_V(\bar{I})(d_T +\beta \bar{V})+ D_T d_V\\&+ \,D_V(\bar{I})d_{I}]k^2 +(d_T +\beta \bar{V})d_{I}+(d_T +\beta \bar{V})d_V >0,\\ b_3(k^2)&= D_T^2D_V({\bar{I}})k^6+[ D_T^2d_V + D_T D_V(\bar{I}) d_{I}+D_T D_V(\bar{I}) (d_T +\beta \bar{V})]k^4\\&+\, [D_T(d_T +\beta \bar{V})d_V+D_V(\bar{I})(d_T +\beta \bar{V})d_{I}]k^2\\&+\, \beta \bar{V}d_{I}d_V >0. \end{aligned}$$
$$\begin{aligned}&b_1(k^2)b_2(k^2)-b_3(k^2)\\&\quad =(2D_T+D_V(\bar{I}))[D_T^2 +2D_T D_V(\bar{I})]k^6\\&\quad \quad \quad +\, \{[(2D_T+D_V(\bar{I}))][D_T d_{I} +(D_T+D_V(\bar{I})) (d_T +\beta \bar{V})\\&\quad \quad \quad +\, 2D_T d_V+D_V(\bar{I})d_{I}]\\&\quad \quad \quad + \,[D_T^2 +2D_T D_V(\bar{I})](d_T+d_{I}+d_V)\}k^4 \\&\quad \quad \quad +\, \{(2D_T+D_V(\bar{I}))[(d_T +\beta \bar{V})d_{I}+(d_T +\beta \bar{V})d_V]\\&\quad \quad \quad +[D_T d_{I} +(D_T+D_V(\bar{I})) (d_T +\beta \bar{V})+2D_T d_V\\&\quad \quad \quad +D_V(\bar{I})d_{I}](d_T+d_{I}+d_V)\}k^2\\&\quad \quad \quad +\, (d_T+d_{I}+d_V)[(d_T +\beta \bar{V})d_{I}+(d_T +\beta \bar{V})d_V]\\&\quad \quad \quad - \,D_T^2D_V({\bar{I}})k^6-[ D_T^2 d_V + D_T D_V(\bar{I}) d_{I}+D_T D_V(\bar{I}) (d_T +\beta \bar{V})]k^4\\&\quad \quad \quad -\, [D_T(d_T +\beta \bar{V})d_V+D_V(\bar{I})(d_T +\beta \bar{V})d_{I}]k^2\\&\quad \quad \quad -\, \beta \bar{V}d_{I}d_V \end{aligned}$$
$$\begin{aligned}&\ge 4D_T^2 D_V(\bar{T})k^6+\{2D_T[D_T d_{I}\\&+\, D_V(\bar{I}) (d_T +\beta \bar{V})+2D_T d_V+ D_V(\bar{I})d_{I}]+[D_T^2+2D_T D_V(\bar{I})]d_T\}k^4\\&+\, \{D_T[(d_T +\beta \bar{V})d_{I}+(d_T +\beta \bar{V})d_V]\\&+ \,[D_T d_{I} +(D_T+D_V(\bar{I})) (d_T +\beta \bar{V})+2D_T d_V+D_V(\bar{I})d_{I}](d_T+d_{I}+d_V)\}k^2\\&+ \,(d_T+d_{I})[(d_T +\beta \bar{V})d_{I}+(d_T +\beta \bar{V})d_V]\\&> 0. \end{aligned}$$

By the Routh–Hurwitz Criterion, we know that all eigenvalues of (28) have negative real parts, and therefore, the positive steady state \(\bar{E}\) is linearly stable if it exists.

Appendix 2

Proof of nonexistence of traveling wavefront solutions for \(c\in (0,c^*)\).

The Jacobian matrix of (16) at \(E'_0\) is

$$\begin{aligned} J_0=\left( \begin{array}{cccc} -1/c &{} 0 &{} -1/c &{} 0 \\ 0 &{} - \rho _1/c &{} 1/c &{} 0 \\ 0 &{} 0 &{} 0 &{} 1 \\ 0 &{} -\rho _2/D_0&{} \rho _3/D_0 &{} c/D_0 \\ \end{array} \right) . \end{aligned}$$

It has an eigenvalue \(\lambda =-1/c\) which is negative for all \(c>0\). So, we only need to consider other eigenvalues which are determined by

$$\begin{aligned} P(\lambda ):=\lambda ^3+a_1\lambda ^2+a_2\lambda +a_3=0, \end{aligned}$$

where

$$\begin{aligned} a_1=\frac{D_0\rho _1-c^2}{c D_0},\, a_2=-\frac{\rho _1+\rho _3}{D_0}<0,\,a_3=\frac{\rho _2-\rho _1\rho _3}{c D_0}>0. \end{aligned}$$

Since \(P(0)=a_3>0\) and \(P(-\infty )=-\infty \), \(P(\lambda )=0\) has a negative root. By the Descarte’s rule of signs and by the Routh–Hurwitz criterion, the other two roots of \(P(\lambda )=0\) are either positive and real or a pair of conjugate complex numbers. In the latter case, the complex eigenvalues imply the oscillations of solutions of (16) near \(E_0'\), implying the \(w\) and \(v\) will take negative values (making solutions biologically meaningless), and thus, (16)–(17) cannot have positive solutions, meaning that (14) cannot have traveling wavefronts connecting \(E_0\) and \(\bar{E}\). Therefore, in order for (14) to have traveling wavefronts connecting \(E_0\) and \(\bar{E}\), it is necessary \(P(\lambda )=0\) to have a pair of positive real roots (counting multiplicity).

Note that \(P'(\lambda )=3\left( \lambda ^2+\frac{2a_1}{3}\lambda +\frac{a_2}{3}\right) \) and \(P'(\lambda )=0\) has a unique positive root

$$\begin{aligned} \lambda ^*=\frac{1}{3}\left( -a_1+\sqrt{a^2_1-3a_2}\right) . \end{aligned}$$

Since \(P(0)=a_3>0\) and \(P'(0)=a_2<0\), we conclude that \(P(\lambda )=0\) has two positive real roots if and only if \(P(\lambda ^*)<0\). From \(P'(\lambda ^*)=0\), we obtain that

$$\begin{aligned} {\lambda ^*}^2+\frac{2a_1}{3}\lambda ^*+\frac{a_2}{3}=0,\quad {\lambda ^*}^3+\frac{2a_1}{3}{\lambda ^*}^2+\frac{a_2}{3}\lambda ^*=0. \end{aligned}$$

Using these equations to simplify the form of \(P(\lambda ^*)\), we obtain

$$\begin{aligned} P(\lambda ^*)&= \frac{a_1}{3}{\lambda ^*}^2+\frac{2a_2}{3}\lambda ^*+a_3\\&= \frac{2}{3}\left( a_2-\frac{a_1^2}{3}\right) \lambda ^*+a_3-\frac{a_1a_2}{9}\\&= \frac{1}{27}\left[ -2 \left( a^2_1-3a_2 \right) ^{3/2}+27a_3+2a^3_1-9a_1a_2\right] . \end{aligned}$$

It then follows that

$$\begin{aligned} P(\lambda ^*)&< 0 \Leftrightarrow 27a_3+2a^3_1-9a_1a_2\le 0, \mathrm{{OR}} \,\,\, 27a_3+2a^3_1-9a_1a_2\\&> 0\, \,\mathrm{{AND}}\,\, 4\left( a^2_1-3a_2\right) ^3>\left( 27a_3+2a^3_1-9a_1a_2\right) ^2;\\ P(\lambda ^*)&> 0\Leftrightarrow 27a_3+2a^3_1-9a_1a_2>0\, \,\mathrm{{AND}}\,\, 4\left( a^2_1-3a_2\right) ^3\\&< \left( 27a_3+2a^3_1-9a_1a_2\right) ^2. \end{aligned}$$

Let \( Q_1(c):=27a_3+2a^3_1-9a_1a_2\), then \(Q_1(c)=\frac{1}{D_0^3c^3}(d_0c^6 +d_1c^4+d_2c^2+d_3)\), where \(d_0=-2,\, d_1=-3D_0(\rho _1-3\rho _3)\), \(d_2=3D_0^2(9\rho _2-6\rho _1\rho _3+\rho ^2_1)>0\) and \(d_3=2\rho ^3_1D_0^3>0\). By the Descarte’s rule of signs, \(\bar{Q}_1(c):=d_0c^6+d_1c^4+d_2c^2+d_3=0\) has a unique positive root \(c_0^*>0\). Since \(\bar{Q}_1(0)=d_3>0\), we see that \(\bar{Q}_1(c)>0\) if \(0<c<c^*_0\), and \(\bar{Q}_1(c)<0\) if \(c>c^*_0\). Furthermore, \(Q_1(c^*_0)=0,\, Q_1(c)>0\) if \(0<c<c^*_0\), and \(Q_1(c)<0\) if \(c>c^*_0\).

Let \(Q_2(c):=4\left( a^2_1-3a_2\right) ^3-\left( 27a_3+2a^3_1-9a_1a_2\right) ^2\), then \( Q_2(c)=\frac{27}{D_0^4c^4}(b_0c^6+b_1c^4+b_2c^2+b_3) \) where \(b_i,\, i=0,1,2,3\), are given by (18).

Note that \(b_0>0,\, b_1>0,\, b_3<0\). Again by the Descarte’s rule of signs, \(Q(c)\) given by (18) has a unique positive root \(c^*>0\). Since \(Q(0)=b_3<0\), we see that \(Q(c)<0\) if \(0<c<c^*\), and \(Q(c)>0\) if \(c>c^*\). Therefore, \(Q_2(c^*)=0,\, Q_2(c)<0\) if \(0<c<c^*\), and \( Q_2(c)>0\) if \(c>c^*\).

Note that \(a_1^2(c)-3a_2>0\) for all \(c>0\). Thus, \(Q_2(c_0^*)=a_1^2(c_0^*)-3a_2>0\), implying \(c^*<c^*_0\). In summary, we have obtained:

$$\begin{aligned} P(\lambda ^*)<0&\Leftrightarrow Q_2(c)\le 0, \,\,\mathrm{{OR}},\,\,\, Q_2(c)>0 \,\, \mathrm{{AND}}\,\,Q_1(c)>0\\&\Leftrightarrow c\ge c^*_0 \,\,(\text{ hence }\,\, c>c^*), \,\,\, \mathrm{{OR}},\,\,\, 0<c<c^*_0 \,\,\mathrm{{AND}}\,\, c>c^*\\&\Leftrightarrow c>c^*; \\ P(\lambda ^*)>0&\Leftrightarrow Q_2(c)>0 \,\, \mathrm{{AND}}\,\,Q_1(c)<0 \\&\Leftrightarrow 0<c<c^*_0 \,\,\mathrm{{AND}} \,\, c<c^*\\&\Leftrightarrow 0<c<c^*. \end{aligned}$$

Thus, for any \(c\in (0, c^*)\), system (14) has no traveling wavefront solutions with speed \(c\) that connects \(E_0\) and \(\bar{E}\).

From the definition of \(Q(c)\), we obtain

$$\begin{aligned} Q(\sqrt{D_{0}\rho _{1}})=D_{0}^{3}\rho _1\left[ 4\rho _1(\rho _1+\rho _3)^3-27 (\rho _2-\rho _1\rho _3)^2\right] . \end{aligned}$$

It is easy to see that \(Q_{3}(\rho _{1}):=4\rho _{1}(\rho _{1}+\rho _{3})^3\) is strictly increasing function of \(\rho _1\), and \(Q_3(0)=0,\, Q_3(+\infty )=+\infty ;\, Q_4(\rho _1):=27(\rho _2-\rho _1\rho _3)^2\) is strictly decreasing function of \(\rho _1\) when \(\rho _2>\rho _1\rho _3\), and \(Q_4(0)=27\rho ^2_2\). Therefore, \(Q(\sqrt{D_0 \rho _1})\) has a unique positive root \(\rho _1^*\), such that \(Q(\sqrt{D_0\rho _1})<0\) for \(0<\rho _1<\rho ^*_1\), and \(Q(\sqrt{D_0\rho _1})>0\) for \(\rho _1>\rho ^*_1\). By the property of \(Q(c)\), we have \(\sqrt{D_0\rho _1}<c^*\) for \(0<\rho _1<\rho ^*_1\); \(\sqrt{D_0\rho _1}>c^*\) for \(\rho _1>\rho ^*_1\) and \(c^*=\sqrt{D_0\rho _1^*}\), where \(\rho _1\) is determined by (21). Therefore, we obtain the information (20) about \(c^*\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lai, X., Zou, X. Repulsion Effect on Superinfecting Virions by Infected Cells. Bull Math Biol 76, 2806–2833 (2014). https://doi.org/10.1007/s11538-014-0033-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-014-0033-9

Keywords

Navigation