Skip to main content

Advertisement

Log in

Backward bifurcations, turning points and rich dynamics in simple disease models

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

Abstract

In this paper, dynamical systems theory and bifurcation theory are applied to investigate the rich dynamical behaviours observed in three simple disease models. The 2- and 3-dimensional models we investigate have arisen in previous investigations of epidemiology, in-host disease, and autoimmunity. These closely related models display interesting dynamical behaviors including bistability, recurrence, and regular oscillations, each of which has possible clinical or public health implications. In this contribution we elucidate the key role of backward bifurcations in the parameter regimes leading to the behaviors of interest. We demonstrate that backward bifurcations with varied positions of turning points facilitate the appearance of Hopf bifurcations, and the varied dynamical behaviors are then determined by the properties of the Hopf bifurcation(s), including their location and direction. A Maple program developed earlier is implemented to determine the stability of limit cycles bifurcating from the Hopf bifurcation. Numerical simulations are presented to illustrate phenomena of interest such as bistability, recurrence and oscillation. We also discuss the physical motivations for the models and the clinical implications of the resulting dynamics.

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

Similar content being viewed by others

References

  • Alexander HK, Wahl LM (2011) Self-tolerance and autoimmunity in a regulatory T cell model. Bull Math Biol 73(1):33–71

    Article  MathSciNet  MATH  Google Scholar 

  • Arino J, McCluskey CC, van den Driessche P (2003) Global results for an epidemic model with vaccination that exhibits backward bifurcation. SIAM J Appl Math 64(1):260–276

    Article  MathSciNet  MATH  Google Scholar 

  • Baecher-Allan C, Wolf E, Hafler DA (2006) MHC class II expression identifies functionally distinct human regulatory T cells. J Immunol 176:4622–4631

    Article  Google Scholar 

  • Blayneh KW, Gumel AB, Lenhart S, Clayton T (2010) Backward bifurcation and optimal control in transmission dynamics of west nile virus. Bull Math Biol 72(4):1006–1028

    Article  MathSciNet  MATH  Google Scholar 

  • Brauer F (2004) Backward bifurcations in simple vaccination models. J Math Anal Appl 298(2):418–431

    Article  MathSciNet  MATH  Google Scholar 

  • Briggs CJ, Godfray HCJ (1995) The dynamics of insect-pathogen interactions in stage-structured populations. Am Nat 145(6):855–887

    Article  Google Scholar 

  • Brown GC, Hasibuan R (1995) Conidial discharge and transmission efficiency of Neozygites floridana, an entomopathogenic fungus infecting two-spotted spider mites under laboratory conditions. J Invertebr Pathol 65(1):10–16

    Article  Google Scholar 

  • Capasso V, Serio G (1978) A generalization of the Kermack–McKendrick deterministic epidemic model. Math Biosci 42(1):43–61

    Article  MathSciNet  MATH  Google Scholar 

  • Castillo-Chavez C, Cooke K, Huang W, Levin S (1989a) On the role of long incubation periods in the dynamics of acquired immunodeficiency syndrome (AIDS). Part 2: Multiple group models. In: Castillo-Chavez C (ed) Mathematical and statistical approaches to AIDS epidemiology, lecture notes in biomathematics, vol 83. Springer, Berlin, pp 200–217

    Chapter  Google Scholar 

  • Castillo-Chavez C, Cooke K, Huang W, Levin SA (1989b) Results on the dynamics for models for the sexual transmission of the human immunodeficiency virus. Appl Math Lett 2(4):327–331

    Article  MathSciNet  MATH  Google Scholar 

  • Collier AC, Coombs RW, Schoenfeld DA, Bassett RL, Timpone J, Baruch A, Jones M, Facey K, Whitacre C, McAuliffe VJ, Friedman HM, Merigan TC, Reichman RC, Hooper C, Corey L (1996) Treatment of human immunodeficiency virus infection with Saquinavir, Zidovudine, and Zalcitabine. N Engl J Med 334(16):1011–1018

    Article  Google Scholar 

  • Conway JM, Coombs D (2011) A stochastic model of latently infected cell reactivation and viral blip generation in treated HIV patients. PLoS Comput Biol 7(4):e1002033

    Article  MathSciNet  Google Scholar 

  • Derrick WR, van den Driessche P (1993) A disease transmission model in a nonconstant population. J Math Biol 31(5):495–512

    Article  MathSciNet  MATH  Google Scholar 

  • Dornadula G, Zhang H, VanUitert B, Stern J, Livornese L Jr, Ingerman MJ, Witek J, Kedanis RJ, Natkin J, DeSimone J, Pomerantz RJ (1999) Residual HIV-1 RNA in blood plasma of patients taking suppressive highly active antiretroviral therapy. JAMA 282(17):1627–1632

    Article  Google Scholar 

  • Dushoff J, Huang W, Castillo-Chavez C (1998) Backwards bifurcations and catastrophe in simple models of fatal diseases. J Math Biol 36(3):227–248

    Article  MathSciNet  MATH  Google Scholar 

  • Fraser C, Ferguson NM, de Wolf F, Anderson RM (2001) The role of antigenic stimulation and cytotoxic T cell activity in regulating the long-term immunopathogenesis of HIV: mechanisms and clinical implications. Proc R Soc B Biol Sci 268(1481):2085–2095

    Article  Google Scholar 

  • Fung ICH, Gambhir M, van Sighem A, de Wolf F, Garnett GP (2012) The clinical interpretation of viral blips in HIV patients receiving antiviral treatment: are we ready to infer poor adherence? J Acquir Immune Defic Syndr 60(1):5–11

    Article  Google Scholar 

  • Fussmann GF, Ellner SP, Shertzer KW, Hairston NG Jr (2000) Crossing the Hopf bifurcation in a live predator–prey system. Science 290(5495):1358–1360

    Article  Google Scholar 

  • Garretta NJ, Apeaa V, Noria A, Ushiro-Lumbb I, Oliverb AR, Bailya G, Clarkb DA (2012) Comparison of the rate and size of HIV-1 viral load blips with Roche COBAS TaqMan HIV-1 versions 1.0 and 2.0 and implications for patient management. J Clin Virol 53(4):354–355

    Article  Google Scholar 

  • Gil L et al (2003) Contribution to characterization of oxidative stress in HIV/AIDS patients. Pharmacol Res 47(3):217–224

    Article  Google Scholar 

  • Grennan JT, Loutfy MR, Su D, Harrigan PR, Cooper C, Klein M, Machouf N, Montaner JSG, Rourke S, Tsoukas C, Hogg B, Raboud J, the CANOC Collaboration (2012) Magnitude of virologic blips is associated with a higher risk for virologic rebound in HIV-infected individuals: a recurrent events analysis. J Infect Dis 205(8):1230-1238

  • Gomez-Acevedo H, Li MY (2005) Backward bifurcation in a model for HTLV-I infection of \({\rm CD}4^+\) T cells. Bull Math Biol 67(1):101-114

  • Hadeler KP, van den Driessche P (1997) Backward bifurcation in epidemic control. Math Biosci 146(1):15–35

    Article  MathSciNet  MATH  Google Scholar 

  • Hethcote HW, van den Driessche P (1991) Some epidemiological models with nonlinear incidence. J Math Biol 29(3):271–287

    Article  MathSciNet  MATH  Google Scholar 

  • Hethcote HW, Lewis MA, van den Driessche P (1989) An epidemiological model with a delay and a nonlinear incidence rate. J Math Biol 27(1):49–64

    Article  MathSciNet  MATH  Google Scholar 

  • Huang W, Cooke KL, Castillo-Chavez C (1992) Stability and bifurcation for a multiple-group model for the dynamics of HIV/AIDS transmission. SIAM J Appl Math 52(3):835–854

    Article  MathSciNet  MATH  Google Scholar 

  • Israel N, Gougerot-Pocidalo MA (1997) Oxidative stress in human immunodeficiency virus infection. Cell Mol Life Sci 53(11–12):864–870

    Article  Google Scholar 

  • Jones LE, Perelson AS (2005) Opportunistic infection as a cause of transient viremia in chronically infected HIV patients under treatment with HAART. Bull Math Biol 67(6):1227–1251

    Article  MathSciNet  MATH  Google Scholar 

  • Korobeinikov A, Maini PK (2005) Non-linear incidence and stability of infectious disease models. Math Med Biol 22(2):113–128

    Article  MATH  Google Scholar 

  • Li MY, Muldowney JS (1995) Global stability for the SEIR model in epidemiology. Math Biosci 125(2):155–164

    Article  MathSciNet  MATH  Google Scholar 

  • Li N, Karin M (1999) Is NF-\(\kappa \)B the sensor of oxidative stress? FASEB J 13(10):1137–1143

    Google Scholar 

  • Liu W, Levin S, Iwasa Y (1986) Influence of nonlinear incidence rates upon the behavior of SIRS epidemiological models. J Math Biol 23(2):187–204

    Article  MathSciNet  MATH  Google Scholar 

  • Liu W, Hethcote HW, Levin SA (1987) Dynamical behavior of epidemiological models with nonlinear incidence rates. J Math Biol 25(4):359–380

    Article  MathSciNet  MATH  Google Scholar 

  • Murray JD (2002) Mathematical biology: I. An introduction. Interdisciplinary applied mathematics. Springer, Berlin

    Google Scholar 

  • Ngonghala CN, Miranda IT, Ngwa GA (2014) Persistent oscillations and backward bifurcation in a malaria model with varying human and mosquito populations: implications for control. J Math Biol 70(7):1581–1622

    Article  MathSciNet  MATH  Google Scholar 

  • Palmer S, Wiegand AP, Maldarelli F, Bazmi H, Mican JM, Polis M, Dewar RL, Planta A, Liu S, Metcalf JA, Mellors JW, Coffin JM (2003) New real-time reverse transcriptase-initiated PCR assay with single-copy sensitivity for human immunodeficiency virus type 1 RNA in plasma. J Clin Microbiol 41(10):4531–4536

    Article  Google Scholar 

  • Palmer S, Maldarelli F, Wiegand A, Bernstein B, Hanna GJ, Brun SC, Kempf DJ, Mellors JW, Coffin JM, King MS (2008) Low-level viremia persists for at least 7 years in patients on suppressive antiretroviral therapy. Proc Natl Acad Sci 105(10):3879–3884

    Article  Google Scholar 

  • Rong L, Perelson AS (2009a) Asymmetric division of activated latently infected cells may explain the decay kinetics of the HIV-1 latent reservoir and intermittent viral blips. Math Biosci 217(1):77–87

    Article  MathSciNet  MATH  Google Scholar 

  • Rong L, Perelson AS (2009b) Modeling latently infected cell activation: viral and latent reservoir persistence, and viral blips in HIV-infected patients on potent therapy. PLoS Comput Biol 5(10):e1000533

    Article  MathSciNet  Google Scholar 

  • Ruan S, Wang W (2003) Dynamical behavior of an epidemic model with a nonlinear incidence rate. J Differ Equ 188(1):135–163

    Article  MathSciNet  MATH  Google Scholar 

  • Schwarz KB (1996) Oxidative stress during viral infection: a review. Free Radic Biol Med 21(5):641–649

    Article  Google Scholar 

  • Simon CP, Jacquez JA (1992) Reproduction numbers and the stability of equilibria of SI models for heterogeneous populations. SIAM J Appl Math 52(2):541–576

    Article  MathSciNet  MATH  Google Scholar 

  • Stephenson CB, Marquis GS, Douglas SD, Wilson CM (2005) Immune activation and oxidative damage in HIV-positive and HIV negative adolescents. J Acquir Immune Defic Syndr 38(2):180–190

    Article  Google Scholar 

  • van den Driessche P, Watmough J (2002) Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission. Math Biosci 180(12):29–48

    Article  MathSciNet  MATH  Google Scholar 

  • van Gaalen RD, Wahl LM (2009) Reconciling conflicting clinical studies of antioxidant supplementation as HIV therapy: a mathematical approach. BMC Public Health 9(Suppl. 1):1–18

    Google Scholar 

  • Wiggins S (2003) Introduction to applied nonlinear dynamical systems and chaos, vol 2. Springer Science and Business Media, Berlin

    MATH  Google Scholar 

  • Yao W, Hertel L, Wahl LM (2006) Dynamics of recurrent viral infection. Proc R Soc B Biol Sci 273(1598):2193–2199

    Article  Google Scholar 

  • Yu P (1998) Computation of normal forms via a perturbation technique. J Sound Vib 211(1):19–38

    Article  MathSciNet  MATH  Google Scholar 

  • Yu P, Zhang W, Wahl LM (2016) Dynamical analysis and simulation of a 2-dimensional disease model with convex incidence. Commun Nonlinear Sci Numer Simul 37:163–192

    Article  MathSciNet  Google Scholar 

  • Zhang W, Wahl LM, Yu P (2013) Conditions for transient viremia in deterministic in-host models: viral blips need no exogenous trigger. SIAM J Appl Math 73(2):853–881

    Article  MathSciNet  MATH  Google Scholar 

  • Zhang W, Wahl LM, Yu P (2014) Viral blips may not need a trigger: How transient viremia can arise in deterministic in-host models. SIAM Rev 56(1):127–155

    Article  MATH  Google Scholar 

  • Zhang W, Wahl LM, Yu P (2014) Modelling and analysis of recurrent autoimmune disease. SIAM J Appl Math 74(6):1998–2025

    Article  MathSciNet  MATH  Google Scholar 

  • Zhou L, Fan M (2012) Dynamics of an SIR epidemic model with limited medical resources revisited. Nonlinear Anal Real 13(1):312–324

    Article  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Wenjing Zhang.

Additional information

This work is supported by the Natural Sciences and Engineering Research Council of Canada (NSERC) (No. 238389-RGPIN and No. R2686A02).

The first draft of this paper has been posted on arXiv.org since April 20, 2015 (No. arxiv:1504.05260).

Appendices

Appendix A

The equilibrium solutions of system (7) are obtained by solving the following algebraic equations: \(f_1(S,\,I)=0\) and \(f_2(S,\,I)=0\), from which the disease-free equilibrium can be easily obtained as \(\bar{\mathrm {E}}_0 = (\varLambda /d,\,0)\). For the infected equilibrium \(\bar{\mathrm {E}}=(\bar{S},\,\bar{I})\), \(\bar{S}\) is solved from \(f_1=0\) as \(\textstyle \bar{S}(I)=\frac{\varLambda (1+kI)}{(d k+\beta ) I+d}\). Then, substituting \(S=\bar{S}(I)\) into \(f_2=0\) yields a quadratic equation of the form

$$\begin{aligned} {\mathscr {F}}(I) = {\mathscr {A}}I^2 + {\mathscr {B}} I + {\mathscr {C}} = 0, \end{aligned}$$
(21)

which in turn gives two roots: \(\bar{I}_{1,\,2} = \frac{-{\mathscr {B}}\pm \sqrt{{\mathscr {B}}^2-4{\mathscr {A}}\,{\mathscr {C}}}}{2{\mathscr {A}}}\), where, \({\mathscr {A}} = (d+\gamma +\epsilon )(dk+\beta )\), \({\mathscr {B}} = [(dk +\beta )\omega +d](d+\gamma +\epsilon ) +(d k+\beta )\alpha -\beta \varLambda \), \({\mathscr {C}} = [(d+\gamma +\epsilon )\omega +\alpha ]d - \beta \varLambda \omega \) for system (7). Since all parameters take positive values, we have \({\mathscr {A}}>0\). To get the two positive roots essential for backward bifurcation, it is required that \({\mathscr {B}}<0\) and \({\mathscr {C}}>0\). Noticing that \(\beta , \, \varLambda ,\, \omega >0\), we can see that the infection force, \(\beta \), the constant influx of the susceptibles, \(\varLambda \), and the effect of medical treatment \(\frac{\alpha I}{\omega + I}\) are indispensable terms for backward bifurcation. The number of positive infected equilibrium solutions changes from two to one when the value of \({\mathscr {C}}\) passes from negative to positive, which gives a critical point at \({\mathscr {C}}=0\), that is, \([(d+\gamma +\epsilon )\omega +\alpha ]d = \beta \varLambda \omega \), which is equivalent to \(R_0^{(7)} = 1\).

On the other hand, we may infer the emergence of backward bifurcation without solving the equilibrium conditions. When we introduce the loss of the infectives due to medical treatment, the dynamics of system (7) differ greatly from system (6). In particular, backward bifurcation emerges and complex dynamical behaviors may occur. To clarify this effect, we obtain the function \(f_4(S,\,I)\) from the equation \(\frac{dI}{dt}=0\) of (7). Note that \(f_4(S,\,I)\) is not an incidence rate. But, if we fix \(S=\tilde{S}>0\), there exist \(0<I_1<I_2<+\infty \), such that \(\frac{\partial f_4(\tilde{S},\,I)}{\partial I} =\frac{1}{(1+kI)^2(\omega +I)^2} [\beta \tilde{S}(\omega +I)^2-\alpha \omega (1+kI)^2]>0\), \(\forall \; I \in (0,\,I_2)\); and \(\frac{\partial ^2 f_4(\tilde{S},\,I)}{\partial I^2} = -2k\beta \tilde{S} (1+kI)^{-3} + 2\alpha \omega (\omega +I)^{-3}>0\), \(\forall \; I \in (0,\,I_1)\), \(\frac{\partial ^2 f_4(\tilde{S},\,I)}{\partial I^2}=0\), for \(I=I_1\), \(\frac{\partial ^2 f_4(\tilde{S},\,I)}{\partial I^2}<0\), \(\forall \; I \in (I_1,\,I_2)\). Thus, \(f_4(\tilde{S},\,I)\) actually has a convex-concave ‘S’ shape, and may have two positive intersection points with the ray line, \(g_1(I)\), in the first quadrant.

The infected equilibrium of (7) is denoted as \(\bar{\mathrm {E}}_1 = (\bar{S},\,\bar{I})\), where \(\bar{I}\) is solved from the equation \({\mathscr {F}}(I) =0\) in (21). The turning point is determined by both the quadratic equation (21) and the relation \(\frac{\mathrm {d}\varLambda }{\mathrm {d}I}=-\frac{\partial {\mathscr {F}}}{\partial I}/ \frac{\partial {\mathscr {F}}}{\partial \varLambda }=0\), which is equivalent to \(\frac{\partial {\mathscr {F}}}{\partial I}=0\). Solving \(\frac{\partial {\mathscr {F}}}{\partial I}=0\) yields the expression of the turning point of I, given in (14). To find the stability of the infected equilibrium \(\bar{\mathrm {E}}_1\), evaluating the Jacobian matrix at \(\bar{\mathrm {E}}_1\), and further denoting it as \(J_2=J|_{(7)}(\bar{\mathrm {E}}_1)\), we obtain the characteristic polynomial in the form of (13), with \(\mathrm {Tr}(J_2)=a_{11}/[(\omega +I)^2(kI+1)(dkI+\beta I+d)]\) and \(\mathrm {Det}(J_2)=a_{21}/[(\omega +I)^2(kI+1)(dkI+\beta I+d)]\), where \(a_{11}=a_{1a}-a_{1b}\) and \(a_{21}=a_{2a}-a_{2b}\), with \(a_{1b}= \beta \varLambda (\omega +I)^2\) and \(a_{2b}=d a_{1b}\), and \(a_{1a}\) and \(a_{2a}\) only contain positive terms (their expressions are omitted here for brevity). Determining whether a Hopf bifurcation can occur from \(\bar{\mathrm {E}}\) is equivalent to finding whether \(\mathrm {Det}(J_2)\) remains positive when \(\mathrm {Tr}(J_2)=0\). Consider the following expression:

$$\begin{aligned} h_1(I)= & {} \Big [\mathrm {Tr}(J_2)- \frac{1}{d} \mathrm {Det}(J_2) \Big ] (\omega +I)^2(kI+1)(dkI+\beta I+d) \nonumber \\= & {} a_{11}-\frac{1}{d}a_{21}=a_{1a}-\frac{1}{d}a_{2a} \nonumber \\= & {} \frac{1}{d}(dkI+\beta I+d)[(kI+1)d^2(\omega +I)^2- \beta \epsilon I (\omega +I)^2-\alpha \beta \omega I],\nonumber \\ \end{aligned}$$
(22)

where the expressions of \(a_{1a}\) and \(a_{2a}\) have been used. This indicates that \(h_1(I)\) may take negative values, for which \(\mathrm {Det}(J_2)>0\).

Appendix B

It is easy to find the uninfected equilibrium of model (4), \(\bar{\mathrm {E}}_0=(\bar{X}_0,\,\bar{Y}_0)=(\frac{1}{D},\,0)\), whose characteristic polynomial has two roots: \(\lambda _1=-D<0\), and \(\lambda _2 = \frac{B}{D}-1\), which gives \(R_0=\frac{B}{D}\). Consequently, \(\bar{\mathrm {E}}_0\) is stable (unstable) for \(R_0<1\,({>}1)\). To find the infected equilibrium solution, setting \(f_6(X,\,Y) =0\) yields \(\bar{X}_1(Y)=\frac{Y+C}{(A+B)Y+BC}\), which is then substituted into \(f_5(X,\,Y)=0\) to give the following quadratic equation:

$$\begin{aligned} {\mathscr {F}}_5(Y)=(A+B)Y^2+(BC+D-A-B)Y+C(D-B)=0. \end{aligned}$$
(23)

In order to have two real, positive roots, two conditions must be satisfied, that is, \(BC+D-A-B<0\) and \(D-B>0\), or in compact form, \(0<D-B<A-BC\). The condition \(D-B>0\) is equivalent to \(0<R_0=\frac{B}{D}<1\), which is a necessary condition for backward bifurcation. Moreover, the positive influx constant, having been scaled to 1, is a necessary term for the positive equilibrium of Y. Therefore, the positive influx rate term and the increasing and saturating infectivity function are necessary for backward bifurcation.

We further examine the incidence function, \(f_7(X,\,Y)\) defined in (12), without solving for the equilibrium solutions. The incidence function \(f_7\) obviously satisfies the condition (3a), as well as the condition (3b) since \(\frac{\partial }{\partial X}f_7(X,\,Y)=[B+AY(Y+C)^{-1}]Y >0\) and \(\frac{\partial }{\partial Y}f_7(X,\,Y)= ACXY(Y+C)^{-2}+[B+AY(Y+C)^{-1}]X>0\) for all \(X,\,Y>0\). However, the second partial derivative of \(f_7(X,\,Y)\) with respect to Y, \(\frac{\partial ^2}{\partial Y^2}f_7(X,\,Y)= 2AC^2X(X+C)^{-3}>0\) for all \(X,\,Y>0\), showing that \(f_7(X,\,Y)\) is a convex function with respect to the variable Y. Consequently, \(f_7(X,\,Y)\) can only have one intersection with \(g_2(Y)=Y\), implying that only one equilibrium solution would exist if we only consider the second equation in (11), as shown Fig. 2a. However, when considering both conditions given in (11) for equilibrium solutions, we will have two intersection points between \(f_7\) and \(g_2\). According to the first equation in (11), that is \(f_5(X,\,Y)=0\), we can use Y to express X in the equilibrium state as \(\bar{X}(Y)=(Y+C)[(A+B)Y^2+(BC+D)Y+DC]^{-1}\). Substituting \(\bar{X}(Y)\) into \(f_7(X,\,Y)\) in (12), we obtain

$$\begin{aligned} f_7(Y)=Y[(A+B)Y+BC][(A+B)Y^2+(BC+D)Y+CD]^{-1}, \end{aligned}$$
(24)

and \(\frac{\partial }{\partial Y}f_7(Y)= D[(A+B)Y^2+2(A+B)CY+BC^2][(A+B)Y^2+(BC+D)Y+CD]^{-2}>0\) for all \(X,\,Y>0\). However, the sign of \(\frac{\partial ^2}{\partial Y^2}f_7(Y)= -2D[(A+B)^2Y^3+3C(A+B)^2Y^2+3(A+B)BC^2Y+(B^2C-AD)C^2] [(A+B)Y^2+(BC+D)Y+CD]^{-3}\), could alter at the inflection point from positive to negative as Y increases. Therefore, with appropriate parameter values, \(f_7(Y)\) can have a convex-concave ‘S’ shape.

If we choose the parameter B as the bifurcation parameter, then \(R_0=\frac{B}{D}=1\) defines \(B_S=D\) where the ‘S’ in subscript stands for static bifurcation. Further, it can be proved that this is a transcritical bifurcation. Therefore, \(\bar{\mathrm {E}}_0\) is stable when \(B<D\) (or \(R_0<1\)), loses its stability and becomes unstable when B increases to pass through \(B_S=D\), that is \(B>D\) (or \(R_0>1\)), and no other bifurcations can happen.

Next, we examine the infected equilibrium \(\bar{\mathrm {E}}_1=(\bar{X},\,\bar{Y})\). Since \(\bar{X}(Y)=\frac{Y+C}{(A+B)Y+BC}\), \(\bar{Y}\) is determined by the quadratic equation (23), which gives the turning point \((B_T,\,Y_T)\), as given in (15). We perform a further bifurcation analysis on its corresponding characteristic polynomial (13), which takes the form

$$\begin{aligned} \displaystyle P|_{\bar{\mathrm {E}}_1}(\lambda ,Y) \!= & {} \! \lambda ^2\!+\!\frac{a_{1a}}{[(A\!+\!B)Y\!+\!BC](Y\!+\!C)} \lambda \!+\!\frac{a_{2a}}{[(A\!+\!B)Y\!+\!BC](Y\!+\!C)},\quad \text {where} \nonumber \\ a_{1a}= & {} (A+B)^2Y^3+(2BC+D)(A+B)Y^2+(B^2C^2+ACD+2BCD\nonumber \\&-AC)Y+BC^2D, \nonumber \\ a_{2a}= & {} (A+B)^2Y^3+2(A+B)BCY^2+(B^2C-AD)CY. \end{aligned}$$
(25)

Therefore, the sign of the subtraction between the trace and determinant is determined by \(h_2(Y)=a_{1a}-a_{2a}=D(A+B)Y^2+[2CD(A+B)-AC]Y+BC^2D\). Here the equilibrium solution of Y and other parameters satisfy the quadratic equation (23), which leads to an explicit expression, given by \(\bar{B}= -\frac{AY^2+(D-A)Y+CD}{Y^2+(C-1)Y-C}\). Substituting \(B=\bar{B}\) into \(h_2(Y)\), we obtain

$$\begin{aligned} \displaystyle h_2(Y)|_{B\!=\!\bar{B}}\!=\! a_{1a} \!-\! a_{2a} \!=\! \frac{[AC(D-1)-D^2]Y^2-[AC(D-1)+2CD^2]Y-C^2D^2}{Y-1}. \end{aligned}$$
(26)

Hopf bifurcation may occur when the trace is zero, while the determinant is still positive. This implies \(h_2(Y)<0\), which is possible with appropriately chosen parameter values. Hence, by solving \(a_{1a}=0\) in (25) together with the quadratic equation (23), we get two pairs of points denoted by \((B_{h1},\,Y_{h1})\) and \((B_{h2},\,Y_{h2})\), which are candidates for Hopf bifurcation. Then validating the above two points by substituting them back into the characteristic polynomial (25), respectively, we denote the Hopf bifurcation point as \((B_H,\,Y_H)\) if this validation confirms their existence. According to Yu et al. (2015), Hopf bifurcation can happen only from the upper branch of the infected equilibrium \(\bar{\mathrm {E}}_1\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Zhang, W., Wahl, L.M. & Yu, P. Backward bifurcations, turning points and rich dynamics in simple disease models. J. Math. Biol. 73, 947–976 (2016). https://doi.org/10.1007/s00285-016-0976-6

Download citation

  • Received:

  • Revised:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00285-016-0976-6

Keywords

Mathematics Subject Classification

Navigation