Skip to main content

Advertisement

Log in

Modelling the Effect of Incubation and Latent Periods on the Dynamics of Vector-Borne Plant Viral Diseases

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

Abstract

Most of the plant viral diseases spread through vectors. In case of the persistently transmitted disease, there is a latent time of infection inside the vector after acquisition of the virus from the infected plant. Again, the plant after getting infectious agent shows an incubation time after the interaction with an infected vector before it becomes diseased. The goal of this work is to study the effect of both incubation delay and latent time on the dynamics of plant disease, and accordingly a delayed model has been proposed. The existence of the equilibria, basic reproductive number (\(\mathcal {R}_0\)) and stability of equilibria have been studied. This study shows the relevance of the presence of two time delays, which may lead to system stabilization.

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.

Institutional subscriptions

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

References

  • Al Basir F, Blyuss KB, Ray S (2018) Modelling the effects of awareness-based interventions to control the mosaic disease of Jatropha curcas. Ecol Complex 36:92–100

    Google Scholar 

  • Anguelov R, Lubuma J, Dumont Y (2012) Mathematical analysis of vector-borne diseases on plants. In: 2012 IEEE 4th international symposium on plant growth modeling, simulation, visualization and applications, pp 22–29

  • Bailey NTJ (1975) The mathematical theory of infectious diseases and its applications, 2nd edn. Charles Griffin & Company Limited, London

    MATH  Google Scholar 

  • Banerjee M, Takeuchi Y (2017) Maturation delay for the predators can enhance stable coexistence for a class of prey-predator models. J Theor Biol 412:154–71

    MathSciNet  MATH  Google Scholar 

  • Bruhn JA, Fry WE (1981) Analysis of Potato Late Blight epidemiology by simulation modelling. Psychopathology 71(6):612–616

    Google Scholar 

  • Buonomo B, Cerasuolo M (2014) Stability and bifurcation in plant-pathogens interactions. Appl Math Comput 232:858–871

    MathSciNet  MATH  Google Scholar 

  • Cao J, Xiao M (2007) Stability and Hopf bifurcation in a simplified BAM neural network with two time delays. IEEE Trans Neural Netw 18(2):416–430

    MathSciNet  Google Scholar 

  • Chan M, Jeger MJ (1994) Analytical model of plant virus disease dynamics with roguing and replanting. J Appl Ecol 31:413–427

    Google Scholar 

  • Cook DWM, Long PJ, Ganesh S (1999) The combined effect of delayed application of yeast biocontrol agents and fruit curing for the inhibition of the postharvest pathogen Botrytis cinerea in kiwifruit. Postharvest Biol Technol 16:233–243

    Google Scholar 

  • Cooke KL (1979) Stability analysis for a vector disease model. Rocky Mountain J Math 9(1):31–42

    MathSciNet  MATH  Google Scholar 

  • Dietzgen RG, Mann KS, Johnson KN (2016) Plant virus-insect vector interactions: current and potential future research directions. Viruses 8(11):303

    Google Scholar 

  • Duffus JE (1987) Whitefly transmission of plant viruses. In: Harris KF (ed) Current topics in vector research. Springer, New York, pp 73–91

    Google Scholar 

  • Haber S, Maxwell DP, Gilbertson RL (1991) Bean golden mosaic. In: Hall R (ed) Compendium of bean diseases. American Phytopathological Society, Eagan, pp 42–43

    Google Scholar 

  • Hassard BD, Hassard DB, Kazarinoff ND, Wan YH, Wan YW (1981) Theory and applications of Hopf bifurcation, vol 41. CUP Archive, Cambridge

    MATH  Google Scholar 

  • Heffernan JM, Smith RJ, Wahl LM (2005) Perspectives on the basic reproductive ratio. J R Soc Interface 2(4):281–93

    Google Scholar 

  • Henneberry TJ, Castle SJ (2001) Bemisia pest status, economics, biology and population dynamics. In: Harris KF, Smith OP, Duffus JE (eds) Virus insects plant interactions. Academic Press, San Diego, pp 247–278

    Google Scholar 

  • Hohn T (2007) Plant virus transmission from the insect point of view. Proc Natl Acad Sci 104(46):17905–17906

    Google Scholar 

  • Holt J, Jeger MJ, Thresh JM, Otim-Nape GW (1997) An epidemilogical model incorporating vector population dynamics applied to African cassava mosaic virus disease. J Appl Ecol 34(3):793–806

    Google Scholar 

  • Horowitz AR (1986) Population dynamics of Bemisia tabaci (Gennadius): with special emphasis on cotton fields. Agric Ecosyst Environ 17:37–47

    Google Scholar 

  • Jackson M, Chen B (2016) Modeling plant virus propagation with delays. J Comput Appl Math 309:611–621

    MathSciNet  MATH  Google Scholar 

  • Jiao J, Chen L (2008) Global attractivity of a stage-structure variable coefficients predator-prey system with time delay and impulsive perturbations on predators. Int J Biomath 1(2):197–208

    MathSciNet  MATH  Google Scholar 

  • Kern H (1956) Problems of incubation in plant diseases. Ann Rev Microbiol 10(1):351–368

    Google Scholar 

  • Kuznetsov YA (2013) Elements of applied bifurcation theory, vol 112. Springer, Berlin

    Google Scholar 

  • Legg JP (2008) African cassava mosaic disease. In: Mahy BWJ, van Regenmortel MHV (eds) Encyclopedia of virology, 3rd edn. Elsevier, Amsterdam, pp 30–36

    Google Scholar 

  • Li Q, Dai Y, Guo X, Zhang X (2018) Hopf bifurcation analysis for a model of plant virus propagation with two delays. Adv Differ Equ 2018:259

    MathSciNet  MATH  Google Scholar 

  • Liu D, Raghothama KG, Hasegawa PM, Bressan RA (2012) Osmotin overexpression in potato delays development of disease symptoms. Proc Natl Acad Sci U S A 91:1888–1892

    Google Scholar 

  • Madden LV, Hughes G, Van Den Bosch F (2007) The study of plant disease epidemics. APS Press, St Paul

    Google Scholar 

  • Meng X, Li Z (2010) The dynamics of plant disease models with continuous and impulsive cultural control strategies. J Theor Biol 266(1):29–40

    MathSciNet  MATH  Google Scholar 

  • Moore SM, Borer ET, Hosseini PR (2009) Predators indirectly control vector-borne disease: linking predator-prey and host-pathogen models. J R Soc Interface 7(42):161–176

    Google Scholar 

  • Murphy AM, Holcombe LJ, Carr JP (2000) Characteristics of salicylic acid-induced delay in disease caused by a necrotrophic fungal pathogen in tobacco. Physiol Mol Plant Pathol 57:47–54

    Google Scholar 

  • Niehaus K, Kapp D, Pühler A (1993) Plant defence and delayed infection of alfalfa pseudonodules induced by an exopolysaccharide (EPS I)-deficient Rhizobium meliloti mutant. Planta 190:415–425

    Google Scholar 

  • Okamoto KW, Amarasekare P (2012) The biological control of disease vectors. J Theor Biol 309:47–57

    MathSciNet  MATH  Google Scholar 

  • Pal PJ, Saha T, Sen M, Banerjee M (2012) A delayed predatorprey model with strong Allee effect in prey population growth. Nonlinear Dyn 68:23–42

    MATH  Google Scholar 

  • Perko L (1996) Differential equations and dynamical systems. Springer, Berlin

    MATH  Google Scholar 

  • Rakshit N, Basir FA, Banerjee A, Ray S (2019) Dynamics of plant mosaic disease propagation and the usefulness of roguing as an alternative biological control. Ecol Complex 38:15–23

    Google Scholar 

  • Rimbaud L, Dallot S, Delaunay A, Borron S, Soubeyrand S, Thébaud G, Jacquot E (2015) Assessing the mismatch between incubation and latent periods for vector-borne diseases: the case of Sharka. Phytopathology 105(11):1408–1416

    Google Scholar 

  • Ruan S (2001) Absolute stability, conditional stability and bifurcation in Kolmogorov-type predator-prey systems with discrete delays. Q Appl Math 59(1):159–73

    MathSciNet  MATH  Google Scholar 

  • Shi R, Zhao H, Tang S (2014) Global dynamic analysis of a vector-borne plant disease model. Adv Differ Equ 2014:59

    MathSciNet  MATH  Google Scholar 

  • Shun-xiang RE, Zhen-zhong WA, Bao-li QI, Yuan XI (2001) The pest status of Bemisia tabaci in China and non-chemical control strategies. Insect Sci 8(3):279–288

    Google Scholar 

  • Siegel A, Ginoza W, Wildman SG (1957) The early events of infection with tobacco mosaic virus nucleic acid. Virology 3:554–559

    Google Scholar 

  • Smrž J (1972) The effect of the transmission of the beet mosaic virus on the variability of its incubation period. Biol Plant 14:1–10

    Google Scholar 

  • Suffert F, Thompson RN (2018) Some empirical arguments demonstrating that the latent period varies over the course of a plant disease epidemic. bioRxiv 148619

  • Van der Plank JE (1963) Plant diseases: epidemics and control. Academic Press, London

    Google Scholar 

  • Venturino E, Roy PK, Basir FA, Datta A (2016) A model for the control of the mosaic virus disease in Jatropha curcas plantations. Energy Ecol Environ 1:360–369

    Google Scholar 

  • Villa F (1992) New computer architectures as tools for ecological thought. Trends Ecol Evol 7(6):179–183

    Google Scholar 

  • Yoshimura A, Kawasaki K, Takasu F, Togashi K, Futai K, Shigesada N (1999) Modeling the spread of pine wilt disease caused by nematodes with pine sawyers as vector. Ecology 80(5):1691–1702

    Google Scholar 

  • Zhang T, Meng X, Song Y, Li Z (2012) Dynamical analysis of delayed plant disease models with continuous or impulsive cultural control strategies. Abstr Appl Anal. https://doi.org/10.1155/2012/428453

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

The authors wishes to thank the anonymous reviewer for his/her careful reading of the manuscript. Fahad Al Basir acknowledges the University Grants Commission, Govt. of India, for Dr. D S Kothari Postdoctoral fellowship, F No: MA/16-17/0029. Sagar Adhurya is thankful to the University Grants Commission, New Delhi, for funding his research by NET-JRF fellowship.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Santanu Ray.

Additional information

Publisher's Note

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

Ezio Venturino: Member of the INdAM research group GNCS.

Appendices

Appendix A

We follow the method established in the paper by Heffernan et al. (2005) for calculating \(\mathcal {R}_0\).

We consider the next-generation matrix G which comprises two parts, namely F and V, where

$$\begin{aligned} {F}= & {} \left[ \frac{\partial {F}_i(E_0)}{\partial x_j}\right] =\left[ \begin{array}{ll} 0 &{} ~~ \lambda K \\ \frac{\beta \Pi }{d} &{}~~0 \\ \end{array} \right] \\ {V}= & {} \left[ \frac{\partial {V}_i(E_0)}{\partial x_j}\right] =\left[ \begin{array}{ll} m~ &{}~~ 0\\ \\ 0 ~&{} ~~ d \\ \end{array}\right] \end{aligned}$$

where \({F}_i\) are the new infections, while the \({V}_i\) transfers of infections from one compartment to another. \(E_0\) is the disease-free equilibrium. We get \(\mathcal {R}_0=\frac{\Pi k\lambda \beta }{md^2}\), and it is the dominant eigenvalue of the matrix \({G} = {F}{V}^{-1}\).

Appendix B

We rewrite the model (1), without delay, as follows:

$$\begin{aligned} \frac{\mathrm{d}S}{\mathrm{d}t}= & {} rS\left( 1-\frac{S+I}{K}\right) -\lambda SV\,\equiv \,f_1(S,I,U,V),\end{aligned}$$
(22a)
$$\begin{aligned} \frac{\mathrm{d}I}{\mathrm{d}t}= & {} \lambda SV-mI\,\equiv \,f_2(S,I,U,V),\end{aligned}$$
(22b)
$$\begin{aligned} \frac{\mathrm{d}U}{\mathrm{d}t}= & {} \Pi -\beta UI-dU\,\equiv \,f_3(S,I,U,V),\end{aligned}$$
(22c)
$$\begin{aligned} \frac{\mathrm{d}V}{\mathrm{d}t}= & {} \beta UI-dV\,\equiv \,f_4(S,I,U,V), \end{aligned}$$
(22d)

and consider \(\lambda \) as bifurcation parameter to check the change in stability of the disease-free equilibrium point \(E_2(K,0,\frac{\Pi }{d},0)\). Jacobian matrix for (22) evaluated at \(E_2\) is given by

$$\begin{aligned} J\left( E_2\right) \,=\,\left[ \begin{array}{llll} -r &{} -r &{} 0 &{} -\lambda K\\ 0 &{} -m &{} 0 &{} \lambda K\\ 0 &{} -\frac{\beta \Pi }{d} &{} -d &{} 0\\ 0 &{} \frac{\beta \Pi }{d} &{} 0 &{} -d\\ \end{array}\right] . \end{aligned}$$
(23)

This matrix has a zero eigenvalue when \(\lambda \,\equiv \,\lambda _*\,=\,\frac{d^2m}{K\Pi \beta }\). The system undergoes a transcritical bifurcation as \(\lambda \) crosses \(\lambda _*\). To prove the transversality condition of transcritical bifurcation, we use standard notations in Perko (1996). The system (22) can be written in compact form as

$$\begin{aligned} \frac{\mathrm{d}{X}}{\mathrm{d}t}=\mathbf{f}({X}), \end{aligned}$$
(24)

and the Jacobian matrix in (23) is denoted as \(A\,=\,J\left( E_2\right) \). For \(\lambda \,=\,\lambda _*\), the matrix A has a zero eigenvalue, eigenvectors corresponding to zero eigenvalue for the matrices A and \(A^t\) are given by

$$\begin{aligned} W_1= & {} \left[ \begin{array}{l} -\frac{r+m}{r} \\ 1 \\ -\frac{\Pi \beta }{d^2} \\ \frac{\Pi \beta }{d^2} \\ \end{array}\right] ,\nonumber \\ W_2= & {} \left[ \begin{array}{l} 0 \\ \frac{\Pi \beta }{md} \\ 0 \\ 1 \\ \end{array}\right] . \end{aligned}$$
(25)

Partial differentiation of \(\mathbf{f}\) with respect to \(\lambda \) gives

$$\begin{aligned} {\mathbf {f}}_\lambda \,=\,\left[ \begin{array}{l} -SV \\ SV \\ 0 \\ 0 \\ \end{array}\right] , \end{aligned}$$
(26)

and the Jacobian of \(\mathbf{f}_\lambda \) can be calculated as follows:

$$\begin{aligned} D\mathbf{f}_\lambda \,=\, \left[ \begin{array}{llll} -V &{}\quad 0 &{}\quad 0 &{}\quad -S \\ V &{}\quad 0 &{}\quad 0 &{}\quad S \\ 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 &{}\quad 0 \\ \end{array}\right] . \end{aligned}$$
(27)

Hence, we can verify the transversality conditions for transcritical bifurcation as follows:

$$\begin{aligned} W_2^t\left[ \mathbf{f}_\lambda (E_2,\lambda _*)\right]= & {} 0, \end{aligned}$$
(28a)
$$\begin{aligned} W_2^t\left[ D\mathbf{f}_\lambda (E_2,\lambda _*)\,W_1\right]= & {} \frac{\Pi ^2\beta ^2K}{md^3}\,\,\ne \,\,0, \end{aligned}$$
(28b)
$$\begin{aligned} W_2^t\left[ D^2\mathbf{f}(E_2,\lambda _*)\,(W_1,W_1)\right]= & {} -\frac{2\beta ^2\Pi }{d^2}\left( 1+\frac{\Pi d(r+m)}{rK\beta }\right) \,\,\ne \,\,0. \end{aligned}$$
(28c)

These ensure that the boundary equilibrium point \(E_2\) loses stability through transcritical bifurcation.

Appendix C

For the values in Table 1, we get the four roots of Eq. (15) as: \(- \,0.0664\), 0.0360 and \(- \,0.0108 \pm 0.0430i\). We get one pair of complex root and the positive root as 0.0360. Here, \(\omega _4 = - 4.6897\times 10^{-6}<0\). According to Lemma 1, Eq. (15) will have at least one positive root.

Equation (16) has three roots: \(- 0.0374\) and \( - 0.0008 \pm 0.0073i\). We have got \(\delta = 2.6996 \times 10^{-11}>0,\)\(H(l)<0\). According to Lemma 2, Eq. (15) has one positive root. Then, Eq. (11) has a pair of purely imaginary root. Finally, we have obtained \(\theta _0= 0.1897\) from (19), as well as the critical value of \(\tau _1\), namely \(\tau _1^*=15.38\), from relation (18).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Al Basir, F., Adhurya, S., Banerjee, M. et al. Modelling the Effect of Incubation and Latent Periods on the Dynamics of Vector-Borne Plant Viral Diseases. Bull Math Biol 82, 94 (2020). https://doi.org/10.1007/s11538-020-00767-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11538-020-00767-2

Keywords

Navigation