Skip to main content
Log in

Basin Transition and Alternative States: Role of Multi-species Herbivores-Induced Volatile in Plant–Insect Interactions

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

Abstract

A simple model on volatile organic compound (VOC)-mediated plant–insect interactions is proposed and examined here, when two different classes of herbivorous insects competing for a common resource (plant) in the presence of a specialist carnivorous enemy, which only predates one of the herbivore species. We, particularly, emphasize the impact of VOCs on plant’s growth fitness. The system experiences several local and global bifurcations with emergent alternative states for variations in recruitment factors and predation rate. Basin transitions and basin of attractions have provided detail descriptions on the selectivity of the alternative states, when only one of the herbivore species can survive depending on the choice of initial population densities of the interacting species and how it provides a steady growth in plant. Additionally, our results support the concept of competitive exclusion principle in an indirect interspecific competition between the two herbivore types for the common resource, plant.

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

Similar content being viewed by others

References

  • Bairagi N, Saha S, Chaudhuri S, Dana SK (2019) Zooplankton selectivity and nutritional value of phytoplankton influences a rich variety of dynamics in a plankton population model. Phys Rev E 99:012406

    Google Scholar 

  • Binyameen M, Ali Q, Roy A, Schlyter F (2021) Plant volatiles and their role in insect olfaction. In: Plant–pest interactions: from molecular mechanisms to chemical ecology. Springer, pp 127–156

  • Birkhoff G, Rota G (1989) Ordinary differential equations. Ginn, Boston

    MATH  Google Scholar 

  • Cardoza YJ, Alborn HT, Tumlinson JH (2002) In vivo volatile emissions from peanut plants induced by simultaneous fungal infection and insect damage. J Chem Ecol 28(1):161–174

    Google Scholar 

  • Chen L, Chen F, Chen L (2010) Qualitative analysis of a predator-prey model with holling type ii functional response incorporating a constant prey refuge. Nonlinear Anal Real World Appl 11(1):246–252

    MathSciNet  MATH  Google Scholar 

  • De Boer JG, Hordijk CA, Posthumus MA, Dicke M (2008) Prey and non-prey arthropods sharing a host plant: effects on induced volatile emission and predator attraction. J Chem Ecol 34(3):281

    Google Scholar 

  • Dent CL, Cumming GS, Carpenter SR (2002) Multiple states in river and lake ecosystems. Philos Trans R Soc Lond B Biol Sci 357(1421):635–645

    Google Scholar 

  • Dicke M (1999) Evolution of induced indirect defense of plants. In: The ecology and evolution of inducible defenses, pp 62–88

  • Dublin HT, Sinclair AR, McGlade J (1990) Elephants and fire as causes of multiple stable states in the serengeti-mara woodlands. J Anim Ecol 1147–1164

  • Ermentrout B (2002) Simulating, analyzing, and animating dynamical systems: a guide to XPPAUT for researchers and students, vol 14. Siam, Philadelphia

    MATH  Google Scholar 

  • Fergola P, Wang W (2011) On the influences of defensive volatiles of plants in tritrophic interactions. J Biol Syst 19(02):345–363

    MathSciNet  MATH  Google Scholar 

  • Hardin G (1960) The competitive exclusion principle. Science 131(3409):1292–1297

    Google Scholar 

  • Hastings A, Powell T (1991) Chaos in a three-species food chain. Ecology 72(3):896–903

    Google Scholar 

  • Holling CS (1959a) The components of predation as revealed by a study of small-mammal predation of the European pine sawfly. Can Entomol 91(5):293–320

    Google Scholar 

  • Holling CS (1959b) Some characteristics of simple types of predation and parasitism. Can Entomol 91(7):385–398

    Google Scholar 

  • Holling CS (1973) Resilience and stability of ecological systems. Annu Rev Ecol Syst 4(1):1–23

    Google Scholar 

  • Izhikevich EM (2000) Neural excitability, spiking and bursting. Int J Bifurcat Chaos 10(06):1171–1266

    MathSciNet  MATH  Google Scholar 

  • Kirillov SY, Nekorkin V (2015) Dynamic saddle-node bifurcation of the limit cycles in the model of neuronal excitability. Radiophys Quantum Electron 57(11):837–847

    Google Scholar 

  • Knowlton N (1992) Thresholds and multiple stable states in coral reef community dynamics. Am Zool 32(6):674–682

    Google Scholar 

  • Laycock WA (1991) Stable states and thresholds of range condition on north American rangelands: a viewpoint. J Range Manag 44(5):427–433

    Google Scholar 

  • Lewontin RC (1969) The meaning of stability. Brookhaven Symp Biol 22:13–24

    Google Scholar 

  • Liu X, Chen L (2003) Complex dynamics of holling type II lotka-volterra predator–prey system with impulsive perturbations on the predator. Chaos Solitons Fractals 16(2):311–320

    MathSciNet  MATH  Google Scholar 

  • Liu Y, Zeng R, Liu D, Luo S, Wu H, An M (2006) Modelling dynamics of plant defence volatiles using the an-liu-johnson-lovett model. Allelopath J 18(2):215

    Google Scholar 

  • Liu YH, Li Liu D, An M, Fu YL, Zeng RS, Luo SM, Wu H, Pratley J (2009) Modelling tritrophic interactions mediated by induced defence volatiles. Ecol Model 220(23):3241–3247

    Google Scholar 

  • May RM (1977) Thresholds and breakpoints in ecosystems with a multiplicity of stable states. Nature 269(5628):471–477

    Google Scholar 

  • Mithöfer A, Boland W, Maffei ME (2018) Chemical ecology of plant–insect interactions. Ann Plant Rev 261–291

  • Moayeri H, Ashouri A, Poll L, Enkegaard A (2007) Olfactory response of a predatory mirid to herbivore induced plant volatiles: multiple herbivory vs single herbivory. J Appl Entomol 131(5):326–332

    Google Scholar 

  • Mondal R, Kesh D, Mukherjee D (2018) Influence of induced plant volatile and refuge in tritrophic model. Energy Ecol Environ 3(3):171–184

    Google Scholar 

  • Mondal R, Kesh D, Mukherjee D (2019) Role of induced volatile emission modelling tritrophic interaction. Differ Equ Dyn Syst. https://doi.org/10.1007/s12591-019-00458-y

    Article  Google Scholar 

  • Nyström M, Folke C, Moberg F (2000) Coral reef disturbance and resilience in a human-dominated environment. Trends Ecol Evol 15(10):413–417

    Google Scholar 

  • Pashalidou FG, Eyman L, Sims J, Buckley J, Fatouros NE, De Moraes CM, Mescher MC (2020) Plant volatiles induced by herbivore eggs prime defences and mediate shifts in the reproductive strategy of receiving plants. Ecol Lett 23(7):1097–1106

    Google Scholar 

  • Perko L (2013) Differential equations and dynamical systems, vol 7. Springer, Berlin

    MATH  Google Scholar 

  • Poelman EH, Dicke M (2018) Plant-mediated interactions among insects within a community ecological perspective. Ann Plant Rev 309–337

  • Price PW, Bouton CE, Gross P, McPheron BA, Thompson JN, Weis AE (1980) Interactions among three trophic levels: influence of plants on interactions between insect herbivores and natural enemies. Annu Rev Ecol Syst 11(1):41–65

    Google Scholar 

  • Rasmann S, Turlings TC (2007) Simultaneous feeding by aboveground and belowground herbivores attenuates plant-mediated attraction of their respective natural enemies. Ecol Lett 10(10):926–936

    Google Scholar 

  • Rodriguez-Saona C, Crafts-Brandner SJ, Cañas LA (2003) Volatile emissions triggered by multiple herbivore damage: beet armyworm and whitefly feeding on cotton plants. J Chem Ecol 29(11):2539–2550

    Google Scholar 

  • Rostás M, Ton J, Mauch-Mani B, Turlings TC (2006) Fungal infection reduces herbivore-induced plant volatiles of maize but does not affect naive parasitoids. J Chem Ecol 32(9):1897

    Google Scholar 

  • Runyon JB, Mescher MC, De Moraes CM (2006) Volatile chemical cues guide host location and host selection by parasitic plants. Science 313(5795):1964–1967

    Google Scholar 

  • Scheffer M, Hosper S, Meijer M, Moss B, Jeppesen E (1993) Alternative equilibria in shallow lakes. Trends Ecol Evol 8(8):275–279

    Google Scholar 

  • Scheffer M, Carpenter S, Foley JA, Folke C, Walker B (2001) Catastrophic shifts in ecosystems. Nature 413(6856):591

    Google Scholar 

  • Shiojiri K, Takabayashi J, Yano S, Takafuji A (2000) Flight response of parasitoids toward plant-herbivore complexes: a comparative study of two parasitoid-herbivore systems on cabbage plants. Appl Entomol Zool 35(1):87–92

    Google Scholar 

  • Silva DB, Urbaneja A, Pérez-Hedo M (2021) Response of mirid predators to synthetic herbivore-induced plant volatiles. Entomol Exp Appl 169(1):125–132

    Google Scholar 

  • Sutherland JP (1974) Multiple stable points in natural communities. Am Nat 108(964):859–873

    Google Scholar 

  • Van Den Boom CE, Van Beek TA, Posthumus MA, De Groot A, Dicke M (2004) Qualitative and quantitative variation among volatile profiles induced by tetranychus urticae feeding on plants from various families. J Chem Ecol 30(1):69–89

    Google Scholar 

  • Verheggen FJ, Capella Q, Schwartzberg EG, Voigt D, Haubruge E (2009) Tomato-aphid-hoverfly: a tritrophic interaction incompatible for pest management. Arthropod-Plant Interact 3(3):141–149

    Google Scholar 

  • Vos M, Berrocal SM, Karamaouna F, Hemerik L, Vet L (2001) Plant-mediated indirect effects and the persistence of parasitoid-herbivore communities. Ecol Lett 4(1):38–45

    Google Scholar 

  • Vucetich JA, Hebblewhite M, Smith DW, Peterson RO (2011) Predicting prey population dynamics from kill rate, predation rate and predator–prey ratios in three wolf-ungulate systems. J Anim Ecol 80(6):1236–1245

    Google Scholar 

Download references

Acknowledgements

We thank the anonymous reviewers and the editor for their insightful comments in improving the manuscript. Ritwika Mondal is supported by DST/INSPIRE Fellowship/2015/IF 150747. Dipak Kesh acknowledges support by RUSA 2.0.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Dipak Kesh.

Additional information

Publisher's Note

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

Appendices

Appendix A: Positivity and Boundedness

We first show that solutions of system (1) are positive for all \(t\ge 0\). Integrating first three equations from (1) with the initial conditions, we get

$$\begin{aligned} x(t)&= x_0 \exp \left[ \int ^t_0 \left( r(1-\frac{x(\theta )}{k})-\frac{c_1y_1(\theta )}{a_1+x(\theta )}\right) d\theta \right] \ge 0 \\ y_1(t)&= y_{10} \exp \left[ \int ^t_0 \left( -d_1+\frac{p_1c_1x(\theta )}{a_1+x(\theta )} -\frac{m_1z(\theta )}{a_3+y_1(\theta )}\right) d\theta \right] \ge 0 \\ y_2(t)&= y_{20} \exp \left[ \int ^t_0\left( -d_2+\frac{p_2c_2x(\theta )}{a_2+x(\theta )}\right) d\theta \right] \ge 0. \end{aligned}$$

From the last equation of system (1), note that

$$\begin{aligned} \frac{\mathrm{d}z}{\mathrm{d}t}\ge -d_3z+\frac{m_1n_1y_1z}{a_3+y_1}.] \end{aligned}$$

Integrating the above inequality, we get

$$\begin{aligned} z(t)= z_0\exp [\int ^t_0\{-d_3+\frac{m_1n_1y_1(\theta )}{a_3+y_1(\theta )}\}d\theta ]\ge 0. \end{aligned}$$

Hence, all solutions of (1) with \((x_0, y_{10}, y_{20}, z_0)\in R^4_+\) remain in for \(t\ge 0\).

To show, all solutions of system (1) are bounded, let, (\(x(t),~y_1(t),~y_2(t),z(t)\)) be any solution of system (1) with positive initial conditions ( \(x_0,y_{10},y_{20},z_0\)). As \(\frac{\mathrm{d}x}{\mathrm{d}t}\le rx(1-\frac{x}{k})\), by a standard comparison theorem we have,

$$\begin{aligned} \lim _{t\rightarrow \infty }\sup {x(t)}\le K_1, \text{ where }~K_1=\max \{x_0,k \}. \end{aligned}$$

Next to show, \(y_1(t)\) is bounded we consider a function \(W_1=x+\frac{y_1}{p_1}\) and by taking the time derivative along the solutions of system (1), we get

$$\begin{aligned} \frac{\mathrm{d}W_1}{\mathrm{d}t}+m_{11}W_1\le (r+1)K_1 \text{ where } m_{11}=\min \{1,d_1,d_2\}. \end{aligned}$$

Applying a theorem in differential inequalities (Birkhoff and Rota 1989), we have

$$\begin{aligned}&0 \le W_1\le (r+1)\frac{K_1}{m_{11}}+\frac{W_1(x_0,y_{10})}{e^{m_{11}t}},\\&\quad \text {and} ~\text {as}~ t\rightarrow \infty ,~~0\le W_1(t)\le (r+1)\frac{K_1}{m_{11}}=\gamma _{11} ~\text {(say)}. \end{aligned}$$

Therefore, \(y_1\le p_1\gamma _{11}\). Similarly, we can show that \(y_2\le p_2\gamma _{11}\). Again, we formulate a function \(W=x+\frac{1}{p_1}y_1+\frac{1}{p_2}y_2+\frac{1}{n_1p_1}z\). Now we have \(0\le W(t)\le \frac{1}{\bar{m}}[(r+1)K_1+\frac{K_1\gamma _{11}}{n_1p_1}(\mu _1p_1+\mu _2p_2)]=\gamma _2\) (say), where \(\bar{m}=\min \{1,d_1,d_2,d_3\}\).

Appendix B: Existence of Equilibria

System (1) admits five equilibrium points:

(i) Population-free equilibrium, \(E_0(0,0,0,0)\);

(ii) \(E_k(k,0,0,0)\);

(iii) \(E_1(\hat{x},0,\hat{y_2},\hat{z})\), where \(\hat{x}=\frac{a_2d_2}{p_2c_2-d_2},~\hat{y_2}=r(1-\frac{\hat{x}}{k})\frac{(a_2+\hat{x})}{c_2}, \hat{z}=\frac{\mu _2 \hat{x}\hat{y_2}}{d_3}\), equilibrium \(E_1\) exists if \(c_2 > \frac{d_2}{p_2}\);

(iv) \(E_2(\bar{x}, \bar{y_1},0,\bar{z})\), where \(\bar{x}\) is the positive real roots of the equation:

$$-r\mu _1 m_1x^4+G_2x^3+G_3x^2+G_4x+G_5=0,$$

where \(G_2=\mu _1m_1r(k-a_1)-(\mu _1m_1a_1+(m_1n_1-d_3)(p_1c_1-d_1))r,\)

\(G_3=\mu _1m_1rka_1+r(k-a_1)(\mu _1m_1a_1+(m_1n_1-d_3)(p_1c_1-d_1)))+(m_1n_1-d_3)d_1a_1r,\)

\(G_4=(\mu _1m_1a_1+(m_1n_1-d_3)(p_1c_1-d_1))rka_1-r(k-a_1)(m_1n_1-d_3)d_1a_1\) \(~~~~~~~~~ -d_3a_3c_1k(p_1c_1-d_1),\)

\(G_5=d_3a_3c_1d_1a_1k-d_1a_1rka_1(m_1n_1-d_3).\)

Therefore, we have \(\bar{y_1}=\frac{r}{c_1}(1-\frac{\bar{x}}{k})(a_1+x)\), and \(\bar{z}=\frac{a_3+\bar{y_1}}{m_1}(\frac{p_1c_1\bar{x}}{a_1+\bar{x}}-d_1)\).

v) \(E^*(x^*,y_1^*,y_2^*,z^*)\), when

$$\begin{aligned}&x^*=\frac{a_2d_2}{p_2c_2-d_2},\\&y_1^*=\frac{c_2kd_3a_3H-\mu _2 m_1 x^* r(k-x^*)(a_2+x^*) }{k[c_2H(m_1 n_1 -d_3)+ \mu _1 m_1 x^* c_2 -\mu _2 c_1m_1x^*]},\\&y_2^*=[r(1-\frac{x^*}{k})-\frac{c_1y_1^*}{a_1+x^*}]\frac{a_2+x^*}{c_2},\\&z^*=\frac{H}{m_1}(a_3+y_1^*) \end{aligned}$$

where \(H=\frac{p_1 c_1 x^*}{a_1+x^*}-d_1\), \(m_1^{**}=\frac{d_3c_2H}{c_2Hn_1+(\mu _1c_2-\mu _2c_1)x^*}\) and \(m_1^*=\frac{c_2kd_3a_3H}{\mu _2x^*r(k-x^*)(a_2+x^*)}\).

\(x^*\), \(y_1^*\), \(y_2^*\) and \(z^*\) are feasible if

(i) \(p_2c_2>d_2\), (ii) \(m_1^{**}<m_1<m_1^*\) and (iii) \(0\le d_1 < \frac{p_1c_1d_2a_2}{p_2c_2a_1-d_2a_1+d_2a_2}\) hold simultaneously.

Appendix C: Stability and Bifurcation Analysis

Variational matrix of system (1) at any arbitrary point \((x, y_1, y_2, z)\) is

$$\begin{aligned}&V(x, y_1, y_2, z)=[v_{ij}]_{4\times 4}, \\&\text {where},~ v_{11}=r-\frac{2rx}{k}-\frac{a_1 c_1 y_1}{(a_1+x)^2}-\frac{a_2 c_2 y_2}{(a_2+x)^2},~ v_{12}=-\frac{c_1x}{a_1+x}, \\&v_{13}=-\frac{c_2x}{a_2+x}, ~v_{21}=\frac{a_1 c_1 p_1 y_1}{(a_1+x)^2},\\&v_{22}=-d_1+\frac{c_1 p_1 x}{a_1+x}-\frac{m_1 z a_3}{(a_3 +y_1))^2}, ~v_{24}= \frac{m_1 y_1}{a_3 +y_1}, v_{31}=\frac{a_2 c_2 p_2 y_2}{(a_2+x)^2}, \\&v_{33}=-d_2+\frac{c_2 p_2 x}{a_2+x}, \\&v_{41}=\mu _1 y_1+\mu _2 y_2, ~v_{42} =\mu _1 x+\frac{m_1 n_1 a_3 z}{(a_3+y_1))^2}, ~v_{43}= \mu _2 x, \\&v_{44}=-d_3+\frac{m_1 n_1 y_1}{a_3+y_1}, ~v_{14}=v_{23}=v_{32}=v_{34}=0. \end{aligned}$$

1.1 Stability Analysis of Equilibrium \(E_0\)

Variational matrix of system (1) at equilibrium \(E_0\) is

$$\begin{aligned} V_0(0,0,0,0)=\left( \begin{array}{cccc} r &{}\quad 0 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad -d_1 &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad -d_2 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad 0 &{}\quad -d_3 \end{array}\right) . \end{aligned}$$

The eigenvalues of \(V_0\) is \(r~(>0)\), \(-d_1~(<0)\), \(-d_2~(<0)\) and \(-d_3~(<0)\). One eigenvalue is positive. Therefore, equilibrium \(E_0\) is unstable always.

1.1.1 Stability Analysis of Equilibrium \(E_k\)

Variational matrix of system (1) at equilibrium \(E_1\) is

$$\begin{aligned}V_1(k,0,0,0)=\left( \begin{array}{cccc} -r &{}\quad -\frac{c_1k}{a_1+k} &{}\quad -\frac{c_2k}{a_2+k} &{}\quad 0 \\ 0 &{}\quad -d_1+\frac{c_1 p_1 k}{a_1+k} &{}\quad 0 &{}\quad 0 \\ 0 &{}\quad 0 &{}\quad -d_2+\frac{c_2 p_2 k}{a_2+k} &{}\quad 0 \\ 0 &{}\quad \mu _1 k &{}\quad \mu _2 k &{}\quad -d_3 \end{array} \right) . \end{aligned}$$

The eigenvalues of \(V_1\) is \(-r~(<0)\), \(-d_1+\frac{c_1 p_1 k}{a_1+k}\), \(-d_2+\frac{c_2 p_2 k}{a_2+k}\) and \(-d_3~(<0)\). Two eigenvalues are negative and rest of two eigenvalues are negative, if \(d_1>\frac{c_1 p_1 k}{a_1+k}\) and \(d_2>\frac{c_2 p_2 k}{a_2+k}\).

Therefore, equilibrium \(E_1\) is asymptotically stable, if \(d_1>\frac{c_1 p_1 k}{a_1+k}\) or \(d_2>\frac{c_2 p_2 k}{a_2+k}\).

1.1.2 Stability Analysis of Equilibrium \(E_1\)

Variational matrix of system (1) at equilibrium \(E_1\) reads as

$$\begin{aligned} V_1(\hat{x},0,\hat{y_2},\hat{z})=\left( \begin{array}{cccc} \hat{a_{11}} &{}\quad \hat{a_{12}} &{}\quad \hat{a_{13}} &{}\quad 0 \\ 0 &{}\quad \hat{a_{22}} &{}\quad 0 &{}\quad 0 \\ \hat{a_{31}} &{}\quad 0 &{}\quad 0 &{}\quad 0 \\ \hat{a_{41}} &{}\quad \hat{a_{42}} &{}\quad \hat{a_{43}} &{}\quad \hat{a_{44}} \\ \end{array} \right) , \end{aligned}$$

where \(\hat{a_{11}}=-\frac{r \hat{x}}{k}+\frac{c_2 \hat{x} \hat{y_2}}{(a_2+\hat{x})^2},~ \hat{a_{12}}=-\frac{c_1 \hat{x}}{a_1+\hat{x}},~ \hat{a_{13}}=-\frac{c_2 \hat{x}}{a_2+\hat{x}} \)\(\hat{a_{22}}=-d_1+\frac{c_1 p_1 \hat{x}}{a_1+\hat{x}}-\frac{m_1 \hat{z}}{a_3} ,~\hat{a_{31}}= \frac{c_2 a_2 p_2 \hat{y_2}}{(a_2+\hat{x})^2} ,~\hat{a_{41}}=\mu _2 \hat{y_2}\)\(\hat{a_{42}}=\mu _1 \hat{x}+\frac{m_1 n_1 \hat{z}}{a_3} ,~\hat{a_{43}}=\mu _2 \hat{x} ,~\hat{a_{44}}=-d_3.\)

The corresponding characteristic equation is \((d_3+\xi _1)(\hat{a_{22}}-\xi _1)(\xi _1^2- \hat{a_{11}}\xi _1-\hat{a_{13}}\hat{a_{31}})=0.\)

From Routh–Hurwitz criterion, equilibrium \(E_1\) will be locally asymptotically stable if \(\hat{a_{22}}<0\) and \(\hat{a_{11}}<0,\) and unstable otherwise.

Analysis for transcritical bifurcation:

One of the eigenvalues of the variational matrix \(V_1\) at equilibrium \(E_1(\hat{x},0,\hat{y_2},\hat{z})\) is

$$\begin{aligned} \hat{a_{22}}=-d_1+\frac{p_1c_1\hat{x}}{a_1+\hat{x}}-\frac{m_1\hat{z}}{a_3}. \end{aligned}$$

If we choose \(m_1=\frac{a_3}{\hat{z}}(\frac{p_1c_1a_2d_2}{a_1(p_2c_2 -d_2)+a_2d_2}-d_1)=\hat{m_1} \text{(say) },\) then one eigenvalue of \(V_1\) will be zero.

Eigenvectors of \(V_1\) and \(V_3^T\) with respect to zero eigenvalue are \(\hat{U}=(0,\hat{u_2},\hat{u_3},\hat{u_4})^T\) and \(\hat{\gamma }_{m_1}=(0,1,0,0)^T\) where \(\hat{u_2}=\hat{\theta }\), \(\hat{\theta } \) is a real number; \(\hat{u_3}=-\frac{\hat{a_{12}}}{\hat{a_{13}}}\hat{\theta }\); and \(\hat{u_4}=-\frac{\hat{\theta }}{\hat{a_{13}}\hat{a_{44}}}(\hat{a_{13}}\hat{a_{42}}-\hat{a_{12}}\hat{a_{43}})\).

After some mathematical manipulation, we get \(\hat{\gamma }_{m_1}^TF_{m_1}(E_1,\hat{m_1})=0\) which implies that saddle node bifurcation does not occur at \(\hat{m_1}\) (Sotomayor’s theorem (Perko 2013).

Further, we get \(\hat{\gamma }_{m_1}^T[Df_{m_1}(E_1,\hat{m_1})v_{m_1}] =-\frac{\hat{z}\hat{\theta }}{a_3}\ne 0,\) and \(\hat{\gamma }_{m_1}^T[D^2F{m_1}(E_1,\hat{m_1})(v_{m_1},v_{m_1})] =\frac{2\hat{m_1}\hat{\theta }}{a_3}(\frac{\hat{z}\hat{u_2}}{a_3}-\hat{u_4}) \ne 0,\) if \(\frac{\hat{z}\hat{u_2}}{a_3}-\hat{u_4} \ne 0.\)

Thus, system (1) possesses a transcritical bifurcation at \(m_1=\hat{m_1}\), if \(\frac{\hat{z}\hat{u_2}}{a_3}-\hat{u_4} \ne 0.\)

Through the transcritical bifurcation, system (1) exchanges the stability between the interior (\(E^*\)) and prey herbivore-free (\(E_1\)) equilibrium points, when \(y_1\) dies out.

1.1.3 Stability Analysis of Equilibrium \(E_2\)

Variational matrix of system (1) at equilibrium \(E_2\) reads as

$$\begin{aligned} V_2(\bar{x},\bar{y_1},0,\bar{z})=\left( \begin{array}{cccc} \bar{a_{11}} &{}\quad \bar{a_{12}} &{}\quad \bar{a_{13}} &{}\quad 0 \\ \bar{a_{21}} &{}\quad \bar{a_{22}} &{}\quad 0 &{}\quad \bar{a_{24}} \\ 0 &{}\quad 0 &{}\quad \bar{a_{33}} &{}\quad 0 \\ \bar{a_{41}} &{}\quad \bar{a_{42}} &{}\quad \bar{a_{43}} &{}\quad \bar{a_{44}} \end{array} \right) , \end{aligned}$$

where

$$\begin{aligned}&\bar{a_{11}}= -\frac{r \bar{x}}{k}+\frac{c_1 \bar{x} \bar{y_1}}{(a_1+\bar{x})^2}, ~\bar{a_{12}}=-\frac{c_1 \bar{x}}{a_1+\bar{x}}, ~\bar{a_{13}}=-\frac{c_2 \bar{x}}{a_2+\bar{x}}, \\&\bar{a_{21}}=-\frac{c_1 p_1 a_1 \bar{y_1}}{(a_1+\bar{x})^2}, ~\bar{a_{22}}= \frac{m_1 \bar{y_1} \bar{z}}{(a_3 +\bar{y_1})^2},\\&\bar{a_{24}}=-\frac{m_1 \bar{y_1}}{a_3 +\bar{y_1}}, ~\bar{a_{33}}=-d_2+\frac{c_2 p_2 \bar{x}}{a_2+\bar{x}}, ~\bar{a_{41}}=\mu _1 \bar{y_1},\\&\bar{a_{42}}=\mu _1\bar{x}+\frac{m_1 n_1 a_3\bar{z}}{(a_3 +\bar{y_1})^2}, ~\bar{a_{43}}=\mu _2 \bar{x}, ~\bar{a_{44}}= -\frac{\mu _1 \bar{x} \bar{y_1}}{\bar{z}}. \end{aligned}$$

The corresponding characteristic equation is

$$\begin{aligned} (\bar{a_{33}}-\xi _2)(\xi _2^3+\bar{C_1}\xi _2^2+\bar{C_2}\xi _2+\bar{C_3})=0, \end{aligned}$$

where \(\bar{C_1}\)=\(- \bar{a_{11}} - \bar{a_{22}} - \bar{a_{44}}\), \(\bar{C_2}\)=\(\bar{a_{11}} \bar{a_{22}} - \bar{a_{12}} \bar{a_{21}} + \bar{a_{11}} \bar{a_{44}} + \bar{a_{22}} \bar{a_{44}} - \bar{a_{24}} \bar{a_{42}}\), \(\bar{C_3}\)=\(\bar{a_{11}} \bar{a_{24}} \bar{a_{42}} - \bar{a_{11}} \bar{a_{22}} \bar{a_{44}} + \bar{a_{12}} \bar{a_{21}} \bar{a_{44}} - \bar{a_{12}} \bar{a_{24}} \bar{a_{41}}\).

Following Routh–Hurwitz criterion, equilibrium \(E_2\) will be locally asymptotically stable, if

$$\begin{aligned} \left( d_2>\frac{c_2 p_2 \bar{x}}{a_2+\bar{x}}, ~\bar{C_1}, ~\bar{C_3}, ~\text {and} ~\bar{C_1}\bar{C_2}-\bar{C_3} \right) >0, ~\text {and unstable otherwise.} \end{aligned}$$

Analysis for saddle node bifurcation:

If \(\bar{C_3}=0\), then one of the eigenvalue of \(V_2\) will be zero. We choose the value of \(m_1=\bar{m_1}\) (say) so that \(\bar{C_3}\) becomes zero.

Eigenvectors of \(V_2\) and \(V_2^T\) with respect to zero eigenvalue are \(\bar{U}=(\bar{\theta },\bar{u_2},0,\bar{u_4})^T\) and \(\bar{\gamma }_{m_1}=(\bar{v_1},\bar{v_2},\bar{v_3},\bar{v_4})^T\) and \(\bar{u_2}=-\bar{\theta }\frac{\bar{a_{11}}}{\bar{a_{12}}}\), where \(\bar{\theta } \) is a real number.

\(\bar{u_4}=-\frac{\bar{\theta } }{\bar{a_{12}}\bar{a_{24}}}(\bar{a_{12}}\bar{a_{21}}-\bar{a_{11}}\bar{a_{22}})\), \(\bar{v_2}=-\frac{\bar{a_{44}}}{\bar{a_{24}}}\), \(\bar{v_1}=-\frac{1}{\bar{a_{11}}}(\bar{a_{21}}\bar{v_2}+\bar{a_{41}})\), \(\bar{v_3}=-\frac{1}{\bar{a_{33}}}(\bar{a_{43}}+\bar{a_{13}}\bar{v_1})\).

After some mathematical manipulation, we get

$$\begin{aligned} \bar{\gamma }_{m_1}^TF_{m_1}(E_2,\bar{m_1})=\frac{\bar{y_1}\bar{z}}{a_3+\bar{y_1}}\{n_1+(a_3+\bar{y_1})\frac{\mu _1\bar{x}\bar{y_1}}{m_1\bar{y_1}\bar{z}}\} \ne 0 \end{aligned}$$

which implies that \(E_2\) possesses saddle node bifurcation at \(\hat{m_1}\) (Sotomayor’s theorem Perko 2013).

1.1.4 Stability Analysis of \(E^*\)

Variational matrix of system (1) at \(E^*\) reads as

$$\begin{aligned}&V_*(x^*,~y_1^*,~y_2^*,~z^*)=\left( \begin{array}{cccc} a_{11} &{}\quad a_{12} &{}\quad a_{13} &{}\quad 0 \\ a_{21} &{}\quad a_{22} &{}\quad 0 &{}\quad a_{24} \\ a_{31} &{}\quad 0 &{}\quad 0 &{}\quad 0 \\ a_{41} &{}\quad a_{42} &{}\quad a_{43} &{}\quad a_{44} \end{array} \right) , \\&\text {where} ~a_{11}=-\frac{r x^*}{K}+\frac{c_1 x^* y_1^*}{(a_1+x^*)^2}+\frac{c_2 x^* y_2^*}{(a_2+x^*)^2} ,~a_{12}=-\frac{c_1 x^*}{a_1+x^*},\\&a_{13}=-\frac{c_2 x^*}{a_2+x^*},~a_{21}=-\frac{c_1 p_1 a_1 y_1^*}{(a_1+x^*)^2},\\&a_{22}=\frac{m_1 y_1^* z^*}{(a_3 +y_1^*)^2}, ~a_{24}=-\frac{m_1 y_1^*}{a_3 +y_1^*},~a_{31}=-\frac{c_2 p_2 a_2 y_2^*}{(a_2+x^*)^2},\\&a_{41}=\mu _1 y_1^* +\mu _2 y_2^*,~a_{42}= \frac{m_1 n_1 a_3 z^*}{(a_3 +y_1^*)^2}+\mu _1 x^*,\\&a_{43}=\mu _2 x^*, ~\text {and}~a_{44}=-d_3+\frac{m_1 n_1 y_1^*}{a_3 +y_1^*}. \end{aligned}$$

So the corresponding characteristic equation will be

$$\begin{aligned} \xi _*^4+ C_1\xi _*^3+C_2\xi _*^2+C_3\xi _*+C_4=0, \end{aligned}$$

where \(C_1\)=\(- a_{11} - a_{22} - a_{44}\), \(C_2\)=\(a_{11}a_{22} - a_{12}a_{21} - a_{13}a_{31} + a_{11}a_{44} + a_{22}a_{44} - a_{24}a_{42}\), \(C_3\)=\(a_{13}a_{22}a_{31} - a_{11}a_{22}a_{44} + a_{11}a_{24}a_{42} + a_{12}a_{21}a_{44} - a_{12}a_{24}a_{41} + a_{13}a_{31}a_{44},\) and \(C_4\)=\(- a_{12}a_{24}a_{31}a_{43} - a_{13}a_{22}a_{31}a_{44} + a_{13}a_{24}a_{31}a_{42}.\)

By Routh–Hurwitz criterion, \(E^*\) will be locally asymptotically stable, when

$$\begin{aligned} \left( C_1, ~C_3, ~C_4, ~\text {and}~C_1C_2C_3-C_3^2C_1^2C_4\right) >0, ~\text {and unstable otherwise}. \end{aligned}$$

Appendix D: Emergence of Chaos

Further, to check possible occurrence of some complex phenomenon we vary the environmental carrying capacity, which is one of the most vital parameter in ecological context. We draw one parameter bifurcation diagram against k in Fig. 7.

The system (1) undergoes through several critical transitions, such as HB at \(k=4.197\), saddle node bifurcation of limit cycle (SNLC) at \(k=5.753\), transcritical bifurcation of limit cycle (TRLC) at \(k=6.94\), torus bifurcation (TB) at \(k=8.536\) and boundary crisis (BC) at \(k=8.714\). Yellow, red and green dots represent stable cyclic states, whereas blue for unstable cycles.

Fig. 7
figure 7

Role of environmental carrying capacity for \(m_1=0.9\); other parameters are as in Table 1. Yellow, red and green dots represent stable cyclic state, whereas blue dots for unstable cycles. Critical transition occurred at HB = 4.197, SNLC = 5.753, TRLC = 6.94, TB = 8.536 and BC = 8.714. Chaotic dynamics are magnified in the inset

At lower carrying capacity, system (1) stays in non-prey herbivore-free equilibrium state \(E_2\). All populations, except non-prey herbivore, transit from steady state to oscillation after HB which is shown by yellow dots. For higher values of k, an unstable cycle (in blue) coexists with the limit cycle (yellow). At SNLC point, stable cycle in red collides with the unstable cycle (blue) and disappears. Further, with the increase in k, two stable limit cycles (red and green) exchange their stability at TRLC. All population survive at oscillatory state shown in red, whereas prey herbivore goes to extinction and remaining three species maintain the oscillatory state, denoted by green dots.

For larger k, regular cycle (yellow) develops quasi-periodicity via TB, which eventually leads to the onset of irregular dynamics through quasi-periodic route to chaos. The chaotic state coexists with the alternate higher amplitude regular cycle (green). With the further increase in k, apparent chaos disappears at the BC. For a range of \(k \in (8.536,8.714)\), the chaotic state coexists with the stable periodic cycle (green). It restores the regularity in temporal dynamics as the higher-amplitude oscillation in green still remains for larger values of carrying capacity.

The system exhibits alternate cyclic states for a range of \(5.753<k<8.714\). Before TRLC, the oscillatory state in red coexists with the limit cycles in yellow which exclude non-prey herbivore. After TRLC, principle of competitive exclusion is somewhat followed as either of the two herbivores can survive in one of the alternate cyclic states (shown in green and yellow).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mondal, R., Saha, S., Kesh, D. et al. Basin Transition and Alternative States: Role of Multi-species Herbivores-Induced Volatile in Plant–Insect Interactions. Bull Math Biol 83, 100 (2021). https://doi.org/10.1007/s11538-021-00930-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11538-021-00930-3

Keywords

Navigation