Skip to main content
Log in

Bifurcation Analysis of Models with Uncertain Function Specification: How Should We Proceed?

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

Abstract

When we investigate the bifurcation structure of models of natural phenomena, we usually assume that all model functions are mathematically specified and that the only existing uncertainty is with respect to the parameters of these functions. In this case, we can split the parameter space into domains corresponding to qualitatively similar dynamics, separated by bifurcation hypersurfaces. On the other hand, in the biological sciences, the exact shape of the model functions is often unknown, and only some qualitative properties of the functions can be specified: mathematically, we can consider that the unknown functions belong to a specific class of functions. However, the use of two different functions belonging to the same class can result in qualitatively different dynamical behaviour in the model and different types of bifurcation. In the literature, the conventional way to avoid such ambiguity is to narrow the class of unknown functions, which allows us to keep patterns of dynamical behaviour consistent for varying functions. The main shortcoming of this approach is that the restrictions on the model functions are often given by cumbersome expressions and are strictly model-dependent: biologically, they are meaningless. In this paper, we suggest a new framework (based on the ODE paradigm) which allows us to investigate deterministic biological models in which the mathematical formulation of some functions is unspecified except for some generic qualitative properties. We demonstrate that in such models, the conventional idea of revealing a concrete bifurcation structure becomes irrelevant: we can only describe bifurcations with a certain probability. We then propose a method to define the probability of a bifurcation taking place when there is uncertainty in the parameterisation in our model. As an illustrative example, we consider a generic predator–prey model where the use of different parameterisations of the logistic-type prey growth function can result in different dynamics in terms of the type of the Hopf bifurcation through which the coexistence equilibrium loses stability. Using this system, we demonstrate a framework for evaluating the probability of having a supercritical or subcritical Hopf bifurcation.

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

  • Adamson, M. W., & Morozov, A Yu. (2012). When can we trust our model predictions? Unearthing structural sensitivity in biological systems. Proc R Soc A Math Phys Eng Sci, 469, 20120500.

    Article  MathSciNet  Google Scholar 

  • Adamson, M. W., & Morozov, A Yu. (2014). Defining and detecting structural sensitivity in biological models: developing a new framework.,. doi:10.1007/s00285-014-0753-3.

    MathSciNet  Google Scholar 

  • Arditi, R., & Ginzburg, L. R. (1989). Coupling in predator-prey dynamics: Ratio-dependence. J Theor Biol, 139, 311–326.

    Article  Google Scholar 

  • Arditi, R., Ginzburg, L. R., & AkV̧akaya, H. R. (1991). Variation in plankton densities among lakes: a case for ratio-dependent models. Am Nat, 138, 1287–1296.

    Article  Google Scholar 

  • Bautin, N., & Leontovich, Y. (1976). Methods and procedures for the qualitative investigation of dynamical systems in a plane. Moscow: Nauka.

    MATH  Google Scholar 

  • Berezovskaya, F., Karev, G., & Arditi, R. (2001). Parametric analysis of the ratio-dependent predator-prey model. J Math Biol, 43, 221–246.

    Article  MathSciNet  MATH  Google Scholar 

  • Berec, L., Angulo, E., & Courchamp, F. (2007). Multiple Allee effects and population management. Trends Ecol Evol, 22, 185–191.

    Article  Google Scholar 

  • Bishop, M. J., Kelaher, B. P., Smith, M., York, P. H., & Booth, D. J. (2006). Ratio-dependent response of a temperate Australia estuarine system to sustained nitrogen loading. Oecologia, 149, 701–708.

    Article  Google Scholar 

  • Cao, J., Fussmann, G. F., & Ramsay, J. O. (2008). Estimating a predator-prey dynamical model with the parameter cascades method. Biometrics, 64, 959–967.

    Article  MathSciNet  MATH  Google Scholar 

  • Chou, I., & Voit, E. O. (2009). Recent developments in parameter estimation and structure identification of biochemical and genomic systems. Math Biosci, 219, 57–83.

    Article  MathSciNet  MATH  Google Scholar 

  • Chow S-N, Li C, Wang D (1994) Normal forms and bifurcation of planar vector fields. Cambridge University Press, Cambridge.

  • Cordoleani, F., Nerini, D., Gauduchon, M., Morozov, A., & Poggiale, J.-C. (2011). Structural sensitivity of biological models revisited. J Theor Biol, 283, 82–91.

    Article  MathSciNet  Google Scholar 

  • de Mazancourt, C., & Dieckmann, U. (2004). Trade-off geometries and frequency-dependent selection. Am Nat, 164, 765–778.

    Article  Google Scholar 

  • Englund, G., & Leonardsson, K. (2008). Scaling up the functional response for spatially heterogeneous systems. Ecol Lett, 11, 440–449.

    Article  Google Scholar 

  • Freckleton, R. P., Watkinson, A. R., Green, R. E., & Sutherland, W. J. (2006). Census error and the detection of density dependence. J Anim Ecol, 75, 837–851.

    Article  Google Scholar 

  • Freedman, H. I., & Mathsen, R. M. (1993). Persistence in predator-prey systems with ratio-dependent predator influence. Bull Math Biol, 55, 817–827.

    Article  MATH  Google Scholar 

  • Fussmann, G. F., & Blasius, B. (2005). Community response to enrichment is highly sensitive to model structure. Biol Lett, 1, 9–12.

    Article  Google Scholar 

  • Gause, G. F. (1934). The struggle for existence. New York: Hafner Publishing Company.

    Book  MATH  Google Scholar 

  • Gilpin, M. E., & Ayala, F. J. (1973). Global models of growth and competition. Proc Natl Acad Sci USA, 70, 3590–3593.

    Article  MATH  Google Scholar 

  • González-Olivares, E., González-Yañez, B., Lorca, J. M., Rojas-Palma, A., & Flores, J. D. (2011). Consequences of double Allee effect on the number of limit cycles in a predator-prey model. Comput Math Appl, 62, 3449–3463.

    Article  MathSciNet  MATH  Google Scholar 

  • Gross, T., Ebenhoh, W., & Feudel, U. (2004). Enrichment and food-chain stability: The impact of different forms of predator-prey interaction. J Theor Biol, 227, 349–358.

    Article  MathSciNet  Google Scholar 

  • Gross, T., & Feudel, U. (2006). Generalized models as an universal approach to the analysis of nonlinear dynamical systems. Phys Rev E, 73, 016205.

    Article  Google Scholar 

  • Haque, M. (2009). Ratio-dependent predator-prey models of interacting populations. Bull Math Biol, 71, 430–452.

    Article  MathSciNet  MATH  Google Scholar 

  • Hopf, F. A., & Hopf, F. W. (1985). The role of the Allee effect in species packing. Theor Popul Biol, 27, 25–50.

    Article  MathSciNet  MATH  Google Scholar 

  • Hsu, S.-B., Hwang, T.-W., & Kuang, Y. (2001). Global analysis of the Michaelis-Menten-type ratio-dependent predator-prey system. J Math Biol, 42, 489–506.

    Article  MathSciNet  MATH  Google Scholar 

  • Jost, C., Arino, O., & Arditi, R. (1999). About deterministic extinction in ratio-dependent predator-prey models. Bull Math Biol, 61, 19–32.

    Article  MATH  Google Scholar 

  • Kuang, Y., & Beretta, E. (1998). Global qualitative analysis of a ratio-dependent predator-prey system. J Math Biol, 36, 389–406.

    Article  MathSciNet  MATH  Google Scholar 

  • Kisdi, E., Geritz, S. A. H., & Boldin, B. (2013). Evolution of pathogen virulence under selective predation: A construction method to find eco-evolutionary cycles. J Theor Biol,. doi:10.1016/j.jtbi.2013.05.023.

    MathSciNet  Google Scholar 

  • Kolmogorov, A. N. (1936). Sulla Teoria di Volterra della Lotta per l’Esistenza. Giornale Instituto Italiani Attuari, 7, 74–80.

    MATH  Google Scholar 

  • Kuznetsov, Y. A. (2004). Elements of applied bifurcation theory. New York: Springer.

    Book  MATH  Google Scholar 

  • Lewis, M. A., & Kareiva, P. (1993). Allee dynamics and the spread of invading organisms. Theor Popul Biol, 43, 141–158.

    Article  MATH  Google Scholar 

  • Li, B., & Kuang, Y. (2007). Heteroclinic bifurcation in the Michaelis-Menten type ratio-dependent predator-prey system. SIAM J Appl Math, 67, 1453–1464.

    Article  MathSciNet  MATH  Google Scholar 

  • Morris, W. F. (1990). Problems in detecting chaotic behavior in natural populations by fitting simple discrete models. Ecology, 71, 1849–1862.

    Article  Google Scholar 

  • Myerscough, M. R., Darwen, M. J., & Hogarth, W. L. (1996). Stability, persistence and structural stability in a classical predator-prey model. Ecol Model, 89, 31–42.

    Article  Google Scholar 

  • Owen, M. R., & Lewis, M. A. (2001). How predation can slow, stop or reverse a prey invasion. Bull Math Biol, 63, 655–684.

    Article  MATH  Google Scholar 

  • Reeve, J. D. (1997). Predation and bark beetle dynamics. Oecologia, 112, 48–54.

    Article  Google Scholar 

  • Rosenzweig, M. L. (1971). Paradox of enrichment: destabilization of exploitation ecosystems in ecological time. Science, 171, 385–387.

    Article  Google Scholar 

  • Savageau, M. A., & Voit, E. O. (1987). Recasting nonlinear differential equations as S-systems: A canonical nonlinear form. Math Biosci, 87, 83–115.

    Article  MathSciNet  MATH  Google Scholar 

  • Sen, M., Banerjee, M., & Morozov, A. (2012). Bifurcation analysis of a ratio-dependent prey-predator model with the Allee effect. Ecol Complex, 11, 12–27.

    Article  Google Scholar 

  • Sen, M., Banerjee, M., & Morozov, A. (2014). Stage-structured ratio-dependent predator-prey models revisited: when should the maturation lag result in systems’ destabilization? Ecol Complex, 19, 23–34.

  • Sibly, R. M., Barker, D., Denham, M. C., Hone, J., & Pagel, M. (2005). On the regulation of populations of mammals, birds, fish, and insects. Science, 309, 607–610.

    Article  Google Scholar 

  • Steuer, R., Gross, T., Selbig, J., & Blasius, B. (2006). Structural kinetic modeling of metabolic networks. PNAS, 103, 11868–11873.

    Article  Google Scholar 

  • Wang, M. E., & Kot, M. (2001). Speeds of invasion in a model with strong or weak Allee effects. Math Biosci, 171, 83–97.

    Article  MathSciNet  MATH  Google Scholar 

  • Wood, S. N., & Thomas, M. B. (1999). Super-sensitivity to structure in biological models. Proc R Soc Lond Ser B Biol Sci, 266, 565–570.

    Article  Google Scholar 

  • Xiao, D., & Ruan, S. (2001). Global dynamics of a ratio-dependent predator-prey system. J Math Biol, 43, 268–290.

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgments

We highly appreciated S. V. Petrovskii (University of Leicester) for comments and suggestions.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to A. Yu. Morozov.

Appendices

Appendix 1: The First Lyapunov Exponent in a Planner System

The stability of a limit cycle is determined by its First Lyapunov number \(L_{1}\). In the two-dimensional system

$$\begin{aligned} \dot{x}&= ax + by + f\left( {x,y} \right) ,\\ \dot{y}&= cx + dy + g\left( {x,y} \right) \end{aligned}$$

with \(f\left( {x,y} \right) =a_{20} x^{2}+a_{11} xy+a_{02} y^{2}+a_{30} x^{3}+a_{21} x^{2}y+a_{12} xy^{2}+a_{03} y^{3}\), and \(g\left( {x,y} \right) =b_{20} x^{2}+b_{11} xy+b_{02} y^{2}+b_{30} x^{3}+b_{21}x^{2}y+b_{12} xy^{2}+b_{03} y^{3}\), the first Lyapunov number is given by the following expression (Bautin and Leontovich 1976; Chow et al. 1994):

$$\begin{aligned} L_1&= -\frac{\pi }{4b\varDelta ^{\frac{3}{2}}}\left\{ \left[ ac\left( a_{11}^2 +a_{11} b_{02} +a_{02} b_{11} \right) +ab\left( b_{11}^2 +b_{11} a_{20} +b_{20} a_{11} \right) \right. \right. \\&\left. \left. +c^{2}\left( a_{11} a_{02} +2a_{02} b_{02} \right) -2ac\left( b_{02}^2 -a_{20} a_{02} \right) -2ab\left( a_{20}^2 -b_{20} b_{02} \right) \right. \right. \\&\left. \left. -b^{2}\left( 2a_{20} b_{20} +b_{11} b_{20} \right) +\left( bc-2a^{2} \right) \left( b_{11} b_{02} -a_{11} a_{20} \right) \right] \right. \\&\left. -\left( a^{2}+bc \right) \left[ 3\left( cb_{03} -ba_{30} \right) +2a\left( a_{21} +b_{12} \right) +\left( ca_{12} -bb_{21} \right) \right] \right\} , \end{aligned}$$

where \(\varDelta \) is the determinant of the Jacobian matrix.

In the case where the first Lyapunov number, evaluated at the equilibrium at the Hopf bifurcation point, is positive, then the resulting limit cycle will be stable, and the Hopf bifurcation is supercritical. If the first Lyapunov exponent is negative, the resulting limit cycle will be unstable and so the Hopf bifurcation is a subcritical one (Kuznetsov 2004).

Appendix 2: Deriving the Projection from Function Space to the Space of Local Values

Here, we derive and sketch a proof of the necessary and sufficient conditions for the existence of a function \(\tilde{r} \) in the \(\varepsilon _Q \)-neighbourhood of the base function \(r\) taking the values \(x^{*},\;\tilde{r} \left( {x^{*}} \right) ,\;\tilde{r} ^{{\prime }}\left( {x^{*}} \right) ,\;\tilde{r}^ {{\prime }{\prime }}(x^{*})\) and \(\tilde{r}^ {{\prime }{\prime }{\prime }}(x^{*})\).

For \(\tilde{r} \) to remain within the \(\varepsilon _Q\)-neighbourhood, it must first satisfy:

$$\begin{aligned} \tilde{r} \left( x \right) <r_{\varepsilon +} \left( x \right)&= 1+\varepsilon -x,\; \hbox {and}\\ \tilde{r} \left( x \right) >r_{\varepsilon -} \left( x \right)&= 1-\varepsilon -x. \end{aligned}$$

Essentially, this means that \(\tilde{r} (x)\) must remain between the red bounds in Fig. 6 over the whole domain—it must remain within distance \(\varepsilon \) of the base function.

\(\tilde{r}\) must also be in \(Q\), so must further satisfy:

  1. (i)

    \(\tilde{r} ^{{\prime }{\prime }{\prime }}\left( x \right) =\tilde{r} ^{{\prime }{\prime }{\prime }}\left( {x^{*}} \right) \;\forall x\in \left( {x^{*}-w,x^{*}+w} \right) ,\)

  2. (ii)

    \(\left| {\tilde{r}^{{\prime }{\prime }}(x)} \right| <A \quad \forall x\in \left[ {0,x_{\hbox {max}} } \right] ,\)

  3. (iii)

    \(\tilde{r} ^{{\prime }}\left( x \right) <0\quad \forall x\in \left[ {0,x_{\hbox {max}} } \right] ,\)

  4. (iv)

    \(\tilde{r} \left( 0 \right) >0.\)

Condition i) tells us that across the interval \(\left( {x^{*}-w,x^{*}+w} \right) ,\; \tilde{r} \) is given by the cubic:

$$\begin{aligned} \tilde{r} \left( x \right) =\tilde{r} \left( {x^{*}} \right) +\tilde{r} ^{{\prime }}\left( {x^{*}} \right) \left( {x-x^{*}} \right) +\tilde{r} ^{{\prime }{\prime }}\left( {x^{*}} \right) \left( {x-x^{*}} \right) ^{2}+\frac{\tilde{r} ^{{\prime }{\prime }{\prime }}\left( {x^{*}} \right) }{6}\left( {x-x^{*}}\right) ^{3} \end{aligned}$$
(9)

Therefore, an initial necessary condition for the existence of a valid function \(\tilde{r} \) attaining the values \(x^{*},\;\tilde{r} \left( {x^{*}} \right) ,\;\tilde{r} ^{{\prime }}\left( {x^{*}} \right) ,\;\tilde{r}^ {{\prime }{\prime }}(x^{*})\) and \(\tilde{r}^ {{\prime }{\prime }{\prime }}(x^{*})\) is that this cubic must stay between \(r_{\varepsilon -}\) and \(r_{\varepsilon +} \) over this interval. In addition, this cubic must always have a negative first derivative and not have a second derivative of magnitude greater than \(A\). Furthermore, \(\tilde{r}\) will be bounded above by the parabolas tangent to the above cubic at \(x^{*}-w\) and \(x^{*}+w\) with second derivative \(A\), and will be bounded below by the tangent parabolas with second derivative \(-A\). These are given by the blue curves in Fig. 6. Taking these upper and lower bounds over the intervals \([0,x^{*}-w)\) and \((x^{*}+w,x_\mathrm{max}]\), together with the fact that \(\tilde{r} \) must equal the cubic (B1) over \(\left( {x^{*}-w,x^{*}+w} \right) \), we can construct the following functions:

$$\begin{aligned}&\!\!\!\!P_1 \left( x \right) \\&=\left\{ \begin{array}{ll} B+C\left( {x-x^{*}+w} \right) -\frac{A}{2}\left( {x-x^{*}+w} \right) ^{2}: &{} x\in [0,x^{*}-w) \\ \tilde{r} \left( {x^{*}} \right) +\tilde{r} ^{{\prime }}\left( {x^{*}} \right) \left( {x-x^{*}} \right) +\frac{\tilde{r} ^{{\prime }{\prime }}\left( {x^{*}} \right) }{2}\left( {x-x^{*}} \right) ^{2}+\frac{\tilde{r} ^{{\prime }{\prime }{\prime }}\left( {x^{*}} \right) }{6}\left( {x-x^{*}} \right) ^{3}:&{} x\in \left( {x^{*}-w,x^{*}+w} \right) \\ D+E\left( {x-x^{*}-w} \right) -\frac{A}{2}\left( {x-x^{*}-w} \right) ^{2}: &{} x\in (x^{*}+w,x_\mathrm{max}]\\ \end{array} \right. \end{aligned}$$

and

$$\begin{aligned}&\!\!\!\! P_2 \left( x \right) \\&=\left\{ \begin{array}{ll} B+C\left( {x-x^{*}+w} \right) +\frac{A}{2}\left( {x-x^{*}+w} \right) ^{2}:&{}x\in [0,x^{*}-w) \\ \tilde{r} \left( {x^{*}} \right) +\tilde{r} ^{{\prime }}\left( {x^{*}} \right) \left( {x-x^{*}} \right) +\frac{\tilde{r} ^{{\prime }{\prime }}\left( {x^{*}} \right) }{2}\left( {x-x^{*}} \right) ^{2}+\frac{\tilde{r} ^{{\prime }{\prime }{\prime }}\left( {x^{*}} \right) }{6}\left( {x-x^{*}} \right) ^{3}:&{}x\in \left( {x^{*}-w,x^{*}+w} \right) \\ D+E\left( {x-x^{*}-w} \right) +\frac{A}{2}\left( {x-x^{*}-w} \right) ^{2}:&{} x\in (x^{*}+w,x_\mathrm{max} ] \\ \end{array} \right. \end{aligned}$$

where:

$$\begin{aligned} B&= \tilde{r} \left( {x^{*}} \right) -w\tilde{r} ^{{\prime }}\left( {x^{*}} \right) +\frac{w^{2}}{2}\tilde{r} ^{{\prime }{\prime }}\left( {x^{*}} \right) -w^{3}\tilde{r} ^{{\prime }{\prime }{\prime }}\left( {x^{*}} \right) ,\\ C&= \tilde{r} ^{{\prime }}\left( {x^{*}} \right) -w\tilde{r} ^{{\prime }{\prime }}\left( {x^{*}} \right) +\frac{w^{2}}{2}\tilde{r} ^{{\prime }{\prime }{\prime }}\left( {x^{*}} \right) ,\\ D&= \tilde{r} \left( {x^{*}} \right) +w\tilde{r} ^{{\prime }}\left( {x^{*}} \right) +\frac{w^{2}}{2}\tilde{r} ^{{\prime }{\prime }}\left( {x^{*}} \right) +w^{3}\tilde{r} ^{{\prime }{\prime }{\prime }}\left( {x^{*}} \right) ,\\ E&= \tilde{r} ^{{\prime }}\left( {x^{*}} \right) +w\tilde{r} ^{{\prime }{\prime }}\left( {x^{*}} \right) +\frac{w^{2}}{2}\tilde{r} ^{{\prime }{\prime }{\prime }}\left( {x^{*}} \right) . \end{aligned}$$

For any function \(\tilde{r} \in Q, P_1 \) and \(P_2 \) necessarily form lower and upper bounds for \(\tilde{r} \), since they are constructed as the extreme cases of functions in \(Q\). So we have \(P_1 \left( x \right) \le \tilde{r}\left( x \right) \le P_2 \left( x \right) \;\forall x\in \left[ {0,x_\mathrm{max}} \right] \) (indeed, \(\tilde{r} ,\; P_1 \) and \(P_2 \) coincide over the interval \(\left( {x^{*}-w,x^{*}+w} \right) )\), and therefore, the conditions:

$$\begin{aligned} P_1 \left( x \right)&< r_{\varepsilon -} \left( x \right) ,\\ P_2 \left( x \right)&< r_{\varepsilon +} \left( x \right) ,\\ P_1^{\prime } \left( x \right)&< 0\;\forall x\in \left( {x^{*}-w,x^{*}+w} \right) ,\\ \left| {P_1^{{\prime }{\prime }}(x)} \right|&< A\;\forall x\in \left( {x^{*}-w,x^{*}+w} \right) , \end{aligned}$$

are necessary (note: \(P_1 ,\; P_2 \) and \(\tilde{r} \) coincide over the interval \(\left( {x^{*}-w,x^{*}+w} \right) \) and so are interchangeable in the 3rd and 4th conditions). In terms of the figure, these conditions can be interpreted as requiring that the red upper and blue lower bounds clearly cannot cross.

It remains to be shown that they are sufficient. In order to prove this, it is enough to provide a method to construct a valid function \(\tilde{r} \) which remains between \(r_{\varepsilon -} \) and \(r_{\varepsilon +}\) given only these conditions. We already have \(\tilde{r} \) equal to \(P_1 \) and \(P_2 \) over \(\left( {x^{*}-w,x^{*}+w} \right) \), so only need to construct \(\tilde{r} \) over \([0,x^{*}-w)\) and \((x^{*}+w,x_{\hbox {max}} ]\). To do this, we can in fact use the exact same approach taken in Appendix 2 of Adamson and Morozov (2014).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Adamson, M.W., Morozov, A.Y. Bifurcation Analysis of Models with Uncertain Function Specification: How Should We Proceed?. Bull Math Biol 76, 1218–1240 (2014). https://doi.org/10.1007/s11538-014-9951-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-014-9951-9

Keywords

Navigation