Skip to main content
Log in

The Jungle Universe: coupled cosmological models in a Lotka–Volterra framework

  • Research Article
  • Published:
General Relativity and Gravitation Aims and scope Submit manuscript

Abstract

In this paper, we exploit the fact that the dynamics of homogeneous and isotropic Friedmann–Lemaître universes is a special case of generalized Lotka–Volterra system where the competitive species are the barotropic fluids filling the Universe. Without coupling between those fluids, Lotka–Volterra formulation offers a pedagogical and simple way to interpret usual Friedmann–Lemaître cosmological dynamics. A natural and physical coupling between cosmological fluids is proposed which preserves the structure of the dynamical equations. Using the standard tools of Lotka–Volterra dynamics, we obtain the general Lyapunov function of the system when one of the fluids is coupled to dark energy. This provides in a rigorous form a generic asymptotic behavior for cosmic expansion in presence of coupled species, beyond the standard de Sitter, Einstein-de Sitter and Milne cosmologies. Finally, we conjecture that chaos can appear for at least four interacting fluids.

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

Similar content being viewed by others

Notes

  1. If one interprets the cosmological constant as the curvature associated to vacuum.

  2. This point doesn’t exclude the possibility of negative values for \(\Lambda \) in the Friedmann–Lemaître equations, but it doesn’t allow changes for \(\Lambda \)’s signum in a given Friedmann–Lemaître universe.

  3. Although both are pressureless with \(\omega =0\), we split both to allow for different couplings.

  4. This is so since gravity is still minimally coupled to matter fluids.

References

  1. Wainwright, J., Ellis, G.F.R. (eds.): Dynamical Systems in Cosmology. Cambridge University Press. Cambridge, MA (1997)

  2. Belinskii, V.A., Khalatnikov, L.M., Lifshitz, E.M.: Oscillatory approach to a singular point in relativistic cosmology. Adv. Phys. 19, 525 (1970)

    Article  ADS  Google Scholar 

  3. Misner, C.W.: Mixmaster universe. Phys. Rev. Lett. 22, 1071 (1969)

    Article  ADS  MATH  Google Scholar 

  4. Wainwright, J., Hsu, L.: A dynamical systems approach to Bianchi cosmologies: orthogonal models of class A. Class. Quantum Gravity 6, 1409 (1989)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  5. Ringström, H.: The Bianchi IX attractor. Ann. Inst. H. Poincaré 2, 405 (2001)

    Article  MATH  Google Scholar 

  6. Damour, T., Henneaux, M., Nicolai, H.: Cosmological billiards. Class. Quantum Gravity 20, R145 (2003)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  7. Caldwell, R.R., Kamionkowski, M., Weinberg, N.N.: Phantom energy and cosmic doomsday. Phys. Rev. Let. 91, 7 (2003)

    Article  ADS  Google Scholar 

  8. Amendola, L., Tsujikawa, S.: Dark Energy: Theory and Observations. Cambridge University Press, Cambridge, MA (2010)

    Book  Google Scholar 

  9. Copeland, E.J., Liddle, A., Wands, D.: Exponential potentials and cosmological scaling solutions. Phys. Rev. D 57, 4686–4690 (1998)

    Article  ADS  Google Scholar 

  10. de la Macorra, A., Piccinelli, G.: General scalar fields as quintessence. Phys. Rev. D 61, 123503 (2000)

    Article  ADS  Google Scholar 

  11. Ng, S.C.C., Nunes, N.J., Rosati, F.: Applications of scalar attractor solutions to cosmology. Phys. Rev. D 64, 083510 (2001)

    Article  ADS  Google Scholar 

  12. Chimento, L.P., Jakubi, A.S., Pavon, D.: Enlarged quintessence cosmology. Phys. Rev. D 62, 063508 (2000)

    Article  ADS  Google Scholar 

  13. Copeland, E.J., Sami, M., Tsujikawa, S.: Dynamics of dark energy. Int. J. Mod. Phys. D 15, 1753 (2006). arXiv:hep-th/0603057

  14. Tsujikawa, S.: Dark energy investigation and modeling. In: Dark Matter and Dark Energy Astrophysics and Space Science Library, vol. 370, pp. 331–402 (2011) arXiv:1004.1493

  15. Kremer, G.M., Sobreiro, O.A.S: Bulk viscous cosmological model with interacting dark fluids, Braz. J. Phys. 42, 77 (2012) arXiv:1109.5068 [gr-qc]

  16. Lip, S.Z.W.: Interacting cosmological fluid and the coincidence problem. Phys. Rev. D 83, 023528 (2011)

    Article  ADS  Google Scholar 

  17. Arévalo, F., Bacalhau, A.P., Zimdahl, W.: Cosmological dynamics with non-linear interactions. Class. Quantum Gravity 29, 235001 (2012)

    Article  Google Scholar 

  18. Zimdahl, W., Pavon, D.: Scaling cosmology. Gen. Relativ. Gravit 35, 413 (2003)

    Article  ADS  MATH  Google Scholar 

  19. Lotka, A.J.: Contribution to the theory of periodc reaction. J. Phys. Chem 14(3), 271–274 (1910)

    Article  Google Scholar 

  20. Volterra, V.: Variazioni e fluttuazioni del numero d’individui in specie animali conviventi. Mem. Acad. Lincei Roma 2, 31–113 (1926)

    Google Scholar 

  21. Goel, N.S., et al.: On the Volterra and Other Non-Linear Models of Interacting Populations. Academic Press, New York (1971)

    Google Scholar 

  22. Uzan, J.-P., Lehoucq, R.: A dynamical study of the Friedmann equations. Eur. J. Phys. 22, 371–384 (2001)

    Article  MATH  Google Scholar 

  23. Hobson, M.P., Efstathiou, G.P., Lasenby, A.N.: General Relativity: An Introduction for Physicists. Cambridge University Press, Cambridge, MA (2006)

    Book  Google Scholar 

  24. Hofbauer, J., Sigmund, K.: Evolutionary Games and Population Dynamics. Cambridge University Press, Cambridge, MA (1998)

    Book  MATH  Google Scholar 

  25. Murray, J.D.: Mathematical Biology I: An Introduction, 3rd edn. Springer, Berlin (2002)

    Google Scholar 

  26. Martin, J.: Inflationary cosmological perturbations of quantum-mechanical origin. In: Lecture Notes Physics, vol. 669, pp. 199–244 (2005)

  27. Caldera-Cabra, G., Maartens, R.: Arturo Urena-Lopez, L.: Dynamics of interacting dark energy. Phys. Rev. D 79, 063518 (2009)

    Article  ADS  Google Scholar 

  28. Wetterich, C.: An asymptotically vanishing time-dependent cosmological “constant”. Astron. Astrophys. 301, 321 (1995)

    ADS  Google Scholar 

  29. Amendola, L.: Coupled quintessence. Phys. Rev. D 62, 043511 (2000)

    Article  ADS  Google Scholar 

  30. Cen, R.: Decaying cold dark matter model and small-scale power. Astrophys. J. Lett 546(L77), 2001 (2000)

    Google Scholar 

  31. Uzan, J.-P.: Dynamics of relativistic interacting gases: from a kinetic to a fluid description. Class. Quantum Gravity 15, 1063 (1998)

    Article  ADS  MATH  MathSciNet  Google Scholar 

  32. Wang, R., Xiao, D.: Bifurcations and chaotic dynamics in a 4-dimensional competitive Lotka–Volterra system. Nonlinear Dyn. 59, 411–422 (2010)

    Article  MATH  MathSciNet  Google Scholar 

  33. May, R.M., Leonard, W.J.: Non linear aspects of competition between three species. SIAM J. Appl. Math. 29(2), 243–253 (1975)

    Article  MATH  MathSciNet  Google Scholar 

Download references

Acknowledgments

J.P. thanks Frédéric Jean for helpful discussions. A.F. and L.G. are supported by ARC Convention No. 11/15-040. This paper presents research results of the Belgian Network DYSCO (Dynamical Systems, Control, and Optimization), funded by the Interuniversity Attraction Poles Programme, initiated by the Belgian State, Science Policy Office. The scientific responsibility rests with its authors.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to André Füzfa.

Appendices

Appendix 1: The Lotka–Volterra equation

The Lotka–Volterra equation is defined by a set of two coupled ordinary differential equation of the form

$$\begin{aligned} \left\{ \begin{array}{c} x_{1}^{\prime }=x_{1}\left( +r_{1}-a_{12}x_{2}\right) \\ x_{2}^{\prime }=x_{2}\left( -r_{2}+a_{21}x_{1}\right) \end{array} \right. \end{aligned}$$

where \(x_{1}\left( t\right) \) and \(x_{2}\left( t\right) \) are two function of the variable \(t\), a prime indicate the \(t\)-derivative and \(r_{1},\,r_{2},\,a_{12}\) and \(a_{21}\) are four positive constants. It is one of the simplest model to describe interactive dynamical population between some prey (which number is \(x_{1}\)) and its predator (which number is \(x_{2}\)). The constant have intuitive interpretation:

  • \(r_{1}\) (resp. \(r_{2}\)) is the growth rate for preys (resp. predators) to increase (resp. decrease) when there is no predators (resp. preys);

  • \(a_{12}\) and \(a_{21}\) are the coupling factors between the two populations; in ecological systems they are related to the mobility, the agressivity and other natural properties of the different concerned populations.

The properties of such a set of equation are very well known and a lot of classical books (see [24] and reference within) explain and demonstrate that starting with strictly positive initial conditions, the solution \(x_{1}\left( t\right) \) and \(x_{2}\left( t\right) \) are periodic functions. The orbits in the phase space are concentric closed curves centered on the equilibrium \( \left( \tilde{x}_{1},\tilde{x}_{2}\right) =\left( \frac{r_{1}}{a_{12}},\frac{ r_{2}}{a_{21}}\right) \). These curves correspond to the isocontours of the Ljapunov function \(V\left( x_{1},x_{2}\right) =a_{21}x+a_{12}y- \left( r_{2}\ln x+r_{1}\ln y\right) \).

When \(r_{1},\,r_{2},\,a_{12}\) and \(a_{21}\) have different signs the situation is changed and the system can describe other situations like competition, symbiose, etc (see [24]).

Lotka–Volterra systems is often considered as an unrealistic model in the context of population dynamics because there is no limitation when one specie is ruled out. A better model could be implemented introducing a logistic limitation such that

$$\begin{aligned} \left\{ \begin{array}{c} x_{1}^{\prime }=x_{1}\left( +r_{1}-a_{11}x_{1}-a_{12}x_{2}\right) \\ x_{2}^{\prime }=x_{2}\left( -r_{2}+a_{21}x_{1}-a_{22}x_{2}\right) \end{array} \right. \end{aligned}$$

The sign of the new parameters \(a_{11}\) and \(a_{22}\) is always related to the physical nature of the dynamical system. When all constants are positives, the system admits a general treatment: the Lotka–Volterra curves are generally replaced by spirals which converge to the equilibrium \(\left( x^{*}\ne 0,y^{*}\ne 0\right) \) solution of the system

$$\begin{aligned} \left\{ \begin{array}{l} r_{1}=+a_{11}x_{1}+a_{12}x_{2} \\ r_{2}=-a_{21}x_{1}+a_{22}x_{2} \end{array} \right. \end{aligned}$$

When the matrix \(A=\left( a_{ij}\right) \) is non invertible \(\left( \text { singular}\right) \ \)the situation is called degenerated and equilibria lie on the axes.

Such models could be generalized to \(n\)-dimensional systems introducing the generalized Lotka–Volterra equation occuring in this paper. Less things can be said about dynamical properties of such systems when the coefficient of \(\mathbf {r}\) and \(A\) are generic ones, in particular the dynamics can change strongly giving rise to cyclic competition or cooperation [33].

Appendix 2: Counter example: a linear center which is actually nonlinearly unstable

Consider the dynamical system

$$\begin{aligned} \left\{ \begin{array}{l} \dot{x}=-y+x^{3}\\ \dot{y}=x+y^{3} \end{array} \right. \end{aligned}$$

The origin is an equilibrium, the eigenvalues of the jacobian near this equilibrium are \(\pm \) i. In the linear approximation the origin seems to be a center. But as the eigenvalues have no real part, the system is not hyperbolic and almost nothing can be assumed about the non linear dynamics considering only the linear one around the equilibrium. In this particular case one can prove that the origin is a repulsive focus, and it is then actually unstable. As a matter of fact, considering the intersection \(M\) between an orbit and the circle \(x^{2}+y^{2}=R^{2}\). The angle \(\alpha \) between the tangent in \(M\) to the orbit and the tangent in \(M\) to the circle is given for any radius \(R\) by the expression

$$\begin{aligned} \cos \alpha =\left( \dot{x},\dot{y}\right) \cdot \left( 2x,2y\right) =\left( -y+x^{3},x+y^{3}\right) \cdot \left( 2x,2y\right) =2x^{4}+2y^{4}>0 \end{aligned}$$

Hence, each orbit is going out from any circle of radius \(R>0\) and the origin is unstable.

Appendix 3: Proof of the stability of the \(\Omega _{\mathrm{d}}-\Omega _{\mathrm{e}}\) plane

The aim of this section is to provide a simple proof of the attractiveness of the \(\Omega _{\mathrm{d}}-\Omega _{\mathrm{e}}\) plane for all orbits whose initial conditions belong to the hyper-tetrahedron:

$$\begin{aligned} T_{4}=\left\{ \Omega _{\mathrm{d}}>0\right\} \cup \left\{ \Omega _{\mathrm{e}}>0\right\} \cup \left\{ \Omega _{\mathrm{r}}>0\right\} \cup \left\{ \Omega _{\mathrm{b}}>0\right\} \cup \left\{ \Omega _{\mathrm{d}}+\Omega _{\mathrm{e}}+\Omega _{\mathrm{r}}+\Omega _{\mathrm{b} }<1\right\} . \end{aligned}$$

Let us recall the ODE system describing the equations of motion:

$$\begin{aligned} \left\{ \begin{array}{l} \Omega _{\mathrm{d}}^{\prime }=\Omega _{\mathrm{d}}\left[ \Omega _{\mathrm{d} }+(\varepsilon +1+3\omega _{\mathrm{e}})\Omega _{\mathrm{e}}+2\Omega _{\mathrm{r} }+\Omega _{\mathrm{b}}-1\right] \\ \Omega _{\mathrm{e}}^{\prime }=\Omega _{\mathrm{e}}\left[ (1-\varepsilon )\Omega _{\mathrm{d}}+\left( 1+3\omega _{\mathrm{e}}\right) \Omega _{\mathrm{e}}+2\Omega _{\mathrm{r}}+\Omega _{\mathrm{b}}-1-3\omega _{\mathrm{e} }\right] \\ \Omega _{\mathrm{r}}^{\prime }=\Omega _{\mathrm{r}}\left[ \Omega _{\mathrm{d} }+\left( 1+3\omega _{\mathrm{e}}\right) \Omega _{\mathrm{e}}+2\Omega _{\mathrm{r}}+\Omega _{\mathrm{b}}-2\right] \\ \Omega _{\mathrm{b}}^{\prime }=\Omega _{\mathrm{b}}\left[ \Omega _{\mathrm{d} }+\left( 1+3\omega _{\mathrm{e}}\right) \Omega _{\mathrm{e}}+2\Omega _{\mathrm{r}}+\Omega _{\mathrm{b}}-1\right] \end{array} \right. \end{aligned}$$
(21)

To prove our claim we need to prove first the invariance with respect to the flow (21) of the hyper-tetrahedron \(T_{4}\). The invariance of each coordinates hyperplanes is trivial and follows straightforwardly from (21). For instance any solution such that \(\Omega _{\mathrm{d} }(0)=0\) will have \(\Omega _{\mathrm{d}}(\lambda )=0\) for all \(\lambda \), then using the uniqueness of the Cauchy problem we can ensure that any solution with \(\Omega _{\mathrm{d}}(0)>0\) will never cross the hyperplane \(\Omega _{\mathrm{d}}=0\). A very similar analysis can be performed for the remaining cases.

Let us now consider the remaining piece of the boundary of \(T_{4}\), that is the hyperplane\(\{\Omega _{\mathrm{d}}+\Omega _{\mathrm{e}}+\Omega _{\mathrm{r} }+\Omega _{\mathrm{b}}=1\}\). A straightforward computation gives:

$$\begin{aligned} \left( \Omega _{\mathrm{d}}\!+\!\Omega _{\mathrm{e}}\!+\!\Omega _{\mathrm{r}} \!+\!\Omega _{\mathrm{b}}\right) ^{\prime }=\left[ \Omega _{\mathrm{d}} \!+\!(1\!+\!3\omega _{\mathrm{e}})\Omega _{\mathrm{e}}\!+\!2\Omega _{\mathrm{r}} \!+\!\Omega _{\mathrm{b}}\right] \left[ \Omega _{\mathrm{d}}\!+\!\Omega _{\mathrm{e} }\!+\!\Omega _{\mathrm{r}}\!+\!\Omega _{\mathrm{b}}-1\right] , \end{aligned}$$

thus any solution with initial conditions

$$\begin{aligned} \Omega _{\mathrm{d}}(0)+\Omega _{\mathrm{e}}(0)+\Omega _{\mathrm{r}} (0)+\Omega _{\mathrm{b}}(0)=1, \end{aligned}$$

will always satisfies the constraint

$$\begin{aligned} \Omega _{\mathrm{d}}(\lambda )+\Omega _{\mathrm{e}}(\lambda )+\Omega _{\mathrm{r} }(\lambda )+\Omega _{\mathrm{b}}(\lambda )=1\quad \forall \lambda . \end{aligned}$$

Thus once again the uniqueness result of the Cauchy problem implies that any solution such that \(\Omega _{\mathrm{d}}(0)+\Omega _{\mathrm{e}}(0)+\Omega _{\mathrm{r}}(0)+\Omega _{\mathrm{b}}(0)<1\), will never reach the hyperplane \(\Omega _{\mathrm{d}}+\Omega _{\mathrm{e}}+\Omega _{\mathrm{r}}+\Omega _{\mathrm{b}}=1\).

Finally putting together the above partial results, we can conclude that any orbit with initial condition inside \(T_{4}\) will never leave it.

A by-product of the invariance of the tetrahedron is that orbits inside \(T_{4}\) will always have positive projections on the axes. This allows us to compute the distance from the plane \(\left( \Omega _{\mathrm{d}},\Omega _{\mathrm{e}}\right) \) using the linear function \(F(\Omega _{\mathrm{r}},\Omega _{\mathrm{b}})=\Omega _{\mathrm{r}}+\Omega _{\mathrm{b}}\), which is zero if and only if \(\Omega _{\mathrm{r}}=\Omega _{\mathrm{b}}=0\), that is the point belongs to the plane \(\left( \Omega _{\mathrm{d}},\Omega _{\mathrm{e}}\right) \).

We can then compute the Lie derivative of \(F\) and prove that its restriction to \(T_{4}\) is strictly negative, hence \(F(\Omega _{\mathrm{r}}(\lambda ),\Omega _{\mathrm{b}}(\lambda ))\rightarrow 0\) for \(\lambda \rightarrow +\infty \) and because of the positiveness of \(\Omega _{\mathrm{r}}\left( \lambda \right) \) and \(\Omega _{\mathrm{b}}\left( \lambda \right) \) we can conclude that both \(\Omega _{\mathrm{r}}\left( \lambda \right) \) and \(\Omega _{\mathrm{b}}\left( \lambda \right) \) goes asymptotically to zero.

To prove the latter claim let us compute the derivative of \(F\) along the flow of (21):

$$\begin{aligned} \left. \frac{dF}{dt}\right| _{\mathrm{flow}}=\left[ \Omega _{\mathrm{d} }+(1+3\omega _{\mathrm{e}})\Omega _{\mathrm{e}}+2\Omega _{\mathrm{r}} +\Omega _{\mathrm{b}}-1\right] \left[ \Omega _{\mathrm{r}}+\Omega _{\mathrm{b} }\right] -\Omega _{\mathrm{r}}, \end{aligned}$$

because of our previous result \(\Omega _{\mathrm{d}}(\lambda )+\Omega _{\mathrm{b}}(\lambda )-1<-\Omega _{\mathrm{e}}(\lambda )-\Omega _{\mathrm{r} }(\lambda )\) for all \(\lambda \). Using this inequality we get

$$\begin{aligned} \left. \frac{dF}{dt}\right| _{\mathrm{flow}}<\left( 3\omega _{\mathrm{e} }\Omega _{\mathrm{e}}\!+\!\Omega _{\mathrm{r}}\right) \left( \Omega _{\mathrm{r} }\!+\!\Omega _{\mathrm{b}}\right) -\Omega _{\mathrm{r}}\,=3\omega _{\mathrm{e} }\Omega _{\mathrm{e}}\left( \Omega _{\mathrm{r}}\!+\!\Omega _{\mathrm{b}}\right) \!+\!\left( \Omega _{\mathrm{r}}\!+\!\Omega _{\mathrm{b}}-1\,\right) \Omega _{\mathrm{r}} \end{aligned}$$

let us observe that the right hand side is strictly negative, in fact

$$\begin{aligned} 3\omega _{\mathrm{e}}\Omega _{\mathrm{e}}(\Omega _{\mathrm{r}}+\Omega _{\mathrm{b}})<0 \end{aligned}$$

provided and \(\Omega _{\mathrm{e}}\) is associated to the dark energy \(\left( \omega _{\mathrm{e}}<-\frac{1}{3}<0\right) \) and

$$\begin{aligned} \Omega _{\mathrm{r}}+\Omega _{\mathrm{b}}-1<\Omega _{\mathrm{d}}+\Omega _{\mathrm{e}}+\Omega _{\mathrm{r}}+\Omega _{\mathrm{b}}-1<0. \end{aligned}$$

This concludes our proof of the attractiveness of the dark plane for all orbits whose initial conditions belong to \(T_{4}\).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Perez, J., Füzfa, A., Carletti, T. et al. The Jungle Universe: coupled cosmological models in a Lotka–Volterra framework. Gen Relativ Gravit 46, 1753 (2014). https://doi.org/10.1007/s10714-014-1753-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10714-014-1753-8

Keywords

Navigation