Abstract
We investigate systems of self-propelled particles with alignment interaction. Compared to previous work (Degond and Motsch, Math. Models Methods Appl. Sci. 18:1193–1215, 2008a; Frouvelle, Math. Models Methods Appl. Sci., 2012), the force acting on the particles is not normalized, and this modification gives rise to phase transitions from disordered states at low density to aligned states at high densities. This model is the space-inhomogeneous extension of (Frouvelle and Liu, Dynamics in a kinetic model of oriented particles with phase transition, 2012), in which the existence and stability of the equilibrium states were investigated. When the density is lower than a threshold value, the dynamics is described by a nonlinear diffusion equation. By contrast, when the density is larger than this threshold value, the dynamics is described by a similar hydrodynamic model for self-alignment interactions as derived in (Degond and Motsch, Math. Models Methods Appl. Sci. 18:1193–1215, 2008a; Frouvelle, Math. Models Methods Appl. Sci., 2012). However, the modified normalization of the force gives rise to different convection speeds, and the resulting model may lose its hyperbolicity in some regions of the state space.
Similar content being viewed by others
References
Aldana, M., Huepe, C.: Phase transitions in self-driven many-particle systems and related non-equilibrium models: a network approach. J. Stat. Phys. 112, 135–153 (2003)
Aoki, I.: A simulation study on the schooling mechanism in fish. Bull. Jpn. Soc. Sci. Fish. 48, 1081–1088 (1982)
Bailey, P.B., Everitt, W.N., Zettl, A.: Computing eigenvalues of singular Sturm–Liouville problems. Results Math. 20, 391–423 (1991)
Bolley, F., Cañizo, J.A., Carrillo, J.A.: Mean-field limit for the stochastic Vicsek model. Appl. Math. Lett. 25, 339–343 (2012)
Carrillo, J.A., D’Orsogna, M.R., Panferov, V.: Double milling in self-propelled swarms from kinetic theory. Kinet. Relat. Models 2, 363–378 (2009)
Carrillo, J.A., Klar, A., Martin, S., Tiwari, S.: Self-propelled interacting particle systems with roosting force. Math. Models Methods Appl. Sci. 20, 1533–1552 (2010)
Chuang, Y.-L., D’Orsogna, M.R., Marthaler, D., Bertozzi, A.L., Chayes, L.S.: State transitions and the continuum limit for a 2D interacting, self-propelled particle system. Physica D 232, 33–47 (2007)
Couzin, I.D., Krause, J., James, R., Ruxton, G.D., Franks, N.R.: Collective memory and spatial sorting in animal groups. J. Theor. Biol. 218, 1–11 (2002)
Cucker, F., Smale, S.: Emergent behavior in flocks. IEEE Trans. Autom. Control 52, 852–862 (2007)
Degond, P.: Macroscopic limits of the Boltzmann equation: a review. In: Degond, P., Pareschi, L., Russo, G. (eds.) Modeling and Computational Methods for Kinetic Equations. Modeling and Simulation in Science, Engineering and Technology, pp. 3–57. Birkhäuser, Basel (2004)
Degond, P., Frouvelle, A., Liu, J.-G.: Phase transitions, hysteresis and hyperbolicity for self-organized alignment dynamics (2012, in preparation)
Degond, P., Liu, J.-G., Motsch, S., Panferov, V.: Hydrodynamic models of self-organized dynamics: derivation and existence theory (2011, submitted)
Degond, P., Motsch, S.: Continuum limit of self-driven particles with orientation interaction. Math. Models Methods Appl. Sci. 18, 1193–1215 (2008a)
Degond, P., Motsch, S.: Large scale dynamics of the persistent turning walker model. J. Stat. Phys. 131, 989–1021 (2008b)
Degond, P., Motsch, S.: A macroscopic model for a system of swarming agents using curvature control. J. Stat. Phys. 143, 685–714 (2011)
Degond, P., Yang, T.: Diffusion in a continuum model of self-propelled particles with alignment interaction. Math. Models Methods Appl. Sci. 20, 1459–1490 (2010)
Doi, M., Edwards, S.F.: the Theory of Polymer Dynamics. International Series of Monographs on Physics. Oxford University Press, London (1999)
D’Orsogna, M.R., Chuang, Y.-L., Bertozzi, A.L., Chayes, L.S.: Self-propelled particles with soft-core interactions: patterns, stability, and collapse. Phys. Rev. Lett. 96, 104302 (2006)
Frouvelle, A.: A continuum model for alignment of self-propelled particles with anisotropy and density-dependent parameters. Math. Mod. Meth. Appl. Sci. 22, 1250011 (2012)
Frouvelle, A., Liu, J.-G.: Dynamics in a kinetic model of oriented particles with phase transition. SIAM J. Math. Anal. 44, 791–826 (2012)
Grégoire, G., Chaté, H.: Onset of collective and cohesive motion. Phys. Rev. Lett. 92, 025702 (2004)
Hsu, E.P.: Stochastic Analysis on Manifolds. Graduate Series in Mathematics. Am. Math. Soc., Providence (2002)
Kulinskii, V.L., Ratushnaya, V.I., Zvelindovsky, A.V., Bedeaux, D.: Hydrodynamic model for a system of self-propelling particles with conservative kinematic constraints. Europhys. Lett. 71, 207 (2005)
Maier, W., Saupe, A.: Eine einfache molekulare Theorie des nematischen kristallinflüssigen Zustandes. Z. Naturforsch. 13, 564–566 (1958)
Mogilner, A., Edelstein-Keshet, L.: A non-local model for a swarm. J. Math. Biol. 38, 534–570 (1999)
Mogilner, A., Edelstein-Keshet, L., Bent, L., Spiros, A.: Mutual interactions, potentials, and individual distance in a social aggregation. J. Math. Biol. 47, 353–389 (2003)
Onsager, L.: The effects of shape on the interaction of colloidal particles. Ann. N.Y. Acad. Sci. 51, 627–659 (1949)
Pego, R.L., Serre, D.: Instabilities in Glimm’s scheme for two systems of mixed type. SIAM J. Numer. Anal. 25, 965–988 (1988)
Ratushnaya, V.I., Bedeaux, D., Kulinskii, V.L., Zvelindovsky, A.V.: Collective behavior of self-propelling particles with kinematic constraints: the relation between the discrete and the continuous description. Physica A 381, 39–46 (2007)
Ratushnaya, V.I., Kulinskii, V.L., Zvelindovsky, A.V., Bedeaux, D.: Hydrodynamic model for the system of self propelling particles with conservative kinematic constraints; two dimensional stationary solutions. Physica A 366, 107–114 (2006)
Toner, J., Tu, Y.: Flocks, herds, and schools: a quantitative theory of flocking. Phys. Rev. E 58, 4828–4858 (1998)
Topaz, C.M., Bertozzi, A.L.: Swarming patterns in a two-dimensional kinematic model for biological groups. SIAM J. Appl. Math. 65, 152–174 (2004)
Topaz, C.M., Bertozzi, A.L., Lewis, M.E.: A nonlocal continuum model for biological aggregations. Bull. Math. Biol. 68, 1601–1623 (2006)
Vicsek, T., Czirók, A., Ben-Jacob, E., Cohen, I., Shochet, O.: Novel type of phase transition in a system of self-driven particles. Phys. Rev. Lett. 75, 1226–1229 (1995)
Watson, G.S.: Distributions on the circle and sphere. J. Appl. Probab. 19, 265–280 (1982). Essays in Statistical Science
Weidmann, J.: Spectral Theory of Ordinary Differential Operators. Lecture Notes in Mathematics. Springer, Berlin (1987)
Acknowledgements
The authors wish to acknowledge the hospitality of the Mathematical Sciences Center and Mathematics Department of Tsinghua University, where this research was completed. The research of J.-G.L. was partially supported by NSF grant DMS 10-11738. A.F. wishes to acknowledge partial support from the FP7-REGPOT-2009-1 project “Archimedes Center for Modeling, Analysis and Computation.”
Author information
Authors and Affiliations
Corresponding author
Additional information
Communicated by Eva Kanso.
Appendices
Appendix A: Poincaré Constant
In this appendix, we prove the following.
Proposition A.1
We have the following Poincaré inequality, for \(\psi\in H^{1}(\mathbb{S})\):
The best constant Λ κ in this inequality is the smallest positive eigenvalue of the operator
We define the linear operator \(L_{\kappa}^{*}\) by
Then one of the following three possibilities is true:
-
(i)
Λ κ is the smallest eigenvalue of the Sturm–Liouville problem
(A.4)for g∈C 2([0,π]) with Neumann boundary conditions (g′(0)=g′(π)=0) and such that \(\int_{0}^{\pi}(\sin\theta)^{n-2} e^{\kappa\cos\theta }g(\theta)\,\mathrm{d} \theta=0\), and the eigenspace of \(L_{\kappa\varOmega }^{*}\) associated to the eigenvalue Λ κ is of dimension 1, spanned by \(\omega\mapsto h^{0}_{\kappa}(\omega\cdot\varOmega )\), where the function θ↦h 0(cosθ) is smooth, positive for 0⩽θ<θ 0, and negative for θ 0<θ⩽π.
-
(ii)
Λ κ is the smallest eigenvalue of the Sturm–Liouville problem
(A.5)for g∈C 2([0,π]) with Dirichlet boundary conditions (g(0)=g(π)=0), and the eigenspace of \(L_{\kappa\varOmega}^{*}\) associated to Λ κ is of dimension n−1, consisting in the functions ψ A of the form \(\psi_{A}(\omega)=h_{\kappa}^{1}(\omega\cdot \varOmega)A\cdot\omega\) for any vector A∈ℝn such that Ω⋅A=0, with \(\theta\mapsto h_{\kappa}^{1}(\cos\theta)\) a smooth positive function for 0<θ<π.
-
(iii)
The two preceding Sturm–Liouville problems have the same smallest eigenvalue Λ κ , and the eigenspace of \(L_{\kappa \varOmega}^{*}\) associated to Λ κ is of dimension n, spanned by the two types of function of the above cases.
Proof
First of all, we have
and
The second inequality of (A.6) follows from the Poincaré inequality on the sphere:
The first inequality of (A.7) follows from the fact that
Equations (A.6) and (A.7) lead to the Poincaré inequality (A.1) with
We use the inner product \((\varphi,\psi)\mapsto\langle\varphi\psi\rangle _{M_{\kappa\varOmega}}\), adapted to M κΩ . We denote by \(\dot{L}^{2}_{\kappa}(\mathbb{S})\) (resp. \(\dot{H}^{1}_{\kappa}(\mathbb{S})\)) the functions \(\psi\in L^{2}(\mathbb{S})\) (resp. in \(H^{1}(\mathbb{S})\)) such that \(\langle\psi\rangle_{M_{\kappa \varOmega}}=0\).
The operator \(L_{\kappa\varOmega}^{*}\) given by (A.2) is self-adjoint since \(\langle\nabla_{\omega}\psi\cdot\nabla_{\omega}\varphi \rangle_{M_{\kappa\varOmega}}=\langle\psi L_{\kappa\varOmega}^{*}\varphi \rangle_{M_{\kappa\varOmega}}\). It is then easy to see, using the Lax-Milgram theorem, that if φ belongs to \(\dot{L}^{2}_{\kappa}(\mathbb{S})\), then there is a unique solution \(\psi\in\dot{H}^{1}_{\kappa}(\mathbb{S})\) to the equation \(L_{\kappa\varOmega}^{*}\psi=\varphi\). The so-obtained inverse operator is then compact and self-adjoint. By the spectral theorem, we get a basis of eigenfunctions, in the Hilbert space \(\dot{L}^{2}_{\kappa}(\mathbb{S})\), which are also eigenfunctions of \(L_{\kappa\varOmega}^{*}\). If we denote \(\varLambda_{\kappa}^{-1}\) the largest eigenvalue of the inverse of \(L_{\kappa\varOmega}^{*}\), then it is easy to see that Λ κ is the best constant for the following Poincaré inequality, in the space \(\dot{H}^{1}_{\kappa}(\mathbb{S})\):
Since the constants trivially satisfy this inequality, this shows that Λ κ is the best constant for the Poincaré inequality (A.1) in \(H^{1}(\mathbb{S})\).
The goal is now to reduce the computation of the eigenvalues to simpler problems, using separation of variables. We write ω=cosθΩ+sinθ v, where v belongs to the unit sphere, orthogonal to Ω. We identify Ω with the last element of an orthogonal basis of ℝn, and we write \(v\in \mathbb{S}_{n-2}\).
By spherical harmonic decomposition in an adapted basis (see for example Frouvelle and Liu 2012, Appendix A), we have a unique decomposition of the form
where is a given orthonormal basis of the spherical harmonics of degree m on \(\mathbb {S}_{n-2}\), for m∈ℕ, with \(k_{m}=\tbinom{n+m-2}{n-2}- \tbinom {n+m-4}{n-2}\). If ψ is continuous, \(g_{m}^{k}\) is given by
We now show that the decomposition (A.8) remains stable under the action of the operator L κΩ , so that its spectral decomposition can be performed independently for each term of the decomposition.
First, we examine the case of dimension n⩾3. Let ψ(ω)=g(θ)Z(v). We have
where the unit vector e θ is given by
We take functions \(\psi(\omega)= g_{m}^{k}(\theta)Z_{m}^{k}(v)\) and \(\varphi (\omega)=\sum_{k,m} f_{m}^{k}(\theta)Z_{m}^{k}(v)\). Since the spherical harmonics are orthonormal, and are eigenfunctions of Δ v for the eigenvalues −m(m+n−3), we get:
Suppose m⩾1. Then, it is easy to see that the function ψ belongs to \(\dot{H}^{1}_{\kappa}(\mathbb{S})\) if and only if \((\sin\theta )^{\frac{n}{2}-1} g'\in L^{2}(0,\pi)\) and \((\sin\theta)^{\frac{n}{2}-2} g'\in L^{2}(0,\pi)\). This condition is equivalent to the fact that g∈V, where V is defined by (4.5), and which we denote by \(V_{\kappa}^{m}\) for convenience:
Suppose now that m=0. Then \(Z^{k}_{m}\) is a constant, and the condition \(\psi\in\dot{H}^{1}_{\kappa}(\mathbb{S})\) is equivalent to the first condition only: \((\sin\theta)^{\frac{n}{2}-1} g'\in L^{2}(0,\pi)\), under the constraint that \(\int_{0}^{\pi}(\sin\theta)^{n-2} e^{\kappa\cos \theta}g(\theta)\,\mathrm{d} \theta=0\). We will denote this space by \(V^{0}_{\kappa}\):
Formula (A.10) then suggests to define the operator \(L_{\kappa,m}^{*}:V_{\kappa}^{m}\to(V_{\kappa}^{m})^{*}\) by
From (A.10), it follows that, if we decompose \(\psi(\omega )=\sum_{k,m} g_{m}^{k}(\theta)Z_{m}^{k}(v)\), then
showing that \(L_{\kappa\varOmega}^{*}\) is block diagonal on each of these spaces \(V_{\kappa}^{m}\) (tensorized by the spherical harmonics of degree m on \({\mathbb{S}}_{n-2}\)). So we can perform the spectral decomposition of \(L_{\kappa\varOmega}^{*}\) by means of the spectral decomposition of each of the \(L_{\kappa,m}^{*}\). It is indeed easy to prove, using the Lax-Milgram theorem, that the operators \(L_{\kappa,m}^{*}\) have self-adjoint compact inverses for the dot product \((f,g)=\int_{0}^{\pi}fg(\sin\theta)^{n-2} e^{\kappa\cos\theta }\,\mathrm{d} \theta\). Therefore, the eigenfunctions and eigenvalues of \(L_{\kappa\varOmega}^{*}\) correspond to those of the operators \(L_{\kappa,m}^{*}\), for all m∈ℕ. If we denote by λ κ,m the smallest eigenvalue of \(L_{\kappa,m}^{*}\), we finally get
We notice that
but since all the \(V_{\kappa}^{m}\) are the same for m⩾1, and since
we get
Finally, Λ κ is the minimum between λ κ,0 and λ κ,1. The eigenfunctions for the operator \(L_{\kappa\varOmega}^{*}\) being smooth, this is also true for the operators \(L_{\kappa,m}^{*}\), by formula (A.9). So we can transform the definitions (A.11) by integration by parts.
Indeed, if g 0 is an eigenfunction (in \(V_{\kappa}^{0}\)) associated to \(L_{\kappa,0}^{*}\) and an eigenvalue λ, then g 0 is smooth and satisfies the Sturm–Liouville eigenvalue problem
Conversely, a smooth function with the condition \(\int_{0}^{\pi}(\sin\theta )^{n-2} e^{\kappa\cos\theta}g(\theta)\,\mathrm{d} \theta=0\) belongs to \(V_{\kappa}^{0}\). Actually, in dimension n⩾3, we do not need to impose the Neumann boundary conditions: they appear naturally, since we have
Therefore, by continuity at θ=0 and π, \(g_{0}'(0)=g_{0}'(\pi)=0\). Then, using classical Sturm–Liouville oscillation theory (see Weidmann 1987, for example), we find that the first eigenspace of \(L_{\kappa}^{*}\) is of dimension 1, spanned by a function g κ,0(θ), which is positive for 0⩽θ<θ 0 and negative for θ 0<θ⩽π.
Similarly, if g 1 is an eigenfunction (in \(V_{\kappa}^{1}\)) associated to \(L_{\kappa,1}^{*}\) and an eigenvalue λ, then g 1 is smooth, with g 1(0)=g 1(π)=0 and satisfies the Sturm–Liouville eigenvalue problem
And conversely, if a function with Dirichlet boundary conditions is in C 2([0,π]), then it belongs to \(V_{\kappa}^{1}\). Once again, if n⩾3, we do not need to impose the Dirichlet boundary conditions in the C 2([0,π]) framework, since we have
So, by continuity at θ=0 and π, g 1(0)=g 1(π)=0, and then a first order expansion shows that continuity holds, whatever the values of \(g_{0}'(\theta)\) at the endpoints are. Again, using classical Sturm–Liouville theory, we find that the first eigenspace of \(L_{\kappa}^{*}\) is of dimension 1, spanned by a function g κ,1(θ), which keeps the same sign on (0,π).
The case λ κ,0<λ κ,1 corresponds to case (i) of the proposition. Since a spherical harmonic of degree 0 on the sphere \(\mathbb{S}_{n-2}\) is a constant, introducing the function \(h_{\kappa}^{0}\) such that \(h_{\kappa}^{0}(\cos\theta)=g_{\kappa ,0}(\theta)\) allows us to state that the eigenspace of \(L_{\kappa \varOmega}^{*}\) associated to the lowest eigenvalue is spanned by \(\omega \mapsto h_{\kappa}^{0}(\omega\cdot\varOmega)\).
The case λ κ,0>λ κ,1 corresponds to case (ii) of the proposition. The spherical harmonics of degree 1 on the sphere \(\mathbb{S}_{n-2}\) are the functions of the form v↦A⋅v, with A⋅Ω=0. Introducing \(h_{\kappa}^{0}\) such that \(h_{\kappa}^{0}(\cos\theta) \sin\theta= g_{\kappa,0}(\theta)\) allows us to state that the eigenspace of \(L_{\kappa\varOmega}^{*}\) associated to the lowest eigenvalue is of dimension n−1, consisting of the functions of the form \(\omega\mapsto h_{\kappa}^{1}(\omega\cdot\varOmega) A\cdot\omega\), with A any vector in ℝn such that A⋅Ω=0.
Finally, the case λ κ,0=λ κ,1 corresponds to case (iii) of the proposition, and this ends the proof in the case of dimension n⩾3.
We now examine the special case of dimension n=2. We identify \(H^{1}(\mathbb{S})\) with the 2π-periodic functions in \(H^{1}_{loc}(\mathbb{R})\). So, Λ κ is the smallest eigenvalue of the periodic Sturm–Liouville problem
for functions g such that \(\int_{-\pi}^{\pi}e^{\kappa\cos\theta }g(\theta)\,\mathrm{d} \theta=0\). Here the decomposition corresponding to (A.8) is the even-odd decomposition (there are only two spherical harmonics on \(\mathbb{S}_{0}\): the constant function of degree 0 and the odd function of degree 1). The odd part g o of g can be identified with a function of \(H^{1}_{0}(0,\pi)\), and it is easy to see that the odd part of \(L_{\kappa}^{*}(g)\) is \(L_{\kappa}^{*}(g_{o})\), and similarly for the even part g e . So, we can perform the spectral decomposition of \(L^{*}_{\kappa}\) separately on the spaces of even and odd functions.
Actually, if g is a solution of the Sturm–Liouville periodic problem, the function \(\widetilde{g}\) defined by \(\widetilde{g}(\theta)=e^{-\kappa \cos\theta}\partial_{\theta}g(\pi-\theta)\) is another solution with the same eigenvalue. Furthermore, if g is odd, then \(\widetilde{g}\) is even and conversely. So the eigenvalues are the same for the odd and even spaces problems. Therefore, in dimension n=2, Proposition A.1 can be refined, and we can state that case (iii) is the only possibility: the eigenspace of \(L_{\kappa\varOmega}^{*}\) associated to Λ κ is of dimension 2, spanned by an odd function \(g_{\kappa}^{o}\), positive on (0,π), and an even function \(g_{\kappa}^{e}=\widetilde{g}_{\kappa}^{o}\), positive for 0<θ<θ 0 and negative for θ 0<θ<π. The proof of Proposition A.1 is complete. □
We can now state a conjecture, which refines Proposition A.1, if true, and which is based on numerical experiments.
Conjecture A.1
-
(i)
When κ>0 and n⩾3, only statement (ii) of Proposition A.1 is true.
-
(ii)
The function κ↦Λ κ is increasing.
We also observe numerically that λ 1∼κ when kappa is large.
Some investigations are in progress to prove the monotonicity of the eigenvalue with respect to κ, based on formal expansions similar to those used in Sect. 5 of Frouvelle (2012).
Remark A.1
At the end of the proof of Proposition A.1, we have seen that in dimension n=2 only statement (iii) is true. The proof uses a transformation of the solution of an eigenvalue problem into the solution of another eigenvalue problem. We can try to find a similar transformation in dimensions n⩾3: if f satisfies L κ,0 f=λf (with Neumann boundary conditions), then \(\widetilde{f}=e^{-\kappa\cos\theta} \partial_{\theta}f(\pi-\theta)\) (with Dirichlet boundary conditions) satisfies
so if we can prove that \(\int_{0}^{\pi}\cos\theta \widetilde{f}^{\,2} (\sin\theta)^{n-2} e^{\kappa\cos\theta} \,\mathrm{d} \theta>0\), we can deduce that λ 0>λ 1. So far we have been unable to prove this estimate.
Appendix B: Numerical Computations of the Coefficients
We adopt a finite difference approach to compute the function g κ associated to the GCIs and defined by (4.6). We consider the function f κ such that \(f_{\kappa}(\theta)=(\sin \theta)^{\frac{n}{2}-1} g_{\kappa}(\theta)\). In particular, since g κ ∈V defined by (4.5), f κ belongs to \(H^{1}_{0}(0,\pi)\). Since g κ satisfies (4.6), f κ satisfies
We discretize the interval (0,π) with N+1 points \(\theta_{i}=\frac{1}{N}i\pi\), and denote by \(f^{i}_{\kappa}\) an approximation of f κ at these points. Since \(f_{\kappa}\in H^{1}_{0}(0,\pi)\), we have \(f^{0}_{\kappa}=f^{N}_{\kappa}=0\). We define \(e_{\kappa}^{i}=e^{\kappa\cos\theta_{i}}\). A second order approximation of \((e^{\kappa\cos\theta}f_{\kappa}')'\) at θ i is then given by
Introducing
the vector is the solution of the linear system AF=S, where the vector S is , and the tridiagonal matrix A is defined by
We use the trapezoidal method to perform the integrations in the definitions (2.11) and (4.9) of c and \(\widetilde{c}\). The other coefficients ρ, λ, and θ c are then directly computed from c and \(\widetilde{c}\). The numerical results provided in Figs. 3 and 4 have been obtained for N=3000.
We now detail how we obtain an approximation of the Poincaré constant Λ κ . By Appendix A, Λ κ is the minimum between λ κ,1 and λ κ,0, which are the smallest eigenvalues of two Sturm–Liouville problems. Several algorithms exist to compute eigenvalues of singular Sturm–Liouville problems (which is the case here whenever n⩾3) with a good precision (Bailey et al. 1991). However, we use a simpler method based on finite differences.
Actually, λ κ,1 is the smallest eigenvalue associated to problem (A.5), with g∈V. So, considering once again the function f such that \(f(\theta)=(\sin\theta)^{\frac{n}{2}-1} g(\theta )\), the vector AF, with A defined by (B.1), gives a second order approximation of \((\sin\theta)^{\frac{n}{2}-1}\widetilde{L}_{\kappa}^{*}g(\theta)=\lambda f(\theta)\) at the points θ i . So we can take the smallest eigenvalue of A as an approximation of λ κ,1.
We now look for an approximation of λ κ,0.
Let g be a solution of the Sturm–Liouville problem (A.4) with Neumann boundary conditions. We introduce , the vector of approximations of g at the points \(\theta_{i+\frac{1}{2}}=\frac{1}{N}(i+\frac{1}{2})\pi\). Introducing \(m_{\kappa}^{i}=(\sin\theta)^{n-2} e^{\kappa\cos\theta_{i}}\), a second order approximation of \(L_{\kappa}^{*}g\) at the point \(\theta _{i+\frac{1}{2}}\), with i∈〚1,N−2〛, is then given by
With the Neumann boundary conditions, the approximations at the points \(\theta_{\frac{1}{2}}\) and \(\theta_{N-\frac{1}{2}}\) are given by
Introducing
a second order approximation of \(L_{\kappa}^{*}g\) is given by BG, where the tridiagonal matrix B is defined by
So we can take the smallest positive eigenvalue of B as an approximation of λ κ,0 (excluding the constant functions). The computations of Fig. 1 have been performed with N=300 points.
Rights and permissions
About this article
Cite this article
Degond, P., Frouvelle, A. & Liu, JG. Macroscopic Limits and Phase Transition in a System of Self-propelled Particles. J Nonlinear Sci 23, 427–456 (2013). https://doi.org/10.1007/s00332-012-9157-y
Received:
Accepted:
Published:
Issue Date:
DOI: https://doi.org/10.1007/s00332-012-9157-y
Keywords
- Self-propelled particles
- Alignment interaction
- Vicsek model
- Phase transition
- Hydrodynamic limit
- Nonhyperbolicity
- Diffusion limit
- Chapman–Enskog expansion