Skip to main content
Log in

Collective Dynamics in the Vicsek and Vectorial Network Models Beyond Uniform Additive Noise

  • Published:
Journal of Nonlinear Science Aims and scope Submit manuscript

An Erratum to this article was published on 01 February 2016

Abstract

In this work, we analyze the coordination of interacting individuals in two nonlinear dynamical models that are subject to a new form of noise. Specifically, we propose extensions both to the classical Vicsek model, whereby each individual averages the orientation of its geographically proximal neighbors, and to the vectorial network model, in which the selection of neighbors is random and independent of the group geometric configuration. In the traditional forms of these models, the update rule for the individuals’ orientations is affected by additive uniform noise. Motivated by biological groups in which individuals’ turn rates exhibit sporadic and large changes, we extend the uniform additive noise model to a turn rate stochastic process. Through comprehensive numerical simulations, we demonstrate the impact of such occasional large deviations (intensity and frequency), along with the role of the neighbors’ selection process, on the coordination of the group. In addition, we establish a closed-form expression for the group polarization for the vectorial network model in the vicinity of an ordered state.

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
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

Notes

  1. Here and in what follows, we omit the explicit dependence on the time step k for any increments to a process that arise from i.i.d. random variables.

  2. In the original formulation by Aldana et al. (2007), the neighbor set is constructed by randomly and uniformly selecting K individuals, so that the individual i does not necessarily belong to \(\mathcal {N}_i(k)\). In the work presented herein, individual i will always select itself as one of its K neighbors, and the other \(K-1\) individuals are chosen randomly, with equal probability, from the remaining \(N-1\) individuals. While this difference does not bear notable consequences for a uniform additive noise, it has a significant effect for a complex turn rate model, as further elaborated in what follows.

References

  • Abaid, N., Porfiri, M.: Consensus over numerosity-constrained random networks. IEEE Trans. Autom. Control. 56, 649–654 (2011)

    Article  MathSciNet  MATH  Google Scholar 

  • Abaid, N., Porfiri, M.: Leader-follower consensus over numerosity-constrained random networks. Automatica 48(8), 1845–1851 (2012)

    Article  MathSciNet  MATH  Google Scholar 

  • Abaid, N., Bollt, E., Porfiri, M.: Topological analysis of complexity in multiagent systems. Phys. Rev. E 85(4), 041907 (2012a)

    Article  Google Scholar 

  • Abaid, N., Igel, I., Porfiri, M.: On the consensus protocol of conspecific agents. Linear Algebra Appl. 437(1), 221–235 (2012b)

    Article  MathSciNet  MATH  Google Scholar 

  • Aldana, M., Dossetti, V., Huepe, C., Kenkre, V.M., Larralde, H.: Phase transitions in systems of self-propelled agents and related network models. Phys. Rev. Lett. 98(9), 95702 (2007)

    Article  Google Scholar 

  • Aldana, M., Huepe, C.: Phase transitions in self-driven many-particle systems and related non-equilibrium models: a network approach. J. Stat. Phys. 112(1–2), 135–153 (2003)

    Article  MATH  Google Scholar 

  • Ariel, G., Rimer, O., Ben-Jacob, E.: Order-disorder phase transition in heterogeneous populations of self-propelled particles. J. Stat. Phys. 1, 188 (2014)

    MathSciNet  MATH  Google Scholar 

  • Attanasi, A., Cavagna, A., Castello, L., Giardina, I., Grigera, T.S., Jelić, A., Melillo, S., Parisi, L., Pohl, O., Shen, E., et al.: Information transfer and behavioural inertia in starling flocks. Nat. Phys. 10(9), 691–696 (2014)

    Article  Google Scholar 

  • Ballerini, M., Cabibbo, N., Candelier, R., Cavagna, A., Cisbani, E., Giardina, I., Lecomte, V., Orlandi, A., Parisi, G., Procaccini, A., et al.: Interaction ruling animal collective behavior depends on topological rather than metric distance: evidence from a field study. Proc. Natl. Acad. Sci. 105(4), 1232–1237 (2008)

    Article  Google Scholar 

  • Bergman, C.M., Schaefer, J.A., Luttich, S.: Caribou movement as a correlated random walk. Oecologia 123(3), 364–374 (2000)

    Article  Google Scholar 

  • Bernstein, D.S.: Matrix Mathematics. Princeton University Press, Princeton (2009)

    Book  MATH  Google Scholar 

  • Borzì, A., Wongkaew, S.: Modeling and control through leadership of a refined flocking system. Math. Models Methods Appl. Sci. 1–28 (2014)

  • Burrow, J.F., Baxter, P.D., Pitchford, J.W.: Lévy processes, saltatory foraging, and superdiffusion. Math. Model Nat. Phenom. 3(3), 115–130 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  • Butail, S., Ladu, F., Spinello, D., Porfiri, M.: Information flow in animal-robot interactions. Entropy 16(3), 1315–1330 (2014)

    Article  Google Scholar 

  • Cachat, J., Stewart, A., Utterback, E., Hart, P., Gaikwad, S., Wong, K., Kyzar, E., Wu, N., Kalueff, A.V.: Three-dimensional neurophenotyping of adult zebrafish behavior. PLoS One 6(3), 17597 (2011)

    Article  Google Scholar 

  • Camazine, S.: Self-Organization in Biological Systems. Princeton University Press, Princeton (2003)

    MATH  Google Scholar 

  • Chaté, H., Ginelli, F., Grégoire, G., Raynaud, F.: Collective motion of self-propelled particles interacting without cohesion. Phys. Rev. E 77(4), 046113 (2008a)

    Article  Google Scholar 

  • Chaté, H., Ginelli, F., Grégoire, G., Peruani, F., Raynaud, F.: Modeling collective motion: variations on the Vicsek model. Eur. Phys. J. B 64(3–4), 451–456 (2008b)

    Article  Google Scholar 

  • Couzin, I.D., Krause, J.: Self-organization and collective behavior in vertebrates. Adv. Study Behav. 32(1), 1–75 (2003)

  • Czirók, A., Vicsek, T.: Collective behavior of interacting self-propelled particles. Phys. A 281(1), 17–29 (2000)

    Article  Google Scholar 

  • Degond, P., Motsch, S.: Large scale dynamics of the persistent turning walker model of fish behavior. J. Stat. Phys. 131(6), 989–1021 (2008)

    Article  MathSciNet  MATH  Google Scholar 

  • DeLellis, P., Polverino, G., Ustuner, G., Abaid, N., Macrì, S., Bollt, E.M., Porfiri, M.: Collective behaviour across animal species. Sci. Rep. 4, 3723 (2014)

    Article  Google Scholar 

  • DeLellis, P., Porfiri, M., Bollt, E.M.: Topological analysis of group fragmentation in multiagent systems. Phys. Rev. E 87(2), 022818 (2013)

    Article  Google Scholar 

  • Gautrais, J., Ginelli, F., Fournier, R., Blanco, S., Soria, M., Chaté, H., Theraulaz, G.: Deciphering interactions in moving animal groups. PLoS Comput. Biol. 8(9), 1002678 (2012)

    Article  MathSciNet  Google Scholar 

  • Gautrais, J., Jost, C., Soria, M., Campo, A., Motsch, S., Fournier, R., Blanco, S., Theraulaz, G.: Analyzing fish movement as a persistent turning walker. J. Math. Biol. 58(3), 429–445 (2009)

    Article  MathSciNet  MATH  Google Scholar 

  • Giardina, I.: Collective behavior in animal groups: theoretical models and empirical studies. HFSP J. 2(4), 205–219 (2008)

    Article  Google Scholar 

  • Heylighen, F.: The science of self-organization and adaptivity. In: Kiel, L.D. (ed.) The Encyclopedia of Life Support Systems, pp. 253–280. Eolss Publishers, Oxford (2001)

    Google Scholar 

  • Jorion, P.: On jump processes in the foreign exchange and stock markets. Rev. Financ. Stud. 1(4), 427–445 (1988)

    Article  Google Scholar 

  • Kalueff, A.V., Gebhardt, M., Stewart, A.M., Cachat, J.M., Brimmer, M., Chawla, J.S., Craddock, C., Kyzar, E.J., Roth, A., Landsman, S., Gaikwad, S., Robinson, K., Baatrup, E., Tierney, K., Shamchuk, A., Norton, W., Miller, N., Nicolson, T., Braubach, O., Gilman, C.P., Pittman, J., Rosemberg, D.B., Gerlai, R., Echevarria, D., Lamb, E., Neuhauss, S.C.F., Weng, W., Bally-Cuif, L., Schneider, H.: Towards a comprehensive catalog of zebrafish behavior 1.0 and beyond. Zebrafish 10(1), 70–86 (2013)

    Article  Google Scholar 

  • Kareiva, P.M., Shigesada, N.: Analyzing insect movement as a correlated random walk. Oecologia 56(2–3), 234–238 (1983)

    Article  Google Scholar 

  • Kelley, D.H., Ouellette, N.T.: Emergent dynamics of laboratory insect swarms. Sci. Rep. 3(1073), 1–7 (2012)

  • Krause, J., Ruxton, G.D.: Living in Groups. Oxford University Press, Oxford (2002)

    Google Scholar 

  • Ladu, F., Butail, S., Macrì, S., Porfiri, M.: Sociality modulates the effects of ethanol in zebra fish. Alcohol. Clin. Exp. Res. 38, 1–9 (2014)

    Article  Google Scholar 

  • Lin, Y., Abaid, N.: Collective behavior and predation success in a predator-prey model inspired by hunting bats. Phys. Rev. E 88, 062724 (2013)

    Article  Google Scholar 

  • Metzler, R., Klafter, J.: The random walk’s guide to anomalous diffusion: a fractional dynamics approach. Phys. Rep. 339(1), 1–77 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  • Muller, U.K., Stamhuis, E.J., Videler, J.J.: Hydrodynamics of unsteady fish swimming and the effects of body size: comparing the flow fields of fish larvae and adults. J. Exp. Biol. 203(2), 193–206 (2000)

    Google Scholar 

  • Mwaffo, V., Porfiri, M.: Turning rate dynamics of zebrafish exposed to ethanol. Int. J. Bifurc. Chaos (2015)

  • Mwaffo, V., Anderson, R.P., Butail, S., Porfiri, M.: A jump persistent turning walker to model zebrafish locomotion. J. R. Soc. Interface 12(102), 20140884 (2015)

    Article  Google Scholar 

  • Othmer, H.G., Dunbar, S.R., Alt, W.: Models of dispersal in biological systems. J. Math. Biol. 26(3), 263–298 (1988)

    Article  MathSciNet  MATH  Google Scholar 

  • Pimentel, J.A., Aldana, M., Huepe, C., Larralde, H.: Intrinsic and extrinsic noise effects on phase transitions of network models with applications to swarming systems. Phys. Rev. E 77(6), 061138 (2008)

    Article  Google Scholar 

  • Porfiri, M.: A master stability function for stochastically coupled chaotic maps. Europhys. Lett. 96(4), 40014 (2011)

    Article  Google Scholar 

  • Porfiri, M.: Linear analysis of the vectorial network model. Express Br. IEEE Trans. Circ. Syst. II 61(1), 44–48 (2014)

    MathSciNet  Google Scholar 

  • Potts, W.K.: The chorus-line hypothesis of manoeuvre coordination in avian flocks. Nature 309, 344–345 (1984)

    Article  Google Scholar 

  • Rusconi, R., Guasto, J.S., Stocker, R.: Bacterial transport suppressed by fluid shear. Nat. Phys. 10, 212–217 (2014)

    Article  Google Scholar 

  • Saintillan, D., Shelley, M.J.: Emergence of coherent structures and large-scale flows in motile suspensions. J. R. Soc. Interface 9(68), 571–585 (2012)

    Article  Google Scholar 

  • Shimoyama, N., Sugawara, K., Mizuguchi, T., Hayakawa, Y., Sano, M.: Collective motion in a system of motile elements. Phys. Rev. Lett. 76(20), 3870 (1996)

    Article  Google Scholar 

  • Sumino, Y., Nagai, K.H., Shitaka, Y., Tanaka, D., Yoshikawa, K., Chaté, H., Oiwa, K.: Large-scale vortex lattice emerging from collectively moving microtubules. Nature 483(7390), 48–452 (2012)

    Article  Google Scholar 

  • Sumpter, D.J.T.: The principles of collective animal behaviour. Philos. Trans. R. Soc. B Biol. Sci. 361(1465), 5–22 (2006)

    Article  Google Scholar 

  • Sumpter, D.J.T.: Collective Animal Behavior. Princeton University Press, Princeton (2010)

    Book  MATH  Google Scholar 

  • Tarcai, N., Viragh, C., Abel, D., Nagy, M., Varkonyi, P.L., Vasarhelyi, G., Vicsek, T.: Patterns, transitions and the role of leaders in the collective dynamics of a simple robotic flock. J. Stat. Mech. Theory Exp. 2011(04), 04010 (2011)

    Article  Google Scholar 

  • Tian, B.-M., Yang, H.-X., Li, W., Wang, W.-X., Wang, B.-H., Zhou, T.: Optimal view angle in collective dynamics of self-propelled agents. Phys. Rev. E 79, 052102 (2009)

    Article  Google Scholar 

  • Tsang, A.C.H., Kanso, E.: Flagella-induced transitions in the collective behavior of confined microswimmers. Phys. Rev. E 90, 021001 (2014)

    Article  Google Scholar 

  • Vicsek, T., Zafeiris, A.: Collective motion. Phys. Rep. 517(3–4), 71–140 (2012)

    Article  Google Scholar 

  • 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(6), 1226–1229 (1995)

    Article  Google Scholar 

  • Wang, X.F., Xu, J.: Cascading failures in coupled map lattices. Phys. Rev. E 70(5), 056113 (2004)

    Article  Google Scholar 

  • Wang, X.R., Md, M.J., Lizier, J.T., Prokopenko, M., Rossi, L.F.: Quantifying and tracing information cascades in swarms. PLoS One 7(7) (2012)

  • Zienkiewicz, A., Barton, D.A.W., Porfiri, M., diBernardo, M.: Data-driven stochastic modelling of zebrafish locomotion. J. Math. Biol. 1–25 (2014). doi:10.1007/s00285-014-0843-2

Download references

Acknowledgments

This work was supported by the National Science Foundation under Grant # CMMI-0745753, Grant # CMMI-1129820, Grant # CMMI-1433670, and GK-12 Fellows Grant # DGE-0741714.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Maurizio Porfiri.

Additional information

Communicated by Eva Kanso.

Appendix

Appendix

In this Appendix, we derive a closed-form expression for the polarization in the VNM-\(\omega \) in the vicinity of the ordered state, that is, when the noise scaling factor \(\eta \) is small.

1.1 Linear Approximation of the VNM-\(\omega \)

We linearize (5) in the neighborhood of the synchronized state \(\theta _0\) by defining \( \psi _i(k) = \theta _i(k) - \theta _0 \) to obtain the linear stochastic dynamics of the individuals’ orientation disagreements, that is,

$$\begin{aligned} \psi _i(k+1)= & {} \frac{1}{K} \sum \limits _{j=1}^{K} \psi _{i_j}(k) + \eta \omega _i(k), \end{aligned}$$
(9a)
$$\begin{aligned} \omega _i(k+1)= & {} \omega _i(k)( 1 -\alpha ) + \varepsilon _i + \gamma \tau _i. \end{aligned}$$
(9b)

For convenience, we define \(\psi (k) = \left[ \psi _1(k), \ldots , \psi _{N}(k)\right] ^\mathrm {T}\), and similarly, we write the intrinsic noise vector as \(\omega (k) = \left[ \omega _1(k),\ldots ,\omega _{N}(k)\right] ^\mathrm {T}\), where \(\mathrm {T}\) indicates matrix transposition. Using this compact notation, the linear stochastic system in (9) can be rewritten as

$$\begin{aligned} \psi (k+1)= & {} W \psi (k) + \eta \omega (k), \end{aligned}$$
(10a)
$$\begin{aligned} \omega (k+1)= & {} \omega (k)( 1 -\alpha ) + \varepsilon + \gamma \tau , \end{aligned}$$
(10b)

where \(\varepsilon = \left[ \varepsilon _1,\ldots ,\varepsilon _N\right] ^\mathrm {T} \) and \(\tau = \left[ \tau _1,\ldots ,\tau _N\right] ^\mathrm {T}\) are \(N\times 1\) vectors.

W is a random variable, which is defined so that its rows are i.i.d. vectors with K randomly selected entries taking value 1 / K, while any other entry equals 0. This corresponds to a specific instance of the numerosity-constrained network construction defined in Abaid and Porfiri (2011), with numerosity equal to \(K-1\) and weight 1 / K; in the context of numerosity-constrained networks, \(K(I_N - W)\) is the graph Laplacian of the network. While linear noisy consensus over numerosity-constrained networks is addressed in Abaid et al. (2012b), the analysis presented therein pertains to a uniform additive noise, similar to the traditional VNM. In what follows, we extend the analysis to the stochastic turn rate model in (10b).

1.2 Steady-State Mean Square Deviation

We introduce the disagreement vector \(\xi (k)= R \psi (k)\) through the orthogonal projection \(R =I_N-\frac{1}{N} 1_N 1_N^\mathrm {T}\). Next, we multiply both sides of (10a) by R to obtain

$$\begin{aligned} \xi (k+1)= & {} R W \xi (k) + \eta R \omega (k), \end{aligned}$$
(11a)
$$\begin{aligned} \omega (k+1)= & {} ( 1 -\alpha )\omega (k) + \varepsilon + \gamma \tau . \end{aligned}$$
(11b)

For any time step k, we define \(\delta (k) = \mathbf {E}\left[ \Vert \xi (k)\Vert ^2\right] \) as the expected value of the disagreement norm. In line with Porfiri (2014), Abaid et al. (2012b), Abaid and Porfiri (2011, 2012) and Porfiri (2011), we consider the autocorrelation matrix \( \varXi (k) = \mathbf {E}\left[ \xi (k)\xi (k)^\mathrm {T}\right] \) to investigate the mean square behavior of the stochastic system (11). Notably, the trace of this matrix corresponds to the expected value of the disagreement norm, that is, \(\delta (k)\). In the absence of noise, the consensus protocol in (10a) is said to be asymptotically mean square consentable if the disagreement dynamics in (11a), with \(\eta =0\), is mean square stable, which is equivalent to \(\lim _{k\rightarrow \infty }\delta (k) = 0\) for any initial condition \(\xi (0)\). In the presence of noise, the disagreement norm does not approach zero, and we define the steady-state mean square deviation as \( \delta _\infty = \lim _{k \rightarrow \infty }\delta (k) \).

Using well-known properties of the Kronecker algebra (Bernstein 2009), we obtain the following recursive relation for the autocorrelation matrix:

$$\begin{aligned} \mathrm {vec}\left( {\varXi }(k+1)\right)= & {} G \mathrm {vec}\left( {\varXi }(k)\right) + \eta \left( R\otimes R\mathbf {E}[W]\right) \mathrm {vec}\left( \mathbf {E}\left[ \xi (k)\omega (k)^\mathrm {T}\right] \right) \nonumber \\&+\, \eta \left( R\mathbf {E}[W] \otimes R\right) \mathrm {vec}\left( \mathbf {E}\left[ \omega (k)\xi (k)^\mathrm {T}\right] \right) \nonumber \\&+\, \eta ^2 R\otimes R \mathrm {vec}\left( \mathbf {E}\left[ \omega (k)\omega (k)^\mathrm {T}\right] \right) , \end{aligned}$$
(12)

where \(\otimes \) is the Kronecker product, \(\mathrm {vec}(\cdot )\) indicates vectorization, and \(G= (R\otimes R) \ \mathbf {E}\left[ W \otimes W\right] \) is a \( {N^2 \times N^2 }\) matrix.

Adapting results from Abaid and Porfiri (2011), it can be verified that

$$\begin{aligned}&\mathbf {E}[W] = I_N -\frac{N(K-1)}{K(N-1)} R, \\&R\mathbf {E}[W] = \mathbf {E}[W] R = \left( 1- \frac{N(K-1)}{K(N-1)}\right) R, \\&R\otimes R\mathbf {E}[W] = \mathbf {E}[W] R \otimes R = \left( 1- \frac{N(K-1)}{K(N-1)}\right) R\otimes R, \\&G=\left( 1-\frac{N(K-1)}{K(N-1)}\right) ^2(R\otimes R)+(I_N\otimes R)F, \end{aligned}$$

where the lengthy expression for the matrix F can be retrieved from Proposition 2 in Abaid and Porfiri (2011).

The evolution of the disagreement autocorrelation matrix is then obtained by iterating (12) from time 0 to time k , and using the fact that W and \( \omega (s) \) are independent random variables, for all s, to find

$$\begin{aligned}&\mathrm {vec}\left( {\varXi }(k+1)\right) \nonumber \\&\quad = G^{k+1} \mathrm {vec}\left( {\varXi }(0)\right) \nonumber \\&\qquad +\, \eta \left( 1- \frac{N(K-1)}{K(N-1)}\right) \sum _{s=0}^{k-1} G^{k-s-1} R\otimes R \Big \{ \mathrm {vec}\left( \mathbf {E}\left[ \xi (s)\omega (s)^\mathrm {T}\right] \right) \nonumber \\&\qquad +\, \mathrm {vec}\left( \mathbf {E}\left[ \omega (s)\xi (s)^\mathrm {T}\right] \right) \Big \} + \eta ^2 \sum _{s=0}^{k} G^{k-s} R\otimes R \ \mathrm {vec} \left( \mathbf {E}\left[ \omega (s)\omega (s)^\mathrm {T}\right] \right) ,\qquad \end{aligned}$$
(13)

where, for consistency with the main text, we have set the initial value of the turn rate to zero, that is, \( \omega (0) = {0}_N\).

From (11a) and (11b), we establish the following recursive relations:

$$\begin{aligned} \mathrm {vec}\left( \mathbf {E}\left[ \xi (s)\omega (s)^\mathrm {T}\right] \right)= & {} (1-\alpha ) (I_N \otimes R) \Big \{ \mathrm {vec}\left( \mathbf {E}\left[ \xi (s-1)\omega (s-1)^\mathrm {T}\right] \right) \nonumber \\&+\, \eta \mathrm {vec}\left( \mathbf {E}\left[ \omega (s-1) \omega (s-1)^\mathrm {T}\right] \right) \Big \} , \end{aligned}$$
(14a)
$$\begin{aligned} \mathrm {vec}\left( \mathbf {E}\left[ \omega (s)\xi (s)^\mathrm {T}\right] \right)= & {} (1-\alpha ) (R \otimes I_N) \Big \{ \mathrm {vec}\left( \mathbf {E}\left[ \omega (s-1)\xi (s-1)^\mathrm {T}\right] \right) \nonumber \\&+\, \eta \mathrm {vec}\left( \mathbf {E}\left[ \omega (s-1) \omega (s-1)^\mathrm {T}\right] \right) \Big \}. \end{aligned}$$
(14b)

Using the independence between \(\varepsilon \) and \(\tau \), one can also verify from (11b) that the evolution of the noise autocorrelation matrix satisfies the following relation:

$$\begin{aligned} \mathbf {E}\left[ \omega (s)\omega (s)^\mathrm {T}\right] = \frac{1 - (1-\alpha )^{2s}}{1-(1-\alpha )^2}\left( 1 + \lambda \gamma ^2\right) I_N, \end{aligned}$$
(15)

where for \(i=1,\ldots ,N\), using the independence between \(\varepsilon _i\) and \(\tau _i\), one can verify that the variance of the zero-mean random variable \(\varepsilon _i+\gamma \tau _i\) equals \( 1 + \lambda \gamma ^2\). By replacing (14) and (15) into (13) and iterating over k, we find

$$\begin{aligned}&\mathrm {vec}\left( {\varXi }(k+1)\right) \nonumber \\&\quad = G^{k+1} \mathrm {vec}\left( {\varXi }(0)\right) \nonumber \\&\qquad +\, 2\eta ^2 \frac{1+\lambda \gamma ^2}{1-(1-\alpha )^2}\Biggl \{ \frac{\varGamma }{1-\varGamma } \left( \sum _{s=0}^{k-1} G^{s} - \varGamma ^k \sum _{s=0}^{k-1} {\varGamma }^{-s}G^s \right) \nonumber \\&\qquad - \, (1-\alpha )^{2k}\sum _{s=0}^{k-1} (1-\alpha )^{-2s}{G}^s\frac{1-\left( \frac{\varGamma }{(1-\alpha )^2}\right) ^{s}}{1-\frac{\varGamma }{(1-\alpha )^2}} \Biggl \}\mathrm {vec}\left( R \right) \nonumber \\&\qquad +\, \eta ^2 \frac{1+\lambda \gamma ^2}{1-(1-\alpha )^2} \left( \sum _{s=0}^{k}G^{s} - (1-\alpha )^{2k} \sum _{s=0}^{k} (1-\alpha )^{-2s}{G}^s\right) \mathrm {vec} \left( R\right) , \end{aligned}$$
(16)

where \(\varGamma = (1-\alpha )\left( 1- \frac{N(K-1)}{K(N-1)}\right) \).

The convergence of the series in (16) depends on the spectral radius of G, which as shown in Abaid and Porfiri (2011) corresponds to the asymptotic convergence factor of the linear consensus protocol in the absence of noise. This quantity can be exactly calculated using Theorem 2 in Abaid and Porfiri (2011), which yields

$$\begin{aligned} r_a= 1 - \frac{\left( K-1\right) \left( N + K(N^2 - N - 1)\right) }{K^2(N-1)^2}. \end{aligned}$$
(17)

Thus, when the consensus protocol in (11) satisfies \(r_a < (1-\alpha )^{2} < 1\), the series in (16) admits a finite limit, that is,

$$\begin{aligned} \lim \limits _{k\rightarrow \infty }\mathrm {vec}\left( {\varXi }(k)\right) = \eta ^2 \mu \, \left( I_N\otimes I_N - G \right) ^{-1} \mathrm {vec}(R), \end{aligned}$$
(18)

where \(\mu \) is defined in (8). We right multiply (18) by \(\mathrm {vec}\left( I_N\right) ^\mathrm {T}\) in order to obtain the steady-state mean square deviation, that is,

$$\begin{aligned} \delta _{\infty } = \eta ^2 \mu \, \mathrm {vec}(I_N)^{\mathrm {T}} \left[ I_N \otimes I_N -G\right] ^{-1} \mathrm {vec}(R). \end{aligned}$$
(19)

The final computation of \(\delta _\infty \) in (19) can be performed by employing Proposition 2 in Abaid et al. (2012b) and the related arguments. Thus, we find that when \(r_a < (1-\alpha )^2 < 1\), the steady-state mean square deviation in (19) reduces to

$$\begin{aligned} \delta _\infty =\eta ^2 \frac{\mu (N-1)}{1-r_a}, \end{aligned}$$
(20)

where \(\mu \) and \(r_a\) depend on K and N through (8) and (17).

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Mwaffo, V., Anderson, R.P. & Porfiri, M. Collective Dynamics in the Vicsek and Vectorial Network Models Beyond Uniform Additive Noise. J Nonlinear Sci 25, 1053–1076 (2015). https://doi.org/10.1007/s00332-015-9260-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00332-015-9260-y

Keywords

Mathematics Subject Classification

Navigation