Model of collective fish behavior with hydrodynamic interactions

1 2 3 4 1, Audrey Filella, Fran¸coisNadal, Cl´ement Sire, Eva Kanso, and Christophe Eloy ∗ 1Aix Marseille Univ, CNRS, Centrale Marseille, IRPHE, 13013 Marseille, France 2Department of Mechanical, Electrical and Manufacturing Engineering, Loughborough University, Loughborough LE11 3TU, UK 3CNRS, Universit´ePaul Sabatier, Laboratoire de Physique Th´eorique,31062 Toulouse, France 4Aerospace and Mechanical Engineering, University of Southern California, 854 Downey Way, Los Angeles, CA 90089, USA (Dated: May 4, 2018) schooling is often modeled with self-propelled particles subject to phenomenological behav- ioral rules. Although fish are known to sense and exploit flow features, these models usually neglect hydrodynamics. Here, we propose a novel model that couples behavioral rules with far-field hydro- dynamic interactions. We show that (1) a new “collective turning” phase emerges; (2) on average individuals swim faster thanks to the fluid; (3) the flow enhances behavioral noise. The results of this model suggest that hydrodynamic effects should be considered to fully understand the collective dynamics of fish.

Collective animal motion is ubiquitous: insects , tices generated by other fish to reduce the energetic costs birds flock, hoofed vertebrates , and even humans of locomotion [27–29]. Fish use their , a hair- exhibit coordination in crowds [1–5]. Among these fas- based sensor running along their side [30], to sense the cinating collective behaviors, schooling refers to the co- surrounding flow, which has also proved crucial for collec- ordinated motion of fish. It is exhibited by half of the tive behavior [31, 32]. Yet, the existing behavioral models known fish species during some phase of their life cy- of fish schooling do not include hydrodynamics. Here, we cle [6] and can generate different disordered phases, such propose to combine a data-driven attraction-alignment as (important cohesion but low polarization of model [19–21], with far-field hydrodynamic interactions. fish heading), or ordered phases: milling (torus or vor- In the context of this new model, several questions arise: tex pattern), bait ball (dense “ball” of fish), or highly Can the swimmers exploit the flow to swim faster on av- polarized schools [7]. erage? Do the hydrodynamic interactions give rise to Interestingly, these collective phases can be achieved novel collective phases? Does the flow play the role of without any leader in the group. This has first been ob- a self-induced noise, as it is the case at low Reynolds served in experiments [8], and later confirmed by the de- number [33]? velopment of self-propelled particle (SPP) models [9, 10]. Fish are modeled as self-propelled particles moving in In the context of fish schooling, these mathematical mod- an unbounded two-dimensional plane. They move at con- els have generally been constructed by assuming simple stant speed v relative to the flow and exhibit no inertia phenomenological behavioral rules, such as the popular [34]. Following behavioral rules inferred from shallow- “three-A rules” of avoidance, alignment, and attraction water tracking experiments [19, 20], we consider that [11–13]. From a physical point of view, these SPP models each individual is attracted to its Voronoi neighbors with 1 1 have an obvious interest because of their simplicity and intensity kp (units m− s− ), tends to align with the same 1 universality [14], and because they allow the derivation neighbors with intensity kv (units m− ), and is subject of continuum equations [15]. Similar approaches have to a rotational noise with standard deviation σ (units been used in soft (e.g., bacteria swarms and microtubule bundles) to derive continuous models (a) (b) taking into account hydrodynamic interactions at vanish- ing Reynolds number [16, 17]. However, they have rarely arXiv:1705.07821v4 [physics.bio-ph] 3 May 2018 been connected quantitatively to experimental observa- tions [18]. It is only recently that it has been possible to infer and model the actual behavioral rules from the individual tracking of fish in a tank [19–21]. Schooling likely serves multiple purposes [22], includ- ing better foraging for patchy resources and increased FIG. 1. (a) Sketch of two interacting swimmers, showing the k protection against predators. Fish are also thought to heading ei of swimmer i, its viewing angle θij , the relative benefit from the hydrodynamic interactions with their alignment angle φij , the inter-swimmer distance ρij , and the ρ θ neighbors [23–26], but it is unclear whether this requires polar coordinates in the reference frame of swimmer j (ej ,ej ). particular configurations or regulations. When swim- (b) Grayscale representation of the anisotropic visual percep- ming in a structured flow, fish can exploit near-field vor- tion, modeled by the term (1 + cos θij ) in Eqs. (2–4). 2

1/2 rad s− ). Moreover, each swimmer responds to the far-field flow disturbance created by all other swimmers. This flow is an elementary dipole, with dipole intensity 2 Sv in two dimensions, where S = πr0 is the swimmer surface and r0 its typical length [25, 35–37]. Note that, except for hydrodynamic interactions involving the swim- mer size r0, swimmers are considered as point-like parti- cles. We use v and kp to make the problem dimension- p less, yielding the length scale v/kp and time scale p p 1/4 1/ vkp. To this end, I = kv v/kp, In = σ(vkp)− k and If = Skp/v characterize the alignment, noise and dipole intensities, respectively. The dimensionless equa- tions of motion are

r˙i = eik + Ui, (1)

θ˙i = ρij sin(θij) + I sin(φij) + Inη + Ωi. (2) h k i Equation (1) expresses that each individual, located at ri, is moving with a constant unit speed along its ori- entation eik (Fig. 1a). An additional drift term, Ui, arises when hydrodynamic interactions are taken into ac- count. A far-field approximation is used to model the flow [37, 38]. Here, we choose to neglect the vorticity shed in the swimmer wakes [39, 40] to keep the model simple and tractable. Under this potential flow approxi- mation, each swimmer generates a dipolar flow field, and we can use the principle of superposition to calculate the flow Ui experienced by a swimmer

θ ρ X If ej sin θji + ej cos θji U = u , with u = , (3) i ji ji π ρ2 j=i ij 6 where uji is the velocity induced by swimmer j at the ρ θ FIG. 2. (a–d) Plots of the swimmer positions and associ- position ri and (ej ,ej ) are the polar coordinates in the ated streamlines for different values of the parameters. On all −2 framework of swimmer j (Fig. 1a). The angular velocity figures, the dipole intensity is If = 10 , the scale bar corre- in Eq. (2) is the sum of an attraction term, an align- sponds to 10 r0, and color scale represents the instantaneous ment term, a standard Wiener process η(t), describing velocity. For clarity, swimmers are represented as “airfoils” of the spontaneous motion of the fish and modeling its “free length 7r0. Four distinct dynamical phases are observed: (a) swarming for In = 0.8, I = 0.5; (b) schooling for In = 0.5, will”, and a rotational term Ωi induced by hydrodynamic k interactions. The behavioral terms (attraction and align- Ik = 9; (c) milling for In = 0.3, Ik = 1.5; and (d) turning for I = 0.2, I = 4. (e) Paths followed by each swimmer during ment) are averaged over the Voronoi neighbors, noted , n k Vi 12 dimensionless time units, the final time corresponding to with the weight (1 + cos θij) modeling continuously a rear (d). For long-time dynamics, see Supplementary Fig. 1 and blind angle [20] (Fig. 1b) Supplementary Movies 1–4.

X  X = (1 + cos θ ) (1 + cos θ ). (4) h◦i ◦ ij ij j i j i ∈V ∈V Another interpretation of this angular velocity is to The rotation induced by hydrodynamic interactions is consider that each swimmer is a dumbbell oriented in the not due to vorticity, since it is zero for a potential flow, swimming direction, with each weight advected by the but to gradients of normal velocities along the swimming flow (Supplementary Fig. 2). The intensity of the hydro- direction. In other words, the angular velocity due to the dynamic interactions (related to both the drift term Ui flow is and the induced rotation Ωi) is proportional to the dipole X intensity If . For the 10 cm-long fish (r0 = 5 cm) consid- Ω = ek ∇u e⊥. (5) 1 i i · ji · i ered in Ref. [19], the cruise speed is v 0.2 m s− , and j=i 1 1 ≈ k = 0.41 m− s− , yielding a dipole intensity I 0.016. 6 p f ≈ 3

(a) No fluid (b) Fluid ( ⌦ i =0 ) (c) Fluid ( ⌦ i =0 ) (d) Color code 6 1 milling turning k I 1. 2 M 0.5 milling alignment swarming schooling 0 0 0.5 1 noise In noise In noise In polarization P

FIG. 3. Phase diagram using the value of the polarization P and the milling M as a color code (d), for three cases: (a) no fluid (Ui = Ωi = 0); (b) fluid without induced rotation (Ui 6= 0, Ωi = 0); (c) full hydrodynamic model (Ui 6= 0, Ωi 6= 0). In all −2 cases, the dipole intensity is If = 10 , the solid white line indicates the M = 0.4 level, and the dashed line the P = 0.5 level. Solid black lines in (b–c) show the V -levels and the solid blue line in (a) shows the milling-schooling transition line found in [20]. In (c), the black dots show the parameter values corresponding to Fig. 2a–d.

In the present model, the flow induces translational and all individuals. The parameter P is the polarization, M rotational motions, whose origin are physical, but, in is the milling and corresponds to the normalized angular principle, it could also elicit a behavioral response (e.g., momentum of the group (straight-line schooling gives a a tendency for fish to go along or against the flow) [21]. value of 0, while perfect milling gives 1). We consider a group of N = 100 individuals,Fig3 with To assess the importance of hydrodynamic interac- random initial orientations and initially distributed in a tions, we performed a systematic parametric study for 20 20 box (in dimensionless length units), although the × three cases (Fig. 3): a pure behavioral model with no subsequent dynamics is not affected by the initial con- effects of the fluid (Ui = 0 and Ωi = 0, in Eqs. (1– ditions. The dynamical system described by Eqs. (1–5) 2)), a simple model of hydrodynamic drift without in- is solved numerically, using an explicit scheme with time duced rotation (Ui = 0 and Ωi = 0), and a full hydrody- 2 6 step δt = 10− . Depending on the values of the three namic model with both induced translation and rotation dimensionless parameters (I , In, and If ), four different (U = 0 and Ω = 0). For each set of parameters, P , M, k i i dynamical phases emerge (Fig. 2). When noise is com- and 6 V are obtained6 after time-averaging over ∆t = 100 parable or larger than the alignment, we observe a disor- (after waiting 100 time units to ensure that the transient dered swarming phase (Fig. 2a): swimmers form a sparse dynamics of few dimensionless time units is over), and group with no preferential orientation. When alignment ensemble-averaging over 100 realizations. intensity is stronger, the group is denser and individu- In the absence of hydrodynamic interactions, the re- als tend to swim in the same direction: this is known sults of Ref. [20] are recovered (Fig. 3a). For P > 0.5, as the schooling phase (Fig. 2b). When alignment and which roughly correspond to I & 2, we observe the attraction are comparable and noise is low or moderate, k schooling phase. For M > 0.4, obtained for I . 2 and the group reaches a milling phase (Fig. 2c): it forms a k In . 0.5, the group exhibits a milling phase. For all other “vortex”. These three phases (swarming, schooling, and cases tested, the swarming phase is observed. We chose milling) can also be observed without any hydrodynamic the threshold values P = 0.5 and M = 0.4 to distin- U interactions [20] ( i = 0 and Ωi = 0, in Eqs. (1–2)). guish the four phases, but P and M vary continuously in However, when the flow is explicitly taken into account, the parameter space. An alternative choice of thresholds turning a novel phase appears that we call the phase would be possible and would yield qualitatively similar (Fig. 2d–e). In this new phase, swimmers tend to align results. along a preferential orientation and, at the same time, When hydrodynamic drift is introduced with I = the group follows a large-scale quasi circular trajectory. f 10 2, but induced rotation is neglected (Fig. 3b), the In order to precisely characterize these different phases, − phase diagram is practically unchanged. The only differ- the global order parameters P and M are introduced [41], ence is that the mean velocity V is now slightly greater along with the average speed V than 1 (V = 1 when hydrodynamic interactions are ne- r ei r˙i glected). When the full hydrodynamic model is con- P = ek ,M = × ,V = r˙ , (6) i r i sidered (Fig. 3c), the new turning phase appears for | | ei r˙i | | M > 0.4 and P > 0.5 (corresponding to In . 0.25 and r where ei = (ri ri)/ ri ri is the unit vector along 3 . I . 5) and V is increased. Looking at the swim- the segment joining− the| center− | of mass of the group and mingk speeds of each individual in the group (Fig. 2e), the i-th swimmer, and the over bar denotes average over we see that some individuals swim slower because of the 4

(a) swarming (b) schooling and swimmers thus tend to spend more time in-line. The role of the fluid is not only to increase the swim- ming speed on average, but also to introduce a source of disorder. To assess if this disorder has the same ef- fective impact as the noise Inη in Eq. (2) that describes 100r0 100r0 the spontaneous angle fluctuations, we performed simu- lations with no noise (In = 0) and with varying dipole in- tensity If (Fig. 5a). The phase diagrams shown in Fig. 3c and Fig. 5a are qualitatively similar, both exhibiting the four phases (schooling, swarming, milling, and turning) (c) milling (d) turning with the same topology. It thus shows that the hydrody- namic interactions also play the role of a rotational noise. There are however some differences between Fig. 3c and Fig. 5a. First, the average velocity increases when the dipole intensity increases whereas it tends to decrease

100r0 100r0 with noise intensity. Second, for large If and small I , even if P and M, are both small, the school does notk behave as in the swarming phase. It can be composed of very dense clusters of quasi static swimmers (Supplemen- tary Fig. 5). This non-realistic behavior is an artifact of the simulations due to the absence of noise. There is a continuous transition between the milling, p/pmax turning,Fig4 and schooling phases (Figs. 3 and 5). In a pure behavioral model (with no fluid), the transition between FIG. 4. Heat maps showing the probability of presence p(ρ, θ) the milling and schooling is also continuous, but M is of all other swimmers in the framework of an individual. The systematically higher when the fluid is present (Fig. 5b). parameters are the same as in Fig. 2. Although the turning and milling phases are similar, their origins are different. As Calovi et al. [20] al- ready noted, the milling phase can only be stabilized fluid, and some others swim much faster with swimming when swimmers have an anisotropic visual perception speed reaching r˙ = 2.5. | i| (Fig. 1b and Eq. (4)). On the contrary, the turning To understand why individuals always swim faster on phase still exists when visual perception is isotropic (Sup- average when hydrodynamic interactions are taken into plementary Fig. 7), but requires the full hydrodynamic account, we computed the probability of presence of other model to be observed. Although experimental data are swimmers in the framework of each individual (Fig. 4). too scarce to support the existence of the turning phase For the same parameter values as in Fig. 2, we collected in real fish schools, we can speculate that the speed en- the position and orientation of each swimmer during hancement achieved in this phase could be advantageous ∆t = 900. For the swarming phase (Fig. 4a), the proba- bility of presence is isotropic. Hence, there is no velocity increase due to dipolar hydrodynamic interactions on av- (a) (b) erage. However, for the schooling phase, individuals tend to swim in-line rather than side-by-side, leading to a den- 1 k I sity distribution polarized along the vertical (Fig. 4b), or V equivalently θji preferentially around 0◦ or 180◦ (Fig. 1). P 0.5 This induces a velocity increase along the swimming di- 0.5 M rection eik (see Eq. (3)). The same is true for the milling alignment and turning phases (Fig. 4c–d). 0 Why do individuals tend to swim in-line in the presence 0510 of the fluid? To address this question, we examined the dipole intensity If alignment I preferential location of the nearest Voronoi neighbors in ∥ the swarming and schooling phases, in the presence of the FIG. 5. (a) Phase diagram in the absence of noise (In = 0). fluid or not (Supplementary Fig. 3). It appears that the The color code and the contour levels are the same as in Fig. 3. preferential in-line configuration is only present in the full (b) Values of the polarization P (red), milling M (green), hydrodynamic model. This is because the side-by-side and mean velocity V (black) for two cases: no fluid (If = 0) and low noise (In = 0.05) with solid lines; full hydrodynamic configuration becomes unstable when hydrodynamic in- −2 teractions are considered [40, 42] (Supplementary Fig. 4), model (If = 10 ) and no noise (In = 0) with open symbols.

Fig5 5 to the group, for energetic considerations, or when con- [11] I. Aoki, Bull. Jap. Soc. Sci. Fish. 48, 1081 (1982). fronted to a danger. Note that the milling and the turn- [12] A. Huth and C. Wissel, J. Theor. Biol. 156, 365 (1992). ing phases, when the number of swimmers is large, can [13] I. D. Couzin, J. Krause, R. James, G. D. Ruxton, and N. R. Franks, J. Theor. Biol. 218, 1 (2002). break into several smaller groups and thus affect the value ´ of the order parameters P and M (Supplementary Movies [14]G. Odor, Rev. Mod. Phys. 76, 663 (2004). [15] J. Toner, Y. Tu, and S. Ramaswamy, Ann. Phys. 318, 5-7, and Supplementary Fig. 6). 170 (2005). In summary, we proposed a new model of collective [16] M. C. Marchetti, J. Joanny, S. Ramaswamy, T. Liver- fish motion that includes behavioral rules and far-field pool, J. Prost, M. Rao, and R. A. Simha, Rev. Mod. hydrodynamic interactions. By simulating numerically Phys. 85, 1143 (2013). the dynamics of this model for a large group of swim- [17] D. Saintillan and M. J. Shelley, in Complex Fluids in mers, we showed that, on average, fish swim faster in a biological systems (Springer, 2015) pp. 319–355. [18] A. Huth and C. Wissel, Ecological modelling 75, 135 school, due to the presence of the fluid. This suggests (1994). that fish would need less energy to swim in a school for [19] J. Gautrais, F. Ginelli, R. Fournier, S. Blanco, M. So- a given swimming speed. This emergent property results ria, H. Chat´e, and G. Theraulaz, PLoS Comput Biol 8, from the preferential in-line pairing of swimmers, more e1002678 (2012). robust than the side-by-side configuration. In addition, [20] D. S. Calovi, U. Lopez, S. Ngo, C. Sire, H. Chat´e, and we observed a new phase, called the turning phase, which G. Theraulaz, New Journal of Physics 16, 015026 (2014). only exists with the full hydrodynamic model. Finally, [21] D. S. Calovi, A. Litchinko, V. Lecheval, U. Lopez, A. P´erezEscudero, H. Chat´e,C. Sire, and G. Theraulaz, we showed that the fluid has similar effect to the sponta- PLoS Comp. Biol. 14, e1005933 (2018). neous cognitive rotational noise. These promising results [22] J. Krause and G. D. Ruxton, Living in groups (Oxford underline the importance of hydrodynamic interactions University Press, 2002). in fish schooling. In future work, it will be important [23] D. Weihs, Nature (London) 241, 290 (1973). to assess the validity of the far-field approximation used [24] S. Alben, in Natural Locomotion in Fluids and on Sur- here by integrating the fish wakes into the fluid model. faces, The IMA Volumes in Mathematics and its Appli- cations No. 155, edited by S. Childress, A. Hosoi, W. W. We are grateful to G. Theraulaz for his valuable Schultz, and J. Wang (Springer New York, 2012) pp. insight. A. F. acknowledges support from A*MIDEX 3–13. (ANR-11-IDEX-0001-02), and the Labex MEC (ANR- [25] A. C. H. Tsang and E. Kanso, J. Nonlin. Sc. 23, 971 10-LABX-0092). E. K. acknowledges sabbatical support (2013). from the Flatiron Institute at the Simons Foundation and [26] C. Hemelrijk, D. Reid, H. Hildenbrandt, and J. Padding, research support from Office of Naval Research (ONR) Fish and Fisheries 16, 511 (2015). through grants N00014-14-1-0421 and N00014-17-1-2287 [27] J. C. Liao, D. N. Beal, G. V. Lauder, and M. S. Tri- antafyllou, J. Exp. Biol. 206, 1059 (2003). and the Army Research Office (ARO) through the grant [28] J. C. Liao, D. N. Beal, G. V. Lauder, and M. S. Tri- W911NF-16-1-0074. antafyllou, Science 302, 1566 (2003). [29] D. N. Beal, F. S. Hover, M. S. Triantafyllou, J. C. Liao, and G. V. Lauder, J. Fluid Mech. 549, 385 (2006). [30] H. Bleckmann, Forts. Zool. (1994). [31] B. L. Partridge and T. J. Pitcher, J. Comp. Physiol. 135, ∗ [email protected] 315 (1980). [1] C. W. Reynolds, in ACM SIGGRAPH Computer Graph- [32] K. Faucher, E. Parmentier, C. Becco, N. Vandewalle, and ics, Vol. 21 (ACM, 1987) pp. 25–34. P. Vandewalle, Animal Behaviour 79, 679 (2010). [2] I. D. Couzin and J. Krause, Adv. Study Behav. 32, 1 [33] S. D. Ryan, B. M. Haines, L. Berlyand, F. Ziebert, and (2003). I. S. Aranson, Physical Review E 83, 050904 (2011). [3] T. Vicsek and A. Zafeiris, Physics Reports 517, 71 [34] Inertia is assumed to be negligible, as it was observed in (2012). the experiments made on Kuhlia mugil [19] (α  1 in [4] M. Moussa¨ıd,D. Helbing, and G. Theraulaz, Proc. Nat. Ref. [20]). Acad. Sc. USA 108, 6884 (2011). [35] J. Lighthill, An informal introduction to theoretical fluid [5] M. Moussaid, E. G. Guillot, M. Moreau, J. Fehrenbach, mechanics (Oxford University Press, New York, NY, O. Chabiron, S. Lemercier, J. Pettr´e,C. Appert-Rolland, 1986). P. Degond, and G. Theraulaz, PLoS Comp. Biol. 8, [36] J. Lighthill, “Mathematical approaches in hydrodynam- e1002442 (2012). ics,” (SIAM, 1991) Chap. Hydrodynamic far fields, pp. [6] E. Shaw, American Scientist 66, 166 (1978). 3–20. [7] U. Lopez, J. Gautrais, I. D. Couzin, and G. Theraulaz, [37] A. A. Tchieu, E. Kanso, and P. K. Newton, Proc. R. Interface focus (2012), 10.1098/rsfs.2012.0033. Soc. London Ser. A 468, 3006 (2012). [8] D. V. Radakov, Schooling in the ecology of fish (J. Wiley, [38] M. Gazzola, A. Tchieu, D. Alexeev, A. de Brauer, and 1973). P. Koumoutsakos, J. Fluid Mech. 789, 726 (2016). [9] T. Vicsek, A. Czir´ok, E. Ben-Jacob, I. Cohen, and [39] G. V. Lauder and E. G. Drucker, Physiology 17, 235 O. Shochet, Phys. Rev. Lett. 75, 1226 (1995). (2002). [10] G. Gr´egoire,H. Chat´e, and Y. Tu, Physica D: Nonlinear [40] A. D. Becker, H. Masoud, J. W. Newbolt, M. Shelley, Phenomena 181, 157 (2003). 6

and L. Ristroph, Nature Communications 6, 8514 (2015). der parameters (P , M), defined with some arbitrariness [41] Following the usage in collective behavior studies, we re- when the system is finite. fer to the different spatial organizations of the swimming [42] E. Kanso and A. C. H. Tsang, Fluid Dyn. Res. 46, 061407 group as “phases”, which should not to be confused with (2014). the phases used in equilibrium statistical physics. Here, they correspond to different regions in the space of or-