Universal Scaling Laws for a Generic Swimmer Model

Bruno Ventéjou bruno.ventejou@univ-grenoble-alpes.fr Univ. Grenoble Alpes, CNRS, LIPhy, 38000 Grenoble, France    Thibaut Métivet thibaut.metivet@inria.fr Univ. Grenoble Alpes, Inria, CNRS, Grenoble INP, LJK, 38000 Grenoble, France    Aurélie Dupont Univ. Grenoble Alpes, CNRS, LIPhy, 38000 Grenoble, France    Philippe Peyla philippe.peyla@univ-grenoble-alpes.fr Univ. Grenoble Alpes, CNRS, LIPhy, 38000 Grenoble, France
Abstract

We have developed a minimal model of a swimmer without body deformation based on force and torque dipoles which allows accurate 3D Navier-Stokes calculations. Our model can reproduce swimmer propulsion for a large range of Reynolds numbers, and generate wake vortices in the inertial regime, reminiscent of the flow generated by the flapping tails of real fish. We performed a numerical exploration of the model from low to high Reynolds numbers and obtained universal laws using scaling arguments. We collected data from a wide variety of micro-organisms, thereby extending the experimental data presented in (M. Gazzola et al., Nature Physics 10, 758, 2014). Our theoretical scaling laws compare very well with experimental data across the different regimes, from Stokes to turbulent flows. We believe that this model, due to its relatively simple design, will be very useful for obtaining numerical simulations of collective effects within fish schools composed of hundreds of individuals.

Introduction The wide variety of means employed by living creatures to move in aquatic environments is fascinating Childress (1981). Motion generally involves a complex interplay between the deformation of the body and the surrounding fluid. From the smallest organisms, like bacteria, to colossal blue whales Berg (2004); S.Powar et al. (2022); Garcia et al. (2011); Smits (2019); Gray (1936); Wolfgang et al. (1999); Dabiri et al. (2006); Müller et al. (2001), the differences in length scale L𝐿Litalic_L, velocity v𝑣vitalic_v and mode of locomotion are so vast that the elaboration of a universal model to describe swimming across these scales might seem impossible.

The importance of inertia with respect to viscous dissipation is quantified by the Reynolds number, which expresses the ratio of stress due to inertia to stress due to viscosity: Re=ρvL/η𝑅𝑒𝜌𝑣𝐿𝜂Re=\rho vL/\etaitalic_R italic_e = italic_ρ italic_v italic_L / italic_η, where ρ𝜌\rhoitalic_ρ and η𝜂\etaitalic_η represent fluid density and viscosity, respectively. The Reynolds numbers associated with aquatic living species span several decades, typically ranging from Refish103106similar-to𝑅superscript𝑒fishsuperscript103superscript106Re^{\rm fish}\sim 10^{3}-10^{6}italic_R italic_e start_POSTSUPERSCRIPT roman_fish end_POSTSUPERSCRIPT ∼ 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT for fish to Remicroorg103less-than-or-similar-to𝑅superscript𝑒microorgsuperscript103Re^{\rm micro-org}\lesssim 10^{-3}italic_R italic_e start_POSTSUPERSCRIPT roman_micro - roman_org end_POSTSUPERSCRIPT ≲ 10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT for micro-organisms like spermatozoa. Consequently, the drag force exerted by the fluid on the swimmer is largely dependent on the species considered, originating from either fully viscous dissipation at the small scale, or turbulent inertia at the large scale.

As a result, each swimming model Taylor (1951); Spagnolie and Lauga (2010); Lauga (2020); Lighthill (1960); Liu and Kawachi (1999); Kern and Koumoutsakos (2006) tends to offer a tailored approach specific to the flow regime considered and the corresponding body deformation. Most models focus on periodic body deformation, coupled with the surrounding fluid, and resolve the full swimming cycle. This provides specific approaches that address swimming at small scales, where viscous forces dominate (e.g. for micro-organisms Lauga and Powers (2009)), differently from the macroscopic strategies of large fish or mammals, which can leverage the inertia of the surrounding fluid to break time reversibility Purcell (1977). The highly diverse physical origins of particular swimming patterns represent an obstacle to the exploration of a more comprehensive and universal viewpoint.

In this letter, we propose a different approach. Deformation kinematics, such as undulations, oscillations and pulsations, are ignored, and locomotion is described using force and torque dipoles applied by a solid body of finite size L𝐿Litalic_L on a fluid. While a similar description has already been used for micro-swimmers at low Re𝑅𝑒Reitalic_R italic_e Hernandez et al. (2005); Mehandia and Nott (2008); Jibuti et al. (2014), to the best of our knowledge, it has never yet been employed for high Re𝑅𝑒Reitalic_R italic_e, when inertia starts to dominate viscous forces. In this work, we study the motion described by our model over 8888 decades of Reynolds numbers (105Re104less-than-or-similar-tosuperscript105𝑅𝑒less-than-or-similar-tosuperscript10410^{-5}\lesssim Re\lesssim 10^{4}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT ≲ italic_R italic_e ≲ 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT), theoretically and numerically, comparing our results with experimental data. Although the direct effect on swimming velocity of the specific deformation of the swimmer’s body is acknowledged, particularly as it can reduce the drag force Gray (1936); Li et al. (2021), we would like to emphasize that we are not attempting to provide a detailed and precise analysis of a particular mode of locomotion, but rather a more universal description in terms of the forces applied to the fluid. While our swimmer model is minimal, the motion of the surrounding fluid is accurately captured using the full numerical resolution of the 3D Navier-Stokes equation, and our approach encompasses the different swimming regimes of a wide variety of aquatic species. Our model can also remarkably reproduce the characteristic wake vortices observed behind fish due to the flapping of their tails Drucker and Lauder (2002).

Our approach furthermore exhibits universal scaling laws which link the swimming Reynolds number Re𝑅𝑒Reitalic_R italic_e to a new dimensionless group, the thrust number defined below. We identify three different regimes: the Stokes regime (Re<1𝑅𝑒1Re<1italic_R italic_e < 1), a laminar regime (1<Re<1031𝑅𝑒superscript1031<Re<10^{3}1 < italic_R italic_e < 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT) and a turbulent regime (Re>103104𝑅𝑒superscript103superscript104Re>10^{3}-10^{4}italic_R italic_e > 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT), and calculate the theoretical exponents of the scaling laws in the three regimes using simple scaling analyses (independently of the space dimension). These compare very well with the numerical simulations produced by our generic swimmer model. We also validate our results with experimental data presented in Gazzola et al. (2014) for the laminar and turbulent regimes, and further extend this validation with data collected on micro-swimmers for the Stokes regime.

Refer to caption

Figure 1: The swimmer is an ellipsoidal rigid body of length L𝐿Litalic_L and width L/ar𝐿subscript𝑎𝑟L/a_{r}italic_L / italic_a start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, with arsubscript𝑎𝑟a_{r}italic_a start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT the aspect ratio. a) Time-dependent dipoles of forces {𝑭,𝑭}𝑭𝑭\{-\bm{{F}},\bm{{F}}\}{ - bold_italic_F , bold_italic_F } and torques {𝑻,𝑻}𝑻𝑻\{-\bm{{T}},\bm{{T}}\}{ - bold_italic_T , bold_italic_T } are applied at the center-of-mass of the body 𝑿subscript𝑿\bm{{X}}_{\mathcal{B}}bold_italic_X start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT and the “phantom tail” 𝑿tsubscript𝑿𝑡\bm{{X}}_{t}bold_italic_X start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, with 𝑭(t)=(π/2)𝑭0|cos(ωt)|𝑭𝑡𝜋2subscript𝑭0𝜔𝑡\bm{F}(t)=(\pi/2)\bm{F}_{0}\lvert\cos(\omega t)\rvertbold_italic_F ( italic_t ) = ( italic_π / 2 ) bold_italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | roman_cos ( italic_ω italic_t ) | and 𝑻(t)=𝑻0cos(ωt)𝑻𝑡subscript𝑻0𝜔𝑡\bm{T}(t)=\bm{T}_{0}\cos(\omega t)bold_italic_T ( italic_t ) = bold_italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_cos ( italic_ω italic_t ). The position of 𝑿tsubscript𝑿𝑡\bm{{X}}_{t}bold_italic_X start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT can be controlled to steer the swimmer, but is kept such that 𝑿𝑿t=L/2+L/arnormsubscript𝑿subscript𝑿𝑡𝐿2𝐿subscript𝑎𝑟\|\bm{{X}}_{\mathcal{B}}-\bm{{X}}_{t}\|=L/2+L/a_{r}∥ bold_italic_X start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT - bold_italic_X start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ∥ = italic_L / 2 + italic_L / italic_a start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT. b) Static model obtained by averaging the time dependent model.
Refer to caption
Figure 2: Re𝑅𝑒Reitalic_R italic_e as a function of Th𝑇Thitalic_T italic_h from 3D numerical simulations (crosses), obtained by solving the full Navier-Stokes equations with our swimmer model. We clearly obtain three regimes: ReTh1.0similar-to𝑅𝑒𝑇superscript1.0Re\sim Th^{1.0}italic_R italic_e ∼ italic_T italic_h start_POSTSUPERSCRIPT 1.0 end_POSTSUPERSCRIPT for Re20less-than-or-similar-to𝑅𝑒20Re\lesssim 20italic_R italic_e ≲ 20; ReTh0.66similar-to𝑅𝑒𝑇superscript0.66Re\sim Th^{0.66}italic_R italic_e ∼ italic_T italic_h start_POSTSUPERSCRIPT 0.66 end_POSTSUPERSCRIPT for 20Re1000less-than-or-similar-to20𝑅𝑒less-than-or-similar-to100020\lesssim Re\lesssim 100020 ≲ italic_R italic_e ≲ 1000 and ReTh0.51similar-to𝑅𝑒𝑇superscript0.51Re\sim Th^{0.51}italic_R italic_e ∼ italic_T italic_h start_POSTSUPERSCRIPT 0.51 end_POSTSUPERSCRIPT for Re1000greater-than-or-equivalent-to𝑅𝑒1000Re\gtrsim 1000italic_R italic_e ≳ 1000. The three lines correspond to fitting curves and give the numerical scaling exponents. Note that the crossovers between the different regimes depend on the geometry of the swimmer; this curve corresponds to ar=4subscript𝑎𝑟4a_{r}=4italic_a start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT = 4.

Swimmer model.

The model uses a time-dependent force dipole combined with a torque dipole (Fig. 1(a)), both attached to a rigid body \mathcal{B}caligraphic_B of ellipsoidal shape. An autonomous swimming body creates its own motion, therefore the total sum of forces and torques must cancel out, due to the third law of Newton of the fluid-body system. Feynman et al. (2006). The model developed by Filella et al. (2018) presents some conceptual similarities. It represents each fish as a point-like active particle bearing a dipole in a potential 2D inviscid fluid, which allows consideration of the hydrodynamic interactions between fish in the far-field limit. However, since our aim is to explore a wide spectrum of Reynolds numbers (from viscous to inertial regimes), we solved the full incompressible Navier-Stokes equations in 3D and 2D (see SM ven (a)).

The swimmer exerts on the fluid a force dipole {𝑭,𝑭}𝑭𝑭\{\bm{-F},\bm{F}\}{ bold_- bold_italic_F , bold_italic_F } like that generally used for a micro-swimmer at a low Re𝑅𝑒Reitalic_R italic_e Hernandez et al. (2005); Mehandia and Nott (2008); Jibuti et al. (2014) (see Fig. 1). We used a pusher-like model that reproduces the force distribution of a fish at high Reynolds numbers. This approach can easily be extended to a more detailed model by using more complex force distributions (e.g. Mehandia and Nott (2008)). As shown in Fig. 1(a), the force dipole is composed of one force applied in the fluid at the rear of the body 𝑿tsubscript𝑿𝑡\bm{{X}}_{t}bold_italic_X start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT, mimicking a swimming organ, and an opposite force exerted inside the body at the center-of-mass 𝑿subscript𝑿\bm{{X}}_{\mathcal{B}}bold_italic_X start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT. The force is time-dependent with pulsation ω𝜔\omegaitalic_ω and pusher-like: 𝑭(t)=(π/2)𝑭𝟎|cosωt|𝑭𝑡𝜋2subscript𝑭0𝜔𝑡\bm{F}(t)=(\pi/2)\bm{F_{0}}\lvert\cos\,\omega t\rvertbold_italic_F ( italic_t ) = ( italic_π / 2 ) bold_italic_F start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT | roman_cos italic_ω italic_t | with 𝐅0(𝑿𝑿t)>0subscript𝐅0subscript𝑿subscript𝑿𝑡0\mathbf{F}_{0}\cdot(\bm{{X}}_{\mathcal{B}}-\bm{{X}}_{t})>0bold_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⋅ ( bold_italic_X start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT - bold_italic_X start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) > 0. The absolute value in the expression of 𝑭(t)𝑭𝑡\bm{F}(t)bold_italic_F ( italic_t ) enforces the pusher nature of the swimmer. We also consider a torque dipole: 𝑻(t)=𝑻𝟎cosωt𝑻𝑡subscript𝑻0𝜔𝑡\bm{T}(t)=\bm{T_{0}}\cos\,\omega tbold_italic_T ( italic_t ) = bold_italic_T start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT roman_cos italic_ω italic_t, collocated with the force dipole. The torque at the back represents the stroke of the swimming organ and causes the vortex street Drucker and Lauder (2002) in the fish’s wake at high Re𝑅𝑒Reitalic_R italic_e. An opposite torque is applied in the body (Fig. 1(a)), and represents the counter-reaction of the rest of the body. For practical reasons in the scaling analyses below, we also introduce force and torque densities f0F0/L3subscript𝑓0subscript𝐹0superscript𝐿3f_{0}\equiv F_{0}/L^{3}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≡ italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT and τ0T0/L3subscript𝜏0subscript𝑇0superscript𝐿3\tau_{0}\equiv T_{0}/L^{3}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≡ italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_L start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT respectively. Averaging these dipoles over one period of time (2π/ω2𝜋𝜔2\pi/\omega2 italic_π / italic_ω) results in a simple static force dipole {𝑭𝟎,𝑭𝟎}subscript𝑭0subscript𝑭0\{\bm{-F_{0}},\bm{F_{0}}\}{ bold_- bold_italic_F start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT , bold_italic_F start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT } (Fig. 1(b)), while the average torque cancels out. This time averaging over one beating period is very similar to models of pushers and pullers beating at low Re𝑅𝑒Reitalic_R italic_e Hernandez et al. (2005); Mehandia and Nott (2008); Jibuti et al. (2014).

Scaling laws. The hydrodynamic nature of our model allows for simple scaling arguments, inspired by Gazzola et al. (2014) for inertial flows but translated to a more generic framework and extended to the non-inertial Stokes regime. In the following, we present 3D3𝐷3D3 italic_D arguments, but they remain valid in 2D2𝐷2D2 italic_D (see SM ven (a)). We also consider that all the lengths scale as L𝐿Litalic_L.

The body of the swimmer is submitted to different dominant drag forces depending on the Reynolds number. To describe this effect across all swimming regimes, we introduce the thrust number Th𝑇Thitalic_T italic_h as the ratio between the applied force density f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT multiplied by inertial forces ρ|D𝐯/Dt|ρv2L1similar-to𝜌𝐷𝐯𝐷𝑡𝜌superscript𝑣2superscript𝐿1\rho\lvert D\mathbf{v}/Dt\rvert\sim\rho v^{2}L^{-1}italic_ρ | italic_D bold_v / italic_D italic_t | ∼ italic_ρ italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_L start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, and the square of viscous forces |ηΔ𝐯|2(ηvL2)2similar-tosuperscript𝜂Δ𝐯2superscript𝜂𝑣superscript𝐿22\lvert\eta\Delta\mathbf{v}\rvert^{2}\sim(\eta vL^{-2})^{2}| italic_η roman_Δ bold_v | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ∼ ( italic_η italic_v italic_L start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, which gives

Thρf0L3η2.𝑇𝜌subscript𝑓0superscript𝐿3superscript𝜂2Th\equiv\frac{\rho f_{0}L^{3}}{\eta^{2}}.italic_T italic_h ≡ divide start_ARG italic_ρ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG start_ARG italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (1)

The thrust number appears naturally at all scales of Reynolds numbers, as shown below. It contains the force term at the origin of the motion, which is characterized by the Reynolds number. It therefore provides a convenient method for evaluating velocity as a function of force. Let us consider classic scaling arguments:

  • At high Reynolds numbers, the boundary layer around the body is turbulent and pressure drag dominates. The corresponding force scales as f0L3ρv2L2similar-tosubscript𝑓0superscript𝐿3𝜌superscript𝑣2superscript𝐿2f_{0}L^{3}\sim\rho v^{2}L^{2}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ∼ italic_ρ italic_v start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Since vηRe/(ρL)similar-to𝑣𝜂𝑅𝑒𝜌𝐿v\sim\eta Re/(\rho L)italic_v ∼ italic_η italic_R italic_e / ( italic_ρ italic_L ), we obtain Re(ρf0L3/η2)1/2=Th1/2similar-to𝑅𝑒superscript𝜌subscript𝑓0superscript𝐿3superscript𝜂212𝑇superscript12Re\sim(\rho f_{0}L^{3}/\eta^{2})^{1/2}=Th^{1/2}italic_R italic_e ∼ ( italic_ρ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT = italic_T italic_h start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT.

  • For small but finite Reynolds numbers, the regime is laminar and the viscous force in the boundary layer dominates: f0L3(ηv/δ)L2similar-tosubscript𝑓0superscript𝐿3𝜂𝑣𝛿superscript𝐿2f_{0}L^{3}\sim(\eta v/\delta)L^{2}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ∼ ( italic_η italic_v / italic_δ ) italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT where δ𝛿\deltaitalic_δ is the thickness of the boundary layer, which obeys the Blasius law Landau and Lifshitz (2013) δLRe1/2similar-to𝛿𝐿𝑅superscript𝑒12\delta\sim L\,Re^{-1/2}italic_δ ∼ italic_L italic_R italic_e start_POSTSUPERSCRIPT - 1 / 2 end_POSTSUPERSCRIPT. This finally leads to ReTh2/3similar-to𝑅𝑒𝑇superscript23Re\sim Th^{2/3}italic_R italic_e ∼ italic_T italic_h start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT.

  • At low Reynolds numbers, the Stokes drag force dominates and the force applied on the body compensates the drag: f0L3ηvLsimilar-tosubscript𝑓0superscript𝐿3𝜂𝑣𝐿f_{0}L^{3}\sim\eta vLitalic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ∼ italic_η italic_v italic_L. This gives Reρf0L3/η2=Thsimilar-to𝑅𝑒𝜌subscript𝑓0superscript𝐿3superscript𝜂2𝑇Re\sim\rho f_{0}L^{3}/\eta^{2}=Thitalic_R italic_e ∼ italic_ρ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / italic_η start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_T italic_h.

From the Stokes to the turbulent regime, we observe that the exponent α𝛼\alphaitalic_α of the scaling ReThαsimilar-to𝑅𝑒𝑇superscript𝛼Re\sim Th^{\alpha}italic_R italic_e ∼ italic_T italic_h start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT is always below one and decreases. It suggests a diminishing swimming performance as the Reynolds number of the swimmer increases. To confirm these three successive regimes, we present below numerical simulations with our swimmer model, exploring a large range of values for the thrust and Reynolds numbers.

Numerical simulations.

We performed direct numerical simulations of the incompressible Navier-Stokes equations in the presence of our model swimmer, a rigid ellipsoid body with the force and torque dipoles attached. The corresponding fluid momentum balance equation writes:

ρD𝒗Dt(2η𝑬(𝒗))+p=𝓕+𝓣𝜌𝐷𝒗𝐷𝑡2𝜂𝑬𝒗𝑝𝓕𝓣\displaystyle\rho\frac{D\bm{v}}{Dt}-\nabla\cdot\left(2\eta\bm{{E}}(\bm{{v}})% \right)+\nabla p=\bm{{\mathcal{F}}}+\bm{{\mathcal{T}}}italic_ρ divide start_ARG italic_D bold_italic_v end_ARG start_ARG italic_D italic_t end_ARG - ∇ ⋅ ( 2 italic_η bold_italic_E ( bold_italic_v ) ) + ∇ italic_p = bold_caligraphic_F + bold_caligraphic_T (2)

where 𝒗𝒗\bm{{v}}bold_italic_v and p𝑝pitalic_p are respectively the velocity and pressure fields, D/Dt𝐷𝐷𝑡D/Dtitalic_D / italic_D italic_t denotes the material derivative D𝒗/Dt𝒗/t+(𝒗)𝒗𝐷𝒗𝐷𝑡𝒗𝑡bold-⋅𝒗bold-∇𝒗D\bm{{v}}/Dt\equiv\partial\bm{{v}}/\partial t+(\bm{v\cdot\nabla})\bm{v}italic_D bold_italic_v / italic_D italic_t ≡ ∂ bold_italic_v / ∂ italic_t + ( bold_italic_v bold_⋅ bold_∇ ) bold_italic_v, 𝑬(𝒗)(𝒗+𝒗t)/2𝑬𝒗bold-∇𝒗bold-∇superscript𝒗𝑡2\bm{{E}}(\bm{{v}})\equiv(\bm{\nabla v}+\bm{\nabla v}^{t})/2bold_italic_E ( bold_italic_v ) ≡ ( bold_∇ bold_italic_v + bold_∇ bold_italic_v start_POSTSUPERSCRIPT italic_t end_POSTSUPERSCRIPT ) / 2 is the strain-rate tensor, ρ𝜌\rhoitalic_ρ and η𝜂\etaitalic_η denote the density and viscosity fields, and 𝓕(t)𝓕𝑡\bm{{\mathcal{F}}}(t)bold_caligraphic_F ( italic_t ) and 𝓣(t)𝓣𝑡\bm{{\mathcal{T}}}(t)bold_caligraphic_T ( italic_t ) respectively represent the – time-varying – force and torque dipoles attached to the swimmer, as introduced above. The rigid body \mathcal{B}caligraphic_B of the swimmer is accounted for with a fictitious domain penalty method inspired by (Janela et al., 2005); in practice, this can be implemented simply with a spatially-variable viscosity (Tanaka and Araki, 2000): η=ηf+(ηbηf)H𝜂subscript𝜂𝑓subscript𝜂𝑏subscript𝜂𝑓subscript𝐻\eta=\eta_{f}+(\eta_{b}-\eta_{f})\,H_{\mathcal{B}}italic_η = italic_η start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT + ( italic_η start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT - italic_η start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT ) italic_H start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT where Hsubscript𝐻H_{\mathcal{B}}italic_H start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT is the indicator – or Heaviside – function of \mathcal{B}caligraphic_B. In practice, a viscosity ratio ηb/ηf=103106subscript𝜂𝑏subscript𝜂𝑓superscript103superscript106\eta_{b}/\eta_{f}=10^{3}-10^{6}italic_η start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT / italic_η start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT = 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT is applied between the fluid and the swimmer body to ensure that the rigid motion constraint 𝑬(𝒗)=0𝑬𝒗0\bm{{E}}(\bm{{v}})=0bold_italic_E ( bold_italic_v ) = 0 is satisfied within \mathcal{B}caligraphic_B. Density ρ𝜌\rhoitalic_ρ is defined as constant inside and outside \mathcal{B}caligraphic_B, thus making the swimmer neutrally buoyant. Note that this fictitious domain penalty method allows the swimmer to be treated directly as part of the Navier-Stokes equations through the viscosity field, thereby avoiding the need to deal with moving boundary conditions and potential remeshing issues at the body interface in the discrete setting. The incrompressible Navier-Stokes equations are solved numerically using an implicit 𝒫2𝒫1𝒫2𝒫1\mathcal{P}2-\mathcal{P}1caligraphic_P 2 - caligraphic_P 1 finite element method (Metivet et al., 2018) implemented in the parallel FEEL++ library Prud’homme et al. (2012). The position and orientation of the swimmer are updated at each time-step using a first order Euler scheme with the translational and rotational velocities computed from the fluid velocity field in \mathcal{B}caligraphic_B. A comprehensive derivation of the numerical model and technical details are provided in SM ven (a).

This numerical framework is used to explore a large range of Reynolds numbers (105<Re<104superscript105𝑅𝑒superscript10410^{-5}<Re<10^{4}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT < italic_R italic_e < 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT) and Thrust numbers (103<Th<106)10^{-3}<Th<10^{6})10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT < italic_T italic_h < 10 start_POSTSUPERSCRIPT 6 end_POSTSUPERSCRIPT ), in order to evaluate the dependence of Re𝑅𝑒Reitalic_R italic_e as a function of Th𝑇Thitalic_T italic_h while varying each of the model’s different parameters (L𝐿Litalic_L, ηfsubscript𝜂𝑓\eta_{f}italic_η start_POSTSUBSCRIPT italic_f end_POSTSUBSCRIPT, f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, ω𝜔\omegaitalic_ω and τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT) separately (see SM ven (a)). As shown in Fig. 2, the numerical simulations are in perfect agreement with the scaling laws presented above, displaying the Stokes regime in the range 105Re102less-than-or-similar-tosuperscript105𝑅𝑒less-than-or-similar-tosuperscript10210^{-5}\lesssim Re\lesssim 10^{2}10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT ≲ italic_R italic_e ≲ 10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, the laminar regime for 102Re103less-than-or-similar-tosuperscript102𝑅𝑒less-than-or-similar-tosuperscript10310^{2}\lesssim Re\lesssim 10^{3}10 start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≲ italic_R italic_e ≲ 10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT and the turbulent regime for 103Reless-than-or-similar-tosuperscript103𝑅𝑒10^{3}\lesssim Re10 start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ≲ italic_R italic_e,

Refer to caption
Figure 3: 2D Vorticity field for different Re𝑅𝑒Reitalic_R italic_e. The simulations were performed with ω=2π𝜔2𝜋\omega=2\piitalic_ω = 2 italic_π, L=32𝐿32L=32italic_L = 32 and 𝑭𝟎=𝑻𝟎normsubscript𝑭0normsubscript𝑻0\|\bm{{F_{0}}}\|=\|\bm{{T_{0}}}\|∥ bold_italic_F start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT ∥ = ∥ bold_italic_T start_POSTSUBSCRIPT bold_0 end_POSTSUBSCRIPT ∥. (a): 𝑻0=4000delimited-∥∥subscript𝑻04000\lVert\bm{T}_{0}\rVert=4000∥ bold_italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∥ = 4000 and Re=960𝑅𝑒960Re=960italic_R italic_e = 960. (b): 𝑻0=600delimited-∥∥subscript𝑻0600\lVert\bm{T}_{0}\rVert=600∥ bold_italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∥ = 600 and Re=300𝑅𝑒300Re=300italic_R italic_e = 300. (c): 𝑻0=100delimited-∥∥subscript𝑻0100\lVert\bm{T}_{0}\rVert=100∥ bold_italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∥ = 100 and Re=50𝑅𝑒50Re=50italic_R italic_e = 50. (d): 𝑻0=0.25delimited-∥∥subscript𝑻00.25\lVert\bm{T}_{0}\rVert=0.25∥ bold_italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∥ = 0.25 and Re=0.1𝑅𝑒0.1Re=0.1italic_R italic_e = 0.1.

with fitting exponents that match the predictions. Note that the Re𝑅𝑒Reitalic_R italic_e ranges of each regime depend on the aspect ratio of the swimmer arsubscript𝑎𝑟a_{r}italic_a start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT, which is kept constant. We also found that ω𝜔\omegaitalic_ω and τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT do not play any role in the Re(Th)𝑅𝑒𝑇Re(Th)italic_R italic_e ( italic_T italic_h ) dependency, which confirms that all the important parameters are embedded in the Th𝑇Thitalic_T italic_h number.

Wake and vortices. Although torque plays no role in the scaling of Re𝑅𝑒Reitalic_R italic_e as a function of Th𝑇Thitalic_T italic_h, it is essential to reproduce the wake at the rear of the swimmer in the inertial regime. The torque dipole can be used to account for flagellum, body undulation or tail beating to generate a reverse von Karman vortex street, as observed in the wake of a fish at high Re𝑅𝑒Reitalic_R italic_e number Kern and Koumoutsakos (2006); Ko et al. (2023).

Figures 3(a,b) illustrate the typical wakes obtained with (T0,F0)=(4000,4000),(600,600)subscript𝑇0subscript𝐹040004000600600(T_{0},F_{0})=(4000,4000),(600,600)( italic_T start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_F start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = ( 4000 , 4000 ) , ( 600 , 600 ) in 2D simulations.

Refer to caption
Figure 4: Reynolds number Re𝑅𝑒Reitalic_R italic_e as a function of the swimming number Sw𝑆𝑤Switalic_S italic_w. Crosses are new experimental data available in the database ven (b). Dots are experimental data collected by Gazzola et al. (2014). The three lines correspond to fitted curves and give the numerical scaling exponents of the three different regimes: ReSw0.98similar-to𝑅𝑒𝑆superscript𝑤0.98Re\sim Sw^{0.98}italic_R italic_e ∼ italic_S italic_w start_POSTSUPERSCRIPT 0.98 end_POSTSUPERSCRIPT for Re<10𝑅𝑒10Re<10italic_R italic_e < 10; ReSw1.33similar-to𝑅𝑒𝑆superscript𝑤1.33Re\sim Sw^{1.33}italic_R italic_e ∼ italic_S italic_w start_POSTSUPERSCRIPT 1.33 end_POSTSUPERSCRIPT for 10<Re<10410𝑅𝑒superscript10410<Re<10^{4}10 < italic_R italic_e < 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT; and ReSw1.0similar-to𝑅𝑒𝑆superscript𝑤1.0Re\sim Sw^{1.0}italic_R italic_e ∼ italic_S italic_w start_POSTSUPERSCRIPT 1.0 end_POSTSUPERSCRIPT for 104<Resuperscript104𝑅𝑒10^{4}<Re10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT < italic_R italic_e.

The vortices created at the same frequency are spaced further apart as the Reynolds number (i.e. velocity) increases from Re=100𝑅𝑒100Re=100italic_R italic_e = 100 to Re=960𝑅𝑒960Re=960italic_R italic_e = 960. The length over which they dissipate becomes shorter towards the viscous regimes (Fig. 3(c)), vanishing completely at low Re𝑅𝑒Reitalic_R italic_e (Fig. 3(d)). Indeed, no vortices are present behind micro-organisms Drescher et al. (2010).

Comparison with experimental data. To compare our scaling results with experimental data, forces must be expressed in terms of observable data (Fig. 4), such as undulation or tail beating frequency. Although force f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, torque τ0subscript𝜏0\tau_{0}italic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and pulsation ω𝜔\omegaitalic_ω are independent quantities in the model, physical constraints exist between these quantities in living organisms. We made the reasonable assumption that the size of the swimming organ scales with the size of the body L𝐿Litalic_L Gazzola et al. (2014). Force and torque generated by the swimming organ are such that τ0f0Lsimilar-tosubscript𝜏0subscript𝑓0𝐿\tau_{0}\sim f_{0}Litalic_τ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L.

In inertial regimes, i.e. excluding the Stokes regime that we address separately, the instantaneous force creates a transient acceleration of the fluid, which scales with Lω2𝐿superscript𝜔2L\omega^{2}italic_L italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, i.e. f0ρLω2similar-tosubscript𝑓0𝜌𝐿superscript𝜔2f_{0}\sim\rho L\omega^{2}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∼ italic_ρ italic_L italic_ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. Introducing the observation-based swimming number Sw=ρωL2/η𝑆𝑤𝜌𝜔superscript𝐿2𝜂Sw=\rho\omega L^{2}/\etaitalic_S italic_w = italic_ρ italic_ω italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / italic_η Gazzola et al. (2014), the theoretical force drive can be related to experimentally measurable data as ThSw2similar-to𝑇𝑆superscript𝑤2Th\sim Sw^{2}italic_T italic_h ∼ italic_S italic_w start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. The scaling laws previously derived from our model can thus be reformulated in terms of Sw𝑆𝑤Switalic_S italic_w: in the laminar regime ReTh2/3Sw4/3similar-to𝑅𝑒𝑇superscript23similar-to𝑆superscript𝑤43Re\sim Th^{2/3}\sim Sw^{4/3}italic_R italic_e ∼ italic_T italic_h start_POSTSUPERSCRIPT 2 / 3 end_POSTSUPERSCRIPT ∼ italic_S italic_w start_POSTSUPERSCRIPT 4 / 3 end_POSTSUPERSCRIPT, while in the turbulent regime ReTh1/2Swsimilar-to𝑅𝑒𝑇superscript12similar-to𝑆𝑤Re\sim Th^{1/2}\sim Switalic_R italic_e ∼ italic_T italic_h start_POSTSUPERSCRIPT 1 / 2 end_POSTSUPERSCRIPT ∼ italic_S italic_w, in accordance with the results of Gazzola et al. (2014).

Refer to caption
Figure 5: Strouhal number St=Aω/v𝑆𝑡𝐴𝜔𝑣St=A\omega/vitalic_S italic_t = italic_A italic_ω / italic_v as a function of the Reynolds number Re𝑅𝑒Reitalic_R italic_e. Dots correspond to the data collected in Gazzola et al. (2014), and crosses are new experimental data. The three lines correspond to guides for the eyes with the three different regimes: St2π=1𝑆𝑡2𝜋1\tfrac{St}{2\pi}=1divide start_ARG italic_S italic_t end_ARG start_ARG 2 italic_π end_ARG = 1 for Re<10𝑅𝑒10Re<10italic_R italic_e < 10; StRe0.25similar-to𝑆𝑡𝑅superscript𝑒0.25St\sim Re^{-0.25}italic_S italic_t ∼ italic_R italic_e start_POSTSUPERSCRIPT - 0.25 end_POSTSUPERSCRIPT for 10<Re<10410𝑅𝑒superscript10410<Re<10^{4}10 < italic_R italic_e < 10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT; and St2π0.3𝑆𝑡2𝜋0.3\tfrac{St}{2\pi}\approx 0.3divide start_ARG italic_S italic_t end_ARG start_ARG 2 italic_π end_ARG ≈ 0.3 for 104<Resuperscript104𝑅𝑒10^{4}<Re10 start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT < italic_R italic_e.

In the Stokes regime, the force exerted by the swimming organ is balanced by viscous drag, so that f0L3ηLvηL2ωsimilar-tosubscript𝑓0superscript𝐿3𝜂𝐿𝑣similar-to𝜂superscript𝐿2𝜔f_{0}L^{3}\sim\eta Lv\sim\eta L^{2}\omegaitalic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_L start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ∼ italic_η italic_L italic_v ∼ italic_η italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ω. Reintroducing again the swimming number Sw𝑆𝑤Switalic_S italic_w, we obtain ThSwsimilar-to𝑇𝑆𝑤Th\sim Switalic_T italic_h ∼ italic_S italic_w, leading to ReSwsimilar-to𝑅𝑒𝑆𝑤Re\sim Switalic_R italic_e ∼ italic_S italic_w. Note that this is a natural consequence of the absence of inertia: each stroke creates a net displacement that scales with Lsimilar-toabsent𝐿\sim L∼ italic_L, inducing a swimming velocity vLωsimilar-to𝑣𝐿𝜔v\sim L\omegaitalic_v ∼ italic_L italic_ω.

Figure 4 shows the experimental data collected from micrometer- to meter-size aquatic organisms along with the corresponding fitted scaling laws. In addition to the results from Gazzola et al. (2014), an excellent agreement between experimental data and hydrodynamic scaling laws is also obtained in the Stokes regime.

Note that vLωsimilar-to𝑣𝐿𝜔v\sim L\omegaitalic_v ∼ italic_L italic_ω in both the Stokes and turbulent regimes, so that the corresponding Strouhal number St=Sw/Re𝑆𝑡𝑆𝑤𝑅𝑒St=Sw/Reitalic_S italic_t = italic_S italic_w / italic_R italic_e is constant, as illustrated in Figure 5. The transition between St/2π0.3𝑆𝑡2𝜋0.3St/2\pi\approx 0.3italic_S italic_t / 2 italic_π ≈ 0.3 Gazzola et al. (2014) and St/2π1𝑆𝑡2𝜋1St/2\pi\approx 1italic_S italic_t / 2 italic_π ≈ 1 occurs in the laminar regime where StRe1/4similar-to𝑆𝑡𝑅superscript𝑒14St\sim Re^{-1/4}italic_S italic_t ∼ italic_R italic_e start_POSTSUPERSCRIPT - 1 / 4 end_POSTSUPERSCRIPT.

Conclusion. By avoiding direct fluid-structure coupling, our generic swimmer model provides an efficient model for hydrodynamic propulsion, while retaining the salient features of swimming organisms across several decades of Reynolds numbers. High numerical stability and efficiency ensure fully tractable 2D and 3D simulations, thereby paving the way to large scale simulations with hundreds of agents. It also proposes a methodology for progressive refinement of the hydrodynamic field, by retaining higher moments of forces and torques, resulting in more complex propulsion models. This approach broadens our understanding of the swimming of aquatic organisms by revealing the universal relationship between the velocity of a swimmer and the force exerted by its swimming organ. The sub-linear dependence demonstrated between Re𝑅𝑒Reitalic_R italic_e and Th𝑇Thitalic_T italic_h suggests diminishing swimming performance as the swimmer’s Reynolds number increases. The scaling laws obtained also match the experimental data obtained from thousands of aquatic animals, ranging from large mammalians to micro-organisms. Our results shed new light on the general mechanisms underlying swimming and provide an efficient and robust numerical framework to investigate the collective behavior of swimmers in complex environments.

Acknowledgements. This project received financial support from the French National Research Agency (ANR-21-CE45-0005, FISHSIF project).

References

  • Childress (1981) S. Childress, Mechanics of Swimming and Flying (Cambridge University Press, 1981).
  • Berg (2004) H. Berg, E.coli in motion (Springer, 2004).
  • S.Powar et al. (2022) S.Powar, F. Parast, A. Nandagiri, A. Gaikwad, D. Potter, M. O’Bryan, R. Prabhakar, J. Soria,  and R. Nosrati, Small Methods 6, e2101089 (2022).
  • Garcia et al. (2011) M. Garcia, S. Berti, P. Peyla,  and S. Rafaï, Phys. Rev. E 83, 035301 (2011).
  • Smits (2019) A. Smits, J. Fluid. Mech., Perspectives 874, 1 (2019).
  • Gray (1936) J. Gray, J. Exp. Biol. 13, 192 (1936).
  • Wolfgang et al. (1999) M. Wolfgang, J. Anderson, M. Grosenbaugh, D. Yue,  and M. Triantafyllou, J. Exp. Biol. 202, 4841 (1999).
  • Dabiri et al. (2006) J. O. Dabiri, S. P. Colin,  and J. H. Costello, J. Exp. Biol. 209, 17 (2006).
  • Müller et al. (2001) U. Müller, J. Smit, E. Stamhuis,  and J. Videler, J. Exp. Biol. 204, 2751 (2001).
  • Taylor (1951) G. I. Taylor, Proc. R. Soc. Lond. A 209, 447 (1951).
  • Spagnolie and Lauga (2010) S. Spagnolie and E. Lauga, Phys. of Fluids 22, 031901 (2010).
  • Lauga (2020) E. Lauga, Phys. Rev. Fluids 5, 123101 (2020).
  • Lighthill (1960) M. J. Lighthill, J. Fluid Mech. 9, 305 (1960).
  • Liu and Kawachi (1999) H. Liu and K. Kawachi, J. Comp. Phys. 155, 223 (1999).
  • Kern and Koumoutsakos (2006) S. Kern and P. Koumoutsakos, J. Exp. Biol. 209, 4841 (2006).
  • Lauga and Powers (2009) E. Lauga and T. R. Powers, Reports on Progess in Physics 72, 096601 (2009).
  • Purcell (1977) E. Purcell, American Journal of Physics 3–11, 1 (1977).
  • Hernandez et al. (2005) J. Hernandez, C. Stoltz,  and M. Graham, Phys. Rev. Lett. 95, 204501 (2005).
  • Mehandia and Nott (2008) V. Mehandia and P. Nott, J. Fluid Mech. 595, 239 (2008).
  • Jibuti et al. (2014) L. Jibuti, L. Qi, C. Misbah, W. Zimmermann, S. Rafaï,  and P. Peyla, Phys. Rev. E 90, 063019 (2014).
  • Li et al. (2021) G. Li, H. Liu, U. Müller, C. Voesenek,  and J. van Leeuwen, Proc. Royal Soc. 288, 1 (2021).
  • Drucker and Lauder (2002) E. G. Drucker and G. V. Lauder, Integrative and Comparative Biology 42, 243 (2002)https://academic.oup.com/icb/article-pdf/42/2/243/1841427/i1540-7063-042-02-0243.pdf .
  • Gazzola et al. (2014) M. Gazzola, M. Argentina,  and L. Mahadevan, Nature Physics 10, 758 (2014).
  • Feynman et al. (2006) R. Feynman, R. Leighton,  and M. Sands, The Feynman Lectures on Physics, Vol. 1 (Addison-Wesley, 2006).
  • Filella et al. (2018) A. Filella, F. Nadal, C. Sire, E. Kanso,  and C. Eloy, Phys. Rev. Lett. 120, 198101 (2018).
  • ven (a) “See supplementary material at [to-be-inserted-by-publisher],”  (a).
  • Landau and Lifshitz (2013) L. D. Landau and E. M. Lifshitz, Fluid mechanics: course of theoretical physics, Vol. 6 (Elsevier, 2013).
  • Janela et al. (2005) J. Janela, A. Lefebvre,  and B. Maury, in ESAIM: Proceedings, Vol. 14 (EDP Sciences, 2005) pp. 115–123.
  • Tanaka and Araki (2000) H. Tanaka and T. Araki, Phys. Rev. Lett. 85, 1338 (2000).
  • Metivet et al. (2018) T. Metivet, V. Chabannes, M. Ismail,  and C. Prud’homme, Mathematics 6, 203 (2018).
  • Prud’homme et al. (2012) C. Prud’homme, V. Chabannes, V. Doyeux, M. Ismail, A. Samake,  and G. Pena, ESAIM: Proc. 38, 429 (2012).
  • Ko et al. (2023) H. Ko, G. Lauder,  and R. Nagpal, J. R. Soc. Interface 20, 1 (2023).
  • ven (b) “Doi will be added after the peer review,”  (b).
  • Drescher et al. (2010) K. Drescher, R. E. Goldstein, N. Michel, M. Polin,  and I. Tuval, Phys. Rev. Lett. 105, 168101 (2010).