\UseRawInputEncoding

Revisiting tea-leaf paradox with a deformable spoon

Benjamin Apffel benjamin.apffel@epfl.ch    Romain Fleury Institute of Electrical and Micro Engineering, Laboratory of Wave Engineering, Ecole Polytechnique Federale de Lausanne (EPFL), Station 11, 1015 Lausanne, Switzerland
(July 4, 2024)
Abstract

The tea-leaf paradox refers to the migration of tea leaves toward the center of a tea cup when the water is put in rotation by a spoon. Although much attention has been given to liquid and particle transport generated by a spoon, the back-action of the generated flow on a deformable spoon has not been considered. In this work, we therefore propose to revisit the tea-leaf paradox with a deformable stirrer consisting of a floating ball attached to a soft string. The later is set in rotation and the ball trajectory results from the balance between string tension and the back action of the surrounding flow, leading to surprising effects. At low rotation speeds, we show that the circular trajectory of the ball shrinks when we stir faster, which contradicts common intuition based on inertia. For higher rotation rates, the ball is either suddenly attracted toward the center, or is repulsed away from it, depending on the string length. A model explaining these observations is proposed, and validated using experimental measurements of the generated flow. Interestingly, the system exhibits strong hysteretic behavior, leading to a modified version of the tea-leaf paradox in which the spoon itself can be trapped at the center in the flow generated by its past motion.

preprint: APS/123-QED

In its simplest version, the tea leaf paradox arises when mixing a liquid with a spoon performing circular trajectories (Fig. 1a). The paradox arises when one notices that particles placed at the bottom tend to aggregate at the center of the cup rather than on the outside, as expected from centrifugal force. The transport arises due to three-dimensional flow generated along the vertical direction by the edges of the container Thomson and Thomson (1997). Such flows imposed by boundary conditions are of prime interest from geophysics to physical chemistry and microfluidics, for example to understand the preferential concentration of sediments concentration near a river bank’s meanders (Einstein, 1926; Friedkin, 1945; Bowker, 2007; Leopold and Wolman, 1960; Jackson, 1975), dissolution mediated by rotating paddles McCarthy et al. (2003), for controlling the aggregation of nanoparticles Zhang et al. (2023) or blood-plasma separation (Yeo et al., 2006; Arifin et al., 2006).

The back action of the flow generated by rotating free-to-move objects has also been investigated. For instance, magnetic disks placed in rotation at a liquid interface were shown to interact and self-assemble under the generated flow’s action (Grzybowski et al., 2000, 2001, 2002), and interact with their environment (Gorce et al., 2019, 2021). More generally, the interplay between deformable bodies and flow generation is at the heart of propulsion mechanisms for living organisms (Gazzola et al., 2014; Taylor, 1997; Wu, 2011; Ashraf et al., 2017) or robots (Oliveira Santos et al., 2023; Esposito et al., 2012; Zhu et al., 2019), an idealized case of such situation being the flow disturbance generated by a flexible plate (Wu, 1961; Alben, 2021; Toomey and Eldredge, 2008).

Refer to caption
Figure 1: (a) Sketch of a rigid and a deformable spoon mixing a liquid. (b) Definition of geometrical parameters for the flexible stirrer consisting of a plastic ball of radius 7 mm pulled by an inextensible string attached to a rotation arm. (c) Experimental idealization of a rigid spoon (left) using a metalic stem, and its deformable counterpart made by replacing the stem by a wool strand (top right). The flow induced by the deformable stirrer transports particles (pepper grains) placed at the bottom near the center, showing that the tea leaf paradox still occurs in this case (bottom right).

Inspired by these works, we propose here to revisit the tea leaf paradox with a deformable spoon, and to focus on the motion of the later rather than on the flow (Fig. 1a). As an experimental model, we study the motion of a ball attached to one end of an inextensible string, the other end being attached to a rotating arm as in Fig. 1b-c. Imposing a circular trajectory results in the emergence of several unexpected phenomena that we describe in this work. First, the radius of the stirrer’s circular trajectory surprisingly decreases when the rotation rate increases. This holds until a critical rotation speed is reached, for which a strong bifurcation occurs and the ball is either suddenly attracted toward the center, or repelled outwards,depending on the string length. A model explaining these observations is developed, and we then show in particular that the lift force generated by the ball’s motion must be taken into account to satisfactory explain the result. A quantitative criteria for the appearance of the collapsed regime is derived, and we show that strong hysteresis is observed once the stirrer is trapped.

Refer to caption
Figure 2: (a) Example of experimentally measured trajectories for different rotation speed Ω=0.2,1.4,2Ω0.21.42\Omega=0.2,1.4,2roman_Ω = 0.2 , 1.4 , 2 round/s and two different string length Ls=4.6subscript𝐿𝑠4.6L_{s}=4.6italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 4.6 cm (top) and Ls=5.3subscript𝐿𝑠5.3L_{s}=5.3italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 5.3 cm (bottom). (b) Experimental kinogrammes showing a circular trajectory (top), a collapsed state (middle) and a diverging state (bottom). (c) Picture from below showing the loss of tension in the string when a collapsed state is reached. (d) Radius of the trajectory as a function of the rotation speed for different string length Lssubscript𝐿𝑠L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. For the shortest strings (Ls4.6subscript𝐿𝑠4.6L_{s}\leq 4.6italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≤ 4.6 cm), the ball reaches a diverging state while longest strings (Ls5subscript𝐿𝑠5L_{s}\geq 5italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≥ 5 cm) lead to collapsing states. (e) Motion of the ball in the laboratory or comoving frame and definition of the forces and angles. (f) Experimental measurement of tan(ϕ)italic-ϕ\tan(\phi)roman_tan ( italic_ϕ ) (with ϕitalic-ϕ\phiitalic_ϕ the rope angle (er,T)subscript𝑒𝑟𝑇(-\vec{e}_{r},\vec{T})( - over→ start_ARG italic_e end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT , over→ start_ARG italic_T end_ARG )) for different string length as a function of rotation speed ΩΩ\Omegaroman_Ω (left) and trajectory radius Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT (right). All data collapse in a single line of slope aexp=0.56subscript𝑎𝑒𝑥𝑝0.56a_{exp}=0.56italic_a start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT = 0.56 cm-1 (dashed line) while theory predicts ath=0.14subscript𝑎𝑡0.14a_{th}=0.14italic_a start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT = 0.14 cm-1 (dot line).

All experiments (except illustrations of Fig. 1) are conducted in a rectangular tank of 55×\times×55 cm2 and filled with tap water with depth 6.5 cm, preventing the ball to get close from any edge. A model rigid spoon has been built using a 3D-printed plastic ball of radius Rs=7subscript𝑅𝑠7R_{s}=7italic_R start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 7 mm and mass m=1.2𝑚1.2m=1.2italic_m = 1.2 g at the end of a metallic stem (Fig 1c). A stepper motor is used to set it up in rotation with rate Ω1similar-toΩ1\Omega~{}\sim~{}1roman_Ω ∼ 1 round/s, and the rotation arm length is fixed to W=4.4𝑊4.4W=4.4italic_W = 4.4 cm. This model spoon is made flexible by replacing the stem by a strand of wool of length Lssubscript𝐿𝑠L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and diameter 0.2absent0.2\leq~{}0.2≤ 0.2 mm as sketched in Fig. 1b. Such strand is, in our regimes, non-extensible, of negligible mass compared to the ball and free of static or plastic deformation after being constrained. As we aim to study here the ball’s motion at the fluid interface, the final version of the setup includes a small tube to guide the strand close from the water surface as in Fig. 1c. This allows to confine the deformable system in the horizontal plane while avoiding the introduction of constrains on the string. When the rotation is turned on, a short transient regime of duration less than 10s is observed during which the ball converges toward a circular trajectory or radius Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT. On a minute scale, articles placed at the bottom converge toward the center of the tank (see Fig. 1c and supplementary movies), showing that three dimensional flows and associated transport (’tea-leaf’ paradox) still occur with our model deformable spoon. In what follows, we will focus on the surprising properties of the stirrer’s motion.

A camera (Basler) placed below the tank records 10 images per revolution of the rotation arm (see Fig. 2c for a typical image). The ball’s position is retrieved on each image using a convolution algorithm detailed in the Supplementary Informations (SI). Examples of measured trajectories are shown in Fig. 2a for two string length Ls=4.6subscript𝐿𝑠4.6L_{s}=4.6italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 4.6 cm, Ls=5.3subscript𝐿𝑠5.3L_{s}=5.3italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 5.3 cm and different rotation speeds Ω=0.2,1.4Ω0.21.4\Omega=0.2,1.4roman_Ω = 0.2 , 1.4 and 2 round/s. The trajectories are circular in all cases but their radius Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT strongly depends on the rotation speed and the string length. Several remarkable features can be observed. First, when the rotation speed increases from 0.20.20.20.2 round/s to 1.41.41.41.4 round/s, the radius of the trajectory decreases. This is opposite from what one occurs when the ball hangs in the air, for which the increasing centrifugal force pushes the ball more on the outside. Second, for the case Ls=5.3subscript𝐿𝑠5.3L_{s}=5.3italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 5.3 cm, the ball reaches a trajectory with zero radius for Ω=2Ω2\Omega=2roman_Ω = 2 round/s, and the string goes from straight to curved shape (Fig. 2c). In this case, the ball simply spins at the center in what we will call a ’collapsed’ or ’self-trapped’ behavior. For a slightly shorter string Ls=4.6subscript𝐿𝑠4.6L_{s}=4.6italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 4.6 cm, the same rotation speed Ω=2Ω2\Omega=2roman_Ω = 2 round/s gives rise to opposite behavior as the radius of the trajectory brutally increases, in what we call a ’diverging’ behavior. Kinogrammes obtained by stacking images from the side are shown in Fig. 2b and display the three regimes discussed above (see also supplementary movies). Experiment snapshots in Fig. 2c show that the string is straight in non-collapsed cases, but gains some curvature in the collapsed state.

More systematic experiments allow to plot the trajectory’s radius as a function of rotation rate for different string length varying from 2.7 cm to 5.9 cm. The results are shown in Fig. 2d. Aside from the smallest values of string length Ls=2.7subscript𝐿𝑠2.7L_{s}=2.7italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 2.7 cm, two regimes can be identified. Starting at low rotation rate, increasing the later leads to diminishing the radius trajectory. This holds until a critical rotation rate, for which the system exhibits strong bifurcation and the radius brutally changes. Depending on the string length, the ball is either suddenly expelled to the outward or attracted in the inward of the circular trajectory, corresponding to diverging and collapsing behavior discussed above. The first case occurs for Ls4.6subscript𝐿𝑠4.6L_{s}\leq 4.6italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≤ 4.6 cm, while the second case is observed for Ls5subscript𝐿𝑠5L_{s}\geq 5italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ≥ 5 cm.

We now propose to perform an analysis of the forces acting on the ball in the comoving frame. We will restrain ourselves to the case where the string is tensed and the immersed volume constant. From experimental observations, this corresponds to the cases where the trajectory radius diminishes with ΩΩ\Omegaroman_Ω without being collapsed (see SI for details on data selection). For now, we will consider the inertial force Finsubscript𝐹𝑖𝑛\vec{F}_{in}over→ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT, the viscous force Fvsubscript𝐹𝑣\vec{F}_{v}over→ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT and the string tension T𝑇\vec{T}over→ start_ARG italic_T end_ARG sketched in Fig. 2e. We assume that the ball performs a circular motion of radius Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT with rotation speed ΩΩ\Omegaroman_Ω. The inertial force due to the acceleration writes Fin=3/2mRtΩ2ersubscript𝐹𝑖𝑛32𝑚subscript𝑅𝑡superscriptΩ2subscript𝑒𝑟\vec{F}_{in}=3/2mR_{t}\Omega^{2}\vec{e}_{r}over→ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT = 3 / 2 italic_m italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over→ start_ARG italic_e end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT where the 3/2 factor accounts for the added mass of fluid that needs to be pushed by the sphere. As Reynolds number Re𝑅𝑒Reitalic_R italic_e typically ranges from 500500500500 to 5000500050005000 in our experiments, we take a quadratic law for the viscous force Fv=1/2ρCDSim(RtΩ)2eθsubscript𝐹𝑣12𝜌subscript𝐶𝐷subscript𝑆𝑖𝑚superscriptsubscript𝑅𝑡Ω2subscript𝑒𝜃\vec{F}_{v}=1/2\rho C_{D}S_{im}(R_{t}\Omega)^{2}\vec{e}_{\theta}over→ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT = 1 / 2 italic_ρ italic_C start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT italic_S start_POSTSUBSCRIPT italic_i italic_m end_POSTSUBSCRIPT ( italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT roman_Ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over→ start_ARG italic_e end_ARG start_POSTSUBSCRIPT italic_θ end_POSTSUBSCRIPT with CD0.4subscript𝐶𝐷0.4C_{D}\approx 0.4italic_C start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT ≈ 0.4 is the dimensionless drag parameter Clanet (2015) and Sim0.8πR2subscript𝑆𝑖𝑚0.8𝜋superscript𝑅2S_{im}\approx 0.8\pi R^{2}italic_S start_POSTSUBSCRIPT italic_i italic_m end_POSTSUBSCRIPT ≈ 0.8 italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the immersed cross-sectional surface of the object in the direction of the motion.

As the ball is steady in the comoving frame, one gets T=FinFv𝑇subscript𝐹𝑖𝑛subscript𝐹𝑣\vec{T}=-\vec{F}_{in}-\vec{F}_{v}over→ start_ARG italic_T end_ARG = - over→ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT - over→ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT. In particular, the angle ϕ=(er,T)italic-ϕsubscript𝑒𝑟𝑇\phi=(-\vec{e}_{r},\vec{T})italic_ϕ = ( - over→ start_ARG italic_e end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT , over→ start_ARG italic_T end_ARG ) (see Fig.3a) can be expressed as

tan(ϕ)=FvFin=αthRtitalic-ϕsubscript𝐹𝑣subscript𝐹𝑖𝑛subscript𝛼𝑡subscript𝑅𝑡\tan(\phi)=\frac{F_{v}}{F_{in}}=\alpha_{th}R_{t}roman_tan ( italic_ϕ ) = divide start_ARG italic_F start_POSTSUBSCRIPT italic_v end_POSTSUBSCRIPT end_ARG start_ARG italic_F start_POSTSUBSCRIPT italic_i italic_n end_POSTSUBSCRIPT end_ARG = italic_α start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT (1)

where αth0.24CD/Rssubscript𝛼𝑡0.24subscript𝐶𝐷subscript𝑅𝑠\alpha_{th}\approx 0.24C_{D}/R_{s}italic_α start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT ≈ 0.24 italic_C start_POSTSUBSCRIPT italic_D end_POSTSUBSCRIPT / italic_R start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. Experimental measurements of tan(ϕd)subscriptitalic-ϕ𝑑\tan(\phi_{d})roman_tan ( italic_ϕ start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ) as a function of ΩΩ\Omegaroman_Ω are shown in Fig. 2f. All measurements collapse on a single curve tan(ϕ)=αexpRtitalic-ϕsubscript𝛼𝑒𝑥𝑝subscript𝑅𝑡\tan(\phi)=\alpha_{exp}R_{t}roman_tan ( italic_ϕ ) = italic_α start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT as shown in 2f, confirming the scaling law in Eq. 1. However, the measured experimental value αexp=0.56subscript𝛼𝑒𝑥𝑝0.56\alpha_{exp}=0.56italic_α start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT = 0.56 cm-1 is much larger than the predicted value αth=0.14subscript𝛼𝑡0.14\alpha_{th}=0.14italic_α start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT = 0.14 cm-1.

Refer to caption
Figure 3: (a) Stack of 15 consecutive images taken at 150 frame/s showing the motion of particles at the water surface in the laboratory (top) and comoving (botton) frame. Direction of lift force FLsubscript𝐹𝐿\vec{F}_{L}over→ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT and fluid motion is also indicated. (b) Experimental measurement of velocity field’s circulation using PIV for different contour sizes and three different rotation rate of the ball. Error bar show the standard deviation of the measurement over 40 consecutive images. (c) Circulation as a function of rotation speed and linear fit Γ=SMΩΓsubscript𝑆𝑀Ω\Gamma=S_{M}\Omegaroman_Γ = italic_S start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT roman_Ω with SM=5.4subscript𝑆𝑀5.4S_{M}=-5.4italic_S start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT = - 5.4 cm-2.

Such discrepancy can be solved by considering the additional lift force produced by the fluid. The later is generated by the complete solid rotation of the sphere during its revolution, shown in Fig. 2e, combined with non-zero relative velocity of the ball with respect to the fluid Falkovich (2018); Rubinow and Keller (1961). A quantitative characterization of the resulting surface flow is performed using particle image velocimetry (PIV) measurement with PIVLab (Thielicke and Stamhuis, 2014; Thielicke and Sonntag, 2021). The liquid surface was seeded with small particles (pepper grains) while water was made opaque using a bit of powder milk to ensure good contrast. A camera placed above the tank recorded 150 frames/s during a full revolution of the ball, ensuring small displacement of the ball and the particles between two images. A stack of 15 consecutive images is shown in Fig. 4a. The same stack in the co-moving frame shows the streamline’s bending by the ball. From this picture, one can expect a lift force directed toward the inward of the trajectory. Its magnitude can be estimated starting from the lift force on a cylinder of height D𝐷Ditalic_D in a two dimensional flow Falkovich (2018); Dupeux et al. (2010)

FL=ρDRtΩΓersubscript𝐹𝐿𝜌𝐷subscript𝑅𝑡ΩΓsubscript𝑒𝑟\vec{F}_{L}=\rho DR_{t}\Omega\Gamma\vec{e}_{r}over→ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_L end_POSTSUBSCRIPT = italic_ρ italic_D italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT roman_Ω roman_Γ over→ start_ARG italic_e end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT (2)

where Γ=v.dlformulae-sequenceΓcontour-integral𝑣𝑑𝑙\Gamma=\oint\vec{v}.\vec{d}lroman_Γ = ∮ over→ start_ARG italic_v end_ARG . over→ start_ARG italic_d end_ARG italic_l is the circulation taken around the object far from the boundary layer. The circulation at the water surface can be measured experimentally from PIV measurement. The contour we chose is a square of side L𝐿Litalic_L centered on the ball as shown in Fig. 3b. The circulation ΓΓ\Gammaroman_Γ for different contour size L𝐿Litalic_L is shown in the same figure for three rotation rates Ω=0.2,0.6,1.0Ω0.20.61.0\Omega=0.2,0.6,1.0roman_Ω = 0.2 , 0.6 , 1.0 rad/s. The error bars are computed from the standard deviation on the measurement over 40 consecutive frames, showing low dispersion of measurement along time. As the size of the contour increases, the circulation increases and eventually reaches a plateau for the largest contours, which is the circulation value ΓΓ\Gammaroman_Γ introduced in Eq. (2) Falkovich (2018). The later is shown as a function of the rotation rate in Fig. 3e and shows excellent agreement with a linear fit Γ=SMΩΓsubscript𝑆𝑀Ω\Gamma=S_{M}\Omegaroman_Γ = italic_S start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT roman_Ω with SM=5.4subscript𝑆𝑀5.4S_{M}=-5.4italic_S start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT = - 5.4 cm2. The lift force can therefore be expressed as FM=MRtΩ2ersubscript𝐹𝑀𝑀subscript𝑅𝑡superscriptΩ2subscript𝑒𝑟\vec{F}_{M}=-MR_{t}\Omega^{2}\vec{e}_{r}over→ start_ARG italic_F end_ARG start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT = - italic_M italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT over→ start_ARG italic_e end_ARG start_POSTSUBSCRIPT italic_r end_POSTSUBSCRIPT with M=ρDSM𝑀𝜌𝐷subscript𝑆𝑀M=\rho DS_{M}italic_M = italic_ρ italic_D italic_S start_POSTSUBSCRIPT italic_M end_POSTSUBSCRIPT homogeneous to a mass and D𝐷Ditalic_D a distance let as a free parameter for now. The lift force scales exactly as the inertial force but with an effective negative mass. One must therefore modify Eq. 1 as

tan(ϕ)=ath12M/3mRtitalic-ϕsubscript𝑎𝑡12𝑀3𝑚subscript𝑅𝑡\tan(\phi)=\frac{a_{th}}{1-2M/3m}R_{t}roman_tan ( italic_ϕ ) = divide start_ARG italic_a start_POSTSUBSCRIPT italic_t italic_h end_POSTSUBSCRIPT end_ARG start_ARG 1 - 2 italic_M / 3 italic_m end_ARG italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT (3)
Refer to caption
Figure 4: (a) Angle ϕitalic-ϕ\phiitalic_ϕ as a function of the trajectory radius Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT for different string length Lssubscript𝐿𝑠L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT according to Eq. (4) (color lines) and to force equilibrium of Eq. (1) (dashed line). Either zero or two equilibrium angles fulfill both constraints depending on Lssubscript𝐿𝑠L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, and only the largest one is stable in the later case (red crosses). (b) Increasing string elongation ΔLsΔsubscript𝐿𝑠\Delta L_{s}roman_Δ italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT measured experimentally as a function of ΩΩ\Omegaroman_Ω of different string length. (c) Elongation of the string in the horizontal plane due to increasing tension with ΩΩ\Omegaroman_Ω. (d) Radius of the equilibrium trajectories when the string is elongated from Lssubscript𝐿𝑠L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT to Ls+ΔLssubscript𝐿𝑠Δsubscript𝐿𝑠L_{s}+\Delta L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT + roman_Δ italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. (e) Radius of the trajectory for the diverging (left, Ls=4.6subscript𝐿𝑠4.6L_{s}=4.6italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 4.6 cm) and collapsing (right, Ls=5.0subscript𝐿𝑠5.0L_{s}=5.0italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 5.0 cm) when the rotation speed is progressively increased (blue circles) or decreased (red squares) (f) Critical rotation speeds ΩusubscriptΩ𝑢\Omega_{u}roman_Ω start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT and ΩdsubscriptΩ𝑑\Omega_{d}roman_Ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT as a function of string length Lssubscript𝐿𝑠L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT.

The linear relation ΓΩproportional-toΓΩ\Gamma\propto\Omegaroman_Γ ∝ roman_Ω ensures that both equations (1) and (3) have similar dependency with Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT but with an increasing prefactor in the second case, which is perfectly consistent with our results. Using the found value αexpsubscript𝛼𝑒𝑥𝑝\alpha_{exp}italic_α start_POSTSUBSCRIPT italic_e italic_x italic_p end_POSTSUBSCRIPT in Eq. (3) fixes the only free parameter of our model D=0.3𝐷0.3D=0.3italic_D = 0.3 cm. For a purely two-dimensional flow, D𝐷Ditalic_D should exactly be the penetration length of the sphere in water, which is 1.01.01.01.0 cm. The mismatch between the two can be attributed to the spherical shape of our object as well as to possible three-dimensional flows generation, in line with the particle’s motion discussed in Fig. 1b.

So far, we have shown that our experimental measurement were consistent with the force analysis expressed in Eq. (1). Nevertheless, this is not enough to predict the trajectories radius Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT shown in Fig. 2d. The later can be fully determined if one assumes that the string is straight. In this case, the triangle formed by the ball, the hooking point and the rotation center (Fig. 4a) imposes an extra relationship between the geometrical parameters that reads

ϕ=arccos(Ls2+Rt2W22LsRt)italic-ϕsuperscriptsubscript𝐿𝑠2superscriptsubscript𝑅𝑡2superscript𝑊22subscript𝐿𝑠subscript𝑅𝑡\phi=\arccos\left(\frac{L_{s}^{2}+R_{t}^{2}-W^{2}}{2L_{s}R_{t}}\right)italic_ϕ = roman_arccos ( divide start_ARG italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_W start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT end_ARG ) (4)

A plot of this geometrical constrain is shown in Fig.4a for Ls=5,5.2subscript𝐿𝑠55.2L_{s}=5,5.2italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 5 , 5.2 and 5.45.45.45.4 cm (plain colored lines), while force equilibrium from Eq. (1) is represented by the black dashed line. For smallest string length, two solutions (ϕ,Rt)italic-ϕsubscript𝑅𝑡(\phi,R_{t})( italic_ϕ , italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT ) are found while no solution exists when Lssubscript𝐿𝑠L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT is too large. Among the two solutions, one can show that only the largest one materialized by red crosses in Fig. 4a is stable (see SI). This finally fully determines the radius trajectory for a given string length.

The last step comes by noticing that the string length in the interface plane (also denoted Lssubscript𝐿𝑠L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT) increases with the rotation rate. This is shown by experimental measurement of the string elongation ΔLs=Ls(Ω)Ls(Ω0)Δsubscript𝐿𝑠subscript𝐿𝑠Ωsubscript𝐿𝑠subscriptΩ0\Delta L_{s}=L_{s}(\Omega)-L_{s}(\Omega_{0})roman_Δ italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( roman_Ω ) - italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT ( roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) in Fig. 4b, which is by definition zero for Ω0=0.2subscriptΩ00.2\Omega_{0}=0.2roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.2 round/s. We observed experimentally that the string’s edges tends to become more horizontal when ΩΩ\Omegaroman_Ω increases, as sketched in Fig. 4c (see also pictures in SI). The tension increase, which scales as TΩ2proportional-to𝑇superscriptΩ2T\propto\Omega^{2}italic_T ∝ roman_Ω start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, leads to three-dimensional reorganization of the inextensible string and to the apparent flexibility in the horizontal plane. It is this varying string length in the horizontal plane that should be taken into account for the geometrical constrain (4). The resulting equilibrium radius Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT therefore depends on the elongation ΔLsΔsubscript𝐿𝑠\Delta L_{s}roman_Δ italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT and is plotted in Fig. 4d for three initial string length and various elongations. The radius Rtsubscript𝑅𝑡R_{t}italic_R start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT progressively decreases when the later increases, which is consistent with our experimental observations. The collapse of the system can also be interpreted as the loss of solutions discussed in Fig. 4a when the string becomes too long. In this case, the ball’s equilibrium and the string tension cannot be guaranteed simultaneously, which leads to a shape change of the rope. The later occurs for smaller values of elongation when the initial string length is larger. Last, the predicted value for the radius before collapse is approximately 2 cm and independent of Lssubscript𝐿𝑠L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. All those features are remarkably consistent with our experimental results of Fig. 2d.

A last aspect of the ball’s motion is the stability of the collapsed state. Starting from a collapsed state obtained for Ls=5subscript𝐿𝑠5L_{s}=5italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT = 5 cm and Ωu=2.6subscriptΩ𝑢2.6\Omega_{u}=2.6roman_Ω start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT = 2.6 round/s (blue circles in Fig. 4e), we progressively decrease the rotation speed. Interestingly, the ball remains trapped until ΩΩ\Omegaroman_Ω goes below Ωd=1.2subscriptΩ𝑑1.2\Omega_{d}=1.2roman_Ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT = 1.2 round/s (red squares), for which circular trajectories appear again. Such hysteresis was observed in all collapsing configurations, but not in the diverging case as shown in Fig. 4f. We plot in Fig. 4g the critical rotation speeds ΩusubscriptΩ𝑢\Omega_{u}roman_Ω start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT and ΩdsubscriptΩ𝑑\Omega_{d}roman_Ω start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT for various string length Lssubscript𝐿𝑠L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT. Their difference decreases with Lssubscript𝐿𝑠L_{s}italic_L start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT, showing that the hysteresis is stronger near the transition between diverging and collapsing behavior for a reason that is still to be understood. The hysteresis is due to the existence of a memory in the system, which can be qualitatively interpreted in two ways. First, the tension lost in the string needs to be recovered to expel the ball from the flow’s center. Second, the surrounding flow exhibits circular symmetry that needs to be broken to recover circular trajectories. Both features require a strong perturbation to be broken, leading to a robust hysteresis.

We have discussed in this work the circular motion of a ball pulled by a string and show the prime importance of the string’s deformation and flow’s action to explain the variety of observed trajectories. This opens several interesting avenues for the future. Our results first strongly suggests to perform the same experiments with several objects to introduce pairwise interactions mediated by the flow as done in other systems Grzybowski et al. (2000). It would also be interesting to replace water with a non-newtonian fluid, which were for instance shown to modify swimmers efficiency Espinosa-Garcia et al. (2013). Last, replacing the inextensible rope by a flexible one could also lead to new physical features that are still to explore.

Acknowledgements.
The authors acknowledge J.B. Gorce for fruitfull discussions and M. Mallejac for his help with Blender.

References

  • Thomson and Thomson (1997) James Thomson and William Thomson, “V. On the origin of windings of rivers in alluvial plains, with remarks on the flow of water round bends in pipes,” Proceedings of the Royal Society of London 25, 5–8 (1997), publisher: Royal Society.
  • Einstein (1926) A. Einstein, “Die Ursache der Mäanderbildung der Flußläufe und des sogenannten Baerschen Gesetzes,” Naturwissenschaften 14, 223–224 (1926).
  • Friedkin (1945) J.F. Friedkin, A Laboratory Study of the Meandering of Alluvial Rivers (United States Waterways experiment station, 1945) google-Books-ID: d4HuAAAAMAAJ.
  • Bowker (2007) Kent Bowker, “Albert Einstein and Meandering Rivers,” Earth Sciences History 7, 45 (2007).
  • Leopold and Wolman (1960) L B Leopold and M. G. Wolman, “River Meanders,” GSA Bulletin 71, 769–793 (1960).
  • Jackson (1975) R G. Jackson, II, “Velocity–bed-form–texture patterns of meander bends in the lower Wabash River of Illinois and Indiana,” GSA Bulletin 86, 1511–1522 (1975).
  • McCarthy et al. (2003) Leonard G. McCarthy, Carolin Kosiol, Anne Marie Healy, Geoff Bradley, James C. Sexton,  and Owen I. Corrigan, “Simulating the hydrodynamic conditions in the united states pharmacopeia paddle dissolution apparatus,” AAPS PharmSciTech 4, 22 (2003).
  • Zhang et al. (2023) Zehui Zhang, Bin Zhou, Mingtao Jia, Chengbin Wu, Tingting Niu, Chen Feng, Hongqiang Wang, Yanfeng Liu, Jialu Lu, Zhihua Zhang, Jun Shen,  and Ai Du, “Einstein’s tea leaf paradox induced localized aggregation of nanoparticles and their conversion to gold aerogels,” Science Advances 9, eadi9108 (2023), publisher: American Association for the Advancement of Science.
  • Yeo et al. (2006) Leslie Y. Yeo, James R. Friend,  and Dian R. Arifin, “Electric tempest in a teacup: The tea leaf analogy to microfluidic blood plasma separation,” Applied Physics Letters 89, 103516 (2006).
  • Arifin et al. (2006) Dian R. Arifin, Leslie Y. Yeo,  and James R. Friend, “Microfluidic blood plasma separation via bulk electrohydrodynamic flows,” Biomicrofluidics 1, 014103 (2006).
  • Grzybowski et al. (2000) Bartosz A. Grzybowski, Howard A. Stone,  and George M. Whitesides, “Dynamic self-assembly of magnetized, millimetre-sized objects rotating at a liquid–air interface,” Nature 405, 1033–1036 (2000), publisher: Nature Publishing Group.
  • Grzybowski et al. (2001) Bartosz A. Grzybowski, Xingyu Jiang, Howard A. Stone,  and George M. Whitesides, “Dynamic, self-assembled aggregates of magnetized, millimeter-sized objects rotating at the liquid-air interface: Macroscopic, two-dimensional classical artificial atoms and molecules,” Physical Review E 64, 011603 (2001), publisher: American Physical Society.
  • Grzybowski et al. (2002) Bartosz A. Grzybowski, Howard A. Stone,  and George M. Whitesides, “Dynamics of self assembly of magnetized disks rotating at the liquid–air interface,” Proceedings of the National Academy of Sciences 99, 4147–4151 (2002), publisher: Proceedings of the National Academy of Sciences.
  • Gorce et al. (2019) Jean-Baptiste Gorce, Hua Xia, Nicolas Francois, Horst Punzmann, Gregory Falkovich,  and Michael Shats, “Confinement of surface spinners in liquid metamaterials,” Proceedings of the National Academy of Sciences 116, 25424–25429 (2019), publisher: Proceedings of the National Academy of Sciences.
  • Gorce et al. (2021) Jean-Baptiste Gorce, Konstantin Y. Bliokh, Hua Xia, Nicolas Francois, Horst Punzmann,  and Michael Shats, “Rolling spinners on the water surface,” Science Advances 7, eabd4632 (2021), publisher: American Association for the Advancement of Science.
  • Gazzola et al. (2014) Mattia Gazzola, Médéric Argentina,  and L. Mahadevan, “Scaling macroscopic aquatic locomotion,” Nature Physics 10, 758–761 (2014), publisher: Nature Publishing Group.
  • Taylor (1997) Geoffrey Ingram Taylor, “Analysis of the swimming of microscopic organisms,” Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences 209, 447–461 (1997), publisher: Royal Society.
  • Wu (2011) Theodore Yaotsu Wu, “Fish Swimming and Bird/Insect Flight,” Annual Review of Fluid Mechanics 43, 25–58 (2011), publisher: Annual Reviews.
  • Ashraf et al. (2017) Intesaaf Ashraf, Hanaé Bradshaw, Thanh-Tung Ha, José Halloy, Ramiro Godoy-Diana,  and Benjamin Thiria, “Simple phalanx pattern leads to energy saving in cohesive fish schooling,” Proceedings of the National Academy of Sciences 114, 9599–9604 (2017), publisher: Proceedings of the National Academy of Sciences.
  • Oliveira Santos et al. (2023) Sara Oliveira Santos, Nils Tack, Yunxing Su, Francisco Cuenca-Jiménez, Oscar Morales-Lopez, P. Antonio Gomez-Valdez,  and Monica M. Wilhelmus, “Pleobot: a modular robotic solution for metachronal swimming,” Scientific Reports 13, 9574 (2023), publisher: Nature Publishing Group.
  • Esposito et al. (2012) Christopher J. Esposito, James L. Tangorra, Brooke E. Flammang,  and George V. Lauder, “A robotic fish caudal fin: effects of stiffness and motor program on locomotor performance,” The Journal of Experimental Biology 215, 56–67 (2012).
  • Zhu et al. (2019) J. Zhu, C. White, D. K. Wainwright, V. Di Santo, G. V. Lauder,  and H. Bart-Smith, “Tuna robotics: A high-frequency experimental platform exploring the performance space of swimming fishes,” Science Robotics 4, eaax4615 (2019), publisher: American Association for the Advancement of Science.
  • Wu (1961) T. Yao-Tsu Wu, “Swimming of a waving plate,” Journal of Fluid Mechanics 10, 321–344 (1961).
  • Alben (2021) Silas Alben, “Collective locomotion of two-dimensional lattices of flapping plates. Part 1. Numerical method, single-plate case and lattice input power,” Journal of Fluid Mechanics 915, A20 (2021).
  • Toomey and Eldredge (2008) Jonathan Toomey and Jeff D. Eldredge, “Numerical and experimental study of the fluid dynamics of a flapping wing with low order flexibility,” Physics of Fluids 20, 073603 (2008).
  • Clanet (2015) Christophe Clanet, “Sports Ballistics,” Annual Review of Fluid Mechanics 47, 455–478 (2015), publisher: Annual Reviews.
  • Falkovich (2018) Gregory Falkovich, Fluid Mechanics, p. 56, 2nd ed. (Cambridge University Press, Cambridge, 2018).
  • Rubinow and Keller (1961) S. I. Rubinow and Joseph B. Keller, “The transverse force on a spinning sphere moving in a viscous fluid,” Journal of Fluid Mechanics 11, 447–459 (1961).
  • Thielicke and Stamhuis (2014) William Thielicke and Eize J. Stamhuis, “PIVlab – Towards User-friendly, Affordable and Accurate Digital Particle Image Velocimetry in MATLAB,” Journal of Open Research Software 2 (2014), 10.5334/jors.bl.
  • Thielicke and Sonntag (2021) William Thielicke and René Sonntag, “Particle Image Velocimetry for MATLAB: Accuracy and enhanced algorithms in PIVlab,” Journal of Open Research Software 9 (2021), 10.5334/jors.334.
  • Dupeux et al. (2010) Guillaume Dupeux, Anne Le Goff, David Quéré,  and Christophe Clanet, “The spinning ball spiral,” New Journal of Physics 12, 093004 (2010).
  • Espinosa-Garcia et al. (2013) Julian Espinosa-Garcia, Eric Lauga,  and Roberto Zenit, “Fluid elasticity increases the locomotion of flexible swimmers,” Physics of Fluids 25, 031701 (2013).