ArticlePublisher preview available

Surface topography measurements of the bouncing droplet experiment

Authors:
To read the full-text of this research, you can request a copy directly from the authors.

Abstract and Figures

A free-surface synthetic Schlieren (Moisy et al. in Exp Fluids 46:1021–1036, 2009; Eddi et al. in J Fluid Mech 674:433–463, 2011) technique has been implemented in order to measure the surface topography generated by a droplet bouncing on a vibrating fluid bath. This method was used to capture the wave fields of bouncers, walkers, and walkers interacting with boundaries. These wave profiles are compared with existing theoretical models and simulations and will prove valuable in guiding their future development. Specifically, the method provides insight into what type of boundary conditions apply to the wave field when a bouncing droplet approaches a submerged obstacle.
Regime diagram [see Molácek and Bush (2013b) for its derivation] displaying the drop’s bouncing or walking mode as a function of the vibration number ω0/σ/ρR3\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\omega _0/\sqrt{\sigma /\rho R^3}$$\end{document} and driving acceleration γ/g\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\gamma /g$$\end{document}. A drop in the (m,n)i\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$(m, n)^i$$\end{document} mode bounces n times in m forcing periods, with the integer i ordering multiple (m, n) states according to their total mechanical energy, with i = 1 being the lowest. Here, we have ω0=80Hz\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\omega _0=80\,{\mathrm{Hz}}$$\end{document}, ν=20cSt\documentclass[12pt]{minimal} \usepackage{amsmath} \usepackage{wasysym} \usepackage{amsfonts} \usepackage{amssymb} \usepackage{amsbsy} \usepackage{mathrsfs} \usepackage{upgreek} \setlength{\oddsidemargin}{-69pt} \begin{document}$$\nu =20\,{\mathrm{cSt}}$$\end{document} and fixed drop radius R = 0.38 mm. The red symbols indicate the experiments performed in our study
… 
This content is subject to copyright. Terms and conditions apply.
1 3
Exp Fluids (2016) 57:163
DOI 10.1007/s00348-016-2251-4
RESEARCH ARTICLE
Surface topography measurements of the bouncing droplet
experiment
Adam P. Damiano1 · P.‑T. Brun1 · Daniel M. Harris1 · Carlos A. Galeano‑Rios2,3 ·
John W. M. Bush1
Received: 15 June 2016 / Revised: 6 September 2016 / Accepted: 10 September 2016 / Published online: 11 October 2016
© Springer-Verlag Berlin Heidelberg 2016
unstable to Faraday waves. Couder et al. (2005) and Protière
et al. (2006) discovered that in certain experimental regimes,
the droplets may self-propel along the surface of the bath due
to interactions with their own wave fields. These walking
droplets, henceforth walkers, are spatially extended objects
that exhibit several phenomena reminiscent of quantum sys-
tems (Couder and Fort 2006; Eddi et al. 2009, 2011; Bush
2010, 2015a, b; Fort et al. 2010; Harris et al. 2013; Perrard
et al. 2014a, b; Oza et al. 2014; Harris and Bush 2014).
While the droplet, of typical radius 0.4 mm, is read-
ily discerned by eye, the waves excited by the droplet,
of typical amplitude of 1–20 μm, are relatively diffi-
cult to observe and quantify. Various theoretical models
have been developed to describe the waves created by a
bouncing droplet (Eddi et al. 2011; Molácek and Bush
2013a, b; Oza et al. 2013; Labousse 2014; Milewski
et al. 2015; Gilet 2016; Blanchette 2016), on the basis
of which much headway has been made in rationalizing
the behavior of the walkers in a variety of settings [see
Bush (2015a, b) for reviews]. Nevertheless, theoretical
developments would benefit from quantitative measure-
ments of the wave field. In particular, walker-boundary
interactions as arise in a number of key quantum ana-
logues (Couder and Fort 2006; Eddi et al. 2009; Harris
et al. 2013; Harris 2015) remain poorly characterized
and understood. Specifically, some theoretical models
of walkers near boundaries apply a zero-wave-amplitude
boundary condition (Gilet 2016; Blanchette 2016) while
others apply a zero slope boundary condition (Duber-
trand et al. 2016).
We here report the results of an experimental effort to
measure the surface topography in the walking drop system
using the surface synthetic Schlieren technique originally
developed by Moisy et al. (2009), as was applied by Eddi
et al. (2009, 2011). Specifically, we utilize the refracted
Abstract A free-surface synthetic Schlieren (Moisy et al.
in Exp Fluids 46:1021–1036, 2009; Eddi et al. in J Fluid
Mech 674:433–463, 2011) technique has been implemented
in order to measure the surface topography generated by a
droplet bouncing on a vibrating fluid bath. This method
was used to capture the wave fields of bouncers, walkers,
and walkers interacting with boundaries. These wave pro-
files are compared with existing theoretical models and
simulations and will prove valuable in guiding their future
development. Specifically, the method provides insight into
what type of boundary conditions apply to the wave field
when a bouncing droplet approaches a submerged obstacle.
1 Introduction
A millimetric drop placed onto a vibrating liquid bath can
bounce indefinitely on the fluid surface due to a thin film of
air that prevents coalescence and is replenished with each
bounce (Walker 1978; Couder et al. 2005; Terwagne et al.
2007; Vandewalle et al. 2006). The drop dynamics depends
critically on the forcing acceleration of the bath γ relative
to the critical threshold γF, at which the interface becomes
Electronic supplementary material The online version of this
article (doi:10.1007/s00348-016-2251-4) contains supplementary
material, which is available to authorized users.
* P.-T. Brun
pierrethomas.brun@gmail.com
1 MIT, Cambridge, MA, USA
2 IMPA/National Institute of Pure and Applied Mathematics,
Est. D. Castorina, 110, Rio de Janeiro RJ 22460-320, Brazil
3 Department of Mathematical Sciences, University of Bath,
Bath BA2 7AY, UK
Content courtesy of Springer Nature, terms of use apply. Rights reserved.
... For instance, Zhang and Cox (1994) extracted the surface topography based on a system of color-coded collimated beams that emerge below the surface. Other approaches, including the present study, relate the surface topography to the optical deformation of a known pattern when it is imaged through a deformed surface (Fouras et al. 2008;Gomit et al. 2013;Moisy et al. 2009;Ng et al. 2011;Murase 1992;Damiano et al. 2016;Morris and Kutulakos 2011;Li et al. 2021). In this approach, the imaged patterns are either random-speckle or regular (e.g., grid) patterns, and they can either be fixed in time (e.g., patterns printed on boundaries of test sections) or change with time (e.g., patterns employed to simultaneously perform velocimetry). ...
... Moreover, the varying path of the light rays in the liquid phase contributes to the reduction in the ray crossing risk (Fig. 3) allowing the resolution of steeper waves than the parallel PP approach could provide (Moisy et al. 2009;Damiano et al. 2016). Figure 3 highlights the risk reduction; the extension (black dashed lines) of the refracted ...
Article
Full-text available
The reconstruction of liquid–air interfaces in three-dimensional (3D) space is of great value for many practical and scientific applications, from manufacturing automation to naval hydrodynamics and much more. This study presents a new, nonintrusive experimental diagnostic for surface reconstruction in the 3D space based on ray tracing and structured illumination. The 3D topography of an air–water interface is obtained based on a robust ray tracing algorithm by imaging, from the gas phase, a vertical 2D laser-induced fluorescent grid generated by Talbot-effect structured illumination. The images of the optically deformed grid are related to the instantaneous surface profiles. The vertical orientation of the structured illumination plane provides boundary conditions for the surface reconstruction and minimizes restrictions of existing schemes. The experimental technique and the profilometry algorithm are presented in detail, and samples of reconstructed surfaces from a droplet impact are discussed. The surface reconstruction performance is assessed by analyzing samples of known topography and estimating residuals for the local surface elevation and slope.
... In Equation (1), drop motion is driven by a wave force, proportional to the local gradient of the underlying wavefield h, and resisted by a linear drag force with coefficient D. The wavefield, defined in Equation (2), is modelled as the superposition of waves of spatial form f (r) generated by each droplet along its past trajectory, as summed over all of the droplets in the lattice. The wave kernel f (r) is based on the wave model of Moláček and Bush [47] but more accurately captures the experimentally observed far-field decay of the wavefield [50], and so the long-range interactions between droplets [32,48,50,51]. For r < 1, f (r) is well approximated by J 0 (r), but then decays more rapidly than J 0 (r) for r > 1, as prescribed by the spatial damping factor ζ which is defined in terms of fluid parameters in Table 1. ...
... In Equation (1), drop motion is driven by a wave force, proportional to the local gradient of the underlying wavefield h, and resisted by a linear drag force with coefficient D. The wavefield, defined in Equation (2), is modelled as the superposition of waves of spatial form f (r) generated by each droplet along its past trajectory, as summed over all of the droplets in the lattice. The wave kernel f (r) is based on the wave model of Moláček and Bush [47] but more accurately captures the experimentally observed far-field decay of the wavefield [50], and so the long-range interactions between droplets [32,48,50,51]. For r < 1, f (r) is well approximated by J 0 (r), but then decays more rapidly than J 0 (r) for r > 1, as prescribed by the spatial damping factor ζ which is defined in terms of fluid parameters in Table 1. ...
Article
Full-text available
We present the results of a theoretical investigation of the stability and collective vibrations of a two-dimensional hydrodynamic lattice comprised of millimetric droplets bouncing on the surface of a vibrating liquid bath. We derive the linearized equations of motion describing the dynamics of a generic Bravais lattice, as encompasses all possible tilings of parallelograms in an infinite plane-filling array. Focusing on square and triangular lattice geometries, we demonstrate that for relatively low driving accelerations of the bath, only a subset of inter-drop spacings exist for which stable lattices may be achieved. The range of stable spacings is prescribed by the structure of the underlying wavefield. As the driving acceleration is increased progressively, the initially stationary lattices destabilize into coherent oscillatory motion. Our analysis yields both the instability threshold and the wavevector and polarization of the most unstable vibrational mode. The non-Markovian nature of the droplet dynamics renders the stability analysis of the hydrodynamic lattice more rich and subtle than that of its solid state counterpart.
... Despite its simplicity, Moisy et al. (2009) reported a precision of 1 μm for a view field of 10 cm . This method has been successfully used in the case of Faraday waves (Eddi et al. 2011;Lau et al. 2020), wave-droplet interactions (Eddi et al. 2009;Damiano et al. 2016) and floating objects (Poty et al. 2014;Metzmacher et al. 2017;Vandewalle et al. 2020). ...
Article
Full-text available
The Schlieren method intends to reveal the elevation of a refractive fluid–fluid interface. The method is based on a comparison of images of a single pattern placed at the bottom of the container. Accurate measurements can be obtained with a simple and low-cost optical setup. However, it is restricted to weak interface deformations, weak slopes and weak paraxial angles. To overcome these limitations, we propose an enhanced optical setup that uses a bitelecentric objective and a double pattern. Thanks to this new setup, we avoid geometrical approximations and we extend the method to moderate/large deformations. Moreover, the proposed method does not depend on the liquid depth and could be used in various applications. Graphical abstract
... The displacement field u(x, y) was determined using the MATLAB-based FCD code provided at [24] and subsequently ∇h(x, y) was inverted using the intgrad2 function in MATLAB [25]. Finally, the surface height profile h(x, y) was processed using a spatial bandpass filter, primarily to remove any nonphysical large wavelength distortions predicted by the method as described in [26]. ...
Article
Nature has evolved a vast array of strategies for propulsion at the air-fluid interface. Inspired by a survival mechanism initiated by the honeybee (Apis mellifera) trapped on the surface of water, we here present the SurferBot: a centimeter-scale vibrating robotic device that self-propels on a fluid surface using analogous hydrodynamic mechanisms as the stricken honeybee. This low-cost and easily assembled device is capable of rectilinear motion thanks to forces arising from a wave-generated, unbalanced momentum flux, achieving speeds on the order of centimeters per second. Owing to the dimensions of the SurferBot and amplitude of the capillary wave field, we find that the magnitude of the propulsive force is similar to that of the honeybee. In addition to a detailed description of the fluid mechanics underpinning the SurferBot propulsion, other modes of SurferBot locomotion are discussed. More broadly, we propose that the SurferBot can be used to explore fundamental aspects of active and driven particles at fluid interfaces, as well as in robotics and fluid mechanics pedagogy.
... The displacement field u(x, y) was determined using the MATLAB-based FCD code provided at [21] and subsequently ∇h(x, y) was inverted using the intgrad2 function in MATLAB [22]. Finally, the surface height profile h(x, y) was processed using a spatial band-pass filter, primarily to remove any nonphysical large wavelength distortions predicted by the method as described in [23]. ...
Preprint
Nature has evolved a vast array of strategies for propulsion at the air-fluid interface. Inspired by a survival mechanism initiated by the honeybee (Apis mellifera) trapped on the surface of water, we here present the SurferBot: a centimeter-scale vibrating robotic device that self-propels on a fluid surface using analogous hydrodynamic mechanisms as the stricken honeybee. This low-cost and easily assembled device is capable of rectilinear motion thanks to forces arising from a wave-generated, unbalanced momentum flux, achieving speeds on the order of centimeters per second. Owing to the dimensions of the SurferBot and amplitude of the capillary wave field, we find that the magnitude of the propulsive force is similar to that of the honeybee. In addition to a detailed description of the fluid mechanics underpinning the SurferBot propulsion, other modes of SurferBot locomotion are discussed. More broadly, we propose that the SurferBot can be used to explore fundamental aspects of active and driven particles at fluid interfaces, as well as in robotics and fluid mechanics pedagogy.
... This functional form serves a good approximation to the droplet-generated waves in the vicinity of the droplet. 35,58,59 Choosing a Bessel function wave form W(x) = J 0 (x) results in the wave form gradient f(x) = J 1 (x). Substituting in Eq. (5) gives ...
Article
Full-text available
Vertically vibrating a liquid bath can give rise to a self-propelled wave–particle entity on its free surface. The horizontal walking dynamics of this wave–particle entity can be described adequately by an integro-differential trajectory equation. By transforming this integro-differential equation of motion for a one-dimensional wave–particle entity into a system of ordinary differential equations (ODEs), we show the emergence of Lorenz-like dynamical systems for various spatial wave forms of the entity. Specifically, we present and give examples of Lorenz-like dynamical systems that emerge when the wave form gradient is (i) a solution of a linear homogeneous constant coefficient ODE, (ii) a polynomial, and (iii) a periodic function. Understanding the dynamics of the wave–particle entity in terms of Lorenz-like systems may prove to be useful in rationalizing emergent statistical behavior from underlying chaotic dynamics in hydrodynamic quantum analogs of walking droplets. Moreover, the results presented here provide an alternative physical interpretation of various Lorenz-like dynamical systems in terms of the walking dynamics of a wave–particle entity.
... where K 1 is the modified Bessel function of the second kind. Our simulations indicate that the fitting parameters are A ≈ 15.4 μm and b ≈ 0.002 mm −2 , which is roughly consistent with previous free-surface synthetic Schlieren experiments 76 . We note that the spatial damping included in equation (27) does not play a crucial role in rationalizing the mechanism responsible for the emergent collective order; thus, we can safely neglect its effects. ...
Article
Full-text available
Macroscale analogues1,2,3 of microscopic spin systems offer direct insights into fundamental physical principles, thereby advancing our understanding of synchronization phenomena⁴ and informing the design of novel classes of chiral metamaterials5,6,7. Here we introduce hydrodynamic spin lattices (HSLs) of ‘walking’ droplets as a class of active spin systems with particle–wave coupling. HSLs reveal various non-equilibrium symmetry-breaking phenomena, including transitions from antiferromagnetic to ferromagnetic order that can be controlled by varying the lattice geometry and system rotation⁸. Theoretical predictions based on a generalized Kuramoto model⁴ derived from first principles rationalize our experimental observations, establishing HSLs as a versatile platform for exploring active phase oscillator dynamics. The tunability of HSLs suggests exciting directions for future research, from active spin–wave dynamics to hydrodynamic analogue computation and droplet-based topological insulators.
... The surface topography generated by a droplet bouncing on a vibrating fluid bath was experimentally measured by Damiano [162] and Damiano et al. [163] using free-surface synthetic Schlieren (FS-SS). Synthetic schlieren [164] is a process that is used to visualize the flow of a fluid of variable refractive index (n = c/v where c is the speed of light in vacuum and v is the light phase velocity in the medium.) ...
Article
Full-text available
This article considers additional phenomena that complement the earlier topics addressed by Ibrahim [(Liquid Sloshing Dynamics: Theory and Applications. Cambridge University Press, Cambridge, 2005), (ASME J Fluids Eng 137(9):090801, 2015)]. The first phenomenon is the localized Faraday waves known as oscillons, which were observed in granular materials and liquid layers subjected to parametric excitation. Extreme waves, known as rogue, generated in the Faraday surface ripples, are related to the increase in the horizontal mobility of oscillating solitons (oscillons), and their horizontal motion is random over a limited range of excitation acceleration amplitude. Parametric excitation of water in a Hele–Shaw cell and the associated localized standing surface waves of large amplitude will be discussed. The surface wave pattern exhibited a certain similarity with the three-dimensional axisymmetric oscillon. Faraday waves in superfluid Fermi–Bose mixtures and their wave function will be addressed in terms of position and time as described by the Schrödinger equation with time-dependent parabolic potential. The phenomenon of walking fluid droplets on Faraday waves constitutes the majority portion of this article. Different regimes of droplet motion in terms of droplet physical properties, the fluid bath excitation acceleration amplitude and frequency will be discussed. The droplet trajectory diffraction, when passes through a slit, shares the same random features of electron diffraction. The duality of the droplet-wave field together with the path-memory-driven nonlocality and other related topics will be assessed. This article is complemented with the fascinating phenomenon of the stone and bombs skipping/ricochet over water surface.
... As presented in Fig. 2(a), the pilot wave accompanying steady rectilinear propulsion has a horseshoe-like form ahead of the particle, with an interference pattern apparent in the particle's wake, reminiscent of that arising in the walking-droplet system. 1,2,17,18,20,21,40 As is evident in Fig. 2(b), the wave field accompanying orbital motion is markedly different in form to that of rectilinear propulsion and serves to confine the particle to its circular trajectory. 26 The evolution of the pilot wave accompanying wobbling, precessing, and erratic states is presented in the supplementary material. ...
Article
Full-text available
We present the results of a theoretical investigation into the dynamics of a vibrating particle propelled by its self-induced wave field. Inspired by the hydrodynamic pilot-wave system discovered by Yves Couder and Emmanuel Fort, the idealized pilot-wave system considered here consists of a particle guided by the slope of its quasi-monochromatic “pilot” wave, which encodes the history of the particle motion. We characterize this idealized pilot-wave system in terms of two dimensionless groups that prescribe the relative importance of particle inertia, drag and wave forcing. Prior work has delineated regimes in which self-propulsion of the free particle leads to steady or oscillatory rectilinear motion; it has further revealed parameter regimes in which the particle executes a stable circular orbit, confined by its pilot wave. We here report a number of new dynamical states in which the free particle executes self-induced wobbling and precessing orbital motion. We also explore the statistics of the chaotic regime arising when the time scale of the wave decay is long relative to that of particle motion and characterize the diffusive and rotational nature of the resultant particle dynamics. We thus present a detailed characterization of free-particle motion in this rich two-parameter family of dynamical systems.
Article
In a pioneering series of experiments, Yves Couder, Emmanuel Fort and coworkers demonstrated that droplets bouncing on the surface of a vertically vibrating fluid bath exhibit phenomena reminiscent of those observed in the microscopic quantum realm. Inspired by this discovery, we here conduct a theoretical and numerical investigation into the structure and dynamics of one-dimensional chains of bouncing droplets. We demonstrate that such chains undergo an oscillatory instability as the system’s wave-induced memory is increased progressively. The predicted oscillation frequency compares well with previously reported experimental data. We then investigate the resonant oscillations excited in the chain when the drop at one end is subjected to periodic forcing in the horizontal direction. At relatively high memory, the drops may oscillate with an amplitude larger than that prescribed, suggesting that the drops effectively extract energy from the collective wave field. We also find that dynamic stabilization of new bouncing states can be achieved by forcing the chain at high frequency. Generally, our work provides insight into the collective behavior of particles interacting through long-range and temporally nonlocal forces.
Article
Full-text available
We aim to describe a droplet bouncing on a vibrating bath. Due to Faraday instability a surface wave is created at each bounce and serves as a pilot wave of the droplet. This leads to so called walking droplets or walkers. Since the seminal experiment by {\it Couder et al} [Phys. Rev. Lett. {\bf 97}, 154101 (2006)] there have been many attempts to accurately reproduce the experimental results. Here we present a simple and highly versatile model inspired from quantum mechanics. We propose to describe the trajectories of a walker using a Green function approach. The Green function is related to Helmholtz equation with Neumann boundary conditions on the obstacle(s) and outgoing conditions at infinity. For a single slit geometry our model is exactly solvable and reproduces some general features observed experimentally. It stands for a promising candidate to account for the presence of any boundaries in the walkers'dynamics.
Article
Full-text available
A deterministic low-dimensional iterated map is proposed here to describe the interaction between a bouncing droplet and Faraday waves confined to a circular cavity. Its solutions are investigated theoretically and numerically. The horizontal trajectory of the droplet can be chaotic: it then corresponds to a random walk of average step size equal to half the Faraday wavelength. An analogy is made between the diffusion coefficient of this random walk and the action per unit mass h/m of a quantum particle. The statistics of droplet position and speed are shaped by the cavity eigenmodes, in remarkable agreement with the solution of Schrödinger equation for a quantum particle in a similar potential well.
Article
Full-text available
Small drops bouncing across a vibrating liquid bath display many features reminiscent of quantum systems.
Article
Full-text available
A millimetric droplet bouncing on the surface of a vibrating fluid bath can self-propel by virtue of a resonant interaction with its own wave field. This system represents the first known example of a pilot-wave system of the form envisaged by Louis de Broglie in his double-solution pilot-wave theory. We here develop a fluid model of pilot-wave hydrodynamics by coupling recent models of the droplet's bouncing dynamics with a more realistic model of weakly viscous quasi-potential wave generation and evolution. The resulting model is the first to capture a number of features reported in experiment, including the rapid transient wave generated during impact, the Doppler effect and walker–walker interactions.
Article
Full-text available
Waves and particles are distinct objects at a macroscopic scale. The existence of walkers, drops bouncing on a vertically vibrated fluid bath is a surprising case of dual objects at our scale. The drop is self-propelled, piloted by the standing surface waves generated by its previous rebounds. These objects exhibit a rich dynamics relying on the concept of path memory. Indeed, the wave field results from the position of the past impacts left all along the walker trajectory. The memory is tunable at will by simply changing the vertical acceleration of the bath. A series of experiments have revealed the surprising dynamical behaviors of this dual drop-wave entity. In this PhD, we give a theoretical understanding of the temporal non local structure of walkers. We explore the dynamics of numerical walkers in a two-dimensional harmonic potential. We observe that the system only reaches a relatively limited set of stable attractors, quantized in both extension and mean angular momentum, in excellent agreement with the experimental results. We investigate how the different time scales are intertwined, which decouples the short-time acting propulsion from the build-up of coherent wave structures at much longer time scales. We analyze the non-local mechanism revealing the internal symmetries of the walker which drives the convergence of the dynamics to a set of low-dimensional eigenstates.
Article
Couder et al. ( Nature , vol. 437 (7056), 2005, p. 208) discovered that droplets walking on a vibrating bath possess certain features previously thought to be exclusive to quantum systems. These millimetric droplets synchronize with their Faraday wavefield, creating a macroscopic pilot-wave system. In this paper we exploit the fact that the waves generated are nearly monochromatic and propose a hydrodynamic model capable of quantitatively capturing the interaction between bouncing drops and a variable topography. We show that our reduced model is able to reproduce some important experiments involving the drop–topography interaction, such as non-specular reflection and single-slit diffraction.
Article
Since their discovery by Yves Couder and Emmanuel Fort, droplets walking on a vibrating liquid bath have attracted considerable attention because they unexpectedly exhibit certain features reminiscent of quantum particles. While the behaviour of walking droplets in unbounded geometries has to a large extent been rationalized theoretically, no such rationale exists for their behaviour in the presence of boundaries, as arises in a number of key quantum analogue systems. We here present the results of a combined experimental and theoretical study of the interaction of walking droplets with a submerged planar barrier. Droplets exhibit non-specular reflection, with a small range of reflection angles that is only weakly dependent on the system parameters, including the angle of incidence. The observed behaviour is captured by simulations based on a theoretical model that treats the boundaries as regions of reduced wave speed, and rationalized in terms of momentum considerations.
Article
We present a first-principles model of drops bouncing on a liquid reservoir. We consider a nearly inviscid liquid reservoir and track the waves that develop in a bounded domain. Bouncing drops are modeled as vertical linear springs. We obtain an expression for the contact force between drop and liquid surface and a model where the only adjustable parameter is an effective viscosity used to describe the waves on the reservoir’s surface. With no adjustable parameters associated to the drop, we recover experimental bouncing times and restitution coefficients. We use our model to describe the effect of the Bond, Ohnesorge, and Weber numbers on drops bouncing on a stationary reservoir. We also use our model to describe drops bouncing on an oscillated reservoir, describing various bouncing modes and a walking threshold.