Fact-checked by Grok 2 weeks ago

Free surface

In physics, a is the surface of a subjected to zero parallel and constant perpendicular normal stress, typically forming the between two homogeneous , such as a and a surrounding gas like air. This boundary allows particles to move tangentially along it without mass flux across, distinguishing it from rigid or confined surfaces in . Free surface flows, often incompressible and gravity-driven, occur in open-channel scenarios where the surface deforms and moves, governed by equations like the Navier-Stokes or that account for the dynamic . These flows exhibit key characteristics such as effects, wave propagation, and sensitivity to external forces like or pressure gradients, making them prevalent in natural and engineered systems. Notable applications include for modeling river flows and dam breaks, for simulating tsunamis and coastal waves, and environmental simulations of flooding or debris transport. In , the arises when partially filled tanks (slack tanks) allow liquid to shift during vessel motion, virtually raising the center of and reducing , a critical factor in ship design and safety assessments. Numerical methods, such as level-set or volume-of-fluid approaches, are essential for accurately capturing the evolving free surface in simulations across these domains.

Fundamentals

Definition

In , a free surface is defined as the between a and an adjacent gas or , where the parallel to the vanishes, permitting the surface to deform dynamically under external influences such as gravity, pressure gradients, and . This boundary condition arises because the surrounding gas exerts negligible viscous on the compared to the liquid's internal stresses, resulting in a stress-free tangential component while the normal stress balances adjusted for effects. Unlike confined surfaces, such as those along solid walls in or channels where no-slip conditions enforce zero , a free surface lacks such rigid constraints, allowing particles to slide freely along the . Similarly, it differs from solid- interfaces, which impose fixed geometric and full stress transmission, whereas free surfaces evolve in shape and position based on the flow dynamics. Gases, however, do not exhibit free surfaces owing to their low with the surrounding medium, which prevents a distinct, deformable from forming. Representative examples include the ocean's surface interfacing with the atmosphere, where wind and gravity drive deformations, and the upper surface of a in an open container, which remains approximately flat under but can ripple or slosh under agitation. The concept developed in hydrodynamics, with foundational contributions from , who explored fluid equilibrium and , and George Gabriel Stokes, who analyzed wave motions and boundary conditions at such interfaces in his seminal works on viscous fluids.

Physical Properties

Surface tension is a fundamental property of free surfaces, representing the cohesive per unit length acting parallel to the , which acts to minimize the surface area of the . This arises from the imbalance of intermolecular attractions at the surface compared to the bulk , quantified by the surface tension \gamma, typically measured in millinewtons per meter (mN/m). For the water-air at 20°C, \gamma is approximately 72.8 mN/m, enabling phenomena such as the support of small objects on the surface despite their exceeding that of . Viscosity and density play key roles in the response of free surfaces to perturbations, particularly in damping oscillatory motions. Viscosity introduces dissipative effects that attenuate surface waves and other deformations, with higher viscosity leading to faster decay of motion amplitude. Density influences the inertial resistance to deformation and contributes to gravitational effects; together, these properties define the capillary length \lambda_c = \sqrt{\frac{\sigma}{\rho g}}, a characteristic scale below which surface tension dominates over gravity, typically around 2.7 mm for water at room temperature. This length scale helps predict the behavior of free surfaces in confined or small-scale geometries. In small-scale systems, free surfaces exhibit meniscus formation, a at the with a solid boundary driven by wettability, which is characterized by the \theta measured through the liquid. The reflects the balance of adhesive forces between the liquid and solid versus the liquid's ; \theta < 90^\circ indicates wetting (e.g., water on ), while \theta > 90^\circ indicates non-wetting (e.g., mercury on ). This is described by Young's : \sigma_{sg} = \sigma_{sl} + \sigma_{lg} \cos \theta, where \sigma_{sg}, \sigma_{sl}, and \sigma_{lg} are the solid-gas, solid-liquid, and liquid-gas interfacial tensions, respectively. The physical properties of free surfaces are sensitive to environmental factors such as and impurities. Surface tension decreases nearly linearly with increasing due to enhanced molecular weakening intermolecular bonds; for , \gamma drops from 75.6 mN/m at 0°C to 72.8 mN/m at 20°C and further to about 59 mN/m at 100°C. Impurities, particularly , typically reduce by adsorbing at the and disrupting , though some solutes like salts can increase it in specific cases. Representative values for common liquids at 20°C are summarized below, illustrating the range across different substances.
LiquidSurface Tension (mN/m)
Water72.8
22.3
Mercury485
63.4
28.2

Equilibrium States

Flat Surfaces

In , the free surface of a aligns perpendicular to the local , forming an surface where the is constant. This configuration ensures that the is across the surface, as any deviation would induce a restoring force due to the in the . For a \mathbf{g} = -g \hat{z}, the potential is V = g z, and the free surface corresponds to V = constant, resulting in a horizontal plane locally. Over small scales, such as in laboratory containers, the free surface appears perfectly flat because the dominates, and the surface height z satisfies \nabla (g z) = 0 for constant g, confirming the planar . However, on Earth's scale, the of the introduces minor deviations from this local flatness; for instance, across a 1-meter span, the central bulge due to is approximately 19.6 nanometers. Globally, the free surface, such as mean , follows the —an irregular surface shaped by Earth's mass distribution and rotation, with undulations up to tens of meters relative to a reference . In confined vessels, the apparent flatness can be altered by surface tension effects, particularly when the container dimensions are comparable to or smaller than the capillary length scale, \lambda_c = \sqrt{\sigma / (\rho g)}, where \sigma is the , \rho the fluid density, and g the . For at standard conditions, \lambda_c \approx 2.7 mm, so in small tubes or vessels below this scale, surface tension induces a curved rather than a flat interface, transitioning to gravity-dominated flatness in larger containers.

Rotational Configurations

In rotating reference frames, the equilibrium free surface of an inviscid deviates from flatness due to the influence of , resulting in a paraboloidal shape that minimizes the . This arises when the undergoes solid-body with \omega, where the surface height z(r) at radial distance r from the is given by z(r) = \frac{\omega^2 r^2}{2g} + h_0, with g denoting and h_0 the height at the center. The paraboloid opens upward, with the lowest point at the and increasing elevation toward the periphery, reflecting the outward centrifugal effect that counteracts . The shape derives from the condition that the free surface aligns with equipotential lines of the effective in the rotating frame. The total potential \Phi combines the gravitational term gz and the centrifugal term -\frac{1}{2}\omega^2 r^2, yielding \Phi = gz - \frac{1}{2}\omega^2 r^2. At , \Phi is constant along the surface, leading directly to the paraboloidal equation upon solving \nabla \Phi = 0. Equivalently, this balance corresponds to an effective \mathbf{g}_{\text{eff}} = -g \hat{z} + \omega^2 r \hat{r}, where the surface remains perpendicular to \mathbf{g}_{\text{eff}} everywhere, varying in both magnitude and with . As \omega \to 0, this reduces to the flat under uniform alone. Stable rotational equilibria persist up to limits imposed by hydrodynamic instabilities, particularly for inviscid fluids where the Rayleigh criterion governs centrifugal stability. This criterion requires that the square of the circulation increase outward (\frac{d}{dr}(r^2 \Omega)^2 > 0, with \Omega = \omega for solid-body rotation), ensuring resistance to axisymmetric disturbances in flows with free surfaces. Beyond this threshold or when the central height h_0 approaches zero—indicating the surface contacts the container bottom—instability ensues, potentially leading to vortex formation or spilling. Historical demonstrations include Isaac Newton's bucket experiment, where a partially filled vessel rotated about its vertical axis produces a curved surface, illustrating relative to distant matter. In laboratory settings, such configurations appear in centrifuges for particle separation, where centrifugal forces enhance efficiency by creating effective fields far exceeding , with open or semi-open rotors exhibiting analogous paraboloidal interfaces to maintain during operation.

Dynamic Behaviors

Surface Waves

Surface waves represent dynamic disturbances on a free surface that propagate away from their generation point, restoring the surface toward through gravitational or forces. These waves arise from perturbations such as , pressure variations, or displacements in bodies like oceans or laboratory tanks. In the for small amplitudes, their behavior is governed by dispersion relations that dictate phase and group velocities as functions of . Gravity waves dominate for longer wavelengths, where buoyancy provides the primary restoring force. The dispersion relation for these waves in water of finite depth h is given by \omega^2 = g k \tanh(k h), where \omega is the angular frequency, g is gravitational acceleration, and k = 2\pi / \lambda is the wavenumber. In the deep-water limit where k h \gg 1, \tanh(k h) \approx 1, simplifying to \omega = \sqrt{g k}, which implies that phase speed c = \omega / k decreases with increasing k, allowing shorter waves to be overtaken by longer ones. This dispersive nature enables wave packets to spread, with group velocity c_g = d\omega / dk = c/2 in deep water carrying energy from the source. For shorter wavelengths, capillary waves, or ripples, emerge where surface tension \sigma restores the surface, as detailed in the physical properties of free surfaces. The dispersion relation in this regime, neglecting , is \omega^2 = (\sigma / \rho) k^3, with \rho the fluid density, yielding phase speeds that increase with k. The full gravity-capillary relation \omega^2 = g k + (\sigma / \rho) k^3 exhibits a minimum speed of approximately $0.23 m/s at a critical \lambda_c \approx 1.7 cm, marking the transition between gravity- and capillary-dominated . Surface waves classify as progressive, which travel across the surface with a propagating crest, or standing, formed by interference of oppositely traveling waves resulting in fixed nodes and antinodes. Ocean swells exemplify long-period progressive gravity waves that persist far from their wind-generated origin, while wind-driven chop consists of shorter, irregular progressive waves blending gravity and capillary modes within the fetch area. Wave propagation is attenuated by damping mechanisms, including viscous dissipation concentrated in boundary layers near the free surface and walls. The Stokes boundary layer, with thickness \delta \sim \sqrt{\nu / \omega} where \nu is kinematic viscosity, accounts for significant energy loss through shear in oscillatory flows, contributing a damping rate proportional to \sqrt{\nu \omega / 2} for low-viscosity cases. Additionally, nonlinear effects cause wave steepening, where higher-order terms sharpen the wave front, eventually leading to breaking when the steepness k a \approx 0.1-0.3 (with a the amplitude) triggers instabilities like Benjamin-Feir modulation, dissipating energy via turbulence and air entrainment.

Instabilities

Free surfaces can exhibit instabilities under specific perturbations or driving forces, leading to the formation of complex patterns, droplet breakup, or turbulent flows that deviate from equilibrium or simple . These instabilities arise due to the interplay of , , , and external forcings, often amplifying small disturbances into macroscopic structures. Understanding these mechanisms is crucial for predicting in fluids. The Rayleigh-Plateau instability describes the breakup of a cylindrical liquid jet into droplets due to surface tension. In this process, axisymmetric perturbations with wavelengths longer than the circumference of the jet (λ > 2πa, where a is the jet radius) grow exponentially, as the reduced surface area of separated droplets lowers the system's energy. Lord Rayleigh derived the dispersion relation for the growth rate of these perturbations in inviscid fluids, given by \omega = \sqrt{\frac{\sigma}{\rho a^3}} f(ka), where \omega is the growth rate, \sigma is the surface tension, \rho is the liquid density, k = 2\pi / \lambda is the wavenumber, and f(ka) is a function that peaks near ka ≈ 0.7 for the fastest-growing mode. This instability underpins applications like inkjet printing and fiber spinning, where controlling perturbation wavelengths stabilizes jets. Faraday instability occurs when a layer with a free surface is subjected to vertical vibrations, parametrically exciting standing surface waves. Above a critical threshold, the subharmonic response—where the wave frequency \omega_wave is half the vibration frequency \omega_vib (\omega_vib = 2 \omega_wave)—leads to the formation of ordered patterns such as rolls or squares on the surface. first observed these waves in 1831 through experiments with mercury and vibrated at audio frequencies, noting their dependence on container geometry and fluid depth. Theoretical analyses confirm that the instability threshold scales with the vibration amplitude and is modulated by and . Marangoni instability in thin liquid films arises from thermocapillary effects, where gradients in surface tension due to temperature variations drive convective flows. In a heated film, warmer regions at the center have lower surface tension, pulling fluid from cooler, higher-tension edges and destabilizing the uniform film into cellular patterns. J.R.A. Pearson's 1958 linear stability analysis for a non-deformable surface showed that the critical Marangoni number Ma_c = \frac{\gamma \Delta T h}{\mu \kappa} \approx 80 (for Prandtl number ≈ 7), where \gamma is the surface tension-temperature coefficient, \Delta T is the temperature difference, h is film thickness, \mu is viscosity, and \kappa is thermal diffusivity, marks the onset of stationary convection cells. This mechanism is prominent in low-gravity environments or thin films where buoyancy is negligible. Turbulence at free surfaces often initiates via the Kelvin-Helmholtz instability, triggered by across the air-water . Velocity differences exceeding a critical value (typically U > \sqrt{g h} for shallow water depth h) generate roll waves that extract from the shear, leading to wave growth and eventual breakdown into turbulent cascades. first described the instability in 1868 for discontinuous velocity profiles, while extended it in 1871 to continuous shear layers, showing unstable modes for wavenumbers below a cutoff determined by contrast and . In oceanic contexts, this instability facilitates an from large-scale wind input to small-scale dissipation, mixing momentum and scalars across the surface .

Applications and Extensions

Engineering Contexts

In naval architecture, the free surface effect significantly impacts ship stability, particularly when partially filled tanks allow liquid to shift during heel, effectively raising the vessel's center of gravity and reducing the metacentric height. This phenomenon arises because the liquid's free surface permits transverse movement, creating a virtual rise in the center of gravity that diminishes the righting moment. The free surface correction to the metacentric height (FSC) is given by \mathrm{FSC} = \frac{i}{\nabla}, where i is the transverse second moment of inertia of the free surface about its longitudinal centerline, and \nabla is the ship's displacement volume. This correction ensures accurate stability calculations during design and loading assessments. According to standards outlined in IMO resolutions, such effects must be accounted for to prevent capsizing risks, with practical mitigation involving tank baffles or filling levels optimized to minimize surface area. In , free surface flows over structures like and weirs are modeled to predict and energy dissipation, adapting Bernoulli's equation to account for varying flow depth and at the surface. The equation, \frac{p}{\rho g} + \frac{v^2}{2g} + z = \constant, is applied along streamlines with the pressure term set to zero at the free surface, enabling estimation of critical depths and flow rates for safe overflow design in and channels. This approach, validated through physical models, supports the sizing of crests to handle events without supercritical flow transitions leading to or . Recent studies emphasize integrating these models with site-specific topography for improved hydraulic efficiency. Sloshing in partially filled tanks poses challenges to , especially in where fuel motion can couple with structural vibrations, altering attitude control and during maneuvers. In such systems, theory is employed, assuming irrotational and , with boundary conditions enforcing no penetration at tank walls and kinematic-dynamic constraints at the free surface to capture wave propagation and pressure impacts. For instance, NASA's analyses of propellant slosh in cylindrical s demonstrate how these models predict damping and resonance frequencies, informing baffle designs to suppress oscillations that could destabilize launch vehicles. This integration is crucial for missions requiring precise orbital insertions, where unmitigated sloshing might exceed limits. Recent advancements in (CFD) have enhanced free surface simulations for applications, particularly through implementations of level-set methods in open-source tools like for multiphase flows. These developments improve interface tracking by evolving a to represent the free surface, reducing numerical diffusion and enabling accurate prediction of complex topologies like breaking waves or droplet formation. Such techniques have been applied to optimize sloshing dampers in marine vessels and fuel management in . These progress support design iterations, minimizing physical prototyping needs.

Astrophysical and Laboratory Uses

In astrophysics, free surfaces play a critical role in liquid-mirror telescopes, where a rotating pool of mercury forms a parabolic shape under centrifugal force, enabling large-aperture optical systems at lower cost than traditional glass mirrors. The International Liquid Mirror Telescope (ILMT), a 4-meter-diameter instrument located in the Devasthal Observatory in India, exemplifies this application; its primary mirror consists of a thin film of liquid mercury spun at precise speeds to achieve the required paraboloidal curvature for focusing light, allowing continuous scanning of a 22-arcminute-wide strip near the zenith for deep photometric and astrometric surveys of transients. This configuration leverages the rotational equilibrium of the free surface, similar to principles discussed in rotational configurations, to produce diffraction-limited performance while avoiding the expense of polishing solid mirrors. Planetary science employs models of free surfaces to understand liquid bodies on moons under low-gravity conditions, such as the methane-ethane lakes on Saturn's moon Titan, where surface tension and weak gravity (about 1/7 of Earth's) dictate equilibrium shapes and dynamics. These lakes, observed by the Cassini mission, exhibit calm, specular reflections indicative of stable free surfaces, with modeling efforts using coupled atmospheric-lake simulations to predict evaporation rates, thermal stratification, and multiphase equilibria involving dissolved nitrogen. For instance, the TITANPOOL numerical model simulates the thermodynamic evolution of these non-ideal mixtures, revealing how low gravity minimizes wave disruptions and promotes flat or gently curved equilibria compared to Earth analogs. Such studies inform broader understandings of volatile retention and climate cycles on Titan, highlighting free surface behaviors in reduced-gravity environments. Laboratory investigations replicate these low-gravity free surface phenomena using drop towers to simulate microgravity, providing short-duration zero-g conditions (typically 5-10 seconds) for observing liquid dominated by . In these experiments, unbound liquids form near-spherical shapes due to minimized gravitational distortion, allowing precise measurement of effects and interface stability without container influences. tests, such as those at facilities like ZARM or NASA's , have demonstrated how sudden free-fall transitions enhance surface tension's role, leading to rapid coalescence or pinning behaviors in droplets, which validates models for fuel management and planetary lake analogs. These controlled setups bridge theoretical predictions with empirical data, emphasizing spherical equilibria as a hallmark of surface tension in weightless regimes. Extensions to granular materials in fluidized beds treat the upper interface as an effective free surface, mimicking liquid-like behaviors in where gas flow suspends particles, enabling uniform mixing and . In vibrated or gas-fluidized granular systems, the bed's top surface exhibits fluid dynamic properties, such as free-flow under pseudo-gravity and wave propagation, analogous to liquid free surfaces but with discrete particle interactions. This analogy is applied in processes like pharmaceutical and catalytic reactions, where computational fluid dynamics-discrete element method (CFD-DEM) simulations optimize bed height and bubbling to maintain a stable "free surface" for efficient or . Such systems scale to industrial volumes, providing cost-effective alternatives to true handling in powder-based .

References

  1. [1]
    LSDYNA Free Surface & Bi-Phasic flows - Ansys
    In physics, a free surface is the surface of a fluid that is subject to constant perpendicular normal stress and zero parallel shear stress, such as the ...Missing: definition | Show results with:definition
  2. [2]
    Free Surface Flow - an overview | ScienceDirect Topics
    Free surface flow is defined as a type of incompressible flow that occurs at a moving, deforming boundary, specifically the air-water interface, where fluid ...
  3. [3]
  4. [4]
  5. [5]
    Free surface effect - Wärtsilä
    A partially filled tank is know as a “slack tank”. The reduction of stability caused by the liquids in slack tanks is known as free-surface effect.Missing: architecture | Show results with:architecture
  6. [6]
  7. [7]
    An Introduction to Fluid Dynamics
    Interaction of vorticity with a free surface in turbulent open channel flow. ... Entropy increase in confined free expansions via molecular dynamics and smooth- ...
  8. [8]
    [PDF] an introduction to - fluid dynamics
    Page 1. AN INTRODUCTION TO. FLUID DYNAMICS. BY. G. K. BATCHELOR, F.R.S.. Prof.,1Or of Applid Mathmatit' in 1M Un;v.,tity of Cambridg• .. :.:. CAMBRIDGE.
  9. [9]
    Free surface - chemeurope.com
    In fluid mechanics a free surface flow, also called open channel flow, is the gravity driven flow of a fluid under a free surface, typically water flowing ...Waves · Rotation · Related terms
  10. [10]
    From Navier to Stokes: Commemorating the Bicentenary of ... - MDPI
    Jan 6, 2024 · The article presents a summarised history of the equations governing fluid motion, known as the Navier–Stokes equations.
  11. [11]
    [PDF] Hydrostatic shapes
    In this chapter the influence of gravity on the shape of large bodies of fluid is analyzed, the primary goal being the calculation of the size and shape of the.
  12. [12]
    [PDF] Chapter 2 - The Earth's Gravitational field
    Or: the field is perpendicular to the equipotential surface. In global gravity one aims to determine and explain deviations from the equipotential surfaces, or.
  13. [13]
    Fluid Statics & the Hydrostatic Equation – Introduction to Aerospace ...
    In fluid mechanics, a primary concern is describing and understanding the motion, or dynamics, of fluids, i.e., the field of fluid dynamics. First, remember ...
  14. [14]
    Straight, Level, and the Curvature of the Earth | Math Encounters Blog
    Nov 16, 2010 · Conclusion. The contractor had stated that the curvature of the Earth causes level to deviate from horizontal by an 1/8th of an inch (125 ...
  15. [15]
    Geoids - an overview | ScienceDirect Topics
    The geoid is defined as an equipotential surface of the Earth's gravity field that corresponds to the undisturbed global sea level, serving as a height ...
  16. [16]
    [PDF] 2. Definition and Scaling of Surface Tension - MIT OpenCourseWare
    Roughly speaking, the capillary length prescribes the maximum size of pendant drops that may hang inverted from a ceiling, water-walking insects, and raindrops.
  17. [17]
  18. [18]
    [PDF] 2. Theory A. Derivation of parabolic formula of rotating liquid In the ...
    The liquid surface should be perpendicular to the combined force, so it takes the concave shape of a paraboloid. For the quantitative calculation, the rotating ...
  19. [19]
    The stability of rotating flows with a cylindrical free surface
    Mar 28, 2006 · A necessary and sufficient condition for stability to axisymmetric disturbances is derived, which requires that Rayleigh's criterion of ...
  20. [20]
    Newton's Bucket: Fluid in a Spinning Tank - UCLA EPSS
    This new SpinLab video shows the Newton's Bucket experiment in which the free surface of water in a rotating tank takes on a paraboloidal shape.
  21. [21]
    Principles of Continuous Flow Centrifugation - Beckman Coulter
    Air is displaced through the center inlet. The cushion or gradient is held against the rotor wall by centrifugal force.
  22. [22]
    [PDF] Gravity waves on water - UMD Physics
    Waves on the surface of water can arise from the restoring force of gravity or of surface tension, or a combination. For wavelengths longer than a couple.
  23. [23]
    [PDF] Chapter 2 - Deep water gravity waves
    Deep water gravity waves. 2.1 Surface motions on water of finite depth. We now move on to consider motions in water that is not shallow, i.e., we will allow ...
  24. [24]
    Capillary Waves
    Capillary Waves. Water in contact with air actually possesses a finite surface tension, $T\simeq 7\times 10^{-2}\,{ (Haynes and Lide 2011b), which allows ...
  25. [25]
    [PDF] Measurements of surface-wave damping in a container
    Stokes layers, and a2C includes the effects of both viscous dissipation in the bulk and a higher order correction to the damping in the Stokes layers. ~At ...
  26. [26]
    [PDF] The Role of Surface-Wave Breaking in Air-Sea Interaction
    Oct 1, 2007 · Breaking serves to limit the height of surface waves, mix the surface waters, generate ocean currents, and enhance air-sea fluxes of heat, mass, ...
  27. [27]
    [PDF] Analysis of Propellant Slosh Dynamics
    The linearized dynamic equations of propellant motion in the regime dominated by gravity force have been developed for a cylindrical tank.
  28. [28]
  29. [29]
    The 4 m International Liquid Mirror Telescope
    The primary components of the ILMT are: (1) a parabolic mirror covered with a thin film of liquid mercury, (2) an air bearing that supports the mirror, and its ...
  30. [30]
    Air–Sea Interactions on Titan: Effect of Radiative Transfer on the ...
    In this work we use a 2D atmospheric mesoscale model coupled to a slab lake model to investigate the effect of solar and infrared radiation on the exchange of ...
  31. [31]
    Stratification Dynamics of Titan's Lakes via Methane Evaporation - NIH
    We built the TITANPOOL numerical model to investigate the thermodynamic, physical, and chemical evolution of non-ideal methane–ethane–nitrogen lakes under ...
  32. [32]
    Air-sea interactions on Titan: Lake evaporation, atmospheric ...
    Nov 15, 2020 · Titan's abundant lakes and seas exchange methane vapor and energy with the atmosphere via a process generally known as air-sea interaction.
  33. [33]
    [PDF] Fluid Interface Phenomena in a Low-Gravity Environment: 3 / 7c
    Drop towers used as experimental facilities have played a major role in the development of fundamental theory, engineering analysis, and the proofing of system ...
  34. [34]
    Effect of gravity on the spreading of a droplet deposited by liquid ...
    Jun 21, 2023 · Under microgravity condition the surface tension dominates over the gravity and as a result the droplet tends to form a spherical shape and will ...
  35. [35]
    Dynamically structured bubbling in vibrated gas-fluidized granular ...
    Aug 26, 2021 · The motion of granular materials, e.g., sand or catalytic particles, underlies many natural and industrial processes.
  36. [36]
    CFD-DEM Fluidized Bed Drying Study Using a Coarse-Graining ...
    Nov 22, 2023 · Fluidized beds are commonly applied to industrial drying applications. Modeling using the computational fluid dynamics-discrete element ...