Fact-checked by Grok 2 weeks ago

Weber number

The Weber number (We) is a in that characterizes the ratio of deforming inertial forces to stabilizing forces in multiphase flows, particularly those involving an between two immiscible fluids. It is mathematically defined as We = \frac{\rho v^2 L}{\sigma}, where \rho is the of the , v is the , L is the scale, and \sigma is the . This parameter plays a crucial role in determining the dominance of over in such systems, helping to predict behaviors like droplet breakup, bubble formation, and interface stability. When the Weber number is low, forces prevail, maintaining coherent structures such as intact droplets or films; conversely, high values indicate that inertial forces overcome , leading to fragmentation or dispersion. The Weber number finds extensive applications in engineering contexts, including the analysis of spray for in systems, emulsion production in chemical processing, and the dynamics of thin liquid films in or processes. It is often used alongside other dimensionless numbers, such as the for viscous effects and the number for gravitational influences, to provide a comprehensive of multiphase phenomena.

Fundamentals

Definition

The Weber number (We) is a in that characterizes the relative importance of inertial forces to forces in flows involving interfaces between fluids. It is defined mathematically as \We = \frac{\rho v^2 L}{\sigma}, where \rho is the of the phase exerting the inertial forces (e.g., the continuous phase in multiphase flows, in kg/m³), v is the characteristic velocity relative to the interface (in m/s), L is the scale (in m, such as the of a droplet or the radius of a liquid ), and \sigma is the coefficient (in N/m). This formulation arises as the ratio of inertial forces, approximated as \rho v^2 L^2, to surface tension forces, approximated as \sigma L. Equivalently, from an energy perspective, the Weber number represents the ratio of kinetic energy, on the order of \rho v^2 L^3, to surface energy, on the order of \sigma L^2. The dimensionless nature of We is confirmed by dimensional analysis: the units of the numerator \rho v^2 L yield kg/(m·s²), matching those of the denominator \sigma (also kg/s²), resulting in a unitless quantity.

Physical Significance

The Weber number characterizes the relative dominance of inertial forces over forces in fluid flows involving interfaces between phases, such as liquid-gas or liquid-liquid systems. When the Weber number exceeds unity (We > 1), inertial effects prevail, promoting deformation, disruption, and breakup of fluid structures like droplets or interfaces. Conversely, when We < 1, forces dominate, favoring stable, minimally deformed configurations such as spherical droplets or bubbles. In low-Weber-number regimes, flows are capillary-dominated, where surface tension maintains coherent shapes, as observed in small bubbles rising in liquids that remain nearly spherical due to negligible inertial deformation. High-Weber-number regimes, on the other hand, are inertia-dominated, leading to phenomena like splashing upon drop impact on surfaces, where kinetic energy overcomes interfacial resistance to eject secondary droplets. The Weber number plays a critical role in multiphase flows by predicting interface stability, the onset of wave formation on free surfaces, and thresholds for structural breakup. For instance, a critical Weber number of approximately 12 marks the transition to bag breakup of droplets in air streams, beyond which inertial forces cause fragmentation.

Historical Development

Origin and Introduction

The Weber number emerged in the early 20th century amid advancements in fluid mechanics, particularly within the fields of naval architecture and wave theory spanning the 1910s to 1930s, as researchers sought to apply principles of dynamic similitude for scaling model experiments to full-scale phenomena. These efforts were driven by the need to predict behaviors in ship hydrodynamics and surface wave propagation, where forces like inertia, viscosity, gravity, and surface tension interact across different scales. Building on foundational ideas from dimensional analysis, the number addressed the challenge of maintaining similarity in model testing for interfacial flows. Moritz Weber, a professor of naval mechanics at the Technische Hochschule Charlottenburg in Berlin, formalized the Weber number in his 1919 work on similitude mechanics, Die Grundlagen der Ähnlichkeitsmechanik und ihre Verwertung bei Modellversuchen. In this treatise, Weber introduced the dimensionless group to quantify the balance between inertial forces and surface tension in fluid flows, particularly in the context of capillary waves and ship hydrodynamics. His formulation arose from applying to problems involving free surfaces, enabling predictions of wave stability and droplet formation without scale-dependent discrepancies. Weber's initial applications focused on analyzing surface waves and interfacial stability in naval mechanics, where the number helped correlate model-scale experiments with prototype behaviors in water wave resistance and spray dynamics. Preceding Weber's formalization, related dimensionless considerations appeared in Lord Rayleigh's 1879 analysis of liquid jet instability, where the ratio of inertial to surface tension effects governed breakup wavelengths, though not explicitly as a named parameter. Weber's contribution built upon this by integrating it into a broader similitude framework for practical engineering applications.

Naming and Legacy

The Weber number is named after Moritz Weber (1871–1951), a German professor of naval mechanics at the Technical University of Berlin, who formalized its use as a dimensionless parameter in similitude analysis and model studies during the early 20th century. In his seminal 1919 paper, Weber developed the parameter as a ratio characterizing capillary effects relative to inertial forces, though it was later designated in his honor. Weber also introduced the term "similitude" to denote the principle of dynamic similarity achieved through such dimensionless groups in physical modeling. Post-World War II, the Weber number saw widespread adoption in engineering literature as computational and experimental techniques advanced, solidifying its role in analyzing interfacial phenomena. It appeared prominently in standard references, including the CRC Handbook of Chemistry and Physics from the 1980s onward, where it is listed among key dimensionless quantities for fluid systems. This recognition stemmed from its utility in scaling complex flows, building on Weber's foundational contributions to similarity principles outlined in his 1930 work on the general similitude principle in physics. The legacy of the Weber number endures in its profound influence on multiphase flow modeling, where it provides a critical balance between inertial and surface tension forces to predict interface stability and breakup. It remains indispensable in (CFD) simulations and experimental setups for deriving scaling laws in diverse systems, from macroscale industrial processes to nanoscale interactions. Notable milestones highlight its evolution: it achieved first widespread application in the 1950s within , notably for simulating rocket fuel atomization and spray dynamics in propulsion systems. In more recent decades, since the 2000s, the parameter has been extended to microscale flows, such as in for controlling droplet formation and transport in lab-on-a-chip devices.

Mathematical Formulation

General Expression

The Weber number We is expressed in its standard form as We = \frac{\rho v^2 L}{\sigma}, where \rho is the density of the fluid, v is the characteristic velocity of the flow, L is the characteristic length scale associated with the fluid interface (such as a droplet diameter or jet radius), and \sigma is the surface tension at the interface. This formulation quantifies the ratio of inertial forces to surface tension forces within multiphase flows. The nondimensional Weber number emerges from scaling analysis of the relevant pressures: the inertial pressure scales as \rho v^2, while the capillary pressure scales as \sigma / L, yielding We as their ratio to compare the dominance of each in deforming interfaces. In scenarios involving two immiscible fluids, the expression accounts for the relative motion across the interface by replacing the velocity with the difference between the phase velocities: We = \frac{\rho (v_1 - v_2)^2 L}{\sigma}, where v_1 and v_2 are the velocities of the two fluids. For gas bubbles rising in a liquid, the characteristic length L is conventionally the bubble diameter, emphasizing deformation under ambient flow. Similarly, for liquid jets issuing into a surrounding medium, L is taken as the jet radius to capture surface wave instabilities leading to breakup. Critical values of the Weber number delineate regimes of stability and instability at interfaces; for instance, in inviscid liquid bridge pinch-off, We \approx 4 signals the onset of capillary-driven rupture.

Derivation via Buckingham Pi Theorem

The derivation of the Weber number employs the Buckingham Pi theorem, a method of dimensional analysis that identifies dimensionless groups from the variables governing a physical phenomenon. This approach is particularly useful for problems involving free-surface flows where surface tension plays a role, such as the breakup of liquid jets or droplet formation. For the Weber number, the analysis assumes an inviscid flow (neglecting viscosity to focus on inertia and surface tension) with a free surface, and gravity effects are excluded (as they lead to a separate dimensionless group, the Froude number). The relevant variables are the fluid density \rho, a characteristic velocity v, a characteristic length scale L, and the surface tension \sigma. These have the following dimensions in terms of mass (M), length (L), and time (T): [\rho] = \mathrm{M L^{-3}}, = \mathrm{L T^{-1}}, [L] = \mathrm{L}, and [\sigma] = \mathrm{M T^{-2}}. There are n = 4 variables and j = 3 fundamental dimensions, so the Buckingham Pi theorem predicts k = n - j = 1 dimensionless \pi group. To form the \pi group, select the repeating variables \rho, v, and L (which collectively span the three dimensions without forming a dimensionless combination on their own). The non-repeating variable \sigma is combined as \pi = \sigma \rho^a v^b L^c, where the exponents a, b, and c are chosen to render \pi dimensionless: [\pi] = [\sigma] [\rho]^a ^b [L]^c = \mathrm{M T^{-2}} \cdot (\mathrm{M L^{-3}})^a \cdot (\mathrm{L T^{-1}})^b \cdot \mathrm{L}^c = \mathrm{M}^{1+a} \mathrm{L}^{-3a + b + c} \mathrm{T}^{-2 - b}. Setting the exponents to zero yields the system of equations:
  • For M: $1 + a = 0 \implies a = -1,
  • For T: -2 - b = 0 \implies b = -2,
  • For L: -3a + b + c = 0 \implies -3(-1) + (-2) + c = 0 \implies 3 - 2 + c = 0 \implies c = -1.
Thus, \pi = \sigma \rho^{-1} v^{-2} L^{-1} = \frac{\sigma}{\rho v^2 L}. The reciprocal of this \pi group defines the Weber number: \mathrm{We} = \frac{1}{\pi} = \frac{\rho v^2 L}{\sigma}, which emerges as the sole dimensionless parameter characterizing the balance between inertial and surface tension forces in the system. This confirms the necessity of We for dynamic similarity in such flows, as any physically similar systems must match in We to ensure comparable behavior.

Role in Fluid Equations

Appearance in Navier-Stokes Equations

The Navier-Stokes equations governing incompressible, Newtonian fluid flows are nondimensionalized using characteristic scales for density \rho, velocity U, length L, and time L/U. The dimensional momentum equation \rho (\partial \mathbf{u}/\partial t + \mathbf{u} \cdot \nabla \mathbf{u}) = -\nabla p + \mu \nabla^2 \mathbf{u} transforms into its nondimensional form \partial \mathbf{u}^*/\partial t^* + \mathbf{u}^* \cdot \nabla^* \mathbf{u}^* = -\nabla^* p^* + (1/\mathrm{Re}) \nabla^{*2} \mathbf{u}^*, where asterisks denote nondimensional variables, \mathrm{Re} = \rho U L / \mu is the , and pressure is scaled by \rho U^2. In the bulk fluid away from interfaces, surface tension does not appear explicitly, as it acts solely at free surfaces or phase boundaries. However, for multiphase flows with interfaces, surface tension is incorporated into the governing equations to capture interfacial dynamics. In models of multiphase flows, such as those using sharp interface representations, surface tension manifests as a curvature-driven stress term in the momentum equation. The dimensional surface tension force is \mathbf{f}_\sigma = \sigma \kappa \mathbf{n} \delta_\Gamma, where \sigma is the surface tension coefficient, \kappa is the interfacial curvature, \mathbf{n} is the unit normal to the interface \Gamma, and \delta_\Gamma is the Dirac delta function concentrated on the interface. Upon nondimensionalization, this term scales as \mathbf{f}_\sigma^* = (\sigma / (\rho U^2 L)) \kappa^* \mathbf{n} \delta_\Gamma^* = \mathrm{We}^{-1} \kappa^* \mathbf{n} \delta_\Gamma^*, where \mathrm{We} = \rho U^2 L / \sigma is the , and starred quantities are scaled (e.g., \kappa^* = \kappa L). Thus, the nondimensional momentum equation includes an additional term \mathrm{We}^{-1} \kappa^* \mathbf{n} \delta_\Gamma^*, highlighting the balance between inertial forces and surface tension. This formulation assumes constant \sigma, neglecting Marangoni effects from gradients in \sigma. For numerical stability and conservation properties in sharp interface models, an equivalent conservative form is often employed, where the surface tension contribution appears in the divergence of a capillary stress tensor. Specifically, the force can be expressed as \mathbf{f}_\sigma = \nabla \cdot [\sigma (I - \mathbf{n} \otimes \mathbf{n}) \delta_\Gamma], which integrates to the same \sigma \kappa \mathbf{n} \delta_\Gamma under suitable conditions, since \kappa \mathbf{n} = \nabla_\Gamma \cdot (I - \mathbf{n} \otimes \mathbf{n}) relates to the surface divergence. In nondimensional terms, this becomes \mathrm{We}^{-1} \nabla^* \cdot [(I - \mathbf{n}^* \otimes \mathbf{n}^*) \delta_\Gamma^*], or equivalently involving \mathrm{We}^{-1} \nabla^* \cdot (\mathbf{n}^* \otimes \mathbf{n}^*) in the stress tensor formulation after integration by parts, ensuring the coefficients the curvature-driven component. These models typically assume high Reynolds numbers where inertial effects dominate viscous dissipation, though the full viscous term (1/\mathrm{Re}) \nabla^{*2} \mathbf{u}^* is retained for completeness.

Free Surface Boundary Conditions

In free surface flows governed by the Navier-Stokes equations, the boundary conditions at the fluid interface consist of kinematic and dynamic components, which incorporate the through nondimensionalization. The kinematic condition ensures that the interface moves with the fluid, expressed as \mathbf{v} \cdot \mathbf{n} = \frac{\partial h}{\partial t} + \mathbf{v} \cdot \nabla h, where \mathbf{v} is the velocity vector, \mathbf{n} is the unit normal to the interface, and h describes the interface height. This condition remains independent of surface tension and thus the Weber number in its dimensional form, but it couples with the dynamic condition to determine overall interface evolution. The dynamic boundary condition enforces stress balance across the interface: -p \mathbf{I} + \boldsymbol{\tau} \cdot \mathbf{n} = \sigma \kappa \mathbf{n}, where p is pressure, \boldsymbol{\tau} is the viscous stress tensor, \sigma is surface tension, and \kappa is the mean curvature. Upon nondimensionalization using characteristic scales for velocity U, length L, density \rho, and pressure \rho U^2, the surface tension term becomes \mathrm{We}^{-1} \sigma \kappa \mathbf{n}, with the Weber number defined as \mathrm{We} = \rho U^2 L / \sigma. This scaling highlights how \mathrm{We}^{-1} weights the relative importance of surface tension forces in the stress balance. The Weber number directly scales the Laplace pressure jump across a curved interface, \Delta p = \sigma / R, where R is the radius of curvature on the order of the characteristic length L. In nondimensional terms, this jump is \Delta \tilde{p} = \mathrm{We}^{-1} / \tilde{R}, emphasizing that surface tension induces a pressure discontinuity proportional to \mathrm{We}^{-1}. For high Weber numbers (\mathrm{We} \gg 1), inertial forces dominate, weakening the enforcement of surface tension and allowing flat interfaces to deform significantly under flow. Conversely, low Weber numbers (\mathrm{We} \ll 1) amplify surface tension effects, enforcing near-spherical or minimally curved shapes to minimize surface energy. In numerical simulations of free surface flows using volume-of-fluid (VOF) or level-set methods, the Weber number parameterizes the surface tension contribution, influencing interface sharpening to accurately resolve curvature-driven dynamics. For instance, in the continuum surface force (CSF) model for VOF, surface tension is distributed as a body force \mathbf{F}_{sv} = \sigma \kappa \nabla C / |\nabla C|, where C is the volume fraction; nondimensionalization yields a term scaled by \mathrm{We}^{-1}, requiring finer grid resolution at low \mathrm{We} to prevent diffusive smearing of sharp interfaces. Similarly, level-set methods incorporate \sigma \kappa \delta(\phi) \mathbf{n} (with \phi the level-set function and \delta the Dirac delta), where high \mathrm{We} permits coarser sharpening due to reduced capillary effects.

Relations to Other Dimensionless Numbers

Comparison with Reynolds and Froude Numbers

The Weber number (We), Reynolds number (Re), and Froude number (Fr) are dimensionless groups that characterize different force balances in fluid flows, particularly in multiphase and interfacial contexts. The Reynolds number, defined as Re = \frac{\rho v L}{\mu}, where \rho is fluid density, v is velocity, L is a characteristic length, and \mu is dynamic viscosity, quantifies the ratio of inertial to viscous forces. In contrast, the Weber number, We = \frac{\rho v^2 L}{\sigma} with \sigma as surface tension, measures inertial forces relative to surface tension forces. Both incorporate the inertial term \rho v^2 L, but Re normalizes it by viscous effects (\mu / L), while We does so by capillary effects (\sigma / L), leading to distinct regimes: high Re and low We favor viscous-capillary balances in flows where surface tension resists deformation against weak viscosity, as seen in thin-film coating processes. The Froude number, Fr = \frac{v}{\sqrt{g L}} where g is gravitational acceleration, represents the balance between inertial and gravitational forces, dominating in free-surface flows like waves or hydraulic jumps. Unlike Fr, which emphasizes gravity-driven phenomena such as spilling breakers in open channels, We highlights capillary-driven instabilities, such as droplet formation where surface tension counters inertia. Their interplay arises in regimes where both gravity and surface tension influence outcomes; for instance, the combined parameter We \cdot Fr^2 = \frac{\rho v^4}{g \sigma} emerges in analyzing total energy balances at interfaces in wave dynamics. In practical multiphase flows, such as open-channel or jet impingement scenarios, interaction regimes are delineated by these numbers: low We coupled with high Fr indicates gravity-dominated propagation with minimal capillary disruption, whereas high We promotes splashing or atomization driven by inertia overcoming surface tension, often independent of Fr if gravity is negligible. These distinctions guide experimental scaling, where matching We, Re, and Fr ensures dynamic similitude for interfacial flows, preventing distortions in model tests of phenomena like ship wakes or bubble columns.

Connection to Ohnesorge Number

The Ohnesorge number, denoted as Oh, is a dimensionless quantity defined as Oh = \frac{\mu}{\sqrt{\rho \sigma L}}, where \mu is the dynamic viscosity of the fluid, \rho is its density, \sigma is the surface tension, and L is a characteristic length scale such as the jet diameter or droplet radius. This formulation captures the ratio of viscous forces to the combined effects of inertial and surface tension forces, providing a measure of how viscosity influences capillary-dominated flows relative to inertia and surface tension. In terms of other dimensionless numbers, Oh relates directly to the (We = \frac{\rho v^2 L}{\sigma}) and the (Re = \frac{\rho v L}{\mu}) through the expression Oh = \frac{\sqrt{We}}{Re}, where v is a characteristic velocity. The derivation of this relationship stems from combining the definitions of We and Re. Starting with We = \frac{\rho v^2 L}{\sigma}, the square root yields \sqrt{We} = v \sqrt{\frac{\rho L}{\sigma}}. Dividing by Re = \frac{\rho v L}{\mu} gives \frac{\sqrt{We}}{Re} = \frac{v \sqrt{\frac{\rho L}{\sigma}}}{\frac{\rho v L}{\mu}} = \frac{\mu \sqrt{\frac{\rho L}{\sigma}}}{\rho L} = \frac{\mu}{\sqrt{\rho \sigma L}}, confirming Oh = \frac{\sqrt{We}}{Re}. This viscous ratio highlights scenarios where capillary and inertial effects dominate over viscosity, typically when Oh < 0.1, indicating low viscosity influence and prevalence of inertial breakup mechanisms. Conversely, higher values signal increasing viscous dominance. The significance of Oh lies in its ability to refine the analysis provided by We alone by incorporating viscosity, particularly in classifying spray and jet breakup regimes. While We scales the balance between inertia and surface tension, Oh distinguishes inertial breakup (low Oh) from viscous breakup ( Oh > 0.1 ), enabling prediction of droplet formation stability in multiphase flows. For instance, in spray , low Oh values promote inertial mechanisms leading to fine droplets, whereas Oh > 0.1 shifts toward viscous-dominated fragmentation. Specific flow regimes illustrate this connection: high Oh combined with low We characterizes dripping faucets, where viscous forces and suppress inertial jetting, resulting in periodic drop detachment rather than continuous streams. In applications like paint spraying, We alone often fails to capture breakup due to varying viscosities; incorporating Oh accounts for these effects, predicting coarser droplets in high-viscosity paints and aiding process optimization.

Applications

Multiphase and Interfacial Flows

In multiphase and interfacial flows, the Weber number serves as a key dimensionless parameter that characterizes the competition between inertial forces driving deformation and forces resisting it, enabling predictions of and in two-phase systems such as gas-liquid mixtures and liquid-liquid dispersions. When We << 1, dominates, maintaining compact shapes, whereas We > 1 signifies inertial dominance, promoting deformation that enhances mixing and efficiency in processes like emulsification. In gas-liquid flows, the Weber number critically influences bubble rise dynamics and stability; at low We (typically below 1), bubbles adopt spherical shapes with minimal deformation, resulting in predictable terminal velocities governed primarily by and . As We increases to order 1 or higher, inertial effects cause bubble deformation into ellipsoidal or cap-like forms, which alters drag coefficients and can lead to path instabilities in flows, such as those in bubbly reactors where jet-like structures form and disrupt under high relative velocities. These transitions are essential for optimizing and gas holdup in industrial applications. For liquid-liquid interfacial flows, including , the Weber number determines thresholds for droplet deformation, coalescence, and under or turbulent conditions; initiates when We surpasses a critical value often ranging from 1 to 10, influenced by the ratio of the phases, allowing dispersed droplets to fragment and achieve finer distributions that stabilize the . At We ≈ 1–10 in flows, interfacial stretching overcomes , facilitating coalescence avoidance and enhanced mixing, as seen in stirred tank emulsifiers. The Weber number also underpins scaling analyses for dynamic similarity in multiphase reactors, ensuring that interfacial phenomena observed in models translate to prototypes by matching We alongside Reynolds and Froude numbers, thereby validating predictions of deformation and phase interactions across scales without complications.

Droplet Breakup and Atomization

The Weber number plays a crucial role in determining the deformation and breakup of liquid droplets subjected to aerodynamic forces, particularly in high-speed gas flows where inertial effects compete with . In multiphase flows, droplets deform when the Weber number exceeds a critical , leading to and fragmentation into smaller daughter droplets. This process is essential for understanding in applications like spray combustion and aerosol generation. Droplet breakup occurs in distinct regimes classified primarily by the Weber number and the gas-to-liquid density ratio. For low density ratios (typically ρ_g/ρ_l < 0.1, as in air-water systems), bag breakup dominates in the moderate range of approximately 10 to 100, where the droplet flattens into a thin bag-like structure that bursts, producing a distribution of smaller droplets. This regime is characterized by the growth of a toroidal rim that supports the expanding bag before rupture. At higher Weber numbers exceeding 100, particularly in high-velocity winds, the shear stripping regime prevails, wherein the droplet boundary layer thins due to intense shear, stripping off small droplets from the periphery while the core remains intact longer. These transitions are well-documented in experimental studies using shock tubes to initiate sudden accelerations. In atomization processes, such as those in fuel injectors, the Weber number governs the transition from primary to secondary breakup. Primary breakup of liquid jets into ligaments and initial droplets requires a gas Weber number on the order of 8 to 20, marking the onset of wind-induced instabilities beyond the Rayleigh-Plateau regime, which leads to the formation of a spray with characteristic drop sizes. Secondary breakup further refines the spray, incorporating modes influenced by the Weber number; for instance, at low to moderate values, Rayleigh-Plateau-like instabilities can contribute to filament thinning and rupture in the resulting droplets. Seminal regime maps, such as the Hsiang-Faeth diagram, delineate these behaviors as functions of the Weber number and density ratio, predicting boundaries like bag breakup for We ≈ 12–50 and shear stripping for We > 350 at high density ratios (ρ_l/ρ_g > 500). Practical examples illustrate the Weber number's impact on droplet stability. For raindrops falling at , the Weber number balances aerodynamic drag against , with deformation becoming noticeable around We ≈ 5, leading to oblate shapes for larger drops without immediate breakup. In inkjet printing, low Weber numbers (typically We < 4) ensure stable droplet ejection without premature fragmentation or satellite formation, allowing precise deposition by maintaining ligament integrity during flight. These cases highlight how the Weber number establishes scale for inertial disruption in environmental and engineering contexts.

Heat Pipes and Capillary Phenomena

In heat pipes, the governs the entrainment limit, a critical operational boundary where high vapor velocities can disrupt the liquid return through the wick structure. When the exceeds unity (We ≥ 1), inertial forces from the vapor flow overcome surface tension, causing liquid droplets to be entrained from the pores of the wick, with the characteristic length scale typically taken as the pore size. This phenomenon limits the maximum heat transport capacity, as entrainment floods the evaporator and reduces thermal performance. In capillary flows within tubes, the Weber number predicts the stability of the during liquid rise or displacement. For low Weber numbers (We < 1), surface tension dominates over inertia, enabling smooth, stable advancement as seen in traditional imbibition processes. Conversely, high Weber numbers (We > 1) introduce significant inertial effects that destabilize the , potentially leading to flooding where the liquid interface breaks and overflows, compromising flow control in confined geometries. For thin liquid films and sheets, such as those in falling film flows, the Weber number indicates the onset of surface instabilities like wave formation. When We > 1, inertial forces promote the growth of surface , transitioning the flow from a uniform laminar state to an unstable wavy regime that enhances mixing but can lead to film rupture if unchecked. This threshold is particularly relevant in moderate flows where surface tension's stabilizing role diminishes. In applications, the Weber number scales wicking limits in micro-heat exchangers, where We ≈ 1 delineates the for effective pumping in porous wicks without disrupting efficiency. Similarly, in flows, low Weber numbers favor uniform film deposition by maintaining dominance, while higher values (We > 1) trigger beading instabilities, resulting in non-uniform that affect product quality in processes like or treatment.

References

  1. [1]
    Weber number | KRÜSS Scientific
    The Weber number is a characteristic number used in fluid mechanics. As a dimensionless quantity, it describes the ratio between deforming inertial forces ...
  2. [2]
    Weber Number - The Engineering ToolBox
    The Weber Number is a dimensionless value useful for analyzing fluid flows where there is an interface between two different fluids.
  3. [3]
    Weber number - Thermopedia
    Feb 2, 2011 · It represents the ratio of inertial force to surface tension force. There are several alternative forms of Weber number, including the square ...
  4. [4]
    Bubble breakup reduced to a one-dimensional nonlinear oscillator
    A Weber number of order unity separates stable (low We) from unstable (large We) bubbles. The effect of viscosity, on its side is quantified by the Reynolds ...
  5. [5]
    Splashing regimes of high-speed drop impact
    Sep 16, 2025 · When a drop impinges onto a deep liquid pool, it can yield various splashing behaviours, leading to a crown-like structure along the free ...
  6. [6]
    Investigation of interface deformation dynamics during high-Weber ...
    RTP is the main instability mode for small Weber numbers, and SIE is the terminal instability mode for increasing Weber numbers. The RTP regime is ...
  7. [7]
    Droplet breakup and size distribution in an airstream: Effect of inertia
    The critical Weber number ( We c r ) for the transition from the vibrational to the bag breakup is approximately 12.
  8. [8]
    Investigation of performance limits in axial groove heat pipes
    In this analysis, the critical value for the entrainment Weber number is found to be 2 pi less than or equal to 3 pi. Perhaps more important to the heat pipe ...
  9. [9]
    [PDF] Note on the History of the Reynolds Number
    In 1908, Arnold Sommerfeld presented a paper on hydrodynamic stability at the 4th International Congress of Mathematicians in Rome (Sommerfeld.
  10. [10]
    Moritz Weber's Fluid Dynamics - Resolved Analytics
    Moritz Weber made significant contributions to wave theory, including his work on the Weber number describing the balance between inertial and surface tension
  11. [11]
    Die Grundlagen der Ähnlichkeitsmechanik - SpringerLink
    Moritz Weber (ord. Professor der Mechanik). Authors ... About this chapter. Cite this chapter. Weber, M. (1919). Die Grundlagen der Ähnlichkeitsmechanik.
  12. [12]
    [PDF] On The Instability Of Jets - Semantic Scholar
    On The Instability Of Jets · L. Rayleigh · Published 1 November 1878 · Mathematics, Physics · Proceedings of The London Mathematical Society.
  13. [13]
    Principal contributors to fluid mechanics
    Moritz Weber (1871–1951) A German professor of naval mechanics who is credited with formalising the use of non-dimensional groups as the basis for similarity ...Missing: similitude | Show results with:similitude
  14. [14]
    Dimensional Analysis and Scale-up in Chemical Engineering
    Moritz Weber comprehensively described these techniques of deriving dimensionless numbers in his two fundamental papers of 1919 and 1930 [A 1, A 2]. See ...<|control11|><|separator|>
  15. [15]
    [PDF] Cold Flow Testing for Liquid Propellant Rocket Injector Scaling and ...
    Consequently, the dimensionless groups, namely the Weber number, are generally smaller for water than for LOX for a given set of operating flow rates and ...
  16. [16]
    [PDF] PREDICTION OF THE BUBBLE SIZE GENERATED BY A ...
    In this paper the critical Weber number is mlated to the average energy dissipation rate per unit volume within the liquid to predict the maximum bubble ...
  17. [17]
    [PDF] Role of Weber number in primary breakup of turbulent liquid jets in ...
    2008, 2009), we demonstrated that the liquid Weber number, rather than the crossflow Weber number, controls the wavelength of liquid surface insta- bilities on ...
  18. [18]
    Breakup of diminutive Rayleigh jets | Physics of Fluids - AIP Publishing
    The vertical axis shows the Weber number We = ρ v 2 r / γ (at the left) and the jetting velocity v (at the right). The lower critical velocity for jet formation ...
  19. [19]
    None
    ### Summary of Buckingham Pi Theorem Application for Weber Number in Fluid Mechanics
  20. [20]
    None
    ### Derivation of Weber Number Using Buckingham Pi Theorem
  21. [21]
    [PDF] Notes on Thermodynamics, Fluid Mechanics, and Gas Dynamics
    Dec 15, 2021 · (3) Determine the number of Π terms using the Buckingham-Pi Theorem. (a) (# of Π terms) = (# of variables) – (# of reference dimensions). (b) ...
  22. [22]
    [PDF] A comparison of different approaches to compute surface tension ...
    Keywords: Navier-Stokes equations, Incompressible flows, Two-phase flows ... ∇nn+γ. Γ. : ∇un+γ. 2. +. 1. 2. 1. 2. ∇nn. Γ : ∇un. (29). +. 1. 2. 1. 2. ∇·. I − n n+γ.
  23. [23]
    A Coupled Level Set and Volume-of-Fluid Method for Computing 3D ...
    We present a coupled level set/volume-of-fluid (CLSVOF) method for computing 3D and axisymmetric incompressible two-phase flows.
  24. [24]
    Advanced partial differential equation models
    where We is the Weber number,. We=ϱU2Lσ. The Weber number measures the importance of surface tension effects and is the ratio of the pressure scale ϱ ...
  25. [25]
    [PDF] 2.2 Similarity Parameters from Governing Equations and Boundary ...
    In this paragraph we will see how we can specify the SP's for a problem that is governed by the Navier-Stokes equations. The SP's are obtained by scaling, non- ...
  26. [26]
    [PDF] Wolfgang von Ohnesorge - MIT
    Aug 24, 2011 · Following Ohnesorge's lead, it makes physical sense to isolate dy- namical effects into a single dimensionless variable (e.g. a Reynolds number,.
  27. [27]
    A Numerical Investigation on Droplet Bag Breakup Behavior of ... - NIH
    Sep 23, 2020 · For a low Ohnesorge number (Oh < 0.1), with increasing the We number, the five breakup typical breakup modes occur: vibrational, bag, multimode ...
  28. [28]
    [PDF] Droplet Breakup in Automotive Spray Painting
    The implication by the large Ohnesorge numbers is that the droplets get successively more difficult to break with the scaling a1. 8 in the asymptotic limit.
  29. [29]
    The applicability of dynamic-similarity criteria to isothermal, liquid ...
    The applicability of dynamic-similarity criteria to isothermal, liquid-gas, two-phase flows without mass transfer ... View PDFView articleView in Scopus ...
  30. [30]
    [PDF] 1 Emulsion Formation, Stability, and Rheology - Wiley-VCH
    An important parameter that describes droplet deformation is the Weber number ... This is consistent with the increase of stress with increasing φ. The drop ...
  31. [31]
    Scaling of Lift Reversal of Deformed Bubbles in Air-Water Systems
    ... (Clift, Grace, Weber, 1978, Tomiyama, Kataoka, Zun, Sakaguchi, 1998). Since ... bubble deformation. The following three effects are taken into account ...
  32. [32]
    Dispersion of water into oil in a rotor–stator mixer. Part 1
    Drop breakup in complex laminar flows is poorly understood due to the ... Weber Number correlation which only applies for L ≫ d ≫ η, where L is the ...
  33. [33]
    Near-limit drop deformation and secondary breakup - ScienceDirect
    The properties of drop deformation and secondary breakup were observed for shock wave initiated disturbances in air at normal temperature and pressure.
  34. [34]
    Modeling of drop breakup in the bag breakup regime - AIP Publishing
    Apr 17, 2014 · In this Letter, study will be focused on the bag-type breakup (BTB) and the corresponding Weber number is in the range of 10 < We < 35. The ...
  35. [35]
    Effect of Mach number on droplet aerobreakup in shear stripping ...
    Aug 9, 2020 · The Weber number, which is constrained at 1100 with a standard deviation of 50, lies within the stripping breakup regime and allows a wide ...
  36. [36]
    A breakup model for transient Diesel fuel sprays - ScienceDirect.com
    In this paper a breakup model for analysing the evolution of transient fuel sprays characterised by a coherent liquid core emerging from the injection nozzle.
  37. [37]
    A deformable liquid drop falling through a quiescent gas at terminal ...
    Jun 28, 2010 · Visually noticeable drop deformation is shown to occur when the Weber number We ~ 5. For μ ≥ 100, drops of Reynolds number Re < 200 tend to ...
  38. [38]
    Physicochemical parameters that underlie inkjet printing for medical ...
    3). If the Weber number is low, the generated droplet does not break up, remains suspended, and periodically oscillates at the tip of the nozzle (region A).
  39. [39]
    FLUID MECHANICS OF HEAT PIPES - Annual Reviews
    limit. The entrainment phenomenon is governed by the Weber number that compares the vapor inertial forces to the liquid surface-tension forces.
  40. [40]
    Analysis of the critical Weber number at the onset of liquid ...
    Entrainment phenomena in capillary-driven heat pipes were studied and both analytical and experimental approaches were utilized to identify and better ...
  41. [41]
    Entrainment Limits in Heat Pipes | AIAA Journal
    Entrainment Limits in Heat Pipes. C.L. Tien and ... Analysis of the critical Weber number at the onset of liquid entrainment in capillary-driven heat pipes.
  42. [42]
    On the dynamics of a meniscus inside capillaries during imbibition ...
    Aug 4, 2021 · Weber number can be large if the combination of ρ , U 2 , and L is large. For typical open-channel flows, all these reference parameters are ...
  43. [43]
    Surface equation of falling film flows with moderate Reynolds ...
    Nov 1, 1999 · When the Reynolds number R is as small as unity, the present equation agrees with the Navier–Stokes equation and also with the traditional long- ...
  44. [44]
    [PDF] Exploring the Limits of Boiling and Evaporative Heat Transfer Using ...
    where d is a characteristic length scale of the wick structure. It is assumed that the entrainment will be serious when the Weber number, We, is ~ 1 [67].
  45. [45]
    [PDF] Bead-on-fibre morphology in shear-thinning flow - Clemson University
    Fibre coating processes benefit from this shear-thinning effect as highly shear-thinning paints can form asymmetric bead-on-fibre patterns, which flow faster ...