Fact-checked by Grok 2 weeks ago

Electrohydrodynamics

Electrohydrodynamics (EHD), also known as electro-fluid-dynamics, is the study of the interaction between and fluid flows, encompassing the effects of electric forces on the motion, deformation, and stability of electrically conducting or polarizable fluids such as liquids and gases. This interdisciplinary field combines principles from , , and continuum physics to describe phenomena where electric fields induce charge separation, accumulation at interfaces, and resultant stresses that drive bulk flows, interfacial deformations, and instabilities. The foundational theoretical framework for EHD was established in the mid-20th century, building on early observations dating back to the ; William Gilbert documented fluid motion under electric influence in 1600, while coined the term "electrohydrodynamics" in 1964 and developed the leaky dielectric model in 1966 to explain droplet deformation in insulating liquids under uniform . In this model, fluids are treated as imperfect dielectrics with finite , leading to free charge accumulation at fluid interfaces that generates tangential electro-osmotic stresses and normal Maxwell stresses, causing prolate or oblate deformations depending on ratios of (S) and (R) between phases. Key governing equations include the Navier-Stokes equations augmented by electric body forces (via and dielectrophoretic terms) and for charge conservation, often simplified under the regime for low Reynolds numbers. Notable EHD phenomena include electrohydrodynamic instabilities such as tip streaming (where cones form at droplet poles leading to jet ejection), Quincke rotation (spontaneous droplet spinning above a critical field strength), and varicose or whipping modes in charged jets that affect breakup into monodisperse droplets. These effects are quantified by dimensionless numbers like the electric Reynolds number (Re_E = \frac{\varepsilon^2 E^2}{\mu \sigma}, measuring charge convection relative to conduction and viscous effects, where \sigma is conductivity) and the electric capillary number (Ca_E = \varepsilon E^2 a / \gamma, comparing electric to surface tension stresses), where \varepsilon is permittivity, E is field strength, \mu is viscosity, \gamma is interfacial tension, and a is characteristic length. EHD has diverse applications across and , including microfluidic pumping without parts (via ion-drag or conduction mechanisms), high-resolution electrohydrodynamic printing for micro/nanofabrication (achieving feature sizes below 100 nm through Taylor cone- modes), emulsion stabilization or breakup for and , and systems like ionic thrusters that convert directly to kinetic in liquids or gases. As of 2020, advances leverage EHD for enhanced in cooling, particle manipulation in (e.g., cell sorting), and sustainable energy technologies such as electrostatic precipitators for air purification. As of 2025, further developments include multimodal EHD printing for high-resolution sensor fabrication and resilient flexible EHD pumps for human-machine interfaces.

Introduction

Definition and Scope

Electrohydrodynamics (EHD) is the study of the interactions between and fluid motion in electrically conducting or polarizable fluids, encompassing phenomena such as charge injection, , and the resulting hydrodynamic effects. This interdisciplinary field integrates principles from hydrodynamics, , , and thermophysics, primarily focusing on weakly conducting liquids and gases where electric forces couple effectively with viscous forces. The scope of EHD extends to both and conducting fluids, with particular emphasis on liquid dielectrics like oils exhibiting conductivities in the range of 10^{-12} to 10^{-7} S/m, where induce flows, instabilities, and enhanced without requiring mechanical components. Central to EHD are the electric forces acting on the fluid: the Coulomb force, which exerts on free charges within the fluid (q\mathbf{E}, where q is and \mathbf{E} is the ); the dielectrophoretic force, arising from gradients in the electric field acting on induced dipoles in polarizable ; and , a volumetric force due to electric field-induced density changes in the fluid. These forces drive fluid motion by coupling with the Navier-Stokes equations, often at micro- and nanoscale regimes where surface effects dominate. EHD phenomena are broadly classified into injection-induced flows, such as ion-drag effects where charges are injected from electrodes to propel the fluid, and polarization-induced flows, exemplified by dielectrophoresis where non-uniform fields manipulate or weakly charged particles via . Unlike (MHD), which addresses fluid motion under magnetic fields in highly conducting plasmas or liquids (with magnetic Reynolds numbers of order unity), EHD emphasizes in weakly conducting where magnetic effects are negligible (σ ε_0 c^2 L^2 ≪ 1, with L as characteristic length). , such as and , constitute specific subsets of EHD involving relative motion between immiscible fluid phases or solids and electrolytes near charged interfaces.

Historical Development

The earliest observations of electrohydrodynamic phenomena date back to the late 16th and early 17th centuries, when scientists began documenting the motion of liquids and particles under . In 1600, William Gilbert described the attraction and movement of liquid droplets toward charged objects, such as rubbed , in his seminal work , marking one of the first recorded instances of electric forces influencing fluid behavior. Similarly, in 1629, Niccolò Cabeo observed the attraction of small particles, like , to electrified bodies, followed by contact and repulsion, providing early evidence of electrodynamic interactions with in air. During the 18th and 19th centuries, foundational work on laid the groundwork for understanding electrokinetic effects in fluids. Alessandro Volta's invention of the in 1800 enabled sustained electric currents, facilitating experiments that revealed electrochemical reactions in liquids. advanced this in the 1830s through his studies of in water and other electrolytes, where he quantified the decomposition of fluids under and observed associated motion of charged species, establishing key principles of electrokinetic transport. The saw the formal development of electrokinetics, with quantitative studies on the of colloidal particles providing insights into particle migration in within fluids. of electrohydrodynamics was rigorously defined in 1969 through a landmark review by J.R. Melcher and , which integrated electrodynamics and hydrodynamics to explain interfacial shear stresses and fluid motion driven by . had coined the term "electrohydrodynamics" in 1964. Following 1969, research emphasized electrohydrodynamic instabilities using the leaky dielectric model developed by in 1966 and extended by Melcher, which accounted for finite in fluids and predicted deformation and circulation in droplets under . In the 1990s and 2000s, electrohydrodynamics integrated with and , enabling precise control of fluid flows at microscales for applications like droplet manipulation and . Recent developments as of 2024 have advanced computational modeling and applications of electrohydrodynamic flows in systems, with key reviews highlighting their role in biological contexts such as vesicle dynamics and cellular processes, as well as enhanced and pumping technologies.

Fundamental Principles

Governing Equations

Electrohydrodynamics couples electromagnetic fields with fluid motion through the interaction of electric charges and fields within conducting or polarizable fluids. The governing equations are derived under the quasi-electrostatic approximation, suitable for low-frequency phenomena where magnetic effects are negligible. The electric field \mathbf{E} is irrotational, satisfying \nabla \times \mathbf{E} = 0, which allows representation as \mathbf{E} = -\nabla \phi with \phi the electric potential. Gauss's law for the electric displacement \mathbf{D} = \varepsilon \mathbf{E}, where \varepsilon is the permittivity, takes the form \nabla \cdot \mathbf{D} = \rho_f, with \rho_f denoting the free charge density. Charge conservation is expressed as \frac{\partial \rho_f}{\partial t} + \nabla \cdot \mathbf{J} = 0, where the \mathbf{J} includes ohmic conduction, convection by fluid velocity \mathbf{v}, and diffusive contributions: \mathbf{J} = \sigma \mathbf{E} + \rho_f \mathbf{v} - D \nabla \rho_f, with \sigma the electrical conductivity and D the diffusion coefficient (often negligible in macroscopic EHD flows). The fluid momentum is governed by the Navier-Stokes equations augmented with electric body forces: \rho \left( \frac{\partial \mathbf{v}}{\partial t} + \mathbf{v} \cdot \nabla \mathbf{v} \right) = -\nabla [p](/page/Pressure) + \mu \nabla^2 \mathbf{v} + \rho_f \mathbf{E} - \frac{1}{2} E^2 \nabla \varepsilon + \nabla \left[ \frac{1}{2} E^2 \left( \frac{\partial \varepsilon}{\partial \rho} \right) \rho \right], where \rho is mass density, p , \mu dynamic viscosity, the term \rho_f \mathbf{E} is the force, -\frac{1}{2} E^2 \nabla \varepsilon the dielectrophoretic force, and the final term the electrostrictive force (relevant in compressible fluids). The \nabla \cdot \mathbf{v} = 0 holds for incompressible flows. Boundary conditions include no-slip \mathbf{v} = 0 at solid walls and, at fluid interfaces, continuity of tangential \mathbf{E} and normal \mathbf{D} \cdot \mathbf{n} (adjusted for surface charge), alongside kinematic conditions for interface tracking. Non-dimensionalization reveals key regimes via the electric \mathrm{Re}_E = \varepsilon E_0^2 / (\mu U), comparing electric to viscous stresses (with E_0 characteristic and U scale), and the charge relaxation time \tau_\varepsilon = \varepsilon / \sigma, contrasting conduction and convection timescales. Low \mathrm{Re}_E and rapid relaxation (\tau_\varepsilon \ll flow time) yield ohmic conduction dominance, while high \mathrm{Re}_E or slow relaxation favor injection-dominated flows. These parameters delineate behaviors such as leaky dielectric models in non-aqueous systems versus electrokinetic effects in confined aqueous geometries.

Electric Forces in Fluids

In electrohydrodynamics, electric forces acting on fluids originate from the coupling between electromagnetic fields and the material properties of the fluid, such as , , and , resulting in both volumetric body forces and interfacial surface forces that induce or modify fluid motion. These forces are fundamental to EHD phenomena and can be derived from combined with thermodynamic considerations of the electric energy in the fluid. Body forces include the Coulomb force, which acts on free charges within the fluid volume and is expressed as \mathbf{f}_C = \rho_f \mathbf{E}, where \rho_f is the volume of free charge and \mathbf{E} is the vector. In nonuniform , an additional dielectrophoretic force arises from the interaction with , particularly relevant for fields, given by \mathbf{f}_{DEP} = \frac{1}{2} \Re \left[ \nabla (\alpha |\mathbf{E}|^2) \right], where \alpha denotes the complex of the fluid, accounting for both real and imaginary components related to and . Surface forces at fluid interfaces or electrodes are primarily described by the divergence of the Maxwell stress tensor, \boldsymbol{\tau}_M = \varepsilon \left( \mathbf{E} \otimes \mathbf{E} - \frac{1}{2} |\mathbf{E}|^2 \mathbf{I} \right), where \varepsilon is the , \mathbf{I} is the identity tensor, and additional polarization terms may contribute at discontinuities in or . This tensor yields tangential stresses that drive flows along interfaces and normal stresses that promote deformation or rupture, such as in electrospraying. Electrostriction introduces a volumetric body force in fluids where permittivity depends on density, formulated as \mathbf{f}_{ES} = -\frac{1}{2} E^2 \nabla \varepsilon, arising from field-induced compression that alters the fluid's dielectric response. In typical liquid EHD systems, this force is often negligible compared to Coulomb or dielectrophoretic effects due to the low compressibility of liquids. EHD processes operate in distinct regimes based on charge generation mechanisms: the injection regime, where free charges are directly emitted from electrodes through processes like , leading to space-charge layers and strong Coulomb-driven flows; and the conduction regime, where charges result from of neutral molecules under the electric field, with transport dominated by ohmic conduction and heterocharge layers near electrodes. In dielectric fluids with low conductivity, polarization-based forces such as dielectrophoresis dominate the force balance, whereas in electrolytes with significant free charge carriers, Coulomb forces prevail and can lead to intense electroconvection. The total electrohydrodynamic can be derived from the variation of the electric density with respect to fluid displacement, yielding \mathbf{f}_{EHD} = \rho_f \mathbf{E} - \frac{1}{2} E^2 \nabla \varepsilon + \frac{1}{2} \nabla \left( E^2 \frac{\partial \varepsilon}{\partial \rho} \rho \right), where the first term captures effects, the second dielectrophoretic contributions from gradients, and the third electrostrictive corrections involving density \rho. These forces enter the Navier-Stokes momentum equation as source terms to govern the overall .

Electrokinetic Phenomena

Electrokinesis

Electrokinesis refers to the bulk motion of a fluid induced by the transfer of from electrically charged ions to neutral molecules, occurring without the presence of phase boundaries or interfaces. This phenomenon arises in liquids where free charges are introduced via unipolar injection from electrodes, leading to forces that drive the overall fluid flow. Unlike interfacial effects, electrokinesis involves volumetric forces distributed throughout the fluid phase. The primary mechanisms of electrokinesis include ion-drag and electroconvection. In the ion-drag process, ions injected at the electrode are accelerated by the applied electric field and collide with surrounding neutral molecules, imparting momentum and thereby entraining the bulk fluid. Electroconvection arises from the Coulomb force on space charge generated by mechanisms such as ion injection or dissociation in applied electric fields, often uniform between electrodes, leading to hydrodynamic instabilities that induce convective patterns, such as plumes or rolls. These mechanisms dominate in poorly conducting liquids, where charge relaxation times are long enough to sustain significant space charge densities. A key dimensionless parameter governing electrokinesis is the injection strength C = \frac{\rho_0 L^2}{\varepsilon V}, where \rho_0 is the injected at the , L is the gap width, \varepsilon is the , and V is the applied voltage. This parameter quantifies the relative importance of injected charge to the field-induced charge; regimes with C > 10 indicate strong injection, where electrokinetic flows become dominant and transition from linear to nonlinear behaviors. Experimental investigations of electrokinesis typically employ insulating oils, such as , confined between parallel plate under high voltages (often exceeding 10 kV). Ion injection occurs at a sharp emitter , creating a that propagates across the gap to a collector. Velocity profiles in these setups exhibit parabolic or plume-like structures, with characteristic speeds scaling as u \sim \frac{\varepsilon}{\mu} \left( \frac{V}{L} \right)^2, derived from balancing the electric against viscous dissipation in the Navier-Stokes equations. Electrokinesis fundamentally differs from dielectrophoresis, as it relies on the conduction of free charges rather than the of dipoles in molecules, enabling sustained flows without requiring alternating fields. In confined geometries, electrokinesis shares conceptual similarities with , where charge-driven slips occur at walls rather than in open .

Electroosmosis and Electrophoresis

Electroosmosis describes the motion of a liquid relative to a stationary charged surface under the influence of an applied tangential , where the field exerts on the counterions within the diffuse part of the electrical double layer at the solid-liquid . This arises in aqueous systems, where the charged surface attracts oppositely charged ions, forming a mobile diffuse layer that shears under the electric field, dragging the bulk fluid. The classical theoretical framework for this flow was established by Helmholtz, who introduced the concept of the double layer, and refined by Smoluchowski, leading to the Helmholtz-Smoluchowski relation for the electroosmotic velocity. Under the thin double-layer approximation, where the Debye length is much smaller than the channel dimension, the electroosmotic velocity is uniform across the channel cross-section and given by u_{eo} = -\frac{\varepsilon \zeta}{\mu} E, where \varepsilon is the electrical of the fluid, \zeta is the at the slipping plane, \mu is the dynamic , and E is the applied strength. This slip velocity represents the electrokinetic coupling coefficient, which quantifies the linear relationship between the flow and the driving field in low-conductivity aqueous electrolytes. Electrophoresis, conversely, involves the motion of charged colloidal particles or macromolecules through a quiescent fluid under an applied , with the velocity formula mirroring that of but with opposite sign due to the relative motion of the particle. In the thin double-layer limit, the electrophoretic velocity is u_{ep} = \frac{\varepsilon \zeta}{\mu} E, assuming no effects or hydrodynamic interactions dominate. This reciprocity between and electrophoresis stems from the underlying electrokinetic mechanism, where the electric force on the particle's double layer drives its translation relative to the surrounding fluid. In aqueous systems, the \lambda_D, which characterizes the thickness of the diffuse double layer, is given by \lambda_D = \sqrt{\frac{\varepsilon [kT](/page/KT)}{2 n e^2}} for a symmetric 1:1 , where k is Boltzmann's constant, T is , n is the , and e is the . The \zeta in strongly depends on , as surface charge arises from / of surface groups (e.g., on silica), shifting the and thus \zeta, while affects screening and \zeta magnitude through specific adsorption. For instance, in typical aqueous s like NaCl, \zeta decreases in magnitude with increasing due to enhanced screening, but adjustments can tune \zeta from positive to negative values across the physiological range. For (AC) fields, particularly in aqueous microsystems, induced-charge electroosmosis (ICEO) emerges as a nonlinear effect where applied fields induce zeta potentials on ideally polarizable surfaces, generating time-averaged tangential flows. In this regime, the induced \zeta scales with the applied voltage, leading to quadratic dependence of the slip velocity on field strength; for microelectrodes, the characteristic ICEO flow speed is on the order of (\varepsilon / \mu) (V_{rms} / d)^2 , where V_{rms} is the root-mean-square voltage and d is the electrode dimension. This mechanism drives vortical flows around electrodes in low-frequency AC fields (kHz range), distinct from linear DC electroosmosis, and is prominent in aqueous electrolytes with Debye lengths comparable to electrode spacing. Zeta potential in aqueous electrokinetic systems is commonly measured via streaming potential techniques, where pressure-driven flow through a charged or generates an difference proportional to the , from which \zeta is derived using \Delta V / \Delta P = -(\varepsilon \zeta / \mu \kappa) , with \kappa the . This method, rooted in the reciprocity of , provides accurate \zeta values in aqueous setups by minimizing polarization effects compared to direct measurements.

Electrohydrodynamic Instabilities

Electrokinetic Instabilities

Electrokinetic instabilities refer to the onset of chaotic or turbulent flows in electrokinetic systems, arising from the coupling between , transport, and fluid motion, particularly in confined geometries such as electroosmotic pumps where conductivity gradients are present. These instabilities disrupt the otherwise laminar electroosmotic flow, leading to enhanced mixing but also potential performance degradation in microfluidic devices. The primary mechanism involves charge separation at conductivity interfaces, where applied amplify small perturbations in concentration, generating electric body forces that drive motion and further distort the conductivity profile. This feedback loop is quantified by the critical electric Rayleigh number, Ra_E = \frac{\varepsilon V \zeta}{\mu D}, where \varepsilon is the , V the applied voltage, \zeta the , \mu the , and D the diffusivity; instability thresholds occur when Ra_E exceeds a critical value, typically around 10-20 depending on . In linear stability analyses, the governing equations—linearized Navier-Stokes coupled with Poisson-Nernst-Planck—reveal that perturbations grow with a rate \sigma \sim \left( \frac{\varepsilon E^2}{\mu} \right) k^2, where E is the strength and k the perturbation , indicating faster growth for longer wavelengths in the initial stages. Key types include overlimiting current instabilities near ion-exchange membranes, where ion depletion creates extended space-charge regions that trigger electroconvective vortices, allowing ion fluxes beyond classical diffusion limits. Another prominent type is electroconvective rolls in dilute electrolytes, where non-equilibrium electro-osmotic slip at charge-selective surfaces destabilizes the diffusion layer, forming pairs of counter-rotating rolls that enhance mass transport. These instabilities relate briefly to bulk electrokinesis by extending similar charge-flow couplings to non-uniform ion distributions. In low-conductivity fluids like deionized water, these instabilities are particularly enhanced due to prolonged charge relaxation times, \tau = \varepsilon / \sigma, where \sigma is the electrical conductivity; longer \tau permits greater charge buildup before dissipation, lowering the critical field for onset compared to higher-conductivity electrolytes.

Tearing and Interfacial Instabilities

In electrohydrodynamics, interfacial instabilities arise when electric fields induce deformations at the boundary between immiscible fluids, often leading to breakup or jet formation. A prominent example is the Taylor cone instability, where electrostatic stresses compete with surface tension to deform a fluid meniscus into a conical shape. This cone, characterized by a semi-vertical angle of approximately 49.3°, forms when the electric field balances the capillary pressure, as derived from the equilibrium condition at the interface. Beyond a critical voltage V_c \sim \sqrt{\gamma d / \varepsilon}, where \gamma is the surface tension, d is the characteristic radius of curvature, and \varepsilon is the permittivity, the cone becomes unstable, resulting in the ejection of a thin liquid jet from the apex due to unbalanced normal stresses. This instability is fundamental to processes involving charged liquid emission and highlights the role of electric forces in overcoming stabilizing capillary effects. In systems modeled as leaky dielectrics, where finite allows free charge accumulation at the , tangential electric stresses further drive interfacial tearing instabilities. These stresses arise from imbalances in charge and conduction across the , inducing flows that wrinkle the surface and promote short-wavelength perturbations. The governing this instability is obtained by solving the normal stress balance equation, incorporating viscous, , and electric contributions, which predicts growth rates that increase with strength and conductivity mismatch. For sufficiently strong fields, these wrinkles evolve into tears, rupturing the and facilitating fluid mixing or droplet formation, distinct from bulk electrokinetic effects that may contribute minimally at charged boundaries. The classical Rayleigh-Plateau instability, which causes axisymmetric breakup of uncharged liquid jets due to , is significantly altered by in electrohydrodynamic regimes. Electric stresses modify the growth rate of perturbations, accelerating jet breakup for high fields, reducing the critical for instability and promoting finer droplet sizes, while low fields may slightly stabilize longer modes. Another key interfacial instability is Quincke rotation, where a spherical droplet or particle in a slightly conducting fluid undergoes spontaneous rotation above a critical strength. This occurs when the charge relaxation time exceeds the viscous hydrodynamic time, leading to tangential electro-osmotic flows that destabilize the no-rotation state. The critical field for onset is given by E_c \sim \sqrt{\frac{\gamma}{\varepsilon R}}, where R is the droplet , and has applications in micromixing and . The nature of these instabilities varies between direct current (DC) and alternating current (AC) fields. Under DC fields, persistent charge accumulation amplifies tangential and normal stresses, driving pronounced deformations and rapid growth of tearing or Plateau modes. In AC fields, however, oscillatory polarization induces dielectrophoretic forces that can suppress instability growth, particularly at frequencies matching charge relaxation times, thereby stabilizing interfaces against wrinkling or jet ejection. Experimental observations in electrospraying setups confirm these dynamics, particularly the cone-jet transition. As voltage increases from the dripping regime, the meniscus deforms into a , beyond which a steady emerges and undergoes varicose breakup via the electrically modified Rayleigh-Plateau mechanism, producing monodisperse charged droplets. High-speed imaging reveals the cone's stability window, with instabilities manifesting as pulsations or multiple jet branches when parameters deviate from optimal flow rates and conductivities.

Applications

EHD Pumping and Propulsion

Electrohydrodynamic (EHD) pumping and leverage the ion-drag mechanism, where asymmetric electrode configurations inject charges into a liquid, creating a net fluid flow through Coulombic forces on the ions that drag surrounding neutral molecules via . In ion-drag pumps, a high-voltage (typically 5–15 kV) accelerates injected ions, generating bulk motion in low-conductivity fluids without mechanical components. Seminal work by and coworkers established the foundational models for charge injection and flow induction in such systems. Velocities up to approximately 1 cm/s can be achieved at 10 kV in liquids, enabling compact actuation for various devices. Historically, EHD thrusters emerged in the as potential concepts, with early ionocraft designs explored by for ionic wind generation in or low-pressure environments, though limited by power constraints. Common configurations include two-dimensional wire-cylinder setups for , where a thin emitting wire and cylindrical collector produce directional thrust via asymmetric , as seen in lifter-style thrusters. For microfluidic applications, three-dimensional arrays—such as needle-to-mesh or interdigitated structures—enable precise flow control in channels, supporting multi-stage pumping for enhanced pressure gradients. Performance is quantified by hydraulic efficiency \eta = \frac{ Q \Delta P}{I V}, where Q is , \Delta P is pressure rise, I is current, and V is applied voltage; typical values remain below 1% due to ohmic losses and charge recombination, though the systems offer advantages in silence and compactness over mechanical pumps. These designs have seen revival in the for , building on 2013 MIT experiments (published 2012) that demonstrated ionic thrusters yielding up to 110 N/kW thrust in air, suitable for lightweight, low-noise unmanned aerial vehicles. Low-conductivity dielectric fluids, such as , are preferred for EHD pumping to minimize and maximize charge injection efficiency, with conductivities typically between $10^{-11} and $10^{-7} S/m. In aqueous environments, however, electrode corrosion poses significant challenges, as corona discharges accelerate material degradation in metals like or aluminum through electrochemical reactions and bombardment. Mitigation strategies include corrosion-resistant electrodes, such as or coated alloys, to sustain long-term operation. Recent advances incorporate hybrid EHD-magnetohydrodynamic (MHD) systems, combining injection with Lorentz forces in conductive fluids to boost rates and pressures in micropumps for biomedical or cooling applications. Numerical optimization using finite element methods has further refined geometries and field distributions, enabling simulations of charge transport and to maximize efficiency in complex 3D arrays.

Electrospraying and Electrospinning

Electrospraying is an electrohydrodynamic process where a high-voltage is applied to a emerging from a , forming a that emits a steady , which subsequently breaks into monodisperse droplets in the cone-jet mode. This mode ensures uniform droplet sizes, typically on the order of nanometers to micrometers, making it suitable for applications requiring precise particle . The flow rate Q in this regime scales approximately as Q \sim \left( \pi d^3 \epsilon \gamma / \rho \right)^{1/2} \left( I / (2\pi \sigma) \right), where d is the , \epsilon is the , \gamma is the , \rho is the , I is the , and \sigma is the electrical . Different operational modes exist, including (, large droplets), stable cone-jet (optimal for monodispersity), and multi-jet (high flow rates, leading to broader size distributions). The stability window for the cone-jet mode is characterized by the We = \rho v^2 d / \gamma \approx 1, balancing inertial and surface tension forces to prevent premature . Interfacial instabilities, such as Rayleigh-Plateau perturbations, enable controlled into droplets in this . Electrospinning extends electrospraying principles to produce nanofibers by incorporating high-molecular-weight into the solution, where electric forces stretch and entangle polymer chains during jet ejection, solidifying into fibers upon solvent evaporation. A critical voltage of approximately 5-10 kV is typically required to initiate the and jet formation, depending on solution and nozzle-to-collector distance. Resulting fiber diameters range from 10 to 1000 , influenced by polymer concentration and processing parameters, enabling the creation of non-woven mats with high surface area-to-volume ratios. These techniques find applications in systems, where electrosprayed particles encapsulate therapeutics for controlled release, and electrospun scaffolds support by mimicking structures. Water-based variants enhance , reducing the need for toxic organic solvents and improving suitability for biomedical uses like and . In the 2020s, advances include multi-nozzle arrays for scalable production, achieving higher throughput while maintaining fiber uniformity through synchronized electric fields. Integration with has enabled hybrid fabrication of complex scaffolds, combining macroscale printed structures with nanoscale electrospun fibers for enhanced mechanical and biological performance in regeneration.

References

  1. [1]
    Electro-Hydrodynamics of Emulsion Droplets: Physical Insights to ...
    Electric capillary number is defined as the ratio between the magnitude of electrical stresses ( ε 2 E 0 2 ) and capillary stresses ( γ / a ), where γ is the ...
  2. [2]
    Mechanisms and modeling of electrohydrodynamic phenomena - PMC
    In the remaining sections of this paper, we provide a brief account of the history of EHDs and related technology. Then, we review the known theories and ...
  3. [3]
    Alternating current electrohydrodynamics in microsystems: Pushing ...
    Taylor was the first to coin the term electrohydrodynamics as a generic label for this fluid flow behaviour arising from applied electric fields. In his ...
  4. [4]
    The Taylor-Melcher Leaky Dielectric Model - Annual Reviews
    Jan 1, 1997 · This review deals with the foundations of the leaky dielectric model and experimental tests designed to probe its usefulness.
  5. [5]
    Electrohydrodynamics
    May 25, 2012 · Abstract. The basic principles of electrohydrodynamics (EHD) are reviewed, including governing equations and boundary con-.
  6. [6]
    Coherent structures in electrokinetic instability with orthogonal ...
    Sep 26, 2017 · Electrokinetics is a branch of electrohydrodynamics (EHD) that deals with transport phenomena in the presence of electric double layers, usually ...<|control11|><|separator|>
  7. [7]
    CHAPTER 1: Electrical Spinning to Electrospinning: a Brief History
    Aug 16, 2018 · In about 1600, Gilbert1 first recorded the movement of liquid under the influence of a triboelectric field produced by rubbing amber. Due to ...
  8. [8]
    [PDF] Ionic wind produced by a millimeter-gap DC corona discharge ...
    Jan 27, 2021 · ... Niccolo Cabeo who noticed in 1629 the attraction, the contact and finally the repulsion of sawdust by electrified body, even if he did not ...
  9. [9]
    The History of Electrochemistry: From Volta to to Edison
    The story of electrochemistry begins with Alessandro Volta, who announced his invention of the voltaic pile, the first modern electrical battery, in 1800.
  10. [10]
    Historical Perspective on the Tools That Helped Shape Soil Chemistry
    Nov 1, 2011 · Today, electrophoresis is the most commonly used form of electrokinetics ... Herbert Freundlich also used this equation extensively in the 1920s ...
  11. [11]
    A Review of the Role of Interfacial Shear Stresses - Annual Reviews
    Electrohydrodynamics: A Review of the Role of Interfacial Shear Stresses. J R Melcher and G I Taylor; Vol. 1:111-146 (Volume publication date January 1969) ...
  12. [12]
    The Taylor-Melcher Leaky Dielectric Model - Annual Reviews
    ABSTRACT. Electrohydrodynamics deals with fluid motion induced by electric fields. In the mid 1960s GI Taylor introduced the leaky dielectric model to ...
  13. [13]
    Non-Linear Electrohydrodynamics in Microfluidic Devices - PMC
    In the late 1990s, droplets, pico-liter to micro-liter in size, have been identified as the microfluidic counterpart of the test tube. Droplet-based ...
  14. [14]
  15. [15]
    Studies in electrohydrodynamics. I. The circulation produced in a ...
    Taylor Geoffrey Ingram. 1966Studies in electrohydrodynamics. I. The circulation produced in a drop by an electric fieldProc. R. Soc. Lond. A291159–166http ...
  16. [16]
    Review Article—Dielectrophoresis: Status of the theory, technology ...
    The DEP force depends on the square of the applied field magnitude, indicating that DEP can be observed using either dc or ac field. Electrode geometry is an ...
  17. [17]
  18. [18]
    Two-dimensional numerical analysis of electroconvection in a ...
    Mar 8, 2012 · In this paper, we restrict the study to the strong injection case, which corresponds to values of the non-dimensional injection parameter C ...
  19. [19]
    Electrokinetics meets electrohydrodynamics | Journal of Fluid ...
    Sep 30, 2015 · The fields of electrokinetics (EK) and electrohydrodynamics (EHD) have developed separately, for different types of fluids and interfaces.Missing: seminal electrokinesis
  20. [20]
    The creation of electric wind due to the electrohydrodynamic force
    Jan 25, 2018 · We report direct evidence that electric wind is caused by an electrohydrodynamic force generated by the charged particle drag as a result of the momentum ...
  21. [21]
    Mechanism of charge injection-based electrohydrodynamic pump ...
    Mar 21, 2023 · In addition, the dimensionless parameters are the electric Rayleigh number T, injection strength C, mobility M, and charge-diffusion number α.
  22. [22]
    Electrohydrodynamic instabilities and electroconvection in the ...
    This paper reviews only the transient regime of unipolar injection into insulating isotropic liquids. Under the action of the Coulomb force, ions injected ...
  23. [23]
    Numerical investigation of injection-induced electro-convection in a ...
    Note that as a key dimensionless parameter that governs the EC flow, the injection strength C is proportional to the square of the characteristic length scale.
  24. [24]
    Disintegration of water drops in an electric field - Journals
    Disintegration of water drops in an electric field. Geoffrey Ingram Taylor ... Taylor Cone in a Pulsating Electrospray Directly Impact Mass Spectrometry Signals?, ...
  25. [25]
    Electrohydrodynamic instability of the interface between two fluids ...
    Aug 2, 2005 · The critical potential is found to be lower for leaky dielectrics than for perfect dielectrics. The weakly nonlinear analysis shows that the ...Missing: tearing | Show results with:tearing
  26. [26]
    Electrohydrodynamic instability of a charged liquid jet in the ...
    Apr 29, 2010 · The cases in the Rayleigh regime belong to the so-called Rayleigh-plateau instability, in which only axisymmetric disturbances can grow and the ...
  27. [27]
    Electrohydrodynamic instabilities at interfaces subjected to ...
    Jun 23, 2010 · Electrohydrodynamic instability, caused at fluid-fluid interfaces by the application of an external electric field, has attracted a lot of ...
  28. [28]
    Cone-jet regime in electrospray: A comprehensive review
    Aug 29, 2025 · Experimental observations indicate a sharp increase in drop charge density during the transition ... analysis for pulsating cone jet caused ...
  29. [29]
    A Review on Electrohydrodynamic (EHD) Pump - PMC - NIH
    This paper reviews the mechanism, structures, fluid types, and current progress of functional fluidic EHD pumps and EHD gas pumps, including soft pumps.
  30. [30]
    Electrohydrodynamic pumping in microsystems - IOP Science
    Abstract. The physical principles behind the electrohydrodynamic (EHD) actuation in microsystems is presented by reviewing five different EHD micropumps.Missing: E² | Show results with:E²<|control11|><|separator|>
  31. [31]
  32. [32]
    Simple in fabrication and high-performance electrohydrodynamic ...
    Dec 5, 2022 · The estimated flow rate and pressure for the proposed modulus EHD pump with dimensions of 1 cm3 can reach values of 82 ml/s and 620 kPa, ...
  33. [33]
    [PDF] A Model of an Ideal Electrohydrodynamic Thruster - DTIC
    Apr 8, 2010 · EHD based propulsion systems were proposed and investigated in the early 1960s. The most prominent effort, called the Ionocraft and led by ...
  34. [34]
    Electrohydrodynamic (EHD) Thrusters - RMCybernetics
    Ionocrafts was the name given to the first kind of vertical takeoff EHD thrusters designed during the early 60's, and form part of the EHD thrusters family.
  35. [35]
  36. [36]
    Electrohydrodynamic effect offers promise for efficient propulsion in air
    Jun 28, 2024 · Researchers at MIT have run their own experiments and found that ionic thrusters may be a far more efficient source of propulsion than conventional jet engines.
  37. [37]
  38. [38]
    Pumping of dielectric liquids using non-uniform-field induced ...
    Nov 30, 2011 · Further, NUF-EHD flow did not produce any observable electrode corrosion or bubble generation. In conclusion, we proposed and demonstrated a ...
  39. [39]
    Current and droplet size in the electrospraying of liquids. Scaling laws
    Scaling laws of the spray current as well as the charge and size of the droplets have been obtained from a theoretical model of the charge transport.
  40. [40]
    Revision of capillary cone-jet physics: Electrospray and flow focusing
    Jun 15, 2009 · We = 1 corresponds to the stability limit Q ≃ Q σ , while for We ≳ 20 there is a significant influence of the Weber number on the axisymmetric ...
  41. [41]
    A review on fabrication of nanofibers via electrospinning and their ...
    In this review paper we discussed the theory and the experimental setup of electrospinning along with the history of this process. We also ...
  42. [42]
    Electrospun Polymer Nanofibers: Processing, Properties, and ...
    It is observed that critical excitation voltage was significantly reduced from 45 KV to 20 KV. ... fiber diameters increased 350 nm, the elastic modulus decreases ...
  43. [43]
    Electrohydrodynamics: A facile technique to fabricate drug delivery ...
    This review discusses the state-of-the-art of using electrohydrodynamic technique to generate nano-fiber/particles as drug delivery devices. Keywords: ...Missing: paper | Show results with:paper
  44. [44]
    Sustainable strategies for waterborne electrospinning of ...
    Biocompatibility is an indispensable property for clinical translation of nanofibers into materials for e.g. drug delivery, tissue engineering or wound healing ...
  45. [45]
    Advanced multi-nozzle electrohydrodynamic printing - IOP Science
    Nov 13, 2024 · This paper reviews the recent development of multi-nozzle EHD printing technology, analyses jet motion with multi-nozzle, explains the origins ...
  46. [46]
    Current Trends and Future Prospects of Integrating Electrospinning ...
    Feb 2, 2025 · This article presents a review of the recent findings on the combination of electrospun nanofibers and three-dimensional (3D)-printed structures for ...