Electrodynamic droplet deformation

Last updated
Schematic of Periodic Droplet Deformation Between Prolate and Oblate Shapes Due to the Presence of an Oscillating Electric Field UPDATED ELECTRODYNAMIC DROPLET DEFORMATION.png
Schematic of Periodic Droplet Deformation Between Prolate and Oblate Shapes Due to the Presence of an Oscillating Electric Field

Electrohydrodynamic droplet deformation is a phenomenon that occurs when liquid droplets suspended in a second immiscible liquid are exposed to an oscillating electric field. Under these conditions, the droplet will periodically deform between prolate and oblate ellipsoidal shapes. The characteristic frequency and magnitude of the deformation is determined by a balance of electrodynamic, hydrodynamic, and capillary stresses acting on the droplet interface. This phenomenon has been studied extensively both mathematically and experimentally because of the complex fluid dynamics that occur. Characterization and modulation of electrodynamic droplet deformation is of particular interest for engineering applications because of the growing need to improve the performance of complex industrial processes(e.g. two-phase cooling, [1] crude oil demulsification). The primary advantage of using oscillatory droplet deformation to improve these engineering processes is that the phenomenon does not require sophisticated machinery or the introduction of heat sources. This effectively means that improving performance via oscillatory droplet deformation is simple and in no way diminishes the effectiveness of the existing engineering system.

Contents

Motivation

The heat transfer dynamics in two-phase two component flow systems are governed by the dynamic behavior of droplets/bubbles that are injected into the circulating coolant stream. [2] [3] The injected bubbles/droplets are typically of a lower density than the coolant and thus experience an upward buoyancy force. They enhance the thermal performance of cooling systems because as they float upwards in heated pipes the coolant is forced to flow around the bubbles/droplets. The secondary flow around the droplets modifies the coolant flow creating a quasi-mixing effect in the bulk fluid that increases the heat transfer from the pipe walls to the coolant. Current two-component, two-phase cooling systems such as nuclear reactors, control the cooling rate by optimizing solely the coolant type, flow rate, and bubble/droplet injection rate. This approach modifies only bulk flow settings and does not provide engineers the option of control of directly modulating the mechanisms that govern the heat transfer dynamics. Inducing oscillations in the bubbles/droplets is a promising approach to improving convective cooling because creates secondary and tertiary flow patterns that could improve heat transfer without introducing significant heat to the system.

Electrodynamic droplet deformation also of particular interest in crude oil processing as a method to improve the separation rate of water and salts from the bulk. In its unprocessed form, crude oil cannot be used directly in industrial processes because the presence of salts can corrode heat exchangers and distillation equipment. To avoid fouling due to these impurities it is necessary to first remove the salt, which is concentrated in suspended water droplets. Exposing batches of crude oil to both DC and AC high-voltage electric fields induces droplet deformation that ultimately causes the water droplets to coalesce into larger droplets. Droplet coalescence improves the separation rate of water from crude oil because the upward velocity of a sphere is proportional to the square of the sphere’s radius. This can be easily shown by considering gravitational force, buoyancy, and Stokes flow drag. It has been reported that increasing both the amplitude and frequency of the applied electric fields can significantly increases water separation up to 90%. [4]

Taylor’s 1966 Solution

Schematic of Velocity Field Corresponding to Taylor's Analytical Solution for Flow Inside and Outside a Droplet. Taylor 1966 Circulation Velocity Field.png
Schematic of Velocity Field Corresponding to Taylor's Analytical Solution for Flow Inside and Outside a Droplet.

Taylor’s 1966 solution [5] to internal and external flow of a sphere induced by an electric field was the first to provide an argument that accounted for pressure induced by fluid flow both inside the droplet and in the external fluid field. Unlike some of his contemporaries, Taylor argued that surface tension and a uniform internal pressure could not balance the spatially varying normal stress on a droplet interface that was resulted from the presence of a steady, uniform electric field. He posited that in order for a droplet interface to remain in a non-deformed state in the presence of an electric field, there must be fluid flow both inside and outside the droplet interface. He developed a solution for the internal and external flow field using a streamfunction approach similar to that of creeping flow past a sphere. [6] Taylor confirmed the validity of his solution by comparing it to images from flow visualization studies that observed circulation both inside and outside the droplet interface.

Torza's solution

Torza's 1971 solution [7] for droplet deformation under the presence of a uniform, time-varying electric field is the most widely reference model for predicting small amplitude droplet deformations. Similar to the solution developed by Taylor, Torza developed a solution for electrodynamic droplet deformation by considering fluid circulation both inside and outside of a droplet interface. His solution is innovative because it derives an expression for the instantaneous droplet deformation ratio by considering separate sub-problems to derive the effects of electric stress, internal hydrodynamic stress, external hydrodynamic stress, and the surface tension on the droplet interface. The droplet deformation ratio D is a quantity that expresses the relative extension and shortening of the vertical and horizontal dimensions of a sphere.

The electric stress sub-problem is formulated by defining electric potential fields on the inside and outside of the droplet interface that are expressible as complex phasors with the oscillation frequency as the imposed electric field.

Since Torza treats the fluid inside the droplet and outside the droplet as having no net charge, the governing equation for the electric stress sub-problem reduces to Gauss's law with a spatial charge density of zero. By re-expressing the electric field in terms of the gradient of the electric potential, the governing electric equation reduces to Laplace's equation. Separation of variables can be used to derive a solution to this equation of the form of a power series multiplied by the cosine of the polar angle taken relative to the direction of the electric field. Using the solutions for the magnitude of the electric potentials on the inside and outside of the droplet, the electric stress created on the bubble/droplet interface can be determined using the definition of the Maxwell stress tensor and neglecting the electric field.

It is worth noting that because the electric field is in the form of a phasor, the scalar product and tensor product of electric field with itself, as are present in the Maxwell stress tensor, result in a doubling of the oscillation frequency. The sub-problem Torza solves to determine the velocity fields and hydrodynamic stresses that result from the electric stress is of exactly the same form as the one Taylor used for his solution for steady electric fields. Specifically, Torza solves the streamfunction formulation of the curl of the Navier–Stokes equations in spherical coordinates by adopting Taylor’s streamfunction solution form and imposing stress balance conditions at the interface. Using the streamfunction solution, Torza derived analytical expressions for the velocity fields that could be used to derive analytical expressions for hydrodynamic stress on the interface for incompressible Newtonian fluids.

To incorporate the effect of surface tension into the periodic deformation of a droplet, Torza calculated the difference in electric and hydrodynamic stresses across the interface and used that as the driving stress in the Laplace Pressure Equation. This is the most important relation for this system because it describes the mechanism by which differences in stress across the droplet interface can induce deformation by inducing a change of the principle radii of curvature.

Using this relation between surface pressures in conjunction with geometrical arguments derived by Taylor for small deformations, Torza was able to derive an analytical expression for the deformation ratio as the sum of a steady component and an oscillating component with a frequency that is twice that of the imposed electric field as shown.

The important terms to recognize in this expression are in the steady term, the cosine in the time-varying term, and gamma in both terms. The phi term is what Taylor and Torza refer to as a “discriminating function” because its value determines whether the droplet will tend to spend more time in either a prolate or oblate shape. It is a function of all the material properties and the frequency of oscillation, but is completely independent of time. The time varying cosine term shows that the droplet does in fact oscillate at twice the frequency of the imposed electric field but is also generally out of phase due to the constant alpha term that arises due to the mathematics. The other variables are constants that depend on the geometric, electric, and thermodynamic properties of the relevant liquids in addition to the oscillation frequency.

In general, it is apparent that the magnitude of the droplet deformation is constrained by the interfacial tension, represented by gamma. As the interfacial tension increases, the net magnitude decreases due to an increase in capillary forces. Since the equilibrium shape of a droplet tends towards the one with the minimum energy, a large value of interfacial tension tends to drive the droplet shape towards a sphere.

Safety and practical considerations

Although periodic droplet deformation is widely studied for its practical industrial applications, its implementation poses significant safety issues and physical limitations due to the use of electric field. In order to induce periodic droplet deformation using an electric field, an extremely large amplitude electric field must be applied. Research studies using water droplets suspended in silicone oil required root-mean-square values as high as 10^6 V/m . Even for a small electrode spacing, this type of field requires electric potentials greater than 500V, which is roughly three times wall voltage in the United States. Practically speaking, this large of an electric field can only be achieved if the electrode spacing is very small (~ O(0.1 mm)) or if a high-voltage amplifier is available. It is for this reason that the majority of studies of this phenomenon are currently being conducted in research laboratories using small diameter tubes; tubes of this size are in fact present in industrial cooling systems, such as nuclear reactors.

Related Research Articles

<span class="mw-page-title-main">Laplace's equation</span> Second-order partial differential equation

In mathematics and physics, Laplace's equation is a second-order partial differential equation named after Pierre-Simon Laplace, who first studied its properties. This is often written as

<span class="mw-page-title-main">Navier–Stokes equations</span> Equations describing the motion of viscous fluid substances

The Navier–Stokes equations are partial differential equations which describe the motion of viscous fluid substances, named after French engineer and physicist Claude-Louis Navier and Irish physicist and mathematician George Gabriel Stokes. They were developed over several decades of progressively building the theories, from 1822 (Navier) to 1842–1850 (Stokes).

<span class="mw-page-title-main">Surface tension</span> Tendency of a liquid surface to shrink to reduce surface area

Surface tension is the tendency of liquid surfaces at rest to shrink into the minimum surface area possible. Surface tension is what allows objects with a higher density than water such as razor blades and insects to float on a water surface without becoming even partly submerged.

In mathematics and science, a nonlinear system is a system in which the change of the output is not proportional to the change of the input. Nonlinear problems are of interest to engineers, biologists, physicists, mathematicians, and many other scientists since most systems are inherently nonlinear in nature. Nonlinear dynamical systems, describing changes in variables over time, may appear chaotic, unpredictable, or counterintuitive, contrasting with much simpler linear systems.

In fluid dynamics, Stokes' law is an empirical law for the frictional force – also called drag force – exerted on spherical objects with very small Reynolds numbers in a viscous fluid. It was derived by George Gabriel Stokes in 1851 by solving the Stokes flow limit for small Reynolds numbers of the Navier–Stokes equations.

In fluid dynamics, Couette flow is the flow of a viscous fluid in the space between two surfaces, one of which is moving tangentially relative to the other. The relative motion of the surfaces imposes a shear stress on the fluid and induces flow. Depending on the definition of the term, there may also be an applied pressure gradient in the flow direction.

<span class="mw-page-title-main">Stokes flow</span> Type of fluid flow

Stokes flow, also named creeping flow or creeping motion, is a type of fluid flow where advective inertial forces are small compared with viscous forces. The Reynolds number is low, i.e. . This is a typical situation in flows where the fluid velocities are very slow, the viscosities are very large, or the length-scales of the flow are very small. Creeping flow was first studied to understand lubrication. In nature, this type of flow occurs in the swimming of microorganisms and sperm. In technology, it occurs in paint, MEMS devices, and in the flow of viscous polymers generally.

In fluid mechanics and mathematics, a capillary surface is a surface that represents the interface between two different fluids. As a consequence of being a surface, a capillary surface has no thickness in slight contrast with most real fluid interfaces.

In fluid dynamics, the Oseen equations describe the flow of a viscous and incompressible fluid at small Reynolds numbers, as formulated by Carl Wilhelm Oseen in 1910. Oseen flow is an improved description of these flows, as compared to Stokes flow, with the (partial) inclusion of convective acceleration.

<span class="mw-page-title-main">Potential flow around a circular cylinder</span> Classical solution for inviscid, incompressible flow around a cyclinder

In mathematics, potential flow around a circular cylinder is a classical solution for the flow of an inviscid, incompressible fluid around a cylinder that is transverse to the flow. Far from the cylinder, the flow is unidirectional and uniform. The flow has no vorticity and thus the velocity field is irrotational and can be modeled as a potential flow. Unlike a real fluid, this solution indicates a net zero drag on the body, a result known as d'Alembert's paradox.

<span class="mw-page-title-main">Jeans equations</span> System of differential equations

The Jeans equations are a set of partial differential equations that describe the motion of a collection of stars in a gravitational field. The Jeans equations relate the second-order velocity moments to the density and potential of a stellar system for systems without collision. They are analogous to the Euler equations for fluid flow and may be derived from the collisionless Boltzmann equation. The Jeans equations can come in a variety of different forms, depending on the structure of what is being modelled. Most utilization of these equations has been found in simulations with large number of gravitationally bound objects.

In the science of fluid flow, Stokes' paradox is the phenomenon that there can be no creeping flow of a fluid around a disk in two dimensions; or, equivalently, the fact there is no non-trivial steady-state solution for the Stokes equations around an infinitely long cylinder. This is opposed to the 3-dimensional case, where Stokes' method provides a solution to the problem of flow around a sphere.

<span class="mw-page-title-main">Elasto-capillarity</span> Physical phenomenon

Elasto-capillarity is the ability of capillary force to deform an elastic material. From the viewpoint of mechanics, elastocapillarity phenomena essentially involve competition between the elastic strain energy in the bulk and the energy on the surfaces/interfaces. In the modeling of these phenomena, some challenging issues are, among others, the exact characterization of energies at the micro scale, the solution of strongly nonlinear problems of structures with large deformation and moving boundary conditions, and instability of either solid structures or droplets/films.The capillary forces are generally negligible in the analysis of macroscopic structures but often play a significant role in many phenomena at small scales.

<span class="mw-page-title-main">Squirmer</span> Model in fluid dynamics

The squirmer is a model for a spherical microswimmer swimming in Stokes flow. The squirmer model was introduced by James Lighthill in 1952 and refined and used to model Paramecium by John Blake in 1971. Blake used the squirmer model to describe the flow generated by a carpet of beating short filaments called cilia on the surface of Paramecium. Today, the squirmer is a standard model for the study of self-propelled particles, such as Janus particles, in Stokes flow.

<span class="mw-page-title-main">Stokes problem</span>

In fluid dynamics, Stokes problem also known as Stokes second problem or sometimes referred to as Stokes boundary layer or Oscillating boundary layer is a problem of determining the flow created by an oscillating solid surface, named after Sir George Stokes. This is considered one of the simplest unsteady problems that has an exact solution for the Navier-Stokes equations. In turbulent flow, this is still named a Stokes boundary layer, but now one has to rely on experiments, numerical simulations or approximate methods in order to obtain useful information on the flow.

In fluid dynamics, Taylor scraping flow is a type of two-dimensional corner flow occurring when one of the wall is sliding over the other with constant velocity, named after G. I. Taylor.

In the larger context of Navier-Stokes Equations but especially in the context of potential theory Elementary flows are a collection of basic flows from which it is possible to construct more complex flows with different techniques. In this article the term flows is used interchangeably to the term solutions due to historical reasons.

In fluid dynamics, Hicks equation, sometimes also referred as Bragg–Hawthorne equation or Squire–Long equation, is a partial differential equation that describes the distribution of stream function for axisymmetric inviscid fluid, named after William Mitchinson Hicks, who derived it first in 1898. The equation was also re-derived by Stephen Bragg and William Hawthorne in 1950 and by Robert R. Long in 1953 and by Herbert Squire in 1956. The Hicks equation without swirl was first introduced by George Gabriel Stokes in 1842. The Grad–Shafranov equation appearing in plasma physics also takes the same form as the Hicks equation.

Schneider flow describes the axisymmetric outer flow induced by a laminar or turbulent jet having a large jet Reynolds number or by a laminar plume with a large Grashof number, in the case where the fluid domain is bounded by a wall. When the jet Reynolds number or the plume Grashof number is large, the full flow field constitutes two regions of different extent: a thin boundary-layer flow that may identified as the jet or as the plume and a slowly moving fluid in the large outer region encompassing the jet or the plume. The Schneider flow describing the latter motion is an exact solution of the Navier-Stokes equations, discovered by Wilhelm Schneider in 1981. The solution was discovered also by A. A. Golubinskii and V. V. Sychev in 1979, however, was never applied to flows entrained by jets. The solution is an extension of Taylor's potential flow solution to arbitrary Reynolds number.

The fracture of soft materials involves large deformations and crack blunting before propagation of the crack can occur. Consequently, the stress field close to the crack tip is significantly different from the traditional formulation encountered in the Linear elastic fracture mechanics. Therefore, fracture analysis for these applications requires a special attention. The Linear Elastic Fracture Mechanics (LEFM) and K-field are based on the assumption of infinitesimal deformation, and as a result are not suitable to describe the fracture of soft materials. However, LEFM general approach can be applied to understand the basics of fracture on soft materials. The solution for the deformation and crack stress field in soft materials considers large deformation and is derived from the finite strain elastostatics framework and hyperelastic material models.

References

  1. Kaji NN, Mori YH, Tochitani YY. Electrically Induced Shape Oscillation of Drops as a Means of Direct-Contact Heat Transfer Enhancement: Part 2—Heat Transfer. J. Heat Transfer. 1988;110(3):700-704.
  2. S. Mostafa Ghiaasiaan. Two-Phase Flow, Boiling, and Condensation: In Conventional and Miniature Systems. 2008. Cambridge University Press
  3. Takaaki Mochizuki. Periodic Deformation of Microsize Droplets in a Microchannel Induced by a Transverse Alternating Electric Field. Langmuir 2013 29 (41)
  4. Byoung-Yun Kim, Jun Hyuk Moon, Tae-Hyun Sung, Seung-Man Yang, Jong-Duk Kim. Demulsification of water-in-crude oil emulsions by a continuous electrostatic dehydrator. Separation Science and Technology. Vol. 37, Iss. 6, 2002
  5. G. Taylor. (1966). Studies in Electrohydrodynamics. I. The Circulation Produced in a Drop by Electrical Field. Proceedings of the Royal Society A: Mathematical, Physical and Engineering Sciences.
  6. Pijush K. Kundu, Ira M. Cohen. Fluid Mechanics. 2010. Academic Press
  7. S. Torza, R. G. Cox, and S. G. Mason "Electrohydrodynamic Deformation and Burst of Liquid Drops" Phil. Trans. R. Soc. Lond. A February 18, 1971 269 1198 295-319