You are on page 1of 36

ANNUAL

REVIEWS Further
Quick links to online content

ELECTROHYDRODYNAMICS: A REVIEW OF THE


ROLE OF INTERFACIAL SHEAR STRESSES
By J. R. MELCHERt AND G. I. TAYLOR
Massachusetts Institute of Technology, Cambridge, Massachusetts
and Farmfield, Huntingdon Road, Cambridge, England

SCOPE

Electrohydrodynamics can be regarded as a branch of fluid mechanics


Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

concerned with electrical force effects. It can also be considered as that part
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

of electrodynamics which is involved with the influence of moving media on


electric fields. ActuaJIy, it is both of these areas combined, since many of the
most interesting problems in electrohydrodynamics involve both an effect
of the fluid motion On the fields and an influence of the fields on the motion.
The word "electrohydrodynamics" is relatively new; the area it repre­
sents is not. The related literature is as venerable as that for the subject of
electricity itself. Even more, to generate an engineering interest there is no
need to emphasize the great technological promise of the area, since applica­
tions already form the basis for major industries. But the center of attention
in almost any discussion is the lack of reproducibility in experiments and the
inadequacies of theoretical models. Electrostatic effects in fluids are known
for their vagaries; often they are 50 extremely dependent on electrical con­
duction that investigators are discouraged from carefully relating analytical
models and simple experiments. Yet the foundations of fluid mechanics are
formed from work that relates carefully designed experiments to analytical
models, and we wish to focus attention on electrohydrodynamic research
having this objective. An historical survey of the subject has been given by
Pickard (1) and is not deemed appropriate here.
ELECTRODYNAMICS
Laws and approximations.-A summary of the pertinent electrical laws

E. Dy­
serves further to define our subject. A salient feature of electrohydrodynamic
interactions is the irrotational nature of the electric field intensity,
namic currents are so small that the magnetic induction is ignorable, and the
appropriate laws are essentially those of electrostatics, as summarized in
Table 1.2 Gauss' law, Equation Ib, relates the free-charge density, q, to the
electric displacement D, while Equation Ie brings in the free-current density
in a dynamic equation that guarantees conservation of charge. As is conven-

lOne of the authors, U. R. M.) acknowledges the support of N.A.S.A. research


grant NGL-22-009-014 ff6, and would like to thank Tsen-Chun g Cheng for his as­
sistance in obtaining the data of Figure 3 and the photograph of Figure 2.
2 Equa tions in tables are referenced with the table number as the prefix; e.g.,
Equation Ib is Equation b in Table I.

111
1 12 MELCHER & TAYLOR

TABLE I
SUMMARY OF DIFFERENTIAL LAWS, TRANSFORMATIONS AND BOUNDARY
CONDITIONS FOR QUASI-STATIC ELECTRIC FIELD SYSTEM

Differential Laws Transformations Boundary Conditionsa

V X E=O Ia E' =E Ie n X [E] == 0 Ij


v·D == q Ib D'=D If n·[D] == Q Ik
aq
V·l+- == 0 Ie q' == q Ig n·!l]+ VI·K Ilb
at
aQ
Id i
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

D==EOE+P i' == - qv Ih == n,v!q]--


at
P' P Ii
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

==

a [A]=Aa-Ab==the jump of A across the interface; b VI=surface Nabla.

tional, the electric displacement is further defined in terms of the polariza­


tion density, P, (Equation Id) with Eo=8.8SXlO-12 in MKS units.
The quasi-static electrical laws of Table I are invariant to a Galilean
transformation (2), which can be used to show that the fields in a primed
frame moving with the velocity v are as given by Equations Ie to Ii, in
Table 1. The transformations reflect the quasi-static approximation implicit
to the differential laws. Thus, the electric field and current density do not
transform as they do in the magnetohydrodynamic approximation, in which
magnetic induction is essential but net charge is negligible. A boundary hav­
ing the unit normal n directed from region (b) to region (a), supporting the
surface-charge density Q and surface-current density K, and having the nor­
Ij to II, which are found by inte­
mal velocity n· v, is described by conditions
grating the differential laws over surfaces and volumes that include the
boundary. (2) The surface-current density K of condition II includes con­
tributions from the convection of surface charge, and, if appropriate, con­

Conduction and polarization.-The quasi-static equations of Table


tributions due to surface conduction.
I are
written in terms of the macroscopic fields with the effects of material motion

J=l* (g, E). Subject to the assumption


accounted for by constitutive laws. For many purposes, the conduction law
for the fluid at rest takes the form
that accelerations do not influence the conduction process, this law holds in

l* (g', E'), and,


the face of fluid motion if it is evaluated in a frame of reference moving with
the fluid velocity v. That is, with motion we must write J' =

in view of Equations Ig and Ih of Table I, the conduction law expressed in


the laboratory frame but with fluid motion becomes
I == i*(q, E) + qv 1.

Equations Ie and Ii show that if the polarization is a function of E, it is


the same whether viewed from the laboratory frame or in the moving frame
of the fluid. Of course, the assumption implicit in using the transformation
ELECTRO HYDRODYNAM I CS 113

laws to generalize constitutive laws to the case of material motion is that


accelerational effects on conduction and polarization can be ignored.
Charge relaxation . Even though electrical conduction in fluids is often
-

poorly characterized by Ohm's law, (3) it is evident from recently reported


research that this simplest of all conduction laws can be used to understand
a surprisingly wide range of electrohydrodynamic phenomena. In this re­
view attention is confined largely to this case where
1* = uE 2.

with the electrical conductivity u of a given fluid element constant. I n addi­


tion, we will take as the polarization constitutive law simply
D=� �
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

where the permittivity fi of a moving fluid particle is constant.


Access provided by 103.147.138.252 on 07/15/21. For personal use only.

In a homogeneous incompressible fluid, where u and fi are constants and


V· v 0, we can make far-reaching conclusions about the distribution of the
=

free-charge density, q. We combine Equations Ib and Ic with Equations 2


and 3 to obtain

[�at + v.v ] q +� q = 0 4.
E

The characteristic lines in (r, t) space are simply the particle lines, hence we
have
aT
q = qoe-t'Ton- = v 5.
dt

where the bulk relaxation time T =e/u. Thus, the free-charge density in the
neighborhood of a given fluid particle decays with the relaxation time T.
M oreover, unless a given element of fluid can be traced via a particle line to
a source of charge, it will support no bulk charge density.
H YDRODYN AMICS
Equations of Motion.-We confine ourselves to cases where the mass
density p of a given fluid element is constant; hence the fluid, having a con­
stant viscosity JJ. and subject to the gravitational acceleration g, has a pres­
sure p and velocity v governed by the equations of Table I I . In addition to
the mechanical pressure and viscous stress Tm, there is an electrical force due
to the free-charge density q (the charges that contribute to conduction and
convection currents) and due to polarization. The boundary conditions
(lId to I If) are found by integrating the conservation of momentum and
mass, Equations I Ia-I I c through the interface.
Electrical Forces.-The electrical force on an incompressible fluid can be
correctly written in alternative forms that differ by the gradient of a pres­
sure. This is true because in the differential laws and implied boundary con­
ditions of Table I I , the pressure, p, appears only in Eq uation l I b and i s
simply redefined by the addition o f an electrically induced pressure. H ence,
we ignore electrostriction forces, since they could be of importance only for
114 MELCHER & TAYLOR

TABLE II
HYDRODYNAMIC EQUATIONS AND BOUNDARY CONDITIONS

Differential Laws Boundary Conditions

pg + v. (Tm + To) n[p] n·[Tm + Tel


Dv
p -
De
= IIa = IId

T;jm
- OijP [v]
(Ov; OVi ) lIb n X
aXi + ax;
= 0 lIe
= p.

v·v = 0 lIe n·[v] = 0 IIf


Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

dilatational fluid motions, and write the force density in the form due to
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

Korteweg & Helmholtz in (4, 5)

F = qE - 1
-£lvE
2
6.

Equation 6 can be written identically as

F = V·Te; 7.

where the Maxwell stress tensor Te accounts for not only forces attributable
to free charges, but because E = E(r, t), those due to polarization as well.

OBJECTIVES

We are now in a position to set limits on the scope of this review. The
dynamics of fluid systems characterized by regions of uniform ohmic conduc­
tivity and permittivity will be highlighted. We have established with Equa­
tion 5 that in the absence of sources of charge that are communicated by

density, q. Moreover, because E is constant in a given region, it is clear from


material convection with the volume of interest, the bulk is free of the charge

Equation 6 that the fluid is not coupled to the electric field in the bulk.
Hence, with the restrictions outlined, we review classes of motion involving
electromechanical coupling at fluid interfaces.
Our observations should serve to illustrate that, if a fluid system includes
interfacial regions where electrical parameters suffer discontinuities, electro­
mechanical coupling at the interfaces is likely to dominate the resulting e\ec­
trohydrodynamics. Surface interactions are of greater significance in elec­
trohydrodynamics than might be expected from much of ordinary hydro­
dynamics. The literature of drops and jets in electric fields is highly de­
veloped, and relates largely to the dynamics of two-phase systems with inter­
faces stressed by electrical surface forces. Due to meteorological interest,
water and air are often considered, and these fluids exemplify cases in
ELECTRO HYDRODYNAMICS 1 15

which one fluid is much more highly conducting than the other. Then, if
the relaxation time in the more conducting fluid is short compared to dy­
namical times of interest, the interface can be regarded as perfectly conduct­
ing; it supports no tangential electric stress. Regardless of interfacial defor­
mation, surface forces always act perpendicularly to the surface in this im­
portant class of interaction.
At the opposite extreme, where the fluids in a two-phase system are

term of Equation 6, the polarization -forc e density, is operative at the inter­


considered as perfectly insulating with no free-charge density, the second

face. Again, as the force-density expression shows, the surface-force density


must act in the direction of -Ve; that is, perpendicular to an interface.
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

By contrast with these two limiting cases, the physical situations re­
viewed relate mainly to the electrohydrodynamics resulting from electrical
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

surface shear forces. Thus, our review is confined to a small corner of the
total area of electrohydrodynamics: ohmic fluids and surface interactions
dominated by interfacial electrical shear-force densities. We begin with
steady motions and conclude with stability problems.

STEADY CONVECTION: DC FIELDS


A SIMPLE EXAMPLE
Consider a case (6) which has the dual virtues of being easily demon­
strated in the laboratory and easily described mathematically, while ex­
emplifying the nature of electrohydrodynamic shear-stress interactions.
To induce an interfacial electrical shear force, the interface must simul­
taneously support a surface-charge density and a tangential electric-field
intensity. This is accomplished in a simple way with the experiment of
Figure 1, where a shallow, slightly conducting liquid, region (b), fills an in­
sulating container A to the depth b. Electrodes B and C, abutting the right
and left ends of the container, make electrical contact with the liquid to
complete an electrical circuit with the source of potential Yo. Thus, one re­
quirement for a shearing-force density at the interface D is provided by the

FIG. 1. Electrode C has the potential Vo relative to electrodes B and F. Surface


charges induced on the interface D act in concert with the field Ez, which drives the
conduction currents in the liquid to induce the counterclockwise cellular convection
shown in Figure 2.
116 MELCHER & TAYLOR

conduction current in the liquid, which insures that there is a tangential


electric field E" at the interface.
For the purpose of fulfilling the second requirement, free charge on the
liquid surface, a third electrode F extends over the interface and is canted
at an angle such that it contacts the interface and electrode at the left and
reaches a height a at the extreme right. The interface assumes a distribution
in electrical potential that varies from Vo at the right to zero at the left.
Because the slanted electrode has zero potential, with a spacing hex) that
varies essentially linearly with x, there is a surface charge induced on the
interface. Both the potential difference and the spacing vary linearly, and
the surface-charge density therefore tends to be uniform-at least if it is
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

not redistributed appreciably by the resulting fluid motions. Thus, there is


an electrical shear force on the interface which acts to the left and tends to
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

produce counter-clockwise cellular convection in the plane of the paper.


There are also normal stresses on the interface, but we assume these are
balanced by gravity and the hydrostatic pressure without a significant al­
teration in the interface geometry.
Quasi-one-dimensional model.-If the length 1 of the experiment is large
compared to both a and b, a simple model suffices to make a quantitative
prediction of the fluid convection. For now, we assume that the convection
of charge at the interface gives rise to a current in the x direction that is
negligible compared to conduction current in the bulk of the liquid. That is,

the requirement that n·uE=O on the upper and lower surfaces, conditions
for now the fields are determined as though the liquid were stationary, with

satisfied by the uniform field3


Vo .
Eb = --1- 1% 8.

This is also the tangential component of E j ust above the interface (Equa­
tion Ij). Hence, the distribution of potential on the interface is
Vo
t/>(y = b) = -
1 9.

and f/> is defined such that E= -Vf/>. It follows that in the region above the
where we take the upper electrode and left edge of the liquid as the reference

interface the field is approximately


Ea [</>(y b)/h(x) 1 iu 10.
We have arranged the experiment such that hex) =ax/l, so that Equations 9
= =

and 10 give
Ea = iuVo/a 11.
The interfacial shear-force density is n · [Te] i,,[Tzuej, Equation I Id.4 Thus,
=

from Equations 8 and 11, the interface is subject to an electrical shear-force


density

3 The components of i are the unit vectors in the coordinate directions.


• [Al=Aa-Ab=the jump of A across the interface.
ELECTROHYDRODYNAMICS 117
12.

That this expression is negative is consistent with Figure 1. A reversal of the


applied potential polarity has no effect on Tx because the signs of both the
surface charge and the tangential field are then reversed.
I n the limit where b«l, the flow in regions of the fl uid bulk removed sev­
eral lengths b from the ends can be approximated as being plane: v=vx(y)ix•
The profile is determined by the no-slip condition on the tank bottom vx(O)

= -EoVo2/la, and the condition that net flow in the x direction be zero. Thus,
=0, the viscous and electrical shear-stress balance at the interface, J.l.avx/ay

) - fo Vo'b [�(!-.-)' - (L)]


x(y 2}la 2
v
= 13.
b b
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

and the profile is as sketched in Figure 1. Of course, the velocity observed is


Access provided by 103.147.138.252 on 07/15/21. For personal use only.

somewhat less than that given by this expression, since the viscous losses
from the reversal in flow direction near the tank ends are not included in the
model.
Charge transport; electric Reynolds nu mber .-A lthough the i nfluence 01
the electrical stresses on the fl uid is included in the model, it ignores the
reciprocal effect of the motion on the fields. We have assumed that the
convection of charge at the interface in the �x direction is negligible com­
pared to the cond uction curren t through the bulk; i.e. , bJx»Qvx, or using
Jx=uVo/1 and Q=EoVo/a,
( )( )
I
ba ----;;-
fOVX
R, « 1; R, "" 14.

The electric Reynolds n u mber Re is defined by Stuetzer (7) as the ratio of a


charge-relaxation time to a time L/vx for the fl uid to move a characteristic
length L at the characteristic velocity vx• From Equation 14, in our example,
the length L=ab/l, a combination of lengths, because the component of E
that gives rise to electrical dissipation is not in the same region of space or in
the same direction as the componen t (above the interface) that represents
much of the energy storage.
In the section on steady, dc-field-induced cellular convection, we will
limit our discussions to cases where Re«1. However, the ac-field-i nduced
motions and cases of i nstability reviewed shortly will include finite-electric­
Reynolds-n u mber effects.
An experiment.-The cellular convection is readily observed in an appa­
ratus having the configuration of Figure 1 by i n trod ucing small particles
that are nearly neutrally buoyant. A streak photograph is shown in Figure
2 where the liquid is corn oil (relaxation time E/U of about 1 sec) and the
dimensions and voltage are as indicated in the figure caption.
In the photograph of Figure 2, note that the vertical poi n t of flow re­
versal is about 2/3 of the distance from the tank bottom to the in terface, as
suggested by the plane-flow model , Equation 13. The nearly symmetric
shape of the cell, together with the observation that small particles placed on
the interface at the right traverse most of the length with nearly constant
118 MELCHER & TAYLOR
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

FIG. 2. Streak photograph of the cellular convection induced in an apparatus hav­


ing the configuration sketched in Figure 1,1=24 cm, b=3.8 cm, Vc=20 kV and the
fluid is corn oil. A comparison of the experimentally determined surface velocity and
the prediction of Equation 13 is shown in Figure 3.

velocity, further indicates that the physical arrangement of electrodes and


interface is successful in producing a nearly u niform electric-shear-stress
distribution over the i n terface.
A plot of the product of observed surface velocity U near the center of
the apparatus and maxi mum spacing a, as a function of the applied voltage
Vo for three angles of inclination, is shown in Figure 3 where the solid line is
predicted by Equation 13 evaluated at y = b. For the maximum velocity Re
as defined by Equation 14 is on the order of unity, so that the assu mption
that the electrical-relaxation process easily keeps up with the motion is not

80
u
Q)
en
......
N 60
E
III

'0
40
,
0

20·

0
0 5 10
Vo- kV

FIG. 3. Product of maximum electrode spacing and the interfacial vel ocity U near
the center of the apparatus shown in Figures 1 a nd 2, as a function of the applied
voltage. Liquid is corn o il, E=3.1Eo, ".",,10-10 mhos/m, 1'=0.055 kg (ms)-l and 1=24
cm, b= 3.8 cm. The solid curve, from Equation 13 is Ua =EO Vo2b/4�.
ELECTROHYDRODYNAMICS 119
well taken for higher voltages. Deviations of the theory and experi ment over
the range of voltages shown can be attributed mainly to the viscous end
effects, which are ignored in the simple quasi-one-dimensional model.

PERIODIC CONVECTION
I t is possible to conceive many variations on the theme of dc-field-in­
duced shear flows. We will concentrate on two fu rther combinations of inter­
facial geometry and field nonuniformity that have well developed and rela­
tively si mple analytical descriptions. In this su bsection we fu rther indicate
how the application of a nonuniform field at an interface leads to bulk con­
vection, while in the next section the geometric configuration of the i n terface
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

leads to a distortion of an initially uniform applied field to secure cellular


convection.
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

The three steps required to analyze a low-electric-Reynolds-n u m ber flow


are ill ustrated in the previous section. First, the fields are computed and,
because Re «1, effects of the fluid are only geometric. Thus the electric inter­
facial stresses are computed independently of the fluid velocity. Second ,
a flow pattern consisten t with the distribution of stresses on the in terface is
found. This pattern is not coupled to the field in the bulk. Finally, the field­
induced stresses and flow are matched at the in terfaces to determine a self­
consistent relationship between the imposed potentials and the flow velocity.
Of cou rse, there are a limited n u mber of situations in which this last step can
be completed in closed form-our reason for discussing the following cases.
A configu ration of i mposed field and fluid studied by Smith & Melcher
(8,9) is shown in Fig u re 4a. A static, spatially periodi c distribution of poten­
tial is imposed on a planar electrode A, which also serves as the bottom for a
container filled to a depth b with a slightly conducting liquid (region b).
This layer of liquid is i n turn covered by a second fluid, i n region (a), which
can, generally, also be slightly conducting. In Figure 4 the u pper fluid is
assu med for purposes of discussion to be the less conducting. The mechanism
for creating cellular convection is basically the same as in the case of Figure
1. With the upper fluid less conducting than the one below (having, for
example, zero conductivity) adjacent positions of positive and negative
polarity on the segmented electrode can be thought of as being joined by a
resistance (the lower liquid) in series with a capacitance (the u pper liquid)
in series with a resistance (the lower liquid again). Thus, charges ind u ced on
the interface have the same sign as the neighboring charges on the seg­
mented electrode. These surface charges are subject to the electric-field in­
tensity, the tangential component of which produces electrical shear forces
on the interface sketched in Figure 4a. Note that the spatial periodicity of
these stresses is such that we expect two cells to form in the length t, with
points of zero interfacial velocity on the in terface having the same x coor­
dinates as both the peak potential and zero potential on the segmented elec­
trode. These physical considerations serve as a guide in guessi ng the appro­
priate flow pattern to match the electrical stresses. As in the case of Figure 1,
120 MELCHER & TAYLOR
y
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

(b)

FIG. 4. (a) A segmented electrode A is constrained by a static periodic potential


and is in electrical contact with a liquid having depth b and an interface at B. Surface
charges are induced with the polarity shown if the lower fluid is much more highly
conducting than the upper one. (b) The surface charges interact with the imposed
electric-field intensity E to produce the cellular streamlines.

it is again assu med that gravity holds the interface essentially flat, so that

Electric stresses.-Our introductory remarks make it clear that in the


normal stresses are not of interest.

bulk of the fl uids (a) and (b), the electric potential cf> satisfies Laplace's equa­
tion. In view of the potential constraint at the electrode, the potential dis­
tributions in each fl uid m ust have an x dependence of the form cos (7rxjl).

y dependence of the potential requires four boundary conditions: (a) that


We take motions as being independent of the z coordinate. To determine the

the potential at y= -b is as given in Figure 4a; (b) that the potential must
be continuous at the fl uid-fl uid interface (Eq. Ij) cf>a(y=O) =cf>b(y=O); (c)
that the convection of surface charge at the interface is ignored, and so the
normal component of the current must be continuous (Eq. II), uaacf>ajay(y
0)
= =Ubacf>bjay(y= 0); and finally, (d) that the upper fl uid is bounded from
above by a sufficiently distant b6undary that cPb->O as y-> 00.
Variable separable sol utions having the required x dependence, while
satisfying Laplace's equation and meeting these boundary conditions, are

15.
ELECTRO HYDRODYNAMICS 121

and
<ph = --;-
lEo [cosh (".y) - smh (".y)] cos ( ".x )
--
1
ITo
ITb . -
-1 -1- 16.

where

Eo = 7r�O [ cosh Ch) : sinh Cb) Jl


+

The electrical shear-force density follows from Equations 7 and Ij


T�' = [T...] = &o[.Eu] evaluated at
y = 0 17.
where the required components of E follow directly from Equations 15 and 16
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

Convection.-As a further restriction on the electrohydrodynamics, we


assume at the outset that the hydrodynamic Reynolds n u mber based o n t
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

and the peak surface velocity U is small compared to u nity, so that the fluid
momentum can be ignored. Then it is appropriate to define a stream func­
tion "" in the usual way as
. a", . a",
V = z,,--Iu-
ay ax 18.

Our approximation is the usual limit of creep flow, and"" satisfies the bi­
harmonic equation (10)
19.
so that v in turn satisfies the'equations of motion IIa-I I c with Dv/Dt==O.

of the x dependence for the stream function. From the sketches of Figure 4a,
The key to matching the shear-stress conditions is the j udicious choice

it is clear that we can expect the x component of the velocity to be an odd


function of x with the wavelength t. Thus, we choose"" to be of the variable,
separable form
'" = f(y)sin (ax/b); a = 2".b/l 20.
S ubstitution of Equation 20 i nto Equation 19 shows that solution in each
fl uid region is a linear combination of four solutions. Only two of the solu­
tions in the upper fluid remain finite as y� co, thus we are left with six
arbitrary coefficients, two for solutions in region (a) and four for solutions i n
region (b), with which t o satisfy the boundary conditions.
Owing to gravity, the surface is not deflected enough in the vertical direc­
tion to affect the field distributions. Thus, i n our boundary conditions, we
ignore the normal-stress balance at the interface, but instead stipulate that

ditions are: (a) and (b): that vl/b(y=O) =0, and vl/b(y=O) =0; (c) that the
the normal velocity not only be continuous, but vanish. The boundary con­

mechanical shear stress balance the surface-force density given by Equations


1S-17, [T:l1r] + T,,' =0 at y=O (Eq. lId); (d) that the tangential component
of velocity be continuous at the i nterface, [v,,](y=O) =0 (Eq. IIe); and, fin­

electrode vanish : v"b (y = -b) =0, and v b (y -b) =0.


ally, (e) and (f) that the normal and tangential velocities at the segmented
l/ ==
sin (ax/b), the stream functions i n the respective fluids are
I f we call the tangential velocity i n the x- direction at the interface U
122 MELCHER & TAYLOR
if/> = Uye-ou/b sin (ax/b) 21.

and

>/l' = U {y sin h a sinh [a(1 +y/b)] - a(y + b) sinh (ay/b)} s_�_ 22.

The shear-stress boundary condition determines U, which is written in terms


of R=Ub/Uo, S=Ea/Eb and M=J.lo/J.lb as

23.

where

'Y(a) a2). [2M(Sinh2 a a2) + sinh 2a


Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

=
7r (sinh2 a - { - - 2a]
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

• [ cosh � +R-1 sinh � J} -1


We are successful in satisfying all of the boundary conditions only because
the x distribution of stress can be matched exactly to the viscous-stress dis­
tribution. The flow pattern represented by Equations 18, 21, and 22 is shown
in Figure 4b.
Note that the velocity U, as given by Equation 23, has the same param­
eter dependence as that found for the canted-electrode experi ment of Figure
1. This is particularly evident if in both cases the uppe� fluid is a noncon­
ducting gas, so that uo-+O, J.la-+O and Ea-+EO. Then Equation 23 becomes:

U = 'Y(a) (�:) Vo2 24.

with 'YCa) only a function of geometry. The surface velocity given by Equa­
tion 13 for the canted-plate experiment is identical to this expression if
'YCa)-+b/a. In the general case, the sign of RS-l discriminates the sense of

Experiment.-An experiment with essentially the configuration of Figure


the cellular convection, since 'Y >0.

4 gives the streak photograph of Figure 5; the sense of the rotation is as


sketched in Figure 4b, consistent with RS> 1. Most of the particles Cair
bubbles) entrained to trace the streamlines are in the upper liquid, although
careful examination shows the expected cells in the lower liquid as well.
Data for the experiment are given with Figure which shows quantita­ 6,
as a function of the applied voltage Vo. I n this experiment, R. based on 1 in
tive comparison (9) of the measured and predicted peak surface velocity, U,

each of the liquids is on the order of 0.01 or less, while the hydrodynamic
Reynolds number is always less than 0.6.
CONVECTION IN DROPS AND AROUND BUBBLES
As shown by Taylor (11) electric shear-induced convection can occur
quite naturally in electrified drops and bubbles, for physical reasons closely
related to those responsible for the convection in the configurations of
Figures 1 and 4. For purposes of developing an analytical model, a spherical
drop of vapor void of radius b is shown in Figure 7a, where the appropriate
ELECTROHYDRODYNAMICS 1 23
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

FIG. 5. Streak photograp h of periodic cellular convection observed in apparatus


having essentially the fluid and field configuration of Figure 4a. The cells have the
spatial period of Figure 4b and velocity-voltage relationship of Figure 6 (after Smith
& Melcher).

spherical coordinates are defined . Electrodes, removed many radii b from the
spherical region (b), make electrical contact with the surrounding fl uid (a).

field intensity to be uniform and of magnitude Eo as r cos 8-+ 00 , play a role


Thus, the outer fluid, together with the electrodes that constrain the electric­

analogous to that of the electrodes and liquids (b) in Figures 1 and 4; by


virtue of the conduction current they i nsure that there is a tangential com­
ponent of E in the neighborhood of the spherical interface.
Once again, convection results because not only is there a tangential
component of E, but there is also a surface charge ind uced on the interface.
The case illustrated in Figure 7a pertains to an insulating fl uid (b)-perhaps
an air bubble-in a somewhat cond ucting liquid. In this extreme limit of the
general case where fl uids (a) and (b) can have arbitrary electrical conductivi­
ties and permittivities, it is appropriate to view the bubble and its surround­
ing conducting fluid as a resistance in series with a capacitance (the bubble),
in turn in series with a resistance. This makes reasonable the polarity of the
i nd uced surface charges indicated in Figure 7a, and the surface shear-force
densities also sketched there.
Our approach to describing a self-consistent shear flow is essentially the
same as in the case of the periodic convection, with one exception. In the
previous case, gravity is used to hold the normal electric stresses in balance.
In the case of the sphere we find a u nique combination of fluid properties
that make possible a spherical equilibri u m of the interface. Then it is possible
to deduce whether the drop tends toward a prolate or oblate geometry for
combi nations of physical parameters other than those required for a spherical

Electrical stresses.-In the limit o f zero electric Reynolds n u m ber, the


equilibriu m .

solution for the electric-field i ntensities is the classic one, ( 1 2) where the
field in the interior of the drop is u niform, while that outside is a su perposi­
tion of a u niform field and a three-di mensional dipole field. There are four
1 24 MELCHER & TAYLOR

'I JI
I
I
I 0
6 Two-d imensionol I
I
theory I
I
I 0
I
5 I
I
Theory corrected
for side walls I
I }
I
I
0
Q) 4 I
en
"-
I 0
I
E I
,
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

E I
I
!:J 3 I
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

I
I 0
I
I
I
2 ,'0
I
I
I

I
15

FIG. 6. Peak tangential velocity U of the interface as a fu nction of applied potential


for the case depicted in Figure 4 and photographed in Figure 5. The liquids are Dow
Corning FS·1265 below (Eb/Eo=6.9, ub=3.3XIO-9, }Lb=0.37) and corn oil above
(Ea/'o = 3.1 , cra = 5 X lo-n, }La 0.055). The cells are somewhat distorted at the top be­
=

cause of a rigid boundary positioned approximately at the topof the picture in Figure 5.

boundary conditions to be satisfied: (a) that the field be finite at the origin;
(b) that the tangential electric field be continuous at the i nterface ( Eq. Ij),
[Ee)(r=b)=O; (c) that the conduction cu rrent normal to the interface be

(d) that E----7Eo(ircos O-iesinO) as r----7OO. Thus, the electric potential in each
conti nuous (Eq. II in the limit where K----70) , [<TEr)(r=b) =0; and finally,

region is
1 -
<Pa = - Eo cos 0 r +
R�)
2 + R r2
( 25.

<Pb - 3Eor co�O/(2 + R) 26.


where again, R=<Tb/<Ta and in spherical coordinates
=

-
E
. a<p . 1 a<p
1,--19-- 27.
ar T ao
=

Direct substitution shows that cp satisfies Laplace's equation and the neces­

1" = n· Ire) in the radial and tangential directions follow by direct substitu-
sary boundary conditions. The components of electric surface-force density
ELECT ROHYDRODYNAMICS 125
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

{oj
(b)

FIG. 7. (a) Spherical drop or vapor void having radius b and comprising region (b) is
immersed in a liquid (region a), which makes electrical contact with electrodes that
impose an electric field that is u niform and of magnitude Eo far from the sphere.
Surface charges, induced as shown for the case in which the outer fluid conducts much
more than the inner one, interact with Ee to create the shear-force density To'. (b)
Flow pattern resulting from field-fluid configuration of (a).

tion of the last three expressions into the relations (see Eq. 7)
Tr' = [T,r'] = ! I [oE?] - [oEe']) (r = b) 28.
Te< = [T,e'] = roE,Ee](r = b) 29.

The choice of appropriate stream functions with corresponding viscous shear


stresses that can hold the tangential and normal components of the electric
surface-force density in equilibrium at each point on the interface must be
made on the basis of the 8 dependences of Te. Observe that T/ is propor­
tional to a constant term and a term in cos'8, while Tee has the 8 dependence
cos e sin e.
Viscous shear stresses.-In spherical coordinates, it is appropriate to use
the Stokes stream function if; defined such that (13)
.[ 1 a", ] [1 a", ]
;2sin 0 ao rsin 0 ar
.
v = I, 30.
- Ie

and for creep flow the eq uations of motion require


[ a' sin 0 a (1 a )J'1/;=0
ar' r' ao sin (J ao
�+ - - --
31.

The boundary conditions dictate the 8 dependence of the variable separable


solutions that we seek from Equation 31. For now, we consider that the in­
terface is in radial force equilibrium. Then the boundary conditions are
essentially of the same nature as those for the previous case of periodic con-
126 MELCHER & TAYLOR

vection, except that, because the origin is included in the interior region,
there are now only two solutions of interest in each region, and only four
boundary conditions at the interface must be met. These are: (a) and (b),
that the radial components of velocity vanish at the interface, vra(r=b)
=vrb(r=b) =0; (c) that lve](r=b) =0; and (d) that u+[Wa/ar(v6/r»)(r
= b) O. In this last condition, we evaluate the viscous shear stresses in
=

spherical geometry (14) and take advantage of the tact that Vr vanishes at
every point on the interface.
As this last boundary condition is expressed in terms of if; (Eq. 30), ob­
serve that, if the condition is to be satisfied at every angle (J, if;/sin (J must
have the same dependence as Tee. Thus we are led to look for solutions of the
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

fer) sin' 8 cos 8


form
.p 32.
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

and substitution of this expression into Equation 31 shows thatf(r) has the

origin and as r-> 00, that the appropriate linear combinations of solutions are
form rn, n = - 2, 0, 3, or 5. It follows from the conditions on the flow at the

.pa = (A b'r-' + Bb') sin' 8 cos 8


>II' (Cb-1y3 + Db-ar5) sin' 8 cos 8
33.
= 34.
The four constants A, B, C, and D are fully determined by the four boundary
conditions. From either Equation 33 or 34, the velocities at the interface

b) = 2U cosO sin 8
take the form
Vu(r = 35·
where U is the peak velocity. The first three boundary conditions give A
= -B C= -D = U. Finally, the balance of interfacial shear-force densi­
=

ties relates the peak surface velocity U to the applied electric-field intensity
to complete the determination of the fluid response:
9fbE02b(RS 1)
U
-

10(2 + R)2(Pa + Pb)


= - --,-----.,.---,- 36.

In the case of Figure 7, the drop is highly insulating compared to the


surrounding fluid. Thus, RS < 1 and Equation 36 shows that the convective
response of the fluid is in the direction expected from the sign of the surface
charges. The streamlines of Figure 7b, based on Equations 33 and 34 for the
case RS < 1, are also as would be expected in view ·of the shear-force densi­
ties sketched in Figure 7a. From Figure 7 or Equation 35, it is evident that
fluid at the interface has its maximum speed at (J= ±11"/4 and ±5 11"/4.
The close relationship between the cellular convection within and around
a spherical drop and the convection produced by a periodic imposed field is
emphasized by a comparison of Figure 4 and 7, or a comparison of Equations
23 and 36. It is not surprising that the sign of (RS-l) determines the sense of
the convection, because if a given electric field is applied to an interface, it
is this function of conductivies and permittivities that determines the sign of
the resulting surface-charge density. Of course, as in the cases of Figures 1
and 4, a reversal of the applied potential polarity reverses the sign of both the
tangential electric-field intensity and that of the surface charge; hence, the
dependence of U on the square of Eo is as expected.
ELECTROHYDRODYNAMI CS 127
Radial stress balance; oblate versus prolate.-By assu ming the interface to
be spherical, we have been able to ignore the radial surface stresses. N ever­
theless, they are present and will now be taken into account. The electric
field produces the surface-force density given by Equations 25 to 28. Further
contributions come from [TTrm], (Equation lIb) , the viscous part of which
follows from Equations 30, 33, 34 and 36. The pressures pa and p b are deter­
mined by using the known velocity to integrate the equations of motion:
l' = ITa - 2U.uab2r-3(3 cos2 0 - 1) 37.
38.
where IIa and lIb are constants. There is also an effective radial-force density
- 2 T/b because of the surface tension T. Finally, we assume that there is a
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

force per unit area To cos28 at our disposal that (for To> 0) tends to elongate
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

the interface in the direction of the applied field.


The balance of radial stresses thus requires that

b) - - + Tocos20
2T
T,e + [T,r'n](r =

b = 0 39.

It is remarkable that each contribution to Equation 39 is either constant


or proportional to cos2 (). The constant part is satisfied by adjustment of the
relative hydrostatic pressures n. The coefficients of the remaining terms in
cos' 0 sum to zero i f the externally applied surface-force density To is ad­

To - 9f&E02if>/2(2 + R)'
justed to be

3 .
=

S(R2 + 1) - 2 + - IRS - 1) ---


(2M + 3)
5 '
if>
= 40.
M+ 1

Thus, if the ratios of the fluid parameters represented by R, S, and ]v[


are adjusted such that <I> = 0, the drop can be in steady-state equilibrium.
Further, if <I> < 0, a positive outward-directed surface-force density at the
poles «()=O, 0=71") is required to retain the spherical shape, and we conclude
that in the absence of To the drop would decrease i ts extent in the direction
of Eo (i.e. , become an oblate elli psoid) . Similar reasoning shows that for
<I> >0, the interface is prolate. The function <I> di sc riminates between equi­

Experiment.-Observations of the convection in drops, virtually as de­


libria of oblate and prolate geometry.

scribed, are documen ted by Allan & Mason (15) and McEwan & Dejong
(11). Experiments are complicated by the need for a neutrally buoyant com­
bination of liquids to obtain a stationary drop, and the tendency of any re­

Figure 8 convincingly show cellular convection streak lines from particles


sidual charge to make the drop migrate. Nevertheless, the photographs of

illuminated over the cross-sections of the drops. The figure caption gives
further information on the experiment.
The model appears to correlate successfully with observations of oblate
and prolate ellipsoidal equilibria (11), Two limiting cases are of particular

the vehicle liquid, so that R-+ 00 (for example, a water drop in insulating
interest in this regard. Suppose the drop is highly conducting compared to

oil). Then <f> >0, and the equilibrium geometry is that of a prolate elli psoid.
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

128

(.. )

(b)
MELCHER & TAYLOR
ELECTRO HYDRODYNAMICS 1 29

Of course, in this limit the shear stresses make n'o conf�ibution because the
electric field acts normal to the i nterface; hence it is not surprising that the
viscosity ratio M does not play a role. Work on the stability of this ellipsoi­
dal equilibrium of a highly conducting drop justifiably excludes the effects
of the electric shear forces (16, 17).
In the opposite extreme, where a void of gas is suspended in a slightly
conducting liquid, and thus R-->O and M--> 00, the discriminating function
becomes <I> 5-16/5, and the geometry of the equilibrium depends on the
ratio of permittivities 5=ea/Eb relative to 16/5.
=

STEADY CONVECTION: AC F IELDS


Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

Steady convection in dc fields, as illustrated by the experi ments of Figure


1, 4, and 7, obtains only if there is an electrical-conduction path between the
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

source of potential and one of the fluids. In the absence of such a path, thr
fluids simply polarize, with the electric field confined to the i nsulating re­
gions and directed perpendicular to interfaces. For example, observe that

= R-->O (Eq. 23).


in the case of the periodic cellular convection where fluid (b) is adjacent to
the electrodes, the interfacial velocity vanishes as Ob/O'a
'
Similarly, convection of the spherical interface, represented by Equation 36,

that i mpose the field Eo, becomes small compared to that of the drop (R--> (0 ) .
vanishes as the conductivity of fluid (a), which is adjacent to the electrodes

I nterfacial electrical shear stresses can be induced by means of ac fields


without the need for electrical conduction between the source of potential
and the fluids, In electrical terms, the coupling is capaci tative and analogous
in many respects to the inductive process by which time-varying magnetic
fields couple to the rotor of an i nduction machine, or to the liquid metal of a
magnetohydrodynamic induction pump. Here we are concerned with ac
electric fields and charge relaxation , rather than with ac magnetic fields and
current diffusion. Early work on rigid-body motions serves not only to give
historical perspective but provides us with a convenient prototype model
for understanding ac-field surface in teractions.
STEADY ROTATIONS IN ROTATING FIELDS
The cross-section of a circular cylindrical rotor (b) immersed i n a fluid
(a) is shown in Figure 9. We illustrate the effects of an ac field by considering
the consequences of subjecting the rotor and fluid to an electric-field in­
tensity Eo, which rotates with the angular velocity w . The rotor, hence the
cylindrical interface, has the angular velocity Q.
Consider for discussion the case where the fluid is much less conducting

-{ ««(

FIG. 8. (a) Cross-sectional view of silicone oil drop in mixture of castor oil and
corn oil with electric field applied vertically, as shown in Figure 7a. Particles of
powder entrained in interior of drop show streak lines with the pattern of Figure 7b.
(b) Particles in exterior liquid showing streak lines essentially similar to those of
Figure 7b. [after M cEwan & Dejong (10].
130 MELCHER & TAYLOR
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

FIG. 9. A Cylindrical rotor (b) is immersed in a fluid (a) and subjected to an


imposed electric-field intensity Eo that rotates with the angular veloc ity w.

than the cylinder (RS> 1). Then, if the applied field is static as viewed from
the rotor, it would be shielded from the interior region and oriented per­
pendicular to the interface; there would be no interfacial electric shear lorces.
By constrast, with the electric field rotating with respect to the cylinder at
the frequency w-Q, the finite relaxation time for charges to polarize on the
interface comes into play, and in particular, if the period of the field as
measured in the rotating frame of the interface is on the order of the electri­
cal relaxation time (21fjW"""Ebjub), surface charges are induced which are not
in spatial phase with the electric-field intensity. As illustrated in Figure 9, if
the field rotates more rapidly than the rotor, th e r e is an effective dipole mo­

ment from the induced charges that lags Eo, the imposed electric-field in­
tensity, and a resultant shear surface-force density in the clockwise direc­
tion. By subjecting a fluid system to a time-varying electrical excitation, it is
possible to create finite-relaxation-time effects, even though the flow is in the
steady state.
The electromechanical effect of static and rotating electric fields on
cylindrical and spherical, slightly conducting rotors has been the point of
both theoretical and experimental investigations since the early work of
Arno (18). Rotations induced by dc fields have seen particular and periodic
interest and form the background for the class of instabilities to be discussed
in the next section. An excellent historical review of the subject is given by
Pickard (19), who also discusses torques induced because of the finite time
required for dipoles to relax. This latter effect, not considered here, becomes
significant at much higher frequencies than are usually of interest in electro­
hydrodynamics, but nevertheless deserves more attention in connection with
the electro mechanics of fluids.
Rotating jields.-It is a simple matter to give quantitative substance to
our discussion of the rotor dynamics. The imposed electric field is taken as
ELECTROHYDRODYNAMICS 131

uniform far from the rotor axis, and hence, as r cos (O-wt) ---'> oo ,
<l>a ....... - Eor cos (wi - 8) 41.
There are three additional boundary conditions : (a) that [cfJ] (r = b) =0 (Eq.
II) ; (b) that charge be conserved at the interface :

(�at !1 �) [EEr] = 0
ao
[uEr] + + 42.

and (c) that the fields be finite at the origin. Note that for the first time in
our discussions, we include finite-electric- Reynolds-number effects by re­
taining the convection surface current represented by the term in Q of
Equation 42.
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

Fields take the form of a superimposed dipole and uniform fields in the
exterior fluid and a uniform field in the interior region :
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

7 exp j(wl - 0)
<1>. =
[
R e - Eor +
]
Ab2
43.
<l>b = Re rB exp j(wt - 0) 44.

Solution 43 already satisfies condition 41, while Equation 44 is finite at r = 0


and substitution in the remaining two boundary conditions determines A
and B. From these solutions it follows that the surface-charge density is
2EoEb [SR - 1 ] cos (wi - 0 'Y)
Q = vI + R�(1 + R)
-
45.

where
46.
The familiar quantity SR- 1 once again determines the sign of the induced
surface charge.
To corroborate our introductory discussion, for the case of a conducting
cylinder in an insulating fluid, SR > 1, and if w >Q, Equation 4S shows that
'Y is positive so that the axis of the charge distribution in fact lags that of the
applied field as sketched in Figure 9.
Note that RE is an electric Reynolds number composed of the ratio of a
hybrid relaxation ti me (f,, +f/» /(U,, + Ub) to a transport time. The latter quan­
tity is the time required for a point on the interface to traverse a peripheral
distance b, relative to the frame of the rotating electric field. It is significant
that RE can be adjusted by controlling the frequency w of the applied field.
Induced torque and rotation.-The electrical torque per unit axial length
of the rotor is bQEe(r = b) , integrated over the surface of the cylinder :

Te =
4trEo2E.b2(RS - 1) --
RJ?
47.
(1 + S)(1 + R) 1 + R�
Note that this torque has maximum value as RE = 1 and can be positive or
negative, depending on the sign of (RS-1).
If the fluid is of essentially infinite extent, the steady-state viscous torque
per axial length of the rotor is P = - 41rJ.'Slb2, and under the assumption that
132 MELCHER & TAYLOR

no other torques are present, the balance of torques requires that


•• Eo2(RS 1)
1l.(1 + R) (1 + S) ( + R�)
RE
n =
-

48.
1
Of course, the dependence on RE makes this expression i mplicit in the angu­
lar rotor velocity n, but because the electrical frequency w also determines
RE, we are justified in regarding the electric Reynolds number as being in­
dependently controlled.
By contrast with the dc-conduction-driven flows, we now have the pos­

pared to the rotor. That is, the limit of Equation 48 as R� 00 does not ap­
sibility of induced motion even if the outer fluid is highly insulating com­
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

proach zero.
A graphical representation of the torque balance is given in Figure 10,
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

where ye and TV are sketched as functions of n. Thus, the i ntersections


between the curves represent the solutions to Equation 48. I n the case i n
-

1
which RS > (a conducting rotor i n a n i nsulating fluid, perhaps) the only
equilibri u m (i) consists i n a positive rotor velocity, less than that of the
field, with the axis of the charge lagging that of the imposed fIeld.
I t is possible to have three equilibria for the case RS < 1. For a weak
field, the only equilibriu m is (ii) , with the rotor and field rotating in opposite
directions and with the charge axis leading the i mposed-field axis by more
than 90°. As the field is increased, two positive velocity equilibria are possi­
ble: (iii) , with the lower velocity, is unstable because any slight i ncrease i n
rotor velocity tends t o increase the electrical torque and hence t o i ncrease

RS > I

"Eo
(0)

RS<I

( b)
�,�tB.
'
.. "
}
(i i J
'
Te I
,

FIG. 10. Electrical torque T- and viscous torque Tv as functions of the rotor
-

velocity n normalized to T. ('a +<1» /(0'. +O'b). Intersections represent possible veloci­
=

ties for stea dy sta te rotation.


-
ELECTROHYDRODYNAM ICS 133

further the rotor velocity ; and (iv) , with the greater velocity, i s stable. I n
these cases, the rotor and field rotate in the same direction, but the rotor
angular velocity exceeds that of the applied field. Of course, equilibria (i)
and (ii) are stable.
I t should be clear that a spherical rotor stressed by a rotating electric
field would be motivated by a torque having a dependence on the physical
parameters similar to that for the cylinder. I n fact, much of the early work
relates to spherical rather than cylindrical rotors. Thus, drops and bubbles
under the influence of rotating or traveling fields can be expected not only
to u ndergo cellular convection, but to suffer rotations as well. As will be
developed in the section on finite-electric- Reynolds-number instability, these
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

rotations can even be expected for drops and bubbles in de fields. Naturally,
a rigid-body model for the spherical region would be j ustified only if it were
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

composed of a highly viscous liquid (M-.O) .


We now consider a case in which each region is occupied by a fluid in a
configuration arranged to give a simple but basic electrohydrodynamic flow.

TRA VELlNG-WA VE- INDUCED CONVECTION


The combination of fields and fluids shown in Figure 11 represents a

inherent to the experiments of Figures 4 and 9 and investigated by Melcher


physical situation strongly suggested by combining the basic interactions

(20) . As in the case of periodic cellular convection, the interface between two
layers of liquid is stressed by a field from a segmented electrode having a
spatially periodic distribution of potential of wavelength 2'Ir/k. By contrast
with the dc case and in a manner suggested by the rotating-field example,
this potential distribution is made to travel in a direction parallel to the in­
terface. Thus, the lower fluid, region (a) , can be regarded as the cylindrical
rotor "laid out flat," and the segmented electrode as a means of producing a
field at the interface having essentially the same space-time properties in
linear geometry as the rotating field has in cylindrical geometry.
For discussion purposes, consider the limiting case where the upper fluid
is an insulating gas and the lower one is a slightly conducting liquid. Then
negative charges induced on the electrode by the applied potential in turn
induce i mage charges of opposite sign on the interface. Because the electrode
charges travel to the right with a velocity such that the field induced in the
frame of the moving interface has a period on the order of the relaxation time,
charges on the interface lag their i mages on the electrodes, as sketched in
Figure 11. Thus, in this case of RS > 1, there is a shear-force density on the
i nterface acting to the right.

do not travel with the same velocity U as the interface, the shear stresses at
Because the potential wave and its attendant charges on the interface

the interface are pulsating with time at the frequency 2 (w-k U) . I n the fol­
lowing we will make the assumption that this frequency is sufficiently high,
compared to characteristic times of fluid-mechanical response, that the fluid
responds only to the time-average electrical shear stress. Thus, the interfacial
velocity U is taken at the outset as being independent of x.
1 34 MELCHER & TAYLOR

_I Electrode potential Re Voexp. j (wt - k x )

..------: x
t
-----....
Y Seg mented electrode with impo s e d

J traveling potent i a l wave

1 1 1 11 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 111 1 1 1 1 1 1 1 1 1 1 1 1 1 11 1 1 1 1 1 1 1 1 1111 1 1 1 1 1 1 1 1 1 1 1 1 1 1 11 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1111111

�----� ��r
' Y� ' .,.· +.·.·•.•· · �.· .i·. -<· . .
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

. ..
.
. .

. :. : . .. .: :::, : : . . ,::" ,: :' ;-"':« ':'. :: . . ...::.,: ::.:: . ,- .: . ..�.: .� . ,. : '( .b )...:' .:· . .
. : " ' b
. '

. .. . .
..
' , : - .,..
,, '•,
• ", '
"
• • • " . " 0 • • • • •
. '

FIG. 1 1 . Cross-section of layers of fluid (a) and (b), with an interface at y = O.


Shear forces are induced on the interface by the traveling potential wave imposed on
the segmented electrode.

Traveling-wave fields and induced shear stresses.-With the assumption

matter to determine the fields, even including the effect of the convection on
that the interface is moving with a constant velocity, it is a straightforward

the interfacial charge distribution. The potentials in each of the fl uids take
the traveling-wave form
cf> = Re(Aa,b sinh ky + Ba,b cosh ky) exp j(wt - kx) 49.
with the four constants Aa, A b , Ba, and Bb determined by the boundary con­
ditions. These conditions require : (a) that <fJa (y = a) = Re Vo exp j (wt kx) ; -

(b) that <fJb (y = - b) = 0 ; (c) that [<fJ](y = 0) = 0 ; and finally, (d) that charge be
conserved at the interface, [eTEy] + (a;at+ Ua;ax) [fEy] = O. An example of

b� 00 and eTa = O.
the electric-field distribution is given in Figure 1 1 , where for the case shown

With the fields determined from the known potential distribution, the
electrical shear-force d en si ty, Exa[EElil is evaluated and ti m e averaged to
obtain
-

(Txe) = !k'VO'EaEb<Ta sinh kb cosh kb(RS - 1)RE/ [ 1 +RE'jA2; 50.


where
RE = (w - k U)A/J:-

A = Ea cosh ka sihn kb+Eb sinh ka cosh kb


J:- = eTa cosh ka sinh kb+eTb sinh ka cosh kb
ELECTROHYDRODYNAMICS 135

Note the similarity of this result and the expression for the rotor torque,

Flow equilibrium and experiments . Consider the case where the response
Equation 47, particularly as it depends on RE and (RS- 1 ) .
-

of the fluid is in plane Couette flow. Then the viscous shear stresses combine
to give an effective surface-force density T"'= U(p.a!a +J.lb!lJ) , and the
-

balance of viscous and electrical shear stresses requires


51.

This last expression takes the same form as Equation 48 in its dependence on
RE, n, and RS- l . Thus, if we simply think of the lower fluid as being the cyl­
inder "laid out" in plane geometry, it is clear that the steady-state equilibria
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

of Figure 10 pertain equally well to the system of two fluid layers. Charges
are distributed over one wavelength on the interface essentially as they are
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

on the periphery of the cylinder.


In the case of a gas over a liquid, RS > 1 and the fluid travels in the same
direction as the wave, but with less velocity. Experiments (20, 2 1 ) verify
this prediction as well as the dependence of the interface velocity on the

measured dependence of U on the driving frequency w . In this experi ment,


applied frequency and peak voltage. As an example, Figure 12 shows the

the potential-wave velocity greatly exceeds that of the fluid. The solid curve
has the predicted frequency dependence, but is normalized to the peak ampli­
tude of the data. Differences between the absolute observed and predicted
velocities in this experiment are on the order of 5 to 30 per cent, depending
on a and b relative to the length of an el ec tr od e segment ( 1 6 m m) . Further
particulars are given in the figure caption.
One case where RS < 1 consists in placing the electrodes under the insulat­
ing bottom of a channel filled with a layer of slightly conducting liquid. In
the terminology of Figure 1 1 , this is essentially equivalent to making region
(b) an insulating gas and region (a) a slightly conducting liquid. So far as the
analysis is concerned, it is irrelevant that the system is turned upside down.
In this configuration of electrodes covered by a liquid and then a gas, the
liquid interface is experi mentally observed to travel in a direction opposite
to that of the traveling wave ( 2 1 ) , as is consistent with the negative velocity
equilibrium (ii) of Figure 10.
The stable flow equilibrium (iv) with RS < 1 and the interface moving in
the same direction and faster than the traveling wave must generally be
established by using external means to i mpel the fluid. At the same time,
equilibrium flow (iv) requires a minimum voltage before the viscous stresses
are balanced by the electrical shear stresses. If by dint of external forces the
fluid reaches the velocity (iii) in Figure lOb, with the required threshold of
applied voltage, it continues to accelerate until it reaches the stable equili­
rium (iv) . Thus, an i mportant exception to the need for an external starting
mechanism is the limiting case in which w�O so that the unstable equilib­
rium coincides with RE = 0 . Such instabilities, in which an initially static
fluid spon taneously establishes the equilibria (iv) or (ii) of Figure lOb, are
the subject of the next section.
1 36 MELCHER & TAYLOR

S H EAR- I N D UCED I NSTA B I L I T I ES


STEADY ROTATIONS IN A STATIC FIELD
The physical mechanism basic to a class of instabilities found with RS < 1
can be described i n terms of the rotor of Figure 9 , now with the applied elec­
tric-field i n tensity constant (w = 0) . I n Figure lOb, the torque curve P passes
through the origin of the rotor-velocity axis. For small applied fields, the
only equilibri u m is with the rotor stationary (at the origin in Figure lOb)

- Tv curve, the equilibri u m at the origin


and that one is stable. As the field is in creased to a level such that the slope
of the P curve exceeds that o f the
becomes u nstable, and stable equilibria (ii) and (iv) are possi ble. The rotor

48)
will spon taneously reach a steady-state rotation in either direction, i f (from
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

vI - RS . /�
Equation

---- = 1 ; 11 --
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

H. H, == Eo 52.
1 + R Jl.aCTa

This expression gives the threshold condition for an i ns t a bil i t y that


would not arise were i t not for the i nfluence of the convection on the charge
distribution (finite-electric- Reynolds- n u mber effects) . With RS < 1, surface
charges induced by Eo have a dipole moment anti-parallel to the direction of
Eo. The induced charges are carried in the direction of the rotation by the
i n terface and this resulting deflection of the i nduced dipole moment gives
rise to a torque that tends further to i ncrease the rotational velocity [see
sketch of Figure lOb for equilibrium (iii)].

...
41
<J)

...
"-
E
I
::>

o I I I I
o 2 3 4 5 6
w I 2 7T - H z

of Figure 1 1 , with region (a) air, and region (b) M onsanto Aroclor 1 232 (Eb = 5.73 <0,
FIG. 12. Surface velocity as a function of traveling wave frequency for the system

CTb "" 10-9 mh os/m ). The channel takes the re-entrant form of a race track of width 5 . 1

c m , length 0.886 m equal to the wavelength, and depth b= 1 .6 cm.


ELECTROHYDRODYNAMICS 137

The condition of Equation 52 for spontaneous rotation represents the


point at which the electrical torque attributable to those charges induced by
the motion competes with the viscous torque. Thus, by its a nalogy with the
Hartmann nu mber of magnetohydrodynamics, He is the electric H artmann
n u mber.
Observations o n rotations of dielectric spheres and cylinders i m mersed
in slightly conducting fluids stressed by dc fields have been recorded pe­
riodically since Quincke's observations (22) . An indication of the early, as
well as the more recent, l i terature o n this subject is given by Pickard ( 19) .
The conditions for incipient instability predicted by Equation 52, and also
the predicted parameter dependence of the steady rotational equili bri u m
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

velocity, have been well docu mented, a t least for certain well-behaved fl u ids.
We confine our further attention here to electrohydrodynamic instabilities
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

which, as wi th the rotor, have a physical basis in the competing processes


of charge convection and viscous dissipation.

I N STABILITY I N THE INTERFACIAL PLANE

When the spatially periodic, i mposed potential is static, the fluid i n ter­
face of Figure 1 1 , like the rotor of Figure 9, can be u nstable. Of course, the
i n terface between the fluids, u nlike the surface of the rotor, is not constrai ned
to rigid-body motions. Even so, we can obtain insights into the nature of
electroconvective instabilities by further considering the i mplications of
Equation 50. This expression, derived under the assu mption that the fluid
res p o nds wi th a u niform i n terfacial velocity
U to the average stress, i s based
on an approximation that is excellent for a rapidly traveling wave, but highly
questionable in the limit w'-'O.
I n a manner analogous to that used i n obtaining Equation 52 from E q u a­

the rate at which < P > i ncreases with U, i n the neighborhood of a static
tion 48, we take the limit of Equation 51 where w.-.O, and then require that

equilibri u m U = 0, j ust equals that with which Tm i ncreases with U. The


-

resulti n g condition for spontancaus translation of the i nterface, either to the


right or left, is

53.
vI RS t.
HE
-

[1 + R tanh ka coth kb ] = 1;

where
t. = [ak sinh kb cosh kb ]1/ [ 1 + alb M]I (cosh ka sinh kb)
Note the similarity of this expression and Equation 52 for the rotor. The
same fundamental processes are at work in each case of instability.
With the i m posed spatially periodic potential static, i t is clear that the
configuration of Figure 1 1 has the same i ngredients as used to produce
_

periodic cellular convection (Fig. 4). In fact, if an attempt is made experi­


mentally to demonstrate the periodic roll-cell convection described in con­
nection with Figure 4, and a sufficiently nonconducting liquid system is used
that Equation 53 is approximately satisfied, the dominant cellular motions
138 MELCHER & TAYLOR

observed are likely not to be the expected ones in the x-y plane, but rather
to be in the x-z plane of the interface. Conditions similar to Equation 53,
involving the electric Hartmann nu mber, would determine whether or not
the cellular convection with the spherical interface of Figure 7 could be es­
tablished without other dominating motions, such as the rotation of the
drop or bubble.
We now discuss experi ments in which these charge-convection instabili­
ties in the plane of an interface have been studied. In the first, Jolly (23)
used a configuration very similar to the one j ust described to produce cellular
motion in the plane of an interface by means of a nonuniform field. In the
second, Malkus & Veronis (24) studied similar rotations but produced by the
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

curvature of the interface in an essentially uniform i mposed tangential field.


Incipience in a nonuniform field.-An experiment which gives graphic
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

evidence of electroconvective instability in the plane of an i nterface is shown


in Figure 13. Electrodes of alternate polarity are spaced along the bottom
of an insulating container. A layer of slightly conducting liquid then covers
these electrodes to a depth a. For the present discussion, the upper region
(b) is air, although as long as RS < 1 it might be a second liquid.
Motions of the interface are observed by entraining strips of bubbles in

As the voltage V. is raised, there is a threshold at which the interface begins


the regions of the interface above the lines of abutment between electrodes.

to swim in i ts own plane. The appearance of the i nterface as motions begin is


shown in Figure 14, and eventually the interface establishes a pattern of
cellular convection. The smaller the depth, the shorter the initial wave­
length of instability and the smaller and more nu merous the cells. Figure 1 4
shows how the wavelength increases with the depth. For (ka) larger than
about 1, cells formed from one pair of electrodes begin to interact with those
from another pair; the patterns on the interface are not so regular in ap­
pearance, and the interface begins to move as a whole, interacting with the
boundaries of the container.
Even if available, an exact analysis of the conditions for i ncipient insta­
bility is more complicated than is appropriate in this review. However, by
making suitable approximations we can establish that the physical phe­
nomenon shown in Figure 14 is of the same nature as gives rise to spontane­
ous rotations of the rotor. We have arranged the coordinate system of
Figure 13 so that the instability experiment can be regarded as a special
case of the arrangement shown in Figure 1 1 . Thus, Equation 53 gives the

the case o f the instability experiment we have the limit b-H.() , M-+ 00 and
condition under which the interface is unstable to uniform translations. I n

R-+O. In addition the imposed potential is a square-wave function of x,

Vo = 4 V./1r. The resulting condition of incipient instability is


which we approximate by considering the fundamental Fourier component ;

/ '.<1>
. 'V -- 54.
4kV •

HE = cosh ka/vlka; HE =
.

11" 2p..u.
We expect that this relation would give its best approximation when tne
ELECTRO HYDRODYNAM ICS 139

z
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

top view

FIG. 13. A layer of liquid makes contact with strip electrodes having alternately
the constant potentials ± VB. Cellular motions in the plane of the interface are estab­
lished by instability and appear as sketched in the top view and as shown by the
photographs of Figure 14.
characteristic motions of the interface overlap several electrodes ; i.e., for
large depths, a. I n any case, it seems likely that it is an upper bound on the
voltage required to produce instability, because it accounts for only one of
many possible motions the i nterface can execute. The predictions of Equa­
tion 54 are compared to the experi mental measurements of the voltage re­
quired for incipient instability in Figure 1 5 . As expected, the si mple theory
appears to provide an upper bound on V •.

To obtain a lower bound on the voltage for instability, it is reasonable to


postulate that if at any point on the interface the electrical stress is a more
rapidly increasing function of surface velocity than is the viscous stress, the
fluid begins to move. Again , the i nterfacial motions are taken as u niform, so
that the electric stress can be calculated in a manner analogous to that used
in determining Equation SO. Now, the stress used in Equation 5 1 is evaluated
at the coordinate origin in Figure 13. Following the same procedure as with
the average stress, the condition for incipient i nstability using this " poin t­
stress" approximation is
lI E = cosh ka/2[ka(1 + tanh ka/S) ]l 55.

The voltage predicted by this expression, also plotted in Figure 1 5 , tends


to bound the measured values from below. At depths less than a 1 cm, the =

higher harmonics in the Fourier expansion make an essential contri bu tion


to the prediction of V" and if these are incl uded in the analysis (23) , the
point-stress prediction remains below the measured value even for small
depths. Yet a third model, based on representing the i ndividual cells as
fluid- mechanical rotors, is successful in predicting the incipience to within
20 per cent over the range 0.2 < ka > 1. For the present purposes, our approxi­
mate analysis supports the physical basis given for the i nstabilities. Jolly
1 40 MELCHER & TAYLOR
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

(:t)

.. . ,t�·, : .. . .. ,f . ' �

. . ',' (''''. . ....' , . ',


,
... .. ; ' J .:� ' v

. "'�

. .
. •
:' . . .: L ." '., .�¥..
. •( • v..l. .
� . ... . 'I. :,.,
" .

, h I ." .�\ � " :.. .' ... " �


• ' • . .,. ·' J

, . . .
' .; .
r
... ' '.t " I ,
.
.' 1" " ,"lot '. �-.. . .. .
, . ' �.
/,tl'...\ . . :. t �.,\... .. "
�'� I." ·"
�"""

� :I, ....�
.. t .. t �. ,,\ , ' \'
..
. ' • I

(b)
FIG. 14. Top view of the experiment of Figure 13. Air bubbles are entrained in
fluid injected in �trips over regions between electrodes ; an instant after incipience of
instability, bubbles are carried by the interface to the positions shown. (a) k a = O.26,
(b) k a = O.5, (c) k a = O.8. I n (b) , the elasped time is sufficient that cells are beginning
to form.
ELECTROHYDRODYNA M ICS 141
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

(c)

the electrohydrodynamic shear stresses, to predict the transverse (z direc­


uses energy argu ments (23) , together with a judiciously chosen model for

of 0 . 2 < ka <
tion) wavelengths of incipient instability to withi n 5 per cent over the range
1 .2.
Incipience o n a curved interface.-M uch current in terest in t h e su bject o f
surface electroconvection can b e attributed t o a paper b y M alkus & Veronis
(24) in which a range of electrohydrodynamic phenomena are highlighted
and detailed attention is given to cellular interfacial motions akin to those
just described. The experi ment sketched in Figure 16 combines an essenti­
ally uniform electric field i mposed tangential to an interface with a poten­
tially unstable distri bution of su rface charge induced because of a slight
interfacial curvature. For analytical purposes, the vertical deflection of the
interface is approximated as �={3(x - d/2)2. For the case shown, {3 > 0. Wi th
the center of curvature on the side of fluid (b) , the lower fl uid m ust be the
more conducting to be consistent with the potentially u nstable distribution
of equilibri u m surface charges shown in Figure 16.
Experi mentally, it is found that, as the voltage is raised, a threshold is
reached at which cells with the appearance shown in Figure 16 are set in
motion . Each cell can be viewed as a rotor in a dc field, with RS < 1 . With
the assumption at the outset that the upper fluid is air, and hence m uch less
conducting and viscous than the liquid below, M alkus & Veronis develop a
rather complicated eigenvalue theory that predicts onset of electroconvec­
tion at
H, = [O.337r/J3d(1 + l /S) ]l; H, = EOV'a'6/<1al'a 56.
1 42 M ELCHER & TAYLOR

4 overage stress t heory } o

� 3
I

7
>'"
o
2 o o
0 0

�o
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

0point stress theory }


o o
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

o 0

o I I
o 1.0 2.0
depth a - em

FIG. 1 5 . Voltage for i ncipie nce of instability in the plane of the interface. The ap­
paratus is sketched in Figure 13, and the appearance of the interface is shown in
Figure 14. Theoretical approximations are Equations 54 and 55. The fluid is corn oil,
ua = 9.5 X l0-n mhos/m, Ea = 3 . 1 EO and iLa = S .46 X I 0-2 kg(ms)-l.

Note the si milarity between this expression and Equation 55. The new in­
gredient i n Equation 5 6 is the surface curvature, represented by f3d.
Using pinene, i t is found that Equation 56 predicts a threshold voltage
75 per cent of that observed. This order of difference is expected since the
theory is based on the assu mption that the electrode boundaries do not re­
tard the shear flow, but that rather, at these walls, the fluid is free to slip.

y
+
+
+ ( bl
+

s i d e view top view


FIG. 1 6. Plane, parallel electrodes having the dc p tential difference V. induce the

potentially unstable distribution of surface charges shown on a slightly curved inter­
face between fluids (a) and (b) ; it is assumed that RS < 1 . Cellular convection in the
plane of the interface is akin to the spontaneous rotor rotations.
ELECTRO HYDRODYNAMICS 143
Malkus & Veronis also predict the transverse ( z direction) wavelength for
incipient instability. I n the best case, again using pinene, the predicted
wavelength is twice the measured wavelength. This discrepancy is also
attributable to effects from the transverse boundaries.

INSTABILITY OF SURFACE WAVES


In the electroconvective forms of instability so far reviewed, the geom­
etry of the interface remains essentially unaltered by the motion. The
effects of interfacial shears on the vertical motions of an interface are still
another avenue of research-one that can only be touched on here. We wish
to emphasize the close tie between the physical processes discussed in the
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

previous sections and the dynamics of gravity-capillary waves coupled to


electric fields u nder circu mstances where effects of finite relaxation time are
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

Overstability of the rotor in a de field.-With the vertical interfacial mo­


important.

tions and their attendant gravity and capillary forces, new modes of poten­
tial-energy storage are added. Again , the rotor of Figure 9 is a convenient
vehicle for establishing insights into the dynamical phenomena that can be
expected.
By way of simulating the effects of capillary and gravity forces, which
depend on the geometry of the interface and not its velocity, we consider the
case in which the rotor is not only subject to viscous and electrical torques,
but to a spring torque as well. This torque, unlike others we have con­
sidered, depends on the absolute angular position of the rotor, (J(t) . An
analysis of the dynamics about a static equilibrium follows the same lines as

that if the relaxation time T = (E.+Eb) /(Ua+Ub) is short compared to dynami­


outlined in the section on steady rotations. For present purposes, we note

47, in the limit where w--+O and fl = d(J/dt is small. We define I and K as the
cal times of interest we are j ustified in using the torque expression, Equation

moment of inertia and the torsional spring constant of a u nit length of the
rotor and write the torque equation of motion :

- + 47rb2l'a 1 -
I
rJ28 [ H.2(l - RS)] - + KfJ
dO
0 57.
(1 + R)2 dt =

dt2
Here we have assumed that motions are slow enough to justify use of the
steady-state viscous-torque expression TV = - 411'JoIaflb2• Thus, the motions
are those of a torsional pendulum having a damping constant that is positive
or negative depending on the magnitude of He. In fact, Equation 52 now
discriminates between a negative and positive damping coefficien t ; between
damping and overstability. Overstabilities of the rotor constrained by a
torsional spring are found to be in reasonable agreement with this simple
model, at least in a restricted number of fluids (9) .
On the basis of the rotor dynamics, it is not surprising that electrical
shear forces can conspire to overstabilize a configuration which, like the
capillary-gravity wave system, behaves in an oscillatory or wave-like fashion
in the absence of the field.
1 44 MELCHER & TAYLOR

(0l';;>�EO:.�i<�
(bl y

""

FIG. 1 7 . Electrical shear stresses induced on interface of gravity-capillary wave system


lead to overstability.

Overstability of an interface.-Melcher & Schwarz (25) show that electrical


Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

shear forces can also lead to overstability on the interface of a gravity­


capillary wave system, depicted in Figure 1 7 . Here motions are in the plane
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

of the paper rather than in the plane of the interface. In general, this situa­
tion is extremely complex, but consider as a limiting case gas (b) over a
liquid (a) , as with the rotor, i n the limit where T = (Ea+ �b) / (O"a+O"b) is small
compared to dynamical times of i nterest. Then, if the liquid has a small
viscosity, in the sense that for surface perturbations of the form �= cos kx
exp st, s and k satisfy the relation s» k2f.LIp, it can be shown that the interface

58.
is overstable as
lI. = 1

where He is defined by Equation 52.


Overstabilities on an interface satisfying the above restrictions are ob­
served in an apparatus having essentially the geometry of Figure 16, except
that the equilibri u m interface is flat. For a typical experiment, hexane doped

m)-l. With an electrode spacing d=4.7 cm. spontaneous oscillations of the


with ethyl alcohol has properties 5=2 . 5 , J.La = 3.2 X 10-4, O"a=3 X I o-s (ohm­

interface are j ust perceptible with an applied voltage of 14 kv. The instabil­

to the applied electric field. The condition that He = 1 predicts that Eo for
ity appears as a standing wave with points of constant phase perpendicular

incipient instability is 2.2 X 105 V 1m, while the measured value is 3.0 X 105
V1m. I n view of the accuracy with which the electrical-conduction process
can be described, this agreement is as good as could be expected.
When the fluid properties and allowed wavelengths are not so circum­
scribed as outlined here, the condition for incipient instability is consider­
ably more complicated than simply He= 1. (25) Nevertheless, the basic
mechanism of overstability is a salient feature of the dynamics.

SUM MARY REMARKS


Our review is restricted to the dynamics of fluids having uniform electri­
cal properties-to cases where the electromechanical coupling is confined to
interfaces. Even more, the theme is interfacial shear effects. Confined as this
class of electrohydrodynamics may seem, it is clear that we have only begun
to form a picture of the dynamics that are possible when there is an influence
of the motion on the field, as well as an effect of the field on the fluid. This
ELECTROHYDRODYNAM I CS 145
review starts with cases where the former coupling can b e ignored, then in­
troduces the finite electric Reynolds nu mber to represent the effect of con­
vection on the charge distribution. We have used the H artmann nu mber as
an aid in recognizing the connection between various interfacial motions
attributable to the competing processes of electro convection and viscous
shear. Thus, we begin with a simple case that has an easily presented model
and end with cases that are somewhat more complicated.
In conclusion we should ask, where are the scientific and engineering
i mplications of this developing area? Our list of applications can only be
represen tative of the spectrum of interests ; the literature of electrohydro­
dynamics connected with each is larger than our own brief list of references.
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

Dielectrophoretic forces, represented by the second term in Equation 6,


are being applied to fluid mechanics problems ranging from the separation of
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

living and dead cells (26) to the orientation of cryogenic liq uid propellants in
the zero-gravity environment of space (2 7 ) . As pointed out in the introduc­
tion, in almost all cases this class of electrohydrodynamics is representable by
coupling at i nterfaces. Even though the effects of free charge are usually
undesirable in dielectrophoretic interactions, they must be understood. Thus,
in this area electrical surface-shear effects are essential to answer such ques­
tions as, can a cryogenic propellant be oriented indefinitely in a dc field ?
Two-phase heat transfer in an electric field (28), another area of e1ectro­
hydrodynamics, is dominated by electromechanical coupling at interfaces.
Shearing effects can be i m portant in determining i nterfacial stability or i n
providing a mechanism similar t o that o f blowing for shearing the fluid
from a surface.
I maging on liquid interfaces by means of charged particles is being devel­
oped and demands careful attention to the effects of electrical shear stresses
(29) . The formation of charged liquid particles, with a multitude of applica­
tions including i mage reproduction and space propulsion (30) , clearly in­
volves electrical relaxation and shear effects ( 3 1 , 3 2) . From the basic view­
point, there are a host of fluid systems in which the electric Reynolds number
is large, even at modest applied-field strengths. Thus, in studies of electrical
conduction, as well as other physical phenomena at interfaces, the electro­
hydrodynamic contributions to measurements can hardly be ignored. It is
now becoming obvious to many (33) that it makes no sense to study mecha­
nisms of current conduction in slightly conducting liquids without paying
heed to the electrohydrodynamics.
Finally, let us recognize that an interface is a particular kind of dis­
continuity in the electrical and mechanical properties of a fluid. Studies of
surface dynamics have clear implications for coupling in the bulk of fluids.
Consider by analogy the connection between gravity surface waves repre­
sented by a discontinuity in density and the internal dynamics of fluids hav­
ing distribu ted density gradients. Many of the types of interactions we have
reviewed are being studied in the context of electrohydrodynamic bulk
interactions (34) .
146 MELCHER & TAYLOR

L I TERATURE CITED

1. Pickard, W. F., in Progress in Di­ Illinois, Urbana, Illinois, Dept. of


electrics, 1-39 (Academic Press, Elec. Engr., 1965)
New York, 334 pp., 1965) 18. Arno, R., Rendi. Alii Reale A ccad.
2. Woodson, H. H., Melcher, J . R., in Lineci, 1, (2), 284 ( 1 892)
Electromechanical Dynamics : Part 19. Pickard, W. F., Nuovo Cimento, 2 1 ,
I, Discrete Systems, 2 5 1 -3 1 7 (John 3 1 6-32 (1961)
Wiley & Sons, Inc. New York, 20. Melcher, J. R., Phys. Fluids, 9, 1 548-55
N. Y., 894 pp., 1968) ( 1966)
3. Watson, P. K., Charbaugh, A. H., in 2 1 . Ochs, H. T. , Traveling-Wave Electro­
Progress in Dielectrics, Vol. IV, hydrodynamic Pumping (M.S.
201-46, (Birks, J. B . , Hart, J., Thesis, Mass. Inst. Tech., Cam­
Eds., Academic Press, Inc., New bridge, Mass., Dept. Elec. Engr.,
Annu. Rev. Fluid Mech. 1969.1:111-146. Downloaded from www.annualreviews.org

York, 309 pp., 1 9 62) 1967)


4. Stratton, J. A., Electromagnetic Theory, 22. Quincke, G., A nn . Phys. Chemie, 59,
Access provided by 103.147.138.252 on 07/15/21. For personal use only.

1 3 7-40 (McGraw-Hill, New York, 4 1 7-85 ( 1 896)


N. Y. , 6 1 5 pp. 1941) 23. Jolly, D. C., Cellular ElectroconvectilJe
5 . Penfield, P. A . , Jr., Haus, H . A., Instability in a Fluid Layer (M.S.
Electrodynamics of Moving Media, Thesis, Mass. Inst. Tech. , Cam­
65-72 (M.1. T. Press, Cambridge, bridge, Mass., Dept. Elec. Engr.,
Mass., 2 7 6 pp., 1967) 1968)
6. Private communication, G. I . Taylor 24. Malkus, W. V. R., Veronis, G., Phys.
to J. R. Melcher, Nov. 23, 1966 Fluids, 4, 13-23 (1961)
7. Stuetzer, O. M., Phys. Fluids, 5, 534-44 25. Melcher, J. R., Schwarz, W. J., Jr.,
(1962) Interfacial Relaxation Overstability
8. Smith, C. V., Jr., Melcher, J. R., in a Tangential Electric Field
Phys. Fluids, 10, 2 3 1 5-22 (1967) (Mass. Inst. Tech. , Center Space
9. Smith, C. V., Jr., Steady Shear-Induced Res. Rept. CSR TR 68-2, Cam­
Electrohydrodynamic Flows (Doc­ bridge, Mass., 1968)
toral Thesis, Mass. Inst. Tech., 26. Crane, J. S., Pohl, H. A., J. Electro­
Cambridge, Mass., Dept. Elec. chem. Soc., 1 1 5, 584-86 (1968)
Engr., 1968) 27. Melcher, J. R. H urwitz, M., J. Space­
10. Sneddon, I. N., Fourier Transforms, craft &> Rockets, 4, 864-81 ( 1 967)
269-270 (McGraw-Hill, New York, 28. Choi, H. Y., Trans. A SME J. Heat
542 pp., 1 9 5 1 ) Transfer, 90, 98-102 ( 1 968)
11. Taylor. G . I . , P,oc. Royal Soc. A, 291, 29. Poritsky, H., in Developments in Me­

1 59-66 ( 1966) chanics, 2, Part I, 145-72 (Ostrach,

1 2 . Fano, R. M., Chu, L. J . , Adler, R. B.,


S., Scanlan, R. H., Eds., Pergamon

Electromagnetic Fields, Energy, and


Press, 830 pp., 1965)
30. Hendricks, C. D., J. Colloid Sci., 1 7,
Forces, 1 50-53 (John Wiley & Sons,
249-59 (1962)
Inc., New York, 520 pp., 1960)
3 1 . Carson, R. S., Hendricks, C. D.,
1 3. Goldstein, S., Modern Developments in
Natural Pulsations in Electrical
Fluid Dynamics, 1 1 4-1 1 5 (Oxford
Spraying of Liquids (Paper 64-
University Press, Oxford, 702 pp.
675, Am. Inst. Aeron. Astronaut.
1938)
Fourth Elec. Propulsion Conf.,
14. Bird, R. B., Stewart, W. E. Lightfoot,
Philadelphia, Penn., 1 964)
E. N . , Transport Phenomena, 90
32. Hogan, J. J., Carson, R. S., Schneider,
(John Wiley & Sons, Inc., New
J. M., Hendricks, C. D., AIAA J.,
York, 780 pp., 1 960)
4, 1460-1461 (1964)
1 5 . Allan, R. S., Mason, S. G., Proc. Roy. 33. Lewis, T. j., Seeker, P. E., 1966 A nnual
Soc., A, 267, 383-97 (1964) Report : Conference on Electrical In­
16. Taylor, G. I. Pro'. Roy. Soc., A, 280, sulation and Dielectric Phenomena,
383-97 ( 1964) 87-94 (Nat. Acad. Sci., Washing­
1 7. Sample, S. B., Static and DynamiC Be­ ton, D. C., 199 pp., 1968)
havior of Liquid Drops in Electric 34. Melcher, J. R., Firebaugh, M. S.,
Fields (Doctoral Thesis, Univ. of Phys. Fluids, 10, 1 1 78-85 (1967)

You might also like