Sei sulla pagina 1di 27

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

Annu. Rev. Mater. Sci. 1999. 29:32752


c 1999 by Annual Reviews. All rights reserved
Copyright

ELECTROPHORETIC DEPOSITION
OF MATERIALS
Omer O. Van der Biest and Luc J. Vandeperre
Departement Metaalkunde en Toegepaste Materiaalkunde, Katholieke Universiteit
Leuven, De Croylaan 2, 3001 Heverlee, Belgium;
e-mail: Omer.Vanderbiest@mtm.kuleuven.ac.be
KEY WORDS:

mechanism, kinetics, applications, suspensions

ABSTRACT
The electrophoretic deposition of materials is reviewed. Numerous applications
of electrophoretic deposition are described, including production of coatings,
free-standing objects, and laminated or graded materials, infiltration of porous
materials, and fabrication of woven fiber preforms. The preparation of electrophoretic suspensions is discussed as are a number of mechanisms of deposition that have been proposed elsewhere. In discussing the kinetics of the process,
primary attention is given to the relation between the evolution of the current and
the electric field strength.

INTRODUCTION
Electrophoretic deposition (EPD) is essentially a two-step process. In the first
step, particles suspended in a liquid are forced to move toward an electrode
by applying an electric field to the suspension (electrophoresis). In the second
step, the particles collect at one of the electrode and form a coherent deposit
on it. It should be noted that the process yields only a powder compact, and
therefore electrophoretic deposition should be followed by a densification step
such as sintering or curing in order to obtain a fully dense material.
The particles in suspension will move only in response to the electric field if
they carry a charge. Four mechanisms have been identified by which the charge
on the particles can develop (1): (a) selective adsorption of ions onto the solid
particle from the liquid, (b) dissociation of ions from the solid phase into the
liquid, (c) adsorption or orientation of dipolar molecules at the particle surface,
327
0084-6600/99/0801-0327$08.00

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

328

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

and (d) electron transfer between the solid and liquid phase due to differences
in work function.
A charged particle in a suspension is surrounded by ions with an opposite
charge in a concentration higher than the bulk concentration of these ions; this
is the so-called double-layer (Figure 1). When an electric field is applied, these
ions and the particle should move in opposite directions. However, the ions are
also attracted by the particle, and as a result, a fraction of the ions surrounding
the particle will not move in the opposite direction but move along with the

Figure 1 Schematic of the double layer surrounding a charged particle and evolution of the electric
potential from the surface potential, 0 , to zero far from the particle. The potential at the surface
of shear, the limit between the liquid moving with the particle and the liquid, which does not
move with the particle, is termed the zeta-potential, , and is the main parameter determining the
electrokinetic behaviour of the particle.

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION

329

particle. Hence, the speed of a particle is not determined by the surface charge
but by the net charge enclosed in the liquid sphere, which moves along with
the particle. The potential at the surface of shear is termed the zeta-potential or
electrokinetic potential. In principle, a particle with a negative surface charge
can show a positive zeta-potential. The latter occurs, for example, when the
charge of specifically adsorbed ions is higher than the surface charge.
In fact, the equilibrium speed of the particle is determined by four forces acting on the particle. The first, which accelerates the particle, is the force caused
by the interaction of the surface charge with the electric field. All other forces
slow the particle: These are viscous drag from the liquid following Stokes law,
the force exerted by the electric field on the counter-ions in the double layer (retardation) and, when a particle moves, the distortion in the double layer caused
by a displacement between the center of the negative and positive charge, (relaxation). A more complete treatment of these forces, along with information
on calculations of their magnitudes, can be found, for example, in Reference 2.
In this paper, we first present a number of applications of electrophoretic
deposition. Next, the preparation of suspensions for electrophoretic deposition is discussed, followed by a synthesis of the available information on the
mechanisms of electrophoretic deposition and the kinetics of the process.

APPLICATIONS
Electrophoretic deposition can be applied to any solid that is available in the
form of a fine powder (< 30 m) or a colloidal suspension. Examples of electrophoretic deposition of materials of almost any material class can be found,
including metals, polymers, carbides, oxides, nitrides, and glasses (see Table 1).
Furthermore, the process can be used for producing coatings, for shaping monolithic, laminated and graded free-standing objects, and for infiltration of porous
materials and woven fiber preforms for composite production.

Coatings
Electrophoretic deposition of coatings has already gained a world-wide acceptance for automotive, appliance, and general industrial (organic) coatings (3).
For example, the construction and operation of an 800 bodies-per-hour cataphoretic coating plant by a German car builder was described recently (4).
The high level of automation, low levels of pollution, high throw power and
homogeneity of the coating are the advantages that have led to the use of electrophoretic deposition (3). Other advantages are the improved adherence of
electrophoretically deposited coatings and their higher density compared with
dipped or sprayed coatings (5).

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

330

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

Table 1 A non-exhaustive list of examples from a broad range of material classes that have been
electrophoretically depositeda
Material class

Examples

Acids and hydroxides

Calcium hydroxide (91), magnesium hydroxide (91), antimonic acid


(96), boric acid (52)

Borides

LaB6 (97)

Carbides

B4C (36), C (3436, 98), Cr3C2 (99), NpC (88), diamond (100, 101),
PuC (88), ThC (88), UC (88), SiC (3436, 57, 98, 102), UWC2 (88),
WC (97)

Carbonates

(87, 91, 97, 103)

Metals

Al (83, 97, 104), Al-Cr (65), Al-Si (65), Al-Ti (65), Au (88, 105),
B (88), Cu (83), Dy (88), Fe (83), Mo (88, 97), MoSi2 (51), Nb (88),
Nb2Sn (88), Ni (83, 97), Sn (83), Re (88), Ru (97), W (88, 97),
Zn (83, 97), Zr (88)

Nitrides

Si3N4 (56), AlN (37)

Organic materials

Starch (47), styrene-acrylic copolymer (106), latex (3, 85), vinyl


copolymers (3), epoxy resins (3), polyamides (3), poly-urethanes (3)

Oxides

Clay (25, 26), Al2O3 (30, 31, 48, 87, 97, 98), -Al2O3 (39, 64), BaTiO3
(107), Cr2O3 (108), Fe2O3 (108), glass (17, 109), In2O3 (97), La2O3
(51), LiAlO2 (110), mica (111), MgO (97, 112), NiO (88, 108),
ReBaCuO (24), SiO2 (53, 58, 59, 97), TiO2 (97, 108), UO2 (88),
YBaCuO (1820), ZnO (97), ZrO2 (41, 44, 48, 51, 54, 85)

Phosphors

(1416, 81, 97, 113, 114)

Numbers in parentheses are references

Throw power is defined as the ability to cover recessed portions of complexly


shaped parts. This property, combined with improved thickness homogeneity,
is a decided asset for application in the enamel industry (611). The claim that
relatively complex shapes can be coated is clearly shown by some examples
given by Ortner (12): These include taps coated with carbides, fasteners coated
with Ni, and metal cones coated with a ceramic glaze. Platinum grids have
been coated completely with ferrite (13).
Other examples of coatings made by electrophoretic deposition include the
deposition of phosphor coatings in the manufacture of screens for cathode ray
tubes (CRTs) for advanced display applications (1416), deposition of insulating glass for electronic applications (17), and superconductive coatings (1823).
For the latter, an automated process has enabled the coating of wires of kilometer length (24). By performing electrophoretic deposition in a magnetic field,
some authors were able to create a desirable texture for superconductance in
the c oatings (20, 21).

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION

331

Shaping Free-Standing Objects


For traditional ceramics, such as sanitary ware, the main advantage of electrophoretic deposition lies in its higher speed and in the low wear
of the moulds compared with slip casting (25). Just as complexly shaped objects can be coated using EPD, so too can it be used for shaping objects. Tiles,
closed and open end tubes, hemispheres, tubes with changes in diameter, and
conical sections are some of the shapes that have been made, albeit mainly on a
laboratory scale, using solid compositions normally used for table and sanitary
ware (2528).
An example of a hybrid process, somewhere between tape casting and electrophoretic deposition, is the ELEPHANT process (29). Continuous tapes of
ceramic material are made by depositing on two rolling cylinders, which then
press the two slabs together in one long tape. The process was developed
for the tile industry. After electrophoretic deposition, the tape is cut, punched,
dried, and sintered. The economic advantages stem mainly from the lower manpower requirements, low-energy consumption, low wear (of moulds) and low
maintenance costs (29).
Technical ceramics such as alumina (Al2O3) (3033), silicon carbide (SiC)
(3436), and aluminium nitride (AlN) (37, 38) can also be shaped by electrophoretic deposition.
Beta-alumina tubes, used as electrolyte in sodium sulfur batteries, are a classical example of the use of electrophoretic deposition. Whereas charging for
many ceramics is related to adsorption or desorption of hydrogen ions, the
preferential dissolution of sodium ions from the beta-alumina during milling
is the main factor leading to charging of such aluminas (39, 40). When these
beta-alumina tubes were studied, it was discovered that the difference in density between two or more powders used to produce a material was unimportant,
and a homogeneous composition could be ensured. Powers (40) showed this
by comparing the composition of the first and the seventh tubes deposited
from a suspension containing a beta-alumina powder with a low soda content (8% Na2O) with one with a higher soda content (1425%) also containing
various additives such as MgO and Li2O.
MONOLITHS

Layered materials can also be produced via EPD.


When the desired thickness of the first layer is reached, the deposition electrode can be moved to a second suspension for deposition of a layer of different
composition. By changing back and forth, a layered material is readily obtained. Nicholson et al (41) have produced ZrO2/Al2O3 laminates with alumina
layers as thin as 12 m and zirconia layers of 2 m from ethanol-based suspensions. The boundaries between the layers were straight and well defined,
LAMINATED MATERIALS

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

332

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

showing that although EPD can be a fast process, good control of the growth of
the layers can be obtained. The same group also produced alumina/lanthanum
aluminate laminates (42). Ferrari et al (43) and Fischer et al (44) produced alumina/zirconia laminates from aqueous suspensions. Vandeperre & Van der Biest
made a range of SiC-based laminates with graphite (45), SiC+graphite (46),
and porous SiC interlayers (47). They also produced laminated SiC/graphite
composite tubes (35), thus combining the ability of EPD to produce laminated
materials with the shaping capabilities of the process.
GRADED MATERIALS While layered materials are obtained by immersing the
deposition electrode in different baths, graded materials can also be made by
gradually changing the composition of the suspension from which EPD is carried out. Sarkar et al demonstrated the ability to form graded materials by slowly
adding an ethanol-based suspension of an alumina powder to an ethanol-based
suspension of an yttria-stabilized zirconia powder during deposition (48). A
gradual increase in the alumina content of the deposit was observed. Later,
the same group also produced Al2O3/MoSi2- and Al2O3/Ni-graded materials
(49). Chao et al produced alumina/ceria-stabilized zirconia-graded rods from
an acetone-based suspension (50).
UNIQUE MICROSTRUCTURES AND SELECTED DEPOSITION The full potential of
EPD in producing unique microstructural features is perhaps not yet realized.
Nicholson et al (51) did some exploratory work in making nonplanar laminates:
By placing a grid before the deposition electrode, they were able to produce
laminates with wavy interlayers. Thus by using auxiliary electrodes, deposition
can be enhanced locally. Scala & Sandor (52) have deposited B2O3 on silicon
wafers as a boron diffusion source. The deposition onto the entire silicon wafer
was prevented by growing a silica film on the wafer and etching the silica away
only where deposition of boron was required.

Infiltration
The throwing power of electrophoretic deposition can also be used to infiltrate
objects with a matrix material or to apply an internal coating. Gal-Or et al (53)
infiltrated porous graphite electrodes with silicon carbide and silicon oxide
particles. For a good infiltration, the particle buildup in an external coating
should be decreased so that the pores are not blocked. Lower applied voltages
and a high-particle concentration in the suspension improved the infiltration.
Ishihara et al (54) applied an internal coating of yttria-stabilized zirconia to
a porous Ni-CaO stabilized ZrO2 cermet containing 40 wt% Ni for a solid
oxide fuel cell. By repetitive infiltration and sintering, a dense film of uniform
thickness could be applied.

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION

333

Another application where electrophoretic deposition is gaining increasing


interest is the infiltration of fiber preforms with matrix material for composite production (5559). For conducting fibers, the preform is used directly
as the deposition electrode, whereas for nonconducting fibers, the preform is
positioned in front of an electrode. The advantage of using electrophoretic deposition for composite fabrication is that it allows a reduction of the costs compared with, for example, chemical vapor infiltration. Moreover, electrophoretic
deposition is much faster, and the composition of the matrix material can be
controlled quite easily, as is shown by the example of infiltration of woven SiC
fiber mats by a mixed sol of mullite composition (58).

GUIDELINES FOR SUSPENSION PREPARATION


Suspensions used in other colloidal-shaping techniques, such as slip casting
and tape casting, can serve as a starting point for development of suspensions
for electrophoretic deposition. However, in contrast to many of these colloidal
processes, suspensions with much lower solids loading can be used for EPD. The
resulting lower viscosity is an advantage in handling the suspensions. Green
densities of about 4060 vol% are obtained via EPD from suspensions with
solids loadings as low as 12 vol% (34).
The key to successful development of suspensions for electrophoretic deposition is to find a systematic approach to making suspensions in which the
particles have a high zeta-potential, while keeping the ionic conductivity of
the suspensions low. A necessary but not sufficient condition for a high zetapotential is a high surface charge. As mentioned above, four mechanisms have
been identified for charging (1): selective adsorption of ions onto the solid
particle from the liquid, dissociation of ions from the solid phase into the liquid, adsorption or orientation of dipolar molecules at the particle surface, and
electron transfer between the solid and liquid phase due to differences in work
function.
For most oxides in the presence of water, the charge will be determined
mainly by adsorption or desorption of protons. A comprehensive model to
describe the charging of oxides in water has been developed by James et al
(60). They describe the charging as chemical reactions of surface groups, which
can be acid, alkaline, or amphoteric in nature. The value of the pH, where the
concentration of negatively and positively charged surface groups is equal, is
termed the point of zero charge (pzc) and is a property of the chemical groups
on the surface of the powder. The pzc should not be confused with the isoelectric point (iep), which is the pH at which the powder does not move when
an electric field is applied in a given suspension. The pzc and iep can differ

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

334

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

when, for example, multivalent metal ions, which adsorb specifically on the
surface, are added to the suspension (61).
Because of the lower solids loading, simply changing the pH to a value far
from the pzc using concentrated acids or bases is generally all that is required
for obtaining a stable suspension in water. Naturally, the ionic concentration
must be kept low as the stability decreases with increasing ionic concentration
(Figure 2), and the pH range must be adapted so that the oxide does not dissolve.
For example, yttria dissolves readily in an acid aqueous environment.
However, the use of water-based suspensions causes a number of problems in
electrophoretic forming (1). Electrolysis of water occurs at low voltages (5 V),
and therefore gas evolution at the electrodes is inevitable at field strengths high
enough to give reasonably short deposition times. If electrophoresis is used to
form objects, the inclusion of gas bubbles in the deposit can be prevented by
depositing on a porous membrane placed before the electrode. However, current
densities are higher compared with non-aqueous media, which leads to Joule
heating of the suspension and sometimes loss of stability of the suspension.

Figure 2 Interaction energy (attraction is negative) between two particles surrounded by a double layer for a surface potential of 25 mV, a particle radius of 0.35 m, a Hamaker constant of
1020 J, and a relative dielectric constant of 20, for various concentrations of a background electrolyte (indicated in the figure). With increasing electrolyte concentration, the height of the repulsion barrier due to double layer overlap decreases, and the repulsion vanishes completely at even
higher electrolyte concentration.

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION

335

Organic liquids generally have a lower dielectric constant, which is a disadvantage because it limits the charge on the particles: The dielectric constant
expresses the dissociating power of the solvent. However, much higher field
strengths (1001000 V/cm) can be used because the problem of electrolytic
gas evolution and Joule heating are greatly reduced. Therefore, organic liquids are generally preferred over water for electrophoretic deposition, although
papers describing successful electrophoretic deposition from aqueous suspensions continue to be published (25, 43, 44, 62). The fact that water is much less
expensive and less harmful for the environment is a driving force toward its
use.
Within the range of non-aqueous media, alcohols are similar to water, i.e.
both can be regarded as neutral amphiprotic solvents. The hydrogen ion concentration continues to be a measure of the acidity of the medium (63). Therefore,
the acid/base chemistry upon changing from water to alcohols does not change
drastically, and consequently many of the principles explained above remain
valid. A stable suspension can be obtained by adjusting the pH value to the stable range. A recent paper by Wang et al (63) describes how the electrophoretic
mobility of alumina in ethanol goes from positive at low pH to negative at
high pH, completely analogous to the behavior in water. They also showed that
electrostatic stabilization in ethanol is possible.
The electrical conductivity rises steeply with the dielectric constant (64),
making suspensions in methanol and ethanol relatively susceptible to ohmic
heating. Therefore, the current should be limited either by working at constantapplied current or by limiting the voltage that is applied.
Ketones, ethers, and hydrocarbons, on the other hand, are known as aprotic solvents. Their dielectric constant and electrical conductivity are generally
lower. Ketones have a dielectric constant close to 15, whereas for ethers and
hydrocarbons the dielectric constant is lower than 10. An empirical study on
electrophoretic deposition of beta-alumina from a series of organic solvents
showed that a dielectric constant within the range of 12 to 25 was required for
electrophoretic deposition (64). The dissociation power of solvents with a lower
dielectric constant is probably not enough to allow for a reasonable charge to
form on the particles, whereas deposits from solvents with dielectric constants
higher than 25 were found experimentally to be more fluid and slide off the
electrodes more easily (64). An explanation was not provided, and it should
be noted that electrophoretic deposits also have been made from suspensions
in water (43, 44, 65) (dielectric constant of water is 78.5) (66).
Labib & Williams (67, 68) found that electron transfer from the solid to
the liquid phase or vice versa can indeed determine the charge of the powder
in non-aqueous media. They measured the zeta-potential of a number of dried
oxides in a series of dried solvents with different values for the electron donicity

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

336

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

(69). Electron donicity is defined as a measure for the tendency of a solvent to


donate electrons. It is determined by measuring the enthalpy of reaction of the
solvent with a reference acid (SbCl5) in 1,2-dichloroethane. The usefulness of
the electron donicity for charging of solids can be illustrated using the example
of the degree of ionic dissociation of a solute in a series of donor solvents.
The ionization process can be broken up into two steps. The first involves a
base displacement whereby the solvent displaces the basic or anionic portion
of the solute, giving a solvated ion pair:
D : + A : B (D : A)+ (: B) .

1.

The second step involves the separation of the ion pair to give a free solvated
cation and anion. The first step is a function of the electron donicity of the
solvent. The second step is a function of the solvents local dielectric constant,
and the resulting degree of dissociation, as measured by the solutions conductivity, is a function of both. If A:B is thought to be a surface group of
the powder, it is clear that charging of the surface through electron exchange
should be possible. Labib & Williams (67, 68) proposed to attribute a donicity
value to the surface of solids (DS). If the solvent has a higher electron donicity
than the solid, electrons will be transferred from the solvent to the solid, resulting in a negatively charged particle. If the electron donicity of the solvent is
lower than the electron donicity of the solid, the particle will acquire a positive
charge. The values of the zeta-potential of dried SiO2, TiO2, Al2O3, and MgO
in a series of dried solvents with increasing electron donicity were consistent
with this view. The sign of the zeta-potential in these solvents changed at some
point within the donicity scale. The interpolated value, where the sign of the
zeta-potential changed, was assumed to correspond to the electron donicity of
the solid. The ranking of the solids based on their electron donicity values
compared very well with the ranking of the iep in water, i.e. from basic to acid
the ranking was MgO, Al2O3, TiO2, and SiO2. When a limited amount of water
was introduced in the system by equilibrating the solids in an environment with
50% relative humidity, an increase in the donicity of the surfaces was observed.
This was attributed to the fact that adsorption of water molecules on the surface makes Lewis acids much weaker and in some cases changes their identity.
This last result indicates that if water is present, exchange of protons becomes
more important compared with electron exchange as a charging mechanism in
non-aqueous solvents.
Okuyama et al (70) dispersed a series of silicon carbide powders in a wide
range of non-aqueous media. They found that the powder dispersability, as determined from its sedimentation behavior, did not correlate well with dielectric
constant, pKa, hydrogen bond index, and Lewis acid/base interaction. However, in work by Bolger (71) on oxide powders, a good correlation was found

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION

337

between powder dispersability and the difference of the iep of the powder in
water and the pKa or the pKb value of the solvent.
Given the importance of the pH in determining the charging of powders in
water, it was only logical that parameters describing the acidity/alkalinity of
non-aqueous media were used in the first attempts to understand charging in
non-aqueous media. However, one important difference has not attracted much
attention: The charge on a powder in aqueous suspensions is not determined
by the pH of the water before the suspension is prepared, but by the pH of the
suspension once prepared. This simple fact is not stressed when dealing with
aqueous suspensions because the pH is regarded as a tuneable experimental
parameter. Indeed, only limited amounts of concentrated acids or bases are
required in practice to tune the pH. Hence, changes in pH upon addition of a
powder to water are given almost no attention. That the pH can change substantially and unpredictably when a powder is added to water becomes apparent if
the natural pH values of several powders are compared. The term natural pH
of a powder is used in this review to describe the pH at which the suspension
equilibrates when a powder is suspended in water. For example, in two batches
of a nominally identical SiC powder, one had a natural pH of 3.2 at 20 g L1,
whereas the other had a natural pH of 8.2 at 20 g L1. A more systematic investigation showed that simply washing powders in an acid or alkaline aqueous
suspension for a short time suffices to adsorb acids or bases onto the powder,
which results in a remarkable difference in charging and deposition behavior in
acetone, ethanol, and iso-propyl alcohol (72, 73). As a result, the charge on the
powder when added to a suspension medium will not only be correlated with
the difference in pzc and the acidity of the medium, but also with the amount of
acids and bases adsorbed on the powder as expressed by the natural pH. Even
for a clean powder surface, the natural pH of a suspension tends toward the
pzc of the powder (72). Figure 3 illustrates these two effects for a number of
commercially available powders (74).
If it is likely that ions will leach from the solid to be deposited, as e.g. Na+
ions from beta-alumina (75), or barium ions from barium titanate powders (76),
the design of the suspension must take into account that the ionic conductivity
will increase, that the solid might be charged through the dissolution of these
ions, and that re-adsorption of these ions can also influence the zeta-potential.

MECHANISM OF DEPOSITION
Although electrophoretic deposition is an old process, the exact mechanisms
that allow a deposit to be formed are still not entirely clear. According to
Hamaker & Verwey (77), the formation of a deposit by electrophoresis is akin to
the formation of a sediment due to gravitation. The pressure exerted by incoming

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

338

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

Figure 3 The natural pH of a range of commercially available powders, at a powder loading of


20 g L1, as a function of their point of zero charge in water. In general, the natural pH indeed
tends to be close to the point of zero charge. Deviations are observed but can be attributed to the
presence of adsorbed acids or bases.

particles enables particles next to the deposit to overcome the interparticle


repulsion.
Koelmans (78) calculated the expected increase in ionic strength next to an
electrode and found that the ionic strength was of the same order as required to
flocculate a suspension. In his view, the interparticle repulsion decreases and the
particles collapse in a deposit due to the increase in the electrolyte concentration.
Since a finite time is required for the electrolyte concentration to build up near
the electrode, one expects that deposition would not occur before a certain time
has passed. Indeed, although MgCO3 particles suspended in methanol were
found to collect on the electrode as soon as a potential was applied, only after
some critical time did these particles remain on the electrode after the voltage
was cut off (78). The magnitude of the critical time is inversely proportional
to the applied voltage to the second power (78). Hence, the critical time can be
very short, which explains why it is frequently not observed.
Grillon suggested that particles would neutralize upon contact with the deposition electrode or deposit (79). Sarkar & Nicholson (80) note that the double

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION

339

layer of a particle in motion is distorted so that the double layer is thinner ahead
of the particle. Further, they argue that since counter ions move along with the
particle, their concentration in the deposit is much higher than normally would
be in the suspension. Thus they probably recombine with incoming co-ions to
reform their original salts, which results in a thinning of the double layers of the
particles in the deposit. Thus an incoming particle with a thinner double layer
meets with the particles in the deposit that also have a thinned double layer
and as a result the repulsion is decreased substantially, allowing the particles to
overcome this repulsion barrier.
A type of suspension probably best documented with respect to the deposition mechanism is one in which the powder is charged positively by adsorption
of metal ions on the powder. These metal ions are introduced in the suspension
through addition of salts such as Mg(NO3)2, La(NO3)3, Y(NO3)3, MgCl2, or
AlCl3. If a current is passed through solutions of these salts, the formation of
a hydroxide (16, 81, 82) is observed. If the water content of such solutions in
iso-propyl alcohol is limited to below approximately 5 vol%, the main electrochemical reaction product becomes an alkoxide for Mg(NO3)2 or Al(NO3)3
(81). Analysis of deposits made from such suspensions has revealed that these
hydroxides also form during deposition (14). Therefore, it is likely that binding by these hydroxides is the mechanism of deposition. Hydroxides provide a
better adhesion than alkoxides (81), and a limited amount of water is thus an advantage for obtaining well-adhering coatings. Consequently, the deposition of
the powder is induced by an electrochemical reaction at the electrode, and one
would expect that no deposition would occur away from an electrode. Indeed,
Brown & Salt (83) had already observed that an aluminium powder charged
through addition of AlCl3 to ethanol did not deposit on a porous membrane
placed away from the electrode, whereas a deposit could be obtained from this
suspension on the cathode.
However, in the absence of such salts, electrophoretic deposition can occur away from the electrode. For example, magnesium oxide suspended in
methanol was found to deposit on a porous membrane placed before the electrode (83), and aluminium oxide particles suspended in ethanol with addition
of an acid deposited on a dialysis membrane placed before a graphite cathode
(84). Thus although the mechanism of hydroxide formation is supported with
a substantial amount of experimental evidence, it clearly is not the only mechanism for formation of a deposit. When Koelmans (78) proposed his mechanism
of flocculation (discussed above), he had already noted that electrolysis could
lead to insoluble or non-ionized products and therefore not to the increase in
the electrolyte concentration near the electrode required for his mechanism.
However, he also argued that for oxides, protons and hydroxyl ions are potential determining ions. Therefore, if during electrolysis their concentration

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

340

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

changes near the electrode, the surface charge and the interparticle repulsion
will decrease. Thus the deposition could also be induced by a shift in pH at
the electrode. The fact that, in the absence of salts such as AlCl3, deposits can
be formed on membranes placed away from the electrode is not necessarily a
contradiction for such a mechanism. Sheppard (85) pointed out in his discussion of electrophoretic deposition of rubber on membranes that generally in
such experiments the side of the membrane toward the anode becomes more
alkaline, whereas the side toward the cathode becomes more acidic. The concomitant increase in hydrogen ions is, in his opinion, a considerable factor in
the deposition of rubbers on such membranes. Note that if deposition is caused
by a change in pH, a critical time will be required before the pH has changed
sufficiently. Pierce (3) discussed the cathodic deposition of resins quite similarly. He postulates that in order for a resin with amine groups to deposit, the
hydroxyl ion concentration at the electrode should reach a critical value related
to the solubility product for the ammonium hydroxide:

2.
RESIN NH+
3 [OH ] = K sp .
Further, in the absence of migration and chemical reactions, the concentration
of hydroxyl ions at the cathode will increase as the square root of time for an
unstirred electrolyte at constant current density. Therefore, the product of the
square root of the critical time and the current density is expected to be constant
for such systems, which is also found experimentally (3).
To summarize, there are many indications that in order for powder to be
incorporated in the deposit, the charge on the surface must be reduced. For
powders charged by adsorption or de-sorption of protons, a different pH in
the suspension, and near the deposition electrode or the membrane on which
deposition is performed, could reduce the surface charge. For powders charged
by adsorption of metal ions, an electrode reaction converting these adsorbed ions
into uncharged metal hydroxides is required. Hence, no deposition is observed
on membranes in such systems. These hydroxides improve the coherence of
the deposit and the adherence to the electrode. Therefore using salts to charge
powders is probably better for making coatings than for making free-standing
objects, which have to be released from the electrode. If other ways are used to
charge a powder, for example by adsorption of a poly-electrolyte (33), whether
deposition has occurred will be determined by the decharging reactions required
for the surface groups of these poly-electrolytes.
Although the decharging due to changes in pH as a consequence of the
current is very appealing, it should be noted that such changes normally do
not extend very far into the liquid. However, it is possible that the change in
the contribution of convection to the transport of ions to the electrode through
the growth of the deposit causes the changes in pH to extend farther into the

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION

341

liquid. Moreover, merely an increase in the concentration of the powder near


the deposit could cause a decrease in the equilibrium surface charge for powders
charged through proton adsorption/de-sorption. Vandeperre (72) has calculated
the natural pH of a suspension of oxide powders as a function of the surface site
concentration in the suspension and found that the pH shifted toward the point of
zero charge with increasing solids loading. Furthermore, the fact that the double
layer of incoming particles is thinner ahead of the particles, as pointed out by
Sarkar & Nicholson (80), can certainly also contribute to the incorporation in a
deposit.
Finally, if anything is clear, it is that almost any material can be laid down
by EPD. The fact that in many reports on electrophoretic deposition little attention is given to the deposition mechanism clearly shows that even if a full
understanding of the exact mechanism is lacking, electrophoretic deposition is
already being used successfully. However, a better understanding is needed to
decrease the amount of experimental work required to determine the optimal
composition of a suspension and the deposition parameters such as applied
voltage or current .

KINETICS OF DEPOSITION
Single-Powder Deposition
The kinetics of electrophoretic deposition has been the subject of numerous investigations. By simply applying the principle of conservation of mass,
Hamaker (77) proposed the following equation
dY
= f..E.S.c.
dt

3.

with Y, yield of deposition (kg); t, deposition time (s); , electrophoretic


mobility (m2 V1 s1); S, surface area of the electrode (m2); c, solids concentration (kg m3); and f a factor taking into account that not all powder
brought to the electrode by electrophoresis is also incorporated in the deposit
( f = 1 or less). This equation is now accepted as describing the basis for the
kinetics of electrophoretic deposition.
The surface area of the electrodes and the concentration of solids in a suspension can easily be determined or calculated during the process. The electrophoretic mobility can also be measured. Although a good qualitative notion
of the electric field strength can be obtained through measurement (86), the
electric field strength is almost always evaluated from the applied voltage and
the inter-electrode distance. However, such a calculation is subject to conditions. The potential drop over an electrophoretic deposition cell with flat equal

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

342

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

Figure 4 Evolution of the potential in an electrophoretic deposition cell: The applied potential is
consumed by a potential drop at each electrode and by an ohmic potential drop over the suspension
and the deposit.

surface area electrodes can in general be described as consisting of four terms


(Figure 4):
Va = 11 + I.Rd .d1 + I.Rs .(d d1 ) + 12 ,

4.

where Va is the applied potential (V); I is the current (A); Rd and Rs are the
resistance per unit length of the deposit and of the suspension, respectively
(Ohm/m); d is the electrode separation (m); d1 the thickness of the deposit
(m); and 1i is the potential drop at one electrode (V). The resistance per
unit length is defined as the specific resistance (Ohm/m) divided by the surface
area of the electrodes (m2). The conditions required in order for the electric
field strength in a suspension to be close to the applied potential divided by the

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION

343

inter-electrode distance are thus: The potential drops at each of the electrodes
should be negligible, and the specific resistance of the deposit should be of the
same order as the specific resistance of the suspension.
The validity of the approximation for the electric field strength is rendered
even more doubtful by the fact that at constant applied potential, the current
is frequently found to decrease during deposition. The following explanations
have been proposed: a change in polarization at the electrode (87, 88), a higher
specific resistance of the deposit compared with the suspension (17, 30, 57,
8992), and changes in the conductivity of the suspension (72, 93).
Starting at the electrode, one should realize that in order for a net current to
flow through the solid/liquid interface, a potential in excess of the equilibrium
potential is required (activation overvoltage). The relation between the excess
potential over the interface and the resulting net current is exponential (94).
Therefore, small variations in activation overvoltage (order of 1 V) result in
large changes in current. Hence, the activation overvoltage can be expected
to remain more or less constant during the process when depositing from nonaqueous suspensions, where the applied voltage is typically of the order of
1001000 V. A second effect to be taken into account is that according to
Nernsts law, the equilibrium potential over a solid/liquid interface depends on
the concentration of the reactants (94):

RT
aox
1 = 10 +
5.
ln
= 10 + c ,
zF
ared
where 10 is the standard equilibrium potential drop over the solid/liquid interface (V), 1 is the equilibrium potential drop over the solid/liquid interface
(V), aox is the activity of the species being oxidized, ared , is the activity of
the species being reduced, R is the gas constant (J mol1 K1), T is the
temperature (K), z is the valence of the reduction reaction, and F is Faradays
constant (C mol1). Changes in polarization due to a change in concentration
of the reactants for the electrode reaction are often referred to as concentration
overvoltage (c ). If during electrophoretic deposition the transport of reactants
for the electrode reaction toward the electrode is slower than the consumption
at the electrode, the concentration at the electrode will drop and the concentration overvoltage will increase. Note that changes in the proton or hydroxil
concentration were expected and observed in some of the investigations into
the mechanism of deposition (3). Moreover, the fact that a deposit forms on
one of the electrodes can be expected to hinder the transport of reactants toward
that electrode, thus enhancing the probability for an increase in concentration
overvoltage (see Figure 5). Cathodic deposits are especially prone to serious
changes in concentration overvoltage. Species capable of being reduced must
travel from the solution toward the cathode. At the anode, one of the electrode

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

344

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

Figure 5 Illustration of a possible influence of the growing deposit on polarization at the electrode.
The concentration of the reactant for the electrode reaction C as a function of the distance X from
the electrode is shown for four conditions: (a) Before any potential is applied, no concentration
gradients exist. (b) As soon as a current starts to flow, the concentration of the species reacting at the
electrode drops by a small amount, and the amount consumed at the electrode is replaced through
diffusion, convection and migration. (c) The presence of the deposit influences the transport of
reactants to the electrode. If this results in a decreased incoming flux, the concentration at the
electrode will start to drop. (d) Because the diffusion flux is determined by the slope of the
concentration gradient, the diffusion term in the flux decreases with increasing deposit thickness

reactions can be the dissolution of the electrode material. Naturally, in that case
the reactant concentration for the oxidation reaction will not drop. Some types
of electrodes, however, can also suffer from a serious increase of the potential drop at the electrode, even when they are used as an anode. For example,
aluminum, when polarized anodically, does not dissolve but forms a protective
oxide layer. Not only does this layer introduce an extra electrical resistance,
the oxidation of new aluminium becomes increasingly more difficult as oxygen
has to travel through the oxide layer. Hence, other oxidation reactions can be
required, and these can require the transport of species from the solution to the
anode. Note that even when the deposit forms on the anode, an increase in
cathodic concentration overvoltage can also occur. For example, during anodic
deposition of SiC from acetone, Vandeperre & Van der Biest (72) found that the
potential drop at the cathode increased during deposition, resulting in a drastic
increase of the total resistance of the deposition cell (Figure 6).

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION

345

Figure 6 (a) Evolution of the apparent resistance (applied potential divided by the measured
current, Rapp) during deposition of SiC from acetone and of the resistance of the suspension, Rsus,
as determined from a measured difference in potential drop between two reference electrodes
immersed in the cell and the measured current. Initially the magnitude of the current is controlled
by the resistance of the suspension, but after a few minutes, concentration polarization at the
counter electrode becomes very important, and the linear relation between amount of powder that
has deposited and the apparent resistance is lost. Initially the magnitude of the current is controlled
by the resistance of the suspension, but after a few minutes, concentration polarization at the
counter electrode becomes very important, and the linear relation between amount of powder that
has deposited and the apparent resistance is lost. (b) Evolution of the potential drop over the
suspension (Vsusp), at the counter electrode (Vcount) and at the deposition electrode (Vdep) for the
same deposition experiment

In regard to the conductivity of the suspension and the deposit, many authors
claim that the deposit has a much higher specific electrical resistance than does
the suspension. In this review the term resistance of the deposit is used only
to indicate the ohmic potential drop over the deposit as defined in Equation 4
and does not include possible effects of deposit formation on the concentration
overvoltage discussed above.
Most authors have simply inferred the higher specific resistance of the deposit
from the decrease in current (17, 30, 57, 8992). Sarkar et al (86) measured the
electric field strength by measuring the potential drop between two platinum
probes placed between the electrodes and also monitored the potential drop
between the deposition electrode and one platinum probe in the suspension.
During cathodic deposition of alumina from ethanol at constant applied current,
these authors found that the electric field strength remained constant and that the
potential drop between the deposition electrode and the probe in the suspension
increased at the same rate as the voltage required to keep the current constant
increased. Since both the required increase of applied voltage, as well as the

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

346

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

growth of the deposit, were linear with time, they concluded that this result
was proof for the higher specific resistance of the deposit. However, changes
in the concentration overvoltage at the cathode would have given the same
result. In fact, Sarkar et al also observed a transient effect when the voltage was
disconnected for 30 s and the experiment continued afterward: The increase
in resistance had disappeared while the deposit was still on the electrode. It
was only after about 30 s that the potential required for the constant applied
current reached the value it had reached before the experiment was interrupted.
Such a transient effect is more easily understood if the increase in potential
drop is caused by concentration overvoltage. During the time the experiment
is paused, the formed concentration gradients can be relieved by diffusion, and
the concentration overvoltage needs to be built up once again.
It should be noted that a higher specific resistance of the deposit compared
with the suspension is indeed expected. In a deposit, the volume fraction of
the liquid is lower than in the suspension, and therefore the effective area for
ionic conductance is lower (17). Pierce (3) found a high resistance for dense
layers of a cathodically applied polymeric primer. If the deposit was a porous
film, however, as sometimes observed during the electrodeposition of latex, the
voltage drop over the film remained very low due to the conductance through
the pores in the film (3). Since during deposition of ceramic or metal powders
the volume fraction of solids in the deposit is typically between 40 and 60
vol%, the resistance of the deposit cannot be many times higher than the ionic
resistance of the suspension.
The resistance of the suspension is also not necessarily constant. Vandeperre
& Van der Biest (93) modeled the conductivity of a suspension as consisting of a contribution of ions outside the double layers of the powder particles
and a contribution by the electric charges associated with the moving particles
(Figure 7). The decrease in current during electrophoretic deposition of SiC
from acetone with addition of N-butylamine was found to be mainly from a decrease in suspension conductivity (93) caused by depletion of the suspension.
Naturally, the decrease in conductivity of the suspension through this mechanism is observed only if the amount of powder deposited is large compared
with the total amount of powder in suspension and if the ionic conductivity is
relatively small. Moreover, another factor contributing to the lower conductivity of the deposit compared with the suspension is that while in the suspension
the movement of the particles contributes to the conductivity, in the deposit the
particles are immobile and therefore do not contribute to the conductivity of the
deposit. Figure 8 shows the evolution of the current and yield as a function of
time for deposition of alumina (Baikowski-type SM8, 100 g L1) from technical ethanol with addition of 1 M acetic acid (72), together with the calculated
current and yield based on the description of the conductivity of suspensions as

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION

347

Figure 7 Schematic diagram for the resistance of the deposition cell. In the suspension, moving
ions as well as moving particles contribute to the conductivity, which can be represented by two
resistors (Rpowder & Rion) in parallel. In the deposit, only ions are moving, hence one resistance,
Rdep, in series with the resistance of the suspension. Finally, there is also a potential drop, 1i , at
each electrode. The value of the resistance related to the contribution of the powder is a function
of the solid concentration.

being composed of a ionic component and a component related to the movement


of the powder. The expected yield of deposition, if the decrease in current was
due to changes in polarization or a very high specific resistance of the deposit,
is also indicated.

Co-Deposition of Two or More Powders


If two or more different powders are deposited simultaneously, and if the volume
fraction of solids in the suspension is high, the powders deposit at an equal rate.

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

348

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

Figure 8 Measured evolution of current (filled symbols) and yield (open symbols) during electrophoretic deposition of alumina from technical ethanol with 1 M acetic acid. The solid lines
were calculated based on the assumption that the decrease in current is caused by the vanishing
contribution of the alumina particles to the conductivity of the suspension, whereas the dotted lines
were calculated assuming that the current decrease was the result of very high specific resistance of
the deposit and no contribution of the powder to the conductivity. Although the calculated currents
coincide, the evolution of the yield is better described by the assumption that the conductivity of
the suspension decreases.

If, however, the volume fraction of solids is low, the particles can deposit at a
rate proportional to their individual electrophoretic mobility (37, 38, 47, 95). In
work by Moritz on deposition of aluminum nitride with incorporation of yttria
as a sintering additive, an equal deposition rate was observed from solids loading higher than 28 vol% (95). These observations are in accordance with the
findings by Massoud (25), who states that in slips for sanitary ware no segregation between the components was found. The suspension used in his work
contained a high volume fraction of solids.

CONCLUSIONS
EPD of any material available in the form of a fine powder or colloidal suspension is possible. The technique allows the application of coatings, the shaping
of objects, and the infiltration of porous structures and woven fiber preforms.
Moreover, laminated and graded materials can also be made. It is a fast and
economical process that can be automated for use on an industrial scale. The

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION

349

full potential of EPD in making controlled microstructures is perhaps not yet


known; e.g. by using special electrode configurations, deposition can be enhanced locally to make wavy multilayers. Suspension preparation for EPD,
whether for aqueous or non-aqueous media, can be based on acid/base chemistry. In water as well as non-aqueous media the charge can be controlled through
addition of an acid or a base. Simply selecting solvents based on acid or base
characteristics of the solvents will often not suffice because the effect of acids
or bases adsorbed on the powder surface can alter the acidity of the medium
drastically. Other charging mechanisms are through adsorption of metal ions,
such as Al3+ or Mg2+, or adsorption of poly-electrolytes.
In order to understand the kinetics of deposition, the electric field strength
must be known. A qualitative notion can be obtained through measurement,
and a number of mechanisms have been identified that can lead to a decrease
in electric field strength during deposition, i.e. changes in polarization at one
of the electrodes and a high electrical resistance of the deposit. Both effects
can be used advantageously if a self-limiting thickness is required and can be
avoided when thick deposits need to be produced.
A homogeneous composition of deposits of two or more powders can be
ensured, provided the volume fraction of solids in the suspension is sufficiently
high. In contrast to almost any other colloidal-forming technique, density differences between the powders do not result in segregation.
Most investigations regarding the mechanism of deposition propose that the
surface charge of the powder is reduced in the process. A difference in pH
near electrodes or membranes compared with the rest of the suspension is
believed to play a major role in this reduction of charge. It was suggested that
the increase in powder concentration in or near the deposit would also change
the equilibrium charge, and it has been pointed out that the double layer is
thinner ahead of a traveling particle, which results in a decreased repulsion
for incoming particles. Although the lack of a complete understanding of the
deposition mechanism has not prevented the use of the process on an industrial
scale, more detailed investigations, sustained by calculations, will be required
to control the process better and to decrease the amount of experimental work
required during development of a deposition system.
Visit the Annual Reviews home page at
http://www.AnnualReviews.org

Literature Cited
1. Heavens SN. 1990. In Advanced Ceramic
Processing and Technology, ed. JGP Binner, 7:25583. Park Ridge, NJ: Noyes

2. Hiemenz PC. 1986. Principles of Colloid and Surface Chemistry. New York:
Dekker. 815 pp. 2nd. ed.

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

350

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE

3. Pierce PE. 1981. J. Coat. Techn. 53(672):


5267
4. Eberhard J. 1997. GalvanoOrgano
66(678):62123
5. Lamb VA, Reid WE. 1960. Plating
(3):29196
6. Engelhardt Th, Rijenen P, Hellkuhl L,
Kaup F, Warnke H. 1990. Mitt. Ver. Dtsch.
Emailfachleute 38(4):3748
7. Engelhardt Th, Rijenen P, Hellkuhl L,
Kaup F, Warnke H. 1990. Mitt. Ver. Dtsch.
Emailfachleute 38(5):6074
8. Engelhardt Th, Rijenen P, Hellkuhl L,
Kaup F, Warnke H. 1990. Mitt. Ver. Dtsch.
Emailfachleute 38(6):8691
9. Engelhardt Th, Rijenen P, Hellkuhl L,
Kaup F, Warnke H. 1990. Mitt. Ver. Dtsch.
Emailfachleute 38(7):8691
10. Engelhardt Th, Rijenen P, Hellkuhl L,
Kaup F, Warnke H. 1990. Mitt. Ver. Dtsch.
Emailfachleute, 38(8):10516
11. Engelhardt Th, Rijenen P, Hellkuhl L,
Kaup F, Warnke H. 1990. Mitt. Ver. Dtsch.
Emailfachleute, 38(9):120124
12. Ortner M. 1964. Plating (9):88589
13. Barraclough M, Bolton NE, Collins AH,
Andrews JH. 1967. IEEE Trans. Magnet.
3(3):53134
14. Shane MJ, Talbot JB, Kinney BG, Sluzky
E, Hesse KR. 1994. J. Colloid Int. Sci.
165:33440
15. Shane MJ, Talbot JB, Schreiber RD, Ross
CL, Sluzky E, Hesse KR. 1994. J. Colloid
Int. Sci. 165:32533
16. Siracuse JJ, Talbot JB, Sluzky E, Hesse
KR. 1990. J. Electrochem. Soc. 137(1):
34648
17. Sussman A, Ward TJ. 1981. RCA Rev.
42(6):17897
18. Miziguchi J, Matsumura M, Suzuki M,
Yamato H. 1991. J. Electrochem. Soc.
138(10):294246
19. Nojima H, Koba M, Shintaku H, Nagata M. 1991. Jpn. J. Appl. Phys. 30(7):
L116668
20. Hein M, Peiniger M, Mahner E, Muller G,
Piel H, et al. 1989. Physica C 162:11112
21. Hein M, Peiniger M, Piel H, Ponto L,
Becks M, Klein U. 1989. J. Appl. Phys.
66(12):594043
22. Chu CT, Dunn D. 1989. Appl. Phys. Lett.
55(5):49294
23. Sharma AD, Sen A, Maiti HS. 1993. Ceram. Int. 19:6570
24. Woolf LD, Fagaly RL, Raggio WA, Elsner
FE, Fisher MV, et al. 1991. Appl. Phys.
Lett. 58(5):53436
25. Massoud E. 1979. Interceram 2:11719
26. Mihailescu M, Emandi M, Vancea V,
Marcu M. 1991. Interceram 40(3):165
70

27. Aveline M. 1966. Ind. Ceram. 581:28


31
28. Boncoeur M, Carpenter S. 1972. Ind. Ceram. 648:7981
29. Chronberg MS, Handle F. 1978. Interceram 1:3334
30. Andrews JM, Dracass J, Collins AH, Cornish DC. 1969. Proc. Brit. Ceram. Soc.
12:21129
31. Nass R, Schmidt H, Storch W, Harbach F,
Neeff R, Nienburg H. 1992. In Ceramic
Powder Processing Science, ed. H Hausner, GL Messing, S Hirano, pp. 62532.
Heidelberg: Bovery
32. Jean JH. 1995. Mater. Chem. Phys. 40:
28590
33. Ferrari B, Farinas JC, Moreno R. 1998.
Presented at World Ceramics Congress
and Forum on New Materials, Florence,
Italy
34. Vandeperre L, Van der Biest O, Bouyer F,
Persello J, Foissy A. 1997. J. Eur. Ceram.
Soc. 17(2):37376
35. Vandeperre LJ, Van der Biest O, Bouyer
F, Foissy A. 1998. Ceram. Bull. 77(1):53
58
36. Wittwer H, Kruger HG. 1995. Ber DKG
72(9):55660
37. Moritz K, Reetz T. 1993. Ber DKG
70(7):34850
38. Moritz K, Reetz T. 1993. In Third Euroceramics, ed. P Duran, JF Fernandez,
1:42530. Castellon de la Plana, Spain:
Faenza Editrice
39. Kennedy JH, Foissy A. 1975. J. Electrochem. Soc. 122(4):48286
40. Powers RW. 1986. Ceram. Bull. 65(9):
127077
41. Nicholson PS, Sarkar P, Huang X. 1993.
J. Mater. Sci. 28:627478
42. Bissinger M, Prakash O, Sarkar P, Nicholson PS. 1994. Ceram. Sci. Eng. Proc.
15:108492
43. Ferrari B, SanchezHerencia AJ, Moreno
R. 1998. Mater. Lett. 35:37074
44. Fischer R, Fischer E, De Portu G, Roncari
E. 1995. J. Mater. Sci. Lett. 14:2527
45. Vandeperre L, Van der Biest O, Clegg
WJ. 1997. In Key Engineering Materials,
127:56774
46. Vandeperre L, Van der Biest O. 1998. Silicates Ind. 62(5):8791
47. Vandeperre L, Van der Biest O. 1998. SiC
laminates with porous interlayers produced by EPD. Presented at 100th Ann.
Meet. Am. Ceram. Soc. Cincinnati, OH
48. Sarkar P, Nicholson PS, Huang X. 1993.
J. Am. Ceram. Soc. 76(4):105556
49. Sarkar P, Datta S, Nicholson PS. 1997.
Composites Part B 28:4956

P1: APR/SPD

May 15, 1999

P2: APR/ARY

14:52

QC: APR

Annual Reviews

AR086-11

ELECTROPHORETIC DEPOSITION
50. Zhao C. Vandeperre L, Vleugels J, Van
der Biest O. 1998. J. Mater. Sci. Lett.
17:145355
51. Nicholson PS, Sarkar P, Datta S. 1996.
Ceram. Bull. 75(11):4851
52. Scala LC, Sandor JE. 1968. Electrochem.
Technol. 6(1112):43437
53. Gal-Or L, Haber S, Liubovich S. 1992. J.
Electrochem. Soc. 139(4):107881
54. Ishihara T, Takita Y, Sato K. 1996. J. Am.
Ceram.Soc. 79(4):91319
55. Streckert HH, Norton KP, Katz JD, Freim
JO. 1997. J. Mater. Sci. 32:642933
56. Kooner S. 1995. Mater. Challenge (8):2729
57. Kawai C, Wakamatsu S. 1996. J. Mater.
Sci. 31:216570
58. Boccaccini AR, Ponton CB, MacLaren
I, Lewis MH. 1997. J. Eur. Ceram. Soc.
17:154550
59. Boccaccini AR, Trusty PA. 1998. J.
Mater. Sci. 33:93338
60. James RO, Davis JA, Leckie JO. 1978. J.
Colloid Int. Sci. 65(2):33144
61. Lyklema J. 1984. J. Colloid Int. Sci. 99(1):
10917
62. Ferrari B, Moreno R. 1997. J. Eur. Ceram.
Soc. 17:54956
63. Wang G, Nicholson PS, Sarkar P. 1997. J.
Am. Ceram. Soc. 80(4):96572
64. Powers RW. 1975. J. Electrochem. Soc.
122(4):48286
65. Fisch HA. 1972. J. Electrochem. Soc.
119(1):5764
66. Fessenden RJ, Fessenden JS. 1986. Organic Chemistry. Monterey, CA: BrooksCole. 3rd ed.
67. Labib ME, Williams R. 1986. Colloid
Polym. Sc. 264:53341
68. Labib ME, Williams R. 1984. J. Colloid
Int. Sci. 97(2):35666
69. Jensen WB. 1978. Chem. Rev. 78(1):1
21
70. Okuyama M, Haggerty JS, Garvey G,
Ring TA. 1989. J. Am. Ceram. Soc.
72(10):191824
71. Bolger JC. 1981. In AcidBase Interactions, ed. KL Mitchel. New York:
Elsevier
72. Vandeperre L. 1998. Electrophoretic
shaping for ceramic laminated composites. PhD thesis, Katholieke Univ., Leuven, Belgium. 220 pp.
73. Vandeperre L, Van der Biest O. 1999. J.
Eur. Ceram. Sci. Submitted
74. Vandeperre L, Zhao C, Van der Biest
O. 1999. 6th Conf. Eur. Ceram. Soc.
Brighton, UK. In press
75. Kennedy JH, Foissy A. 1977 J. Am. Ceram. Soc. 60(1):3336

351

76. Blanco Lopez MC, Rand B, Riley F.


1996. Key Eng. Mater. 132136:305
8
77. Hamaker HC, Verwey EJW. 1940. Trans.
Farad. Soc. 36:18085
78. Koelmans H. 1955. Philips Res. Rep.
10:16193
79. Grillon F, Fayeulle D, Jeandin M. 1992.
J. Mater. Sci. Lett. 11:27275
80. Sarkar P, Nicholson PS. 1996. J. Am. Ceram. Soc. 79(8):19872002
81. Russ BE, Talbot JB. 1998. J. Electrochem.
Soc. 145(4):125356
82. Bouyer F, Foissy A. 1999. J. Am. Ceram.
Soc. In press
83. Brown DR, Salt FW. 1965. J. Appl. Chem.
15(1):4048
84. Sarkar P, Nicholson PS, Prakash O,
Wang G. 1994. Ceram. Sci. Eng. Proc.
15(5):101927
85. Sheppard SE. 1927. Trans. Am. Electrochem. Soc. 52:4782
86. Sarkar P, Huang X, Nicholson PS. 1993.
Ceram. Eng. Sci. Proc. 14:70717
87. Hamaker HC. 1940. Trans. Farad. Soc.
36:27987
88. Gutierrez CP, Wallace TC, Mosley JR.
1962. J. Electrochem. Soc. 109:923
27
89. Sarkar P, Nicholson PS. 1995. J. Am. Ceram. Soc. 78(11):316566
90. Zhang Z. 1995. J. Am. Ceram. Soc.
78(11):316768
91. Dunbar JC, Mitchell TJ. 1953. J. Sci. Food
Agr. 4:19096
92. Ferrari B, Moreno R. 1995. In Fourth
EuroCeramics, ed. C. Galassi, 1:421
28. Faenza, Italy. Gruppo Editoriale
Faenza Editrice
93. Vandeperre LJ, Van der Biest OO. 1998.
In Innovative Processing and Synthesis of
Ceramics, Glasses, and Composites, Ceram. Proc., ed. NP Bansal, KV Logan, JP
Singh, Ceram. Trans. Vol. 85. Westerville
OH: Am. Ceram. Soc.
94. Bockris JOM, Reddy AKN. 1973. In
Modern Electrochemistry, New York:
Plenum Rosseta. 1432 pp. Vol. 1 & 2
95. Moritz K, Muller E. 1997. Werkstoff
und Verfahrenstechnik (DGM) 563
68
96. Kuwabara K, Noda Y. 1993. J. Mater. Sc.
Lett. 28:525761
97. Mizuguchi J, Muchi T, Sumi K. 1983. J.
Electrochem. Soc. 130(9):181925
98. Wakamatsu KS. 1995. J. Mater. Sci. Lett.
14:46769
99. Furman VV, Gaiduchenko GK, Vlastuk
RZ, Deimontovich VB. 1983. Porosh.
Metall. 251(11):8387
100. Valdes JL, Huggins H, Mitchel, JW,

P1: APR/SPD

P2: APR/ARY

May 15, 1999

14:52

352

101.
102.
103.
104.
105.
106.
107.

QC: APR

Annual Reviews

AR086-11

VAN DER BIEST & VANDEPERRE


Mucha JA, Seibles L. 1991. J. Electrochem. Soc. 138(2):63536
Zhitomirsky I. 1998. Mater. Lett. 37(9):
7278
Mahapatra AK, Dhananjayan N. 1981.
Trans. Ind. Inst. Metals, 34(6):495500
Hill CGA, Rees ALG, Lovering PE. 1947.
Trans. Farad. Soc. 43:40717
Pearlstein F, Gallacio A, Wick R.
1963. J. Electrochem. Soc. 110(7):843
46
Giersig M, Mulvaney P. 1993. J. Phys.
Chem. 97(24):633436
Standish JV, Boerio FJ. 1980. J. Coat.
Technol. 52(663):2939
Nagai M, Takuma Y, Yamashita K, Umegaki T. 1993. J. Am. Ceram. Soc. 76(1):
25355

108. Caley WF, Flengas SN. 1976. Can. Met.


Quart. 15(4):37582
109. Hang KW, Anderson WM, Andrus J.
1981. RCA Rev. 42(6):15977
110. Baumgartner CE, Grimaldi JJ, DeCarlo
VJ, Glugla PG. 1985. J. Electrochem. Soc.
132(1):5763
111. Hirayama C, Berg D. 1963. Electrochem.
Technol. 1(78):22427
112. Hosseinbabaei F, Raissidehkordi B. 1998.
Proc. 9th CIMTEC Conf. TECHNA Srl,
Faenza, Italy. In press
113. Grosso PF, Rutherford RE, Sargent DE.
1970. J. Electrochem. Soc. 117(11):1456
59
114. Williams EW, Jones K, Griffiths J, Roughley DJ, Bell JM, et al. 1979. Solar Cells
1:35766

Potrebbero piacerti anche