Sie sind auf Seite 1von 16

Journal

J. Am. Ceram. Soc., 90 [11] 33733388 (2007)


DOI: 10.1111/j.1551-2916.2007.02013.x
r 2007 The American Ceramic Society

Charging Behavior at the AluminaWater Interface and Implications


for Ceramic Processing
George V. Franks*,w and Yang Ganz
Chemical and Biomolecular Engineering, University of Melbourne, Victoria 3010, Australia

The interaction of water and the alumina surface is comprehensively reviewed. Water can be incorporated in the alumina crystal structure resulting in the formation of aluminum hydroxides
such as gibbsite. Alumina dissolves into water to an extent that
depends primarily upon the solution pH and temperature. The
soluble Al (III)aq species (hydrolysis products) likewise depend
upon the solution pH, temperature, aluminum, and other salt
concentrations. The development of charge on the surface of
alumina is controlled by amphoteric surface ionization reactions.
The charging behavior of both alumina powders and single crystal faces is compared. The differences can be explained by the
reactivities of different types of surface hydroxyl groups. The
substantial difference in surface charging behavior of single
crystal sapphire and alumina powders indicates that experiments and modeling conducted on single crystals is of limited use
in predicting suspension behavior. The atomic scale structure of
the hydroxylated sapphire (0001) basal plane is nearly identical
to the gibbsite (001) basal plane. The observed surface structures
are consistent with the charging behavior of the surfaces. The
role of surface charge on the adsorption of processing additives
is briey discussed. How surface charge and processing additives
at the alumina aqueous solution interface inuence surface forces between particles is reviewed. The inuence of these forces on
suspension properties such as rheological behavior is outlined.
The importance of controlling these behaviors to improve colloidal ceramic powder processing is stressed.

I. Introduction

interface between alumina and water is important in a


wide number of applications. Alumina particles are frequently prepared as aqueous suspensions for colloidal processing1,2 of slipcast,3 gelcast,4 and tape cast5 ceramic components.
Aqueous solgel processing is useful for production of alumina
HE

D. Greencontributing editor

Manuscript No. 23068. Received April 10, 2007; approved July 26, 2007.
Financial support for our research into alumina surfaces has been provided over a
number of years from the Australian Research Council, the University of Melbourne and
the University Newcastle.
*Member, American Ceramic Society.
w
Author to whom correspondence should be addressed. e-mail: gvfranks@unimelb.
edu.au
z
Current address: Department of Applied Chemistry, Harbin Institute of Technology,
ygan@hit.edu.cn.

powders.6 Alumina is often used to coat titania pigment particles


that are widely used in aqueous based latex paints.7 The coatings
are usually derived from aqueous solutions of aluminum sulfate
or sodium aluminate. The transition aluminas are useful in catalysis8 and as supports for applications including automotive
catalytic convertors.9 Alumina also has important applications in
epitaxial growth of lms for example in blue LEDs.10 Natural
corundum (a-Al2O3) has also been intensively studied for its importance in minerals processing11,12 and adsorption of heavy
metal elements.13,14 The precipitation of gibbsite from supersaturated aqueous solutions is an important step in Bayer processing of bauxite into alumina.1522 Gibbsite is also abundant in soil
as silicate minerals stably weathering product,23 its abundance in
natural fresh water aquatic systems is important to ecology.24,25
A key parameter inuencing the behavior of many of these
applications relates to the interaction of the surface of the aluminum oxide or hydroxide and solution species such as ions,
polymers or surfactants. In some cases, the interaction of the
surface is of direct interest, such as in catalysis, adsorption for
water treatment and precipitation from pregnant Bayer solutions. In other applications, the important inuence of water
and adsorbed additives at the interface is the effect on higherlevel behavior such as viscosity or green body properties. In
these applications, the surface of alumina plays the key role in
controlling the behavior. The primary factor that inuences adsorption of a solution species to a solid surface is the charge on
both the surface and solution species. For this reason investigation of the charging behavior of alumina surfaces is the focus
of the present paper.
Before discussing the charging behavior, we rst review the
incorporation of water into the crystal structure of alumina and
the resulting aluminum hydroxide phases. We then discuss the
solubility of alumina in water and the equilibrium solution
speciation. After a thorough discussion of several aspects of
surface charging behavior, we discuss the surface structure of the
basal plane of single crystal sapphire and compare it with submicron powder surface behavior. A brief description of some
adsorption phenomena and interparticle surface forces follows.
Finally, we discuss the inuence of the interparticle forces on the
macroscopic properties of alumina suspensions and the implications for colloidal ceramic powder processing.

II. Aluminum Oxide and Hydroxide Phases


The rst interaction between alumina and water which we consider is when water is incorporated into the crystal structure of

Feature

3374

Journal of the American Ceramic SocietyFranks and Gan

alumina. The various alumina and aluminum hydroxide phases


have different crystal structures. In the present paper we use the
term alumina to refer to the anhydrous aluminum oxides of the
chemical formula Al2O3, which include corundum (a), and a
number of transition aluminas (g, w, k, d, y, and Z). The aluminum hydroxides include the trihydroxides (Al(OH)3) gibbsite
and bayerite, and monohydroxides (AlOOH) boehmite and diaspore. The thermodynamically stable form of alumina is corundum while the transition aluminas are metastable phases,
which occur during the thermal dehydration of the precursor
hydroxides. Panel A describes the bulk crystal structures of the
primary aluminum oxide and hydroxides phases although other
less technologically important phases are also known to exist.
The hydroxides contain water within their structure
primarily as hydroxide ions. In effect, a water molecule is incorporated into the crystal structure in place of an AlO bond
such that one hydrogen atom of the water molecule bonds to an
oxygen anion while the oxygen and other hydrogen bond to an
aluminum ion resulting in the formation of two aluminum
hydroxyls. This is not in fact how the hydroxide phases actually form. Naturally occurring hydroxides form hydrothermally
and/or by weathering dissolution reactions of clay minerals.25
Synthetically, most hydroxide phases form by the condensation
of soluble aluminum hydroxide species28 which will be discussed
in more detail in Section III. The most technologically important example of synthesis of hydroxide by condensation of
solution species is the precipitation of gibbsite during the
Bayer process.1522
In order to produce alumina, hydroxides are produced from
solution and thermally dehydrated. Panel B provides a summary
of the routes from the hydroxides to corundum.26,29 Typically
the hydroxides with cubic close packing, boehmite and bayerite,
form transition phases with face-centered cubic close packing
(g, Z, d, y). Likewise, diaspore (with hexagonal close packing)
and gibbsite form the transition aluminas with hexagonal close
packing (w and k). At high enough temperature the equilibrium
phase corundum (a-alumina) will form in all cases. Upon cooling the a phase remains stable. One approach to making ne
grained dense a-alumina is to use commercially available nanosized g-alumina powders which can be sintered at lower temperature than corundum powders due to the phase transition
between g and a.3032

III. Solubility and Solution Species


When alumina surfaces are immersed in water there is a nite
solubility of alumina into the water. The solubility limit and the
type of solution species depend upon the solution pH, temperature, other ionic species, and the solid phase. Solubility generally increases with temperature as is common for most materials.
The total Al (III)aq concentration in equilibrium with various
aluminas and aluminum hydroxides has a minimum pH
about 67 that ranges from about 106 to 107 M depending
on the solid phase.25,33,34 Figure 3 is an example of the solubility
of gibbsite as a function of pH at 251C. Solubility of gibbsite
increases dramatically as pH is either decreased or increased. At
pH about 3.5 the concentration of Al (III)aq in solution in equilibrium with gibbsite is about 0.1 M and about 0.01 M at pH 12.
Corundum and amorphous aluminum hydroxide solids are typically slightly more soluble than gibbsite.25,34
The aluminum containing species present in solution generally depend on the total Al (III)aq concentration, other ionic
concentrations, temperature and the pH. Only at low pH (below
about pH 4 or 5) do the species Al31 exist as the majority solution species. As pH increases the metal cation reacts with hydroxide anions in solution progressively to produce soluble
hydrolysis products according to the solution equilibria:3335

Al3 ! AlOH2 ! AlOH
2 ! AlOH3 ! AlOH4

The hydrolysis reactions of aluminum with hydroxide occur


between about pH 5 and pH 8 at room temperature and low

Vol. 90, No. 11

Al(III) concentration so that the dominant solution species is the


negatively charged Al(OH)
4 above about pH 8. The mononuclear (one Al ion) species form rapidly and reversibly. These
species dominate at low ionic strength and aluminum concentration at room temperature (see Fig. 4(a)). At pH o3, the Al31
ion prevails in a state where it is coordinated by six water molecules. At pH 47, the Al(OH)
4 ion has the aluminums coordinated by four hydroxide ions. Recent kinetic evidence has
been found for ve-fold coordination (four water molecules and
one hydroxide ion) of AlOH21 ions at pH between 4.3 and 7.36
A number of polynuclear solution species (ions that contain
more than one Al ion) are known to exist under certain solution
conditions.33,37 The most studied of these include Al2(OH)41
2 ,
71
Al3(OH)51
4 and Al13O4(OH)24 (the so called keggin ion). These
polynuclear ions form slowly but are metastable for extended
periods in supersaturated solutions. At low pH in room temperature solutions of low Al (III)aq and electrolyte concentration, the polynuclear species occur only in low concentrations.
When the concentration of Al (III)aq is increased, the occurrence
of Al31 is pushed to lower pH values and the Al(OH)
4 species
do not dominate until higher pH values because the cationic
polynuclear species dominate behavior when the pH is near
neutral (see Fig. 4(b)). Al (III)aq solution species at higher temperatures, pressures, and electrolyte concentrations are important in Bayer processing19 and in geothermal mineral
formation.38,39 The concentration and type of background electrolyte ions (due to added salt, acid, or base) inuence the type
and concentrations of solution species in a way that is still being
investigated. It is known that even the subtle difference between
Na1, K1, and Cs1 in solution inuences the solution species
and solid precipitate in the Bayer process.17,18
When alumina solid phases are immersed in water at ambient
conditions, the solution concentration of Al (III)aq will increase
slowly with time until the solution concentration reaches the
solubility limit. Because the dissolution process consumes H1 at
low pH or OH at high pH, the pH of the suspension shifts
toward neutral as the Al (III)aq concentration increases. Eventually equilibrium will be reached between the pH and the Al
(III)aq concentration in solution. For example, submicron
alumina powders dispersed in aqueous solution at pH around
3 will increase in pH over a period of several hours to days to a pH
above pH 4 as dissolution slowly occurs. If the pH is maintained
at an extremely low or high pH value for a period of several
days, the solution concentration can increase to near the equilibrium. If the pH is then adjusted to neutral where the solubility
limit is lower, the dissolved aluminum species will either precipitate out of solution usually as amorphous solids or as crystalline
solids, or remain in solution as metastable polynuclear species
such as keggin ions. The hydrolysis reaction of Al (III)aq containing salts (such as Al2(SO4)3) in neutral and slightly basic
water is a technologically important method for removing contaminants from drinking water.34,35,40 In this application, the
precipitation of amorphous aluminum hydroxide species traps
contaminants within a oc or aggregate of the precipitate. The
ocs have much greater mass than the individual contaminants
so the ocs containing the contaminants can be removed from
water by sedimentation.

IV. Surface Hydration, Charging, and Surface


Ionization Reactions
Clean surfaces of metal oxides such as alumina in vacuum are
comprised of atoms that have unsatised bonds. In vacuum,
these unfullled bonds generally result in an equal number of
positively charged metal ions and negatively charged oxygen
ions for many substances as shown schematically in Fig. 5(a).
(Other substances can be polar, i.e., all cations on one face and
all anions on the opposite face.) For a-alumina, the actual surface termination in vacuum4146 is more complicated and will be
considered in Section VII. When the metal oxide surface is immersed in water or exposed to ambient air (which usually has at

November 2007

Charging Behavior at the AluminaWater Interface and Implications

3375

Panel A. Crystal Structure of Alumina Oxide and Hydroxide Phases. (After Levin and Brandon, 1998)26
a-Al2O3, corundum: The bulk structure of corundum has oxygen atoms arranged in approximately hexagonal close packed layers. Between any two layers of
oxygen (O) atoms, two-thirds of the octahedral sites are lled by aluminum (Al) atoms in an ordered array. The aluminum atoms do not lie on the mid-plane
between the oxygen layers. The aluminum atoms displace slightly toward the unoccupied octahedral site in the cation layer either above or below. The result is that
an ordered half of the aluminum ions lie just above and half of the aluminum ions lie just below the mid-plane between the oxygen layers. The hexagonal unit cells
size is a 5 4.75 A, c 5 13 A.26,27 The structure of a-Al2O3 is shown in Fig. 1.
g- and Z-Al2O3: The structures of g and Z alumina have oxygen in approximately face centered cubic packing. These phases exist as a defect spinel structure. In the
ideal spinel structure (AB2O4) the cations reside in one half of the octahedral and one eighth of the tetrahedral sites. In the g and Z alumina phases there is
distortion of the cubic lattice and there are a number of disordered cation vacancies in order to maintain the stoichometry Al2O3. Subtle differences in the oxygen
sub-lattice distortion and amount of cation ordering are the difference between g and Z alumina.
d-Al2O3: d-Alumina has oxygen ions in approximately face centered cubic packing. d-Alumina is a superstructure of the spinel structure with ordered cation
vacancies. The unit cell is either orthorhombic or tetragonal.
y-Al2O3: y-Alumina has FCC packing of the oxygen anions and a monoclinic symmetry. The aluminum cations are equally distributed over octahedral and
tetrahedral sites.
k- and w-Al2O3: Many researchers believe that k and w alumina have oxygen atoms in hexagonal close packing and that the aluminum cations reside in octahedral
sites. Other researchers suggest that these phases may be either cubic or orthorhombic and some cations may exist in tetrahedral sites.
g-Al(OH)3, gibbsite: Gibbsite is a pillared structure where each pillar has a double layer of nearly close packed oxygen ions with aluminum ions lling 2/3 of the
octahedral sites between the two layers. The aluminum ions all lie on the midplanes between adjacent oxygen layers. Each oxygen atom is bound to a hydrogen
atom. The stacking sequence of the double layers is AB BA. Because there are no aluminum ions between the adjacent hydroxyl layers, the basal plane is a weak
cleavage plane. The resulting distortion results in a monoclinic structure. The gibbsite structure is shown in Fig. 2.
a-Al(OH)3, bayerite: Bayerite has the same pillared double layered structure as gibbsite but in this phase, the stacking sequence is AB AB. The basal plane is an
easy cleavage plane between the two hydroxide layers. The resulting distortion results in a monoclinic structure.
g-AlOOH, boehmite: The oxygen ions in boehmite are arranged in face centered cubic close packing with aluminum ions in between adjacent layers. Due to the
arrangement of hydrogen ions, the structure is orthorhombic.
a-AlOOH, diaspore: In diaspore the oxygen layers are stacked in hexagonal close packing and the cations are in octahedral sites between layers. Diaspore has an
orthorhombic structure.

Fig. 1. The a-alumina corundum crystal structure viewed along, (a) the /0001S direction (c plane) and (b) the /1010S direction. Blue spheres
represent oxygen atoms and red spheres represent aluminum atoms.

Fig. 2. (a) View of a single layer of the gibbsite structure looking down upon the basal plane. (b) Side view of the gibbsite structure orthogonal to the
basal plane. The light gray spheres represent the hydrogen atoms.

3376

Journal of the American Ceramic SocietyFranks and Gan

Panel B. Transformation Sequences of Hydroxides Through


Transition Aluminas to Corundum Upon Heating. (After
Brandon and Levin, 1998)26
diaspore
gibbsite

alpha
chi

boehmite
bayerite
0C

kappa
gamma

heta
300C

delta

alpha
theta

theta
600C

alpha
alpha

900C

1200C

Vol. 90, No. 11

surface hydroxyl groups. The primary difference between the


types of surface hydroxyl groups is the number of aluminum
ions to which the hydroxide ion is coordinated. The hydrogen
bonded water is removed at temperatures around 3001C, multiply coordinated sites dehydrate next at temperatures between
about 5001 and 7001C and the singly coordinated sites only at
temperatures above about 7001C. The last sites to dehydrate are
the defect sites. No hydroxyl groups are found at temperatures
above about 9501C.50,51,53 The fully dehydroxylated alumina
surface has a structure with oxygen anions bridging two aluminum cations.50
A single amphoteric site surface hydroxyl group reacts with
acid and base at low and high pH, respectively, via surface ionization reactions as follows,5456
Ka1

Me2OH
2 ! Me2OH H
Ka2

Me2OH ! Me2O H
least 15% relative humidity), the surface reacts with water to
produce surface hydroxyl groups13,25 (denoted MeOH) as
shown in Fig. 5(b). The reaction of a clean surface of alumina
and other metal oxides with water depends upon the partial
pressure of water in contact with the surface. Most work has
been conducted on low-index crystallographic surfaces such as
(0001) a-alumina, (0001) a-hematite, (100) and (110) rutile-titania.13,25,47,48 In high vacuum the surface has not reacted with
water so it is not hydroxylated. At very low water pressures,
only the defect sites of a surface react with water to form surface
hydroxyl groups. The defects include atoms at the edge of steps
between each atomic layer and vacancies in the structure. These
defects have higher energy and are more reactive than the majority of surface atoms, which lie on the defect free crystalline
terraces. At higher water partial pressures, the sites on the terraces can also react with water resulting in a fully hydroxylated
surface at high water partial pressures. For alumina the partial
pressure of water required to initiate reaction of terrace sites on
the (0001) basal plane is about 1 Torr.13,25 The surface of submicron powder is somewhat different than single crystal samples
because the majority of the surface sites on the powders are defect sites.49,50 Also, the surface of alumina powders prepared by
grinding or milling (the majority of alumina is used after some
milling) can have surface hydroxyl groups that are the same as
those on gibbsite.51,52
It is also possible to dehydrate a fully hydroxylated surface by
thermal treatment either in vacuum or air.50,51,53 The alumina
surface dehydrates in a step-wise fashion such that different type
of surface hydroxyl groups are removed sequentially. As described in more detail in Section VIII, there are several type of

resulting in either a positively charged surface (MeOH1


2 ) as in
Fig. 5(c), or a negatively charged surface (MeO) as in
Fig. 5(d). The values of the surface ionization reaction constants (Ka1 and Ka2) determine the point of zero charge (PZC)
of the surface and depend upon the particular type of material
(for example Al2O3, SiO2, TiO2, etc.). The PZC of different oxides can be estimated from the dielectric constant of the solid,
the Pauling metaloxygen bond strength and the metalhydroxide bond length.57 These parameters inuence the reactivity of
the surface hydroxyl site with acid and base. In fact, as discussed
in more detail in Section VIII, the surface ionization constants
depend upon the particular coordination environment of the
metal ion at the surface for any particular material.5861 Some of
the surface charge is compensated by bound ions which are
present when there is salt in the solution. Site binding models are
used to reconcile the apparent difference between surface charge
measurements from titration data and z potential results. Frequently a large fraction of ions are bound. The methods of accounting for the surface charge, ion binding and z potentials are
called surface complexation models and are discussed in detail
by Hunter.55
For each type of material, there is a pH known as the PZC,
where the majority of surface sites are neutral (MeOH) and the
net charge on the surface is zero. At a pH below or above the
PZC, the particles surfaces become positively (MeOH1
2 ) or
negatively (MeO) charged due to the reaction with either acid
(H1) or base (OH), respectively. Figure 6 shows how the concentration of surface sites changes with pH.
The surface charge at the solidsolution interface produces a
surface potential. The surface charge is neutralized by counterions (ions with opposite charge to the surface). Due to thermal
energy and entropy, the majority of counterions are not bound
to the charged surface sites. The counterions dissociate from the
surface sites for the same reason sodium chloride placed in water
dissociates into sodium cations and chloride anions. The counterions form a diffuse cloud around the surface. The surface potential drops off as a function of distance from the surface as
determined by the PoissonBoltzmann equation. (See one of the
colloid texts55,56,62 for a detailed treatment.)

V. Electrical Double Layer (EDL) Repulsion and


Van Der Waals Attraction

Fig. 3. Solubility of gibbsite in water at 251C (after Baes and Mesmer33). The maximum equilibrium solubility of Al (III)aq containing
solution species is indicated by the curved solid line.

Because a layer of immobile ions and water molecules exists at


the surface, it is not straightforward to directly measure the
surface potential of particles. Instead, a closely related potential
known as the z potential is usually measured. The z potential is
measured through experiments where relative motion between
the surface and the solution is required. The z potential is the
potential at the plane of shear between the immobilized surface
layer and the bulk solution. This plane is typically located only a
few angstroms from the surface so the difference between the z

November 2007

Charging Behavior at the AluminaWater Interface and Implications

3377

Fig. 4. (a) Fraction of solution species as a function of pH for 105 molal Al (III)aq at a total ionic strength of 1.0 in equilibrium with gibbsite at 251C.
(b) Fraction of solution species as a function of pH for 101 molal Al (III)aq at a total ionic strength of 1.0 in equilibrium with gibbsite at 251C (after Baes
and Mesmer33).

potential and the surface potential is usually not much. The z


potential corresponds even more closely with the diffuse layer
potential, which is largely responsible for controlling the EDL
interaction force between surfaces. The pH where the z potential
is zero is known as the IEP. If the counterions are indifferent,
i.e., they do not specically adsorb, the IEP and the PZC of the
surface correspond.

As mentioned earlier, counterions form a diffuse cloud that


shrouds the surface in order to maintain electrical neutrality of
the system. When two surfaces (or particles) are forced together
their counterion clouds begin to overlap and increase the concentration of counterions in the gap between their surfaces. If
both surfaces have the same charge, this gives rise to a repulsive
potential due to the osmotic pressure of the counterions, which

Fig. 5. Schematic representation of the surface of metal oxides. The blue spheres represent oxygen, the red spheres a metal cation and the small gray
spheres represent hydrogen. (a) In vacuum, unsatised bonds lead to positive and negative sites associated with metal and oxygen atoms, respectively.
(b) The surface sites react with water or water vapor in the environment to form surface hydroxyl groups (MeOH). At the isoelectric point (IEP) the
neutral sites dominate, and the few positive and negative sites present exist in equal numbers. (c) At pH lower than the IEP, the surface hydroxyl groups
react with H1 in solution to create a positively charged surface composed mainly of (MeOH1
2 ) species. (d) At pH higher than the IEP, the surface
hydroxyl groups react with OH in solution to create a negatively charged surface composed mainly of (MeO) species.

3378

Journal of the American Ceramic SocietyFranks and Gan

Fig. 6. Number density per unit area of neutral (MeOH), positive



(MeOH1
2 ) and negative (MeO ) surface sites as a function of pH.

is known as the EDL repulsion. If the particles are of opposite


charge an EDL attraction will result.
The Debye length (k1) is a measure of the thickness of the
counterion cloud (and thus the range of theprepulsion).
The

Debye screening parameter (k) is k 3:29 c (nm1) for


62
monovalent salts, where [c] is the molar concentration of
monovalent electrolyte. An approximate expression for the
EDL potential energy (VEDL) versus the surface to surface separation distance (D) between two spherical particles of radius
(R) with the same surface charge is VEDL 2peeo RC2d ekD ;
where Cd is the diffuse layer potential (typically equivalent to
the z potential), e the relative permittivity of water and eo the
permittivity of free space which is 8.854  1012C2  (J  m)1.
This expression is valid when the surface potential is constant
and below about 25 mV and the separation distance between the
particles is small relative to their size.62
The van der Waals interaction is the other force that needs to
be considered to determine the overall interaction force between
two surfaces in aqueous solution. The van der Waals force is
primarily caused by the electrodynamic interactions between
correlated, uctuating, instantaneous dipoles within the atoms
that comprise the material of interest. Pairwise summation of
the interactions between two spherical particles of the same size
when the distance between the particles (D) is much less than the
radius of the particles (R) results in the surprisingly simple relationship for the van der Waals potential energy, VvdW 5 AR/
12D. More complicated relationships arise if the particles are
not the same size or are small relative to the distance between
them.62 The sign and magnitude of the interaction for a particular pair of particles interacting in a given medium is expressed
as the numerical value of the Hamaker constant (A). When the
Hamaker constant is greater than zero the interaction is attractive and when the Hamaker constant is less than zero the interaction is repulsive. The Hamaker constant between two solid
surfaces depends upon the dielectric properties of the solids and
intervening material. Several methods are available for estimating the Hamaker constant depending on the level of approximation desired and the electro-optical data available.6267
Typical estimates for two alumina surfaces interacting across
water are that A 5 5.270.5 depending on the model and data
used.

Vol. 90, No. 11

many different techniques for measuring the z potentials and


IEP have been investigated. The general conclusion of the body
of work is that equiaxed alumina powders irrespective of their
crystal phase have IEP between about 8 and 10. There may be
subtle differences between a-alumina and gibbsite particles although the variation in measurements between individual research publications is on the order of the possible observed
differences. Table I below is a summary of some of the measurements. The IEP is dened as the pH where the z potential is
zero, while the PZC is determined from titration measurements.
The IEP and PZC are the same when the electrolyte is indifferent, that is, the ions are not specically adsorbing.
Investigation of the z potentials and IEPs of individual crystallographic planes of alumina surfaces is more recent, only
having signicantly commenced less than 20 years ago. The bestcharacterized surfaces are the basal c plane (0001) of sapphire
and the basal plane (001) of gibbsite. See Table II for a summary
of the measurements. Streaming potential measurements and
tting diffuse layer potentials from surface force apparatus and
atomic force microscopy (AFM) force data are the most frequently used methods for determining the z potentials and IEPs
of macroscopic surfaces and single crystal planes. A newly developed technique utilizing second harmonic generation has also
been used.92,93,96 These methods consistently produce IEP values for sapphire (0001) and gibbsite (001) surfaces 35 pH units
below that typically found for powder. The only exceptions are
the results of Veeramasuneni et al.78 and Tulpar et al.,97 which
present sapphire IEP values similar to powder for reasons
described in Section IX.
The charging behavior of other crystallographic planes is less
well characterized and there is less agreement between measurements from different laboratories and techniques.49,95,96,98,100,101
It is likely that differences in surface preparation and cleaning
techniques result in different surface structures and thus different charging behavior.98,102 One well documented example of
the importance in surface preparation is the difference in surface
structure of the sapphire (1
102) surface found by Trainor
et al.100 (missing Al termination) and Catalano et al.101 (stoichiometric termination) which is believed to be due to two different surface preparation and cleaning methods.102
The difference between the IEP of a-alumina powder and
c-plane sapphire has been a topic of debate since the difference
was rst observed. Concerns were raised that the lower IEP of
sapphire crystals was due to contamination of the alumina
surface with soluble silicate species that can be present in the
solution that was in contact with glassware.71 If silicate species
deposited on the surface of alumina, the IEP of the alumina
would in fact be shifted to lower pH values.103 This was found to
occur when high surface area silica and alumina powders were
stored together as suspension over periods of time greater than
weeks. This long time allowed the silica to dissolve and adsorb
on to alumina powders lowering its IEP.103 There is now
signicant evidence that the low IEP of (0001) a alumina (and
(001) gibbsite) surfaces is not due to silica contamination. Several investigations referred to in Table II were conducted under
conditions that the crystals were never in contact with solution
that was contained in glassware. Several investigators examined
the surface of the crystals with XPS49,91,95,96,98 and Auger electron spectroscopy92,93 to prove that no silica was present on the
surface of the crystals. The IEPs of pure (0001) sapphire and
(001) gibbsite clearly occur at pH values about 3 to 5 pH units
below those of the micron sized powders of the same materials.
The reason for this difference is explained in Section IX.

VI. f Potential and IEP Measurements


The IEP, z potential and surface charging of alumina have been
investigated by many researchers over the past 50 or more years.
Kosmulski6871 has produced a series of reviews in recent years
that summarize the results. Before Kosmulskis compilations,6871
the most comprehensive review of z potentials was published by
Parks in 1965.72 Many different brands of alumina powders and

VII. Atomic Scale Surface Structure: Corundum and Gibbsite


The bulk structures of corundum and gibbsite are described in
Section II. The surface of a-Al2O3 (0001) in vacuum has been
extensively modeled10,13,41,42 and characterized.10,13,4346 The reconstruction of the surface which occurs because the crystal
structure is incomplete leads to a variety of surface structures

November 2007

3379

Charging Behavior at the AluminaWater Interface and Implications


Table I. The pHIEP Values of a-Al2O3 and Gibbsite Particles

Name and size

Measurement method

Linde A (0.1 mm)

Electrophoresis
potentiometric titration
Electrophoresis
Electrophoretic mass
transport
Electrophoresis

E. Merck AG (o20 mm)


Alcoa A16 SG (0.2 mm)
Sumitomo AKP
(0.51.3 mm)
Buehler LTD (1.0 mm)

Alcoa A-16SG (0.21 mm)


Sumitomo AKP-30
(0.30 mm)
Sumitomo AKP
(0.31 mm)
JFCC RP-1 (1.85 mm)

10 mM NH4Cl

9.0

5 mM, 30 mM, 139 mM


NaNO3
1 mM KNO3

8.9
(PZC)
9.1

10 mM KNO3
10 mM KNO3

8.8
9.070.3

Ramakrishnan et al.79
Franks and Lange80

Water/unknown

8.7

Water/unknown
0.011 M LiNO3, NaNO3,
KNO3, CsNO3, KBr, KCl, KI
0.01 M NaNO3

8.29.0
9.2

Yang and
Troczynski81
Costa et al.82
Johnson et al.83

9.09.3

Hackley et al.84

0.001 M NaNO3

7.67.8

Wasche et al.85

Sumitomo AKP (0.23 mm)


Gibbsite particles,
hexagonal crystals,
(0.5 mm)
Synthetic gibbsite
particles, (2040 m2/g)
Synthetic gibbsite
particles, (200 nm
diameter, 10 nm thick
plates)
Synthetic gibbsite
particles, (200 nm
diameter, 10 nm thick
plates)
Gibbsite particles,
crystallized from Bayer
liquor, (surface area,
4 m2/g)

10 mM KBr, KNO3
0.5 M NaCl

8.7
9.8
(PZC)

Franks and Meagher49


Hiemstra et al.59

Potentiometric titration

0.0050.1 M NaNO3

Hiemstra et al.86

Potentiometric titration

0.020.1 M NaCl

10.0
(PZC)
9.0
(PZC)

Electrophoresis

0.020.1 M NaCl

10.0
(IEP)

Rosenqvist et al.87

Potentiometric titration

0.0010.1 M NaNO3

8.49.2
(PZC)

Jodin et al.88

Electro acoustic method


Electro acoustic method
Electro acoustic, streaming
potential
Electro acoustic, streaming
potential
Electro acoustic method
Potentiometric titration

9.170.1

Reference

8.870.2
8.0

LaserDoppler
electrophoresis
Electrophoresis
LaserDoppler
electrophoresis
Electrophoresis

Alcoa A-16 (0.40 mm)


Sumitomo AKP
(0.230.7 mm)
Alcoa A16 (0.36 mm)

pHIEP

1100 mM KCl, KNO3,


KClO4
2 mM KCl
Water/unknown

Titration

Alfa-Aesar (1 mm)

Salt concentration and type

Yopps and
Fuerstenau73
Dick et al.74
Hashiba et al.75
Velamakanni and
Lange76
Hayes et al.77
Veeramasuneni et al.78

Rosenqvist et al.87

The particles are a alumina unless otherwise noted. PZC is the point of zero charge as determined by titration. IEP, isoelectric point.

Table II. Isoelectric Points of (0001) Basal Plane Sapphire and (001) Gibbsite
Surface

c-Plane sapphire
c-Plane sapphire
Sapphire, orientation not
specied
c-Plane sapphire
c-Plane sapphire
Sapphire, orientation not
specied
c-Plane sapphire
c-Plane sapphire
c-Plane sapphire
c-Plane sapphire
c-Plane sapphire
c-Plane sapphire
Gibbsite

Method

Check for silica contamination

SFA
SFA
AFM

No
No
Plasma cleaned

Streaming potential and


AFM
Second harmonic
generation and AFM
AFM

XPS

Streaming potential and


AFM
Streaming potential
Second harmonic
generation
AFM
AFM
Streaming potential
AFM

Auger electron
Spectroscopy
Care was taken to avoid
silica contamination
XPS
XPS
XPS
Plasma cleaned
Not plasma cleaned
XPS
Cleaved surface

SFA, surface force apparatus; AFM, atomic force microscope; IEP, isoelectric point.

IEP

o6.7
3.0
9.3
4.2
56

Reference
89

Horn et al.
Ducker et al.90
Veeramasuneni et al.78
Larson et al.91
Stack et al.92 and Stack et al.93

5.05.9

Meagher et al.94

4.85.4

Franks and Meagher49

67
4.1
8.5
56
4.2
5.9

Kershner et al.95
Fitts et al.96
Tulpar et al.97
Tulpar et al.97
Lutzenkirchen98
Gan and Franks99

3380

Journal of the American Ceramic SocietyFranks and Gan

Vol. 90, No. 11

Fig. 7. Schematics of two types of surface termination of a-Al2O3 (0001) plane. (a) (O3  O3) single Al termination found in vacuum; (b) (1  1) O
termination found in water. Large blue spheres represent oxygen atoms and small red spheres represent aluminum atoms. Note that the hydrogen atoms
of the surface hydroxyl groups of the oxygen-terminated surface in water are not shown.

under differing conditions. Most work has been conducted on


sapphire (0001) and gibbsite (001) surfaces. Low energy electron
diffraction (LEED)43 and X-ray methods44,45 have shown that
the sapphire (0001) surface is terminated in vacuum with a single
layer of aluminum atoms under most partial pressures. This
single layer of aluminum atoms contains half of those which lie
between two adjacent close-packed oxygen layers in the bulk
structure. Remember from Fig. 1 how the aluminum atoms lie
on either side of the mid-plane between the close packed oxygen
layers. The surface structure in vacuum is illustrated in Fig. 7(a).
Some earlier reports41,46 indicate that the (0001) sapphire surface can be terminated in a close-packed layer of oxygen atoms
in some near vacuum conditions as shown in Fig. 7(b). Knowledge of the surface structure in vacuum is of limited use in understanding the behavior of the surface of alumina in water
because the surface structure in water is signicantly different
than in vacuum.25
In a humid atmosphere or water, partially and fully
hydroxylated a-Al2O3 (0001) surfaces have been characterized
by LEED,43,104 X-ray methods,44,45 and dynamic-mode scanning force microscopy (SFM).105 The experimental results are
supported by calculations based on rst-principles that indicate
strong surface relaxation occurs in the presence of water
vapor due to interaction between water molecules and the surface.106108 When exposed to water vapor, the unhydrated

alumina surface reacts to form a hydroxyl group (OH) terminated surface (the coverage of OH depends on the partial
pressure of water).13,25 Figure 7(b) shows the arrangement of the
oxygen atoms on the (0001) sapphire surface in water (note, the
hydrogen atoms are not shown). Eng et al.45 used Crystal Truncation Rod X-ray diffraction to determine the surface atom positions. They found that signicant surface relaxation occurs
such that the aluminum atoms between the top two oxygen layers shift positions such that they lie closer to the mid plane between the oxygen layers. Relaxation also occurs in the topmost
oxygen layer. The result of the surface relaxations is that the
oxygen and aluminum atoms in the surface layers reside in
nearly the same positions as the surface atoms in gibbsite. The
hydrogen atom positions were not determined by Eng et al.s45
X-ray work but simulations106108 have indicated that the hydrogen atoms of the surface hydroxyl groups are gibbsite like as
well. At any particular instant, two of the three surface hydroxyl
groups bound to any aluminum ion are nearly vertical to the
surface, while the remaining one lies down at on the surface in
a position just above the unoccupied octahedral site just below.
These rearrangements produce a surface structure closer to that
of gibbsite than that of bulk sapphire. This viewpoint has also
gained support from infrared (IR) spectra studies of water
adsorption on a-Al2O3 (0001) single crystal surface.109 Figure
8 is a schematic comparison of the hydrated a-alumina surface

Fig. 8. Comparison of the surface structure of a-Al2O3 and gibbsite in water (after Eng et al.45). Blue spheres represent oxygen atoms, red spheres
represent aluminum atoms and the light gray spheres represent the hydrogen atoms. (a) The gibbsite surface has all the aluminum atoms along the
midplane between the two uppermost oxygen layers. Two thirds of the surface hydroxyl groups are oriented orthogonal to the surface and one third are
parallel to the surface. (b) Surface relaxation of the sapphire surface occurs such that the top layers of aluminum atoms move closer to the mid plane of
the two uppermost oxygen layers and the oxygen atom positions relax from the close packed structure slightly to positions like on the gibbsite surface.
Like gibbsite, two thirds of the surface hydroxyl groups are oriented orthogonal to the surface and one third are parallel to the surface.106108

November 2007

Charging Behavior at the AluminaWater Interface and Implications

Fig. 9. The step-terrace surface structure (topographical image) of the


sapphire basal (0001) plane in water with 0.51 off-cut angle. Image size:
200  200 nm2; scanning rate: 1 Hz. Each step is about 2 A in height,
which corresponds to the distance between adjacent oxygen layers.

and the gibbsite surface. The surface of hydrated a-alumina


powders prepared by grinding were also found to contain surface hydroxyl groups that are gibbsite like after hydrothermal
hydration.51,52
The advent of scanning probe microscopy (SPM) and AFM
has allowed for direct observation of the surface structure of
alumina in vacuum, low-pressure vapors,105 air,110,111 and water.111,112 Using a dynamic SFM technique Barth and Reichling105 investigated a reconstructured surface of sapphire (0001).
They imaged a (O31  O31)R791 structure produced at 13001C
in ultra high vacuum with atomic resolution. After exposure to
very low (milliPascal) levels of water vapor they observed many
clusters of size 0.40.8 nm. They supposed these clusters were
most probably crystalline Al(OH)3. As early as 1993, Steinberg
et al.111 achieved unit cell resolution images of nanocrystalline aalumina that was grown on mica by the van der Waals epitaxy
method. The 10nm-thick alumina lms were grown on freshly
cleaved mica sheets by vacuum depositing Al2O3 from a pellet
vaporized using an electron beam. They observed a hexagonal
lattice (period of 0.4770.02 nm) at various places on the lm in
both force and height modes in air, water and salt solutions. To
our knowledge, there is only one paper that reports the surface
structure of gibbsite obtained using AFM.113 However, few experimental details are supplied. It is believed that the image
published was taken in dry air.

3381

Recently we have examined sapphire samples cut and polished parallel to the single crystal a-Al2O3 (0001) basal plane
(miscutting angle between about 0.51 to 0.81) both in water112
and in air.110 In water, these surfaces have terraces about 20 nm
wide112 as shown in Fig. 9. The height of the steps is about 2 A,
which is equivalent to the distance between adjacent oxygen
layers.112 As shown in Fig. 10(a), closer inspection of the terraces reveal hexagonal arrays of corrugations. A two dimensional fast Fourier transformation of the image data indicates
the distance between the nearest corrugations is 4.7 A. The hexagonal periodicity of the structure is clearly shown in Fig. 10(b).
The hexagonal structure and the 4.7 A period is equivalent to
the known bulk unit cell size and structure for a-alumina,
a 5 4.75 A.26,27 Although we have not been able to resolve individual atoms in our studies,110,112 our investigations reveal low
spots (dark spots in Fig. 10(a)) in the surface with hexagonal
periodicity of 4.7 A. Figure 10(c) helps to illustrate how these
features correspond to the atomic arrangements on the surface
of the sapphire basal plane. The hollows in the surface structure
appear to correspond with the position where the surface hydroxyl groups are lying down (just above the unoccupied octahedral site in the sublayer). The position where the surface
hydroxyl groups stand up from the surface perhaps correspond
with the high points (bright spots) in the surface structure.

VIII. Multisite Complexation Models of Alumina Surfaces


In Section IV the concept of surface hydroxyl groups (MeOH)
was introduced. In this section we discuss specically the reactivity of the aluminum surface hydroxyl groups (AlOH) found
on alumina surfaces. As early as the 1970s8 it was recognized
that there are more than one type of aluminum surface hydroxyl. The difference between the different types of surface
hydroxyls primarily depends upon the number of aluminum atoms to which the hydroxyl is bound8,50 as well as the coordination of the aluminum ion (octahedral or tetrahedral). A
hydroxyl bound to one aluminum atom is known as a singly
coordinated surface hydroxyl (denoted AlOH), likewise, when
bound to two or three aluminum ions the hydroxyl is referred to
as doubly (Al2OH) or triply (Al3OH) coordinated, respectively. It has also been recognized that these different types of surface hydroxyl groups will have different pKa values controlling
their charging behavior.8,58,59,114
In the late 1980s, Hiemstra and coworkers58,59 developed the
multisite complexation (MUSIC) model to account for the possibility of different type of surface hydroxyl groups with differing reactivity (pKa values). They recognized that intrinsic
equilibrium constants, where no effects of long-range electrostatic effects exist, must be constrained before model tting in

Fig. 10. (a) Atomic scale image of single crystal a-Al2O3 (0001) surface in water showing hexagonal arrangement of bright (high) and dark (low) points
with a periodicity of 4.7 A. The image is an unltered deection image. (b) Overlay of lines highlights the hexagonal structure of the surface. (c) The
correspondence between the dark regions and the position, where the alumina surface hydroxyls lie down above the octahedral vacancies in the sublayer,
is highlighted by overlay of the top layer of a fully hydrated sapphire basal plane.

3382

Vol. 90, No. 11

Journal of the American Ceramic SocietyFranks and Gan

Table III. pKa Values Estimated by Applying the Bond-Valence Method to Gibbsite (001) and (100) Surfaces
Estimate of pKa for each type of surface hydroxyl61,115
Site number
Plane

Basal (001)
Basal (001)
Prismatic (100)
Prismatic (100)

Reaction

1
Al2OH1
2 -Al2OH1H
Al2OH-Al2O1H1
1
Al2OH1
2 -Al2OH1H
AlOH11/2-AlOH1/21H1

2.3
3.9
2.2
11.6

1.6
11.4

5.1
3.2

0.4
8.8

5.2
14.4

10.8
20

order to obtain realistic molecular-scale insights from surface


complexation models. The MUSIC model constrained many of
the adjustable parameters used in earlier models.77 The model
parameters are pKa values, capacitance values and electrolyte
binding constants as well as site densities. The pKa values are
estimated based on the electrostatic approach, site densities
come from the crystal structure and the rest is tted to titration
data. Additional improvements in the 1990s60 incorporated
equilibrium constants that are calculated based on Paulings valence rules and an empirical relationship (calibrated on some
hydroxy-acid solution species). Using the improved MUSIC
model, Hiemstra et al.86 predicted the pKa values for the deprotonation reactions of singly and doubly coordinated gibbsite
surface hydroxyl groups.
For singly coordinated:
pKa 9:9

Al2OH1=2 ! Al2OH1=2 H
And for doubly coordinated:
pKa1 0

Al2 2OH
2 ! Al2 2OH H

pKa2 11:9

!

Al2 2O 2H

Careful examination of Figs. 8, 1 and 2 indicates that the


basal planes of sapphire and gibbsite are composed of doubly
coordinated surface hydroxyl groups. Singly coordinated surface hydroxyl groups occur in greater proportion on prismatic
and other high index planes. Singly coordinated surface hydroxyl groups also occur along the edge of steps on the basal
plane. Gibbsite particles are typically hexagonal prismatic in
shape. The overall charge of the particle results from the combination of contributions from the basal plane and the edge
faces.
Hiemstra et al.s86 predictions of pKa values of aluminum
surface hydroxyl groups were successful in predicting the IEP of
some gibbsite particles. Hiemstras results predict IEP around
9.9 for gibbsite and alumina particles which is not a bad estimate. Moreover, for a number of gibbsite samples studied, experimental surface charge measured from potentiometric
titrations correlate fairly well with samples edge surface area.
But according to this model, all the surface charge at pH in the
usual range of interest (pH 212) is due to the singly coordinated
surface hydroxyls. According to Hiemstras model, the basal
planes which contain doubly coordinated surface hydroxyl
groups are neutral over the usual pH range of interest. More
recent experimental results presented in Table II indicate that
the basal planes of sapphire and gibbsite are in fact at least
weakly charged.99
In addition, some researchers61,87,88 have found that the original MUSIC model can not explain some potentiometric titration results on gibbsite. They have been particularly concerned
with the effect of titration time and the role of doubly coordinated sites in the overall charge and proton balance. Recently,
Bickmore et al.61 emphasized the need to consider the limitations of the MUSIC model. They argued that the MUSIC model
does not properly treat long-range electrostatic, short-range and
solvation contributions. Bickmore et al.61 also proposed an improved method of pKa prediction that combine bond-valence

theory and ab initio determined molecular structure of (hydr)oxide surfaces. The bond valance and metaloxygen bond lengths
of the surface hydroxyl groups determined by ab initio surface
structure calculations are used as input parameters for the MUSIC model. Initially the ab initio calculations were performed for
surfaces terminated in vacuum; the inadequacy of this approach
is discussed later.
Bickmore and co workers61,115 chose a surface unit cell with
six unique hydroxyl groups and used ab initio calculations to
optimize the vacuum terminated gibbsite (001) surface. Six
different doubly coordinated surface hydroxyl sites resulted
due to subtle differences in bond length and ionicity which
emerge from the ab initio calculations. Table III gives the pKa
values estimated by Bickmore and coworkers61,115 for (001) basal and (100) edge surfaces of gibbsite. The six different doubly
coordinated surface hydroxyl groups on the basal plane produced six different pKa values over the range shown in Table III.
The choice of six surface hydroxyl groups as the unit cell for
modeling is somewhat arbitrary and if a larger unit cell were
used it is likely that more unique pKa values would result, but
over a similar range as those presented in Table III. Although
the analysis has not been performed for the a-alumina (0001)
surface, the similar surface structures of the basal gibbsite and
corundum structures suggest that similar pKa values are likely to
result. The prismatic planes of gibbsite contain both doubly coordinated and singly coordinated surface hydroxyl groups. The
doubly coordinated groups are set deeper in the surface and
are believed to be sterically hindered from reacting so that they
do not participate signicantly in the charging behavior of the
prismatic plane surfaces. (Note, that the similarity which we
have been drawing between the basal planes of sapphire and
gibbsite cannot be made for the prismatic faces of the two materials because of the different surface structures of these faces.)
Jodin et al.88 used a detailed IR spectroscopic characterization to determine the bond lengths and valences of the surface
hydroxyl groups on the basal and prismatic planes of gibbsite.
This information was used to determine the pKa values of the
surface ionization reactions that dominate the basal and prismatic planes of gibbsite. A range of pKa values were determined
which depended upon the amount of relaxation of the angle of
the surface hydroxyl group. These researchers found for the
doubly coordinated basal sites:88
Al2 2OH
2

pKa 2 to 4

!

Al2 2OH H

And for singly coordinated edge sites on the prismatic planes:


Al2OH1=2

pKa 7:9 to 9:9

!

Al2OH1=2 H

Although there are some discrepancies between the various


models, one general feature of all the multisite surface charging
models is universal. The singly coordinated surface hydroxyl
groups have higher pKa values than doubly coordinated surface
hydroxyls.8,61,86,88,114 In addition, the results of Bickmore61,115
and Jodin88 and coworkers reveal that there are some doubly
coordinated surface hydroxyl groups on the basal surface that
have pKa values around 45.2. Experimental results from our

November 2007

Charging Behavior at the AluminaWater Interface and Implications

3383

lab suggest that these sites are most likely dominating the behavior of gibbsite basal surface producing the weak surface
charging.99 In light of the signicant similarity between the basal
planes of gibbsite and sapphire we believe that these results also
are consistent with the experimental results on the basal plane of
sapphire (see Table II) which show the surface has IEP at pH
around 46.
However, there are still signicant challenges for improving
MUSIC and bond valence models. The most recent work of
Bickmore and coworkers116 explains that the methods can be
improved by better prediction of bond valance and length using
molecular dynamics ab initio models that include the interaction
of the surface with water. As the surface reconstructions in vacuum are not the same as in water, accurate prediction of the pKa
values of different types of surface hydroxyl groups will rely on
accurate bond properties of the surface in water.

IX. Single Crystal Alumina Surfaces Versus Powder


The difference between the IEP of (0001) plane sapphire (Table
II) and submicron a-alumina particles (Table I) can be explained
by the different types of surface hydroxyl groups found on the
two different surfaces. The surface structure of the basal planes
of hydrated sapphire and gibbsite are comprised primarily of
doubly coordinated surface hydroxyl groups.43,45,50,106108 These
doubly coordinated surface hydroxyl groups have pKa around
pH 46 as discussed in the previous section. On the other hand,
colloidal sized particles with radius of curvature on order of a
few microns or less cannot express a large fraction of their total
area in low index planes, such as (0001) that have surface hydroxyl groups coordinated to multiple aluminum ions. Instead,
the surface of powder is composed of plane edges, steps, vacancies and other defects.49,50 This means that the surfaces of colloidal powders are composed primarily of singly coordinated
surface hydroxyl groups which have much higher pKa values,
typically around pH 911. Figure 11 shows part of an a-alumina
crystal that has been cut into a spherical like shape. One can see
that there are likely to be many singly coordinated surface hydroxyl groups. Investigators have identied the surface hydroxyl
groups found on several types of submicron a-alumina powders
to be similar to those found on gibbsite and/or other aluminum
hydroxides.51,52
The concern that the sapphire surfaces may have low IEP due
to contamination by soluble silicate species can be dismissed.
Several investigators whose results are presented in Table II,
examined the surface of the crystals with XPS 49,91,92,95 and Auger electron spectroscopy92,93 to prove that no silica was present
on the surface of the crystals. Also the (0001) sapphire investigated by Franks and coworkers49,110,112 has been imaged by
high-resolution AFM and has been found to have at terraces
about 1520 nm wide with steps of height 0.2 nm, equal to a
single layer of oxygen atoms as shown in Fig. 9. High-resolution
images such as those shown in Fig. 10 demonstrate that the at
surfaces are crystalline with few defects because the lattice scale
structure can be observed across the majority of the terrace.
There is no evidence in these images for any contamination, inorganic (such as silica), or organic. The IEP of such surfaces are
around pH 5.49
Only two of the recent reports presented in Table II indicate
that the (0001) surfaces of sapphire have IEP similar to powder.78,97 These results only occur when the surfaces have been
treated with oxygen plasma. Although the authors of those papers 78,97 recognised the inuence of plasma treatment, they did
not give an explanation for this interesting fact. In other
work,104 it was found that single crystal (0001) sapphire surfaces can be roughened when exposed to a short (30 s) plasma
treatment. The impact of plasma treatment, according to Elam
et al.,104 is that the surface contains an increased number of reactive sites such as edges and vacancies. The roughened surface
produced by the plasma treatment must contain a signicant
fraction of singly coordinated surface hydroxyl groups and increase the IEP toward that found for powders.

Fig. 11. Schematic illustration of a submicron particle surface indicating that a high fraction of the surface sites may be singly coordinated
surface hydroxyls. Note this image is generated by cutting a corundum crystal and no accounting has been made for surface relaxation
or interaction with water.

z Potential measurements on a-alumina platelets also help us


to understand the role of low index planes such as (0001) in the
charging of submicron particles. Platelet-shaped a-alumina particles have a lower IEP than the typical potato-shaped particles such as Sumitomo AKP particles. z Potential measurements
were preformed on Sumitomo AKP-50 potato-shaped particles and a-alumina platelets from ATO Chem (circa 1994) in
0.01M NH4Cl using a Zeta Meter 3.0 microelectrophoresis apparatus. The results are shown in Fig. 12. The size of the platelets was about 520 mm across and about 13 mm thick as shown
in Fig. 13. Quite clearly the z potential of the platelets is dominated by the contribution from the basal plane. There was signicant scatter in the z potential measurements as indicated by
the wide red region in Fig. 12. This is because some to the
smaller equiaxed particles had only a small fraction of basal
plane relative to total particle surface area and had higher IEP
than the high aspect ratio plate-like particles.
The charge on the alumina surface and the charge of the solution species are the primary factors that control adsorption of
the dissolved processing additives at the aluminawater interface. Of course, dissolved species with charge opposite to that of
the particles surface tend to adsorb while species with like
charge to the surface tend not to adsorb. Other factors such
as van der Waals interactions, hydrogen bonding and hydrophobic interactions inuence adsorption, but usually are second
order effects.117 Because the IEP of single crystal sapphire and
a-alumina powders are substantially different, the adsorption
behavior of solution species will be signicantly different on the
two different types of surfaces. For example actinide cations
which adsorb only weakly to alumina powders at pH below
about six will readily adsorb on to the basal plane of sapphire at
pH values as low as 4.5.118,119 Also, anionic polymers, which are
typically used to disperse alumina powders, adsorb to alumina
particles over a wide range of pH while they would not be
expected to adsorb to single crystal sapphire except at pH below
about 4 or 5.

3384

Journal of the American Ceramic SocietyFranks and Gan

Fig. 12. z Potential as a function of pH for a-alumina platelets and


potato-shaped particles. The solid black line is for Sumitomo AKP-50
potato-shaped particles. The red region indicates the range of measurements for a sample of ATO-chem a-alumina platelets. The platelets
are shown in Fig. 13. The larger platelets had z potentials near the lower
left boundary of the region while the smaller particles had z potentials
near the upper right boundary.

Experimental measurements and modeling of interaction


forces, spectroscopy, and adsorption, performed on a-alumina
single crystal samples is not likely to provide useful information
for improving ceramic processing of alumina powders.
Although studies on single crystal sapphire will aid in improving fundamental knowledge about adsorption, surface charging
behavior and interaction forces, these studies are of limited value in predicting specic alumina suspension behavior due to the
vastly different surface charging behavior. The authors believe
that the development of macroscopic molecularly smooth alumina surfaces with IEP near pH 9 will be benecial to linking
surface force and spectroscopy measurements for example with
suspension behavior. Preliminary results120 indicate that glass
slides coated with aluminum nitrate solutions and heated
at moderate temperatures (30018001C) have IEP in the range
7.58.0, but are not nearly as smooth as commercially available
sapphire samples.

X. Processing Additives at the AluminaWater Interface and


Implications for Ceramic Powder Processing
Most processing additives are used to improve suspension stability and reduced viscosity through increasing the interparticle
repulsive forces. Dispersants (aka deocculants) can be either
organic (polymers,2,121127 surfactants2,121,128132 or other small

Fig. 13. Scanning electron micrograph of the ATO chem a-alumina


platelets whose z potential is presented in Fig. 12. Note that both large
platelets and small equiaxed particles can be found in the sample.

Vol. 90, No. 11

molecules,133137) or inorganic (silicate, phosphate, etc.).2,3,126


The adsorption of these additives to alumina and the resulting
interparticle interactions are primarily controlled by the surface
charging behavior. The interaction between particles in suspension may be repulsive or attractive. A combination of the surface
properties (including charging behavior), solution chemistry and
adsorption of processing additives at the alumina water interface
inuences the interactions between individual particles.2,55,62
These interparticle interactions (forces) as described in Section
V control the behavior of suspensions of the particles such as
suspension stability, settling rate, nal sediment moisture, consolidation behavior and shear ow properties such as shear yield
stress and viscosity.12,55,56
Particles that have repulsive forces between each other remain
as individual particles in the liquid. When the particles are very
small (less than about 5 mm) they have very little mass and will
settle very slowly under the inuence of gravity because Brownian motion keeps them suspended. Because the settling rate of
particles depends upon the particle (or aggregate) size and density,138 attraction, which produces larger, more massive aggregates should be avoided in order to keep suspensions stable
against sedimentation. Dispersed particle suspensions are also
relatively easy to consolidate to relatively high volume fractions
of solids (low-moisture contents) even at low applied consolidation pressures compared with occulated suspensions.139141
The rheological behavior of aqueous suspensions depends upon
the volume fraction of solids, the size and shape of the particles,
and the interparticle forces.2,142,143 Strong attraction between
particles results in high viscosities and yield stresses usually making processing difcult. Repulsion results in stable suspensions
that have low viscosities, which typically facilitates processing.
It is important to have a well-dispersed suspension during
colloidal processing because repulsive interactions produce low
viscosity suspensions and homogeneous, uniform and highdensity green bodies. Low viscosity is important in suspension
handling such as spraying ceramic glaze suspensions3,129,144146
and in lling molds for complex shape forming such as in slip
casting,3 tape casting,5 and gelcasting.4,147,148 Repulsive interactions also allow for high and uniform green density in bodies
shaped by suspension pressure forming such as in slip casting
and pressure ltration. These properties of the green body are
crucial to ensure that shrinkage during drying and sintering is
low and, more important, uniform. If shrinkage is not uniform,
distortion of the body can occur and stresses can develop which
lead to cracking of the component during drying or ring.
A dispersant will typically act to create either an EDL repulsion as described in Section V by increasing the charge on the
particles surface or a steric repulsion that prevents particles
from being drawn into contact by van der Waals attraction.
EDL repulsion 55,56,62 can be implemented by adjusting solution
conditions so that there is a similar surface charge on all the
particles. Surface charge can be created by adjusting the pH to
values away from the IEP of the surface. One can readily recognize that the different IEP values of sapphire single crystals
and alumina powders will produce signicantly differing EDL
behavior for the two different types of alumina. EDL repulsion
can also be developed by adding a charged species (usually surfactant or polymer) that adsorbs to the particles. The adsorption
of the charged species to the alumina surface will depend strongly on the charge of the surface so dramatically differing results
are expected for single crystal and powdered alumina. Although
adjustment of pH to control surface charge of alumina particles
to create EDL repulsion can be used, this method becomes less
effective as particle size decreases.149
A more robust method, steric repulsion150 which relies on
the adsorption of a soluble polymer on the particles surface is
frequently used to disperse alumina. The polymer must be adsorbed in sufcient quantity to completely cover the surface of
the particles; typically about 0.52 wt% of polymer to weight of
solid. Water must be a good solvent for the polymer so that the
polymer extends out from the surface and creates a cushion that
prevents particles from coming together (repulsion).55,56,62,150152

November 2007

Charging Behavior at the AluminaWater Interface and Implications

The polymers that work best in creating steric repulsion are


typically low to moderate molecular weight (about 2000100 000
MW). Electrosteric stabilization occurs when the polymer creating the steric repulsion is also charged so as to create an EDL
repulsion.
The most commonly used polymeric dispersants for alumina
in water are based on acrylic acid groups including sodium or
ammonium poly acrylates and methacrylates.3,122,123,127 These
molecules are negatively charged over much of the pH range
commonly used for processing so they adsorb well on the alumina particles surfaces which are positively charged at the
typical processing pH but are less likely to adsorb to single
crystal sapphire surfaces. The resulting overall charge of the
particles surface with the adsorbed polymer becomes highly
negative when the optimum dose is used. The result is a strong,
long-ranged electrosteric repulsion. The success of the poly acids
in dispersing alumina is due to the ability of these polymers to
create both EDL and steric repulsion. When polyelectrolytes
with charge opposite to the surface are used as dispersants care
must be taken to avoid bridging occulation which can occur if
the dispersant is underdosed.151155 Flocculation is usually to be
avoided in ceramic processing1,2 although weak attraction may
be useful in preventing mass segregation for particles of greatly
differing size or density.126,156
When surfactants are added to alumina suspensions, the orientation of the adsorbed molecules and the adsorbed amount
affect the type of forces between the particles. Typically, when
the surfactant and particle surface have opposite charge, the
hydrophilic head-group will adsorb to the particles surface due
to Coulombic attraction. When the solvent is water and the
surfactant adsorbs in the head-to-surface conguration with the
hydrophobic tails forced to stick out into the aqueous solution,
hydrophobic attraction is created between the surfactant-coated
particles if the concentration of surfactant is low such that only
a monolayer or less is formed.128,129 Due to the resulting hydrophobic attraction, this situation is usually to be avoided. At
higher levels of added surfactant, the molecules will adsorb as
surface micelles or a bilayer, shielding the hydrophobic tails
from the water. This type of adsorption can create repulsion via
two mechanisms: rst, there is a short-range steric repulsion,
and if the surfactant is ionic and sufcient surface charge develops an EDL repulsion can also exist.128,129
Several other small molecules can be used to disperse alumina
in water. Most notably, citric acid and its sodium and ammonium salts134136 have been used to produce highly negatively
charged alumina suspensions resulting in EDL repulsion. The
citrate dispersed alumina suspensions are stable and have low
viscosity even at pH 9 the pristine IEP of alumina powders.
Other physically adsorbed molecules such as sugars and low
molecular weight polysaccharides have been used to induce steric stability of alumina suspensions.137 Because the alumina surface is composed of surface hydroxyl groups, molecules such as
silanes and alcohols can be chemically bound to the surface by
condensation reactions. Silanes with short chains can be reacted
to the surface of alumina either in organic solvent with small
additions of water or directly in aqueous solution.133 If the silane
contains a water soluble group such as an amine, carboxylate or
ethylene oxide, it can be used to stabilize alumina in water. Although both physically and chemically adsorbed small molecules can impart short range repulsion and improve stability of
suspensions, the chemically adsorbed molecules are better than
physically adsorbed molecules because they are harder to remove from the surface during processing.133,157159 Another factor important, particularly in the dispersion of nanoalumina
particles is that the length of the stabilizing molecule relative to
the particles size needs to be carefully controlled for optimum
dispersion.160

XI. Conclusions
Although much research has shed light on the aluminawater
interface and charging of alumina surfaces, there is still much

3385

that we need to better understand. Solution speciation, surface


charging behavior and surface interactions, particularly at extremes of pH and at high electrolyte concentration, still need to
be explored. Better understanding of these phenomena may aid
in improving productivity in Bayer processing of alumina. Additional research on the inuence of terrace and edge sites on the
charging behavior of alumina, particularly investigations that
rule out organic contamination would be benecial. In any case,
because the charging behavior of the sapphire basal plane and
submicron a-alumina particle surfaces are different, it should be
noted that measurement and modeling on the sapphire (0001)
surface such as of polymer adsorption, surface forces, etc., are
only of limited usefulness in predicting the behavior of concentrated alumina powder suspensions. Suspension z potential measurements combined with knowledge of polymer adsorption and
an understanding of surface forces provides a strong basis for
understanding, predicting, and control of macroscopic suspension behavior such as stability, ow, and consolidation. Control
of these behaviors is useful in developing cost effective ceramic
powder processing schemes to produce reliable high performance ceramics with minimal reject rates.
Note added in proof
Recently published ab initio calculations (E. M. Fernandez,
R. I. Eglitis, G. Borstel, L. C. Balbas, Ab initio calculations of
H2O and O2 adsorption on Al2O3 substrates, Comput. Mater.
Sci., 39, 587592 (2007)) indicate that the basal plane of sapphire and amorphous clusters have different surface hydroxyl
groups with different dissociation energy, bond length and
charge distribution.

Acknowledgments
Y. G. thanks the University of Newcastle for his University of Newcastle Research Fellowship, 20042007. G. F. would like to acknowledge the informative sessions with Prof. Tom Healy and Dr. Victoria Bitter of Melbourne. Both authors
would like to thank A/Prof. Erica Wanless of Newcastle for assistance with and access to AFM facilities and Prof. Barry Bickmore of Brigham Young University and
Dr. Johannes Lutzenkirchen of Karlsruhe for providing comments on an early version of the manuscript. The photo of George Franks was taken by Grant Hobson.

References
1

F. F. Lange, Powder Processing Science and Technology for Increased


Reliability, J. Am. Ceram. Soc., 72 [1] 315 (1981).
2
J. A. Lewis, Colloidal Processing of Ceramics, J. Am. Ceram. Soc., 83 [10]
234159 (2000).
3
J. S. Reed, Principles of Ceramic Processing, 2nd Ed, John Wiley and Sons,
New York, 1995.
4
W. M. Sigmund, N. S. Bell, and L. Bergstrom, Novel Powder-Processing
Methods for Advanced Ceramics, J. Am. Ceram. Soc., 83 [7] 155774 (2000).
5
R. E. Mistler and E. R. Twiname, Tape Casting, Theory and Practice. The
American Ceramic Society, Westerville, OH, 2000.
6
B. E. Yoldas, Alumina Sol Preparation from Alkoxides, Am. Ceram. Soc.
Bull., 54 [3] 28990 (1975).
7
G. E. Morris, W. A. Skinner, P. G. Self, and R. St.C. Smart, Surface Chemistry and Rheological Behavior of Titania Pigment Suspensions, Colloids Surf. A:
Physicochem. Eng. Aspects, 155 [1] 2741 (1999).
8
H. Knozinger and P. Ratnasamy, Catalytic Aluminas: Surface Models and
Characterization of Surfaces Sites, Catal. Rev. Sci. Eng., 17 [1] 3169 (1978).
9
J. B. Wachtman, Ceramic Innovations in the 20th Century. The American Ceramic Society, Westerville, OH, 1999.
10
V. E. Henrich and P. A. Cox, The Surface Science of Metal Oxides. Cambridge
University Press, Cambridge, 1994.
11
H. J. Modi and D. W. Fuerstenau, Streaming Potential Studies on
Corundum in Aqueous Solutions of Inorganic Electrolytes, J. Phys. Chem.,
61 [5] 6403 (1957).
12
S. B. Johnson, G. V. Franks, P. J. Scales, D. V. Boger, and T. W. Healy,
Surface ChemistryRheology Relationships in Concentrated Mineral Suspensions, Int. J. Miner. Process., 58, 267304 (2000).
13
G. E. Jr. Brown, V. E. Henrich, W. H. Casey, D. L. Clark, C. Eggleston, A.
Felmy, D. W. Goodman, M. Gratzel, G. Maciel, M. I. McCarthy, K. H. Nealson,
D. A. Sverjensky, M. F. Toney, and J. M. Zachara, Metal Oxide Surfaces and
their Interactions with Aqueous Solutions and Microbial Organisms, Chem. Rev.,
99, 77174 (1999).
14
J. R. Bargar, S. N. Towle, G. E. Jr. Brown, and G. A. Parks, XAFS and
BondValance Determination of the Structures and Compositions of Surface
Functional Groups and Pb(II) and Co(II) Sorption Products on Single Crystal
Alpha-Alumina, J. Colloid Interface Sci., 185, 47392 (1997).
15
S. J. Freij and G. M. Parkinson, Surface Morphology and Crystal Growth
Mechanism of Gibbsite in Industrial Bayer Liquors, Hydrometallurgy, 78, 24655
(2005).

3386
16

Journal of the American Ceramic SocietyFranks and Gan

S. J. Freij, G. M. Parkinson, and M. M. Reyhani, Direct Observation of the


Growth of Gibbsite Crystals by Atomic Force Microscopy, J. Crystal Growth,
260, 2324 (2004).
17
J. Li, C. A. Prestige, and J. Addai-Mensah, Secondary Nucleation of Gibbsite Crystals from Synthetic Bayer Liquors Effect of Alkali Metal Ions, J. Crystal
Growth, 219, 45164 (2000).
18
H. C. Sweegers, C. de Coninck, H. Meekes, W. J. P. van Enckevort, I. D. K.
Hiralal, and A. Rijkebor, Morphology, Evolution and Other Characteristics of
Gibbsite Crystals Grown from Pure and Impure Aqueous Sodium Aluminate Solutions, J. Crystal Growth, 233 [3] 56782 (2001).
19
H. Li, J. Addai-Mensah, J. C. Thomas, and A. R. Gerson, The Inuence of
Al(III) Supersaturation and NaOH Concentration on the Rate of Crystallization
of Al(OH)3 Precursor Particles from Sodium Aluminate Solutions, J. Colloid Interface. Sci., 286, 5119 (2005).
20
D. S. Rossiter, P. D. Fawell, D. Iliveski, and G M. Parkinson, Investigation
of the Unseeded Nucleation of Gibbsite Al(OH)3 from Synthetic Bayer Liquors,
J. Cryst. Growth, 191, 52536 (1998).
21
I. N. Battacharya, J. K. Pradhan, P. K. Gochhayat, and S. C. Das, Factors
Controlling Precipitation of Finer Size Alumina Trihydrate, Int. J. Miner. Process., 65, 10924 (2002).
22
J. A. Counter, J. Addai-Mensah, and J. Ralston, The Formation of Al(OH)3
Crystals from Supersaturated Sodium Aluminate Solutions Revealed by Cryovitrication Transmission Electron Microscopy, Colloids Surf. A: Physicochem.
Eng. Aspects, 154, 38998 (1999).
23
M. E. Essington, Soil and Water Chemistry: An Integrative Approach. CRC
Press, Boca Raton, 2004.
24
G. E. Jr. Brown, How Minerals React with Water, Science, 294, 6770
(2001).
25
G. E. Jr. Brown, T. P. Trainor, and A. M. Chaka, Geochemistry of Mineral
Surfaces and Factors that Affect their Chemical Reactivity; Pages 457509 in
Chemical Bonding at Surfaces and Interfaces, Edited by A. Nilsson, L. G. M. Pettersson, and J. Norskov. Elsevier, New York, 2007.
26
I. Levin and D. Brandon, Metastable Alumina Polymorphs: Crystal Structures and Transition Sequences, J. Am. Ceram. Soc., 81 [8] 19952012 (1998).
27
L. Bragg and G. F. Claringbull, Crystal Structures of Minerals. G. Bell and
Sons Ltd., London, 1965.
28
M. Mozloumi, R. Khalifehzadeh, S. K. Sadrnezhaad, and H. Arami, Alumina Nanopowder Production from Synthetic Bayer Liquor, J. Am. Ceram.
Soc., 89 [12] 36547 (2006).
29
W. H. Gitzen, Alumina as a Ceramic Material. American Ceramic Society,
Westerville, OH, 1970.
30
C. Legros, C. Carry, P. Bowen, and H. Hofmann, Sintering of a Transition
Alumina: Effects of Phase Transformation, Powder Characteristics and Thermal
Cycle, J. Eur. Ceram. Soc., 19, 196778 (1999).
31
P. Bowen and C. Carry, From Powders to Sintered Pieces: Forming, Transformations and Sintering of Nanostructured Ceramic Oxides, Powder Tech., 128,
24855 (2002).
32
P. Bowen, C. Carry, D. Luxemburg, and H. Hofmann, Colloidal Processing
and Sintering of Nanosized Transition Aluminas, Powder Tech., 157, 1007
(2005).
33
C. F. Baes and R. E. Mesmer, The Hydrolysis of Cations. Robert E. Krieger
Publishing Co., Malabar, FL, 1986.
34
B. A. Dempsey, Coagulant Characteristics and Reactions; Pages 524 in
Interface Science in Drinking Water Treatment, Theory and Applications, Edited by
G. Newcombe and D. Dixon. Elsevier, Amsterdam, 2006.
35
J. Duan and J. Gregory, Coagulation by Hydrolyzing Metal Salts, Advances
in Colloid and Interface Science, 100102, 475502 (2003).
36
T. W. Swaddle, J. Rosenqvist, P. Yu, E. Bylaska, B. L. Phillips, and W. H.
Casey, Kinetic Evidence for Five-Coordination in AlOH(aq)[2]1 Ion, Science,
308, 14503 (2005).
37
W. H. Casey, Large Aqueous Aluminum Hydroxide Molecules, Chemical
Reviews, 106 [1] 116 (2006).
38
V. A. Pokrovskii and H. C. Helgeson, Thermodynamic Properties of Aqueous Species and the Solubilities of Minerals at High Pressures and Temperatures:
The System Al2O3-H2O-NaCl, American Journal of Science, 295, 1255342
(1995).
39
E. L. Shock, D. C. Sassani, M. Willis, and D. A. Sverjensky, Inorganic
Species in Geologic Fluids Correlations among Standard Molal Thermodynamic
Properties of Aqueous Ions and Hydroxide Complexes, Geochim. Cosmochim.
Acta, 61 [5] 90750 (1997).
40
J.-Q. Jiang and N. J. D. Graham, Pre-Polymerised Inorganic Coagulants and
Phosphorus Removal by CoagulationA Review, Water SA, 24 [3] 23744
(1998).
41
M. Gautier, G. Fenaud, L. P. Van, B. Villette, M. Pollak, N. Thromat, F.
Jollet, and J.-P. Durand, Alpha-Al2O3 (0001) Surfaces: Atomic and Electronic
Structure, J. Am. Ceram. Soc., 77 [2] 32334 (1994).
42
I. Manassidis and M. J. Gillan, Structure and Energetics of Alumina Surfaces
Calculated form First Principles, J. Am. Ceram. Soc., 77 [2] 3358 (1994).
43
E. A. Soares, M. A. Van Hove, C. F. Walters, and K. F. McCarty, Structure
of the Alpha-Al2O3 (0001) Al Termination and Evidence for Surface from LowEnergy Electron Diffraction: Anomalously Large Thermal Vibrations, Phys. Rev.
B, 65, 195405 (2002).
44
P. Liu, T. Kendelewicz, G. E. Jr. Brown, E. J. Nelson, and S. A. Chambers,
Reaction of Water with a-A12O3 and a-Fe2O3 (0001) Surfaces: Synchrotron Xray Photoemission Studies, Surf. Sci., 417 [1] 5365 (1998).
45
P. J. Eng, T. P. Trainor, G. E. Jr. Brown, G. A. Waychunas, M. Newville, S.
R. Sutton, and M. L. Rivers, Structure of the Hydrated Alpha-Al2O3 (0001)
Surface, Science, 288 [5468] 102933 (2000).
46
J. Toofan and P. R. Watson, The Termination of the Alpha-Al2O3 (0001)
Surface: A LEED Crystallography Determination, Surf. Sci., 401, 1627 (1998).

47

Vol. 90, No. 11

M. A. Henderson, The Interaction of Water with Solid Surfaces Fundamental Aspects Revisited, Surf. Sci. Rep., 46, 1308 (2002).
48
U. Diebold, The Surface Science of Titanium Dioxide, Surf. Sci. Rep., 48,
53229 (2003).
49
G. V. Franks and L. Meagher, The Isoelectric Points of Sapphire Crystals
and Alpha-Alumina Powder, Coll. Surf. A, 214, 99110 (2003).
50
A. A. Tsyganenko and P. P. Mardilovich, Structure of Alumina Surfaces,
J. Chem. Soc., Faraday Trans., 92 [23] 484352 (1996).
51
T. Shirai, J. W. Li, K. Matsumaru, C. Ishizaki, and K. Ishizaki, Surface
Hydration States of Commercial High Purity Alpha-Al2O3 Powders Evaluated by
Temperature Programmed Desorption Mass Spectrometry and Disuse Reectance
Infrared Fourier Transform Spectroscopy, Sci. Tech. Adv. Mater., 6, 1238 (2005).
52
T. Shirai, C. Ishizaki, and K. Ishizaki, Effects of Manufacturing Processes on
Hydration Ability of High Purity Alpha-Al2O3 Powders, J. Ceram. Soc., Japan,
114 [3] 2869 (2006).
53
T. H. Ballinger and J. T. Jr. Yates, IR Spectroscopic Detection of Lewis Acid
Sites on Al2O3 Using Adsorbed CO. Correlation with Al-OH Group Removal,
Langmuir, 7, 30415 (1991).
54
R. O. James, Characterization of Colloids in Aqueous Systems; Pages 349
407 in Ceramic Powder Science. Advances in Ceramics, Vol. 21, Edited by G. L.
Messing, K. S. Mazdiyasni, J. W. McCauley, and R. A. Haber. American Ceramic
Society, Westerville, 1987.
55
R. J. Hunter, Foundations of Colloid Science, 2nd Ed, Oxford University Press,
Oxford, UK, 2001.
56
W. B. Russel, D. A. Saville, and W. R. Schowalter, Colloidal Dispersions.
Cambridge University Press, Cambridge, 1989.
57
D. A. Sverjensky, Zero-Pointof-Charge Prediction from Crystal Chemistry
and Salvation Theory, Geochim. Cosmochim. Acta, 58, 31239 (1994).
58
T. Hiemstra, W. H. van Riemsdijk, and G. H. Bolt, Multisite Proton Adsorption Modeling at the Solid/Solution Interface of (hydr)oxides: A New
Approach. I. Model Description and Evaluation of Intrinsic Reaction Constants,
J. Colloid Interface Sci., 133 [1] 91104 (1989).
59
T. Hiemstra, J. C. M. de Wit, and W. H. van Riemsdijk, Multisite Proton
Adsorption Modeling at the Solid/Solution Interface of (Hydr)Oxides: A New
Approach. II. Application to Various Important (Hydr) Oxides, J. Colloid Interface Sci., 133 [1] 10517 (1989).
60
T. Hiemstra, P. Venema, and W. H. van Riemsdijk, Intrinsic Proton Afnity
of Reactive Surface Groups of Metal (Hydr)Oxides: The Bond Valance Principle,
J. Colloid Interface Sci., 184 [2] 68092 (1996).
61
B. R. Bickmore, C. J. Tadanier, K. M. Rosso, W. D. Monn, and D. L. Eggett,
Bond-Valence Methods for pKa Prediction: Critical Reanalysis and a New Approach, Geochim. Cosmochim. Acta, 68, 202542 (2004).
62
J. N. Israelachvili, Intermolecular and Surface Forces, 2nd Ed, Academic Press,
London, UK, 1992.
63
D. B. Hough and L. R. White, The Calculation of Hamaker Constants from
Lifshitz Theory with Applications to Wetting Phenomena, Adv. Colloid Interface
Sci., 14, 314 (1980).
64
L. Bergstrom, A. Meurk, H. Arwin, and D. J. Rowcliffe, Estimation of
Hamaker Constants for Ceramic Materials from Optical Data using Lifshiz Theory, J. Am. Ceram. Soc., 79 [2] 33948 (1996).
65
L. Bergstrom, Hamaker Constants of Inorganic Materials, Adv. Colloid
Interface Sci., 70, 12569 (1997).
66
H. D. Ackler, R. H. French, and Y.-M. Chiang, Comparison of Hamaker
Constants for Ceramic Systems with Intervening Vacuum or Water: From Force
Laws and Physical Properties, J. Colloid Interface Sci., 179, 4609 (1996).
67
R. H. French, Origins and Applications of London Dispersion Forces and
Hamaker Constants in Ceramics, J. Am. Ceram. Soc., 83 [9] 211746 (2000).
68
M. Kosmulski, Chemical Properties of Material Surfaces. Dekker, New York,
2001.
69
M. Kosmulski, pH-Dependent Surface Charging and Points of Zero
Charge, J. Colloid Interface Sci., 253 [1] 7787 (2002).
70
M. Kosmulski, pH-Dependent Surface Charging and Points of Zero Charge
II. Update, J. Colloid Interface Sci., 275 [1] 2142 (2004).
71
M. Kosmulski, pH-Dependent Surface Charging and Points of Zero Charge
III. Update, J. Colloid Interface Sci., 298 [2] 73041 (2006).
72
G. A. Parks, Isoelectric Points of Solid Oxides Solid Hydroxides and Aqueous Hydroxo Complex Systems, Chem. Rev., 65 [2] 17798 (1965).
73
J. A. Yopps and D. W. Fuerstenau, The Zero Point of Charge of AlphaAlumina, J. Colloid Sci., 19, 6171 (1964).
74
S. G. Dick, D. W. Fuerstenau, and T. W. Healy, Adsorption of Alkylbenzene
Sulfonate (A. B. S.) Surfactants at the Alumina-Water Interface, J. Colloid Interface Sci., 37 [3] 595602 (1971).
75
M. Hashiba, H. Okamoto, Y. Nurishi, and K. Hiramatsu, The Zeta-Potential Measurement for Concentrated Aqueous Suspension by Improved Electrophoretic Mass Transport Apparatus Application to Al2O3, ZrO3 and SiC
Suspensions, J. Mater. Sci., 23 [8] 28936 (1988).
76
B. V. Velamakanni and F. F. Lange, Effect of Interparticle Potentials and
Sedimentation on Particle Packing Density of Bimodal Particle Distributions during Pressure Filtration, J. Am. Ceram. Soc., 74 [1] 16672 (1991).
77
K. F. Hayes, G. Redden, W. Ela, and J. O. Leckie, Surface Complexation
Models: An Evaluation of Model Parameter Estimation using Fiteql and Oxide
Mineral Titration Data, J. Colloid Interface Sci., 142 [2] 44869 (1991).
78
S. Veeramasuneni, M. R. Yalamanchili, and J. D. Miller, Measurement of
Interaction Forces between Silica and Alpha-Alumina by Atomic Force Microscopy, J. Colloid Interface Sci., 184 [2] 594600 (1996).
79
V. Ramakrishnan, P. Pradip, and S. G. Malghan, Yield Stress of Alumina
Zirconia Suspensions, J. Am. Ceram. Soc., 79 [10] 256776 (1996).
80
G. V. Franks and F. F. Lange, Mechanical Behavior of Saturated, Consolidated, Alumina Powder Compacts: Effect of Particle Size and Morphology on the
Plastic-to-Brittle Transition, Colloids Surf. A, 146, 517 (1999).

November 2007
81

Charging Behavior at the AluminaWater Interface and Implications

Q. Yang and T. Troczynski, Dispersion of Alumina and Silicon Carbide


Powders in Alumina Sol, J. Am. Ceram. Soc., 82 [7] 192830 (1999).
82
A. L. Costa, C. Galassi, and R. Greenwood, Alpha-Alumina-H2O Interface
Analysis by Electroacoustic Measurements, J. Colloid Interface Sci., 212 [2] 3506
(1999).
83
S. B. Johnson, P. J. Scales, and T. W. Healy, The Binding of Monovalent
Electrolyte Ions on Alpha-Alumina. I. Electroacoustic Studies at High Electrolyte
Concentrations, Langmuir, 15 [8] 283643 (1999).
84
V. A. Hackley, J. Patton, L.-S. H. Lum, R. J. Wasche, M. Naito, H. Abe, Y.
Hotta, and H. Pendse, Analysis of the Isoelectric Point in Moderately Concentrated Alumina Suspensions using Electroacoustic and Streaming Potential Methods, J. Dispersion Sci. Tech., 23 [5] 60117 (2002).
85
R. Wasche, M. Naito, and V. A. Hackley, Experimental Study on Zeta Potential and Streaming Potential of Advanced Ceramic Powders, Powder Tech.,
123, 27581 (2002).
86
T. Hiemstra, H. Yong, and W. H. van Riemsdijk, Interfacial Charging Phenomena of Aluminum (Hydr)oxides, Langmuir, 15, 594255 (1999).
87
J. Rosenqvist, P. Persson, and S. Sjoberg, Protonation and Charging of
Nanosized Gibbsite (a-Al(OH)3) Particles in Aqueous Suspension, Langmuir, 18,
4598604 (2002).
88
M.-C. Jodin, F. Gaboriaud, and B. Humbert, Limitations of Potentiometric
Studies to Determine the Surface Charge of Gibbsite g-Al(OH)3 Particles,
J. Colloids Interface Sci., 287, 5819 (2005).
89
R. G. Horn, D. R. Clarke, and M. T. Clarkson, Direct Measurement of
Surface Forces between Sapphire Crystals in Aqueous Solutions, J. Mater. Res.,
3 [3] 4136 (1988).
90
W. A. Ducker, Z. Xu, D. R. Clarke, and J. N. Israelachvili, The Forces between Sapphire Surfaces in Salt Solutions: Non-DLVO Forces and the Implications for Colloid Processing, J. Am. Ceram. Soc., 77 [2] 43743 (1994).
91
I. Larson, C. J. Drummond, D. C. Y. Chan, and F. Grieser, Direct Force
Measurements between Silica and Alumina, Langmuir, 13 [7] 210912 (1997).
92
A. G. Stack, S. R. Higgins, and C. M. Eggleston, Point of Zero Charge of a
Corundum-Water Interface Probed with Optical Second Harmonic Generation
(SHG) and Atomic Force Microscopy (AFM): New Approaches to Oxide Surface
Charge, Geochim. Cosmochim. Acta, 65 [18] 30552063 (2001).
93
A. G. Stack, S. R. Higgins, and C. M. Eggleston, Response to Comment on
Point of Zero Charge of a Corundum-Water Interface Probed with Optical Second Harmonic Generation (SHG) and Atomic Force Microscopy (AFM): New
Approaches to Oxide Surface Charge, Geochim. Cosmochim. Acta, 67 [2] 3212
(2003).
94
L. Meagher, G. Maurdev, and M. L. Gee, Interaction Forces between a Bare
Silica Surface and an Alpha-Alumina Surface Bearing Adsorbed Polyelectrolyte
and Surfactant, Langmuir, 18 [7] 264957 (2002).
95
R. J. Kershner, J. W. Bullard, and M. J. Cima, Potential Orientation Dependence of Sapphire Substrates, Langmuir, 20 [10] 41018 (2004).
96
J. P. Fitts, X. M. Shang, G. W. Flynn, T. F. Heinz, and K. B. Eisenthal,
Electrostatic Surface Charge at Aqueous/Alpha Alumina Single-Crystal Interfaces as Probed by Second Harmonic Generation, J. Phys. Chem. B, 109, 79816
(2005).
97
A. Tulpar, D. B. Henderson, M. Mao, B. Caba, R.M Davis, K. E. Van Cott,
and W. A. Ducker, Unnatural Proteins for the Control of Surface Forces, Langmuir, 21 [4] 1497506 (2005).
98
J. Lutzenkirchen Karlsruhe, personal communication, 2007.
99
Y. Gan and G. V. Franks, Charging Behavior of the Gibbsite Basal (001)
Surface in NaCl Solution Investigated by AFM Colloidal Probe Technique, Langmuir, 22, 608792 (2006).
100
T. P. Trainor, P. J. Eng, G. E. Jr. Brown, I. K. Robinson, and M. DeSantis,
Crystal Truncation Rod Diffraction Study of the Alpha-Al2O3 (1102) Surface,
Surf. Sci., 496, 23850 (2002).
101
J. G. Catalano, C. Park, Z. Zhang, and P. Fenter, Termination and Water
Adsorption at the Alpha-Al2O3 (012) Aqueous Solution Interface, Langmuir,
22, 466873 (2006).
102
K. S. Tanwar, J. G. Catalano, S. C. Petitto, S. K. Ghose, P. J. Eng, and T. P.
Trainor, Hydrated Alpha Fe2O3 (1102) Surface Structure: Role of Surface Preparation, Surf. Sci., 601, L59L64 (2007).
103
D. N. Furlong, P. A. Freeman, and A. C. M. Lau, The Adsorption of Soluble Silica at SolidAqueous Solution Interfaces I. Leaching from Glass and
Electrokinetic Study, J. Colloid Interface Sci., 80, 2031 (1981).
104
J. W. Elam, C. E. Nelson, M. A. Cameron, M. A. Tolbert, and S. M. George,
Adsorption of H2O on a Single-Crystal Alpha-Al2O3 (0001) Surface, J. Phys.
Chem. B, 102, 700815 (1998).
105
C. Barth and M. Reichling, Imaging the Atomic Arrangements on the High
Temperature Reconstructed Alpha-Al2O3 (0001) Surface, Nature, 414, 547
(2001).
106
K. C. Hass, W. F. Schneider, A. Curioni, and W. Andreoni, The Chemistry
of Water on Alumina Surfaces: Reaction Dynamics from First Principles, Science, 282 [5387] 2658 (1998).
107
K. C. Hass, W. F. Schneider, A. Curioni, and W. Andreoni, First-Principles
Molecular Dynamics Simulations of H2O on Alpha-Al2O3 (0001), J. Phys. Chem.
B, 104 [23] 552740 (2000).
108
Z. Lodziana, J. K. Norskov, and P. J. Stolze, The Stability of the
Hydroxylated (0001) Surface of Alpha-Al2O3, J. Chem. Phys., 118, 1117988
(2003).
109
H. A. Al-Abadleh and V. H. Grassian, FT-IR Study of Water Adsorption
on Aluminum Oxide Surfaces, Langmuir, 19, 3417 (2003).
110
Y. Gan, E. J. Wanless, and G. V. Franks, Lattice-Resolution Imaging of the
Sapphire (0001) Surface in Air by AFM, Surf. Sci., 601 [4] 106471 (2007).
111
S. Steinberg, W. Ducker, G. Vigil, C. Hyukjin, C. Frank, M. Z. Tseng, D. R.
Clarke, and J. N. Israelachvili, Van der Waals Epitaxial Growth of AlphaAlumina Nanocrystals on Mica, Science, 260, 6569 (1993).

112

3387

Y. Gan and G. V. Franks, High Resolution AFM Images of the SingleCrystal Alpha-Al2O3 (0001) Surface in Water, J. Phys. Chem. B., 109 [25] 12474
9 (2005).
113
S. Lloyd, S. M. Thurgate, R. M. Cornell, and G. M. Parkinson, Atomic
Force Microscopy of Gibbsite, Appl. Surf. Sci., 135 [1] 17882 (1998).
114
C. Contescu, J. Jagiello, and J. A. Schwarz, Heterogeneity of Proton Binding Sites at the Oxide/Solution Interface, Langmuir, 9 [7] 175465 (1993).
115
S. C. Mitchell An Improved MUSIC Model for Gibbsite, Master Thesis,
Brigham Young University, USA, 2005.
116
B. R. Bickmore, K. M. Rosso, C. J. Tadanier, E. J. Bylaska, and D. Doud,
Bond-Valence Methods for pKa Prediction. II. Bond-Valence, Electrostatic, Molecular Geometry, and Solvation Effects, Geochim. Cosmochim. Acta, 70 [16]
405771 (2006).
117
G. J. Fleer, M. A. Cohen Stuart, J. M. H. M. Scheutjens, T. Cosgrove, and B.
Vincent, Polymers at Interfaces. Chapman and Hall, London, 1993.
118
Th. Rabung, D. Schild, H. Geckeis, R. Klene, and Th. Fanghanel, Cm(III)
Sorption Onto Sapphire (Alpha-Al2O3) Single Crystals, J. Phys. Chem. B., 108
[44] 171605 (2004).
119
Th. Stumpf, Th. Rabung, R. Klenze, and J. I. Kim, Spectroscopic Study of
Cm(III) Sorption Onto Gamma Alumina, J. Colloid Interface Sci., 238, 21924
(2001).
120
K. E. Bremmell, unpublished experiments, 2000.
121
R. G. Horn, Surface Forces and their Action in Ceramic Materials, J. Am.
Ceram. Soc., 73, 111735 (1990).
122
J. Cesarano III, I. A. Aksay, and A. Bleier, Stability of Aqueous Alpha
Al2O3 Suspensions with Poly(Methacrylic Acid) Polyelectrolyte, J. Am. Ceram.
Soc., 71 [4] 2505 (1988).
123
J. Cesarano III and I. A. Aksay, Processing of Highly Concentrated Aqueous Alpha Alumina Suspensions Stabilized with Polyelectrolytes, J. Am. Ceram.
Soc., 71 [12] 10627 (1988).
124
V. A. Hackely, Colloidal Processing of Silicon Nitride with Poly (Acrylic
Acid): I, Adsorption and Electrostatic Interactions, J. Am. Ceram. Soc., 80 [9]
23152 (1997).
125
V. A. Hackely, Colloidal Processing of Silicon Nitride With Poly (Acrylic
Acid): I, Rheological Properties, J. Am. Ceram. Soc., 81 [9] 24218 (1998).
126
W. M. Carty and U. Senapati, Porcelain-Raw Materials, Processing,
Phase Evolution and Mechanical Behavior, J. Am. Ceram. Soc., 81 [1] 320
(1998).
127
D. Hotza and P. Greil, Review: Aqueous Tape Casting of Ceramic Powders, Mats. Sci. Eng. A, 202, 20617 (1995).
128
P. Panya, O. Aquero, G. V. Franks, and E. J. Wanless, Dispersion Stability
of a Ceramic Glaze Achieved Through Ionic Surfactant Adsorption, J. Colloid
Interface Sci., 279, 2335 (2004).
129
P. Panya, O. Aquero, E. J. Wanless, and G. V. Franks, The Effect of Ionic
Surfactant Adsorption on the Rheology of Ceramic Glaze Suspensions, J. Am.
Ceram. Soc., 88 [3] 5406 (2005).
130
R. Moreno, The Role of Slip Additives in Tape Casting Technology: Part
1Solvents and Dispersants, Am. Ceram. Soc. Bull., 71, 152131 (1992).
131
J. M. F. Ferreira and H. M. M. Diz, Effect of Solids Loading on Slip Casting Performance of Silicon Carbide Slurries, J. Am. Ceram. Soc., 82, 19932000
(1999).
132
L. Bergstrom, K. Shinozaki, H. Tomiyma, and N. Mizutani, Colloidal Processing of a Very Fine BaTiO3 PowderEffect of Particle Interactions on the
Suspension Properties, Consolidation and Sintering Behavior, J. Am. Ceram.
Soc., 80 [2] 291300 (1997).
133
M. Colic, G. Franks, M. Fisher, and F. Lange, Chemisorption of Organofunctional Silanes on Silicon Nitride for Improved Aqueous Processing, J. Am.
Ceram. Soc., 81, 215763 (1998).
134
E. P. Luther, J. A. Yanez, G. V. Franks, and F. F. Lange, Effect of Ammonium Citrate on the Rheology and Particle Packing of Alumina Slurries,
J. Am. Ceram. Soc., 78, 1495500 (1995).
135
P. C. Hidber, T. J. Graule, and L. J. Gauckler, Citric AcidA
Dispersant for Aqueous Alumina Suspensions, J. Am. Ceram. Soc., 79, 1857
67 (1996).
136
Y. K. Leong, P. J. Scales, T. W. Healy, D. V. Boger, and R. Buscall, Rheological Evidence of Adsorbate Mediated Short Range Steric Forces in Concentrated Dispersions, J. Chem. Soc. Faraday Trans., 89, 24738 (1993).
137
C. H. Schilling, M. Sikora, P. Tomasik, C. Li, and V. Garcia, Rheology of
Alumina Nanoparticle Suspensions Effect of Lower Saccharides and Sugar Alcohols, J. Euro. Ceram. Soc., 22, 91721 (2002).
138
M. Rhodes, Introduction to Particle Technology, 2nd Ed, Wiley, Chichester,
UK, 2007.
139
F. F. Lange and K. T. Miller, Pressure Filtration: Consolidation Kinetics
and Mechanics, Am. Ceram. Soc. Bull., 66, 1498504 (1987).
140
L. Bergstrom, C. H. Schilling, and I. A. Aksay, Consolidation Behavior of Flocculated Alumina Suspensions, J. Am. Ceram. Soc., 75, 330514
(1992).
141
G. V. Franks and F. F. Lange, Plastic-to-Brittle Transition of
Saturated, Alumina Powder Compacts, J. Am. Ceram. Soc., 79 [12] 31618
(1996).
142
L. Bergstrom, Rheology of Concentrated Suspensions; pp. 193244 in Surface and Colloid Chemistry in Advanced Ceramic Processing. Surfactant Science
Series, Vol. 51, Edited by R. J. Pugh and L. Bergstrom. Marcel Dekker, New
York, 1994.
143
S. Liu and J. H. Masliyah, Rheology of Suspensions; pp. 10776 in Suspensions, Fundamentals and Applications in the Petroleum Industry. Advances in
Chemistry Series, Vol. 251, Edited by L. Schramm, The American Chemical
Society, 1996.
144
J. R. Taylor and A. C. Bull, pp. 2067 in Ceramics Glaze Technology.
Pergamon Press, UK, 1986.

3388

Journal of the American Ceramic SocietyFranks and Gan

145

Vol. 90, No. 11

154

F. H. Norton, Elements of Ceramics, 2nd Ed, Addison-Wesley, Reading, MA,


1974.
146
C. W. Parmelee, Ceramic Glazes, 3rd Ed, revised by C. G. Harman. Cahners
Books, Boston, MA, 1973.
147
A. C. Young, O. O. Omatete, M. A. Janney, and P. A. Menchhofer, Gelcasting of Alumina, J. Am. Ceram. Soc., 74, 61218 (1991).
148
S. B. Johnson, D. E. Dunstan, and G. V. Franks, Rheology of Crosslinked
Chitosan/Alumina Suspensions Used for a New GelCasting Process, J. Am.
Ceram Soc., 85, 1699705 (2002).
149
S. Jailani, G. V. Franks, and T. W. Healy, Zeta Potential of Nanoparticle
Suspensions: Effect of Electrolyte Concentration, Particle Size, and Volume Fraction, J. Am. Ceram. Soc., (2007) submitted.
150
D. H. Napper, Polymeric stabilization of colloidal dispersions. Academic
Press, London, 1983.
151
J. Gregory, Effects of Polymers on Colloid Stability; pp. 10130 in The
Scientic Basis of Flocculation, Edited by K. J. Ives. Sijthoff and Noordhoff, The
Netherlands, 1978.
152
J. Gregory, Particles in Water. IWA Publishers, London, 2006.
153
T. W. Healy, Principles of Polymer Flocculation; pp. 120 in Polymer
Flocculation Principles and Applications. Edited by T. W. Healy, Royal Australian
Chemical Institute, Melbourne, 1973.

B. A. Bolto, p. 65 in Polymeric Flocculants in Water and Wastewater


Treatment in Modern Techniques in Water and Wastewater Treatment,
Edited by L. O. Kolarik, and A. J. Priestley. CSIRO Publishing, Melbourne,
1995.
155
T. W. Healy and V. K. La Mer, Energetics of Flocculation and Redispersion
by Polymers, J. Colloid Sci., 19, 32332 (1964).
156
J. C. Chang, B. V. Velamakanni, F. F. Lange, and D. S. Pearson,
Centrifugal Consolidation of Alumina/Zirconia Composite Slurries Vs.
Interparticle Potentials: Particle Packing and Mass Segregation, J. Am. Ceram.
Soc., 74 [9] 22014 (1991).
157
T. Kramer and F. F. Lange, Rheology and Particle Packing of Chem- and
Phys-Adsorbed, Alkylated Silicon Nitride Powders, J. Am. Ceram. Soc., 77 [4]
9228 (1994).
158
F. F. Lange, Colloidal Processing of Powder for Reliable Processing, Curr.
Opin. Solid State. Mater. Sci., 3, 496500 (1998).
159
G. V. Franks, M. Colic, M. L. Fisher, and F. F. Lange, Plastic-to-Brittle
Transition of Consolidated Bodies: Effect of Counterion Size, J. Colloid. Interface Sci., 193, 96103 (1997).
160
A. R. Studart, E. Amstad, M. Antoni, and L. J. Gauckler, Rheology of
Concentrated Suspensions containing Weakly Attractive Alumina Particles,
J. Am. Ceram. Soc., 89 [8] 241825 (2006).
&

George Vincent
Franks, Associate
Professor, completed his undergraduate degree
at the Massachusetts Institute of
Technology
(MIT) in Materials Science and
Engineering
in
1985. He worked
for 7 years in the ceramic processing industry as a process development engineer for Norton Company and Ceramic Process
Systems Incorporated. His industrial work focused mainly on
near net shape forming of ceramic green bodies and nonoxide
ceramic ring. He then completed a PhD at the University of
California at Santa Barbara in Materials in 1997 under the
guidance of Prof. Fred. Lange. He then went to Australia as a
post doctoral researcher at the University of Melbourne where
his research concentrated on surface chemistry effects in suspension rheology. Between 1999 and 2005, George taught chemical engineering at the University of Newcastle, Australia.
During this period he developed a novel gelcasting chemistry,
investigated relationships between aggregate properties and suspension rheology and developed stimulant responsive occulants for mineral tailing dewatering. He is now Associate
Professor in the Department of Chemical and Biomolecular Engineering at the University of Melbourne, Australia. He is a
member of the Particulate Fluids Processing Centre and the
Australia Mineral Science Research Institute. In addition to a
longstanding interest in the alumina water interface, his research
interests include advanced ceramics powder processing and shape
forming, mineral processing (particularly occulation), colloid
and surface chemistry, ion specic effects and suspension rheology. He is Associate Editor of the Journal of the American Ceramic Society and referee for a number of other learned journals.
He has 50 peer reviewed journal publications and three patents.

Prof. Yang Gan is an appointed


professor in the Department of
Applied Chemistry at Harbin Institute of Technology, China. He received his PhD in materials science
in 2001 from the Institute of Metal
Research (China), where his thesis
research was involved with the
study of mechanical property
of nanomaterials. From 2001 to
2003, he was a postdoctoral research fellow at the University of
Science and Technology Beijing, studying the surface defects
using AFM and STM. From 2004 to 2007, he was appointed
a University Research Fellow at the Department of Chemical
Engineering and Chemistry at the University of Newcastle,
Australia. Before returning to China, he spent several months
at University of Melbourne as a visiting scholar. His current
research interests involve the surface structure and chemistry of
inorganic oxides at nano and atomic scale using AFM surface
force measurements and high-resolution imaging. He has been
writing articles on AFM application for a microscopy magazine
Microscopy Today. He is a reviewer for journals Langmuir,
Journal of Colloids and Interface Science and AIChE Journal.
He is also on the book review panel for Physical Sciences
Educational Reviews. He has been an active member of SPM
discussion forum SPM Digest since 2002.

Das könnte Ihnen auch gefallen