Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Colloids and Surfaces

A: Physicochemical and Engineering Aspects 161 (2000) 243 – 257


www.elsevier.nl/locate/colsurfa

Interaction of binders with dispersant stabilised alumina


suspensions
Asad U. Khan, Brian J. Briscoe, Paul F. Luckham *
Particle Technology Group, Department of Chemical Engineering and Chemical Technology, Imperial College of Science,
Technology and Medicine, London, UK

Abstract

The rheological response of selected aqueous alumina suspensions, stabilised with a polyelectrolyte or with an
organic polyvalent salt dispersant, and including poly(vinyl) alcohol (PVA) as a binder, are described in this study.
The polymer dispersant was composed of an ammonium salt of poly(methacrylate) and the organic polyvalent
compound was a sodium salt of an aromatic sulphate. The results show that the addition of PVA, without any
included dispersant does not significantly influence the rheology of the system. However, in the presence of a
dispersant the rheology is greatly affected. At a given concentration of the dispersant, the viscosity, storage and loss
moduli all increase, as the PVA concentration is increased. Also, for a given concentration of the PVA, it is observed
that the viscosity, storage and loss moduli values increase as the concentration of the dispersant is increased. It is
argued that at low PVA concentrations, an excess concentration of the unadsorbed dispersant causes flocculation of
the particles in the suspension by a reduction of the repulsive electrostatic (double layer) effect. In contrast, at higher
concentrations of the PVA the flocculation of the suspension is promoted via a depletion mechanism. © 2000 Elsevier
Science B.V. All rights reserved.

Keywords: PVA; Suspensions; Binders

1. Introduction In this paper, we utilise these ideas to explore the


rheological characteristics of ceramic suspensions.
The rheological properties of concentrated sus- Organic binders are an essential component for
pensions are complex and much of the early work the effective processing of many commercial high
was phenomenological in nature. Hunter [1–3] performance ceramics. These binders are used to
and Goodwin [4 – 7] were really the first to relate provide sufficient strength to the body such that
rheologically determined parameters, the yield the green bodies can be moulded and retained in
value in the first instance and latter the elastic the desired shape without breaking or damage,
modulus, to the microstructure of the suspension. before and during sintering. Many ceramic form-
ing processes make use of a binder, including
* Corresponding author. Tel.: +44-171-594-5583; fax: + those of dry pressing, slip casting, tape casting,
44-171-594-5604. extrusion, roll forming, thick film printing, injec-
E-mail address: p.luckham01@ic.ac.uk (P.F. Luckham) tion and compression moulding.

0927-7757/00/$ - see front matter © 2000 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 7 - 7 7 5 7 ( 9 9 ) 0 0 3 7 4 - X
244 A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257

There are various organic substances that have The changes in the rheology of the binder/liquid
been used, or categorised, as potentially useful solution directly affect the behaviour of the sus-
binders for ceramics and many useful compila- pension formed upon the addition of particulates
tions and descriptions of these different organic to the solution.
binders have been published [8 – 17]. They include The usefulness of a given binder for a specific
poly(vinyl) alcohol, cellulose based materials, nat- system depends very much upon the interactions
ural gums, starches, sodium and ammonium algi- developed between the different components, e.g.
nates etc. The use of polymer latex systems has the dispersant, the particles, etc. of the suspen-
also been reported recently [18]. sion. In this paper, the influence of a poly(vinyl)
The suitability of a binder for a given system alcohol (PVA) addition, in the presence of two
depends upon on a number of factors including different dispersants, upon the rheology of an
the particle/solvent interactions with a dispersant, alumina AES-11 aqueous suspension is reported.
the affinity with the processing liquid, the polarity This binder and these dispersants are frequently
and the interaction with the particulate material. used in commercial practice.
The ‘ideal’ binder should be compatible with the The amount of dispersant and PVA used in this
dispersant, perhaps also function as a stabiliser, paper is expressed on a dry weight of the powder
induce no interference with the solvent quality, basis (dwb), which means it is equivalent to the
function as a lubricant between the particles and wt./wt. basis of the anhydrous solid.
should not produce foaming on air entrainment. It has been shown, for various particulate sys-
Furthermore, an effective burnout profile without tems such as polymer latex particles and surfac-
the formation of a deleterious residue is also tant micelles, that when particles are dispersed in
essential. A low glass transition temperature and a a dispersion medium, there are at least four kinds
high mechanical strength to molecular weight ra- of interactions operating between the different
tio are also desirable. Naturally, cost is also a components of the suspensions. Van der Waals
significant factor. Inevitably, the binders used in attractive forces are the most ubiquitous. The van
practice do not have the properties of the ‘ideal’ der Waals attractive forces are very dominant for
binder; a compromise is required. Usually, differ- high surface area particulate materials, and cause
ent substances are mixed in order to obtain, or to aggregation of the particles. If the particulate
approximate to the desired properties. Naturally, material has an electrical charge induced, say
the mixing of the different components greatly caused by the adsorption of a charged dispersant,
complicates the rheology of the system. the aggregation of the particles in a suspension
The binders are introduced in a liquid phase, can be suppressed through the formation of an
except in the cases of injection and compression electrical double layer. Adsorption, or grafting, of
moulding, and are employed in the formulation of polymer molecules can also prevent aggregation
a particulate system. The liquid phase provides of the particles by the mechanism of steric stabili-
the vehicle that is necessary for the uniform dis- sation. The addition of a non-adsorbing polymer
persing of the binder molecules. The dry green to a stabilised suspension may cause depletion
strength in the cast body is developed by the flocculation, whereby the concentration of non-
evaporation of the supernatant liquid. The resid- adsorbing free polymer is effectively reduced in
ual binder remains in place within the body form- between the particles, and the higher concentra-
ing the necessary interparticle bridges required to tion outside the particle–particle approach zone,
provide a strong adhesion amongst the ceramic exerts an osmotic pressure causing the particles to
particles. flocculate; Russel et al. [19], Hunter [20], Napper
The addition of the organic binder often affects [21], Everett [22] and Fleer et al. [23]. Fig. 1 shows
the rheology of the liquid phase. An increase in schematically different effects which may cause
the viscosity, as well as a change in flow charac- instability in a stabilised suspension (central re-
teristics, from Newtonian (for water) to shear gion). The arrow direction shows the increasing of
thinning, are generally the main consequences. that specific effect.
A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257 245

Fig. 1. Different forces that cause instability in a stabilised suspension. The arrow direction indicates the increasing of the effect.

2. Experimental 61 000, with a degree of hydrolysis \ 98% (den-


sity=1290 kg m − 3). All the suspensions de-
scribed in this paper were composed of 40% by
2.1. Materials volume of alumina and deionised water was used
to prepare the suspensions. The concentration of
Alumina AES-11 (Mandoval, Surrey, UK), these substances (PVA+ dispersants) are ex-
99.8% pure, crystalline with a surface area of 8.14 pressed in ‘dwb’, i.e. on dry weight basis of the
m2 g − 1 (BET values), a mean particle size of 0.4 alumina powder. It means that if the alumina
mm was used as received. Two dispersant systems powder used is 100 g in the suspension, 1% dwb is
were studied, ‘Darvan C’, an ammonium salt of equal to 1 g substance in the suspension.
poly(methacrylate), (25% by wt. polymer in
water of MW 10 000 – 16 000) (R.T. Vanderbilt
Company, USA), and ‘Tiron’ (Fluke Chemicals), 2.2. Methods
which is sodium salt of 4 – 5 dihydroxy-1, 3-ben-
zene disulfonic acid (C6H8Na2O8S2, MW 332.2) All of the measurements were performed with a
also called as pyrocatechol-3, 5-disulfonic acid Bohlin VOR rheometer (Bohlin Rheologi, Lund,
disodium salt. The chemical structures of these Sweden). Before performing rheological charac-
two dispersants are given in references [24,25]. terisation, the suspensions were presheared at a
The PVA was ‘Mowiol 10-98’ (Halow Chemical, comparatively high shear rate of 960 s − 1 for
UK), and had an average molecular weight  3–4 min followed by at least a 2 min ‘rest’
246 A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257

period in order to attempt to provide a common


and consistent shear history for the systems
[24,25]. The viscosity at low shear rates was mea-
sured first, stepping up to higher shear rates,
covering the shear rate range of  10 − 1 to 103
s − 1. All of the experiments were performed at
25°C. Initially, for the dynamic measurements a
‘strain sweep’ measurement was performed in
which the imposed strain amplitude was varied
while the frequency was maintained constant at 1
Hz. The values of the G% and G%% moduli were
measured as a function of the strain amplitude in
order to identify the so-called ‘linear viscoelastic
region’ in which the viscoelastic properties of the Fig. 2. Viscosity of alumina AES-11 40% v/v suspensions
system were relatively independent of the applied against PVA concentration % dwb (dry weight of powder
strain (deviationB 5% in G% values). After identi- basis) at three different shear rates (SR)/(1.46, 14.6 and 146
fying the linear viscoelastic region (strain range), s − 1); no specific dispersant.
the imposed strain was fixed within this region
and the deformation frequency was varied (usu- 3.2. Dispersant without binder
ally from 0.05 to 10 Hz). This is termed here the
‘oscillatory measurement’. The specific results re- Addition of a dispersant to alumina suspen-
ported in this paper are for a frequency of 1 Hz. sions causes changes in the viscosity of the sus-
pensions. This effect is shown in Fig. 3 as the
viscosity at a shear rate of 1.46 s − 1 against the
3. Results dispersant concentration (see also [25]). The opti-
mum concentration, i.e. the concentration which
3.1. Binders without dispersant gives the lowest viscosity, is 1% dwb for ‘Darvan
C’ and 0.1% (or 0.125%) dwb for ‘Tiron’. Below
The suspensions prepared from the alumina this optimum concentration, the surface coverage
powder and water, without the addition of any of the alumina particles with the dispersant
dispersant or binder, were very flocculated and molecules is not complete, and the particles aggre-
dried very quickly and were therefore not suitable gate due to the domination of van der Waals
for the production of ‘good’ quality ceramic mate- attractive forces. Above this optimum concentra-
rials. The addition of external processing agents tion value there is an excess amount of dispersant
changes the interparticle interactions appreciably,
which changes the rheology of the resulting sus-
pensions [26]. PVA is one of the most frequently
used water soluble binders and it has been re-
ported [27] that PVA molecules adsorb weakly
onto the alumina particles (adsorbed concentra-
tion 0.1 mg m2). Fig. 2 reports the viscosity
against the PVA concentration at three different
shear rates of 1.46, 14.6 and 146 s − 1 for 40% (v/v)
alumina AES-11 suspensions. From the Figure, it
appears that, for the concentration range of the
PVA investigated, the viscosities of the suspen-
sions are virtually unchanged as the PVA concen- Fig. 3. Viscosity at a shear rate of 1.46 s − 1 against dispersant
tration is changed, at a given shear rate. concentration of alumina AES-11 40% v/v suspensions.
A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257 247

which is not adsorbed onto the particles, remains


in the aqueous phase, functions as an electrolyte,
hence reduces the range of electrical double layer
repulsion and, therefore, increase in the viscosity
results. The fact that the optimum concentration
corresponds to the monolayer coverage of the
particles was established using adsorption
isotherms [24,25]. Fig. 3 also shows the point of
critical coagulation concentration (c.c.c) for the
‘Tiron’ case. The c.c.c. was calculated using
DLVO theory (Eq. (1)) [28].
9.85 · 104o 3k 5T 5g 4
c.c.c.= (1)
Nae 6A 2z 6
Fig. 4. Relative viscosity against PVA concentration, for alu-
where Na is Avogadro’s constant, k is Boltz- mina AES-11 40% v/v suspensions stabilised with 1% dwb
mann’s constant, o is permitivity, z is the charge ‘Darvan C’, at three different shear rates (SR)/(1.46, 14.6 and
number of the ion, T is absolute temperature, A is 146 s − 1).
Hamaker’s constant, and e charge on an electron
and g is given by Eq. (2) in which 8d is the surface ogy of the suspensions significantly. Figs. 4 and 5
potential of the particles. represent the relative viscosity, hr (hr = ho, where
hs and ho are the viscosities of the suspension and
exp[ze8d/2kT] −1
g= (2) the continuous medium (PVA+water+ non-ad-
exp[ze8d/2kT] +1 sorbed dispersant), respectively) at three different
To calculate the c.c.c., the surface potential was shear rates (1.46, 14.6 and 146 s − 1) as a function
taken equal to 50 mV which is the measured value of the PVA concentration for the suspensions
of the zeta-potential at the suspension pH ( 9). containing respectively 1% dwb ‘Darvan C’ and
The zeta-potential was measured using an elec- 0.125% dwb ‘Tiron’. These concentration values
trophoresis experiment. Hamaker constant was of the ‘Darvan C’ and ‘Tiron’ dispersants are
taken equal to 6.7×10 − 20 J K − 1. Using these those which provide the lowest viscosity, at a
values, a c.c.c. of 35.9 × 10 − 3 mol dm − 3 was given shear rate, when incorporated in a 40% v/v
estimated which corresponds to a 0.225% dwb
This concentration is above the minimum viscos-
ity concentration of 0.125% dwb which would
corresponds to the maximum stabilisation part of
the dispersant. Adding more dispersant increases
the electrolyte concentration of the bulk and
would be expected to cause of aggregation. Thus
experimentally we estimate the c.c.c. to be above
0.125% dwb of the dispersant. The estimated
value of 0.225% dwb is in accord with this. ‘Tiron’
concentration (‘Tiron’ molecule was considered as
2:1 salt and Na+ ions as counter ions).

3.3. Dispersant – binder combinations


Fig. 5. Relative viscosity against PVA concentration, for alu-
In contrast to the PVA without dispersant case, mina AES-11 40% v/v suspensions stabilised with 0.125% dwb
the addition of PVA to a dispersant stabilised ‘Tiron’, at three different shear rates (SR)/(1.46, 14.6 and 146
alumina AES-11 suspension can change the rheol- s − 1).
248 A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257

suspension containing no added binder [24,25].


For both cases, at the lowest shear rate (1.46 s − 1)
the relative viscosity initially increases monotoni-
cally as the PVA concentration is increased from
0 to 0.5% dwb. Thereafter, the relative viscosities
at this shear rate remain virtually unchanged. The
increase in the relative viscosity is approximately
four times for ‘Darvan C’ and about an order of
magnitude for ‘Tiron’, as the PVA concentration
is increased from 0 to 0.5% dwb. At the shear rate
of 14.6 s − 1, the relative viscosity still increases
with the increase of the PVA concentration. How-
ever, the rate of the increase of the viscosity with
the increase of the PVA concentration is less than
that observed at the shear rate of 1.46 s − 1 for the Fig. 6. Relative viscosity against PVA concentration, for alu-
respective suspensions. The increase in the relative mina AES-11 40% v/v suspensions stabilised with different
viscosity at the former shear rate is 2 –3 times concentrations of ‘Darvan C’ (% dwb), at a shear rate 1.46
higher than the later shear rate when the PVA s − 1.
concentration is increased from 0 to 0.5% dwb.
There is a less significant change in the relative caused by an electrical double layer screening
viscosity at the shear rate of 146 s − 1 as the PVA effect, such as that which has been discussed
concentration is varied; the major effects are elsewhere [25]. The increase of the dispersant con-
noted at the lower shear rate. centration in the presence of PVA significantly
The relative viscosity of the alumina suspen- modifies the interactions between the alumina
sions as a function of the PVA concentration at a particles. This result is shown in Figs. 8 and 9 for
shear rate of 1.46 s − 1, is reported in Figs. 6 and ‘Darvan C’ and ‘ Tiron’, respectively. These are
7 for the different concentrations of the disper-
plots of the viscosity at a shear rate of 1.46 s − 1 as
sants ‘Darvan C’ and ‘Tiron’, respectively. For
a function of the dispersant concentration for the
the comparatively low concentrations of the dis-
various concentrations of PVA. The figures show
persants (e.g. for ‘Darvan C’ from 1 to 2% dwb
and for ‘Tiron’ from 0.10 to 0.50% dwb), the
relative viscosity generally increases gradually as
the PVA concentration is increased from 0 to
0.5% dwb, and thereafter levels off or even
slightly decreases. For the 4% dwb concentration
of the ‘Darvan C’ and 1% dwb for ‘Tiron’, the
relative viscosity does not significantly change as
the PVA concentration is increased from 0 to
0.5% dwb. However, increasing the PVA concen-
tration above 0.5% dwb slightly lowers the rela-
tive viscosity with the increase of the PVA
concentrations. As before, the influence of the
dispersants is similar.
If the concentration of the dispersant is in-
creased, in the absence of any binder, above the
optimum amount (i.e. the amount which gives the Fig. 7. Relative viscosity against PVA concentration, for alu-
lowest viscosity), the viscosity of the suspension mina AES-11 40% v/v suspensions stabilised with different
increases, due to the aggregation of the particles concentrations (% dwb) of ‘Tiron’, at a shear rate of 1.46 s − 1.
A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257 249

Fig. 8. Viscosity as a function of ‘Darvan C’ concentration (%


dwb) for different concentrations of PVA (% dwb), of alumina Fig. 10. Viscosity against PVA concentration (% dwb), for
AES-11 suspensions at a shear rate of 1.46 s − 1. alumina AES-11 40% v/v suspensions stabilised with different
concentrations of ‘Darvan C’ (% dwb), at a shear rate of 1.46
that, as the dispersant concentration is increased s − 1.
for the two dispersants, the viscosity of the alu-
mina suspensions simply increases, for a given higher concentration values of the dispersant the
PVA concentration. However, as the concentra- viscosities of the alumina suspensions are higher
tion of the PVA is increased, the rate of the than those for the lower dispersant concentration
increase in the viscosity with the increasing of the for a given PVA concentration. Figs. 12 and 13
dispersant concentration decreases. In Figs. 10 describe the normalised/relative viscosity, hn
and 11, for ‘Darvan C’ and ‘Tiron’ cases, respec- against the PVA concentration for each concen-
tively, the viscosities of the 40% alumina AES-11 tration of the dispersant (corresponding to Figs.
suspensions, at a shear rate of 1.46 s − 1 are plot- 10 and 11, respectively) at a shear rate 1.46 s − 1.
ted against the PVA concentration for the differ- This parameter corresponds to the viscosity of the
ent concentrations of dispersants. The viscosity alumina suspensions, divided by the viscosity of
values, at this shear rate, increase steadily as the the suspension having no PVA for the corre-
PVA concentration is increased. Also, for the

Fig. 9. Viscosity as a function of ‘Tiron’ concentration (% Fig. 11. Viscosity against PVA concentration (% dwb), for
dwb) for different concentrations of PVA (% dwb), of alumina alumina AES-11 40% v/v suspensions stabilised with different
AES-11 suspensions at a shear rate of 1.46 s − 1. concentrations of ‘Tiron’ (% dwb), at a shear rate of 1.46 s − 1.
250 A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257

Fig. 13. Normalised viscosity as a function of PVA concentra-


tion, for alumina AES-11 suspensions stabilised with different
Fig. 12. Normalised viscosity as a function of PVA concentra- concentrations of ‘Tiron’ (5 dwb), at a shear rate of 1.46 s − 1.
tion, for alumina AES-11 suspensions stabilised with different Inset is plot of same variables for different ranges and on
concentrations of ‘Darvan C’ (5 dwb), at a shear rate of 1.46 different scale.
s − 1.
The results of the oscillatory strain measure-
1 ments, at a frequency of 1 Hz., for the alumina
sponding dispersant concentration suspension.
AES-11 suspensions with 1% dwb ‘Darvan C’ and
The normalised/relative viscosity value, hn as a
0.125% dwb ‘Tiron’ are shown in Figs. 15 and 16.
function of the PVA concentration, increases al-
For the ‘Darvan C’ case, the storage modulus
most linearly. However, as the dispersant concen-
gradually increases as the PVA concentration is
tration is increased successively, the gradient of increased. This increase is particularly pro-
the line (normalised viscosity vs. PVA concentra- nounced for the lower concentrations of PVA.
tion) decreases. Fig. 14 is a plot of the relative The loss modulus is always less than the storage
viscosity against the PVA concentration, for the modulus. However, for ‘Tiron’ suspensions the
0.05 and 2% dwb ‘Tiron’ concentrations. These storage modulus is only less than the loss modulus
concentrations are respectively rather below and for the PVA concentrations less than 0.20% dwb.
much above compared, to the optimum concen- The storage modulus also increases relatively
tration (0.10 – 0.125% dwb) of ‘Tiron’ that steeply in the initial stages as the PVA concentra-
which provides the minimum viscosity. Also, plot-
ted in the same Figure are the data for the opti-
mum concentration (0.125% dwb) of ‘Tiron’ for
comparison. The Figure shows that with the in-
creasing of the PVA concentration there is first a
decrease in the relative viscosity followed by a
steady increase. The effect is more pronounced for
the 0.05% dwb ‘Tiron’ concentration. This viscos-
ity-PVA concentration trend for these ‘Tiron’
concentrations is in stark contrast to the optimum
concentration trend for the ‘Tiron’ system plotted
against the PVA concentration.

1
The normalised/relative viscosity hn = hp/hpo, where hp and
hpo are the viscosities of the suspension with and without PVA Fig. 14. Relative viscosity against PVA concentration, for
for a given dispersant concentration, respectively. Further- alumina AES-11 40% v/v suspensions stabilised with different
more, when the PVA concentration is zero, hp/hpo = 1. concentrations (% dwb) of ‘Tiron’, at a shear rate of 1.46 s − 1.
A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257 251

Fig. 17. G% against PVA concentration, for alumina AES-11


Fig. 15. G% and G%% against PVA concentration, for alumina
40% v/v suspensions stabilised with different concentrations of
AES-11 40% v/v suspensions stabilised with 1% dwb ‘Darvan
‘Darvan C’ (% dwb): ", G% 4%; , G% 2%; , G% 1.5%;
, G%
C’, at a frequency of 1 Hz.
0.125%, at a frequency of 1 Hz.

tion is increased and the slopes of the curves observed by other researchers [6], and the chang-
decreases as the PVA concentration is further ing shape of the curve (storage modulus versus
increased. This trend (but not the absolute values) volume fraction) was attributed to the changing
is similar to that of the viscosity at a shear rate of geometry of the fractal structure of the particles
1.46 s − 1 that is shown in Figs. 4 and 5 as the formed due to the flocculation.
relative viscosity against the PVA concentration It is seen in Figs. 17 and 18 that, for all the
curve. The loss modulus, G%%, and the dynamic adopted dispersant concentrations, the G% elastic
viscosity follow the same general trend as the G% modulus values increase with the increasing of the
modulus as the PVA concentration is varied. A PVA concentration. However, the slopes of G%
similar trend of the storage modulus against vol- against the PVA concentration curves decrease
ume fraction of the polystyrene suspensions was (especially for the PVA concentration between 0

Fig. 16. Dynamic viscosity, G%, G%% and dynamic viscosity Fig. 18. G% against PVA concentration, for alumina AES-11
against PVA concentration, for alumina AES-11 40% v/v 40% v/v suspensions stabilised with different concentrations of
suspensions stabilised with 0.125% dwb ‘Tiron’, at a frequency ‘Tiron’ (% dwb): , G% 0.1%; , G% 0.25%; ", G% 0.5%; , G%
of 1 Hz. 1%;
, G% 0.125%, at a frequency of 1 Hz.
252 A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257

Fig. 21. Storage modulus against PVA concentration, for


Fig. 19. G%% against PVA concentration, for alumina AES-11
alumina AES-11 40% v/v suspensions stabilised with different
40% v/v suspensions stabilised with different concentrations of
concentrations of ‘Tiron’, at a frequency of 1 Hz.
‘Darvan C’ (% dwb):
, 1%; , 1.5%; , 2%; ", 4%, at a
frequency of 1 Hz.

tion which is shown in Figs. 19 and 20 follows


and 0.5% dwb) as the dispersant concentration is
nearly the same trend as that observed in the G%
raised from 1 to 4% dwb for ‘Darvan C’ and from
against PVA profile for the various concentrations
0.1 to 1.0% dwb for ‘Tiron’. A similar trend was
of dispersant. There are a few exceptions, notably
observed in the computed viscosity values at a
for the 4% dwb ‘Darvan C’ and 0.05% dwb PVA,
shear rate of 1.46 s − 1 which is expressed as the
and 1% ‘Tiron’ and 0.2% PVA suspensions, for
relative viscosity against the PVA concentration
which there is a sudden decrease in the loss mod-
for the different concentrations of dispersant
ulus value. A similar type of decrease, but to a
(Figs. 6 and 7). At the higher PVA concentra-
lesser degree, was also observed, in the G% values.
tions, the G% moduli for all of the concentrations
The difference between the G% and the G%% values
of the dispersants approach each other. The loss
is almost equal to one order of magnitude
modulus, G%%, as a function of the PVA concentra-
throughout with the G%% values being always less
than the G% values, for the ‘Darvan C’ concentra-
tions of 1.5% dwb and above and for ‘Tiron’
concentrations greater than 0.125% dwb. In gen-
eral, the dynamic viscosity, at a frequency of 1
Hz., and the viscosity at the shear rate of 1.46 s − 1
are observed to be similar, for suspensions of the
similar compositions.
For suspensions containing 0.05 and 2% dwb
‘Tiron’, the storage modulus as a function of PVA
concentration is shown in Fig. 21. The storage
moduli data, for 0.125% dwb ‘Tiron’, are also
plotted in the same figure for the purpose of
comparison. The storage modulus against PVA
concentration curves follow the same general
Fig. 20. G%% against PVA concentration, for alumina AES-11
40% v/v suspensions stabilised with different concentrations of
trend as that shown by the relative viscosity
‘Tiron’ (% dwb):
, G%% 0.1%; , G%% 0.25%; , G%% 0.5%; ", against PVA concentration data for the same
G%% 1%; , G%% 0.125%, at a frequency of 1 Hz. concentrations of ‘Tiron’ (Fig. 14).
A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257 253

4. Discussion tion starting from  0.2% by volume. These


concentrations are comparable with those values
The increase in the viscosity of the dispersant reported here where we observe an increase in low
stabilised suspensions at the shear rate of 1.46 shear viscosity and G% at PVA concentration
s − 1, and the G%, G%% values as well as the dynamic 0.05% dwn ( 0.1% by volume (see Figs. 4, 5, 15
viscosity at the frequency of 1 Hz., with increasing and 16).
of the PVA concentration suggests that the addi- When the dispersant concentration in the alu-
tion of the PVA causes a flocculation of the mina AES-11 suspensions, without any PVA ad-
alumina stabilised particles with the 1% dwb dition, is increased beyond 1% dwb for ‘Darvan
‘Darvan C’ and 0.125% dwb ‘Tiron’ dispersant C’ and 0.125% dwb for ‘Tiron’ (the optimum
concentrations. The degree of the flocculation in- concentrations), the alumina particles commence
creases as the PVA concentration is increased, aggregating; this aggregation is due to the fact
and this effect of flocculation is more pronounced that any dispersant introduced above the opti-
for the PVA concentration range between 0 and mum concentration level remains unadsorbed and
0.5%. Above a 0.5% PVA concentration, there is then functions as an electrolyte which reduces the
now little change in the degree of flocculation. As range and extent of the electrostatic repulsion, as
was discussed earlier, it was found that PVA does was discussed earlier in Section 2. When the PVA
not adsorb, to any measurable extent, onto the is added to the alumina suspension, containing
alumina AES-11 particles. Therefore, the prospect the dispersant above the optimum concentration,
of a flocculation of this system by a bridging there are at least two kinds of interactions in-
mechanism can be reasonably eliminated. The volved. One causes flocculation, due to the reduc-
alumina AES-11 particles, in the presence of tion in the repulsion between the particles (excess
aqueous ‘Darvan C’ and ‘Tiron’ solutions, are dispersant), and the other is the process where
highly surface charged. This was confirmed by the ‘free’ PVA brings about the flocculation probably
use of electrophoresis experiments [21,22]. It was through a depletion mechanism. As a combined
also observed, in a few test cases, that the addi- result of these two effects, the slope of the relative
tion of a simple salt (KNO3), in the ‘Darvan C’ viscosity (and viscoelastic properties) against PVA
and ‘Tiron’ stabilised alumina AES-11 suspen- concentration curves gradually decreases as the
sions containing different concentrations of PVA, dispersant concentration is increased (Figs. 6 and
enhances the extent of, flocculation. This indicates 17 for ‘Darvan C’, and Figs. 7 and 18 for ‘Tiron’).
that these particles are stabilised largely by a For the 4% ‘Darvan C’ and 1% ‘Tiron’ concentra-
classical electrostatic repulsion mechanism and tion level alumina suspensions, there is virtually
that this flocculation is due to the compression of no change in the slope of relative viscosity against
the electrical double layer giving rise to the domi- PVA concentration curves. It is reasonable to
nation of van der Waals attractive forces. How- conclude that the degree of flocculation, caused
ever, in the case of stabilised particles in the by the electrolyte effect (an excess of dispersant),
presence of non-adsorbing PVA, the weak floccu- is so high that the addition of the PVA does not
lation that is observed is unlikely to be due to van greatly change the nature of the flocculation.
der Waals forces as the particles are stable in the The results given in Figs. 8 and 9 may be
absence of the PVA. It is therefore highly proba- explained as follows. The slope of the viscosity, at
ble, based upon elimination, that the aggregation the shear rate of 1.46 s − 1, against the dispersant
of the ‘Darvan C’ and ‘Tiron’ stabilised alumina concentration curves gradually decreases as the
suspensions is caused by a depletion flocculation PVA concentration is increased. The change (de-
process when PVA is added to these suspensions. crease) of the slope of the viscosity against disper-
It has been reported [29] that hydroxyethyl cellu- sant concentration, as the PVA concentration is
lose (HEC) of MW 70 000 and more does not increased from 0 to 1.5% dwb, signifies that, at
adsorb to polystyrene latex particle, causes deple- low concentrations of PVA, the aggregation of
tion flocculation of the particles at a concentra- the particles is principally brought about by the
254 A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257

electrostatic mechanism. However, as the PVA the present system. Therefore, one expects that
concentration is increased the depletion mecha- the PVA will not change the rheology of the
nism predominates. alumina AES-11 suspensions stabilised with
The results shown in Figs. 12 and 13 also 0.05% dwb ‘Tiron’. However, the results (Figs. 14
highlight the two mechanism of flocculation of and 21) show that this is not the case. The relative
these alumina AES-11 suspensions. The nor- viscosity drops as the PVA concentration is in-
malised viscosity, without any PVA addition, is creased. Despite these changes, the suspensions
taken as unity. For the 1% dwb ‘Darvan C’ and remain well flocculated. These changes in the rhe-
0.125% dwb ‘Tiron’ suspensions without any ological properties are probably due to changes
PVA, the particles are well dispersed and there is induced in the structure of the flocs, an effect that
a low level of aggregation. As the PVA concentra- may not be readily quantified.
tion is increased from 0 to 1.5% dwb there is an For the case where the ‘Tiron’ concentration
approximate 25-fold increase for ‘Darvan C’ sus- was 2.0% dwb, which is  20 times higher than
pensions, and  100 times for ‘Tiron’ suspen- the optimum concentration ( 0.10–0.125%
sions, in the viscosity; and this increase in the dwb), the behaviour is rather different. At this
viscosity is brought about primarily by the aggre- high concentration of ‘Tiron’, the suspensions
gation of particles through a depletion mechanism were highly flocculated and the relative viscosity
as was explained above. However, as the disper- and G% values were the highest as compared to the
sant concentration is increased the increase in the other ‘Tiron’ concentrations used for the alumina
normalised viscosity with the increase of the PVA AES-11 suspensions. The addition of the PVA
concentration (the slope of normalised viscosity slightly lowers the value of the relative viscosity
against PVA concentration) decreases. This is due and the other rheological properties. However,
to the fact that, at the higher ‘Darvan C’ concen- this decrease is comparatively small and the sus-
trations, there is already some aggregation of the pensions remain very flocculated after the addi-
primary particles brought about by the reduction tion of PVA. This decrease in the rheological
in the electrostatic stabilisation effect. Therefore, parameters, with the increasing of the PVA con-
the increase in the normalised viscosity with the centration, may be due to the changes in the
increasing of the PVA concentration is reduced. consolidation of the floc structure. The effect is
For the highest concentration of the dispersants again not amenable to a quantitative prediction.
(4% dwb for ‘Darvan C’ and 1% dwb ‘Tiron’) To summarise the discussion of the results of
reported in Figs. 12 and 13, the effect is a small the two dispersant-PVA combinations for the alu-
increase in the normalised viscosity as the PVA mina AES-11 suspensions; the addition of the
concentration is changed from 0 to 1.5% dwb. PVA increases the degree of aggregation and this
The same conclusion can be drawn from the aggregation of particles is probably brought
examination of the viscoelastic properties as was about by the depletion flocculation mechanism.
described earlier. However, as the dispersant concentration is in-
For the 0.05% ‘Tiron’ concentration case, the creased above the optimum value (1% dwb for
surface concentration of the ‘Tiron’ molecules ‘Darvan C’ and 0.125% dwb ‘Tiron’) in the pres-
adsorbed onto the alumina is not sufficient to ence of the PVA, there are two kinds of mecha-
completely cover the surface of the alumina AES- nisms involved which cause the flocculation of the
11 particles. At this less than optimum concentra- particles: (1) the reduction of the electrostatic
tion of the ‘Tiron’, the van der Waals attractive stabilisation effect caused by the excess amount of
forces are significant enough to cause the aggrega- the free dispersant which functions as an elec-
tion of the particles. It has also been observed trolyte; and (2) the depletion effect of the non-ad-
that the addition of PVA to the alumina AES-11 sorbing PVA between the particles. At very high
suspensions, without any dispersant, does not infl- concentrations of the dispersant, the reduction in
uence the rheology. As a result, it may be con- the electrostatic repulsion mechanism dominates
cluded that PVA is a non-adsorbing polymer for due to the stronger degree of flocculation of the
A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257 255

particles induced by the excess electrolyte. De-


pletion flocculation is known, in comparison, to
be a relatively weak interaction.

5. Comparison of the two dispersant – poly(vinyl)


alcohol combinations

For the two dispersants, ‘Darvan C’ and


‘Tiron’ used to stabilised the alumina AES-11
suspensions, the viscosity at a shear rate of 1.46
s − 1 for the respective dispersant’s optimum con-
centrations, as a function of the PVA concentra-
tion, is shown in Fig. 22. This figure shows that
Fig. 23. Normalised viscosity against PVA concentration of
as the PVA concentration is increased, the vis-
alumina AES-11 40% suspensions for three different dispersant
cosities of the suspension stabilised with the two (
, ‘Tiron’ and , ‘Darvan C’) concentrations at their opti-
dispersants start increasing. This increasing of mum concentration. Also normalised viscosity of PVA solu-
the viscosities, with increasing of the PVA con- tions against PVA concentration.
centrations, is due to the alumina particles ag-
gregating, which is probably caused by a that the normalised/reduced viscosity is the
depletion flocculation mechanism. This effect on higher for the ‘Tiron’ suspensions than for the
the viscosity, caused by changing the PVA con- suspensions containing ‘Darvan C’; the degree
centrations, is greater in the ‘Tiron’ stabilised of depletion flocculation, with respect to an in-
system compared to the ‘Darvan C’ stabilised crease in the PVA concentration, is greater in
suspensions. To explain this point further the the case of the ‘Tiron’ stabilised suspensions. In
normalised/reduced viscosity, hn, (hn = hs/h0 Fig. 23, the normalised viscosity2 of the
where hs and h0 are the viscosities of the sus- aqueous–PVA solution is also plotted as a func-
pensions with and without PVA, respectively) is tion of PVA concentration (upper X-axis). In
plotted against the PVA concentration in Fig. the case of the aqueous–PVA solution, there is
23. It is clear from an inspection of the Figure
no depletion effect and the increase of the vis-
cosity of the solution is caused by the PVA–sol-
vent interactions alone. However, in the case of
suspensions there is an addition effect i.e. deple-
tion flocculation, and therefore the slopes are
different for the three systems. The upper X-axis
expresses the PVA concentration in g l − 1 of the
continuous medium for the different 40% (v/v)
suspensions containing the PVA. A comparison
of the viscoelastic properties for the two disper-
sants and the PVA combinations indicates a
similar (to viscosity) level for the PVA interac-
tions in the three systems.

Fig. 22. Viscosity against PVA concentration of alumina AES-


11 40% v/v suspensions for two different dispersants (‘Darvan 2
The normalised/reduced viscosity, hn of aqueous PVA
C’ and ‘Tiron’) at their optimum concentrations, at a shear solution is hn =hp/hw where hp and hw are the viscosities of
rate of 1.46 s − 1. the PVA solution and water, respectively.
256 A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257

6. Conclusion the relative viscosity is probably due to a deple-


tion flocculation mechanism. A 0.5% dwb PVA
In this paper, selected rheological properties of concentration gives the maximum interactions
alumina suspensions in the presence of a PVA which becomes constant above this PVA concen-
binder and two dispersants ‘Darvan C’ and tration. Thus, in summary, the rheological re-
‘Tiron’ have been described. When the PVA was sponses of these systems may be rationalised using
used without any dispersant, the resulting viscos- established wisdom; Fig. 1 is a summary. The
ity of the 40% by volume suspensions remained most marked effect arise from the electrostatic
virtually unaffected for the 0 – 1.5% dwb PVA screening induced by excess dispersant addition
concentration range. The behaviour of many dif- but excessive polymer addition also promote de-
ferent concentration combinations of the PVA pletion flocculation. In the case of both, an opti-
and the dispersants has been described. The use of mal amount of material is required to provide the
‘Darvan C’ and ‘Tiron’ as dispersants with PVA most satisfactory formulation.
for the 40% (v/v) suspensions progressively in- The system we have studied is a complex one
creases the relative viscosity, at a shear rate of such as those frequently encountered industrially.
1.46 s − 1, as the PVA concentration is 0–0.5% That we can explain the rheological data in terms
dwb. Above a PVA concentration of 0.5% dwb, of the microstructure of the system is testament to
there is virtually no change in the relative viscos- the work of Dr Goodwin and his colleagues who
ity. At a higher shear rate (=146 s − 1), the rela- have been pre-eminent in trying to explain what
tive viscosity is almost constant for the 0–1.5% rheological data mean. Without his pioneering
PVA concentration range. For a given PVA con- studies, we could not have been so bold.
centration, the viscosity at the shear rate of 1.46
s − 1 for ‘Tiron’ system is less than that of the
‘Darvan’ system. As the concentration of the dis-
References
persants is increased above the optimum value,
the viscosity of the resulting alumina suspensions [1] B.A. Frith, R.J. Hunter, J. Colloid Interface Sci. 57
is higher than the viscosity of the next lower (1976) 248.
concentration level of the dispersant, for a given [2] B.A. Frith, R.J. Hunter, J. Colloid Interface Sci. 57
value of the PVA concentration. The viscoselastic (1976) 257.
behaviour, i.e. the G%, G%% parameters and dy- [3] B.A. Frith, R.J. Hunter, J. Colloid Interface Sci. 57
(1976) 266.
namic viscosity against the PVA concentration, [4] R. Buscall, J.W. Goodwin, M.W. Hawkins, R.H. Ot-
show similar trends to the corresponding viscosity tewill, J. Chem. Soc. Faraday Trans. I 78 (1982) 2873.
against PVA concentration behaviour. The high [5] R. Buscall, J.W. Goodwin, M.W. Hawkins, R.H. Ot-
viscosities noted for the ‘above optimum’ disper- tewill, J. Chem. Soc. Faraday Trans. I 78 (1982) 2889.
sant concentrations were consistent with the ac- [6] R. Buscall, P.D.A. Mills, J.W. Goowin, D.W. Lawson, J.
Chem. Soc. Faraday Trans. I 84 (1988) 4249.
tion of an electrostatic shielding mechanism. The [7] J.W. Goodwin, R.H. Ottewill, J. Chem. Soc. Faraday
excess amount of the dispersant remains in the Trans. 87 (1991) 357.
continuous medium and functions as an elec- [8] J.W. Whittemore, Am. Ceram. Soc. Bull. 23 (1944) 427.
trolyte and ‘compresses’ the range of the double [9] E.P. McNamara, J.E. Comefora, J. Am. Ceram. Soc. 28
layer and hence changes the rheological proper- (1945) 25.
[10] C.C. Treischel, E.W. Emrich, J. Am. Ceram. Soc. 29
ties. The increase of the viscosity with the increas- (1946) 129.
ing of the PVA concentration is due the increase [11] A. Wild, Am. Ceram. Soc. Bull. 33 (1954) 368.
of interparticle interactions within the system and [12] T. Knapp, Am. Ceram. Soc. Bull. 33 (1954) 11.
these enhanced interactions cause an aggregation [13] H. Thurnauer, in: W.D. Kingery (Ed.), Ceramic Fabrica-
of the particles by a depletion mechanism. The tion Processes, Wiley, New York, 1958.
[14] S. Levine, Ceram. Age 75 (1960) 39.
addition of the PVA causes increase in the relative [15] S. Levine, Ceram. Age 75 (1960) 25.
viscosity of both the ‘Darvan C’ and ‘Tiron’ [16] T.A. Smith, Trans. Br. Soc. 61 (1962) 523.
stabilised alumina suspensions. This increase of [17] A.R. Teter, Ceram. Age 82 (1966) 30.
A.U. Khan et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 161 (2000) 243–257 257

[18] R.W. Greenwood, Research Report, E.U., E.H.C.M. Pro- [24] A.U. Khan, PhD Thesis, Imperial College of Science and
gramme, Preparation of Concentrated Aqueous Alumina Technology and Medicine, London, 1997.
Suspensions Suitable for Tape Casting, I.R.T.E.C., [25] B.J. Brisocoe, A.U. Khan, P. Luckham, J. Eur. Ceram.
Faenza, Italy. Soc. 18 (1998) 2141.
[19] W.B. Russel, D.A. Saville, W.R. Schowalter, Colloidal [26] G.Y. Onoda, Jr., in: G.Y. Onoda, L.L. Hench (Eds.),
Dispersion, Cambridge University Press, Cambridge, Ceramic Processing Before Firing, Wiley-Interscience
1989, p. 258. Publication, New York, 1978.
[20] R.J. Hunter, Foundations of Colloid Science, Oxford [27] B.C. Bonekamp, W.H. Van’T Van, M.J. Schoute, H.J.
Science Publications, 1991. Veringa, in: G. de With, R.A. Terpstra, R. Metselaar
[21] D.H. Napper, Polymeric Stabilisation of Colloidal Dis- (Eds.), EuroCeramic, vol. I, Elsevier Applied Science,
persions, Academic Press, 1983. London, 1989, p. 1.145.
[22] D.H. Everett, Basic Principles of Colloids, Royal Society [28] D.J. Shaw, Introduction to Colloid and Surface Chem-
of Chemistry, 1989. istry, fourth ed., Butterworth-Heinemann, Oxford, 1992,
[23] G.H. Fleer, M.A. Cohen Sturat, J.M.H.M. Scheutjens, T. pp. 225 – 227.
Cosgrove, B. Vincent, Polymer at Interfaces, Chapman [29] W. Liang, Th. F. Thadros, P.F. Luckham, J. Colloid
and Hall, London, 1993. Interface Sci. 160 (1993) 183.

You might also like