Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Colloids and Surfaces A: Physicochem. Eng.

Aspects 436 (2013) 325–332

Contents lists available at ScienceDirect

Colloids and Surfaces A: Physicochemical and


Engineering Aspects
journal homepage: www.elsevier.com/locate/colsurfa

A population balance equation model to predict regimes of controlled


nanoparticle aggregation
Anand K. Atmuri a , Michael A. Henson a , Surita R. Bhatia a,b,c,∗
a
Department of Chemical Engineering, University of Massachusetts, Amherst, MA 01003, United States
b
Department of Chemistry, Stony Brook University, Stony Brook, NY 11794, United States
c
Center for Functional Nanomaterials, Brookhaven National Laboratory, Upton, NY 11793, United States

h i g h l i g h t s g r a p h i c a l a b s t r a c t

• Formation of stable finite-sized clus-


ters observed in charged colloidal
dispersions.
• A population balance equation (PBE)
model was developed to capture this
phenomenon.
• Model successfully predicts aggrega-
tion regimes and final aggregate size.

a r t i c l e i n f o a b s t r a c t

Article history: Forming stable clusters or aggregates of nanoparticles is of interest in a number of emerging applications.
Received 15 February 2013 While formation of unstable fractal aggregates and flocs has been well-studied with both experiments
Received in revised form 27 June 2013 and theory, the conditions that lead to stable, finite-sized clusters is not as well understood. Here, we
Accepted 3 July 2013
present an integrated experimental and modeling study to explore aggregation in concentrated attrac-
Available online 12 July 2013
tive colloidal suspensions. A population balance equation (PBE) model is used to predict the aggregation
dynamics of quiescent colloidal suspensions. A DLVO (Derjaguin–Landau–Verwey–Overbeek) type poten-
Keywords:
tial is used to describe the interparticle potential, with attractive interactions arising from van der Waals
Cluster
Aggregation
forces and long-range repulsive interactions caused by electrostatics. The PBE model includes a full cal-
Aggregate culation of stability ratio variations as a function of aggregate size, such that the energy barrier increases
Processing with increasing size. As the ionic strength is decreased, the model predicts three regimes of behavior:
Theory uncontrolled aggregation into large flocs, controlled aggregation into stable clusters, and no aggrega-
Modeling tion. The model is tested experimentally using latex particles at different salt concentrations and particle
concentrations. When the Hamaker constant and surface potential are fit to aggregate size measure-
ments collected at one salt concentration, the model accurately predicts the final mean aggregate size
and regimes of aggregation at other salt concentrations and the same particle concentration. This result
suggests that van der Waals and electrostatic forces are the dominant particle interactions in determining
the final aggregate state. The mean aggregate size and aggregation regimes at different particle concen-
trations could be accurately predicted by adjusting the surface potential. This parameter adjustment
is consistent with the expectation that increasing colloid weight fractions cause aggregates to have a
more fractal nature and hence have a lower effective repulsion. However, the model predicts much faster
aggregation rates than what are observed experimentally. This discrepancy may be due to hydrodynamic
effects or another slow dynamical process which is not accounted for in the model. Nevertheless, this
study presents the first PBE model that can successfully predict stable aggregate size and aggregation
regimes of charged colloidal particles over a range of salt concentrations and particle concentrations.
© 2013 Elsevier B.V. All rights reserved.

∗ Corresponding author at: Department of Chemistry, Stony Brook University, Stony Brook, NY 111794, United States. Tel.: +1 631 632 7788.
E-mail addresses: surita.bhatia@stonybrook.edu, sbhatia@ecs.umass.edu (S.R. Bhatia).

0927-7757/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.colsurfa.2013.07.002
326 A.K. Atmuri et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 325–332

1. Introduction (i.e., no energy barrier) to the rate when particle interactions are
present. Thus, a higher value of the stability ratio represents a
Aggregation of colloids has long been studied for processes such less efficient collision. We can also consider collision efficiency
as formation of large flocs for water treatment [1] and formation of in the context of two commonly considered limits of aggrega-
fractal colloidal gels. A more recently studied phenomenon is the tion phenomena, diffusion-limited cluster aggregation (DLCA) and
formation of dense, finite-sized clusters of colloidal particles for reaction-limited cluster aggregation (RLCA). For DLCA, the collision
applications such as toner for digital printing [2] microparticles for efficiency is very high and the stability ratio is the order of unity
drug delivery [3] and probes for cellular imaging [4]. We refer to [25,26]. By contrast, for RLCA the collision efficiency is much lower
processes leading to these dense clusters as “controlled aggrega- and the stability ratio is greater than unity [25,26].
tion,” as conditions must be such that aggregates grow to a certain Morbidelli and co-workers have performed extensive exper-
desired size and then remain stable in size. iments [5,27], modeling [24,28,29], and simulations [30] to
The interaction potential between particles plays a major role understand the kinetics of aggregation and structure of the result-
in determining the nature and rate of aggregation in colloidal ing aggregates. However, these and other studies [9,26,31,32] have
systems. Clustering and aggregation arising from a number of dif- generally focused on aggregation leading to gel phases, and the
ferent mechanisms have been explored in the literature, including time scale reported is typically on the order of a few minutes or
addition of salt to screen electrostatics [5,6], additional of a non- for some studies up to a few hours. Thus, formation of finite-sized
adsorbing polymer to induce depletion attraction [7,8] polymer clusters is not discussed, and the time scales are such that it is
bridging [9] and “patchy” electrostatic interactions [10]. In colloidal difficult to determine whether the aggregate size would remain
systems with a short-range or moderate-range attraction and long- stable over longer times. Additionally, these models generally have
range repulsion, there is competition between aggregation from the several parameters that need to be fit to experimental data. For
attractive part of the potential and stabilization from the repulsive example, in some studies the stability ratio must be fit at different
part. This competition can lead to the formation of stable clusters conditions. Such fitting makes it difficult to extend the model to
at low volume fractions [11,12], but at higher volume fractions, a larger parameter space and use the model in a predictive man-
suspensions undergo structural arrest, either via percolation [13] ner. Additionally, some studies [9,19,26] report PBE-type models of
or a glass transition [14] depending on the range of the repulsion. colloidal aggregation using the stability ratio only for the primary
Interestingly, stable clusters of particles have also been observed particles. These types of models do not allow the stability ratio to
in systems with short-range attractions and a screened long-range change as the aggregate size increases, and do not predict a regime
repulsion [7,10] and in systems with steric repulsive interactions of controlled aggregation.
[15], and some theoretical studies predict stable clusters even in More detailed molecular dynamics simulations of clustering
purely repulsive systems that have a soft shoulder in the interpar- phenomena [11,14,33–35] yield some insight into the type of
ticle potential [16]. physics that must be incorporated into a PBE model to predict the
In this study we focus on a process-level description of aggre- formation of stable clusters. Both Sciortino and co-workers [11]
gation of charged particles that are destabilized with addition of a and Groenwold and Kegel [33,34] describe stable cluster formation
salt. We use a DLVO-type potential to describe interactions between arising from a balance of short-range attractions and long-range
particles, which includes van der Waals attraction between parti- electrostatic repulsion. Short-range attractions initially drive pri-
cles and a repulsive component arising from overlapping electrical mary particles to aggregate. Further addition of particles to the
double layers [17,18]. Addition of a salt screens the repulsive com- cluster increases the charge on the cluster. After clusters have
ponent of the potential, leading to aggregation. There are several grown to a certain size, they have accumulated so much charge that
model experimental systems that show stability and aggregation addition of additional primary particles is unfavorable [11,33,34].
behavior that are well-described by the classical DLVO theory [18]. In other words, the clusters become so strongly charged that
Under certain conditions, additional interactions that cannot be they repel primary particles, and this strong long-range repul-
explained by DLVO theory may be present, sometimes referred sion dominates overcomes any particle-cluster attraction and sets
as non-DLVO forces. These include short-range hydration forces, the equilibrium cluster size [11]. In the model of Sciortino and
capillary condensation, and specific ion adsorption [17]. However, coworkers, the radius of the cluster is explicitly renormalized as
due to the difficulty in quantifying these forces and experimen- the aggregation process proceeds [11,14] to account for changes
tally determining the additional parameters that would be needed in interactions as the aggregation process proceeds. Experimental
for their quantification, their utilization in process models is less studies have indirectly confirmed these simulations by demon-
feasible [19]. strating the presence of stable clusters [11,12], although to our
Work by Smoluchowski [20] laid the foundation for the use knowledge no direct measurements of the interparticle potential
of particle population balance equation (PBE) models to describe between clusters of varying sizes and primary particles have been
aggregation in colloidal suspensions. The two important functions performed. Nevertheless, we believe molecular simulations of this
in the PBE are the collision frequency factor, or collision kernel, phenomenon [11,14,33–35] suggest that any process-level model
and the collision efficiency factor, or stability ratio. These kernels must allow the interaction potential between aggregates and par-
take into account aggregation due to Brownian motion and shear ticles, and hence the stability ratio, to change as aggregate size
[9,21–23] with modified kernels available to account for differ- increases.
ential settling [9]. In our present study, we focus on aggregation Here we describe our efforts to develop a PBE model with a min-
that occurs under quiescent conditions [24]. Throughout the paper, imum number of fitted parameters capable of mapping out three
we use the term “particles” to include both primary particles and regimes: uncontrolled aggregation, controlled aggregation, and no
aggregates formed of primary particles. aggregation (a stable suspension of primary particles). By uncon-
In the case of particles with a DLVO-type potential, under certain trolled aggregation, we refer to systems where aggregates continue
conditions particles need to cross an energy barrier to aggregate to grow until either gelation occurs, meaning that the aggregates
with other particles. So, every collision experienced by a particle are large enough to span the system, or precipitation occurs, mean-
need not make it to stick to its collision partner, and the collision ing that very large aggregates form that are too large to remain
outcome is governed in part by the height of the barrier. To take this suspended in solution. By controlled aggregation, we refer to sys-
barrier into account, we use the stability ratio, which is defined as tems where aggregates grow to a specified size and remain stable
the rate of aggregation when the interaction is diffusion-limited at that size while being suspended in solution. To parameterize and
A.K. Atmuri et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 325–332 327

validate our model, we have performed experiments on colloidal where kB is Boltzmann’s constant, T is absolute temperature and
suspensions with moderate particle concentrations of 5–15 wt%,  is the suspending fluid viscosity. The stability ratio Wj,k defines
relevant to some industrial processes that rely on controlled aggre- the efficiency of aggregation when particles of radii rj and rk col-
gation. Recently, we reported experimental evidence of controlled lide, which is governed by the interaction potential between the
aggregation in a shear environment [36], but here we focus on particles given by [38]
aggregation under quiescent conditions where there is no breakage  ∞
exp(VT /(kB T ))
term in the PBE model. Thus, the particles aggregate irreversibly to Wj,k = (rj + rk ) dR (4)
form clusters and the final size of clusters is solely determined by rj +rk
R2
the interaction potential. To our knowledge, this study represents
VT represents the total interaction potential, and in the present case
the first attempt to use PBE models to predict stable aggregate size
using DLVO theory is given as the sum of van der Waals attrac-
and aggregation regimes of charged colloidal particles over a range
tion (VA ) and electrostatic repulsion (VE ). The attractive part of the
of salt and particle concentrations.
potential is given by:

2. Materials and methods A 2r1 r2 2r1 r2 R2 − (r1 + r2 )2
VA = − + + (5)
6 R2 − (r1 + r2 )2 R2 (r1 − r2 )2 R2 − (r1 − r2 )2
Aqueous suspensions of electrostatically stabilized polystyrene
latex particles with an approximate density of 1.05 g/cm3 and a where A represents the Hamaker constant and R is the center-to-
mean diameter of 200 nm and a colloid weight fraction of 45 wt%, center distance between the particles. The electrostatic repulsion
supplied by Xerox Corporation, were used for the aggregation is given by
experiments. Potassium chloride was used to destabilize the latex  k T 2  
B zc e 01 zc e 02 r1 r2
suspension. Colloid weight fractions of 5, 10 and 15 wt% at different VE = 64εr ε0 tanh tanh
zc e 4kB T 4kB T r1 + r2
ionic strengths were prepared by diluting the stock suspension with
the required amount of water. The pH of the suspensions was not exp(−(R − r1 − r2 )) (6)
altered. After addition of salt, samples were taken out periodically
and diluted gently to measure the particle mean size and size distri- where e is the elementary charge, zc is valence of the counterion,
bution using dynamic light scattering (DLS) with a 200-mW Innova and ε0 and εr are the dielectric constants of vacuum and the solvent,
Ar-ion laser of wavelength 488 nm with a Brookhaven Instruments respectively. The Debye–Huckel parameter  is a function of elec-
BI-9000AT correlator. We report the volume mean diameter as a trolyte concentration, valence of electrolyte ions, and temperature,
function of time, as opposed to number mean diameter, such that while the surface potential 0 depends on pH and temperature. The
the formation of large aggregates in emphasized. unknown parameters in Eqs. (5) and (6) are the Hamaker constant
and surface potential, which we have used as fitting parameters.
3. Theory In our model, we consider aggregates as “particles” with a larger
radius, and thus use Eqs. (5) and (6) to compute the interaction
The population balance equation for pure aggregation of parti- potential not only between primary particles, but also between
cles (e.g., no breakage of aggregates) under quiescent conditions is aggregates and single particles, between aggregates of different
given as [20]: sizes, and so on. As discussed further below, this simplification
 v yields an interparticle potential, and hence a stability ratio, that
∂n(v) 1 varies as aggregation proceeds.
= ˇ(v
∂t 2 0 The discretized PBE model (Eq. (2)) was solved numerically
 ∞ using the fixed pivot technique [39] with 64 node points for dis-
− v , v )n(v − v , t)n(v , t)dv − ˇ(v, v )n(v, t)n(v t)dv cretizing particle size. This method was chosen due to its relatively
0 low computational cost and ability to calculate the particle size
(1) distribution with great precision [37]. The PBE model has been
where v is the particle volume and n(v, t) is the density of particles discretized at every node point, yielding 64 nonlinear ordinary dif-
with volume between v and v + dv at time t. In Eq. (1), the term on ferential equations (ODEs) in time which were integrated with the
the left represents the rate of change of the number density (n), MATLAB code ode15s to calculate the number distribution at each
the first term on the right represents the birth rate when parti- node point. The number distribution was converted to volume dis-
cles aggregate, the second term on the right represents the death tribution [40] to calculate the volume mean diameter.
rate when particles undergo aggregation with other particles, and
ˇ is the aggregation frequency. For numerical solution, we use a 4. Results and discussion
discretized version of the population balance equation [34]:
⎧ ⎫ 4.1. Interaction potential between particles and aggregates
⎨ 
j≥k   ⎬
dNi (t) 1
= 1 − ıj,k ˇj,k Nj (t)Nk (t) Fig. 1a shows a representative interaction potential between
dt ⎩ 2 ⎭
xi−1 ≤(xj =xk )≤xi+1 equal-sized particles as a function of their distance of separation.
The form of the DLVO potential results in a deep primary mini-

M
mum and a shallower secondary minimum [17]. The secondary
− Ni (t) ˇi,k Nk (t) (2) minimum is difficult to discern in Fig. 1a due to the large scale
k=1 on the y-axis which was used to show the relative contribution of
where ı is the delta function which avoids double counting of par- electrostatic and van der Waals forces; it is more readily apparent
ticles,  is used to partition particles into their respective size bins, in Fig. 1b. Understanding of aggregation processes typically focuses
and ˇ for quiescent aggregation is given by a Brownian-type kernel: on the particles entering the primary minimum, where they are in
near-contact and are bound essentially irreversibly. The secondary
minimum can also play a role in aggregation; however, it is thought
2kB T 1 1
ˇj,k = (r + rk ) + (3) to be related to formation of looser, reversibly bound aggregates
3Wj,k j rj rk
[17]. Under appropriate conditions, there is an energy barrier that
328 A.K. Atmuri et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 325–332

Fig. 3. Experimental measurements of particle aggregation for a 5 wt% suspension at


different salt concentrations, with aggregate sizes measured via DLS. Uncontrolled
aggregation is observed at 0.3 M (red asterisks), controlled aggregation is observed
for 0.24–0.29 M (black triangles, stars, diamonds, and circles), and no aggregation is
observed at 0.2 M (blue stars). (For interpretation of the references to color in this
figure legend, the reader is referred to the web version of this article.)

increasingly difficult to add to an aggregate as its size increases. As


mentioned above, this mechanism of formation of stable clusters
has been detailed by Groenwold and Kegel [33,34]. Their results
highlight the need to include a size-dependent calculation of the
stability ratio in our PBE model.

4.2. Aggregation experiments


Fig. 1. Interaction energy as a function of distance of separation with param-
eters A = 3.08 × 10−20 J and 0 = 59 mV for a particle size of 200 nm (a) ionic
Aggregation experiments were carried out at colloid weight
strength = 0.2 M and (b) different ionic strengths.
fractions of 5, 10 and 15 wt% and at different salt concentrations.
Addition of salt decreases electrostatic repulsion and thus trigg-
must be crossed before particles enter the primary minimum and ers aggregation. Fig. 3 shows the evolution of the volume mean
begin to aggregate. The size of this barrier impacts the aggrega- aggregate diameter as a function of time for salt concentrations of
tion efficiency of colliding particles. Addition of salt decreases the 0.2–0.3 M at a colloid weight fraction of 5 wt%, with aggregate sizes
magnitude of electrostatic repulsion and hence the barrier height measured via DLS. At the highest salt concentration used, 0.3 M, the
decreases and aggregation is enhanced (Fig. 1b). suspension undergoes uncontrolled aggregation and forms a gel
Fig. 2 shows the total interaction potential energy between a phase. We stopped measuring particle size when the gel phase was
single particle, fixed in size at 200 nm, and particles of different formed. Suspensions with ionic strengths in the range 0.24–0.29 M
sizes as a function of their distance of separation. Many groups show controlled aggregation that ceases after the clusters reach
[9,19,26] have performed PBE-type modeling of colloidal aggrega- some stable size that depends on the salt concentration. At ionic
tion using the stability ratio only for the primary particles. However, strengths of 0.2 M and less (data not shown), we do not observe
as the interparticle potential is dependent on size, the stability any aggregation, and the suspensions remain as primary particles.
ratio changes as aggregation proceeds. This physics is important in Similar results were observed for other particle concentrations.
understanding the mechanism of controlled aggregation of charged Fig. 4 shows aggregation data for a 10 wt% sample at different salt
colloids [33,34]. As the aggregates grow in size, although the sur- concentrations. This system also displays uncontrolled aggrega-
face potential remains the same, the effective repulsion increases tion at 0.3 M, controlled aggregation with aggregates of a stable
because of the increased number of particles in the aggregate.
Growth increases the barrier against aggregation, so particles find it

Fig. 4. Experimental measurements of particle aggregation for a 10 wt% suspension


at different salt concentrations, with aggregate sizes measured via DLS. Uncontrolled
aggregation is observed at 0.3 M (red asterisks), controlled aggregation is observed
Fig. 2. Total interaction energy as a function of distance of separation between a for 0.22–0.28 M (black stars diamonds, circles, and triangles), and no aggregation is
particle of size 200 nm and particles of different sizes. Parameters used are the same observed at 0.2 M (blue stars). (For interpretation of the references to color in this
as in Fig. 1 at an ionic strength of 0.2 M. figure legend, the reader is referred to the web version of this article.)
A.K. Atmuri et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 325–332 329

Fig. 5. Experimental measurements of particle aggregation for a 15 wt% suspension


at different salt concentrations, with aggregate sizes measured via DLS. Uncontrolled Fig. 7. Experimental and predicted stable mean aggregate size as a function of
aggregation is observed at 0.26 M (red asterisks), controlled aggregation is observed ionic strength for dispersions at 5 wt%. Predicted values are based on the Hamaker
for 0.20–0.24 M (black circles, triangles, and stars), and no aggregation is observed constant and surface potential fit at a single ionic strength, 0.29 M.
at 0.19 M (blue triangles). (For interpretation of the references to color in this figure
legend, the reader is referred to the web version of this article.)

0.3 M, the mean diameter of aggregates increases very rapidly, and


the aggregate size exceeds the maximum grid size used (unlike
Table 1
the other ionic strengths). This predicted behavior is consistent
Fit model parameters at different colloid weight fractions.
with gel formation, as we observed experimentally (Fig. 3). At
Parameter Value lower ionic strengths of 0.24–0.28 M, aggregates that reach a stable,
Hamaker constant, A 3.08 × 10−20 J finite size are predicted, in good agreement with our experimental
Surface potential, 0 (5 wt% suspension) −59.5 mV observations. An ionic strength of 0.20 M is predicted to produce
Surface potential, 0 (10 wt% suspension) −59 mV no aggregation, consistent with our experiments. However, for all
Surface potential, 0 (15 wt% suspension) −57 mV
cases the predicted rate of aggregation is much faster than observed
experimentally. We discuss possible reasons for the mismatch in
the predicted aggregation kinetics below. Nevertheless, the goal of
diameter at 0.22–0.28 M, and no aggregation at 0.2 M. For a higher this work is to develop a PBE model that correctly predicts the con-
particle concentration of 15 wt% (Fig. 5), uncontrolled aggregation ditions that lead to stable aggregation and the equilibrium size of
starts at a lower ionic strength (0.26 M), and the window of con- stable aggregates, and in this sense the predictions agree very well
ditions leading to controlled aggregation narrows (0.20–0.24 M). with experimental results.
This result is to be expected, since particle collisions will occur It should be noted that the Hamaker constant and surface poten-
more frequently in denser suspensions, increasing the likelihood tial were fit at only one ionic strength, 0.29 M, and these fit values
of aggregation and formation of large aggregates. were used to predict aggregation behavior at six different ionic
strengths. At a colloid concentration of 5 wt%, the PBE model clearly
4.3. PBE modeling captures the phenomenon of controlled aggregation and can pre-
dict the different regimes of aggregation observed experimentally.
To utilize the PBE model, we fit experimental data of the final To quantitatively test the PBE model predictions, we compare the
aggregate size at a particle concentration of 5 wt% and a salt predicted final mean aggregate size in the controlled aggrega-
concentration of 0.29 M to obtain the value of the Hamaker con- tion regime with our experimental results at 5 wt%. As shown in
stant and surface potential, which were 3.08 × 10−20 J and 59.5 mV Fig. 7, the agreement between the final experimental and predicted
respectively (Table 1). These parameter values are in reasonable aggregate sizes is excellent.
agreement within values cited in literature for similar latex par- Fig. 8 shows the measured and simulated aggregate size distri-
ticles [18]. Using the parameters fit at 0.29 M, the PBE model was butions for a sample at 5 wt% and salt concentration of 0.29 M after
used to predict the mean aggregate size for 5 wt% suspensions at the stable size is reached. While the mean size values are in agree-
five different ionic strengths (Fig. 6). At a salt concentration of ment, the experimental distribution is broader than the distribution

Fig. 6. PBE model predictions for the mean size of aggregates at different salt con- Fig. 8. Comparison of measured and predicted size distributions after stabilization
centrations for a 5 wt% suspension, with values for the Hamaker constant and surface of a 5 wt% suspension at a salt concentration of 0.29 M. The volume mean sizes are
potential fit at a single ionic strength, 0.29 M. 870 nm for the measured distribution and 865 nm for the predicted distribution.
330 A.K. Atmuri et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 325–332

Fig. 9. Experimental and predicted stable mean aggregate size as a function of ionic Fig. 10. Predicted aggregation state diagram as a function of colloid volume fraction
strength for dispersions at 10 wt% and 15 wt%. and ionic strength, compared to experimental results, showing regions of uncon-
trolled aggregation, controlled aggregation, and no aggregation.

predicted by the model. This is often observed in PBE modeling of


aggregation-type processes [37]. In our case, one possible cause of experimentally we find that uncontrolled aggregation occurs at
the mismatch between the experimental and predicted size poly- 0.25 M and above. However, the PBE model predicts the onset of
dispersities is that the PBE model assumes the aggregates formed uncontrolled aggregation at 0.26 M. We suspect this disagreement
are compact and spherical. Any fractal structure in the aggregates may be due to hydrodynamic effects, which are not included in the
would likely broaden the experimentally measured distribution. current PBE model. Stokesian dynamics simulations have shown
The Hamaker constant is an inherent property of the particle that hydrodynamic effects, which are stronger for more con-
and should not change as the colloid weight fraction is increased. centrated suspensions, decrease the volume fraction at which
However, there is some evidence that as particle concentration is percolation occurs [41,42]. This effect could potentially widen the
increased, the aggregates formed become more fractal in nature window of conditions where gelation (e.g., uncontrolled aggrega-
[14] and the effective repulsion experienced by a particle appro- tion) experimentally occurs.
aching the aggregate increases. This effect can be captured by a As noted above, for all salt and particle concentrations, the pre-
slight decrease in the surface potential. So, to apply the PBE model dicted rate of aggregation is much faster than what is observed
to experimental data taken at other weight fractions, we used the experimentally. One possible cause is that there is some physical
Hamaker constant value from above, but re-fit the surface poten- process with slow dynamics not accounted for in our model. There
tial as before. Table 1 shows that the fitted surface potential does are two dynamical processes that could cause the measured clus-
decrease slightly with particle concentration, as expected. For the ter size to be smaller than the predicted cluster size and slow the
10 wt% suspension, data at 0.28 M was used for the fit, resulting aggregation dynamics. One potential process is that particles may
in a surface potential of 59 mV. For the 15 wt% suspension, data at be stuck reversibly, so they may break free from the cluster and dif-
0.24 M was used for the fit, resulting in a surface potential of 57 mV. fuse away due to thermal motions. This could arise from patchiness
These values of the surface potential, fit at a single ionic strength, on the particle surface, which is most certainly present (e.g., regions
were used to predict the aggregation regime and final stable aggre- on the surface that have more charge than other regions and there-
gate sizes for 6–7 other salt concentrations. The predicted stable fore experience stronger repulsion). So even if the overall attraction
mean aggregate size in the controlled aggregation regime is shown between particles is strong enough that we would expect particles
in Fig. 9, compared with experimental results. Again, the model pre- to be irreversibly stuck, there may be local regions that are not
dicts the different regimes of aggregation behavior and the mean experiencing as strong of an attraction. The second process is that
size of stable aggregates fairly well. At lower salt concentrations particles that have already joined a cluster may re-arrange within
where the aggregates are relatively small, the measured sizes are the cluster. This motion again could be due to patchiness. A particle
slightly larger than those predicted by the model. This discrepancy that is stuck to a cluster may have some freedom to roll into a posi-
is likely due to differences between how size is calculated in the tion that makes the cluster denser and could potentially be more
model versus determined experimentally. DLS gives the effective energetically favorable. If either of these processes occurs on a time
hydrodynamic radius, whereas aggregate size in the model is rep- scale that is slower than aggregation, the effect would be that the
resented by volume mean diameter. If we consider the simple case final aggregate size could be predicted accurately but the formation
where the aggregates are doublets, the model will predict size by rate seen experimentally would be slower than predicted by the
summing of volumes of both particles, whereas the effective hydro- model. We attempted to incorporate the first effect into the model
dynamic radius from DLS will be larger. However, at moderate and (e.g., reversible aggregation due to particle patchiness), but our pre-
high salt concentrations where aggregates are larger, this effect liminary work showed that this did not appreciably slow down the
becomes less significant, and there is very good agreement between aggregation rate and rendered the model unable to predict stable
experiment and model predictions. aggregation.
The PBE model can be used to generate a “state diagram” An alternate explanation is that hydrodynamic effects, which
showing conditions leading to different aggregation regimes (e.g., are also not included in the PBE model, may be acting to slow
no aggregation, controlled aggregation, and uncontrolled aggre- the aggregation process. Recent Stokesian dynamics simulations
gation) as a function of colloid concentration and ionic strength. of charge-stabilized particles [42] show that hydrodynamic effects
Fig. 10 compares the predicted aggregation regimes to experi- slow movement of particles into the primary minimum and make
mental results. At each colloid concentration, the surface potential particle rearrangement difficult at small interparticle separations.
was fit at a single ionic strength, and predictions were performed These effects would impact the aggregate structure and percolation
at 6–7 additional values of the ionic strength. The agreement volume fraction, but also the overall rate of aggregation. However, it
between the predicted state diagram and experimental results is extremely difficult to incorporate hydrodynamics into a PBE-type
is very good. There is slight disagreement in the predicted model. Even state-of-the-art molecular simulations of aggregation
and experimental aggregation regimes at 15 wt%. For example, that incorporate hydrodynamics are limited to a relatively small
A.K. Atmuri et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 325–332 331

number of particles due to the complexity of the interactions, and [5] P. Sandkuhler, J. Sefcik, M. Morbidelli, Kinetics of aggregation and gel formation
to accurately model these effects in the concentrated dispersion in concentrated polystyrene colloids, J. Phys. Chem. B 108 (2004) 20105–20121.
[6] P. Sandkuhler, J. Sefcik, M. Morbidelli, Kinetics of gel formation in dilute dis-
we describe here would be very computationally expensive if not persions with strong attractive particle interactions, Adv. Colloid Interface Sci.
infeasible. 108–109 (2004) 133–143.
Although our PBE model does not predict the rate of aggrega- [7] P.J. Lu, J.C. Conrad, H.M. Wyss, A.B. Schofield, D.A. Weitz, Fluids of clusters in
attractive colloids, Phys. Rev. Lett. 96 (2006) 028306.
tion accurately, our goal was to develop a process model capable of [8] A.K. Atmuri, S.R. Bhatia, Polymer-mediated clustering of charged anisotropic
capturing the formation of stable aggregates. The model presented colloids, Langmuir 29 (2013) 3179–3187.
is able to quantitatively predict the final aggregate mean size at dif- [9] V. Runkana, P. Somasundaran, P.C. Kapur, A population balance model for floc-
culation of colloidal suspensions by polymer bridging, Chem. Eng. Sci. 61 (2006)
ferent ionic strengths and to accurately predict different regimes of
182–191.
aggregation, using only one experimental data set at each particle [10] J. Sabin, G. Prieto, F. Sarmiento, Stable clusters in liposomic systems, Soft Matter
concentration to fit the surface potential. Thus, the model has util- 8 (2012) 3212–3222.
[11] F. Sciortino, S. Mossa, E. Zaccarelli, P. Tartaglia, Equilibrium cluster phases and
ity in designing products based on stable aggregation of charged
low-density arrested disordered states: The role of short-range attraction and
nanoparticles. Additionally, our results demonstrate the necessary long-range repulsion, Phys. Rev. Lett. 93 (2004) 055701.
physics that must be included in a process model to capture con- [12] A.I. Campbell, V.J. Anderson, J.S. van Duijneveldt, P. Bartlett, Dynamical arrest in
trolled aggregation, providing further insight into the mechanism attractive colloids: the effect of long-range repulsion, Phys. Rev. Lett. 94 (2005)
208301.
of this phenomenon. [13] F. Sciortino, P. Tartaglia, E. Zaccarelli, One dimensional cluster growth and
branching gels in colloidal systems with short-range depletion attractive and
screened electrostatic repulsion, J. Phys. Chem. B 109 (2005) 21942–21953.
5. Conclusion [14] J.C.F. Toledano, F. Sciortino, E. Zaccarelli, Colloidal systems with competing
interactions: from an arrested repulsive cluster phase to a gel, Soft Matter 5
We present experimental and process modeling results for the (2009) 2390–2398.
[15] K. Larson-Smith, D.C. Pozzo, Scalable synthesis of self-assembling nanopar-
aggregation of charge-stabilized colloids. Although classical col- ticle clusters based on controlled steric interactions, Soft Matter 7 (2011)
loidal aggregation processes are well-studied, here we explore the 5339–5347.
phenomena of controlled aggregation; e.g., aggregates that grow [16] H. Shin, G.M. Grason, C.D. Santangelo, Mesophases of soft-sphere aggregates,
Soft Matter 5 (2009) 3629–3638.
to a fixed, stable size as opposed to uncontrolled aggregation into [17] W.B. Russel, D.A. Saville, W.R. Schowalter, Colloidal Dispersions, Cambridge
fractal gels or flocs. Experimentally, we observe three regimes of University Press, 1989.
aggregation behavior as ionic strength is increased: no aggrega- [18] P.C. Hiemenz, R. Rajagopalan, Principles of Colloid and Surface Chemistry, Mar-
cel Decker Inc., 1997.
tion, controlled aggregation, and uncontrolled aggregation. Our PBE [19] L. Ehrl, Z. Jia, H. Wu, M. Lattuada, M. Soos, M. Morbidelli, Role of counterion
model includes the full calculation of the stability ratio as well as association in colloidal stability, Langmuir 25 (2009) 2696–2702.
variations in DLVO interactions between particles and aggregates [20] M.v. Smoluchowski, Versuch einer mathematischen theorie der koagulations
kinetik kolloider losungen, Z. Phys. Chem. 92 (1917) 129–168.
as aggregates grow in size. Incorporation of this physics into the
[21] M. Soos, J. Sefcik, M. Morbidelli, Investigation of aggregation, breakage and
PBE model is important for its ability to predict controlled aggrega- restructuring kinetics of colloidal dispersions in turbulent flows by popula-
tion. As the aggregates reach a certain size, the effective repulsion tion balance modeling and static light scattering, Chem. Eng. Sci. 61 (2006)
2349–2363.
between the aggregate and a single particle becomes large enough
[22] P.T. Spicer, S.E. Pratsinis, Shear induced flocculation: the evolution of floc struc-
to prevent any further addition of particles. With minimal param- ture and the shape of the size distribution at steady state, Water Res. 30 (1996)
eter fitting, the model shows good agreement with experiments in 1049–1056.
predicting the three regimes of aggregation behavior and is also [23] S.N. Maindarkar, N.B. Raikar, P. Bongers, M.A. Henson, Incorporating emulsion
drop coalescence into population balance equation models of high pressure
able to quantitatively predict the final mean aggregate size. The homogenization, Colloids Surf. A 396 (2012) 63–73.
model predicts a much faster aggregation rate than is observed [24] P. Sandkuhler, J. Sefcik, M. Lattuada, H. Wu, M. Morbidelli, Modeling struc-
experimentally, which we believe could be due to hydrodynamic ture effects on aggregation kinetics in colloidal dispersions, AIChE J. 49 (2003)
1542–1555.
effects. Nevertheless, to our knowledge this is the first PBE model [25] M.Y. Lin, H.M. Lindsay, R.C. Weitz, D.A. Ball, R. Klein, P. Meakin, Universality in
able to predict stable aggregate size and aggregation regimes of colloid aggregation, Nature 339 (1989) 360–362.
charged colloidal particles over a range of salt and particle concen- [26] V. Runkana, P. Somasundaran, P.C. Kapur, Reaction-limited aggregation in pres-
ence of short-range structural forces, AIChE J. 51 (2005) 1233–1245.
trations. [27] P. Sandkuhler, M. Lattuada, J. Wu, H. Sefcik, M. Morbidelli, Further insights into
the universality of colloidal aggregation, Adv. Colloid Interface Sci. 113 (2005)
65–83.
Acknowledgments [28] Z. Jia, H. Wu, M. Morbidelli, Application of the generalized stability model to
polymer colloids stabilized with both mobile and fixed charges, Ind. Eng. Chem.
The authors gratefully acknowledge financial support from an Res. 46 (2007) 5357–5364.
[29] M. Lattuada, P. Sandkuhler, J. Wu, H. Sefcik, M. Morbidelli, Kinetic modeling
NSF GOALI award (CBET-0853551), the NSF-funded Center for Hier- of aggregation and gel formation in quiescent dispersions of polymer colloids,
archical Manufacturing (CMMI-1025020) and Xerox Corporation. Macromol. Symp. 206 (2004) 307–320.
These sponsors had no role in study design; in the collection, anal- [30] M. Lattuada, P. Sandkuhler, J. Wu, H. Sefcik, M. Morbidelli, Aggregation kinetics
of polymer colloids in reaction limited regime: Experiments and simulations,
ysis and interpretation of data; in the writing of the report; and in
Adv. Colloid Interface Sci. 103 (2003) 33–56.
the decision to submit the article for publication. [31] R. Amal, J.R. Coury, J.A. Raper, W.P. Walsh, T.D. Waite, Structure and kinetics of
aggregating colloidal haematite, Colloids Surf. A 46 (1990) 1–19.
[32] E.V. Golikova, Y.M. Chernoberezhskii, V.S. Grigorev, M.P. Semov, Aggregate
References stability of the sol prepared from crystalline quartz in aqueous solutions of
potassium chloride, Glass Phys. Chem. 32 (2006) 646–655.
[1] A.D. Karathanasis, D.M.C. Johnson, Stability and transportability of biosolid [33] J. Groenewold, W.K. Kegel, Anomalously large equilibrium clusters of colloids,
colloids through undisturbed soil monoliths, Geoderma 130 (2006) 334–345. J. Phys. Chem. B 105 (2001) 11702–11709.
[2] A.J. Turner, S. Nair, Z. Lai, C.-M. Cheng, S.R. Bhatia, Controlled aggregation of col- [34] J. Groenewold, W.K. Kegel, Colloidal cluster phases, gelation and nuclear matter,
loidal particles for toner applications, J. Appl. Polym. Sci. 122 (2011) 1358–1363. J. Phys.: Condens. Matter 16 (2004) 4877–4886.
[3] Y.H. Kim, S.H. Gihm, C.R. Park, K.Y. Lee, T.W. Kim, I.C. Kwon, H. Chung, S.Y. [35] A.J. Archer, C. Ionescu, D. Pini, L. Reatto, Theory for the phase behaviour of a
Jeong, Structural characteristics of size-controlled self-aggregates of deoxy- colloidal fluid with competing interactions, J. Phys.: Condens. Matter 20 (2008)
cholic acid-modified chitosan and their application as a DNA delivery carrier, 415106.
Bioconjug. Chem. 12 (2001) 932–938. [36] A.K. Atmuri, Ph.D. Thesis, University of Massachusetts Amherst, 2012.
[4] L.L. Ma, M.D. Feldman, J.M. Tam, A.S. Paranjape, K.K. Cheruku, T.A. Larson, J.O. [37] M. Tourbin, C. Frances, Experimental characterization and population balance
Tam, D.R. Ingram, V. Paramita, J.W. Villard, J.T. Jenkins, T. Wang, G.D. Clarke, R. modelling of the dense silica suspensions aggregation process, Chem. Eng. Sci.
Asmis, K. Sokolov, B. Chandrasekar, T.E. Milner, K.P. Johnston, Small multifunc- 63 (2008) 5239–5251.
tional nanoclusters (nanoroses) for targeted cellular imaging and therapy, ACS [38] G.H. Bogush, C.F. Zukoski, Uniform silica particle precipitation: an aggregative
Nano 3 (2009) 2686–2696. growth model, J. Colloid Interface Sci. 142 (1991) 19–34.
332 A.K. Atmuri et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 436 (2013) 325–332

[39] S. Kumar, D. Ramkrishna, On the solution of population balance equations by [41] A. Furukawa, H. Tanaka, Key role of hydrodynamic interactions in colloidal
discretization – I. A fixed pivot technique, Chem. Eng. Sci. 51 (1996) 1311–1332. gelation, Phys. Rev. Lett. 104 (2010) 245702.
[40] N.B. Raikar, S.R. Bhatia, M.F. Malone, M.A. Henson, Experimental studies and [42] X.J. Cao, H.Z. Cummins, J.F. Morris, Hydrodynamic and interparticle potential
population balance equation models for breakage prediction of emulsion drop effects on aggregation of colloidal particles, J. Colloid Interface Sci. 368 (2012)
size distributions, Chem. Eng. Sci. 64 (2009) 2337–2433. 86–96.

You might also like