reaction-limited aggregation in presence of short-range structural forces

13
Reaction-Limited Aggregation in Presence of Short-Range Structural Forces Venkataramana Runkana NSF Industry/University Cooperative Research Center for Advanced Studies in Novel Surfactants, School of Engineering and Applied Science, Columbia University, New York, NY 10027 Tata Research Development and Design Centre, 54-B, Hadapsar Industrial Estate, Pune, 411013 India P. Somasundaran NSF Industry/University Cooperative Research Center for Advanced Studies in Novel Surfactants, School of Engineering and Applied Science, Columbia University, New York, NY, 10027 P. C. Kapur Tata Research Development and Design Centre, 54-B, Hadapsar Industrial Estate, Pune, 411013 India DOI 10.1002/aic.10375 Published online March 3, 2004 in Wiley InterScience (www.interscience.wiley.com). A geometrically discretized sectional population balance model for reaction-limited aggre- gation of colloidal suspensions is presented. The two important model parameters are collision frequency factor and collision efficiency factor. The collision frequency factor is derived from physically realistic arguments proposed for collision of fractal aggregates. The collision efficiency factor is computed as a function of total interaction energy between particles, including short-range structural repulsion forces. The irregular and open structure of aggregates is taken into account by incorporating their mass fractal dimension. The characteristic time constant of reaction-limited aggregation, derived from dynamic scaling of mean aggregate size-aggregation time data, is found to correlate with electrolyte concentra- tion. The population balance model is tested with published experimental data for aggregation of -alumina suspensions in the presence of different electrolytes. It is shown that the slow kinetics of aggregation under certain conditions of pH and electrolyte concentration require inclusion of short-range structural repulsion forces along with van der Waals attraction and electrical double layer repulsion forces in an extended DLVO theory. The model predictions are in good agreement with experimental data for time evolution of mean aggregate diameter in the reaction-limited aggregation regime. © 2005 American Institute of Chemical Engineers AIChE J, 51: 1233–1245, 2005 Keywords: reaction-limited aggregation, population balance modeling, surface forces, fractal aggregates, colloidal suspensions Introduction Aggregation of colloidal suspensions plays an important role in treatment of municipal and industrial water, beneficiation of minerals, manufacture of pulp and paper, processing of paints and pharmaceuticals and in fabrication of ceramics 1, 2, 3, 4 . Transport and fate of contaminants in the aquatic environment depend to a large extent on aggregation of colloids in the presence of naturally occurring salts and polymers 5 . The ag- gregation phenomenon is usually classified in two regimes - diffusion-limited aggregation (DLA) and reaction-limited ag- gregation (RLA) 6 . DLA occurs when collision efficiency of Correspondence concerning this article should be addressed to P. Somasundaran at [email protected] © 2005 American Institute of Chemical Engineers AIChE Journal 1233 April 2005 Vol. 51, No. 4

Upload: venkataramana-runkana

Post on 06-Jun-2016

212 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Reaction-limited aggregation in presence of short-range structural forces

Reaction-Limited Aggregation in Presence ofShort-Range Structural Forces

Venkataramana RunkanaNSF Industry/University Cooperative Research Center for Advanced Studies in Novel Surfactants, School of Engineering and

Applied Science, Columbia University, New York, NY 10027

Tata Research Development and Design Centre, 54-B, Hadapsar Industrial Estate, Pune, 411013 India

P. SomasundaranNSF Industry/University Cooperative Research Center for Advanced Studies in Novel Surfactants, School of Engineering and

Applied Science, Columbia University, New York, NY, 10027

P. C. KapurTata Research Development and Design Centre, 54-B, Hadapsar Industrial Estate, Pune, 411013 India

DOI 10.1002/aic.10375Published online March 3, 2004 in Wiley InterScience (www.interscience.wiley.com).

A geometrically discretized sectional population balance model for reaction-limited aggre-gation of colloidal suspensions is presented. The two important model parameters arecollision frequency factor and collision efficiency factor. The collision frequency factor isderived from physically realistic arguments proposed for collision of fractal aggregates. Thecollision efficiency factor is computed as a function of total interaction energy betweenparticles, including short-range structural repulsion forces. The irregular and open structureof aggregates is taken into account by incorporating their mass fractal dimension. Thecharacteristic time constant of reaction-limited aggregation, derived from dynamic scaling ofmean aggregate size-aggregation time data, is found to correlate with electrolyte concentra-tion. The population balance model is tested with published experimental data for aggregationof �-alumina suspensions in the presence of different electrolytes. It is shown that the slowkinetics of aggregation under certain conditions of pH and electrolyte concentration requireinclusion of short-range structural repulsion forces along with van der Waals attraction andelectrical double layer repulsion forces in an extended DLVO theory. The model predictionsare in good agreement with experimental data for time evolution of mean aggregate diameterin the reaction-limited aggregation regime. © 2005 American Institute of Chemical EngineersAIChE J, 51: 1233–1245, 2005Keywords: reaction-limited aggregation, population balance modeling, surface forces,fractal aggregates, colloidal suspensions

Introduction

Aggregation of colloidal suspensions plays an important rolein treatment of municipal and industrial water, beneficiation of

minerals, manufacture of pulp and paper, processing of paintsand pharmaceuticals and in fabrication of ceramics1, 2, 3, 4.Transport and fate of contaminants in the aquatic environmentdepend to a large extent on aggregation of colloids in thepresence of naturally occurring salts and polymers5. The ag-gregation phenomenon is usually classified in two regimes -diffusion-limited aggregation (DLA) and reaction-limited ag-gregation (RLA)6. DLA occurs when collision efficiency of

Correspondence concerning this article should be addressed to P. Somasundaran [email protected]

© 2005 American Institute of Chemical Engineers

AIChE Journal 1233April 2005 Vol. 51, No. 4

Page 2: Reaction-limited aggregation in presence of short-range structural forces

two clusters is relatively high or close to unity, while RLAprevails at very low collision efficiencies. Because of differ-ences in hydrodynamics of particle/cluster interactions, thekinetics of aggregation are dissimilar in the two regimes. DLAfollows power-law growth kinetics (Figure 1a) while RLAkinetics conforms to an exponential growth law (Figure 1b).

Starting with the classical work of Smoluchowski7, particlepopulation balance has been widely employed to model aggre-gation of colloidal suspensions and evolution of aggregate sizedistribution in time. The two important model parameters in thepopulation balance equation (PBE) are collision frequencyfactor or collision kernel and collision efficiency factor. It isknown that DLA kinetics by Brownian motion is well repre-sented by the Smoluchowski kernel7. Although several expres-sions, both empirical and theoretical, have been proposed, thereis as yet no general agreement on an appropriate kernel forRLA kinetics. Recently, Sandkuhler et al.8 evaluated a numberof kernels and concluded that even though none could representthe experimental data accurately, the product kernel or itsmodification by Odriozola et al.9 is the most appropriate for theprocess. Other authors have concluded that RLA kinetics arebest described by a sum kernel10, 11. In fact, the earlier exper-imental studies on aggregation of polystyrene12 and gold col-loids13 had shown that RLA kinetics could be representedreasonably well by an empirical sum kernel.

The second parameter in the PBE is collision efficiencyfactor. Except in a few instances14, 15, this is either assumed tobe unity or treated as a floating parameter for fitting experi-

mental aggregate size distribution data. The collision efficiencyis a function of the interaction forces between suspended par-ticles and aggregates or clusters, which may include van derWaals or dispersion attraction, electrical double layer (EDL)repulsion and short-range structural repulsion forces. Despiteits known limitations, the classical DLVO theory16, 17 is quitesuccessful in accounting for the role of dispersion and EDLforces in the stability of colloids in the presence of inorganicelectrolytes. However, under certain conditions of pH and saltconcentration, an additional repulsive force, generally termedas structural or hydration force, has been observed betweenseveral materials at short separation distances (� 3–5 nm)18, 19,

20, 21. This short-range hydration repulsion is thought to play animportant role in the fabrication of ceramic components22 andin the stability of latex immunoassays23. More pertinently,there now exists considerable evidence which suggests that theshort-range structural forces can modify the kinetics of aggre-gation, as elaborated in the following paragraph.

Recently, Waite and coworkers24, 25 and van Bruggen et al.26

studied the stability and aggregation behavior of several alu-minum oxide suspensions in the presence of different salts.Beattie et al.24 examined aggregation of suspensions of �-alu-mina (�-Al2O3), boehmite (�-AlOOH, aluminum oxide hy-droxide or alumina monohydrate) and gibbsite (�-Al(OH)3,aluminum trihydroxide or alumina trihydrate). They concludedthat the stability of gibbsite suspensions was similar to that ofoxides, such as hematite, and the critical coagulation concen-tration was close to 50 mM NaCl in pH range 5-6. The otheraluminum oxides, however, exhibited somewhat anomalousbehavior. In neutral pH conditions the suspensions coagulatedat low salt concentrations, while at lower pH levels coagulationtook place only at much higher salt concentrations. At a pH of4.5, aggregation of �-alumina was observed only above 0.5 MKCl concentrations, and of boehmite at 0.13 M and above.Waite et al.25 reported similar results for �-alumina in thepresence of NaCl and KNO3, that is, aggregation was observedat salt concentrations above 0.5 M only. This behavior wasattributed to the adsorption of aluminum tridecamer, Al13O4

(OH)24 (H2O)127� on alumina surface. In addition to the aggre-

gation studies, force measurements by atomic force microsco-py21 between an aluminum-coated silica sphere and a flataluminum substrate have shown the existence of short-range(� 3 nm) repulsive forces in 1 mM KCl at different pH levels.At high pH, the short-range repulsion was attributed to theswelling of sapphire surface and formation of a thick, approx-imately 15 nm, hydrated gel layer adjacent to the aluminasurface. No swelling was observed at low pH and repulsionwas assumed to be due to hydration. The DLVO type forcescould explain the interaction only up to surface separationdistance of about 3 – 5 nm. In order to describe the force-distance profiles at shorter separations, it was necessary toinvoke an additional repulsive force. Similar short-range forceswere also observed between sapphire surfaces in 0.1 M NaBrsolutions at pH 320.

Interestingly, in the studies of Beattie et al.24 and Waite etal.25, aggregation kinetics followed an exponential growth,suggesting that the phenomenon took place under RLA condi-tions, presumably due to low efficiency of the particle/aggre-gate collisions. Although it is not clear whether adsorption ofaluminum tridecamer or any other polymeric species on parti-cle surface or hydration of the surface was responsible for the

Figure 1 Representative aggregate growth curves (a)diffusion-limited aggregation - power lawgrowth, and (b) reaction-limited aggregation -exponential growth.

1234 AIChE JournalApril 2005 Vol. 51, No. 4

Page 3: Reaction-limited aggregation in presence of short-range structural forces

anomalous aggregation behavior, it seems very likely thatrepulsive forces, other than the classical DLVO forces, were inoperation in the aforementioned experiments.

The objective of this work is to develop a population balancemodel for aggregation processes that follow the slow exponen-tial growth kinetics. Toward this end, we employ a geometri-cally sectioned discrete population balance equation for aggre-gation. The collision kernel in the model is based on theoreticalarguments advanced by Ball et al.1, and the collision efficiencyfactor is computed as a function of interaction forces betweenparticles including the short-range structural forces. The modelis tested and verified with experimental data for slow aggrega-tion of alumina suspensions in the presence of different salts asreported by Beattie et al.24 and more recently by Waite et al.25.

Model of Reaction-Limited Aggregation KineticsPopulation balance equation (PBE)

The evolution of floc or aggregate size distribution in time,which is the focus of this study, can be simulated within theframework of particle population balance. The size-continuousform of population balance equation for aggregation in a spa-tially homogeneous batch process is given by7

�n�v, t�

�t� ��

0

��v, u���v, u�n�v, t�n�u, t�du

�1

2 �0

v

��v � u, u���v � u, u�n�v � u, t�n�u, t�du (1)

where n is number concentration of aggregates (or particles), vand u denote aggregate volume, t is aggregation time, � iscollision frequency factor and � is collision efficiency factor.The first term on right-hand side accounts for the loss ordisappearance of aggregates of size v due to their interactionwith aggregates of all sizes. The second term represents the rateof formation of aggregates of size v by interaction between twoaggregates of smaller sizes. This partial integrodifferentialequation does not admit of close form analytical solutionsunless drastically simplified and generally unrealistic collisionkernels are employed. Consequently, the aggregation equationis invariably solved numerically after discretizing and lumpingit into a set of nonlinear ordinary differential equations. Of theseveral discretization and lumping schemes available in theliterature27, 28, we have chosen the method of geometric sec-tions proposed by Hounslow et al.29 because it is computation-ally less intensive and easy to implement. The rate of change ofparticle or floc concentration due to aggregation alone, that is,for processes that do not involve fragmentation of flocs, isgiven by29

dNi

dt� Ni�1 �

j�1

i�2

2j�i�1�i�1, j�i�1, jNj �1

2�i�1,i�1�i�1,i�1Ni�1

2

� Ni �j�1

i�1

2j�i�i, j�i, jNj � Ni �j�i

max

�i, j�i, jNj (2)

where Ni is number concentration of particles or aggregates insection i. The first term on right-hand side accounts for growthdue to aggregation of clusters belonging to all the smallersections, except the immediately adjacent smaller one. Thesecond term represents growth due to aggregation of clustersbelonging to the section immediately smaller than section i.The third term accounts for loss of aggregates due to theirinteraction with those belonging to all the smaller sections, andthe fourth term represents loss of aggregates due to mutualaggregation, and due to their interactions with larger aggre-gates belonging to higher sections. The smallest section corre-sponds to primary particles and max is the section that containsaggregates with the largest characteristic volume. The fre-quency of collisions is directly related to collision volume ofthe aggregates. The characteristic volume Vi of an aggregate insection i is computed as the arithmetic mean of its lower andupper bounds30

Vi �bi � bi�1

2(3)

where bi-1 and bi are the smallest and the largest volumes insection i, respectively. Moreover, the upper bound is computedas a function of the lower one with a geometric sectionalspacing factor of 2

bi � 2bi�1 (4)

The primary particle volume is denoted by V1. An aggregate ofvolume Vi comprises 2i�1 primary particles. The rate of aggre-gation depends on particle concentration and frequency andefficiency of particle/aggregate collisions. The frequency ofcollisions is governed by suspension hydrodynamics and, inaddition, depends on a number of factors such as particle/aggregate size, temperature and solvent viscosity. The effi-ciency of collisions, on the other hand, depends critically onsurface and colloid chemistry of the suspension. It is a functionof pH, electrolyte concentration and electrochemical nature ofparticle surface, besides the variables mentioned previously.

Collision Kernel

In case of DLA, particles are expected to aggregate in thefirst instance of collision only. Hence, the collision frequencyis directly related to collision cross-section and particle diffu-sion coefficient. In case of RLA, however, it is generallybelieved that repeated collisions are required for a successfulevent of aggregation. Accordingly, the aggregates involved inthe collision process sample several sites on each other’s sur-face before attachment actually occurs. Unlike the collisionkernel for DLA, which was derived by Smoluchowski7 ontheoretical considerations quite some time ago, a rigorousexpression does not exist for the RLA kernel. Instead of usingan ad hoc empirical kernel, we have applied the theoreticalarguments of Ball et al.1 to compute the collision frequencyfactor. These authors considered two regimes of aggregatecollisions: first, collisions between clusters of approximatelyequal masses, and second, collisions between large clusters andvery small clusters. It was assumed that these two regimes aresufficient to describe the aggregation rate. Moreover, they

AIChE Journal 1235April 2005 Vol. 51, No. 4

Page 4: Reaction-limited aggregation in presence of short-range structural forces

concluded on basis of theoretical scaling arguments that thecluster sizes are determined mainly by collisions involvinglarge and small clusters rather than those involving clusters ofequal masses. However, Sandkuhler et al.8 recently concludedthat a kernel, derived from such scaling arguments, which isessentially a function of mass of the larger of the two interact-ing clusters, is not accurate enough for reaction-limited aggre-gation. This is not surprising since the final aggregate massdistribution depends strongly on interactions involving clustersof all sizes rather than those involving very small and verylarge clusters only. We therefore modify the treatment of Ballet al.1 in order to account for masses or sizes of both interactingclusters.

Ball et al.1 proposed that the rate coefficient for RLA isdirectly proportional to the volume of phase-space Vf, overwhich the center of one cluster can be positioned such that thetwo clusters are in bondable contact. Two clusters are assumedto be in bondable contact if they are within a microscopicdistance hw, of touching each other, but not overlapping. Thisdistance can be of the order of the repulsive barrier betweentwo colliding particles and is significantly less than the clustersize. For small hw, Vf is directly proportional to the contactsurface of two interacting aggregates. For two solid spheres ofradii r1 and r2, the volume of phase-space is given by1

Vf � 4��r1 � r2�2hw ; hw��r1, r2 (5)

For two fractal aggregates i and j with collision radii rciand rcj

respectively, the collision frequency is then given by

�i, j � �rci � rcj�2 (6)

The above equation can be rewritten as follows

�i, j � KRLA�rci � rcj�2 (7)

where KRLA is a proportionality constant or lumped parameter,which takes into account the effect of temperature and viscosityof the suspension, besides the parameter hw. The collisionradius of an aggregate containing n0 monodisperse primaryparticles of radius r0 is given by31

rci � r0�n0i

CL� 1/dF

(8)

where dF and CL are aggregate fractal dimension and structureprefactor, respectively. Experimental studies on aggregation oflatex suspensions indicate that CL is of the order one32, 33.Finally, substituting Eq. 8 into Eq. 7 yields the followingexpression for the RLA collision kernel

�i, j � KRLAr02� 1

CL� 2/dF

�n0i

1/dF � n0j

1/dF�2 (9)

This equation is substituted into the PBE in Eq. 2 for simulat-ing the RLA kinetics.

Collision efficiency

The probability of attachment, when two aggregates col-lide, depends mainly on the interaction between primaryparticles in the colliding surfaces rather than on particlesresiding in the interior. This is because forces betweenparticles drop rapidly with distance and, as such, interactionbetween aggregates can be approximated by interaction be-tween the primary surface particles34, 35. We take explicitcognizance of this fact by incorporating in our PBE acollision efficiency factor for aggregates which is computedas reciprocal of the modified Fuchs’ stability ratio W be-tween two primary particles k and l36, 37, 38, 39

Wk,l �

�r0k�r0l

Dk,l

exp�VT/kBT�

s2 ds

�r0k�r0l

Dk,l

exp�VvdW/kBT�

s2 ds

(10)

where Dk,l is hydrodynamic correction factor, given by40

Dk,l �6h� 0

2 � 13h� 0 � 2

6h� 02 � 4h� 0

(11)

and VT is total energy of interaction, VvdW is van der Waalsenergy of attraction between two primary particles of radii r0k

and r0l(assumed spherical), s � r0k

� r0l� h0 is distance

between particle centers, h0 is distance of closest approachbetween particle surfaces and h� 0 � 2h0/(r0k

� r0l). Note that

the influence of hydrodynamic forces at very short separationdistances is accounted for in Eq. 10 by including the correctionfactor in Eq. 11.

Since our focus is on aggregation processes in which short-range structural forces play an important role, the classicalDLVO theory is extended by incorporating short-range struc-tural repulsion forces with van der Waals attraction and elec-trical double layer repulsion in the total energy of interactionbetween particles. The expressions employed for these forcesare summarized below.

Interaction forces

The van der Waals energy of attraction between particlesmay be computed from either the Hamaker microscopic theo-ry41 or the Lifshitz macroscopic theory42. The relative merits ofthese theories are discussed elsewhere43. The latter is morerigorous than the former but requires accurate dielectric datafor the materials involved. Because it is easier to implement,we have employed the Hamaker theory to compute the van derWaals energy of attraction Vvdw between a pair of sphericalparticles41

1236 AIChE JournalApril 2005 Vol. 51, No. 4

Page 5: Reaction-limited aggregation in presence of short-range structural forces

VvdW � �A

6 � 2r0kr0l

s2 � �r0k � r0l�2 �

2r0kr0l

s2 � �r0k � r0l�2

� ln�s2 � �r0k � r0l�2

s2 � �r0k � r0l�2�� (12)

where A is Hamaker constant of solids across the solventmedium. This expression does not take into account the retar-dation phenomenon, which sets in at about 5 nm from theparticle surface44. The retardation effect is incorporated bymultiplying the unretarded van der Waals force with a correc-tion function fR proposed by Gregory45

fR�h0� � 1 �gRh0

Rln�1 �

R

gRh0� (13)

where R is characteristic wavelength of interaction and gR is afitting parameter. The value of R is typically 100 nm and thatof gR is 5.3245.

The main reason for the slow kinetics in RLA regime is thelow collision efficiency. In other words, the repulsive forces aregreater in magnitude than the attractive forces although notlarge enough to prevent aggregation completely. One of theseforces arises from the interaction between electrical doublelayers (EDL), which exist at the solid-solution interface. Whentwo particles of same material approach each other, they aresubjected to electrostatic repulsion whose magnitude dependson surface charge density which, in turn, is a function ofsolution pH, temperature, and type and concentration of elec-trolyte species. The EDL interaction between particles can betreated rigorously by the nonlinear Poisson-Boltzmann equa-tion. The solution of this equation is, however, a complex task,which is why several approximations based on linearization ofthe Poisson-Boltzmann equation have been proposed46, 47, 48.We have used the analytical expression proposed by Bell etal.47, which is given by following expression for two sphericalparticles of radii r0k

and r0land surface potentials 0k

and 0l

Vedl � 64��r�0�kBT

zce� 2� r0kr0l

r0k � r0l

� tanh�zce0k

4kBT � tanh�zce0l

4kBT �exp���h0� (14)

where e is elementary charge and zc is valence of the counte-rion and �0 and �r are dielectric constants of a vacuum andsolvent, respectively. The Debye-Huckel parameter � is a func-tion of temperature T, ionic strength I and dielectric constant �r

of solution. It is computed in the usual manner44

� � �2NAvIe2

�0�rkBT�1/ 2

(15)

where NAv is Avagadro’s number and I is ionic strength of thesolution, which is a function of electrolyte concentration andvalence of electrolyte ions.

In addition to the EDL force, short-range repulsion alsocontributes to the low collision efficiency under certain condi-

tions of pH and electrolyte concentration. Force measurementsbetween surfaces of several materials, such as mica, silica andalumina have shown the existence of this force18, 19, 20, 21. Itsupposedly arises due to the structure of water or solventmolecules at the solid-liquid interface and is termed as hydra-tion or solvation repulsion or more generally structural force. Anumber of theories have been proposed for this force but thesehave not been able to explain various experimental observa-tions under different conditions49, 50. In the absence of a defin-itive theoretical expression, the following simple exponentialdecay function is deemed adequate for representing the exper-imental force-distance profiles20, 44

Fhyd � Phydexp��h0/hyd� (16)

where Phyd is termed structural or hydration force constant andhyd is called hydration decay length. Their values have beenfound to be in the ranges of 106 – 5 108 N/m2 and 0.2 – 1.0nm, respectively51. The equation for hydration interaction en-ergy between two spherical particles can be readily derivedfrom Eq. 16, by applying the Derjaguin approximation52, whichfor two unequal spherical particles of radii r0k

and r0iis given

by

Vhyd � �2�r0kr0l

r0k � r0l

�Phydhyd2 exp��h0/hyd� (17)

where Vhyd is short-range hydration or structural interactionenergy.

Model ImplementationAggregation data

The population balance model was implemented and vali-dated against published data24, 25 for aggregation of �-aluminasuspensions in the presence of different salts. Beattie et al.24

conducted a series of experiments at different salt (KCl) con-centrations and measured mean hydrodynamic diameter ofaggregates as a function of time by photon correlation spec-troscopy (PCS). Waite et al.25 also studied aggregation of�-alumina using NaCl and KNO3 electrolytes. The details ofexperimental procedures and results can be found in the citedreferences. Here we limit ourselves to summarizing the exper-imental conditions for the data used in model simulations. Themean radius of primary particles r0 was 60 nm and solidsconcentration in suspension was 12.5 g/L. This corresponds toa primary particle number concentration of approximately4.32 1018 m�3. The experiments were conducted at pH 4.5in solutions of NaCl, KCl and KNO3, at different salt concen-trations while maintaining temperature at 25 1 °C. Themeasured zeta potential of alumina particles was approximately30 mV at pH 4.5 in KNO3 solutions.

Dynamic scaling

It is well known that cluster or aggregate mass distributionsexhibit dynamic scaling, irrespective of whether aggregationoccurs in DLA or RLA regime6, 13. In former case, aggregatemass grows proportionately with time while it increases expo-nentially in the RLA regime. This characteristic behavior is

AIChE Journal 1237April 2005 Vol. 51, No. 4

Page 6: Reaction-limited aggregation in presence of short-range structural forces

widely observed in a variety of particulate materials such asgold, silica and polystyrene latex6, 53, 54. We have utilized theprinciple of dynamic scaling to verify if aggregation of alumi-num oxides took place under the universal slow RLA kinetics,wherein aggregate mass grows exponentially with time infollowing manner6

M � exp� t

RLA� (18)

where M is aggregate mass and RLA is characteristic timeconstant of RLA, which depends on experimental conditions. Ifan aggregate is a fractal object and its mass scales with size,then the characteristic aggregate size rn also has an exponentialrelationship with time

rn � exp� 1

RLAdFt� (19)

Hence, a plot of mean aggregate size against time should be astraight line on a log-linear plot if aggregation follows the RLAkinetics. The experimental data of Beattie et al.24 and Waite etal.25 are plotted in Figures 2–4. It will be observed that �-alu-mina aggregates exhibit dynamic scaling behavior at all times,and the aggregation phenomenon does follow the exponentialgrowth kinetics provided salt concentrations are less than 0.8M. The experimental data show a nonlinear relationship be-tween aggregate size and time when salt concentrations areequal to or more than 0.8 M. It would seem that the processgoes through a transition between RLA and DLA regimes as itapproaches the fast DLA kinetics. However, in view of the factthat the plots are piece-wise linear, there is also a strongpossibility of aggregate restructuring during aggregation. Inthis instance, how and when restructuring is triggered re-mains a question mark. Waite et al.25 have also measuredaggregate fractal dimensions and found them to be 2.09 and2.1 at 0.6 M and 0.7 M NaCl, respectively. These values aresimilar to fractal dimensions observed for aggregates result-ing by RLA54, 55, 56. A constant value of 2.1 for fractal

dimension of all aggregates is employed in our analysissince it does not change with electrolyte concentration in theexperimental studies of Waite et al.25. Substitutions of frac-tal dimension and slope of straight lines in Figures 2– 4 inEq. 19 yields an estimate of RLA time constant RLA, whichis plotted against salt concentration in Figure 5. It will beseen that RLA decreases with salt content in a nonlinearfashion. This is an interesting result in the sense that abovea critical concentration of 0.5 M the aggregation kinetics iscontrolled primarily by salt concentration. This behavior issimilar to that normally observed in case of coagulation ofcolloidal suspensions. We are, therefore, obliged to con-clude that there is a repulsive barrier, besides the electricaldouble layer force, that must be overcome in order foraggregation to take place.

Figure 2 Semi-log plots for dynamic scaling of meanhydrodynamic diameter of �-alumina aggre-gates in NaCl solutions at pH 4.5, showing ex-ponential growth kinetics (data from Waite etal.25).

Figure 3 Semi-log plots for dynamic scaling of meanhydrodynamic diameter of �-alumina aggre-gates in KNO3 solutions at pH 4.5, showingexponential growth kinetics and shift in slopeat high salt concentrations (data from Waite etal.25).

Figure 4 Semi-log plots for dynamic scaling of meanhydrodynamic diameter of �-alumina aggre-gates in KCl solutions at pH 4.5, showing ex-ponential growth kinetics and shift in slope athigh salt concentrations (data from Beattie etal.24).

1238 AIChE JournalApril 2005 Vol. 51, No. 4

Page 7: Reaction-limited aggregation in presence of short-range structural forces

Interaction energy between particles and hydration forceparameters

The aggregation model has three adjustable parameters, hy-dration force constant (Phyd), hydration decay length (hyd) andproportionality constant in the collision kernel (KRLA). Becauseof mutual compensation, the conventional approach to estima-tion of parameters together by minimizing an error function ofa highly nonlinear model can lead to correlated and misleadingvalues. We have, therefore, employed a systematic two-stagesearch strategy in order to ensure physically meaningful pa-rameter values of the underlying phenomenon. In first stage,the interaction energy-surface separation diagram was utilizedas a guide to determine hydration force parameters. In thesecond stage, the proportionality constant KRLA was adjusted tomatch model simulation results with experimental data asclosely as possible.

In order to map the interaction energy-surface separationdiagram, it is necessary to have representative values for var-ious inputs, such as Hamaker constant and particle surfacepotential. Ducker et al.20 fitted forces between sapphire sur-faces using a nonretarded Hamaker constant of 6.7 10�20 J.The same value was used in the Hamaker theory in Eq. 12 forcomputing van der Waals attraction between �-alumina parti-cles across water medium. The retardation factor was calcu-lated from Eq. 13. The particle surface potential was assumedequal to zeta potential. The experimentally measured zeta po-tential of approximately 30 mV at pH 4.525 was employed forestimating the EDL repulsion, that is, 0k

� 0l� 30 mV. The

energy curves were calculated for interaction between sphericalparticles of r0k

� r0l� 60 nm radius. As shown in Figure 6, at

a KNO3 concentration of 0.5 M the classical DLVO theory - inwhich total interaction energy is sum of van der Waals attrac-tion and EDL repulsion energies only - predicts a strongattraction. It was, however, experimentally observed that ag-gregation did not take place until salt concentration is greaterthan 0.5 M. It is then reasonable to conclude that it is necessaryto include a short-range repulsion force so as to precludecoagulation at this salt concentration. For this purpose, initialguesses of the two parameters in the structural repulsion equa-tion were chosen from the ranges of their values that have beenreported in literature. These values were next adjusted to en-sure that the structural repulsion dominates the interaction.

After some trial and error, the values of Phyd and hyd werefound to be approximately 1.9 107 N/m2 and 0.6 nm,respectively, which are well within the ranges determined bysurface force measurements51. Interestingly, the fitted hydra-tion decay length of 0.6 nm is approximately equal to twice thediameter of water molecules, 0.275 nm. Using these parametervalues, total interaction energy was computed as a sum of vander Waals attraction, EDL repulsion and structural repulsion. Itwill be observed from Figure 7 that the suspension is nowpredicted to be stable at 0.5 M KNO3, as observed experimen-tally. However, it is close to the region where the suspensionbecomes unstable if double layer repulsion is suppressed. Sim-ulation studies show that stability ratio W is quite sensitive tohydration force parameters. An increase in Phyd results in anincrease in W and enhanced suspension stability. Similarly, areduction in Phyd reduces W and increases the instability ofsuspension. On the other hand, hyd has an opposite effect onW.

Figure 5 Variation of characteristic time constant of re-action-limited aggregation with salt concentra-tion.

Figure 6 Interaction energy-surface separation diagramfor �-alumina particles computed from DLVOtheory without including short-range structuralrepulsion, using Eq. 12 for VvdW and Eq. 14 forVedl.

Figure 7 Interaction energy-surface separation diagramfor �-alumina particles computed from ex-tended DLVO theory, including short-rangestructural repulsion, using Eq. 12 for VvdW, Eq.14 for Vedl and Eq. 17 for Vhyd.

AIChE Journal 1239April 2005 Vol. 51, No. 4

Page 8: Reaction-limited aggregation in presence of short-range structural forces

Model Validation

The discretized population balance equation in Eq. 2 wasdivided into 30 geometric sections, and the resulting 30 non-linear ordinary differential equations were solved simulta-neously by Gear’s predictor-corrector numerical technique57.The stability ratio in Eq. 10 was evaluated by numerical inte-gration using the Romberg algorithm58, with r0k

� r0l� 60 nm

and 0k� 0l

� 30 mV. During integration of populationbalance equations, the computed aggregate size distributionwas tested at each time step for conservation of solid volume.The loss of total volume was less than 1% for the resultspresented here. The model was tested by comparing experi-mental data with model simulated time evolution of mass meanaggregate size, calculated in the following manner14

r� �

�i

yNirci

4

�i

yNirci

3 (20)

where yNi is number fraction of aggregates belonging to sectioni and r� is mass mean aggregate radius.

The simulated and experimental aggregate growth curves areshown in Figures 8, 9 and 10 for aggregation of �-alumina inthe presence of NaCl, KNO3 and KCl, respectively. It will beseen that the computed growth curves are in reasonable agree-ment with the experimental data. The discrepancy betweenmodel predictions and experimental data at 0.8 M NaCl andKNO3 is mainly because the aggregation process was assumedto follow RLA kinetics whereas it was actually in the transitionregion from RLA to DLA kinetics. That the process is in thetransition regime is indicated by the linear growth behaviorobserved experimentally at 0.8 M in Figures 8 and 9. Similarresults were reported by Weitz et al.56 for aggregation of goldcolloids. Since alumina is heavier than water, there is also apossibility of differential sedimentation influencing aggregategrowth. In order to verify this, the collision kernel for aggre-

gation at 0.8 M NaCl and KNO3 was computed as a sum ofcontributions due to perikinetic aggregation (Eq. 9) and differ-ential sedimentation. The collision frequency in the lattermechanism is given by59

�i, jDS �

2�g

9��rci � rcj�

2rci

2��i � �l� � rcj

2��j � �l� (21)

where �i and �l are densities of aggregate and fluid, respec-tively, and g is acceleration due to gravity. The aggregatedensity �i decreases as its size increases and it can be estimatedusing the following60

�i � �0�rci

r0� dF�3

(22)

Figure 8 Simulated and experimental time evolution ofmean hydrodynamic diameter of �-alumina ag-gregates in NaCl solutions at pH 4.5 (solidlines - Eq. 9; dotted lines - Eq. 23; symbols -data from Waite et al.25).Dashed lines are simulation results for 0.8 M NaCl includingcontribution from differential sedimentation, Eq. 21.

Figure 9 Simulated and experimental time evolution ofmean hydrodynamic diameter of �-aluminaaggregates in KNO3 solutions at pH 4.5 (solidlines - Eq. 9; dotted lines - Eq. 23; symbols -data from Waite et al.25).Dashed lines are simulation results for 0.8 M KNO3 includingcontribution from differential sedimentation, Eq. 21.

Figure 10 Simulated and experimental time evolution ofmean hydrodynamic diameter of �-aluminaaggregates in KCl solutions at pH 4.5 (solidlines - Eq. 9; dotted lines - Eq. 23; symbols -data from Beattie et al.24).

1240 AIChE JournalApril 2005 Vol. 51, No. 4

Page 9: Reaction-limited aggregation in presence of short-range structural forces

where �0 is density of primary particles. This equation isstrictly applicable for linear portion of the aggregate size-density curve. The simulation results obtained with this mod-ification are shown as dashed lines in Figures 8 and 9 at 0.8 MNaCl and KNO3, respectively. It can be observed that differ-ential sedimentation does not contribute significantly to aggre-gate growth, presumably because aggregate density decreasesas its size increases and aggregates in this study do not growlarge enough for differential sedimentation to become signifi-cant. Consequently, transition regime is a more likely expla-nation for the process behavior under the stated conditions.

The deviation between computed and experimental results at0.7 M KNO3 in Figure 9 at longer aggregation times could bedue to restructuring of aggregates. This is evident from theslight change in slope in Figure 3 for the same data after about40 min of aggregation. Aggregate restructuring could alsoexplain deviation between theory and experiment for the data at0.8 M KCl in Figure 10. It can be seen from Figure 4 that theslope changes after about 30 min of aggregation. Our simula-tions of the aggregation process assume a constant fractaldimension of 2.1. It is, however, possible to obtain a better fitby employing a fractal dimension smaller than 2.1 at shorteraggregation times and a larger value at longer times. In fact,Selomulya et al.61 developed an empirical expression for theevolution of aggregate fractal dimension with time, and fitted itto their experimental data for flocculation of latex particles incouette-flow. In view of the fact that the fitted parameters arespecific to the experimental conditions employed, it was notpossible to utilize the empirical expression of Selomulya etal.61 in our analysis. Although there have been some studies onaggregate restructuring62, 63, the phenomenon is not understoodwell enough to predict as to when and under what conditions itoccurs.

The fitted proportionality constant in the collision kernelKRLA and stability ratio, computed from Eq. 10, are included inTable 1. It is worth noting that computed stability ratios arecomparable to experimentally determined values, although fora different material, polymer latex64. The reported stabilityratios are within the range of 2.67 105 – 4.41 106 fordifferent solid and electrolyte concentrations. Note also theconsistent trend in the variation of KRLA and W with saltconcentration in all cases analyzed. The results in Figures 8 –10 strongly suggest that the collision kernel, proposed in Eq. 9,

is able to represent the RLA kinetics quite accurately, at leastfor aggregation of alumina suspensions under the specifiedexperimental conditions. Nevertheless, it is of some interest tocarry out model simulation of these data with other prominentkernels in the literature. As mentioned earlier, Sandkuhler etal.8 suggested that the following kernel by Odriozola et al.9 isthe most suitable among the kernels proposed for RLA

�i, j ��Br�n0i

1/dF � n0j

1/dF��n0i

�1/dF � n0j

�1/dF� N11�n0in0j�O

4�1 � W�1�N11�n0in0j�O � 1 �

(23)

This kernel contains two adjustable parameters N11, whichrepresents mean number of monomer collisions per encounterand O, which is a numerical constant. Initially, we simulatedaggregation of alumina suspensions employing this kernel withN11 � 6.1 and O � 0.4, as reported by Lattuada et al.64 forpolystyrene latex, and the stability ratio values given in Table1. The computed results turned out to be significantly differentfrom the experimental data. Consequently, the two kernel pa-rameters were adjusted for an acceptable fit. After some trialand error, O was found to be 0.5, and N11 was found todecrease with increasing electrolyte concentration (see Table1). The results obtained with the Odriozola et al.9 kernel areshown as dotted lines in Figures 8 – 10. Although this kernelmatches experimental data at lower salt concentrations well, italso fails to predict aggregation kinetics accurately at 0.8 MNaCl and KNO3 in Figures 8 and 9, respectively. It should benoted here that the kernel in Eq. 9 employed by us involvesonly one adjustable parameter KRLA, while the Odriozola et al.9 kernel contains two parameters N11 and 0, which need to beadjusted depending on experimental conditions. As shown inTable 1, N11 varied with electrolyte concentration, while Lat-tuada et al.64 found that O changed with initial particle con-centration.

Our kernel was next compared with a class of kernels em-ployed previously for describing the RLA kinetics. There canbe recast in a general form as follows8

�i, j � �BrBi, jPi, j (24)

Table 1. Computed Stability Ratio and Fitted RLA Kinetics Parameters

FigureNo. Reference Salt

SaltConcentration

(M)Stability

Ratio KRLA (m/s) N11

8 [25] NaCl 0.5 9.49 106 7.0 10�3 360.6 1.78 106 5.0 10�3 210.65 0.88 106 3.5 10�3 150.7 0.46 106 3.4 10�3 140.8 0.15 106 2.2 10�3 9

9 [25] KNO3 0.5 9.49 106 7.0 10�3 360.6 1.78 106 5.0 10�3 210.7 0.46 106 2.1 10�3 90.8 0.15 106 2.1 10�3 9

10 [24] KCl 0.5 9.49 106 7.0 10�3 400.6 1.78 106 2.5 10�3 120.7 0.46 106 1.0 10�3 4.50.8 0.15 106 1.0 10�3 4.5

Note: The stability ratio was computed using Eq. 10, for r0k� r0l

� 60 nm and 0k� 0l

� 30 mV.

AIChE Journal 1241April 2005 Vol. 51, No. 4

Page 10: Reaction-limited aggregation in presence of short-range structural forces

where

Bi, j ��n0i

1/dF � n0j

1/dF��n0i

�1/dF � n0j

�1/dF�

4(25)

and �Br is the constant aggregation kernel or the collisionfrequency factor for two equal spheres in Brownian motion

�Br �8kBT

3�(26)

The kernels differ mainly in the expression for Pi,j (see Table2). Simulation results obtained using different kernels are com-pared with experimental data for aggregation with 0.7 M NaClin Figure 11. It was possible to match the experimental datawith our kernel, Eq. 9 and that of Odriozola et al., Eq. 23 only,with stability ratio and kernel parameters given in Table 1, asshown previously in Figure 8. All the other kernels, given byEq. 24 along with Pi,j and parameters in Table 2, predictextremely slow aggregate growth at the same stability ratio, asseen in lines at the bottom and the insert in Figure 11. It was notpossible to fit these kernels to experimental data even afteradjusting the exponent values in them. The model predictionsare, however, quite sensitive to parameters in these kernels.

Since Figure 11 does not reveal enough information about thebehavior of different kernels to permit a meaningful compari-son among them, simulations were carried out at a lowerstability ratio of 104 where noticeable aggregate growth occurswith all kernels, as shown in Figure 12. With KRLA � 10�4, ourkernel, Eq. 9 predicts exponential growth, whereas the sumkernel, Eq. 30 leads to linear growth. The Smoluchowskikernel, Eq. 31 predicts power-law growth even at this highstability ratio. The rest predict exponential growth. The modelpredictions are quite sensitive to parameters in Odriozola et al.kernel, Eq. 23. It was already shown in Figures 8 –10 thatexponential growth is predicted by this kernel with stabilityratios and parameter values given in Table 1. However, it wasalso possible to obtain both linear as well as power-law growthby manipulating the parameters. For example, the model pre-dicts power-law growth with N11 � 2 and O � 0.2, as shownin Figure 12 (dotted line). Hence, only highly specializednumerical values of these parameters can lead to reaction-limited aggregation kinetics. On the other hand, our kernel, Eq.9 predicts exponential growth at different stability ratios rep-resentative of RLA kinetics over a reasonably large range ofnumerical values of KRLA.

Discussion

It is evident from the results shown in Figures 8-10 that thekernel employed in our model does represent the slow, expo-nential growth kinetics, typical of RLA, quite well. Unlike thekernels used previously for RLA, which are mostly empiricalin nature, this kernel is based on theoretical proposition that thefrequency of collisions between fractal aggregates depends onthe total surface area of the interacting aggregates. The param-eter KRLA in the collision kernel is not completely arbitrary asit is apparently related to salt concentration. Even thoughsolvent viscosity increases, the repulsive barrier between par-ticles decreases with an increase in salt concentration and thisis possibly the reason for decrease in KRLA with salt content, asseen in Table 1. This is also an indirect indication that the

Table 2. Kernels Proposed in Literature forReaction-Limited Aggregation Kinetics

EquationNo. Pi,j Parameters Reference

27[n0i

� (n0i

1/dF � 1)dF] [n0j

� (n0j

1/dF � 1)dF] [77]28 (max(n0i

, n0j))B Bi,j � 1 [1]

B � 1.029 (n0i

n0j)P P � 0.5 [78]

30 (n0i

1/dF � n0j

1/dF)/2 [8]31 1 [7]

Figure 11 Time evolution of mean aggregate diameterpredicted by various RLA kernels for aggre-gation of �-alumina in 0.7 M NaCl solutions atpH 4.5 (solid lines - Eq. 9; dotted lines - Eq. 23;symbols - data from Waite et al.25).The same plot on an expanded y-axis in insert shows theextremely slow kinetics predicted by kernels other than inEqs. 9 and 23. See text for details of other kernels.

Figure 12 Time evolution of mean aggregate diameterpredicted by various RLA kernels at a stabilityratio of 104.The general expression for kernels is given in Eq. 24 andexpressions for Pi,j in Table 2. Eq. 27 - Schmitt et al.77; Eq.28 - Ball et al.1; Eq. 29 - product kernel78; Eq. 9 - kernel inpresent work, with KRLA � 10�4; Eq. 23 - Odriozola et al.9with N11 � 2 and O � 0.2; Eq. 30 - sum kernel8; Eq. 31 -Smoluchowski kernel7.

1242 AIChE JournalApril 2005 Vol. 51, No. 4

Page 11: Reaction-limited aggregation in presence of short-range structural forces

repulsive forces dictate the kinetics of reaction-limited aggre-gation, rather than diffusion of particles/aggregates. The sen-sitivity of aggregate growth to KRLA is shown in Figure 13 for0.6 M and 0.7 M NaCl concentrations. The solid lines areobtained with best-fit values given in Table 1 and the dashedlines on either side of the solid lines are with 5% of best-fitKRLA. The mean aggregate diameter increases with KRLA, and itis more sensitive to KRLA at 0.7 M, presumably because thegrowth rate is higher at 0.7 M than at 0.6 M NaCl. Althoughthis model can be used to predict aggregation kinetics thatfollows exponential growth, it is necessary to adjust KRLA tomatch experimental data at different electrolyte concentrations.Experimental studies on RLA are relatively scarce as comparedto DLA. One could possibly derive a useful expression forKRLA as a function of experimental conditions, especially elec-trolyte concentration, if more data on RLA kinetics are avail-able.

The hydration force parameters are not only physicallymeaningful, but are also within the range determined by sur-face force measurements. Moreover, their values are decoupledfrom the proportionality constant in the collision kernel KRLA.The stability ratios computed from force parameters are rea-sonably close to experimentally determined values64. Unfortu-nately, with the current status of theories of structural forces, itis not possible to predict a priori under what conditions of pHand electrolyte concentration the short-range structural forcesbecome operative and or predominant. It is well known thatpH, electrolyte type and its concentration have a profoundeffect on coagulation and flocculation. For example, in case ofmetal or ceramic oxides, rate of aggregation increases as pHapproaches the isoelectric point (IEP) of the oxide and de-creases as pH moves away from IEP. Tjipangandjara et al.65

found that the percent of alumina settled is maximum at pH 9,the IEP of alumina. This is mainly because aggregate mean sizeis also maximum at the IEP, as demonstrated by Rattanakawinand Hogg66. Similarly, aggregate growth rate of hematite sus-pensions increased with pH, became maximum at the IEP ofhematite and then decreased at higher pH67, 68. The main reason

for the dependence of aggregation kinetics on pH is due to itseffect on surface potential of oxide materials. Both alumina andhematite are positively charged below the IEP and negativelycharged at higher pH69. The magnitude of electrical doublelayer repulsion depends strongly on surface potential and es-sentially determines the kinetics of aggregation and aggregatefractal dimension. For example, Herrington and Midmore70

found that the fractal dimension of kaolinite flocs decreasedsharply from 2.8 at pH 3 to about 1.8 at pH 4 in pure water andfrom 2.9 at pH 3.5 to about 1.8 at pH 4.5 in 1 mM KCl solution.Similar results were reported for bentonite flocs71. The aggre-gate fractal dimension depends not only on pH but also onother variables, such as electrolyte concentration14, 72, initialparticle concentration72 and temperature73, 74. Since a phenom-enological model to predict fractal dimension is not available,we are obliged to treat it as an experimentally measured modelinput, as have other researchers also.

Finally, even though it is possible to predict reaction-limitedaggregation kinetics in the presence of different electrolytesreasonably well, this model does not explicitly account for theinfluence of type or nature of co- or counter-ions. In principle,however, the model can be extended by incorporating a surfacecomplexation or site-binding model of Leckie and coworkers75,76

which can predict the surface or solid-liquid interface potentialas a function of the nature of counterions.

Conclusions

Reaction-limited aggregation follows slow, exponentialgrowth kinetics. This is mainly because efficiency of particle oraggregate collisions is very low, which is due to the presenceof repulsive electrical double layer and short-range structural orhydration forces. We have presented a population balancemodel for reaction-limited aggregation of colloidal suspen-sions, and tested it with experimental data for aggregation ofalumina suspensions in the presence of different electrolytes.Instead of employing an empirical collision kernel, a theoret-ical kernel, derived from arguments of Ball et al.1 and based ontotal surface area of interacting aggregates, is proposed tomodel the slow, exponential aggregate growth kinetics. Thiskernel represents experimental data more accurately than thekernels previously suggested in the literature.

Another feature of this model is that it incorporates a theo-retical model for predicting the collision efficiency factor,instead of using it as a fitting parameter. The collision effi-ciency factor was computed as a function of surface forces,namely, van der Waals attraction, electrical double layer repul-sion and short-range structural repulsion. This allows us to takeinto account the effect of important process variables, such assolution pH and electrolyte concentration on aggregation,which was not possible with previous population balance mod-els.

AcknowledgmentsThis work was supported by the National Science Foundation (NSF

Grants # INT-96-05197 and INT-01-17622) and the NSF Industry/Univer-sity Cooperative Research Center (IUCRC) for Advanced Studies in NovelSurfactants at Columbia University (NSF Grant # EEC-98-04618). Theauthors thank the management of Tata Research Development and DesignCentre for the permission to publish this article. VR thanks Prof.E.C.Sub-barao, Prof.Mathai Joseph and Dr.Pradip for their advice and encourage-ment.

Figure 13 Sensitivity of model predictions to KRLA usingthe present kernel, Eq. 9 for aggregation of�-alumina in 0.6 M and 0.7 M NaCl solutions atpH 4.5.Symbols are experimental data from Waite et al.25. Solidlines are model predictions using the best-fit KRLA given inTable 1. Dashed lines are simulation results with 5% ofbest-fit KRLA. Aggregate growth rate increases with KRLA.

AIChE Journal 1243April 2005 Vol. 51, No. 4

Page 12: Reaction-limited aggregation in presence of short-range structural forces

Notation

A � Hamaker constant of solid particles across a solvent, Jbi � upper boundary volume of section i, m3

CL � aggregate structure prefactor, dimensionlessD � hydrodynamic correction factor, dimensionlessdF � fractal dimension of flocs, dimensionless

e � elementary charge, CFhyd � hydration/structural repulsion force between two flat surfaces,

N/m2

fR � retardation correction function, dimensionlessg � acceleration due to gravity, m/sec2

gR � parameter in the retardation correction function, dimensionlessh0 � distance of closest approach between particle surfaces, mh�0 � normalized distance of closest approach, dimensionlesshw � distance over which two clusters touch each other, m

I � ionic strength of solution, mol/m3

KRLA � proportionality constant in RLA kernel, m/skB � Boltzmann constant, J/KM � mass of an aggregate, kg

max � maximum number of sections considered in population balancen � number concentration of particles or aggregates, m�3

n0 � number of primary particles in an aggregate, dimensionlessNAv � Avagadro’s number, mol�1

Ni � number concentration of particles or aggregates in section i, m-3

N11 � mean number of monomer collisions per encounter, dimension-less

Phyd � hydration/structural force constant, N/m2

rci� collision radius of an aggregate belonging to section i, m

r0 � mean radius of primary particles, mrn � characteristic size of an aggregate, mr� � mass mean radius of an aggregate, m

r1, r2 � radii of two spherical particles 1 and 2, ms � particles’ center-to-center separation distance, mt � aggregation time, sec

T � suspension temperature, Ku � volume of a particle or an aggregate, m3

v � volume of a particle or an aggregate, m3

Vf � volume of phase-space in which two clusters are in bondablecontact, m3

Vhyd � hydration/structural repulsion energy between two particles, JVi � characteristic volume of an aggregate in section i, m3

Vint � interaction energy between two primary particles, JVvdW � van der Waals attraction energy between two primary particles, JVedl � electrical double layer repulsion energy between two primary

particles, JVT � total interaction energy between two primary particles, JW � stability ratio, dimensionless

yNi� number fraction of aggregates belonging to section i, dimension-

lesszc � counterion valence, dimensionless

Greek letters

� � collision efficiency factor, dimensionless� � collision frequency factor, m3/s

�iDS � collision frequency due to differential sedimentation, m3/s

�Br � collision frequency factor for two equal spheres in Brownianmotion, m3/s

�0 � dielectric constant of free space C/m-V�r � dielectric constant of water or solvent, C/m-V� � Debye-Huckel parameter, m�1

O � numerical constant in Odriozola kernel, dimensionlesshyd � hydration decay length, m

R � characteristic wave length of interaction, m� � dynamic viscosity of the suspending fluid, kg/m-sec�0 � density of primary particles, kg/m3

�i � density of an aggregate in section i, kg/m3

�l � fluid density, kg/m3

0 � particle surface potential, V RLA � characteristic time constant of RLA, min

Literature Cited1. Ball RC, Weitz DA, Witten TA, Leyvraz F. Universal kinetics in

reaction-limited aggregation. Phys Rev Lett. 1987;58:274-277.2. Dobias B. Coagulation and Flocculation: Theory and Applications,

New York: Marcel Dekker, Inc., 1993.3. Pugh RJ, Bergstrom L. Surface and Colloid Chemistry in Advanced

Ceramics Processing. New York: Marcel Dekker, Inc., 1994.4. Somasundaran P, Das KK, Yu X. Selective flocculation. Curr Opinion

Colloid Interface Sci. 1996;1:530-534.5. Jackson GA, Burd AB. Aggregation in the marine environment. Env

Sci Tech. 1998;32:2805-2814.6. Lin MY, Lindsay HM, Weitz DA, Ball RC, Klein R, Meakin P.

Universality in colloid aggregation. Nature. 1989;339:360-362.7. Smoluchowski, Mv. Versuch einer mathematischen theorie der koagu-

lationskinetik kolloider losungen. Zietschrift fur Physikalische Che-mie. 1917;92:129-168.

8. Sandkuhler P, Sefcik J, Lattuada M, Wu H, Morbidelli M. Modelingstructure effects on aggregation kinetics in colloidal dispersions.AIChE J. 2003;49:1542-1555.

9. Odriozola G, Moncho-Jorda A, Schmitt A, Callejas-Fernandez J, Mar-tinez-Garcia R, Hidalgo-Alvarez R. A probabilistic aggregation kernelfor the computer-simulated transition from DLCA to RLCA. EurophysLett. 2001; 53:797-803.

10. van Dongen PG, Ernst MH. Dynamic scaling in the kinetics of clus-tering. Phys Rev Lett. 1985;54:1396-1399.

11. Martin JE, Wilcoxon JP, Schaefer D, Odinek J. Fast aggregation ofcolloidal silica. Phys Rev A. 1990;41:4379-4391.

12. von Schulthess GK, Benedek GB, De Blois RW. Measurement of thecluster size distributions for high functionality antigens crosslinked byantibody. Macromolecules. 1980;13:939-945.

13. Weitz DA, Lin MY. Dynamic scaling of cluster-mass distributions inkinetic colloid aggregation. Phys Rev Lett. 1986;57:2037-2040.

14. Amal R, Coury JR, Raper JA, Walsh WP, Waite TD. Structure andkinetics of aggregating colloidal hematite. Colloids Surfaces A. 1990;46:1-19.

15. Runkana V, Somasundaran P, Kapur PC. Mathematical modeling ofpolymer-induced flocculation by charge neutralization. J Colloid In-terface Sci. 2004;270:347-358.

16. Derjaguin BV, Landau LD. A theory of the stability of stronglycharged lyophobic sols and of the adhesion of strongly charged par-ticles in solutions of electrolytes. Acta Physiochim. USSR. 1941;14:633-662.

17. Verwey EJW, Overbeek JThG. Theory of the Stability of LyophilicColloids. Amsterdam: Elsevier, 1948.

18. Pashley RM. DLVO and hydration forces between mica surfaces inLi�1, Na�1, K�1, and Cs�1 electrolyte solutions in the range 1-100nm. J Colloid Interface Sci. 1981;83:531-546.

19. Pashley RM, Israelachvili JN. DLVO and hydration forces betweenmica surfaces in Mg�2, Ca�2, Sr�2, and Ba�2 chloride solutions. JColloid Interface Sci. 1984;97:446-455.

20. Ducker WA, Xu Z, Clarke DR, Israelachvili JN. Forces betweenalumina surfaces in salt solutions: non-DLVO forces and the implica-tions for colloidal processing. J Am Ceram Soc. 1994; 77:437-443.

21. Karaman ME, Pashley RM, Waite TD, Hatch SJ, Bustamante H. Acomparison of the interaction forces between model alumina surfacesand their colloidal properties. Colloids Surfaces A. 1997;129-130:239-255.

22. Velamakanni BV, Chang JC, Lange FF, Pearson DS. New method forefficient colloidal particle packing via modulation of repulsive lubri-cating hydration forces. Langmuir. 1990;6:1323-1325.

23. Molina-Bolivar JA, Ortega-Vinuesa JL. How proteins stabilize colloi-dal particles by means of hydration forces. Langmuir. 1999;15:2644-2653.

24. Beattie JK, Cleaver JK, Waite TD. Anomalous aggregation behaviorof aluminium oxyhydroxides. Colloids Surfaces A. 1996;111:131-138.

25. Waite TD, Cleaver JK, Beattie JK. Aggregation kinetics and fractalstructure of �-alumina assemblages. J Colloid Interface Sci. 2001;241:333-339.

26. van Bruggen MPB, Donker M, Lekkerkerker HNW, Hughes TL.Anomalous stability of aqueous boehmite dispersions induced byhydrolyzed aluminium poly-cations. Colloids Surfaces A. 1999;150:115-128.

27. Ramkrishna D. Population Balances: Theory and Applications to

1244 AIChE JournalApril 2005 Vol. 51, No. 4

Page 13: Reaction-limited aggregation in presence of short-range structural forces

Particulate Systems in Engineering. New York: Academic Press,2000.

28. Vanni M. Approximate population balance equations for aggregation -breakage processes. J Colloid Interface Sci. 2000;221:143-160.

29. Hounslow MJ, Ryall RL, Marshall VR. A discretized populationbalance for nucleation, growth, and aggregation. AIChE J. 1988;34:1821-1832.

30. Spicer PT, Pratsinis SE. Coagulation and fragmentation: universalsteady-state particle-size distribution. AIChE J. 1996;42:1612-1620.

31. Flesch JC, Spicer PT, Pratsinis SE. Laminar and turbulent shear-induced flocculation of fractal aggregates. AIChE J. 1999;45:1114-1124.

32. de Boer GBJ. Coagulation in Stirred Tanks. PhD Thesis. EindhovenUniversity of Technology. The Netherlands. 1987.

33. Ducoste JJ. A two-scale PBM for modeling turbulent flocculation inwater treatment processes. Chem Eng Sci. 2002;57:2157-2168.

34. Firth BA, Hunter RJ. Flow properties of coagulated colloidal suspen-sions I. energy dissipation in the flow units. J Colloid Interface Sci.1976;57:248-256.

35. Kusters KA, Wijers JG, Thoenes D. Aggregation kinetics of smallparticles in agitated vessels. Chem Eng Sci. 1997;52:107-121.

36. Fuchs N. Ueber die stabilitat und aufladung der aerosole. Zeitschriftfur Physik. 1934;89:736-743.

37. Derjaguin BV, Muller VM. Slow coagulation of hydrophobic colloids.Dokl Akad Nauk SSSR. 1967;176:738-741.

38. McGown DN, Parfitt GD. Improved theoretical calculation of thestability ratio for colloidal systems. J Phys Chem. 1967;71:449-450.

39. Spielman LA. Viscous interactions in Brownian coagulation. J ColloidInterface Sci. 1970;33:562-571.

40. Honig EP, Roebersen GJ, Wiersema PH. Effect of hydrodynamicinteraction on the coagulation rate of hydrophobic colloids. J ColloidInterface Sci. 1971;36:97-109.

41. Hamaker HC. The London-van der Waals attraction between sphericalparticles. Physica. 1937;IV:1058-1072.

42. Lifshitz EM. The theory of molecular attractive forces between solids.Soviet Physics JETP. 1956;2:73-83.

43. Bowen WR, Jenner F. The calculation of dispersion forces for engi-neering applications. Adv Colloid Interface Sci. 1995;56:201-243.

44. Israelachvili JN. Intermolecular and Surface Forces. 2nd ed. NewYork: Academic Press, 1991.

45. Gregory J. Approximate expressions for retarded van der Waals inter-action. J Colloid Interface Sci. 1981;83:138-145.

46. Hogg R, Healy TW, Fuerstenau DW. Mutual coagulation of colloidaldispersions. Trans Faraday Soc. 1966;62:1638-1651.

47. Bell GM, Levine S, McCartney LN. Approximate methods of deter-mining the double-layer free energy of interaction between twocharged colloidal spheres. J Colloid Interface Sci. 1970;33:335-359.

48. Sader JE, Carnie SL, Chan DYC. Accurate analytic formulas for thedouble-layer interaction between spheres. J Colloid Interface Sci.1995;171:46-54.

49. Johnson SB, Franks GV, Scales PJ, Boger DV, Healy TW. Surfacechemistry - rheology relationships in concentrated mineral suspen-sions. Int J Miner Process. 2000;58:267-304.

50. Manciu M, Ruckenstein E. Role of the hydration force in the stabilityof colloids at high ionic strengths. Langmuir. 2001;17:7061-7070.

51. Thompson DW, Collins IR. Electrolyte-induced aggregation of goldparticles on solid surfaces. J Colloid Interface Sci. 1994;163:347-354.

52. Derjaguin BV. Friction and adhesion. IV. The theory of adhesion ofsmall particles. Kolloid-Z. 1934;69:155-164.

53. Broide ML, Cohen RJ. Experimental evidence of dynamic scaling incolloidal aggregation. Phys Rev Lett. 1990;64:2026-2029.

54. Lin MY, Lindsay HM, Weitz DA, Ball RC, Klein R, Meakin P.Universal reaction-limited colloid aggregation. Phys Rev A. 1990;41:2005-2020.

55. Schaefer DW, Martin JE, Wiltzius P, Cannell DS. Fractal geometry ofcolloidal aggregates. Phys Rev Lett. 1984;52:2371-2374.

56. Weitz DA, Huang JS, Lin MY, Sung J. Limits of the fractal dimensionfor irreversible kinetic aggregation of gold colloids. Phys Rev Lett.1985;54:1416-1419.

57. Gear CW. Numerical Initial Value Problems in Ordinary DifferentialEquations. Englewood Cliffs, NJ, USA: Prentice-Hall, 1971.

58. Press WH, Teukolsky SA, Vetterling WT, Flannery BP. NumericalRecipes in FORTRAN: the Art of Scientific Computing. 2nd ed. NewYork: Cambridge University Press, 1992.

59. Camp TR, Stein PC. Velocity gradients and internal work in fluidmotion. J Boston, Soc Civ Eng. 1943;30,219-237.

60. Jiang Q, Logan BE. Fractal dimensions of aggregates determined fromsteady-state size distributions. Env Sci Tech. 1991;25,2031-2038.

61. Selomulya C, Bushell G, Amal R, Waite TD. Understanding the roleof restructuring in flocculation: The application of a population bal-ance model. Chem Eng Sci. 2003;58:327-338.

62. Aubert C, Cannell DS. Restructuring of colloidal silica aggregates.Phys Rev Lett. 1986;56:738-741.

63. Selomulya C, Amal R, Bushell G, Waite TD. Evidence of shear ratedependence on restructuring and breakup of latex aggregates. J ColloidInterface Sci. 2001;236:67-77.

64. Lattuada M, Sandkuhler P, Wu H, Sefcik J, Morbidelli M. Aggregationkinetics of polymer colloids in reaction limited regime: experimentsand simulations. Adv Colloid Interface Sci. 2003;103:33-56.

65. Tjipangandjara K, Huang YB, Somasundaran P, Turro NJ. Correlationof alumina flocculation with adsorbed poly(acrylic acid) conformation.Colloids Surfaces A, 1990;44:229-236.

66. Rattanakawin C, Hogg R. Aggregate size distributions in flocculation.Colloids Surfaces A. 2001;177:87-98.

67. Schudel M, Behrens SH, Holthoff H, Kretzschmar R, Borkovec M.Absolute aggregate rate constants of hematite particles in aqueoussuspensions: a comparison of two different surface morphologies. JColloid Interface Sci. 1997;196:241-253.

68. Gardner KH, Theis TL, Young TC. Colloid aggregation: numericalsolution and measurements. Colloids Surfaces A. 1998;141:237-252.

69. Hunter RJ. Zeta Potential in Colloid Science: Principles and Appli-cations. New York: Academic Press, 1981.

70. Herrington TM, Midmore BR. Investigation of scaling effects in theaggregation of dilute kaolinite suspensions by quasi-elastic light scat-tering. Colloids Surfaces A. 1993;70:199-202.

71. Axford SDT, Herrington TM. Determination of aggregate structuresby combined light-scattering and rheological studies. J Chem Soc.,Faraday Trans. 1994;90:2085-2093.

72. Burns JL, Yan Yd, Jameson GJ, Biggs S. A light scattering study of thefractal aggregation behavior of a model colloidal system. Langmuir.1997;13:6413-6420.

73. Tang P, Colflesh DE, Chu B. Temperature effect on fractal structure ofsilica aggregates. J Colloid Interface Sci. 1988;126:304-313.

74. Amal R, Raper JA, Waite TD. Fractal structure of hematite aggregates.J. Colloid Interface Sci. 1990; 140,158-168.

75. Davis JA, James RO, Leckie JO. Surface ionization and complexationat the oxide/water interface. I. computation of electrical double layerproperties in simple electrolytes. J. Colloid Interface Sci. 1978;63:480-499.

76. Davis JA, Leckie JO. Surface ionization and complexation at theoxide/water interface. II. surface properties of amorphous iron oxyhy-droxide and adsorption of metal ions. J. Colloid Interface Sci. 1978;67:90-107.

77. Schmitt A, Odriozola G, Moncho-Jorda A, Callejas-Fernandez J, Mar-tinez-Garcia R, Hidalgo-Alvarez R. Multiple contact kernel for diffu-sionlike aggregation. Phys. Rev. E, 2000;62:8335-8343.

78. Family F, Meakin P, Vicsek T. Cluster size distribution in chemicallycontrolled cluster-cluster aggregation. J. Chem. Phys. 1985;83,4144-4150.

Manuscript received Jan. 4, 2004, and revision received Aug. 1, 2004.

AIChE Journal 1245April 2005 Vol. 51, No. 4