Vous êtes sur la page 1sur 29

This article was downloaded by: [Washington University in St Louis]

On: 07 October 2014, At: 08:39


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954
Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH,
UK

Chemical Engineering
Communications
Publication details, including instructions for
authors and subscription information:
http://www.tandfonline.com/loi/gcec20

TAILORING PARTICLE SIZE


THROUGH NANOPARTICLE
PRECIPITATION
a b
H.-C. SCHWARZER & W. PEUKERT
a
Institute of Particle Technology , Garching,
Germany
b
Institute of Particle Technology , Erlangen,
Germany
Published online: 10 Aug 2010.

To cite this article: H.-C. SCHWARZER & W. PEUKERT (2004) TAILORING PARTICLE SIZE
THROUGH NANOPARTICLE PRECIPITATION, Chemical Engineering Communications,
191:4, 580-606, DOI: 10.1080/00986440490270106

To link to this article: http://dx.doi.org/10.1080/00986440490270106

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the
information (the “Content”) contained in the publications on our platform.
However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness,
or suitability for any purpose of the Content. Any opinions and views
expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the
Content should not be relied upon and should be independently verified with
primary sources of information. Taylor and Francis shall not be liable for any
losses, actions, claims, proceedings, demands, costs, expenses, damages,
and other liabilities whatsoever or howsoever caused arising directly or
indirectly in connection with, in relation to or arising out of the use of the
Content.

This article may be used for research, teaching, and private study purposes.
Any substantial or systematic reproduction, redistribution, reselling, loan,
sub-licensing, systematic supply, or distribution in any form to anyone is
expressly forbidden. Terms & Conditions of access and use can be found at
http://www.tandfonline.com/page/terms-and-conditions
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014
Chem. Eng. Comm., 191: 580 606, 2004
Copyright # Taylor & Francis Inc.
ISSN: 0098-6445 print/1563-5201online
DOI: 10.1080/00986440490270106

TAILORING PARTICLE SIZE THROUGH


NANOPARTICLE PRECIPITATION
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

H.-C. SCHWARZER
Institute of ParticleTechnology,
Technische Universita«t Mu«nchen,
Garching, Germany

W. PEUKERT
Institute of ParticleTechnology,
Universita«t Erlangen-Nˇrnberg,
Erlangen, Germany

Precipitation of nanoscaled particles and the size-determining precipitation


parameters are investigated experimentally as well as numerically using bar-
ium sulfate as a reference substance. The objective of this work is to suc-
cessfully understand and predict precipitation kinetics. Optimization and
tailoring of product properties to specific needs would then be possible
without the need of extensive experimentation and its costs. Special attention
is paid to the influences of mixing as well as stabilization on the formed PSD.
To simulate particle formation the population balance equation, including the
terms for nucleation, growth, and agglomeration, is coupled with an specially
developed extended version for equi-volumetric mixing of the Engulfment-
Deformation-Diffusion-model of micromixing of Baldyga and Bourne (1999).
The proposed predictive model for nanoparticle precipitation is explained in
detail and simulation results are presented, discussed, and compared to ex-
perimental results.

Keywords: Precipitation; Nanoparticle; Population balance; Micromixing;


Particle size distribution; Barium sulfate

Address correspondence to W. Peukert, Institute of Particle Technology, Universität


Erlangen-Nürnberg, D-91058 Erlangen, Germany. E-mail: w.peukert@lfg.uni-erlangen.de

580
TAILORING PARTICLE SIZE 581

INTRODUCTION
Many properties of solid particles are not only a function of the materials
bulk properties but also depend on particle size and particle size
distribution (PSD), especially for sizes below 1 mm. Examples are optical
properties, melting point, and rate of dissolution. These property changes
arise from the increasing influence of surface properties in comparison to
volumetric bulk properties as the particle size decreases. Nanoscaled
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

particles especially show altered properties and therefore have already


been used in widespread applications, e.g., as pigments, pharmaceuticals,
cosmetics, ceramics, catalysts, and filling materials. Since the desired
product properties might vary with particle size, control of the PSD
during production is key to product quality. Furthermore, new products
can be designed by adjusting and optimizing the PSD.
Precipitation is a promising method for the economic production of
commercial amounts of nanoparticles as it is fast and operable at am-
bient temperature. However, process control — due to the rapidity of the
involved processes of mixing, nucleation, growth, and agglomeration —
and stabilization against agglomeration represent challenges. This article
shows how these challenges can be successfully handled. Precipitation
experiments with barium sulfate in a T-mixer are presented. The focus of
this work therefore is how to tailor the PSD in continuous precipitation.
In addition, the population balance is solved including the terms for
nucleation, growth, and agglomeration to study the evolution of the PSD
and the competing kinetics of these mechanisms.

PHYSICAL CONCEPTS
Overview
The PSD of a precipitation process is the result of several parallel and
successive processes: nucleation, growth, and secondary processes such as
agglomeration, attrition, and breakage. In the case of nanoscaled parti-
cles the latter two are negligible due to low collision energies. Conse-
quently, nucleation, the formation of new particles; growth, the increase
of the particles’ characteristic length; and agglomeration, the combina-
tion of two or more particles to larger units have to be considered when
investigating precipitation. A chart of the precipitation process is depicted
in Figure 1 showing the various influencing parameters on the three
mentioned processes.
Nucleation and growth rates depend strongly on supersaturation, a
measure for the system’s thermodynamic offset from equilibrium. Due to
the highly nonlinear dependence of the nucleation rate on super-
saturation, it has a strong influence on the precipitated PSD. While at
high supersaturations a large number of particles are formed that remain
582 H.-C. SCHWARZER AND W. PEUKERT
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

Figure 1. Precipitation processes and influencing parameters on particle size distribution.

small since the available total mass is limited, fewer but larger particles
are generated at lower supersaturations. Both nucleation and growth
reduce supersaturation.
Unless a slow chemical reaction is involved, which is not the case in
the investigated precipitation of an inorganic and sparingly soluble salt
such as barium sulfate, supersaturation is ‘‘generated’’ through mixing on
the molecular scale, called micromixing. Micromixing is preceded by
macromixing, a flow field controlled reduction of segregation length
scales. Thus, the kinetics of micromixing can be a function of the mac-
romixing kinetics. It is a function of the turbulence intensity and the
specific power input, which determine the size of the smallest eddies and
thereby the length necessary to surmount through diffusion for molecular
mixing. Details on mixing can be found in Baldyga and Bourne (1999). In
addition, the flow field in the reactor controls not only the process of
mixing but also the rate at which particles collide with each other and
thereby the agglomeration rate, since particle-particle collisions are
assumed a prerequisite for agglomeration.
While the flow field can be considered as a process parameter that
can be altered and optimized according to specific needs, a completely
independent set of influencing parameters is given through the com-
position of the educt solutions. For example, the interfacial tension
that influences the nucleation rate is a function of ionic strength, the
particle’s surface charge, and thus the concentrations of the potential
determining ions and of macromolecule adsorption. These parameters,
which could vary strongly through a precipitation process, also
determine the particle interaction and thereby the agglomeration rate.
It can be concluded that precipitation is a complex system of parallel
and subsequent processes with various interacting and changing
parameters.
TAILORING PARTICLE SIZE 583

In recent years various groups have investigated the precipitation


process experimentally as well as through simulation at different levels of
detail and process conditions. In many cases, the focus of investigation
was the influence of mixing on the mean particle size and the PSD in
batch or semi-batch precipitation using the classic stirred tank reactor
(Baldyga et al., 1995; Muhr et al., 1995; Phillips et al., 1999; Wei et al.,
2001) or similar reactors (Rousseaux et al., 2001; Torbacke and
Rasmuson, 2001). Some recent investigations were carried out in tubular
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

coaxial mixers (Baldyga and Orciuch, 2001; Marchisio et al., 2002) as


well. The disadvantages of such reactors are relatively low mixing
intensities, i.e., long mixing times (compared to the investigated T-mixer),
and in the case of tank reactors long residence times as well as broad
residence time distributions. As a consequence, comparatively lower
supersaturations can be realized and particles with sizes in the micro-
meter range were obtained (and desired) in all these investigations. As a
further consequence, successful simulation of the development of the
PSD under such conditions frequently requires detailed temporal and
spatial data on the influencing parameters, which can be obtained
through computational fluid dynamics (CFD) in combination with a
mixing model.
Simulation strategies of precipitation processes in recent years are
sometimes based on CFD with the consequence that, due to limitations
of computational power, particle formation is simulated using a
moment method and often with neglect of agglomeration (Baldyga and
Orciuch, 2001; Rousscaux et al., 2001; Wei et al., 2001). One of the
most advanced works following this strategy is the one by Marchisio
et al. (2002), who simulated precipitation based on CFD, including
mixing though a finite-mode probability density function and the
particle formation as well as agglomeration using the standard moment
method.
The other simulation strategy is to use average parameters to
describe the flow field or the mixing as result of the flow field. This
approach was followed by Baldyga et al. (1995); Muhr et al. (1995); and
Phillips et al. (1999) (agglomeration was neglected in these publications),
and various strategies to solve the population balance and thus calculate
the PSD can be applied, such as a finite-differences method or the
Galerkin-h-p method, which is used in this work. Both simulation stra-
tegies have advantages and there is no general rule which strategy is best.
None of the above mentioned publications, however, focus on the gen-
eration of nanoparticles.
In this work, precipitation is investigated as continuous process with
the main focus on the generation of nanoparticles in a T-mixer that
features a relatively narrow residence time distribution. Therefore a
simulation strategy for the particle size distribution based on the
584 H.-C. SCHWARZER AND W. PEUKERT

assumption of plug flow through the mixer, a mixing model using pressure
drop data, and detailed modeling of particle formation under consi-
deration of agglomeration was chosen in this work.

Supersaturation
The driving force for nucleation and growth is supersaturation, a mea-
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

sure for the system’s thermodynamic offset from its equilibrium. For the
investigated precipitation of barium sulfate from aqueous solutions of
barium chloride and sulfuric acid according to Equation (1),

BaCl2 þ H2 SO4 ! BaSO4 þ 2  HCl ð1Þ

a concentration-based supersaturation Sc can be calculated as defined


in Equation (2) with the value of the solubility product
KSP ¼ 1:08  1010 kmol2  m6 at 25 C taken from literature (Gmelins
Handbuch, 1960).
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
½Ba2þ   ½SO2 4 
Sc ¼ ð2Þ
KSP

Under the investigated experimental conditions, however, two major


additional factors have to be taken into account when calculating the real
supersaturation: due to the high ionic strengths, activities have to be used
instead of concentrations and the incomplete dissociation of sulfuric acid
according to Equation (3) has to be considered. For the dissociation of
sulfuric acid, the first stage can be considered complete and for the sec-
ond stage a value of Kdis;2 ¼ 1:2  102 kmol  m3 at 25 C can be found
in literature (Gmelins Handbuch, 1960). Using this value, the sulfate ion
concentration ½SO2 4 dis can be calculated by solving a second-order
polynomial equation.

H2 SO4 ! HSO
4 þH
þ
! SO2
4 þ2H
þ
ð3Þ

Activities can be described as a product of concentration times mean


activity coefficient. These mean activity coefficients g in the super-
saturated solution and g at saturation can be calculated as a function of
the ionic strength I using Equation (4), which is proposed by Bromley
(1973) as an advanced version of the Debye-Hückel limiting law valid for
ionic strengths up to 6 kmol  m3 . For barium sulfate the value of B in
Equation (4) is given in Bromley (1973) as B ¼ 70.037.
TAILORING PARTICLE SIZE 585

pffiffiffi
0:511  jzþ  z j  1 ð0:06 þ 0:6  BÞ  jzþ  z j  I
log g ¼ pffiffiffi þ  2 þBI ð4Þ
1þ 1 1:5
1þ I
jzþ  z j

This leads to Equation (5) for the (activity-based) supersaturation S,


which is thus not only a function of concentrations but through the
activity coefficients also a function of total ionic strength as well as a
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

function of pH-value, which shifts the dissociation equilibrium and


thereby the sulfate concentration. Generally speaking, the super-
saturation S is always smaller than the concentration-based super-
saturation calculated in Equation (2).

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
½Ba2þ   ½SO2 4 dis
S ¼ g  ð5Þ
KSP

Table 1 shows the increasing difference between concentration-based


supersaturation Sc and (activity-based) supersaturation S with increasing
concentrations calculated for barium sulfate precipitated from barium
chloride and sulfuric acid at surplus of barium ions of R ¼ 1.5. The non
dimensional parameter R is defined according to Equation (6) as the ratio
of barium to sulfate concentrations in the mixture, assuming instanta-
neous and ideal mixing. Values above unity indicate barium surplus,
values below unity sulfate surplus. As can be seen, the values of activity-
based supersaturations are almost an order of magnitude below those of
concentration-based supersaturations. Thus an enormous error would be
made if concentration-based supersaturations were used instead of
activity-based ones.

½Ba2þ 0
R¼ ð6Þ
½SO2
4 0

Table I Deviation of (Activity-Based) Supersaturation from Concentration-Based


Supersaturation Calculated for Barium Sulfate at R ¼ 1.5

Set 1 Set 2 Set 3 Set 4 Set 5 Set 6 Set 7


½Ba2þ 0 Kmol/m3 1 0.69 0.5 0.33 0.21 0.15 0.06
½SO24 0 Kmol/m3 0.67 0.46 0.33 0.22 0.14 0.1 0.04
Sc 34893 24076 17359 11515 7238 5234 2094
S 4732 3971 3375 2688 2072 1692 1152
S/Sc 0.136 0.165 0.194 0.233 0.286 0.323 0.550
586 H.-C. SCHWARZER AND W. PEUKERT

Nucleation
According to Mersmann et al. (2000) and supported by the experimental
findings of Schubert (1998), homogeneous nucleation can be considered
the dominant nucleation mechanism at the studied high supersaturations.
According to Schubert (1998), the rates for homogeneous and hetero-
geneous nucleation are equal at a supersaturation of about 150 for bar-
ium sulfate. The nucleation rate of homogenous nucleation can be
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

described according to Equation (7), which is based on the classical


theory of nucleation and can be found in Mersmann (2000). The
nucleation rate strongly depends on supersaturation, increasing in value
from 4.8 1010m 7 3s 7 1 to 8.6 1022 m 7 3s 7 1 to 5.5 1028 m 7 3s 7 1 as the
supersaturation is increased from 100 to 1000 to 10000.

rffiffiffiffiffiffiffi !
pffiffiffiffiffiffiffiffiffi gCL 16p gCL 3 V2m
Bhom ¼ 1:5DAB ð KSP SNA Þ7=3  Vm exp   
kT 3 kT ðnlnSÞ2
ð7Þ

Since the exponential term in the nucleation rate equation (7) is a third-
order function of the interfacial tension gCL , this parameter also exhibits
a strong influence on the nucleation rate. According to Mersmann (1990),
the interfacial tension can be calculated by the following equation:

 
1 rc
gCL ¼ 0:414  kT  pffiffiffiffiffiffiffiffiffi  In pffiffiffiffiffiffiffiffiffi ð8Þ
3 V 2 M  KSP
m

Not only the rate at which new nuclei are formed but also the size of the
formed nuclei are highly nonlinear functions of supersaturations as well
as interfacial tension. Assuming that the size xC of the newly formed
particles is equal to the size of critical nuclei, Equation (9), which can be
found in Mersmann (2000), can be applied.

4  gCL  Vm
xc ¼ ð9Þ
n  kT  ln S

Growth
According to Mersmann et al. (2000), growth at high supersaturations
can be considered diffusion controlled. Experimental data by Angerhöfer
(1994) as well as by Nielsen (1958) indicate a transition from integration-
controlled to diffusion-controlled growth for barium sulfate at a
TAILORING PARTICLE SIZE 587

supersaturation of about 35 to 40, i.e., at supersaturations far below the


investigated supersaturations of this study. Compared to the nucleation
rate, the growth rate is —as given in Equation (10) — much less depen-
dent on supersaturation. Equation (10) was derived for spherical particles
based on the transport equation of mass stating that the mass flux equals
mass transfer coefficient times surface area times difference of activities
between bulk solution and at the surface. The mass flux is a measure for
the growth rate, and the linear growth rate G can be calculated through
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

derivation from the mass flux. The mass transfer coefficient is expressed
through the Sherwood number Sh and the diffusion coefficient DAB.
Activities are expressed through root of the solubility product times
supersaturation. At the surface equilibrium is assumed, i.e., S ¼ 1.
Supersaturations are used since two ion species, Ba2 þ and SO2 4 , con-
tribute to growth and both contributions interact through charging of the
particles; thus mean activities have to be applied when calculating the
difference.

pffiffiffiffi
@x Sh  DAB  KSP  M S  1
G¼ ¼2  ð10Þ
@t rc x

Agglomeration and Stabilization

In order to obtain nanoparticulate precipitates agglomeration has to be


prevented. There are two reasons for particle-particle collisions of
nanoscaled particles: Brownian motion and shear forces in the turbulent
flow field in the mixer. Differences in sedimentation velocities can be
neglected for nanoscaled particles. The agglomeration rate is usually
described as a second-order differential equation of the total particle
number concentration using a mechanism-dependent proportionality
constant called the agglomeration kernal. The proper equation was
proposed by Smoluchowski (1917) and is given as Equation (11) for
mono as well as polydisperse systems.

dN b ðx1 ; x2 Þ
¼  coll  N2 ð11Þ
dt W

bcoll ðx1 ; x2 Þ in this equation is the mechanism- and size-dependent pro-


portionality constant that describes the binary collision rate of two
particles of sizes x1 and x2 without considering particle interaction. The
influences of particle interaction are accounted for by the stability factor
W, which describes the relative number of collisions with and without
588 H.-C. SCHWARZER AND W. PEUKERT

interaction. Values larger than one account for repulsive interaction and
result in smaller agglomeration rates, i.e., the particles are more stable
against agglomeration. The total particle number concentration N is
related to the particle number density n(x) by the following equation:

Z 1
N¼ nðxÞ  dx ð12Þ
0
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

In literature there are many suggestions on how to calculate bcoll for the
different collision mechanisms: for Brownian motion a cause of particle
collisions the kernel proposed by Smoluchowski (1917), and given in
Equation (13) seems reasonable, while for collisions in highly turbulent
flows the kernel proposed by Saffman and Turner (1956) as given in
Equation (14) is used.

 
2kT 1 1
bcoll;Brown ¼  ðx1 þ x2 Þ  þ ð13Þ
3:nF  rF x1 x2
rffiffiffiffiffiffiffiffiffi rffiffiffiffiffi
p 3 e
bcoll;turb ¼  ðx1 þ x2 Þ  ð14Þ
8:15 nF

A comparison of these collision kernels bcoll is displayed in Figure 2


calculated for a particle of x1 ¼ 50 nm in diameter. As can be seen,
Brownian motion (calculated using Equation (13)) is the dominant agglo-
meration mechanism in motionless to weakly turbulent (e ¼ 1 W/kg)
flows. In highly turbulent (e ¼ 106 W/kg) flows, however, agglomeration
is dominated by turbulent collisions, calculated using Equation (14).
Furthermore, it can be seen that the Brownian motion kernel has a
minimum concerning the collision partner size at equal sizes. Thus it is
least likely that two particles of equal sizes collide with each other. For
the turbulent mechanism, however, the likeliness of collision increases
with the collision partner size. As mentioned previously, sedimentation,
calculated using the equation given in Mersmann et al. (2000), as a
mechanism for collisions can be neglected.
The stability factor W can be calculated according to Equation (15)
by Fuchs (1934) as a function of the interaction potential, which can be
calculated by applying the DLVO theory. Thus W is a function of the
surface potential, created among others by ion adsorption on the surface,
the ionic strength, the particle surface curvature, the Hamaker constant,
and other parameters. Further details can be found in Israelachvili
(1991).
TAILORING PARTICLE SIZE 589
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

Figure 2. Comparison of collision kernels: Brownian motion, sedimentation, turbulent flow


at e ¼ 1 W/kg, turbulent flow at e ¼ 106 W/kg.

 
Z 1 exp jtotal ðaÞ
kT
W ¼ x da ð15Þ
0 ðx þ aÞ2

If every collision would lead to agglomeration, i.e., W = 1, agglomerates


would reach sizes of microns and larger within seconds, as calculations
using the collision kernels by Smoluchowski (1917) for Brownian motion
and Saffman and Turner (1956) for turbulent flows, respectively, have
shown. Thus in order to prevent or control agglomeration the particle-
particle interactions have to be tailored to specific needs. In the case of
preventing agglomeration, highly repulsive particle-particle interactions
have to be applied.
In the case of barium sulfate, it is known from the work of Eble
(2000) that barium ions adsorb much more readily than sulfate ions on
barium sulfate particles, leading to positive surface charges and repulsive
potentials. By varying the molar ratio R as defined in Equation (6), i.e.,
by changing the barium ion concentration, the surface charge and
thereby the electrostatic repulsion forces are altered. Thus it is possible to
control agglomeration by means of stabilization via the molar ratio R.
590 H.-C. SCHWARZER AND W. PEUKERT

Mixing
As already mentioned, supersaturation is ‘‘generated’’ through mixing, a
process that occurs on scales from the mixer scale down to the molecular
level. The relevant mixing process for supersaturation generation is
micromixing, i.e., mixing at length scales below the Kolmogorov scale. If
micromixing is the only limiting mixing step, as it is in the deployed
mixer, its kinetics can be modeled by the Engulfment-Deformation-
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

Diffusion model (EDD model) of Baldyga and Bourne (1999). Since this
model was originally derived for the situation of mixing a small volume
to a much larger volume, it had to be extended for equi-volumetric
mixing. Baldyga and Bourne propose an extension of their micromixing
model by introducing a ‘‘contact zone.’’ For the initial contact zone
volume fraction they assume a value that is proportional to the turnover
time of the eddies and thus unequal to zero. They also propose that
mixing does not start during that first turnover time, which equals
ln(2)/E, the engulfment parameter. This approach is notfeasible for the
coupling of mixing and population balance. In order to avoid this initial
condition, a different extension of the EDD model of micromixing also
based on the concept of a contact zone was derived.
The new as well as the Baldyga EDD model are based on volume
fractions X and assumes ideally mixed zones, one of solution A, one of
solution B, and the contact zone, which consists of a zone of solution A
and the mixing zone. For simplification these zones are called and
indexed with A, B, and the contact zone C, which consists of AC and the
mixing zone M. Thus XA þ XB þ XC ¼ 1 and XAC þ XM ¼ XC. The new
model assumes that XC increases when A is mixed with B, A is mixed
with C, or B is mixed with C. For an increases of M, it has to be mixed
either with B or with AC. In order to describe the rate of mixing, the
engulfment parameter E as in the original models by Baldyga and Bourne
is used. This parameter is a function of the specific power input, a
characteristic turbulence parameter that determines the Kolmogorov
scale and thereby via the Batchelor scale the necessary diffusion length.
Values for this parameter can either be taken from CFD simulations or
calculated from pressure drop measurements with averaging over the
mixer volume. Concentrations and the supersaturation in the mixing zone
can then be calculated via species balances. Using this model the fol-
lowing system of partial differential equations has to be solved:
rffiffiffiffiffi
e
E ¼ 0:058  ð16Þ
nF

dXc
¼ E  ðXC  XA þ XC  XB þ XA  XB Þ ð17Þ
dt
TAILORING PARTICLE SIZE 591

 
dXA XA
¼ E  XC  XA þ XA  XB  ð18Þ
dt XA þ XB
 
dXB XB
¼ E  XC  XB þ XA  XB  ð19Þ
dt XA þ XB
 
dXM XB
¼ E  XC  XB þ XA  XB  þ XM  XAC ð20Þ
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

dt XA þ XB
 
dXAC XA
¼ E  XC  XA þ XA  XB   XM  XAC ð21Þ
dt XA þ XB

Figure 3 shows a comparison of the calculated development of the mixing


zone volume fraction XM for the three mixing models for equi-volumetric
mixing. The equations for the simple and the extended EDD model of
Baldyga can be found in Baldyga and Bourne (1999). As can be seen, the
development of the mixing zone volume fraction with time of the two
extended models is very similar. The simple model predicts much faster

Figure 3. Comparison of micromixing models calculated for a specific power input of


e ¼ 1000 W/kg: simple EDD model, extended EDD model of Baldyga and Bourne, and ex-
tended EDD model as proposed in this work.
592 H.-C. SCHWARZER AND W. PEUKERT

mixing, but it is not valid for equi-volumetric mixing. The advantage of


the proposed extended EDD model is that no assumptions concerning
initial values have to be made.

Population Balance
Since particle formation and mixing occur on similar time scales, that are
in the order of ms to ms, both processes have to be coupled in the model
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

to accurately calculate the PSD. In order to simulate the development of


the PSD including nucleation, growth, and agglomeration, the popula-
tion balance equation as given in Equation (22), coupled via species
balances with the proposed extended EDD model, has to be solved.

@nðxÞ
¼ Bhom ðS;gCL Þ  fðxc ðS;gCL ÞÞ
@t
@ðGðS; xÞ  nðxÞÞ
 þ Bagglo ðn; xÞ  Dagglo ðn; xÞ ð22Þ
@x

In this equation n(x) is a number density concentration of particles of size


x. The first term on the right-hand side accounts for nucleation and the
second for the growth, and Bagglo as well as Dagglo represent the
agglomeration process. The function f(xC) describes the number density
distribution of the formed nuclei. According to Equation (9), this func-
tion would be a Dirac delta function. However, it is likely that there are
fluctuations of nuclei sizes even under constant conditions, and thus a
Gaussian distribution for random fluctuations is appropriate. Therefore
it was assumed that f(xC) is a narrow Gaussian distribution, and a
standard deviation of 5% was chosen to characterize the shape of the
distribution.
The concept of modeling agglomeration is that a particle of size x3 is
formed when particles of sizes x1 and x2 agglomerate (x1 and x2 vanish).
Via mass balance x3 is a function of x1 and x2 . Since there are many
possible combinations of x1 and x2 to form a particle of size x3 , it is
necessary to integrate over all these combinations. This lead to Equations
(23) and (24) to calculate the birthrate Bagglo and death rate Dagglo of
agglomeration. Thus the population balance equations turns out to be a
highly nonlinear integro-differential equation, which is additionally
coupled to the system of differential equations describing the mixing
process.

Z pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
3
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
3
x2 x
bð x3  l3 ; lÞ  nð x3  l3 Þ  nðlÞ
Baggl ðn; xÞ ¼  2=3 dl ð23Þ
2 xc x3  l3
TAILORING PARTICLE SIZE 593

Z 1
Daggl ðn; xÞ ¼ nðxÞ  bðx; lÞ  nðlÞdl ð24Þ
0

The precipitation process was simulated by solving the population bal-


ance equation numerically for the mixing zone XM using a Galerkin-h-p
method implemented in the software package PARSIVAL1 by CiT
GmbH. The system of differential equations describing the mixing pro-
cess was solved within this software and used to adjust the mixing zone
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

volume fraction and its composition by controlling feed streams into the
mixing zone.
In conclusion, it is expected that an increase in supersaturation by
increasing the educt concentration as well as more intense mixing will
lead to smaller particles. Furthermore, it is expected that agglomeration
can be reduced, controlled, and possibly even completely prevented by
increasing the molar ratio R. i.e., increasing the barium concentration in
the suspension. Hence, primary particle size and agglomeration can be
adjusted independently within certain ranges to meet PSD product
requirements.

EXPERIMENTAL RESULTS
Precipitation experiments, some of which have previously been published
by the authors (Schwarzer and Peukert, 2002), were carried out as con-
tinuous experiments in the apparatus shown in Figure 4, which was

Figure 4. Experimental setup for the precipitation experiments.


594 H.-C. SCHWARZER AND W. PEUKERT

designed and built in-house. Two pistons are moved in cylinders of high-
precision Duran glass by Schott Geräte GmbH using a stepping motor
and gear reduction in order to generate two constant volume flows for the
educt solutions. As educts, aqueous solutions of barium chloride Merck
101719, analytical grade, dissolved in deionized water), and sulfuric acid
(Merck 109912, Combi-Titrisol, diluted with deionized water) were used
and precipitated in the T-mixer, shown in Figure 5, which includes a
schematic drawing of the tubes in the interior.
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

The T-mixer was designed and built in-house from stainless steel
(1.4571) with a circular mixing zone of 1 mm in diameter and 10 mm in
length. The feed tubes are 0.5 mm in a diameter and are positioned
opposite each other. In order to characterize the flow through the mixer,
a mixer Reynolds number is defined according to Equation (25) based on
the mixing zone diameter and the volumetric flow rate through the
mixing zone.

UMZ  dMZ V_
Re ¼ ¼ ð25Þ
nF p=4  dMZ nF

Figure 5. Picture of the used mixer showing the tubes in the interior.
TAILORING PARTICLE SIZE 595

In the presented experiments, the mixer was deployed at flow rates ran-
ging from 0.1 mL/s to 12 mL/s, which correspond to Reynolds numbers
ranging from Re = 127 to Re = 15280. The flow through the mixer can
be considered turbulent, dominated by the impinging of the educts. The
upper limit of the applied flow rates is due to the pressure drop in
the mixing zone, which is about 14 bar at Re = 15280. At this flow rate,
the operational capacity of the mixers is larger than one cubic meter
of educt solution per day and more than 100 kg of nanoscaled barium
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

sulfate particles.
During the experiments, the generated suspensions were collected in
empty tanks, the initial as well as the final collections were discarded, and
the remaining main part, which corresponds to steady-state conditions,
taken for analysis. The PSD was measured off-line using a commercial
device (UPA 150 by Microtrac) based on quasi-elastic light scattering.
During the experiments, the pressure drop was recorded and the mean
specific power input in the mixing zone calculated. In addition, trans-
mission electron microscopy (TEM) and dry samples were prepared.
As expected due to the use of dilute educts, no temperature change
was detected as result of the mixing. X-ray diffraction patterns (Cu Ka
radiation, PW 1732 by Philips Analytical) confirmed the b-BaSO4 mod-
ification of the formed particles and thereby their crystallinity. Rheolo-
gical measurements of the suspension after precipitation CVO 120 by
Bohlin Instruments Ltd showed only a minor and thus negligible increase
in viscosity compared to the educts.
Measured PSD from five precipitation experiments at different
Reynolds numbers, i.e., flow rates through the mixer, are presented in
Figure 6 as volume density distributions. The theoretic initial super-
saturation S ¼ 3375 and the molar ratio R ¼ 1.5 were kept constant in
these experiments. As can be seen, the measured mean particle sizes as
well as the sizes of the smallest particles decrease as the Re number and
thereby the pressure drop as well as the mixing intensity increase. Particle
sizes were confirmed by TEM pictures as well as through nitrogen
adsorption via BET surface area and pore size distribution (Nova 2000
by Quantachrome Corp.).
The reason for the observed decrease with increasing Re number is
the different dependencies of nucleation and growth rate on super-
saturation and the ‘‘generation’’ of supersaturation through mixing. With
increasing supersaturation, the nucleation rate increases stronger than the
growth rate and thus more and smaller (size of nucleus) particles are
formed, which grow until supersaturation is completely reduced. At
larger Re numbers, mixing is faster and thus higher levels of
supersaturation are generated.
Figure 7 shows TEM pictures of barium sulfate nanoparticles pre-
cipitated at S ¼ 4405 and R ¼ 3 and mixing conditions corresponding to
596 H.-C. SCHWARZER AND W. PEUKERT
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

Figure 6. Measured density distributions of nanoscaled barium sulfate precipitated at


S ¼ 3375, R ¼ 1.5, and varying Reynolds number, i.e., mixing conditions.

Re ¼ 382 (left) and Re ¼ 6360 (right), respectively. The number mean


particle sizes of the corresponding experiments measured using the quasi-
elastic light scattering apparatus are about 77 nm and 30 nm, respectively,
which agrees well with the sizes of the primary crystals in the TEM

Figure 7. TEM pictures of precipitated barium sulfate nanoparticles at S= 4405, R ¼ 3, as


well as Re ¼ 382 (left) and Re ¼ 6360 (right).
TAILORING PARTICLE SIZE 597

pictures. TEM diffraction patterns showed that the primary particles are
monocrystalline. It is difficult to reach a conclusion on agglomeration
from these pictures since cluster formation during TEM sample pre-
paration (i.e., drying) is very likely. In any case, agglomeration at R ¼ 3
could not have been strong, or single primary particles would not have
occurred.
Reproducibility checks for S ¼ 2275, R ¼ 1.5, and Re ¼ 1270 showed
that the standard deviation of the volume-weighted mean particle size
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

was less than five nanometers in 10 precipitation experiments. The


volume-weighted mean particle size of these measurements was 72.6
nanometers. Thus, taking a possible shift of two nanometers due to the
time lag between precipitation and measurement into consideration, the
deviation between measured and real value is estimated to be below 10%.
In order to investigate the influence of mixing relative to the influence
of supersaturation, several combinations of educt concentrations were
precipitated at varying Re number (volumetric flow rate), i.e., at varying
mixing conditions. The results are shown in Figure 8, depicting the
volume-weighed mean particle size over the Re number.
As can be seen, the mean particle size can be shifted enormously by
altering the mixing conditions. For a supersaturation of S ¼ 3375, for
example, the volume-weighted mean particle size can be shifted from

Figure 8. Influence of mixing and supersaturation on particle size.


598 H.-C. SCHWARZER AND W. PEUKERT

about 50 nm to more than 200 nm, and for a supersaturation of S ¼ 1692,


the mean particle size ranges from 135 nm up to about 400 nm. Thus
particle size distributions with mean sizes ranging from about 50 nm up
to 400 nm can be tailored through mixing conditions and educt con-
centrations.
To determine the influence of agglomeration and stabilization, dif-
ferent levels of barium excess were used. Figure 9 shows cumulative
volume distributions precipitated at S  3370 and a flow rate corre-
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

sponding to Re ¼ 1270.
With decreasing barium excess, i.e., decreasing value of R and thus
decreasing stability factor W, larger PSDs are measured, confirming that
agglomeration can play a major role in precipitation and confirming the
necessity to stabilize against agglomeration in nanoparticle precipitation.
From PSD-time measurements it was found that for R ¼ 1.5 the change
in mean particle size was below one nanometer per minute in a motionless
suspension (Brownian mechanism). Thus stabilization at R  1.5 can be
considered complete for the purposes of this investigation. The arrows in
the chart for the plot of R ¼ 1 indicate that the real PSD of that
experiment is out of the used device’s measurement range and that
therefore the real plot is shifted to some extent in the direction of the
arrows. Figure 9 clearly shows that slight changes of the suspension
composition have a strong effect on the resulting PSD.

Figure 9. Stabilization against agglomeration through barium excess.


TAILORING PARTICLE SIZE 599

SIMULATION RESULTS
The PSD of precipitation is calculated by solving the population balance
equation coupled with the extended EDD model of micromixing
numerically using the Galerkin-h-p method implemented in the com-
mercial software PARSIVAL1 by CiT GmbH. The flow through the
mixer is assumed to be plug flow, allowing a 1-D spatial resolution in the
simulations. Furthermore it is assumed that macromixing is much faster
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

than micromixing, thus having no effect on the mixing kinetics. The


specific power input —an important parameter for the EDD model of
micromixing— was calculated from pressure drop measurements and
averaged over the mixer volume. The assumption of average conditions
in the mixer does not represent the real conditions in the mixer where all
variables such as the composition and the specific power input are spa-
ciotemporally distributed. But, as will be shown, the chosen assumptions
are a good approach for the predictive modeling of nanoparticle pre-
cipitation.
Figure 10 compares the kinetics of particle formation due to
nucleation and growth under the assumption of ideal mixing and negli-
gible agglomeration to supersaturation generation through micromixing

Figure 10. Comparison of calculated kinetics of supersaturation generation through micro-


mixing and supersaturation reduction through nucleation and growth.
600 H.-C. SCHWARZER AND W. PEUKERT

for three different levels of specific power input. Two conclusions can be
drawn from this figure: mixing time scales vary strongly with changing
power input and mixing is — for the investigated supersaturations— not
as fast as particle formation. Thus mixing has a strong influence on the
kinetics of precipitation processes and thereby on the PSD, which is
consistent with experimental findings.
Additionally, it is interesting to note that the complete reduction of
supersaturation is the faster the higher the initial value. This effect can be
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

explained as follows: at high supersaturations large numbers of particles


and thereby a large particle surface area is formed, which acts as a
supersaturation sink through particle growth. More than 99.9%
of supersaturation is reduced through growth under the investigated
conditions.
By solving the population balance equation including nucleation,
growth, and agglomeration with the varying stability factor W, the
influence of agglomeration is found, and the obtained results for S ¼ 3375
and R ¼ 1.5 are displayed in Figure 11 as plots over the specific power
input of the volume-weighted mean particle size.

Figure 11. Calculated influence of agglomeration in the mixing zone as a function of the spe-
cific power input e on the volume-weighted mean particle size for different agglomeration
mechanisms and different stability factors W.
TAILORING PARTICLE SIZE 601

For these calculations it was assumed that agglomeration occurs only


in the mixer, i.e., for the duration of the residence time. Thus the depicted
influence is not only a function of the agglomeration rate, which is
proportional to e1=2 , but also of the residence time in the mixer, which
are connected via the measured pressure drop. The residence time is
inversely proportional to e1=3 . These two dependencies lead to the
maximum in mean particle size, which can be found at an specific power
input of about 105 W/kg. Increasing the stability against agglomeration,
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

i.e., increasing the value of W from 1 to 10, leads to a drastic reduction of


agglomeration. Increasing W even further up to 100 leads to complete
stabilization against agglomeration on the investigated time scale of
residence time, which is in the order of milliseconds. Values of W of 100
are not uncommon. Values for the stability factor of 105 and much higher
can frequently be found.
Comparing the number- and the volume-weighted mean particle sizes
for turbulent agglomeration without interaction influence, it is interesting
to note that the maximum is much more developed in the volume-
weighed case. This indicates — as can also be seen in Figure 12 from the
calculated PSDs— that the shape of the PSD is changed extremely
through agglomeration. Only few very large particles dominate the

Figure 12. Calculated evolution of particle size distribution including nucleation, growth,
and turbulent agglomeration for S ¼ 3375, R ¼ 1.5, e ¼ 104 W/kg, and W ¼ 1.
602 H.-C. SCHWARZER AND W. PEUKERT

volume-weighted distribution. This finding is expected due to the func-


tion of the turbulent agglomeration rate, given in Equation (14), which
supports the agglomeration of large particles. Additionally, as can be
seen from these figures, mean particle sizes can easily reach the mm range
during the residence time, i.e., in the order of milliseconds, and surpass
mixing effects. Figure 11 also shows the influence of agglomeration due
to the Brownian mechanism, which is independent of the specific power
input. Since only the agglomeration time is a function of the specific
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

power input, the deviation of this plot from the plot for ‘‘no agglo-
meration’’ decreases as the specific power input increases, i.e., at decreas-
ing specific power inputs.
Figure 12 shows an example of the evolution of the PSD calculated
for a specific power input of e ¼ 104 W/kg, S ¼ 3375, R ¼ 1.5, and W ¼ 1,
including nucleation, growth, and turbulent agglomeration. As can be
seen, the shape as well as the position of the PSD changes in less than one
millisecond from the size of the nuclei, which is, according to Equation
(9), in the range of one nanometer up to several hundred nanometers.
After 10 microseconds sizes between 1 to 10 nanometers are reached, and
after 100 microseconds the particles sizes are in the range of 10 to 30
nanometers. Sizes in the range of 60 to 300 nanometers are reached after
one millisecond. By that time, supersaturation is completely reduced and
only agglomeration contributes to changes in the particle size distribu-
tion. The width of the PSD increases because of two reasons. First,
particles are formed at different points in time, and thus the PSD con-
tains young and small particles as well as older and larger particles. This
effect increases with time until supersaturation is reduced completely.
The second reason is that turbulent agglomeration enhances the
agglomeration of large particles relative to the agglomeration of smaller
ones as can be seen in Equation (14).
Figure 13 shows a comparison of simulation and experimental results
for S ¼ 3375, R ¼ 1.5, and W > 100. Both curves agree fairly well con-
sidering the various assumptions that have been made and the uncer-
tainties within the used barium sulfate specific data. It is especially
noticeable that the applied concept of using the extended EDD model of
micromixing in combination with pressure drop measurements to
describe micromixing seems appropriate. The mixing effect is qualita-
tively predicted correctly. From the combination of simulation and
experiment it can be concluded that for specific power inputs larger than
about 104 W/kg the mixing process does not have an significant influence
on the resulting particle size. However, compared with Figure 10 it is
clear that the mixing process at e ¼ 104 W=kg is about as fast as super-
saturation reduction. Only if the mixing is much faster than
supersaturation reduction, i.e., at about e > 106 W=kg for the investi-
gated case, can mixing be considered ideal.
TAILORING PARTICLE SIZE 603
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

Figure 13. Comparison of experimental and simulation results for S ¼ 3375 and R ¼ 1.5
plotted as a function of the specific power input e.

CONCLUSIONS
Nanoparticle formation was investigated experimentally as well as
numerically using continuous barium sulfate precipitation in a T-mixter.
First, it was shown that the generation of barium sulfate nanoparticles by
precipitation is possible. Two important aspects were found that strongly
determine the resulting PSD: mixing and agglomeration. The faster the
mixing on the molecular scale (i.e., micromixing), the smaller the formed
primary particles. Agglomeration has to be suppressed by stabilization,
or particles sizes well in the mm range would occur on time scales in the
order of milliseconds, depending on turbulence intensity. The precipita-
tion process was simulated by coupling an extended version of the EDD
model of micromixing one-dimensionally with the population balance
equation. Both simulation results and experimental findings agree fairly
well. It is important to state that no model parameters were adjusted to fit
experimental data.
Simulations of the precipitation process led to new insights into the
kinetics of the process. It was found that the time scales of precipitation
depend strongly on the effective supersaturation, i.e., on educt con-
centrations. Usual time scales for the reduction of supersaturation under
604 H.-C. SCHWARZER AND W. PEUKERT

the assumption of ideal mixing are in the order of a few microseconds to


more than 100 milliseconds.
The remaining offset between experimental and simulation results
can be attributed to imprecise material data and model assumptions that
have to be improved. The l-D coupling using data averaged over the
mixing zone volume for the specific power input, seems especially to be a
good starting point for further improvement since the power input is far
from uniformly distributed throughout the mixer, and local PSDs
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

therefore depend strongly on local conditions. As a consequence a 3-D


spatial resolved coupling of the population balance with CFD data is
currently under investigation.

NOMENCLATURE
a distance between particle surfaces, m
Bagglo birthrate of agglomeration, m4  s1
Bhom nucleation rate, m3  s1
dMZ diameter of mixing zone, m
DAB diffusion coefficient, m2  s1
Dagglo death rate of agglomeration, m4  s1
E engulfment parameter, s1
fðxC Þ function describing the density distribution of nuclei
G linear growth rate, m  s1
I ionic strength, kmol  m3
k Boltzmann constant (1:381  1023 ), J  K1
Kdis:2 dissociation constant, kmol  m3
KSP solubility product, kmol2  m6
M molecular weight, kg  kmol1
n(x) particle number density of size x, m4
N total particle number concentration, m3
NA Avogadro’s number ð6:023  1026 Þ, kmol1
R molar ratio
Re Reynolds number
S supersaturation
Sc concentration-based supersaturation
Sh Sherwood number
T temperature, K
t time, s
uMZ flow velocity through mixing zone, m  s1
Vm molecular volume, m3
V_ volumetric flow rate, m3  s1
W stability factor
x particle size, m
xC size of critical nucleus, m
X volume fraction
zþ ; z: ionic charge of cations or anions
[] concentration, kmol  m3
Greek letters
bcoll agglomeration kernel, m3  s1
gCL interfacial tension, J  m2
TAILORING PARTICLE SIZE 605

g activity coefficient
e specific power input, W  kg1
jtotal ðxÞ total particle-particle interaction potential, J
rC density of particles, kg  m3
rF density of fluid, kg  m3
n dissociation number
nF kinematic viscosity of fluid, m2  s1
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

REFERENCES
Angerhöfer, M. (1994). Untersuchung zur Kinetic der Fällungskristallisation von
Bariumsulfat, Ph.D. diss., Technische Universität München.
Baldyga, J. and Bourne, J. R. (1999). Turbulent Mixing and Chemical Reactions,
John Wiley, Chichester.
Baldyga, J. and Orciuch, W. (2001). Barium sulfate precipitation in a pipe — An
experimental study and CFD modelling, Chem. Eng. Sci., 56, 24352444.
Baldyga, J., Podgorska, W., and Pohorecki, R. (1995). Mixing-precipitation
model with application to double feed semibatch precipitation, Chem. Eng.
Sci., 50, 8, 12811300.
Bromley, L. A. (1973). Thermodynamic properties of strong electrolytes in aqu-
eous solutions, AIChE J., 19(2), 313320.
Eble, A. (2000). Precipitation of Nanoscale Crystals with Particular Reference to
Interfacial Energy, Ph.D. diss., Technische Universität München.
Fuchs, N. (1934). Üher die Stabilität and Aufladung der Aerosole, Z. Physik, 89,
736743.
Gmelins Handbuch der anorganischen Chemie. (1960). Verlag Chemie, Weinheim.
Israelachvili, J. (1991). Intermolecular and Surface Forces, 2nd ed., Academic
Press, London.
Marchisio. D. L., Barresi, A. A., and Garbero, M. (2002). Nucleation, growth
and agglomeration in barium sulfate turbulent precipitation, AIChE J., 48,
20392050.
Mersmann, A. (2000). Crystallization Technology Handbook, 2nd ed., Marcel
Dekker, New York.
Mersmann, A., Bartosch, K., Braun, B., Eble, A., and Heyer, C. (2000). Chem.
Ing. Tech., 72(1/2), 1730.
Mersmann, A. (1990). Calculation of interfacial tensions, J. Cryst. Growth, 102,
841847.
Muhr, H., David, R., and Villermaux, J. (1995). Crystallization and precipitation
engineering V: Simulation of the precipitation of silver bromide octahedral
crystals in double-jet semi-batch reactor, Chem. Eng. Sci., 50(2), 345355.
Nielsen, A. E. (1958). The kinetics of crystal growth in barium sulfate pre-
cipitation, Acta Chem. Stand., 12(5), 951958.
Phillips, R., Rohani, S., and Baldyga, J. (1999). Micromixing in a single-feed
semi-batch precipitation process, AIChE J., 45(1), 8292.
606 H.-C. SCHWARZER AND W. PEUKERT

Rousseaux, J.-M., Vial, C., Muhr, H. and Plasari, E. (2001). CFD simulation of
precipitation in a sliding-surface mixing device, Chem. Eng. Sci., 56,
16771685.
Saffman, P. G. and Turner, J. S. (1956). On the collision of drops in turbulent
clouds, J. Fluid Mech., 1, 1630.
Schubert, H. (1998). Keimbildung bei der Kristallisation schwerlöslicher Feststoffe,
Ph.D. diss., Technische Universität München.
Schwarzer, H.-C. and Peukert, W. (2002). Experimental investigation into the
Downloaded by [Washington University in St Louis] at 08:39 07 October 2014

influence of mixing on nanoparticle precipitation, Chem. Eng. Technol.,


25(6), 657661.
Smoluchowski, M. von (1917). Versuch einer mathematischen Theorie der Koa-
gulationskinetik kolloidaler Lösungen, Z. Phys. Chem., 92, 129168.
Torbacke, M. and Rasmuson, A. C. (2001). Influence of different scales of mixing
in reaction crystallization, Chem. Eng. Sci., 56, 24592473.
Wei, H, Zhou, W. and Garside, J. (2001). Computational fluid dynamics mod-
eling of the precipitation process in a semibatch crystallizer, Ind. Eng. Chem.
Res., 40, 52555261.

Vous aimerez peut-être aussi