Vous êtes sur la page 1sur 8

1516

Langmuir 2005, 21, 1516-1523

Smart Control of Monodisperse Sto1 ber Silica Particles:


Effect of Reactant Addition Rate on Growth Process
K. Nozawa,,, H. Gailhanou,, L. Raison, P. Panizza, H. Ushiki, E. Sellier,|
J. P. Delville,,# and M. H. Delville*,
Institut de Chimie de la Matie` re Condense e de Bordeaux, UPR 9048-CNRS,
Universite Bordeaux I, 87 Avenue du Dr. A. Schweitzer, 33608 Pessac Cedex, France,
Centre de Physique Mole culaire Optique et Hertzienne, UMR CNRS 5798,
Universite Bordeaux I, 351 Cours de la Libe ration, 33405 Talence Cedex, France,
Laboratory of Molecular Dynamics and Complex Chemical Physics, Department of
Environmental and Natural Resource Science, 3-5-8, Saiwai-cho, Tokyo 183-8509, Japan, and
Centre de Ressources en Microscopie Electronique et Microanalyse, Universite Bordeaux I,
351 Cours de la Libe ration, 33405 Talence Cedex, France
Received June 10, 2004. In Final Form: October 29, 2004
Control over the synthesis of monodisperse silica particles up to mesoscopic scale is generally made
difficult due to intrinsic limitation to submicrometric dimensions and secondary nucleation in seeded
experiments. To investigate this issue and overcome these difficulties, we have implemented single step
processing by quantifying the effects of the progressive addition of a diluted tetraethyl orthosilicate solution
in ethanol on the size and monodispersity of silica particles. Contrary to particles grown in seeded
polymerization, monodisperse particles with size up to 2 m were synthesized. Moreover, the particles
exhibit a final diameter (df), which varies with V-1/3 over more than 2 orders of magnitude in rate of
addition (V). On the basis of a kinetic study in the presence of addition showing that particle growth is
limited by the diffusion of monomer species, we developed a diffusion-limited growth model to theoretically
explain the observed df(V) behavior and quantitatively retrieve the measured amplitude and exponent.
Using a single parameter procedure, we can therefore predict and generate in the room temperature range,
monodisperse particles of a targeted size by simply adjusting the rate of addition.

Introduction
Monodisperse colloidal silica particles with uniform size,
shape, and composition have wide application not only in
the field of physical chemistry dealing with dynamic
behavior and stability of particle systems1 but also in
industries including pigments, pharmacy,2 photographic
emulsions,3 ceramics,4 chromatography,5 catalysts,6 and
chemical mechanical polishing.7 Silica particles are also
used as stabilizers,8 coatings,9 glazes10 and binders.11 The
need for well-defined silica nanoparticles is thus constantly
* To whom correspondence may be addressed. Phone: (33)
5 40 00 84 60. Fax: (33) 5 40 00 27 61. E-mail: delville@
icmcb-bordeaux.cnrs.fr.
Institut de Chimie de la Matie
` re Condensee de Bordeaux, UPR
9048-CNRS, Universite Bordeaux I.
Centre de Physique Mole
culaire Optique et Hertzienne, UMR
CNRS 5798, Universite Bordeaux I.
Laboratory of Molecular Dynamics and Complex Chemical
Physics, Department of Environmental and Natural Resource
Science.
| Centre de Ressources en Microscopie Electronique et Microanalyse, Universite Bordeaux I.
# E-mail: jp.delville@cpmoh.u-bordeaux1.fr.
(1) Wiese, G. R.; Healy, T. W. Trans. Faraday Soc. 1970, 66, 490.
(2) Paci, A.; Mercier, L.; Bourget, P. J. Pharm. Biomed. Anal. 2003,
30, 1603.
(3) Overbeek, J. Th. G. Adv. Colloid Interface Sci. 1982, 15, 251.
(4) Sacks, M. D.; Tseng, T. Y. J. Am. Ceram. Soc. 1984, 67, 526.
(5) Unger, K. K.; Kumar, D.; Grun, M.; Buchel, G.; Ludtke, S.; Adam,
Th.; Schumacher, K.; Renker, S. J. Chromatogr., A 2000, 89, 47.
(6) Badly, R. D.; Ford, W. T. J. Org. Chem. 1989, 54, 5347.
(7) Iler, R. K. The Chemistry of Silica; Wiley: New York, 1979.
(8) Gosa, K. L.; Uricanu, V. Colloids Surf., A 2002, 197, 257.
(9) Leder, G.; Ladwig, T.; Valter, V.; Frahn, S.; Meyer, J. Prog. Org.
Coatings 2002, 45, 139.
(10) Adl, S.; Rahman, I. A. Ceram. Int. 2001, 27, 681.
(11) Payne, C. In The Colloid Chemistry of Silica; Bergna, H., Ed.;
American Chemical Society: Washington, DC, 1994.

increasing, as high-tech industries (e.g., biotechnology/


pharmaceuticals12 or photonics13) provide a tremendous
demand for such materials.
The preparation of monodisperse silica particles generally proceeds with the hydrolysis and condensation of
alkoxysilanes (often tetraethyl orthosilicate (TEOS)
[Si(OR)4 with R ) C2H5]) in a mixture of alcohol, water,
and ammonia used as a catalyst. Since its discovery by
Kolbe,14 many studies have been performed based on this
reaction system.15-17 The so-called Stober synthesis, (the
ammonia-catalyzed reaction of TEOS with water in low
molecular weight alcohols), is known to produce monodisperse spherical silica nanoparticles. As the maximum
particle size achievable with high monodispersity from
TEOS seems to be submicrometric in the Stober synthesis,
18 a great deal of efforts has therefore been devoted to
extend this limit up to 2 m by using different silicon
alkoxides and solvent, with a high cost for many of them
such as tetrapentyl orthosilicate. Some authors succeeded
in the preparation of monodisperse silica particles with
sizes above 1 m in acidic emulsion media.19
(12) Caruso, F.; Caruso, R. A.; Molwald, H. Science 1998, 282, 1111.
(13) (a) Xia, Y.; Gates, B.; Ying, Y.; Lu, Y. Adv. Mater. 2000, 12, 693.
(b) Sun, H. B.; Song, J.; Matsuo, Y. X. S.; Misawa, H.; Liu G. D. S. J.
Opt. Soc. Am. B 2000, 17, 476.
(14) Kolbe, G. The Complex Chemical Behavior of Silica. Dissertation,
Jena, Germany, 1956.
(15) Stober, W.; Fink, A.; Bohn, E. J. Colloid Interface Sci. 1968, 26,
62.
(16) Van Helden, A. K.; Jansen, J. W.; Vrij, A. J. Colloid Interface
Sci. 1981, 81, 354.
(17) Van Blaaderen, A.; Geest, V. J.; Vrij, A. J. J. Colloid Interface
Sci. 1992, 154, 481.
(18) Bogush, G. H.; Tracy, M. A.; Zukoski, C. F. J. Non-Cryst. Solids
1988, 104, 95.
(19) Esquena, J.; Pons, R.; Azemar, N.; Caelles, J.; Solans, C. Colloids
Surf., A 1997, 123-124, 575.

10.1021/la048569r CCC: $30.25 2005 American Chemical Society


Published on Web 12/18/2004

Synthesis of Monodisperse Silica Particles

In general, the hydrolysis reaction gives the singly


hydrolyzed TEOS monomer [(OR)3Si(OH)]

Si(OR)4 + H2O f (OR)3Si(OH) + ROH


Subsequently, this intermediate reaction product condenses to eventually form silica according to

(OR)3Si(OH) + H2O f SiO2V + 3ROH


This reaction scheme is, of course, a simplification of
the very complex condensation processes that lead to the
formation of the silica particles. Some of the earliest
research on Stober particles was primarily concerned with
empirically predicting the final particle size for a range
of the initial reactant concentrations (0.1-0.5 M TEOS,
0.5-17.0 M H2O, and 0.1-3.0 M NH3) that lead to
monodisperse colloids, the largest achievable size being
800 nm.15,18 Two models, monomer addition20,21 and
controlled aggregation,22 have been proposed to elucidate
the chemical and/or physical growth mechanisms of silica.
To explain the highly monodisperse nature of these
particles, the first authors divided the formation of silica
into two events: nucleation and growth. Bogush18,22 and
co-workers considered the nucleation and growth of silica
particles as an aggregation process of small subparticles
several nanometers in size. In contrast, Matsoukas and
Gulari20,21 proposed that particle nucleation was the result
of the reaction between two hydrolyzed monomers, such
that the particles grew only by a molecular addition
mechanism.23 More recently, many other investigations
have been devoted to the understanding of the particle
growth mechanism24-26 including microgravity experiments.27
Important parameters in preparing silica particles seem
to be the water and ammonia concentrations. However,
depending on the authors, these parameters have an
opposite influence on the SiO2 particle size explaining
why several groups attempted to rationalize this highly
complicated system using experimental design and multivariate analyses.28
As large monodisperse silica particles (typically above
1 m) are difficult to synthesize according to the Stober
procedure with TEOS,15 some authors preferred to use
(20) Matsoukas, T.; Gulari, E. J. Colloid Interface Sci. 1988, 124,
252.
(21) (a) Matsoukas, T.; Gulari, E. J. Colloid Interface Sci. 1989, 132,
13. (b) Matsoukas, T.; Gulari, E. J. Colloid Interface Sci. 1991, 145, 557.
(22) (a) Bogush, G. H.; Zukoski, C. F. J. Colloid Interface Sci. 1991,
142, 1. (b) Bogush, G. H.; Zukoski, C. F. J. Colloid Interface Sci. 1991,
142, 19. (c) Okudera, H.; Hozumi, A. Thin Solid Films 2003, 434, 62.
(23) (a) Kim, K. S.; Kim, J. K.; Kim, W. S. J. Mater. Res. 2001, 16,
545. (b) Chen, S. L.; Dong, P.; Yang, G. H.; Yang, J. J. Ind. Eng. Chem.
Res. 1996, 35, 4487.
(24) (a) Van Blaaderen, A.; Kentgens, A. P. M. J. Non-Cryst. Solids
1992, 149, 161. (b) Burneau, A.; Humbert, B. Colloids Surf., A 1993,
75, 111. (c) Chen, S. L.; Dong, P.; Yang, G. H.; Yang, J. J. J. Colloid
Interface Sci. 1996, 180, 237.
(25) (a) Walcarius, A.; Despas, C.; Bessie`re, J. Microporous and
Mesoporous Mater. 1998, 23, 309. (b) Brinker, C. J.; Scherer, G. W
Sol-Gel Science; Academic Press: San Diego, 1990.
(26) (a) Green, D. L.; Lin, J. S.; Lam, Y. F.; Hu, M. Z. C.; Schaefer,
D. W.; Harris, M. T. J. Colloid Interface Sci. 2003, 266, 346. (b)
Vogelsberger, W.; Seidel, A.; Breyer, T. Langmuir 2002, 18, 3027. (c)
Pontoni, D.; Narayanan, T.; Rennie, A. R. Langmuir 2002, 18, 56. (d)
Boukari, H.; Lin, J. S.; Harris, M. T. J. Colloid Interface Sci. 1997, 194,
311. (e) Boukari, H.; Long, G. G.; Harris, M. T. J. Colloid Interface Sci.
2000, 229, 129. (f) Boukari, H.; Lin, J. S.; Harris, M. T. Chem. Mater.
1997, 9, 2376.
(27) Smith, D. D.; Sibille, L.; Cronise, R. J.; Hunt, A. J.; Oldenburg,
S. J.; Wolfe, D.; Halas, N. J. Langmuir 2000, 16, 10055.
(28) (a) Lindberg, R.; Sundholm, G.; Pettersen, B.; Sjoblom, J.; Friberg,
S. E. Collods Surf., A 1997, 123-124, 549. (b) Giesche, H. J. Eur.
Ceram. Soc. 1994, 14, 189. (c) Dingsoyr, E.; Christy, A. A. Prog. Colloid
Polym. Sci. 2000, 116, 67.

Langmuir, Vol. 21, No. 4, 2005 1517

the so-called seed polymerization reaction, which is based


on the hydrolysis and condensation of this alkoxide onto
preformed silica seeds. They claimed that growth occurs
without the formation of a new generation of silica
particles.29 However, as illustrated in the following, our
experiments (and others30) are clearly at variance with
this conclusion beyond a certain particle size.
To prevent secondary nucleation effects, the method
was extended by coupling seed polymerization to controlled
growth reaction in a continuous production process.31,32
Reports on such preparations of silica particles33 qualitatively established the increasing impact of the rate of
addition of the TEOS reactant in the reactor vessel on
final size and polydispersity of the particles. Consequently,
as opposed to the classical Stober synthesis (also qualified
as a batch process34), the use of a so-called semibatch
process,33,35 in which one reactant (TEOS/EtOH) is added
into a reactor containing the other ones (H2O/NH4OH/
EtOH) at a constant rate, was claimed to give greater
control over the resulting particle size, shape, and size
distribution. This was illustrated for the synthesis of silica
nanoparticles in multivariate analyses at room temperature36 by comparing two runs performed at a rate of
addition of the TEOS reactant of 18 mL/min (i.e., almost
in batch conditions) and at 0.6 mL/min. Experiments
showed that the final particle size is a decreasing function
of the addition rate. This was qualitatively explained by
a reduction of the nucleation period and better control of
the reaction speed because it proceeds in a starved system
as reactants are added. However, as large addition rate
values were used, the final particle size typically ranged
from 60 to 200 nm. It was concluded that the addition rate
was the preponderant factor (compared to the temperature, the pH, or the TEOS/H2O ratio) on final size and
polydispersity of the particles.
In fact, even if a constant addition of reactant proved
to be beneficial, up-to-now, no systematic study and no
quantitative interpretation of its influence have, so far,
been reported. This is the purpose of the present work.
We explore the role of the addition rate of one of the
reactants in the reaction vessel on the final size and
dispersity of the particles. To get significant results,
experiments were performed over nearly 3 orders of
magnitude in addition rate. Monodisperse particles with
sizes up to 2 m were synthesized. The results of this
single step continuous approach are compared with those
obtained from the seeded polymerization technique, which
also allows for growth of large particles. As we were
interested in quantitatively predicting observations, we
(29) (a) Chen, S. L.; Dong, P.; Yang, G. H.; Yang, J. J. J. Colloid
Interface Sci. 1997, 189, 268. (b) Chen, S. L. Colloids Surf., A 1998, 142,
59. (c) Okubo, T.; Miyamoto, T.; Umemura, K.; Kobayashi, K. Colloid
Polym. Sci. 2001, 279, 1236.
(30) Chen, S. L.; Dong, P.; Yang, G. H.; Yang, J. J. J. Colloid Interface
Sci. 1996, 180, 237.
(31) Giesche, H. J. Eur. Ceram. Soc. 1994, 14, 205.
(32) (a) Zhang, J. H.; Zhan, P.; Wang, Z. L.; Zhang, W. Y.; Ming, N.
B. J. Mater. Res. 2003, 18, 649. (b) Reculusa, S.; Poncet-Legrand, C.;
Ravaine, S.; Mingotaud, C.; Duguet, E.; Bourgeat-Lami, E. Chem. Mater.
2002, 14, 2354.
(33) (a) Kim, K. D.; Bae, H. J.; Kim, H. T. Colloids Surf., A 2003, 224,
119. (b) Kim, K. D.; Kim, H. T. Colloids Surf., A 2002, 207, 263. (c) Kim,
K. D.; Kim, H. T. J. Am. Ceram. Soc. 2002, 85, 1107. (d) Kim, K. D.;
Kim, H. T. J. Sol-Gel Sci. Technol. 2002, 25, 183. (e) Kim, K. D.; Kim,
H. T. Mater. Lett. 2003, 57, 3211.
(34) Fogler, H. S. Elements of Chemical Reaction Engineering: Rate
Laws and Stoichiometry; Prentice-Hall: Englewood Cliffs, NJ, 1986; p
59.
(35) (a) Park, S. K.; Kim, K. D.; Kim, H. K. J. Ind. Eng. Chem. 2000,
6, 365. (b) Reculusa, S.; Mass, P.; Ravaine, S. J. Colloid Interface Sci.
2004, 279, 471.
(36) Park, S. K.; Kim, K. D.; Kim, H. K. Colloids Surf., A 2002, 197,
7.

1518

Langmuir, Vol. 21, No. 4, 2005

Nozawa et al.

Table 1. Size and Dispersity Evolutions of Silica Particles in Seeded Growth Experiments
initial
solution

final
solution

solvent
EtOH (mL)

catalyst 28%
NH4OH (mL)

25 mL of S
25 mL of Sa

S
Sa
Sb

50
175
175

3
12
12

1.5 mL
6.0 mL
6.0 mL

25 mL of Sa
50 mL of Sc

Sc
Sd

175
350

12
24

6 1 mL/45 min
12 1 mL/45 min

Table 2. Experimental Conditions Used for Obtaining


Silica Particles under Continuous Addition
TEOS (mL)
EtOH (mL)
NH4OH (mL)
rates of addition of TEOS
(mL/min)
stirring rates (rpm)
reaction time after addition (h)
reaction temperature (C)

5 (solution I)
30 (solution I)/50 (solution II)
9.5 (solution II)
0.005-1.0
300-700
12
10-40

performed kinetics experiments to find the mechanisms


that govern the growth of silica particles under TEOS
addition. This allows us to build a model that predicts the
variation of the final particle size versus the addition rate.
In view of the quantitative agreement observed between
experiments and predictions, our study allows us to deduce
the optimal conditions for preparing tailored particles in
a single step process by simply controlling one parameter,
i.e., the addition rate of one reactant (TEOS/EtOH).
Experimental Section
Starting Solutions. TEOS (Si(OC2H5)4, 99%, Aldrich Chemical Co.), ethanol (EtOH, J.T. Baker, 99.9% v/v), and ammonia
(NH4OH, 28%, Aldrich Chemical Co.) were used as starting
materials without any further purification. The solutions were
prepared at room temperature under inert atmosphere. An oil
bath was used to control the temperature of the reactions with
an accuracy of (0.05 C (LAUDA, E200, ecoline SRE2312).
Preparation and Analysis of SiO2 Particles. In the few
seeded growth experiments, the preparation of the first seed
suspension was performed according to Stober synthesis.15 After
the reaction came to completion (12 h after the addition of TEOS),
the resulting solution (S) was used as the first seed solution for
a further growth reaction leading to a second generation of seeds
Sa. We reiterate the process as long as the new solution was
found to be monodisperse by transmission electron microscopy
(TEM) measurements, according to details given in Table 1.
The desired amounts of TEOS were added, in a single step or
by small fractions, over various periods of time, into the seed
suspension to give solutions Sb, Sc, and Sd, respectively. The
reaction conditions were kept for another 6 h to make sure that

Figure 1. Schematic drawing of the experimental setup used


for the controlled particle growth procedure under continuous
addition.

reactant TEOS

av diameter
(nm)
110 ( 20
220 ( 20
380 ( 20
100 ( 20
470 ( 20
870 ( 10
420 ( 20
200 ( 20

particles reached their final size. The SiO2 dispersions were then
transferred out of the reactor, and the powders were washed
with ethanol and ultrapure water by repeated centrifugation (at
5000 rpm for 15 min) and further dried at 70 C for 12 h.
In the other set of experiments, the monodispersed spherical
silica particles were prepared by the hydrolysis of TEOS according
to the following procedure. Two solutions, I (TEOS in ethanol)
and II (ammonia in ethanol), were prepared separately. To
effectively investigate the rate of addition effect of TEOS in
ethanol, the total volumes of solutions I and II as well as the
reactant concentrations were kept the same in all the experiment
series (Table 2). Solution I was added, via a micro feed pump
with chosen constant flow rates, under an argon blanket into a
round-bottom flask that contains solution II, under various
stirring at controlled temperatures. The whole mixture was
allowed to react for 12 h. The schematic diagram of the experiment
is shown in Figure 1.
Analysis of SiO2 Particles. TEM was performed at room
temperature on a JEOL JEM-2000 FX transmission electron
microscope, using an accelerating voltage of 200 kV. Scanning
electron microscopy (SEM) was performed on a JEOL JSM-840A,
scanning electron microscope (diameters of about a hundred
particles were used to evaluate the average size and the standard
deviation for each sample). Particle sizes were also checked
optically using a particle size analyzer, Mastersizer 2000
(Malvern Instruments). For diluted suspensions, measurements
were confirmed from the temporal behavior of the angular
variation of the correlation function of the scattered field obtained
with a homemade dynamic light scattering apparatus using a
continuous wave Ar+ laser (wavelength in a vacuum 0 ) 514
nm) and an ALV5000 correlator.

Results and Discussion


Particle Growth via Seeded Experiments. In our
objective to produce monodisperse silica particles of
various controlled sizes, we first performed experiments
based on the seeded growth method from Stober synthesis,
as shown in Table 1. The ratios of added reactants were
calculated in order to double the size of the particles from
one generation to another according to Van Blaaderen et
al.24a,37 Therefore, nanoparticles of 200 nm (Sa) were
obtained from silica suspensions of 100 nm seeds (S)
(Figure 2a,b).
In the same way, nanoparticles of 400 nm (Sb) were
grown from the 200 nm particle suspension (Sa). However,
as shown in Figure 2c, the nucleation of a second
population is observed when the TEOS is added in one
step. To avoid this second population, it was necessary to
add the TEOS by small fractions over a long period of
time (Table 1). As an illustration, results are shown with
an addition of 1 mL every 45 min. In this case; highly
monodisperse particles (Sc) are formed and no secondary
population is observed (see Figure 2d). The same strategy
was used to once more double the silica size starting from
Sc, the same rate of 1 mL every 45 min was applied (12
times) and led to a polydisperse distribution with three
major populations for Sd (Figure 3).
(37) Van Blaaderen, A.; Vrij, A. J. J. Collod Interface Sci. 1993, 156,
1.

Synthesis of Monodisperse Silica Particles

Langmuir, Vol. 21, No. 4, 2005 1519

Figure 3. Transmission electron micrograph of silica particles


showing the limits of the seeded method. The seed solution is
the one presented in Figure 2d. The fractional addition of TEOS
used in this case leads to three major populations (P1, P2, P3).

Figure 2. Transmission electron micrographs of silica particles


illustrating the particle size obtained via the seeded technique:
(a) starting solution (S); (b) doubling of the particle size in a
single step (Sa); (c) starting from Sa, attempt of doubling the
particle size for a single addition of TEOS, formation of a second
population of particles; (d) starting from Sa, doubling of the
particle size by fractional addition of TEOS (see details in Table
1).

These results strongly suggest that a systematic approach of the role of the addition rate of the TEOS reactant
is necessary to bring new insight on the conditions
necessary to confidently control the particle size over a
large variation. Such an approach will help define the
appropriate requirements to produce a direct synthesis of
large monodisperse particles without using any seeds and
time-consuming multistep reaction scheme.
Addition Rate Effect on Final Particle Size.
Experiments were performed at controlled temperature
with a systematic molar ratio for TEOS/NH3/H2O of 1/6.3/
15.2, which is an overall ratio since TEOS is added
dropwise via a micro feed pump and therefore in very
small quantities as compared to water and ammonia
amounts. To determine the relevant mechanisms leading
to the final particle size, we first checked whether the
stirring speed could play a role for a given rate of addition
of TEOS. In a second step, we analyzed by kinetic
measurements, the particle growth rate under controlled
continuous addition. The kinetic evolution of the particle
size is susceptible to provide important information about
the involved growth mechanism. Finally, we investigated
the influence of the variation of the rate of addition of the
TEOS reagent within the reactor on the final size of the
particles.
Effect of Stirring Speed on Final Particle Size.
We performed a set of experiments for a TEOS addition
rate of 0.1 mL/min at T ) 20 C with stirring speeds
varying between 300 and 700 rpm. Using dynamic light

scattering (see below), we found a mean particle diameter


950 ( 20 nm for the whole set of attempts even for the
extreme values (300 and 700 rpm). Despite a ratio larger
than 2 between these extremes, the reproducibility of the
result is not surprising since the associated hydrodynamic
Peclet number, Pe, which compares the particle advection
versus the solute diffusion in a reactor,38 is much smaller
than unity in both cases. Indeed, by definition, one has
Pe ) r2/Dm, where , r, and Dm are respectively the
angular velocity associated to the stirring, the particle
radius, and the molecular diffusion coefficient of the solute
within the mixture. Using the ethanol viscosity ) 1.2
10-3 Pas at T ) 20 C as it constitutes the major part
of the solvent phase and a typical solute size 2 , we find
Dm ) 10-9 m2/s. On the other hand, the angular velocity
varies over more than a factor of 2 from 5 Hz (at 300
rpm) to 12 Hz (at 700 rpm). Then, as experiments lead to
a final particle size rf 500 nm, we finally find 10-3 e Pe
e 3 10-3. Consequently, for classical stirring velocities,
the particle growth is totally dominated by diffusion and
hydrodynamic effects can thus be discarded. This point is
clearly illustrated by the particle growth law presented
below because flow effects are known to accelerate the
kinetics compared to diffusion or reaction-limited growth.38
Particle Growth: Theoretical Background. As in
any growth process performed when all the reactants are
added simultaneously, the particle size strongly depends
on the coarsening mechanism39 because the reactant must
be transported to the interface, generally by diffusion,
and then incorporated into the particle by interface
interaction. If the monomer incorporation (the diffusion)
is the fastest process, then the growth is limited by
diffusion (the interface kinetic). In theses conditions, the
growth rate dr/dt of a spherical particle of radius r can be
described by a general expression, which includes both
bulk diffusion and interface reactions40
(38) Baumberger, T.; Perrot, F.; Beysens, D. Phys. Rev. A 1992, 46,
7636.
(39) Marqusee, J. A.; Ross, J. J. Chem. Phys. 1983, 79, 373.
(40) (a) Leubner, I. H.; Jagannathan, R.; Wey, J. S. Photogr. Sci.
Eng. 1980, 24, 268. (b) Leubner, I. H. J. Imaging Sci. 1985, 29, 219. (c)
Leubner, I. H. J. Phys. Chem. 1987, 91, 6069.

1520

Langmuir, Vol. 21, No. 4, 2005

(C(t) - Ceq(r))
dr
) Ki
dt
1 + r

Nozawa et al.

(1)

where C(t) is the concentration at a given time, t, and


Ceq(r) is the equilibrium concentration around the particle.
Ki is the rate constant for the surface integration of
monomer, and -1 is a screening length that compares
bulk diffusion to surface integration effect;  is defined by

 ) Ki/(DmVM)

(2)

where VM is the molar volume of the precipitate. On the


other hand, Ceq(r) is given by the Gibbs-Thomson relation

Ceq(r) ) CS exp(R/r) CS(1 + R/r)

(3)

CS represents the bulk solubility and R is a capillary length

R)2

VM
RT

(4)

is the liquid/particle interfacial surface energy and R is


the universal gas constant. Consequently, by defining the
supersaturation S as S(t) ) (C(t) - CS)/CS, the general
expression of the particle growth rate becomes

dr KiCSR (1 - rC/r)
)
dt
rC
1 + r

(5)

where rC ) R/S represents the critical radius above which


a particle spontaneously grows and below which it
dissolves. The transition between the interface kinetic
limited and the diffusion-limited growth is controlled by
the product r, i.e., from r , 1 to r . 1. This means that
growth is usually limited by the monomer incorporation
at the particle surface at an early stage and eventually
becomes diffusion-limited at large particle radius.
Equation 5 shows that the particle growth is strongly
dependent on the temporal behavior of the supersaturation
S. The kinetics of precipitation can be divided in four main
stages.41 If initially the system is at equilibrium, then
spontaneous nucleation occurs as soon as the supersaturation S reaches its critical value SC. Particles are formed,
and growth will compete with nucleation until S reaches
a maximum. Then, S starts to decrease and drops below
SC. After this nucleation regime, a transient period appears
where S continues to decrease to reach a quasi-steady
state corresponding to the so-called free-growth regime.
Finally, due to the mass conservation, this free-growth
regime cannot survive indefinitely, and growth switches
to the well-known Ostwald ripening where S starts to
decrease again toward its asymptotic value S ) 0.
To retrieve the particle growth law in the free-growth
regime, we assume a constant supersaturation S. In the
interface kinetic limited case (r , 1), eq 5 reduces to

rC
dr KiCSR
)
1dt
rC
r

(6)

Integration of eq 6 for r/rC . 1 shows that the particle


size increases linearly with time during the free-growth
regime. This r t regime is then followed by the Ostwald
ripening regime where mass conservation leads to a
slowing down characterized by r t1/2.39
(41) Lamer, V. K.; Dinegar, R. H. J. Am. Chem. Soc. 1950, 72, 4847.

On the other hand, growth limited by diffusion presents


different kinetic behaviors. In this case (r . 1), eq 5
becomes

dr DmVMCSR (1 - rC/r)
)
dt
r/rC
r 2

(7)

For r/rC . 1, eq 7 shows that the diffusion-limited freegrowth is characterized by a first behavior r t followed
by an r t1/2 regime.38,42 These regimes are also followed
by the Ostwald ripening which is described by r t1/3.39,43
Finally, the particle number remains constant in the free
growth regime,44 whatever the mechanism that governs
the growth, either interface kinetic or diffusion limited.
While these growth mechanisms were successfully
investigated in experiments involving phase transition
in fluid media,39 their extension to inorganic colloid
dispersions is much more recent. At first, an extensive
study17 showed that silica beads grow by the incorporation
of hydrolyzed monomers instead of aggregation of smaller
particles. Observation of this mechanism supports the
fact that classical ripening theories should also apply to
growth of inorganic colloids. It has effectively been
demonstrated that late stage growth of Stober particles
was limited by diffusion. Using small-angle X-ray scattering, Pontoni et al.26c have investigated the early stage
growth of silica particles and found an r t regime followed
by the behavior r t1/2. These authors also showed that
the particle number as well as their mass density remains
nearly constant over the investigated reaction time, in
agreement with the predictions deduced from the classical
ripening theories.44 However, while saturation in growth
is observed at the late stage, the transition toward the
Ostwald ripening regime, characterized by a r t1/3
behavior, is not clearly evidenced. This regime has
nevertheless been observed by Oskam et al.,45 during the
coarsening of other types of metal oxide nanoparticles
(i.e., zinc and titanium oxide particles).
Particle Growth in the Presence of the Progressive Addition of One Reactant. Since our main
motivation is to analyze the influence of one of the
reactants addition on Stober type silica particle growth,
we extende Pontonis investigation to a study of the
coarsening while specifically controlling the rate of
addition of the TEOS reactant in the solution. The kinetic
evolution of the particle size in a typical experiment carried
out at 20 C and for a rate of addition of TEOS of V )
0.125 mL/min is presented in Figure 4.
The particle growth was characterized according to the
following procedure. During the TEOS addition, we
extracted 1 or 2 drops of solution at regular time intervals
and quickly dilute it in 10 mL of alcohol in order to
instantaneously quench the reaction. Then, for each
sample, we deduced the mean particle size by dynamic
light scattering. To increase accuracy, the correlation
function was measured by varying the scattering angle
every 5 between 0 and 90. Assuming that growing
particles behave as Rayleigh scatterers, the relaxation
time associated to the temporal behavior of the
correlation function is given by ()-1 ) 2Dpq2, where Dp
is the mass diffusion of the particles and q ) 4n/0 sin(/
2) the modulus of the transfer wave vector (n is the index
(42) Cumming, A.; Wiltzius, P.; Bates, F. S. Phys. Rev. Lett. 1990,
65, 863.
(43) Perrot, F.; Guenoun, P.; Baumberger, T.; Beysens, D.; Garrabos,
Y.; Le Neindre, B. Phys. Rev. Lett. 1994, 73, 688.
(44) Tokuyama, M.; Enomoto, Y. Phys. Rev. Lett. 1992, 69, 312.
(45) Oskam, G.; Hu, Z.; Lee Penn, R.; Pesika, N.; Searson, P. C. Phys.
Rev. E 2002, 66, 011403.

Synthesis of Monodisperse Silica Particles

Langmuir, Vol. 21, No. 4, 2005 1521

Figure 4. Growth law of silica particles performed at 20 C


and for a rate of addition of TEOS of 0.125 mL/mn. The line is
a guide for the eyes. Inset: overview of the early stage growth.
A regime r t1/2 is clearly evidenced. The arrow indicates the
end of the TEOS addition.

of refraction of the solvent, here mainly ethanol). Then,


by fitting the linear behavior of ()-1 versus q2, we obtain
a reliable value of Dp and thus of r. The inset in Figure
4, clearly shows that the growth regime corresponds to an
r t1/2 behavior when applying a progressive TEOS
addition (the temporal exponent measured in Figure 4 is
0.53 ( 0.03). Bearing in mind that the supersaturation
reaches a constant value during the TEOS addition, the
observed particle growth necessarily corresponds to the
free-growth regime; the particle growth had already
switched from interface (r , 1) to diffusion limited (r .
1) at the beginning of the investigated temporal window.
Indeed, for reaction-driven growth, we would instead
observe an r t free-growth regime, as a law r t1/2 can
only appear at the end of the growth during Ostwald
ripening,39 i.e., well after the end of the addition of TEOS.
In these experiments, the diffusive origin of the growth
is also confirmed by the effective persistence of the r t1/2
regime after the end of the addition. Finally, as in Pontonis
experiments,26c we do not clearly identify the Ostwald
ripening regime r t1/3 between the free-growth regime
and the saturation to the final particle size.
Effect of the Rate of Addition on the Final Size of
the Particles. Comforted by the results presented in
Figure 4, we then undertook a systematic study of the
impact of the rate of addition of TEOS on the final size
of the particles. The SEM pictures presented in Figure 5
show the final size of silica particles obtained in strictly
identical conditions at 25 C except for the addition rates
of TEOS that were 0.005, 0.05, and 0.5 mL/min.
They illustrate the simplicity as well as the strength of
the method. First, the particle size is an obviously
decreasing function of the addition rate of TEOS. Then,
the particle size can clearly be monitored over a large
range by playing with the addition rate (typically a factor
of 5 over 2 orders of magnitude in addition rate). This
approach leads in one step and without renucleation to
particle size larger than those obtained in the seeded
process (Table 1). Finally, the particle distribution in the
three different experiments appears to be highly monodisperse with a standard deviation varying from 5% to
2% with decreasing rate of addition. To analyze the
influence of different independent parameters, we investigated the evolution of the final particle size versus

Figure 5. Scanning electron micrographs of silica particles


obtained at 25 C over 2 orders of magnitude in addition rates
of TEOS: (a) 0.005 mL/min (1820 nm, standard deviation 2%);
(b) 0.05 mL/min (1330 nm, standard deviation 3%), 0.5 mL/min
(635 nm, standard deviation 5%).

addition rate of TEOS for three temperatures centered on


ambient temperature (10, 25, and 40 C). The corresponding evolutions are shown in Figure 6.
To increase the accuracy of the measurements, particle
diameters were characterized by three different techniques, SEM, dynamic light scattering, and a Malvern
particle size analyzer. As illustrated in Figure 6, the results
given by the three techniques are in very good agreement
with each other. The general trend confirms that the final
particle diameter, df, decreases with the rate of addition
(V), whatever the temperature.
From a quantitative point of view, a power law fit of the
data sets shows that the general behavior df V-1/3
emerges. Power law fits give, (a) df ) 0.38 V-0.32 for T )
10 C, (b) df ) 0.39 V-0.32 for T ) 25 C, and (c) df ) 0.29

1522

Langmuir, Vol. 21, No. 4, 2005

Nozawa et al.

is the significant increase in duration of the free-growth


regime due to a constant supersaturation resulting from
the equilibrium between reactant addition and particle
growth. This is clearly illustrated in Figure 4 in which the
r t1/2 regime is observed over a much larger time period
than that measured by Pontoni et al. when all the reactants
are added in a single step.
Since the main effect of addition is to increase the
duration of the free growth regime, the number density
of particles should then be time independent, as predicted52
and observed26c in the absence of addition. Different models
of jet precipitation47,53,54 based on particle growth limited
by diffusion, also retrieve this expectation and show that
the particle number N versus addition rate behaves as

N)

Figure 6. Evolution of the particle diameter with the rate of


addition of TEOS at (a) 10 C, (b) 25 C, and (c) 40 C determined
by (4) scanning electron microscopy, (O) dynamic light scattering, and (0) particle size analyzer.

V-0.31 for T ) 40 C when df is expressed in micrometers


and V in mL/min; error on df is less than 5%. Contrary to
the strong dependence on addition rate, these results also
show that temperature variations have almost no influence
on the particle final size over the investigated range. This
result is at variance with experiments performed without
TEOS addition.46
To understand the very simple scaling measured as well
as its robustness, we developed an analogy between the
growth of Stober type spherical silica particles with a
controlled addition rate of TEOS and the formation of
photographic colloids, typically silver halide crystals, in
controlled jet precipitation (see for instance Leubners
review47). Indeed, as it is necessary to control the final
crystal size with an error smaller than 5% in order to get
reliable photographic emulsions, numerous methods have
been implemented including semibatch crystallization,48,49
and it appeared that an interesting way to achieve this
goal was the use of the jet precipitation technique.50,51 It
consists of introducing one of the reactants at a carefully
controlled rate over a defined period into a stirred solution.
In these conditions, the kinetics of precipitation can still
be divided in four stages.47 The main difference with those
presented above for growth under a single step addition
(46) Tan, C. G.; Bowen, B. D.; Epstein, N. J. Colloid Interface Sci.
1987, 118, 290.
(47) Leubner, I. H. Curr. Opin. Colloid Interface Sci. 2000, 5, 151.
(48) Tavare, N. S.; Garside, J. Chem. Eng. Sci. 1993, 48, 475.
(49) Torbacke, M.; Rasmuson, A. C. Chem. Eng. Sci. 2001, 56, 2459.
(50) Stavek, J.; Spek, M.; Hirasawa, I.; Toyokura, K. Chem. Mater.
1992, 4, 545.
(51) Zhong, Q.; Matijevic, E. J. Mater. Chem. 1996, 6, 443.

VRT
8DmVM2CS

(8)

where is a numeric factor depending on the chosen model.


This prediction for free-growth limited by diffusion,
particularly the behavior of N vs addition rate, V, was
experimentally verified in double-jet precipitation (zinc
oxide colloids,51 silver halides40,47,53). However, when
addition stops, as in our experiments, we could imagine
that the constant character of N should break down due
to the change in growth conditions. Owing to Ostwald
ripening, we would expect a continuous reduction of N by
particle coalescence and evaporation, as predicted39 and
observed in liquid-phase transitions43 or some metal oxide
colloids.45 However, since the seminal Stober investigation,15 this scheme has never been observed when experiments involve growth of silica particles. A monodisperse
assembly of silica particles is often found at the end of the
coarsening, a point that makes those particles so attractive
but still misses a quantitative explanation.55 This nonexpected late stage growth behavior is also kinetically
illustrated in both Pontonis investigation26c and Figure
4, in the absence and the presence of TEOS addition,
where the r t1/3 regime that characterizes Ostwald
ripening is clearly absent. This result could be explained
by the fact that the experimental conditions of the reaction (especially in terms of the pH of the solution) are not
strong enough to allow silica dissolution.7 Consequently,
both cross-related aspects, monodisperse particle assembly
and lack of observable Ostwald ripening, strongly suggest
that the constant number of particles during the freegrowth regime N is preserved until the end of the
coarsening and thus corresponds to the final particle
number. By use of mass conservation, this particle number
N is then related to the final radius rf by

4
r 3 ) nSiO2VM
3 f

(9)

where nSiO2 is the mole number of silica corresponding to


the volume of added TEOS. By combining eq 8 and eq 9,
we finally find the expected relation between the final
particle radius and the addition rate

df ) 2rf ) 2

6DmnSiO2VM3CS
RT

1/3

V-1/3

(10)

Consequently, our analogy with jet precipitation leads


to the behavior rf V-1/3, in very good agreement with
(52) Tokuyama, M.; Enomoto, Y. Phys. Rev. Lett. 1992, 69, 312.
(53) Sugimoto, T. J. Colloid Interface Sci. 1992, 150, 208.
(54) Chong, J. J. Imaging Sci. Technol. 1995, 39, 120.
(55) Matijevic, E. J. Langmuir 1994, 10, 8.

Synthesis of Monodisperse Silica Particles

measurements. To compare the measured amplitude factor


between df and V -1/3 with its predicted value, we used the
following data. From literature,7 we find that ) 46 erg/
cm2 and Cs ) 2 10-6 mol/cm3 for 6 e pH e 10, at room
temperature. Assuming a particle density of 1.8 g/cm3,
corresponding to the typical density of Stober particles,7
we deduce VM ) 33.3 cm3. Moreover, as we added 5 mL
of TEOS (molar mass 208 g and density 0.934 g/cm3)
within the reactor, the corresponding mole number of
silica is nSiO2 ) 2.25 10-2 mol. Finally, considering
the different existing jet precipitation models, one has53
1 e e 3; note that this dispersion in values has
only a weak influence on the final results due to the
one-third power appearing in the amplitude factor of eq
10. Consequently, when df and V are respectively expressed in micrometers and cm3/min, we find 0.3 e
2[6DmnSiO2VM3Cs/(RT))]1/3 e 0.43, in excellent agreement
with the measured values (0.3-0.4).
Conclusion
A Stober-like synthesis of silica particles was performed
in which the control of the addition rate of one of the
reactant (TEOS) was the main varying parameter. This
investigation was motivated by the necessity to obtain
the synthesis of low cost reproducible mesoscopic monodisperse silica particles (i.e., of diameter typically larger
than 1 m) based on the standard Stober method. The
method presented here fulfills this requirement since it
strongly lowers the time spent to increase particle size by
using the seeded method which also implies much more
manipulation as well as dilution of starting solution at
each step. The cost is also decreased in terms of reactants
consumption and solvent recycling. Furthermore this
method allows the use of classical EtOH and Si(OEt)4
even when sizes as big as 2 m are targeted. It is therefore
cheaper than the classical Stober method, which uses more
expensive reactants, such as pentyl ester silicon derivative,
and often exhibits a wider size distribution.
Besides the use of seeds to induce further growth which,
as illustrated in the first part of our investigation,
eventually leads to the nucleation of new generations of
particles, we rationalize here a method which totally
circumvents this renucleation. Indeed, our systematic
study of the impact of the rate of addition on the final size
of the particles for three temperatures (10, 25, and 40 C)
shows that the particle size decreases as the rate of
addition increases whatever the temperature, according
to a very simple power law. As a corollary, this means
that one can work at any temperature fulfilling the
condition of nonevaporation of reactant. Moreover, monodispersity in particle distribution is always observed (from
5% down to 2%). Finally, the production of the desired
size particle is performed in one step, instead of several
ones as in seeded experiments, just by adjusting the
addition rate. These two last points are of very practical

Langmuir, Vol. 21, No. 4, 2005 1523

importance in the sense that particle manipulation is no


more required and the time consumption is strongly
reduced within the processing, as already mentioned. To
quantify these results, we first analyze the particle growth
rate maintaining a specific addition rate of the reactant.
These measurements reveal that growth of silica particles
is limited by diffusion, as already demonstrated when all
the reactants are added in a single step. Considering this
property and using an analogy with crystal formation in
jet precipitation, we build a growth model that takes the
addition rate into consideration. As for the free-growth
regime in classical phase transitions, we show that the
particle number remains constant during the addition.
The main advantage of this model is to demonstrate that
the particle number is directly proportional to the addition
rate and to give an expression for the proportionality
constant. Using this approach and taking into account
the mass conservation, we retrieve the observed power
law for the variation of the final particle radius versus
addition rate. The amplitude of the predicted power law,
associated to the proportionality constant between particle
number and addition rate, was also confronted in experiments. We found very good quantitative agreement
between theory and experiments. Even if data are missing
for a quantitative comparison, particularly the mole
number of silica corresponding to the volume of added
TEOS, we also find good qualitative agreement with the
two experiments performed at 0.6 and 18 mL/min,36 since
(i) a decrease in particle size was observed for increasing
addition rates and (ii) the addition rate was considered
as the dominant factor on final size and polydispersity of
the particles, as compared to the temperature for instance.
Therefore, the agreement observed between our measurements with results obtained at large addition rates for
synthesis of nanoparticles36 illustrates over almost 4 orders
of magnitude in addition rate (from V ) 5 10-3 to 18
mL/min) how monodisperse silica particles can be synthesized using a single step process with size ranging from
a few tens of nanometers to a few micrometers. Consequently, the simplicity of this method opens new prospects
in the synthesis of silica particles and offers a flexible
level of particle size reproducibility that could be very
appealing for nanoscopic to mesoscopic particle assembling
(photonics, chromatography).
Acknowledgment. The authors want to thank Dr. S.
Reculusa for helpful discussions and A. S. Beaumont and
C. Sableaux for some experimental work.
Note Added after ASAP Publication. References 32b
and 35b and the Acknowledgment paragraph were omitted
in the version published ASAP December 18, 2004; the
corrected version was published ASAP January 18, 2005.
LA048569R

Vous aimerez peut-être aussi