In Situ Production of Spherical Aerogel Microparticles

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

J.

of Supercritical Fluids 55 (2011) 1118–1123

Contents lists available at ScienceDirect

The Journal of Supercritical Fluids


journal homepage: www.elsevier.com/locate/supflu

In situ production of spherical aerogel microparticles


M. Alnaief ∗ , I. Smirnova
Hamburg University of Technology, Thermal Separation Processes, Eißendorfer Straße 38, D-21073 Hamburg, Germany

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

Article history: Due to their high surface area, low density, open pore structure and excellent insulation properties aero-
Received 4 June 2010 gels are intensively investigated since the past decades for a diverse range of applications. The current
Received in revised form 4 October 2010 methods of silica aerogel production by supercritical extraction produce monolithic aerogels, where
Accepted 5 October 2010
the sol is aged in molds and dried by extraction with supercritical CO2 . Aerogels in the form of spheri-
cal microparticles would be beneficial for many applications, for instance, drug delivery for respiratory
Keywords:
route; or as insulating materials. However, because of aerogel’s mechanical properties, it is difficult,
Spherical nanoporous microparticles
rather impossible, to obtain spherical microparticles by milling or crushing of the monolithic aerogels.
Emulsion
Supercritical extraction
This work presents a new method to produce biocompatible spherical aerogel microparticles using an
Silica aerogel emulsion technique (in situ production) followed by supercritical extraction of the resulted dispersion
(gel–oil). Water in oil emulsion was produced by mixing the sol (dispersed phase) with a vegetable oil
(continuous phase) followed by the gelation of the dispersed phase. The size distribution of the final gel
particles was found to be influenced by agitation, surfactant concentration and sol:oil volume ratios. The
gel–oil dispersion was subsequently extracted with supercritical CO2 , Silica aerogel spherical micropar-
ticles with a surface area of 1100 m2 g−1 , pore volume of 3.5 cm3 /g and different mean particle diameters
ranging from 200 ␮m to a few millimeters were produced using the presented method.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction from the gel is a challenge in aerogel production. Thus, a modifi-


cation of the established emulsion process is required to overcome
Aerogels are nanoporous materials with a high surface area and this challenge [15].
an open pore structure. They are usually produced by supercritical The aim of this work is to establish the emulsion-based method
extraction of a stable gel, which is prepared following the sol–gel to produce microspherical aerogel particles with controlled parti-
technology. Because of its extraordinary properties, silica aero- cle size distribution. The suggested process consists of four basic
gels have been gaining increasing attention since the past decays, steps: (1) preparation of the disperse phase by the sol–gel process;
for instance, in space engineering [1], thermal insulation [2], ultra (2) emulsification of the disperse phase in a continuous phase (oil,
high-density capacitors [3], solar-energy collectors [4], Cherenkov immiscible with the first one); (3) cross linking/gelation reaction
detector [5], waste treatment [6], drug delivery and targeting sys- within the dispersed phase (liquid microdroplets) to form stable gel
tems [7–11]. microspheres; (4) CO2 supercritical extraction of the oil–gel disper-
Microspherical shaped aerogel particles would be highly desir- sion to obtain the final microsphere aerogel particles. These steps
able for some of those applications, especially for drug delivery. are schematically represented in Fig. 1.
Microspheres can be principally produced using various well-
known techniques [12–14], however, because of the sensitivity of
aerogels, it is necessary to modify the used processes to meet the 2. Experimental methods
demands of aerogel production. One possible method to produce
microspheres is based on emulsion formation. Emulsion formation 2.1. Reagents
is a well-known process that would allow production of a large
amount of microspherical droplets in a robust and controlled man- Carbon dioxide with a purity of 99.9% was supplied by AGA
ner. Still, production of a stable gel and extraction of the solvent Gas GmbH (Hamburg, Germany). Tetramethoxysilane (TMOS) 98%
was purchased from Fluka Germany, methanol 99.5%, ethanol
(p.A.), hydrochloric acid 30% and ammonium hydroxide 25% were
∗ Corresponding author. Tel.: +49 40 42878 4392; fax: +49 40 42878 4072.
purchased from Merck Germany, domestic grade of canola oil was
E-mail addresses: mohammad.alnaief@tu-harburg.de,
used. All chemicals were used as provided without any further
mohammad.alnaief@tuhh.de (M. Alnaief). processing.

0896-8446/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.supflu.2010.10.006
M. Alnaief, I. Smirnova / J. of Supercritical Fluids 55 (2011) 1118–1123 1119

Fig. 1. Schematic overview over the four main steps in aerogel microsphere preparation by emulsion method combined with supercritical extraction.

2.2. Methods arator. Typically the recycled CO2 was exchanged for fresh CO2 at
least four times during the extraction process to ensure the com-
2.2.1. Preparation of silica gel microspheres plete extraction of the gel. The extraction of a gel using supercritical
Silica sol was produced following the two step sol–gel process CO2 is a mass transfer problem in which the diffusion rate of the
[10,11]. In the first step tetramethylorthosilicate (TMOS), methanol, ethanol from the gel pores is inversely proportional to the square
water, and hydrochloric acid were mixed together with a molar of the radius (cross sectional radius of the cylinder, or the radius of
ratio of: 1 mol TMOS:2.4 mol MeOH:1.3 mol H2 O:10−5 mol HCl. the sphere) [18]. Thus, the extraction time needed to produce aero-
The mixture was stirred at room temperature for 30 min. gels in form of microparticles is much shorter than that needed for
After that, the mixture was diluted with ethanol to obtain the the supercritical extraction of the same gel volume in form of a
desired density of the aerogel. Then, additional water and ammo- monolith.
nia solution were added to obtain the following molar ratio: 1 mol It should be mentioned that it was also possible to separate
TMOS:2.4 mol MeOH:4 mol H2 O:10−5 mol HCl:10−2 mol NH4 OH. oil from the gel phase by filtration before supercritical extraction,
Finally the sol solution was stirred for 3 more minutes. however, since oil is very good extractable with supercritical CO2 it
was more convenient to skip this step and conduct the extraction
2.2.2. Emulsification of the silica sol in an oil phase for the complete oil–gel dispersion, since the extracted oil–ethanol
Canola oil (continuous phase) saturated with ethanol was placed mixture can be recycled.
in a 600 ml vessel and mixed using a stirrer with a constant stir-
ring rate. The sol (dispersed phase) was then poured into the oil 2.3. Analysis
phase at once. Subsequently, droplets of the dispersed phase were
formed under continuous stirring. After 20–30 min gelation of the The produced aerogel microspheres were characterized using
dispersed phase took place. At this stage the stirring was stopped different methods to evaluate the production process and the effect
since the emulsion converted into dispersion of spherical gel parti- of process parameters. Particle size distribution was measured
cles in the oil phase. Thus the described method is differ from that using laser diffraction spectrometer (Sympatec Helos). Surface area,
were CO2 plays as antisolvent to precipitate the dispersed phase pore size and pore size distribution were analyzed using nitro-
from an emulsion [16], furthermore, the supercritical CO2 phase gen adsorption/desorption (surface analyzer Nova 3000e). Particle
will not affect the emulsion properties [17], since simply at super- shape and form were analyzed using light microscopy (Zeiss Micro-
critical extraction stage there is no emulsion anymore. Finally the scope Axio Scope) and scanning electron microscopy (SEM) (Leo
gel spheres were left over night in the oil phase for aging. (Zeiss) 1530).
Different process parameters were investigated to obtain the
desirable particles size and shape. 3. Results and discussion

2.2.3. Supercritical extraction of the gel–oil dispersion Particle shape and size distribution produced by the emulsion
A process flow diagram of the supercritical extraction is shown process depend on different parameters, these parameters can be
in Fig. 2. In a typical experiment, the gel–oil dispersion was placed classified to three main groups: (1) geometrical parameter, like
into the 4 l cylindrical stainless steel vessel. Supercritical (SC) the mixer shape and diameter; (2) physical parameters like vis-
CO2 was delivered using a high pressure diaphragm pump and cosity and density; (3) process related parameter, like stirring rate
was introduced from the top of the vessel at constant flow rate [19]. Mixer shape and speed are the easiest parameters to con-
(100–200 g/min). Temperature was maintained constant (40 ◦ C) trol the dispersion of the dispersed phase in the continuous phase.
using an oil heating jacket. At the outlet of the vessel SC-CO2 loaded However, the minimum attainable droplet size depends on many
with solvent and oil was directed to another 2 l cylindrical stainless other parameters like the viscosity of the dispersed and continu-
steel vessel (separator), where oil and solvent were separated from ous phases, the interfacial tension between the two phases, their
CO2 . The pressure and temperature of the separator were main- relative volume ratio, the size ratio of mixer to the mixing vessel,
tained constant at 40 ◦ C and 60 bar respectively. Solvent-lean CO2 number of mixers, etc. [1].
was then recycled for the process. After 8 h the extraction was com- In this work, the following parameters were investigated in
pleted and the aerogel microspheres were removed from the 4 l order to control the final product properties: stirrer geometry,
autoclave and the solvent–oil mixture from the bottom of the sep- phase ratio, surfactant concentration, stirring rate. The influence
1120 M. Alnaief, I. Smirnova / J. of Supercritical Fluids 55 (2011) 1118–1123

Fig. 2. Schematic diagram of the 4 l supercritical extraction unit.

of these parameters on the final properties of the resulting silica cles. Based on these findings, the sol:oil ratio of 1:1 was chosen for
aerogel’s microparticles is discussed in details in the following. further experiments, since smaller amount of oil is preferable for
further extraction.
3.1. Effect of the phase ratio (dispersed to continuous phase)
3.2. Particle shape and size distribution: effect of agitator shape
The impact of the volume ratio between the dispersed and the
continuous phase on the PSD of the resulting microspheres is not For the first set of experiments, a four bladed radial stirrer
fully understood. Various studies report the reduction of the mean (diameter = 5 cm) was used to emulsify 200 ml sol–oil mixture (1:1
particles size with a reduction in the volume of the continuous volume ratio) in a 600 ml vessel (9 cm internal diameter). The
phase [20], while other studies report that no significant effect was resulting aerogel particles were observed by light microscopy. It
observed [21]. In this study, four different sol:oil volume ratios was noticed that for a high stirring rate (1000–1800 rpm) parti-
were investigated, namely: 2:1, 1:1, 1:2 and 1:3, for all experi- cles were not uniform in shape (Fig. 3a) and some breakages were
ments a constant total volume (100 ml) and stirring rate (500 rpm) observed. However, at a lower stirring rate, 800 rpm, microspheri-
was used. For the sol:oil ratio of (2:1) no microspheres could be cal gel particles with the diameter between 100 and 200 ␮m were
obtained. Since the sol phase is larger than the oil phase, there was obtained (Fig. 3b). Although the particles were spherical, for many
no chance to get sol droplet dispersed in the oil phase. When the applications it is beneficial to have a smaller mean particle size (less
gelation took place the sol phase transformed to a separate con- than 100 ␮m) than those obtained with these conditions. How-
tinue gel phase. For all other investigated phases volume ratios it ever, it is not possible using this kind of stirrer due to the high
was possible to obtain microspherical particles. It was noticed that shear stresses applied to gel particles on the tip of the blade dur-
the mean particle size increases as the oil phase increases (Table 1). ing mixing. Moreover, radial mixers result in such a flow profile
Taking into consideration that the viscosity of the oil phase is about that particles collide with the vessel walls. This amplifies high
50 cP [22], and that of aqueous phase is about 1 cP, then by increas- collision energy which leads to the abrasion and destruction of
ing the oil phase ratio the total viscosity of the system increases, the spherical gel particles. For this reason another type of stirrer,
consequently more energy is needed to form the same micro parti- which allows high energy input and smooth flow pattern, namely,
a two bladed axial stirrer, was chosen. For this type of stirrer it
Table 1 was possible to produce spherical particles at much higher revolu-
Particle size distribution of the aerogel microspheres as a function of water/oil phase tion speeds (up to 1800 rpm) (Fig. 3c). To understand the different
ratio (0% surfactant, 500 rpm). behaviors of these stirrers, the power number and the stirrer power
Phase ratio w/o 1:1 w/o 1:2 w/o 1:3 of these two different stirrers were compared. The power num-
Average [␮m] 722 ± 15 803 ± 16 853 ± 17
ber for the 4 bladed stirrer is 5 with a stirrer power of 38.8 W,
D10 [␮m] 197 ± 24 338 ± 15 317 ± 48 whereas that of the axial stirrer was only 0.4 and 7.7 W as stirrer
D50 [␮m] 703 ± 16 813 ± 13 876 ± 17 power, keeping all other process parameters unchanged (viscos-
D90 [␮m] 1255 ± 26 1243 ± 21 1345 ± 27 ity, density, dimensions, rpm). The values results from the two
D32 [␮m] 269 ± 11 339 ± 8 368 ± 9
bladed axial stirrer at 1800 rpm correspond to a stirring rate of
M. Alnaief, I. Smirnova / J. of Supercritical Fluids 55 (2011) 1118–1123 1121

Table 2
Particle size distribution of the aerogel microspheres as a function of the revolution speed during emulsification.

200 rpm 300 rpm 400 rpm 500 rpm 600 rpm 800 rpm 1300 rpm

Average [␮m] 1720 ± 23 1615 ± 70 996 ± 95 908 ± 32 601 ± 35 323 ± 27 155 ± 7


D10 [␮m] 1166 ± 6 987 ± 100 522 ± 36 495 ± 40 281 ± 60 194 ± 22 46 ± 2
D50 [␮m] 1729 ± 18 1645 ± 60 957 ± 80 842 ± 18 566 ± 28 310 ± 22 143 ± 2
D90 [␮m] 2267 ± 14 2214 ± 36 1509 ± 100 1386 ± 100 957 ± 18 494 ± 16 277 ± 14

Fig. 3. Effect of stirrer type and revolution on the form of silica gel particles: (a) 4 blade stirrer at 1800 rpm; (b) 4 blade stirrer at 800 rpm; (c) propeller at 1800 rpm.

1000 rpm of the four bladed stirrer [19]. Based on these findings, 3.4. Effect of surfactant concentration
all further experiments were conducted using two bladed axial
stirrer. A stabilizer (surfactant) is generally added to the continu-
ous phase to prevent coalescence of the droplets. Increasing the
3.3. Effect of stirring rate surfactant concentration leads to a reduction of the interfacial
surface tension, as a result smaller droplets sizes are usually pro-
In order to investigate the effect of the stirring, the two bladed duced (under the same stirring conditions). Taking account the
axial stirrer (diameter = 6 cm) was used to emulsify 200 ml of sys- hydrophilic–lipophilic balance (HLB) it was found that span® 40
tem (sol:oil phases volume ratio = 1:1) with different revolution (HLB 6.7) is a suitable surfactant for the used system(HLB Canola
speeds. The particle size distributions (PSD) as a function of the oil 7 ± 1) [29]. Keeping all process parameters constant, the con-
stirring speed are determined by laser scattering analysis reported centration of the surfactant was varied between 0 and 10 wt%
in Table 2. The mean particle diameter (D50 ) varies from 1730 ␮m related to the continuous phase (oil). The resultant particles were
to 143 ␮m, when the stirrer revolution rate varies from 200 to characterized using laser diffractometer to estimate the particles
1300 rpm respectively (Table 2). Increasing the mixer speed results size distribution (Table 3). It can be seen that by increasing the
in a higher energy input to the system, this allows the building of surfactant concentration a reduction of the mean particle size is
larger interfacial surface area, thus the dispersed droplets become obtained. This is also confirmed by the decreasing the Sauter diam-
smaller, and consequently the gel microspheres will have a smaller
mean size [23]. These results confirmed that the main factor in Table 3
Particle size distribution of the aerogel microspheres as a function of surfactant
controlling the PSD is the stirrer revolution rate [23–28]. It can be
concentration (w/o 1:1, 500 rpm).
noticed that the PSD is relatively broad in all cases. Among many
other factors, viscosity of the dispersed as well as the continuous Surfactant concentration [wt%]
phase plays a major role in controlling PSD in the emulsion process 0 1 3 5 10
[23]. In our process, the viscosity of the continuous phase remains
Average [␮m] 861 ± 22 762 ± 41 733 ± 12 684 ± 28 584 ± 72
nearly constant (50.8 cP) throughout the experiment; however, the D10 [␮m] 233 ± 81 180 ± 32 298 ± 17 215 ± 22 169 ± 69
viscosity of the dispersed phase (sol phase) increases continuously D50 [␮m] 863 ± 37 788 ± 24 736 ± 32 677 ± 37 582 ± 41
with time until the formation of the gel particles. This might lead to D90 [␮m] 1516 ± 18 1260 ± 31 1152 ± 45 1154 ± 19 988 ± 60
the broadening of the resulting particle size distribution (Table 2). D32 [␮m] 290 ± 21 274 ± 19 294 ± 16 251 ± 15 221 ± 35
1122 M. Alnaief, I. Smirnova / J. of Supercritical Fluids 55 (2011) 1118–1123

Table 4
Internal surface characterization of aerogel microspheres produced under different parameters.

Production process BET surface area [m2 /g] BJH pore volume [cm3 /g] Average pore diameter [nm] C constant BET

Monoliths silica aerogel 982 ± 54 4.13 ± 0.34 16 ± 2 95 ± 12


1:1 phase ratio 1123 ± 56 3.42 ± 0.17 16 ± 1 47 ± 3
1:2 phase ratio 1107 ± 55 – – 41 ± 4
1:3 phase ratio 1068 ± 53 3.54 ± 0.27 15 ± 2 48 ± 3
0 wt% span® 40 1189 ± 58 3.53 ± 0.24 15 ± 2 41 ± 2
3 wt% span® 40 1191 ± 60 3.36 ± 0.18 16 ± 1 45 ± 3
10 wt% span® 40 1110 ± 55 – – 46 ± 4

eter D32 , which indicate that larger surface area (smaller particles) 4.0 16.0
is obtained from the same volume of particles at higher surfactant

Cumulative Pore Volume [cm /g]


3.5 14.0
concentration (Table 3).

Dinsity Distribution q3 [cm /g]


3
3.0 12.0

3.5. Structural properties of the aerogel microparticles 2.5 Cumulative Pore Volume 10.0

Density Distribution q3
2.0 8.0
For the first set of experiments, pure vegetable oil was used as
a continuous phase. It was noticed that the produced particles had 1.5 6.0
extremely high densities in the range of (0.3–0.4 g/cm3 ); moreover,
the surface areas were very low (100–300 m2 /g). The reason for this 1.0 4.0

3
is that approx. 15 wt% of the ethanol used in the sol–gel process dis-
0.5 2.0
solves in the oil phase at 25 ◦ C and up to 25 wt% at 40 ◦ C; As a result,
the gel matrix shrinks leading to a denser particle with smaller sur- 0.0 0.0
face area. This problem was solved by saturating the oil phase with 0 50 100 150 200 250 300
ethanol prior to the emulsification. Fig. 4 shows typical silica aero- Pore Size Å
gel microspheres produced by the suggested process. The particles
clearly display a spherical shape. For all samples, density, surface Fig. 5. Typical pore size distribution of silica aerogel microspheres.

area, pore size and pore volume were measured. It was observed
that these properties do not depend on the specific parameter of the
4. Conclusions
emulsion process, but rather on the sol–gel process itself (catalysts,
component ratio etc.). Average values for the produced aerogel
Combining the sol–gel process with supercritical extraction of
micropartiles resulting from the multiple measurements (n ≥ 3) are
the oil–gel dispersion is proposed as a versatile method for indus-
shown in Table 4. A typical pore size distribution is shown in Fig. 5.
trially relevant production of aerogel microspheres in a robust and
The average packed density of all produced aerogel microspheres
reproducible manner allowing shortening the extraction time in
is 0.07 ± 0.02. These internal structural properties are compara-
comparison to monolithic aerogels. The process was demonstrated
ble with those of the controlled monolithic silica aerogel samples
for silica aerogel microspheres. The influence of the process param-
produced using the same sol–gel procedure without emulsifica-
eters (stirrer shape, stirring speed, surfactant, phase volume ratio)
tion step [7]. Thus, the proposed method allows the production
on the properties of the produced particles was investigated. Stir-
of aerogels in form of spherical microparticles by supercritical
rer shape and revolution rate were found to be the main factors
extraction of dispersion, keeping the internal structural proper-
that control the PSDs of the produced microspheres. Depending
ties similar to that of the monolithic samples which represent the
on the production conditions, silica aerogel microspheres with a
state of the art so far. Since this form of the aerogel is beneficial
mean diameter from 155 ␮m to 1.7 mm were produced. The inter-
for many applications, the process can be applied for the fast pro-
nal structural properties of the aerogels (surface area, pore size)
duction of large amount of aerogels microspheres of corresponding
were less affected by the emulsion preparation, but rather by the
size.
composition of the sol itself and showed similar values to the cor-
responding monolithic aerogels. Principally, the process suggested
in this work allows for the large scale production of supercritically
dried aerogel microspheres of different origin. The extension of this
process to organic aerogels (i.e. alginate aerogel microspheres), as
well as the effect of other process parameters on the final properties
of the produced aerogel microspheres are under investigation.

Acknowledgment

The first author (Mohammad Alnaief) is thankful for the


German-Jordan University for supporting him with a personal
scholarship.

References

[1] M.A.B. Meador, E.F. Fabrizio, F. Ilhan, A. Dass, G.H. Zhang, P. Vassilaras, J.C. John-
ston, N. Leventis, Cross-linking amine-modified silica aerogels with epoxies:
Fig. 4. SEM image of silica aerogel microspheres. Experimental conditions: mechanically strong lightweight porous materials, Chemistry of Materials 17
(700 rpm, two bladed axial stirrer (D: 6 cm), 1:1 phase ratio). (2005) 1085–1098.
M. Alnaief, I. Smirnova / J. of Supercritical Fluids 55 (2011) 1118–1123 1123

[2] S.A. Omer, S.B. Riffat, G. Qiu, Thermal insulations for hot water cylinders: a [16] F. Mattea, Á. Martín, A. Matías-Gago, M.J. Cocero, Supercritical antisolvent
review and a conceptual evaluation, Building Services Engineering Research & precipitation from an emulsion: [beta]-carotene nanoparticle formation, The
Technology 28 (2007) 275–293. Journal of Supercritical Fluids 51 (December) (2009) 238–247.
[3] A. Burke, Ultracapacitors: why, how, and where is the technology, Journal of [17] F. Mattea, Á. Martín, C. Schulz, P. Jaeger, R. Eggers, M.J. Cocero, Behavior of an
Power Sources 91 (2000) 37–50. organic solvent drop during the supercritical extraction of emulsions, AIChE
[4] M Reim, A. Beck, W. Korner, R. Petricevic, M. Glora, M. Weth, T. Schliermann, J. Journal 56 (2010) 1184–1195.
Fricke, C. Schmidt, F.J. Potter, Highly insulating aerogel glazing for solar energy [18] R.B. Bird, W.E. Stewart, E.N. Lightfoot (Eds.), Transport phenomena, second
usage, Solar Energy 72 (2002) 21–29. Wiley International ed., John Wiley, New York, 2002.
[5] P. Krizan, Recent progress in particle identification methods, Nuclear Instru- [19] M. Zlokarnik, Stirring: Theory and practice, Wiley-VCH, Weinheim, New York,
ments & Methods in Physics Research Section A-Accelerators Spectrometers 2001.
Detectors and Associated Equipment 598 (2009) 130–137. [20] H. Jeffery, S.S. Davis, D.T. O’Hagan, The preparation and characterisation
[6] M.S. Ahmed, Y.A. Attia, Aerogel materials for photocatalytic detoxification of of poly(lactide-co-glycolide) microparticles: I. Oil-in-water emulsion solvent
cyanide wastes in water, Journal of Non-Crystalline Solids 186 (1995) 402– evaporation, International Journal of Pharmaceutics 77 (1991) 169–175.
407. [21] F. Gabor, B. Ertl, M. Wirth, R. Mallinger, Ketoprofen- poly(d,l-lactic-co-glycolic
[7] M. Alnaief, I. Smirnova, Effect of surface functionalization of silica aerogel on acid) microspheres: influence of manufacturing parameters and type of poly-
their adsorptive and release properties, Journal of Non-Crystalline Solids 356 mer on the release characteristics, Journal of Microencapsulation 16 (1999)
(2010) 1644–1649. 1–12.
[8] I. Smirnova, S. Suttiruengwong, W. Arlt, Feasibility study of hydrophilic and [22] H. Abramovic, The temperature dependence of dynamic viscosity for some
hydrophobic silica aerogels as drug delivery systems, Journal of Non-Crystalline vegetable oils, Acta Chimica Slovenica 45 (41) (1998) 1969–1978.
Solids 350 (2004) 54–60. [23] Y.Y. Yang, T.S. Chung, X.L. Bai, W.K. Chan, Effect of preparation conditions on
[9] I. Smirnova, S. Suttiruengwong, M. Seiler, W. Arlt, Dissolution rate morphology and release profiles of biodegradable polymeric microspheres con-
enhancement by adsorption of poorly soluble drugs on hydrophilic sil- taining protein fabricated by double-emulsion method, Chemical Engineering
ica aerogels, Pharmaceutical Development and Technology 9 (2004) 443– Science 55 (2000) 2223–2236.
452. [24] H. Rafati, A.G. Coombes, J. Adler, J. Holland, S.S. Davis, Protein-loaded
[10] I. Smirnova, J. Mamic, W. Arlt, Adsorption of drugs on silica aerogels, Langmuir poly(d,l-lactide-co-glycolide) microparticles for oral administration: formu-
19 (2003) 8521–8525. lation, structural and release characteristics, Journal of Controlled Release 43
[11] I. Smirnova, W. Arlt, Synthesis of silica aerogels: Influence of the supercritical (1997) 89–102.
CO2 on the sol–gel process, Journal of Sol–Gel Science and Technology 28 (2003) [25] P. Johansen, H.P. Merkle, B. Gander, Technological considerations related to the
175–184. up-scaling of protein microencapsulation by spray-drying, European Journal of
[12] Y. Tunc, K. Ulubayram, Production of highly crosslinked microspheres by the Pharmaceutics and Biopharmaceutics 50 (2000) 413–417.
precipitation polymerization of 2-(diethylamino)ethyl methacrylate with two [26] J. Jung, M. Perrut, Particle design using supercritical fluids: literature and patent
or three functional crosslinkers, Journal of Applied Polymer Science 112 (2009) survey, Journal of Supercritical Fluids 20 (2001) 179–219.
532–540. [27] M. Moner-Girona, A. Roig, E. Molins, J. Llibre, Sol–gel route to direct formation
[13] K.J. Klabunde, R.M. Richards, Nanoscale Materials in Chemistry, second ed. (rev. of silica aerogel microparticles using supercritical solvents, Journal of Sol–Gel
and updated), Wiley, Hoboken, NJ, 2009. Science and Technology 26 (2003) 645–649.
[14] S.C. Gad, Pharmaceutical Manufacturing Handbook: Production and Processes, [28] T. Mateovic, B. Kriznar, M. Bogataj, A. Mrhar, The influence of stirring rate on
John Wiley & Sons, New Jersey, 2008. biopharmaceutical properties of Eudragit RS microspheres, Journal of Microen-
[15] S. Freitas, H.P. Merkle, B. Gander, Microencapsulation by solvent extrac- capsulation 19 (2002) 29–36.
tion/evaporation: reviewing the state of the art of microsphere preparation [29] G.S. Banker, C.T. Rhodes (Eds.), Modern Pharmaceutics, fourth (rev. and
process technology, Journal of Controlled Release 102 (2005) 313–332. expanded. ed.), Marcel Dekker, New York, NY, 2002.

You might also like