ArticlePDF Available

Synthesis of hollow polymeric nanoparticles for protein delivery via inverse miniemulsion periphery RAFT polymerization

Authors:

Abstract and Figures

Hollow polymeric nanoparticles with a hydrophilic liquid core have been synthesized in a one-pot approach via a novel inverse miniemulsion periphery RAFT polymerization process. Successful encapsulation and release of a model protein is reported as a potential application.
Content may be subject to copyright.
This journal is cThe Royal Society of Chemistry 2012 Chem. Commun., 2012, 48, 11103–11105 11103
Cite this:
Chem. Commun
., 2012, 48, 11103–11105
Synthesis of hollow polymeric nanoparticles for protein delivery via
inverse miniemulsion periphery RAFT polymerizationw
Robert H. Utama, Yi Guo, Per B. Zetterlund* and Martina H. Stenzel*
Received 23rd August 2012, Accepted 25th September 2012
DOI: 10.1039/c2cc36116g
Hollow polymeric nanoparticles with a hydrophilic liquid core
have been synthesized in a one-pot approach via a novel inverse
miniemulsion periphery RAFT polymerization process. Successful
encapsulation and release of a model protein is reported as a
potential application.
The administration of therapeutically active proteins has shown
great potential in the treatment of many diseases, including
cancer and diabetes. However, the delivery of proteins has been
extremely limited through the parenteral route due to the rapid
degradation of the protein. This has prompted the development
of suitable carriers, such as polymer conjugates,
1,2
lipids
3,4
and
nanoparticles,
5–7
that allow for a more robust delivery process.
Hollow polymeric nanoparticles offer several advantages in
comparison to other types of carriers, e.g. higher encapsulation
permassofpolymer,processversatility,
8,9
and (compared to
many hydrogels) their relatively small size, which allows the
penetration into cells.
10
The synthesis of protein-loaded poly-
meric nanoparticles is often challenged by denaturation of the
protein.
11,12
Miniemulsion polymerization is a technique where-
by polymerization is conducted within 50–500 nm droplets
dispersed in a continuous phase (usually water).
13,14
Synthesis
of hollow nanoparticles with a liquid hydrophobic
15–18
or a
hydrophilic
19–21
core utilising radical polymerization in minie-
mulsion has been reported. However, this process is deemed
inappropriate for protein encapsulation because the presence of
radicals can potentially cause protein denaturation.
22
Further-
more, limitations of the aforementioned process include: (i) the
formation of some solid or collapsed nanoparticles; (ii) the use of
non-biocompatible surfactants which prompt the necessity for
further purification steps; (iii) the dependency of the core size on
the polymerization due to the inward growth of the shell and (vi)
limited possibilities for further surface modification, thereby
reducing the versatility of the process and of the synthesized
nanoparticles. All of the above limitations can be overcome by
using the novel synthetic route reported herein.
The term miniemulsion periphery polymerization (MEPP) was
first coined by Wang and co-workers,
23–25
who reported the
synthesis of hollow nanoparticles by essentially crosslinking poly-
meric surfactant species on the outer periphery of oil droplets by
using metal coordination polymerization. We have built on this
concept, and have developed a technique for synthesis of hollow
nanoparticles with a hydrophilic interior by combining the inverse
miniemulsion concept with surface-initiated RAFT polymerization.
An inverse miniemulsion system was specifically chosen to enable
synthesis of hollow nanoparticles with a hydrophilic core suitable
for protein delivery. RAFT crosslinking polymerization is
employed to provide control and functionalization of the nano-
shell. The crosslinking polymerization process, which takes place at
the periphery of the droplets in the oil phase, will ensure the
formation of stable hollow nano-spheres, with the proteins being
encapsulated within the reaction free aqueous interior. As depicted
in Scheme 1, protein encapsulation and production of the nano-
particles are achieved in one-pot via this novel approach, which is
termed inverse miniemulsion periphery RAFT polymerization
(IMEPP).
Scheme 1 Schematic description of the synthesis of hollow nanoparticles
via inverse miniemulsion periphery RAFT polymerization (IMEPP) of
methyl methacrylate/ethylene glycol dimethacrylate using a poly(HPMA)-
b-poly(MMA) macroRAFT agent at 60 1C.
Centre for Advanced Macromolecular Design (CAMD), School of
Chemical Engineering, The University of New South Wales, Sydney,
NSW 2052, Australia. E-mail: p.zetterlund@unsw.edu.au,
m.stenzel@unsw.edu.au; Fax: +61-2-93856250;
Tel: +61-2-93854344
wElectronic supplementary information (ESI) available: Syntheses,
IMEPP processes, TEM images, GPC, DLS, fluorescence and UV-Vis
characterisations. See DOI: 10.1039/c2cc36116g
ChemComm Dynamic Article Links
www.rsc.org/chemcomm COMMUNICATION
Downloaded by UNSW Library on 17 October 2012
Published on 26 September 2012 on http://pubs.rsc.org | doi:10.1039/C2CC36116G
View Online
/ Journal Homepage
/ Table of Contents for this issue
11104 Chem. Commun., 2012, 48, 11103–11105 This journal is cThe Royal Society of Chemistry 2012
A typical IMEPP recipe comprised toluene (a good solvent for
the poly(MMA) block and non-solvent for poly(HPMA)) as the
continuous phase, water as the dispersed phase, an amphiphilic
block-copolymer fulfilling the dual roles of a colloidal stabiliser
and a macroRAFT agent, and sodium chloride as a lipophobe to
minimize Ostwald ripening. For the subsequent polymerization,
methyl methacrylate (MMA), ethylene glycol dimethacrylate
(EGDMA) and AIBN were used as a monomer, a cross-linker
and an initiator, respectively. Poly(N-(2-hydroxylpropyl meth-
acrylamide)-b-poly(methyl methacrylate)) (polyHPMA
19
-b-poly-
MMA
79
), synthesized via RAFT polymerization
26
(see ESIw),
was used as a stabiliser/macroRAFT agent. Note that the
hydrophilic–lipophilic balance (HLB) value of the stabiliser
(4.9) was, calculated using Griffin’s method, well within the
acceptable range for an inverse miniemulsion application.
27,28
The initial inverse miniemulsion was prepared by dissolving
the stabiliser, monomer, cross-linker and initiator in the organic
phase and the lipophobe in the aqueous phase. The combined
mixture was then ultrasonicated, followed by polymerization in
sealed glass ampoules under vacuum at 60 1C. The size and
stability of the droplets (prior to polymerization) were initially
monitored using varying amounts of dispersed phase, stabiliser,
lipophobe, and ultrasonication period. The combinations that
led to a suitable particle size and sufficient stability for the desired
application were used in all further experiments (Table 1).
Using the conditions listed in Table 1, an overall monomer
conversion of 51% was reached in 7 h. The polymerization was
accompanied by no changes in turbidity or viscosity, implying the
preservation of the original miniemulsion characteristics. However,
monomer conversions significantly beyond 51% led to a marked
increase in viscosity, consistent with inter-particle crosslinking.
The droplet/particle size distributions were monitored by
DLS, revealing similar monomodal distributions before and
after polymerization with number-average diameters (d
n
)of
170 and 220 nm, respectively, consistent with a miniemulsion
mechanism whereby the initial droplets are converted to polymer
particles (Fig. S2, ESIw). The high degree of preservation of droplet
identity further confirms the absence of significant inter-particle
crosslinking. Overall, the results also showed the efficacy of the
chosen block-copolymers in stabilising droplets and controlling
the subsequent RAFT polymerization.
TEM analysis of the polymerized miniemulsion (diluted
with toluene) revealed hollow nanoparticles with an average
diameter in the range of 190–210 nm, perfectly in line with the
DLS analysis (Fig. 1). The well-defined spherical morphology
provided strong evidence that the shell remained intact during
the post-treatment and drying. The relatively narrow particle size
distribution obtained by DLS analysis was reflected in the TEM
images by the uniformity of the particles in terms of their size and
shape (Fig. S4, ESIw). The hollow core was clearly visualized,
with the methacrylate-based shell exhibiting a lighter contrast
than the background. Based on the TEM images, the thickness of
the shell was found to be B20 nm. Theoretical calculations of the
shell thickness resulted in values in the range of 4–24 nm, in good
agreement with the TEM images (the minimum value obtained
by considering the amount and density of the reacted monomer
(including the PMMA ‘‘hairs’’), and the maximum value corres-
ponding to the total contour length of the chain extended
PMMA ‘‘hairs’’ (see ESIw).
In order to test the use of these hollow nanoparticles as potential
drug carriers, a model protein was incorporated into the inverse
miniemulsion recipe. Bovine serum albumin (BSA) was added to
the dispersed phase (2 wt%), while keeping all other parameters
constant. BSA can be denatured in various ways, including via
temperature and pH.
29–31
However, there has not been any
conclusive evidence that shows denaturation of BSA by ultra-
sonication. Therefore, the effects of ultrasonication on both the
activity and the structure of BSA were investigated.
The activity of BSA was tested by examining the catalytic
activity towards the hydrolysis of p-nitrophenyl ester.
31
The BSA
catalysed hydrolysis of p-nitrophenyl acetate was carried out for
17 h, using both ultrasonicated BSA and native BSA. Comparison
of the UV-Vis spectra of the two samples (Fig. S6, ESI;wsignal
intensity at 405 nm corresponding to the characteristic wavelength
of the product) revealed no differences, thus indicating the same
catalytic activity (Fig. S7, ESIw). The BSA structural integrity was
confirmed based on the tryptophan (Trp) fluorescence emission
profile.
32
The fluorescence emission profiles of both the native and
the sonicated BSA exhibited an emission maximum at 349 nm,
which corresponds to the intact Trp residues (Fig. S8, ESIw).
Subsequently, IMEPP in the presence of BSA was carried out.
BSA in the prescribed quantity was observed to have no
significant impact on the initial miniemulsion droplet size and
stability (Tables S1 and S2, ESIw). Consistent results with the
experiments without BSA were also obtained in the subsequent
polymerization, with an overall conversion of 22% achieved in
3 h. Measurement of d
n
before and after polymerization resulted
in 170 and 183 nm, respectively. The absence of any secondary
peaks in the particle size distributions (Fig. S3, ESIw) is consistent
with successful encapsulation.
TEM analysis showed that the nanoparticles exhibited
similar morphologies to the hollow nanoparticles without BSA.
Table 1 Optimized IMEPP recipe for synthesis of hollow polymeric
nanoparticles
Continuous phase Dispersed phase Quantity
Toluene 6.5 g
MacroRAFT
a
10 wt% relative to water
MMA [MMA] : [RAFT] = 200 : 1
EGDMA [EGDMA] : [RAFT] = 25 : 1
AIBN [AIBN] : [RAFT] = 0.5 : 1
DI water 10 wt% relative to toluene
NaCl 2 wt% relative to water
BSA
b
2 wt% relative to water
a
PolyHPMA
19
-b-polyMMA
79
,M
n,theo
=10600gmol
1
,PDI=1.26.
b
Bovine serum albumin.
Fig. 1 TEM images of hollow nanoparticles synthesized via IMEPP
of MMA and EGDMA.
Downloaded by UNSW Library on 17 October 2012
Published on 26 September 2012 on http://pubs.rsc.org | doi:10.1039/C2CC36116G
View Online
This journal is cThe Royal Society of Chemistry 2012 Chem. Commun., 2012, 48, 11103–11105 11105
The clearly visible additional dark features correspond to
encapsulated BSA (Fig. 2; 80% loaded spheres calculated
from Fig. S5, ESIw). Diameter measurement based on the
TEM images gave a result of 170 nm with a shell thickness of
B12 nm, which was again in agreement with the theoretical
minimum and maximum thicknesses of 3 and 16 nm, respectively.
The thinner shell obtained at lower conversion confirms that
the shell thickness is controlled by the monomer conversion.
Subsequent experiments also confirmed the possibility of higher
protein loadings. Protein fractions of up to 5 wt% were incorpo-
ratedwithnonoticeableaggregationandchangesinsizecontrol
and stability of the droplets (confirmed visually and by DLS).
To confirm the integrity of the nano-shells, purified nano-
particles were re-dispersed in 1,4-dioxane, a common solvent
for both the HPMA and the MMA block. Without cross-
linking, the hollow nanoparticles would disappear, resulting
in unimolecularly dissolved block copolymers. DLS analysis of
the re-dispersed nanoparticles showed the presence of stable
particles comparable with the particle size obtained from the
original (prior to re-dispersion) miniemulsion. In addition, the
absence of any precipitate confirmed the protection provided by
the shell on the encapsulated BSA. Consistent results obtained
from the DLS and TEM analyses showed the efficiency and
convenience of the IMEPP process in synthesising well-defined,
protein-loaded polymeric nanoparticles (Fig. 2).
To test the hollow particles as potential carriers for therapeutic
agents, the release of BSA was investigated during continuous
mixing of the protein-loaded particles in PBS-buffer solution.
HPLC analysis of the filtrate of the solution revealed that 41%
of the originally encapsulated BSA was released in 3 days. This
confirmed that the extent of crosslinking of the hollow particles
was sufficient to ensure structural stability, yet the crosslink density
was low enough to allow diffusion of BSA through the shell.
Subsequent fluorescence analysis of the encapsulated BSA revealed
a Trp emission peak at 355 nm. Comparison of this peak with that
of the native (355 nm) and denatured BSA (361 nm), together with
the very minor difference in catalytic activity to that of the native
BSA (87% of the original activity), confirmed the preservation of
the BSA structure, despite exposure to ultrasonication at elevated
temperature (Fig. S10 and S11, ESIw).
In summary, a convenient one-pot synthetic route to synthesize
hollow polymeric nanoparticles having a hydrophilic aqueous
core, suitable for protein delivery, has been developed. The
method is based on inverse miniemulsion periphery polymeriza-
tion (IMEPP) using a block-copolymer as a macroRAFT agent
and a stabilizer in the absence of a conventional surfactant.
Successful encapsulation and release of BSA was demonstrated,
revealing no detrimental denaturation of BSA during particle
synthesis. Further work to improve the stability of the nano-
particles against aggregation in aqueous media is underway.
Notes and references
1 S. Kim, J.-H. Kim, O. Jeon, I. C. Kwon and K. Park, Eur. J.
Pharm. Biopharm., 2009, 71, 420–430.
2 H. Maeda, G. Y. Bharate and J. Daruwalla, Eur. J. Pharm.
Biopharm., 2009, 71, 409–419.
3 V. Balasubramanian, O. Onaca, R. Enea, D. W. Hughes and
C. G. Palivan, Expert Opin. Drug Delivery, 2009, 7, 63–78.
4 M. L. Tan, P. F. M. Choong and C. R. Dass, Peptides, 2010, 31,
184–193.
5 A. Kumari, S. K. Yadav and S. C. Yadav, Colloids Surf., B, 2010,
75, 1–18.
6 B. Mishra, B. B. Patel and S. Tiwari, Nanomed.: Nanotechnol.,
Biol. Med., 2010, 6, 9–24.
7 R. C. Mundargi, V. R. Babu, V. Rangaswamy, P. Patel and
T. M. Aminabhavi, J. Controlled Release, 2008, 125, 193–209.
8 Z. Sun and Y. Luo, Soft Matter, 2011, 7, 871–875.
9 P. B. Zetterlund, Y. Kagawa and M. Okubo, Chem. Rev., 2008,
108, 3747–3794.
10 K. S. Soppimath, T. M. Aminabhavi, A. R. Kulkarni and
W. E. Rudzinski, J. Controlled Release, 2001, 70, 1–20.
11 P. Ahlin Grabnar and J. Kristl, J. Microencapsulation, 2011, 28,
323–335.
12 Y. Yeo and K. Park, Arch. Pharmacal Res., 2004, 27, 1–12.
13 J. M. Asua, Prog. Polym. Sci., 2002, 27, 1283–1346.
14 K. Landfester, Angew. Chem., Int. Ed., 2009, 48, 4488–4507.
15 F.Lu,Y.LuoandB.Li,Macromol. Rapid Commun., 2007, 28, 868–874.
16 E. T. A. van den Dungen and B. Klumperman, J. Polym. Sci.,
Part A1, 2010, 48, 5215–5230.
17 A. J. P. van Zyl, R. F. P. Bosch, J. B. McLeary, R. D. Sanderson
and B. Klumperman, Polymer, 2005, 46, 3607–3615.
18 P. B. Zetterlund, Y. Saka and M. Okubo, Macromol. Chem. Phys.,
2009, 210, 140–149.
19 E.-M. Rosenbauer, K. Landfester and A. Musyanovych, Langmuir,
2009, 25, 12084–12091.
20 Y. Wang, G. Jiang, M. Zhang, L. Wang, R. Wang and X. Sun, Soft
Matter, 2011, 7, 5348–5352.
21 F. Lu, Y. Luo, B. Li, Q. Zhao and F. J. Schork, Macromolecules,
2009, 43, 568–571.
22 C.-C. Lin, S. M. Sawicki and A. T. Metters, Biomacromolecules,
2007, 9, 75–83.
23 S. Ye, Y. Liu, S. Chen, S. Liang, R. McHale, N. Ghasdian, Y. Lu
and X. Wang, Chem. Commun., 2011, 47, 6831–6833.
24 R. McHale, N. Ghasdian, Y. Liu, M. B. Ward, N. S. Hondow,
H. Wang, Y. Miao, R. Brydson and X. Wang, Chem. Commun.,
2010, 46, 4574–4576.
25 G. Liang, J. Xu and X. Wang, J. Am. Chem. Soc.,2009,131, 5378–5379.
26 A. Gregory and M. H. Stenzel, Prog. Polym. Sci., 2012, 37, 38–105.
27 R. C. Pasquali, M. P. Taurozzi and C. Bregni, Int. J. Pharm., 2008,
356, 44–51.
28 K. Shinoda and H. Saito, J. Colloid Interface Sci., 1969, 30, 258–263.
29 J. T. Tildon and J. W. Ogilvie, J. Biol. Chem.,1972,247, 1265–1271.
30 Y. Sakurai, S.-F. Ma, H. Watanabe, N. Yamaotsu, S. Hirono,
Y. Kurono, U. Kragh-Hansen and M. Otagiri, Pharm. Res., 2004,
21, 285–292.
31 J. Co
´rdova, J. D. Ryan, B. B. Boonyaratanakornkit and
D. S. Clark, Enzyme Microb. Technol., 2008, 42, 278–283.
32 Y. Moriyama, D. Ohta, K. Hachiya, Y. Mitsui and K. Takeda,
J. Protein Chem., 1996, 15, 265–272.
Fig. 2 TEM images of protein-loaded nanoparticles (dark spots are
encapsulated BSA) and number-based droplet/particle size distributions
before and after IMEPP (in toluene), and after re-dispersion in 1,4-dioxane.
Downloaded by UNSW Library on 17 October 2012
Published on 26 September 2012 on http://pubs.rsc.org | doi:10.1039/C2CC36116G
View Online
... This process involves triggering polymerization at the periphery of the dispersed drop with an amphiphilic macroinitiator, suitable to stabilize the emulsion. For example, the chain extension of poly(N-(2-hydroxylpropyl methacrylamide)-b-poly(methyl methacrylate) (PHPMAm-co-MMA)) macro-CTA with MMA and the crosslinking agent EGDMA via reverse emulsion polymerization in a toluene/water mixture under ultrasonication was demonstrated (Figure 27).[131] The reaction was triggered at 60°C using AIBN as radical initiator in the presence of BSA in the aqueous phase. ...
Preprint
Full-text available
The encapsulation of biologically active ingredients (e.g., peptides, proteins, enzymes, drugs) into polymer particles is extensively used for drug delivery purposes. However, this strategy relies mainly on emulsification processes from preformed polymers, which leads to strong limitations such as low particle concentrations (typically a few wt %), poor active ingredient loadings, as well as a rather limited structural diversity of the polymers usually used. Conversely, polymerizations in dispersed media, which allow for the formation of scalable suspensions of (nano)particles during the polymerization process, have been advantageously used for the in situ encapsulation of active ingredients. In this review, the in situ encapsulation of active ingredients, such as peptides, proteins, enzymes or drugs, in polymer particles obtained by polymerization in dispersed media for potential biomedical applications, is covered. Their physical and chemical encapsulations were both considered as function of the polymerization technique used. Several polymerization and encapsulation parameters will be discussed in view of adjusting the drug loading and the encapsulation efficiency of the active agent considered.
... Polymer nanocapsules with an aqueous core, suitable for protein delivery, were prepared by so-called inverse miniemulsion periphery RAFT polymerization (IMEPP) [44]. The authors used a poly (N-(2-hydroxylpropyl methacrylamide)-b-poly(methyl methacrylate)) (polyHPMA 19 -b-polyMMA 79 ) block copolymer as a stabilizer and mac-roRAFT agent, avoiding any conventional surfactant. ...
Article
Nanoemulsions are kinetically stabilized emulsions with droplet sizes in the nanometer scale. These nanodroplets are able to confine spaces in which reactions of polymerization or precipitation can take place, leading to the formation of particles and capsules that can act as nanocarriers for biomedical applications. This review discusses the different possibilities of using nanoemulsions for preparing biomedical nanocarriers. According to the chemical nature, nanocarriers prepared in nanoemulsions are classified in polymeric, inorganic, or hybrid. The main synthetic strategies for each type are revised, including miniemulsion polymerization, nanoemulsion–solvent evaporation, spontaneous emulsification, sol–gel processes, and combination of different techniques to form multicomponent materials.
Article
Stimuli-responsive core–shell microgels are of significant interest because of their fascinating applications due to the different swelling/shrinkage properties of their core and shell networks. Because such stimuli-responsive core–shell microgels are...
Article
pH-responsive capsule particles show promise for various applications, such as self-healing materials, micro/nanoreactors, and drug delivery systems. Herein, carboxy-functionalized capsule polymer particles possessing neutral-alkaline pH responsive controlled release capability were newly fabricated by interfacial photocrosslinking of spherical photoreactive polymer [poly(2-carboxyethyl acrylate-co-2-cinnamoylethyl methacrylate): P(CEA-CEMA)] particles and a subsequent encapsulation process. Using P(CEA-CEMA) particles, the shell-crosslinked hollow polymer particles were fabricated by the particulate interfacial photocrosslinking procedure. Furthermore, the encapsulation of sulforhodamine B as a model dye into the hollow particles was also performed. Under acidic pH conditions, encapsulated molecules were stably retained in the P(CEA-CEMA) capsules with negligible release of sulforhodamine B. However, the encapsulated sulforhodamine B was gradually or drastically released from the capsule particles under neutral or basic conditions, respectively, indicating that the neutral-alkaline pH responsive controlled release from the capsules was successfully achieved by regulating the release kinetics. These results demonstrate that the fabrication routes of hollow and capsule particles based on particulate interfacial photocrosslinking can be successfully applied to carboxy-functionalized photoreactive polymer particles, and the capsule polymer particles possessing pH-responsive release properties under neutral-basic conditions were successfully fabricated.
Article
Hollow and capsule polymer particles have great potential for use in various applications such as coatings, microreactors, and drug delivery systems. The strategy of the interfacial photocrosslinking of spherical polymer particles is promising for creating a wide range of functional hollow and capsule particles. Herein, polystyrene-based particles possessing nucleobases in polymer side chains were prepared and nucleobase groups were applied to the interfacial photocrosslinking as photoreactive groups for the first time. In this interfacial photocrosslinking, [2π + 2π] photoinduced dimerization between nucleobases proceeds only at the particle/water interface under photoirradiation, and subsequent removal of non-crosslinked polymers results in the formation of shell-crosslinked hollow particles. 4-Vinylbenzyl thymine (VBT) was prepared as a styrene-based photoreactive monomer, and the [2π + 2π] photoreactive dimerization property of poly(VBT-co-styrene) was confirmed, leading to the fabrication of shell-crosslinked hollow polymer particles. Owing to the strong hydrogen bonding property of thymine groups, the washing process for removing non-crosslinked polymers from photoirradiated particles was optimized. Furthermore, the encapsulation of a model dye was successfully performed.
Article
We propose a synthesis method for hollow copolymer nanoparticles, in which the size is controllable by the wettability of the materials designed by relative energy difference (RED). We investigated the influence of cross-linkers in RED and the hollow polymer nanoparticle synthesis. The size of the nanoparticles was characterized by scanning electron microscopy and transmission electron microscopy images. The diameter size of the hollow copolymer (styrene-co-methyl methacrylate) changes from 400 to 141 nm and the average core-vacancy sizes changes from 330 to 71 nm as increasing the feed ratio of the cross-linker, divinyl benzene, from 0.07 to 0.43. Cross-linkers in polymerization precipitates a polymerization reaction to produce seed copolymer particles quickly. The seed copolymer is a more transferrable medium through the surfactants across emulsion droplets and inhibits emulsion growth by unstable concentration variations of seed copolymers in emulsions. Therefore, Ostwald ripening was reduced by a higher feeding ratio of the cross-linker in the copolymer, which tends to produce smaller sized hollow nanoparticles.
Article
Full-text available
The results of a kinetic investigation of the reactions of bovine serum mercaptalbumin with p-nitrophenyl acetate are described. At pH 6.3 to 7.8 these reactions are biphasic and consist of (a) an initial phase during which the release of p-nitrophenol results from a monoacetylation of the albumin and from an albumin-catalyzed hydrolysis of the p-nitrophenylacetate and (b) a second phase during which the release of p-nitrophenol results only from the albumin-catalyzed hydrolysis of the ester. Second order rate constants for both the acetylation reaction and albumin-catalyzed hydrolytic reaction are reported. Results are also presented which indicate that the acetylation site and the catalytic site on the albumin are separate sites and that the conformational integrity of the protein is necessary for its catalytic activity. Analysis of albumin which has reacted with p-nitrophenyl acetate indicate that two tyrosyl hydroxyl groups on the protein are rapidly acetylated by the ester. One of these tyrosyl groups is at the acetylation site of the protein and the other is postulated to be adjacent to the catalytic site. A model compatible with the kinetic and analytical data is proposed in which the acetyl group of an acetyl protein intermediate in the catalytic reaction is transferred to a tyrosyl residue adjacent to the catalytic site.
Article
Full-text available
Unlabelled: Colloidal nanocarriers, in their various forms, have the possibility of providing endless opportunities in the area of drug delivery. The current communication embodies an in-depth discussion of colloidal nanocarriers with respect to formulation aspects, types, and site-specific drug targeting using various forms of colloidal nanocarriers with special insights to the field of oncology. Specialized nanotechnological approaches like quantum dots, dendrimers, integrins, monoclonal antibodies, and so forth, which have been extensively researched for targeted delivery of therapeutic and diagnostic agents, are also discussed. Nanotechnological patents, issued by the U.S. Patent and Trademark Office in the area of drug delivery, are also included in this review to emphasize the importance of nanotechnology in the current research scenario. From the clinical editor: Colloidal nanocarriers provide almost endless opportunities in the area of drug delivery. While the review mainly addresses potential oncological applications, similar approaches may be applicable in other conditions with a requirement for targeted drug delivery. Technologies including quantum dots, dendrimers, integrins, monoclonal antibodies are discussed, along with US-based patents related to these methods.
Article
Miniemulsion polymerization has recently exploded in terms of publications and the development of a wide range of useful polymer materials only accessible through this polymerization technique. In the first part of this article, the fundamental aspects involved in the preparation and polymerization of monomer miniemulsions are reviewed. The second part deals with the application of miniemulsion polymerization for the production of high solids low viscosity latexes, dispersion of polymers of well defined microstructure through controlled radical polymerization, use of catalytic polymerization in aqueous media, encapsulation of inorganic solids, incorporation of hydrophobic monomers, preparation of hybrid polymer particles, implementation of anionic and step polymerization in aqueous dispersed media, and process intensification by using continuous reactors.
Article
A series of novel shell cross-linked nanocapsules were facilely fabricated in an inverse miniemulsion reversible addition–fragmentation chain transfer (RAFT) system, in which poly(2-(dimethylamino)ethyl methacrylate) (PDMAEMA) acted as a stabilizer and macroinitiator. The as-prepared nanocapsules with PDMAEMA cores and poly(methylacrylic acid) (PMAA) shells, had a size of 220 nm and can be used as a carrier for the encapsulation of hydrophilic components such as inorganic salts.
Article
The non-collapsed hollow polymeric nanoparticles with shell thicknesses in the order of 10 nm are prepared by interfacially confined reversible addition fragmentation transfer (RAFT) miniemulsion polymerization. The void fraction and average diameter of the hollow polymeric nanoparticles could be largely tuned up to 0.58 and from 68 nm to 180 nm, respectively. The non-collapsed hollow polymeric nanoparticles could be a building block for nanoporous materials, which hold promise in many fields such as anti-reflection coatings, ultra-thermal insulation materials, catalysis and sensors.
Article
The synthesis of temperature responsive nanoencapsulation of inorganic salts through reversible addition-fragmentation transfer (RAFT) interfacial confined inverse miniemulsion polymerization was reported by using a poly(ethylene oxide)-MacroRAFT (PEO-RAFT) agent. It was found that the nanocapsules (NC) could be applied in medical diagnose or pharmaceuticals therapy because of the thermo-sensitive polymer shell that can release the core contents of the NCs around the body temperature. It was also found that as the polymerization was a radical living/controlled polymerization confined in the interface, the polymer shell could be functionally modified via post reaction. The results show that the strength of the polymer shell can be increased by cross-linking polymerization to obtain rigid hollow NCs.
Article
Nanocapsules (below 100 nm) with liq. cores and molar mass controlled polystyrene shells were synthesized by an in situ miniemulsion polymn. reaction in the presence of a RAFT (reversible addn. fragmentation chain transfer) agent. The formation of structured particles with the targeted core/shell morphol., i.e. liq. cores and polymeric shells, is extremely dependent on the type of RAFT agent used in conjunction with the type of initiator used. Different RAFT agents lead to different polymn. rates, thus resulting in different chain lengths as a function of time. This will influence viscosity and consequently chain mobility and can therefore cause a deviation from the desired morphol. The type of initiator used influences the surface activity of entering oligomers and is therefore also an important factor in obtaining the correct structure. Results showed that a RAFT agent that causes no rate retardation (Ph 2-Pr Ph dithioacetate, PPPDTA), used in conjunction with a surface active initiating species
Article
Network formation in the nitroxide‐mediated cross‐linking copolymerization of styrene and divinylbenzene (3 or 8.2 mol‐% relative to total monomer) using the nitroxide 2,2,6,6‐tetramethylpiperidinyl‐1‐oxy (TEMPO) at 125 °C can proceed markedly differently in aqueous miniemulsion compared to the corresponding solution polymerization depending on the organic‐phase composition. When the organic phase comprises 54 vol‐% of the hydrophobe tetradecane, gelation occurs at much lower conversion in miniemulsion than in solution, and at significantly lower conversion than predicted by Flory–Stockmayer gelation theory. This is proposed to be a result of an effect of the oil–water interface, whereby the concentration of polymer with pendant unsaturations is higher near the interface than in the particle interior. magnified image
Article
Miniemulsion polymerization with an amphiphilic poly(acrylic acid)‐ block ‐polystyrene reversible addition–fragmentation chain transfer agent as a surfactant and polymerization mediator is used to synthesize highly uniform nanocapsules. The nanocapsules with uniform structures, which include particle size, shell thickness, and shape symmetry, could be achieved by the post‐addition of a small amount of sodium dodecyl sulfate. Although the solid particles seem unavoidable, the ‘pure’ uniform core–shell structures are easily collected by centrifugation. magnified image
Article
This study describes the synthesis of well-defined nanocapsules via the miniemulsion technique. Pentaerythritol tetrakis(3-mercaptopropionate) (TetraThiol) or 1,6-hexanediol di(endo, exo-norborn-2-ene-5-carboxylate) (DiNorbornene) is used as the oil phase. TetraThiol is encapsulated via the miniemulsion technique without polymerization, as this monomer would simultaneously act as a chain-transfer agent, and DiNorbornene is encapsulated via miniemulsion polymerization of styrene. Various styrene-maleic anhydride (PSMA) copolymers and poly(styrene-maleic anhydride)-block-polystyrene (PSMA-b-PS) block copolymers were used as surfactant for the synthesis of well-defined nanocapsules with TetraThiol as the core material. The nanocapsules had a diameter of 150–350 nm and the particle size distribution was narrow. The use of PSMA-b-PS block copolymers as surfactant in combination with post-addition of formaldehyde provided improved stability to the nanocapsules. DiNorbornene was encapsulated via miniemulsion polymerization of styrene, and a stable latex with a bimodal particle size distribution was obtained. The distribution of small particles had a size of 60 nm and the distribution of large particles had a size of 150 nm. © 2010 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem, 2010