ArticlePDF Available

Synthesis and Physicochemical Properties of Cationic Microgels Based on Poly(N-isopropylmethacrylamide)

Authors:

Abstract

Surfactant-free, radical precipitation co-polymerization of N-isopropylmethacrylamide (NIPMAm) and the cationic co-monomer N-(3-aminopropyl) methacrylamide hydrochloride (APMH) was carried out to prepare microgels functionalized with primary amines. The morphology and hydrodynamic diameter of the microgels were characterized by atomic force microscopy (AFM) and photon correlation spectroscopy (PCS), with the effect of NaCl concentration and initiator type on the microgel size and yield being investigated. When a V50-initiated reaction was carried out in pure water, relatively small microgels (~160 nm diameter) were obtained in low yield (~20%). However, both the yield and size increased if the reaction was carried out in saline or by using APS as initiator instead of V50. Stable amine-laden microgels in the range from 160 nm to 950 nm in diameter with narrow size distributions were thus produced using reaction media with controlled salinity. Microgel swelling and electrophoretic mobility values as a function of pH, ionic strength and temperature were also studied, illustrating the presence of cationic sidechains and their influence on microgel properties. Finally, the availability of the primary amine groups for post-polymerization modification was confirmed via modification with fluorescein-NHS.
Synthesis and Physicochemical Properties of Cationic Microgels
Based on Poly(N-isopropylmethacrylamide)
Xiaobo Hua,b, Zhen Tongb, and L. Andrew Lyona,*
a School of Chemistry & Biochemistry and the Petit Institute for Bioengineering & Bioscience,
Georgia Institute of Technology, Atlanta, GA 30332, USA
b Research Institute of Materials Science, South China University of Technology, Guangzhou
510640, P.R. China
Abstract
Surfactant-free, radical precipitation co-polymerization of N-isopropylmethacrylamide (NIPMAm)
and the cationic co-monomer N-(3-aminopropyl) methacrylamide hydrochloride (APMH) was
carried out to prepare microgels functionalized with primary amines. The morphology and
hydrodynamic diameter of the microgels were characterized by atomic force microscopy (AFM)
and photon correlation spectroscopy (PCS), with the effect of NaCl concentration and initiator
type on the microgel size and yield being investigated. When a V50-initiated reaction was carried
out in pure water, relatively small microgels (~160 nm diameter) were obtained in low yield
(~20%). However, both the yield and size increased if the reaction was carried out in saline or by
using APS as initiator instead of V50. Stable amine-laden microgels in the range from 160 nm to
950 nm in diameter with narrow size distributions were thus produced using reaction media with
controlled salinity. Microgel swelling and electrophoretic mobility values as a function of pH,
ionic strength and temperature were also studied, illustrating the presence of cationic sidechains
and their influence on microgel properties. Finally, the availability of the primary amine groups
for post-polymerization modification was confirmed via modification with fluorescein-NHS.
Keywords
Controlled salinity; Microgel; N-isopropylmethacrylamide; cationic polymer
Introduction
Temperature-sensitive poly(N-isopropylacrylamide) (pNIPAm) microgels, which were first
prepared by emulsion polymerization in 1986 [1], have received a significant amount of
attention due to dramatic changes in physicochemical properties observed when temperature
is raised above their volume phase transition temperature (VPTT, ~31 °C). In recent years,
such microgels have been used in drug delivery [2–4], biosensors [5–7], tissue regeneration
[8,9], and chemical separations [10]. In many of these applications, the introduction of
specific chemical functionalities is desired, such that the polymer can be elaborated upon
following synthesis. Toward that end, functionalities such as carboxyl [11], azido [12] and
glycidyl [13] groups have been copolymerized into pNIPAm microgels. However, the
primary amine, a reactive group that is particularly useful for bioconjugation [14], has been
far less studied in microgel synthesis. The utility of this group is associated with the
potential for binding with anionic moieties on oligonucleotides, proteins, or enzymes [14–
*To whom correspondence should be addressed. lyon@gatech.edu.
NIH Public Access
Author Manuscript
Colloid Polym Sci. Author manuscript; available in PMC 2011 March 16.
Published in final edited form as:
Colloid Polym Sci
. 2010 December 4; 289(3): 333–339. doi:10.1007/s00396-010-2347-y.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
16], as well as its use in the covalent attachment to molecules carrying isothiocyanates or
succinimidyl esters.
Given the utility of amine chemoligation sites, the incorporation of amines into microgels
has been studied in a few cases. Xu et al. prepared primary amine functionalized pNIPAm
microgels by copolymerization of N-isopropylacrylamide (NIPAm) with N-vinylformamide
(NVF) by a semi-batch route, followed by conversion of the formamide moieties to a
primary amine by acid hydrolysis [17]. Unfortunately, this method requires multiple steps
and very strong acid hydrolysis conditions, thereby limiting its broad application. Leung et
al. prepared microgels by graft copolymerization of NIPAm from either PEI or chitosan,
which results in core-shell structures with the amines mainly becoming shell-localized [18].
This approach produces a very thin polymer shell, with the amine concentration being
limited due to the stabilization effect of polymer during the synthesis. Meunier et al.
prepared amine-laden pNIPAm microgels by using 2,2-azobis (2-amidinopropane)
dihydrochloride (V50) as initiator and 2-aminoethylmethacrylate hydrochloride (AEMH) as
a functional monomer [19]. However, the ester group in the AEMH monomer can be easily
hydrolyzed to carboxyl during the synthesis because of the acid solution and high
temperature, which can cause these microgels to be unstable. In this paper, we discuss
alternate routes for the preparation of amine functionalized microgels based on poly(N-
isopropylmethacrylamide) (pNIPMAm). The use of pNIPMAm, which has a volume phase
transition temperature of ~43 °C was motivated by our recent use of this polymer in drug
delivery applications [4]. Thus, we are very interested in broadening the utility of those
particular microgels by increasing the range of functionalities that can be efficiently
incorporated.
Experimental
Materials
All materials were purchased from Sigma-Aldrich unless otherwise noted. The monomer N-
isopropylmethacrylamide (NIPMAm) was recrystallized from n-hexane (J. T. Baker).
Comonomer N-(3-aminopropyl)methacrylamide hydrochloride (APMH, Polysciences Inc.),
cross-linker N,N-methylene(bisacrylamide) (BIS), cationic initiator 2,2-azobis(2-
amidinopropane) dihydrochloride (V50) and anionic initiator ammonium persulfate (APS)
were all used as received. NHS-Fluorescein (Thermo Sci.) and NaCl were used as received.
Pure water was produced by deionization to a resistance of 18 MΩ·cm (Barnstead E-Pure
system), followed by filtration through a 0.2-μm filter to remove particulate matter.
Microgel Synthesis
Microgels were synthesized by surfactant-free, radical precipitation co-polymerization using
methods similar to those previously reported [20]. For all reactions, the molar composition
was 89% NIPMAm, 9% APMH, and 2% BIS, with a total monomer concentration of 140
mM (Table 1). In a typical synthesis, 50 mL of a filtered, aqueous solution of NIPMAm,
APMH, BIS and NaCl was added to the reaction flask, which was then heated to 70 °C. The
solution was purged with N2 gas with continuous stirring until the temperature remained
stable. The reaction was initiated by addition of 2 mL of a V50 or APS solution. The
reaction was allowed to continue for 20 h while being purged by N2 gas with constant
stirring. After synthesis, the solution was filtered through glass wool to remove any
coagulum and then centrifuged/redispersed several times for purification. After purification,
the microgel dispersion was lyophilized for at least 72 hours.
Hu et al. Page 2
Colloid Polym Sci. Author manuscript; available in PMC 2011 March 16.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Atomic Force Microscopy (AFM)
Microgels were imaged using an Asylum Research MFP-3D Instrument (Santa Barbara,
CA). Imaging was performed and processed using the MFP-3D software under the IgorPro
(WaveMetrics Inc., Lake Oswego, OR) environment. Non-contact mode aluminum-coated
silicon nitride cantilevers were purchased from NanoWorld (force constant = 42 N/m,
resonant frequency = 320 kHz). All images were taken in air under ambient conditions.
AFM samples were prepared by centrifugal deposition at a maximum rotor speed of 2250 ×
g for 5 min on glass coverslips [21]. Glass coverslips (22 mm × 22 mm) were cleaned using
an air plasma (Harrick Plasma, Ithaca, NY) for ~ 15 min. Each cleaned slide was
functionalized by exposure to a 1% solution of 3-aminopropyltrimethoxysilane (APTMS) in
absolute ethanol (200 proof) for ~3 h. The glass was then rinsed with ethanol and dried
under a gentle stream of N2. The amine-functionalized coverslips were further modified by
immersion in a 0.1 mg/mL poly(sodium styrenesulfonate) solution before the centrifugal
deposition of microgels. After microgel deposition, the sample was gently rinsed with
deionized water and dried under a gentle stream of N2.
Particle size characterization
Particle sizes were determined by photon correlation spectroscopy (PCS; Protein Solutions,
Inc.) using an instrument equipped with an integrated Peltier temperature control device,
which provides temperature accuracy within ±0.1 °C. The instrument collects scattering
light at 90° by a single-mode optical fiber coupled to an avalanche photodiode detector. The
samples were thermally equilibrated at each temperature for 30 min before each
measurement. The data presented are the averaged values of 20 measurements, with a 10 s
integration time for each measurement.
Zeta-potential Determination
The ζ-potential was measured with a Zetasizer Nano-ZS (Malvern). Before the
measurements, the 1 mL microgel dispersion in the designated buffer was thermally
equilibrated between parallel electrodes in cuvette for 10 min. The ζ-potential value was the
average of at least three successive measurements.
Fluorescence labeling and imaging
For conjugating the amine reactive fluorescein, 2.5 mg of amine microgels were dispersed in
5 mL of a pH 7.4 PBS buffer, and 1.6 mL of 1 mg/mL DMF solution of fluorescein-NHS
was added. The solution was stirred at 4 °C overnight and then purified by dialysis against
DI water for ~2 weeks with the water being changed twice per day, using Spectra-Pro
10,000 MW cut-off dialysis tubing (VWR). The dyed microgels were imaged by
fluorescence microscopy, which was conducted on an Olympus IX-70 inverted microscope
equipped with a high numerical aperture, oil immersion 100 × objective (NA = 1.30). The
excitation irradiation was a mercury lamp filtered by excitation band-pass filters of 450–490
nm. Images were captured using a color CCD camera (PixelFly, Cooke Corporation).
Results and Discussion
During the preparation of pNIPAm or pNIPMAm microgels, an anionic initiator APS or
cationic initiator V50 is typically degraded thermally to initiate the polymerization, which
results in anionic or cationic microgels, respectively [1,22]. During our experiments, initial
attempts to prepare poly(NIPMAm-co-APMH) microgels (μG) in pure water using a
cationic initiator (V50) resulted in microgels of a very small size with poor yield (Table 1).
A possible reason for the poor yield lies in unfavorable reactivity ratio between the two co-
monomers, which inhibits the formation of copolymer chains that can easily collapse at high
Hu et al. Page 3
Colloid Polym Sci. Author manuscript; available in PMC 2011 March 16.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
temperature [23]. Randomly positioned APMH along the copolymer chain creates
sufficiently strong Coulombic repulsion between the cationic groups such that thermo-
induced chain collapse is inhibited. A similar phenomenon has been described previously by
Meunier et al., where high AEMH concentrations accelerated the incorporation rate of BIS
during particle formation, making the capture of polymer formed later in the reaction
unfavorable [19]. As a result, many free polymer chains may be formed instead of collapsed
precursor particles, thereby reducing the yield of microgels, and limiting the subsequent
polymer size due to a decrease in the amount of incorporated polymer/particle. Fortunately,
the reactivity ratio of ionic monomers can be notably influenced by the ionic strength
[23,24]. We therefore investigated the influence of NaCl concentration on microgel
preparation.
From Table 1, we can see that the diameter of μG0 microgels in pH 3 buffer (10 mM) is
~715 nm, with a yield of ~45%. However, the particle diameter decreased to ~157 nm after
the addition of 9 mol% APMH, with the yield of cross-linked microgels dropping to 20%. In
the past, the presence of salt in alkylacrylamide emulsion polymerization media has been
avoided due to the propensity of deswollen particles to aggregate under high salinity
conditions [25]. However, in this study, our results show that increasing the salinity of the
reaction medium increases the microgel size along with a modest increase in microgel yield.
We hypothesize that the salt affects the polymerization in at least two ways. After the
addition of NaCl, the Coulombic repulsion between the positively charged APMH units
becomes sufficiently screened to allow a greater tendency for homopropagation, which will
result in polymers with a longer sequence of APMH units in the copolymer chains. The
block-like copolymer chains will undergo thermally-induced collapse more readily, as has
been observed for other block copolymers containing pNIPAm [26]. Consequently, more
polymer chains will become incorporated into the growing microgels instead of remaining in
solution as uncrosslinked “free” polymer chains. Additionally, it is likely that growing
microgels will undergo aggregation with the other precursor particles under these charge-
shielding conditions, thereby producing a larger average particle size. However, when the
NaCl concentration reached 200 mM, the reaction system became unstable and macroscopic
coagulation of the polymer was observed.
In addition to the impact of salinity on microgel formation, the influence of different
initiators was studied, where the anionic initiator APS was used to initiate the
polymerization instead of V50. Despite the obvious potential for forming zwitterionic
microgels when APS is used, no colloidal instability was observed during the syntheses
undertaken in this work. Particles produced under the conditions for μG5 had a diameter of
~500 nm with a yield of 60%, which is much higher than that obtained from the
corresponding V50 synthesis. When the APS concentration was increased to 2 mM, a
particle diameter of 740 nm and 65% yield was obtained (μG6). Whereas these results are
too preliminary to completely explain this phenomenon, we tentatively ascribe these results
to the presence of the oxygen-centered radical associated with APS, which is more effective
in abstracting a hydrogen atom from the carbon next to the amine group in APMH. This
chain transfer reaction will provide additional reactive sites, which will induce continuous
graft polymerization and self cross-linking [27,28], resulting in larger microgels and higher
polymer yields.
The ζ-potentials of the microgels in pure water are listed in Table 1; all microgels display a
positive ζ-potential indicating cationic character. For μG0 and μG1, the ζ-potentials are of
similarly low magnitude. For μG0, the positive charge is due to the amidine end groups
arising from the V50 initiator. The lack of a significant change in this value for μG1, which
was synthesized using 9 mol% APMH, suggests that the degree of APMH incorporation was
very poor. In contrast, there is a large increase in the ζ-potential value for microgels
Hu et al. Page 4
Colloid Polym Sci. Author manuscript; available in PMC 2011 March 16.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
synthesized with increasing salinity or with APS as the initiator, which is indicative of
surface charges imparted by more efficient incorporation of APMH. It is also interesting to
note that μG6 displays a slightly higher zeta-potential than the other samples, even though
this particle must also contain some anionic initiator residues. This, coupled with the
observation that these conditions display the highest yield, suggests that APMH
incorporation is maximized for μG6 relative to the other conditions.
Analysis of the microgels via AFM permits visualization of the particle morphology and
polydispersity estimation (Figure 1). It should be noted that in all cases, the strong microgel-
surface interactions, combined with the fact that the particles are dehydrated, result in
severely flattened particles. Of particular interest in these data is the observation that μG3
and μG4 do not appear to be purely spherical particles. Instead, these microgels appear
somewhat irregular and “bumpy”, as if formed via the aggregation of multiple smaller
particles. Such precursor particle aggregation would be predicted to occur more readily in
high salt. Alternatively, the addition of NaCl likely results in the formation of block-like
copolymers during the microgel synthesis [23,24], as described above. If this is indeed the
case, the resultant particles might contain relatively long polyelectrolyte chains at the
microgel surface. This shell-localized polymer will randomly spread on the substrate during
the drying process, resulting in a non-spherical particle appearance. Conversely, microgels
μG5 and μG6 appear spherical, which perhaps suggests a homogeneous particle architecture.
Line profiles taken from these images permit estimation of the dried microgel heights and
diameters. The height/diameter ratio provides a rough, qualitative estimate of microgels
stiffness [29]; this analysis suggests that the APS-initiated microgels (μG5 and μG6) are
somewhat stiffer than those prepared with V50 (Figure 1f), presumably due to increased
cross-linking brought about by the aforementioned chain transfer reactions.
In addition to microgel morphology, it is critical to know how the different synthetic
conditions impact the temperature and pH dependent swelling and charge characteristics.
The physicochemical properties of selected microgels are summarized in Figure 2. As
expected from their pNIPMAm content, all microgels display a volume phase transition
temperature (VPTT) between 40 and 50 °C (Figure 2a). Those lacking cationic co-monomer
(μG0) display a transition very close to the LCST of pure pNIPMAm (~44 °C), with the
VPTT of all other microgels being shifted to higher temperatures, as is typical for microgels
containing hydrophilic or charged groups [30,31]. The swelling ratio is also presumably
influenced by the presence of charged co-monomers (Figure 2b), as the swelling ratios of the
amine containing microgels are all smaller than that of μG0. Amine incorporation is also
evident in the influence of ionic strength on the measured microgel size at pH 3 (Figure 2c).
Below the VPTT, the diameter of μG3 is observed to decrease as the ionic strength increased
from 2 mM to 100 mM. Furthermore, the VPTT is shifted to lower temperatures with a
concomitant sharpening of the transition. These effects arise largely from charge screening,
competitive solvation, and osmotic effects [25,32,33]. By measuring the microgel swelling
ratio as a function of pH (defined as the ratio of the microgel diameter at a particular pH
relative to that at pH 11.5), we observe the expected result that the swelling ratio gradually
decreases at pH values higher than 7.4 and plateaus above the pKa (~10) of APMH (Figure
2d). The pH dependence of the ζ-potential closely mirrors the swelling data (Figure 2e),
where in the absence of APMH (μG0), no pH dependence is observed, but a strong pH effect
is observed in the presence of APMH (μG3 and μG6). For the APMH containing microgels,
the ζ-potential decreases as the pH increases. For example, the ζ-potential of μG6 suddenly
decreases with a pH change from 9.5 to 10.5, with this pH corresponding to the pKa of
primary amine group. However, for μG3, the ζ-potential gradually decreases once the pH
increases above 7.4. This difference in pKa between the polyelectrolyte and the parent
monomer is common due to the decreased propensity for deprotonation in the presence of
the neighboring ionic groups [34]. For example, poly(2-aminoethyl methacrylate), which has
Hu et al. Page 5
Colloid Polym Sci. Author manuscript; available in PMC 2011 March 16.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
a similar structure to pAPMH, has a pKa of approximately 7.6. This pKa is significantly
lower than that of its monomer (pKa =10) [35]. This result is therefore also suggests that the
addition of NaCl results in the formation of block-like copolymers. Due to the cationic
initiator fragments, the ζ-potential is always positive for the μG0 and μG3 microgels over
the pH range studied here. However, due to the neutralization of the amine sidechains and
the presence of anionic initiator groups, the ζ-potential of μG6 microgels is negative at pH
11.5. Finally, to illustrate that the amine groups are accessible to small molecule coupling
reactions, fluorescein-NHS was used as a coupling probe for microgels μG3 and μG6; a
representative image of μG6 is shown (Figure 2f). Studies investigating the accessibility of
those groups in a wider range of bioconjugation reactions are currently underway.
Conclusions
In this investigation, we have shown that primary amine containing pNIPMAm microgels
can be produced by a surfactant free radical precipitation copolymerization. By controlling
the reaction salinity or using an anionic initiator APS, low polydispersity, primary amine
containing pNIPMAm microgels can be obtained in high yield. The increase of particle size
and yield by the salt addition originates from the screening of Coulombic repulsion between
APMH units, which results in more favorable polymer incorporation. The resultant amine-
laden microgels show the expected swelling properties of thermoresponsive cationic
microgels as a function of temperature, pH and ionic strength, as well as reactivity in
standard amide bond-forming reactions. Studies of these microgels in the context of
biomolecule conjugation and thin film self-assembly are currently underway.
Acknowledgments
This work was partially supported by the National Institutes of Health (1 R01 GM088291-01). XH thanks China
Scholarship Council (CSC) for fellowship support. We thank the Kröger group at GT for the use of their zeta
potential equipment.
References
1. Pelton RH, Chibante P. Preparation of Aqueous Lattices with N-Isopropylacrylamide. Colloids and
Surfaces. 1986; 20:247–256.
2. Lee CF, Lin CC, Chi WY. Thermosensitive and control release behavior of poly (N-
isopropylacrylamide-co-acrylic acid) latex particles. Journal of Polymer Science Part a-Polymer
Chemistry. 2008; 46:5734–5741.
3. Choi SH, Yoon JJ, Park TG. Galactosylated poly(N-isopropylacrylamide) hydrogel submicrometer
particles for specific cellular uptake within hepatocytes. Journal of Colloid and Interface Science.
2002; 251:57–63. [PubMed: 16290701]
4. Blackburn WH, Dickerson EB, Smith MH, McDonald JF, Lyon LA. Peptide-Functionalized
Nanogels for Targeted siRNA Delivery. Bioconjugate Chemistry. 2009; 20:960–968. [PubMed:
19341276]
5. Kim J, Nayak S, Lyon LA. Bioresponsive hydrogel microlenses. Journal of the American Chemical
Society. 2005; 127:9588–9592. [PubMed: 15984886]
6. Kim JS, Singh N, Lyon LA. Displacement-induced switching rates of bioresponsive hydrogel
microlenses. Chemistry of Materials. 2007; 19:2527–2532.
7. Zhang YJ, Guan Y, Zhou SQ. Synthesis and volume phase transitions of glucose-sensitive
microgels. Biomacromolecules. 2006; 7:3196–3201. [PubMed: 17096551]
8. Jia XQ, Yeo Y, Clifton RJ, Jiao T, Kohane DS, Kobler JB, Zeitels SM, Langer R. Hyaluronic acid-
based microgels and microgel networks for vocal fold regeneration. Biomacromolecules. 2006;
7:3336–3344. [PubMed: 17154461]
9. Freemont TJ, Saunders BR. PH-responsive microgel dispersions for repairing damaged load-bearing
soft tissue. Soft Matter. 2008; 4:919–924.
Hu et al. Page 6
Colloid Polym Sci. Author manuscript; available in PMC 2011 March 16.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
10. Grabstain V, Bianco-Peled H. Mechanisms controlling the temperature-dependent binding of
proteins to poly(N-isopropylacrylamide) microgels. Biotechnology Progress. 2003; 19:1728–1733.
[PubMed: 14656148]
11. Snowden MJ, Chowdhry BZ, Vincent B, Morris GE. Colloidal copolymer microgels of N-
isopropylacrylamide and acrylic acid: pH, ionic strength and temperature effects. Journal of the
Chemical Society-Faraday Transactions. 1996; 92:5013–5016.
12. Meng ZY, Hendrickson GR, Lyon LA. Simultaneous Orthogonal Chemoligations on
Multiresponsive Microgels. Macromolecules. 2009; 42:7664–7669.
13. Suzuki D, Kawaguchi H. Hybrid microgels with reversibly changeable multiple brilliant color.
Langmuir. 2006; 22:3818–3822. [PubMed: 16584261]
14. Hermanson, GT. Bioconjugate Techniques. Academic Press; San Diego: 1995.
15. Dyer MA, Ainslie KM, Pishko MV. Protein adhesion on silicon-supported hyperbranched
poly(ethylene glycol) and poly(allylamine) thin films. Langmuir. 2007; 23:7018–7023. [PubMed:
17506587]
16. Ramos J, Martin-Molina A, Sanz-Izquierdo MP, Rus A, Borque L, Hidalgo-Alvarez R, Galisteo-
Gonzalez F, Forcada J. Amino-functionalized latex particles obtained by a multistep method:
Development of a new immunoreagent. Journal of Polymer Science Part a-Polymer Chemistry.
2003; 41:2404–2411.
17. Xu JJ, Timmons AB, Pelton R. N-Vinylformamide as a route to amine-containing latexes and
microgels. Colloid and Polymer Science. 2004; 282:256–263.
18. Leung MF, Zhu JM, Harris FW, Li P. New route to smart core-shell polymeric microgels:
Synthesis and properties. Macromolecular Rapid Communications. 2004; 25:1819–1823.
19. Meunier F, Elaissari A, Pichot C. Preparation and Characterization of Cationic Poly(N-
Isopropylacrylamide) Copolymer Latexes. Polymers for Advanced Technologies. 1995; 6:489–
496.
20. Meng Z, Cho JK, Breedveld V, Lyon LA. Physical Aging and Phase Behavior of Multiresponsive
Microgel Colloidal Dispersions. Journal of Physical Chemistry B. 2009; 113:4590–4599.
21. South AB, Whitmire RE, Garcia AJ, Lyon LA. Centrifugal Deposition of Microgels for the Rapid
Assembly of Nonfouling Thin Films. Acs Applied Materials & Interfaces. 2009; 1:2747–2754.
[PubMed: 20356152]
22. Bao LY, Zha LS. Preparation of poly(N-isopropylacrylamide) microgels using different initiators
under various pH values. Journal of Macromolecular Science Part a-Pure and Applied Chemistry.
2006; 43:1765–1771.
23. Mai-ngam K, Boonkitpattarakul K, Sakulsombat M, Chumningan P, Mai-ngam B. Synthesis and
phase separation of amine-functional temperature responsive copolymers based on poly(N-
isopropylacrylamide). European Polymer Journal. 2009; 45:1260–1269.
24. Bokias G, Hourdet D. Synthesis and characterization of positively charged amphiphilic water
soluble polymers based on poly(N-isopropylacrylamide). Polymer. 2001; 42:6329–6337.
25. Lopez-Leon T, Ortega-Vinuesa JL, Bastos-Gonzalez D, Elaissari A. Cationic and anionic poly(N-
isopropylacrylamide) based submicron gel particles: Electrokinetic properties and colloidal
stability. Journal of Physical Chemistry B. 2006; 110:4629–4636.
26. Zhao JP, Zhang GZ, Pispas S. Morphological Transitions in Aggregates of Thermosensitive
Poly(ethylene oxide)-b-poly(N-isopropylacrylamide) Block Copolymers Prepared via RAFT
Polymerization. Journal of Polymer Science Part a-Polymer Chemistry. 2009; 47:4099–4110.
27. Gao J, Frisken BJ. Cross-linker-free N-isopropylacrylamide gel nanospheres. Langmuir. 2003;
19:5212–5216.
28. Gao J, Frisken BJ. Influence of secondary components on the synthesis of self-cross-linked N-
isopropylacrylamide microgels. Langmuir. 2005; 21:545–551. [PubMed: 15641822]
29. Huang X, Misra GP, Vaish A, Flanagan JM, Sutermaster B, Lowe TL. Novel Nanogels with Both
Thermoresponsive and Hydrolytically Degradable Properties. Macromolecules. 2008; 41:8339–
8345.
30. Ma XM, Xi JY, Xian Z, Tang XZ. Deswelling comparison of temperature-sensitive poly(N-
isopropylacrylamide) microgels containing functional-OH groups with different hydrophilic long
side chains. Journal of Polymer Science Part B-Polymer Physics. 2005; 43:3575–3583.
Hu et al. Page 7
Colloid Polym Sci. Author manuscript; available in PMC 2011 March 16.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
31. Debord JD, Lyon LA. Synthesis and characterization of pH-responsive copolymer microgels with
tunable volume phase transition temperatures. Langmuir. 2003; 19:7662–7664.
32. Daly E, Saunders BR. A study of the effect of electrolyte on the swelling and stability of poly(N-
isopropylacrylamide) microgel dispersions. Langmuir. 2000; 16:5546–5552.
33. Fernandez-Nieves A, Fernandez-Barbero A, de las Nieves FJ. Salt effects over the swelling of
ionized mesoscopic gels. Journal of Chemical Physics. 2001; 115:7644–7649.
34. van de Wetering P, Zuidam NJ, van Steenbergen MJ, van der Houwen OAGJ, Underberg WJM,
Hennink WE. A mechanistic study of the hydrolytic stability of poly(2-(dimethylamino)ethyl
methacrylate). Macromolecules. 1998; 31:8063–8068.
35. Thompson KL, Read ES, Armes SP. Chemical degradation of poly(2-aminoethyl methacrylate).
Polymer Degradation and Stability. 2008; 93:1460–1466.
Hu et al. Page 8
Colloid Polym Sci. Author manuscript; available in PMC 2011 March 16.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 1.
AFM height images and line profiles of (a) μG0, (b) μG3, (c) μG4, (d) μG5, and (e) μG6. (f)
Histogram of calculated height/diameter ratios.
Hu et al. Page 9
Colloid Polym Sci. Author manuscript; available in PMC 2011 March 16.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Figure 2.
(a) Hydrodynamic diameter and (b) swelling ratio vs. T of the indicated microgels in pH 3
buffer (I = 10 mM). (c) Diameter vs. T for μG3 microgels in pH 3 buffers with different
ionic strengths. μG0 microgels in pH 3 (10 mM) is shown for comparison. The dependence
of: (d) swelling ratio and (e) ζ-potential of the indicated microgels as a function of pH with
the ionic strength held constant at 2 mM. (f) Fluorescence microscopy image of fluorescein-
labeled μG6 microgels.
Hu et al. Page 10
Colloid Polym Sci. Author manuscript; available in PMC 2011 March 16.
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
NIH-PA Author Manuscript NIH-PA Author Manuscript NIH-PA Author Manuscript
Hu et al. Page 11
Table 1
Synthesis conditions, yields, and microgel properties.
CodeaAPMH/mol % V50/mM APS/mM [NaCl]/mM Dh/nmbζ-potential in DI water/mV Yield/%
μG0 0 2 - 0 714 13.4 45
μG1 9 2 - 0 157 12.9 20
μG2 9 2 - 50 318 33.8 25
μG3 9 2 - 100 508 37.8 35
μG4 9 2 - 150 950 38.9 38
μG5 9 - 1 0 498 33.8 60
μG6 9 - 2 0 744 41.7 65
aReaction conditions: total monomer concentration = 140 mM, temperature = 70 °C, reaction time: 20 h.
bSolution conditions: pH 3, 10 mM ionic strength.
Colloid Polym Sci. Author manuscript; available in PMC 2011 March 16.
... They found that when the AEM concentration was increased above 0.5% of the total monomer concentration, the water-soluble oligomer polyelectrolyte fraction of the reaction mixture that had not been incorporated into the microgel particles started to steeply increase and when more than 1 mol% AEM was applied the microgel particles became rather polydisperse. Similar results were found both in the case of the copolymerization of dimethylaminoethyl methacrylate (DMAEMA) [32,33] and N-(3-aminopropyl) methacrylamide hydrochloride (APMH) [34]. Very recently, Annegarn et al. [35] made a systematic investigation to characterize if the increasing pH and the consequent deprotonation of the APMH repeat units in the water-soluble oligomer chains could counterbalance the negative effects of APMH copolymerization into the microgel particles. ...
Article
Full-text available
Responsive cationic microgels are a promising building block in several diagnostic and therapeutic applications, like transfection and RNA or enzyme packaging. Although the direct synthesis of cationic poly(N-isopropylacrylamide) (PNIPAm) microgel particles has a long history, these procedures typically resulted in low yield, low incorporation of the cationic comonomer, increased polydispersity, and pure size control. In this study, we investigated the possibility of the post-polymerization modification of P(NIPAm-co-acrylic acid) microgels to prepare primary amine functionalized microgels. To achieve this goal, we used 1-ethyl-3-(3-(dimethylamino)propyl)carbodiimide hydrochloride (EDC) mediated coupling of a diamine to the carboxyl groups. We found that by controlling the EDC excess in the reaction mixture, the amine functionalization of the carboxyl functionalized microgel could be varied and as much as 6–7 mol% amine content could be incorporated into the microgels. Importantly, the reaction was conducted at room temperature in an aqueous medium and it was found to be time efficient, making it a practical and convenient approach for synthesizing primary amine functionalized PNIPAm microgel particles.
... We selected PNIPAAm-based networks crosslinked with N,N'-methylenebis(acrylamide) (BIS) as model system, since NIPAAm is one of the most widely used monomers and exhibits temperature-induced VPT close to human body temperature. [17a,31] Weakly acidic methacrylic acid (MAAc, pK A � 5.0 [32] ) and weakly basic N-(3-aminopropyl)methacrylamide (APMH, pK A � 10.0 [33] ) were introduced as pH-responsive co-monomers, because they are water-soluble and oppositely charged at neutral pH value. Polyelectrolyte NG with a hydrodynamic diameter of 400-500 nm and a co-monomer content of 10 mol% were synthesized by precipitation polymerization (Supporting Information, Table S1). ...
Article
Full-text available
In pH‐responsive drug carriers, the distribution of charges has been proven to affect delivery efficiency but is difficult to control and verify. Herein, we fabricate polyampholyte nanogel‐in‐microgel colloids (NiM−C) and show that the arrangement of the nanogels (NG) can easily be manipulated by adapting synthesis conditions. Positively and negatively charged pH‐responsive NG are synthesized by precipitation polymerization and labelled with different fluorescent dyes. The obtained NG are integrated into microgel (MG) networks by subsequent inverse emulsion polymerization in droplet‐based microfluidics. By confocal laser scanning microscopy (CLSM), we verify that depending on NG concentration, pH value and ionic strength, NiM−C with different NG arrangements are obtained, including Janus‐like phase‐separation of NG, statistical distribution of NG, and core–shell arrangements. Our approach is a major step towards uptake and release of oppositely charged (drug) molecules.
... They have attracted a lot of attention in the past ten years because of the prospective applications for smart microgels in the nanotechnological [12], biomedical [13], and catalytic domains [14,15]. Generally, the applications of smart microgels depend on fast responsive behavior of these materials (in swelling or deswelling form) under stimuli conditions like ionic strength [16], temperature [17], and pH [18,19] in the medium. Additionally, their use as micro-reactors for the production and stability of inorganic metal nanoparticles has been reported in the documentation frequently [20,21]. ...
Article
The bimetallic nanoparticles and smart microgels/hydrogels are a particularly interesting combination that have received a lot of attention over the past ten years. These composite materials (bimetallic nanoparticles loaded in smart microgels/hydrogels) are a worthy candidate for a wide range of biological, environmental, and catalytic applications due to the sensitive behavior of microgels in their hybrid systems. There are numerous morphologies and architectural designs for hybrid microgels based on bimetallic nanoparticles reported in the literature. The morphology of microgels in hybrid systems is the key for their usage. For this reason, a specific assembly of hybrid microgels based on bimetallic nanoparticles is designed, specifically, for a particular purpose. This article reviews current developments in the synthesis, classification, characterization, properties, and applications of bimetallic nanoparticles loaded in microgels. The applications of bimetallic nanoparticles loaded in microgels in the environment, biomedicine, sensing, and catalysis have been covered in a tutorial approach.
Article
Die Verteilung von Ladungen innerhalb pH‐responsiver Wirkstoffträger beeinflusst nachweislich die Effizienz der Freisetzung, ist aber schwer zu kontrollieren und zu verifizieren. Wir stellen nun polyampholytische Nanogel‐in‐Mikrogel Kolloide (NiM−C) vor und zeigen, dass die Anordnung der Nanogele (NG) abhängig von den Synthesebedingungen ist. Dazu wurden pH‐responsive Polyelektrolyt‐NG durch Fällungspolymerisation synthetisiert und mit Fluoreszenzfarbstoffen markiert. Die NG wurden dann durch inverse Emulsionspolymerisation mittels tropfenbasierter Mikrofluidik in Mikrogel‐Netzwerke (MG) integriert. Mit konfokaler Laser‐Scanning‐Mikroskopie (CLSM) wurde nachgewiesen, dass, je nach NG‐Konzentration, pH‐Wert und Ionenstärke, NiM−C mit unterschiedlichen NG‐Anordnungen, wie Janus‐artiger Phasentrennung, statistischer Verteilung und Kern‐Schale‐Architektur, erhalten werden. Unser Ansatz ist ein Meilenstein auf dem Weg zum simultanen Transport entgegengesetzt geladener Moleküle.
Article
Full-text available
Stimuli-responsive hydrogel-microgel composites (HMC) were prepared by embedding poly(N-isopropylacrylamide)-based microgels in a poly(N-isopropylacrylamide)-based hydrogel. When the microgels were pre-loaded with the small molecule model drug crystal violet (CV) via electrostatics, the HMC was able to release the CV in a pH-triggered fashion. We found that the CV release rate was dependent on the solution temperature and the dimension of the material. Also, by changing the chemical composition and/or pore size of the hydrogel matrix, the CV release kinetics can be tuned. Moreover, when multiple microgels loaded with different model drugs were embedded in a single HMC, the HMC can be used to control the release rate of each drug analog individually in a pH-dependent fashion. By understanding how properties of a hydrogel can alter the release of small molecules from embedded microgels, new materials capable of controlled and triggered release of multiple small molecule drugs can be designed with myriad uses in the biomedical field.
Article
Responsive poly-N-isopropylacrylamide-based microgels are commonly used as model colloids with soft repulsive interactions. It has been shown that the microgel-microgel interaction in solution can be easily adjusted by varying the environmental parameters, e.g., temperature, pH, or salt concentration. Furthermore, microgels readily adsorb to liquid-gas and liquid-liquid interfaces forming responsive foams and emulsions that can be broken on-demand. In this work, we explore the interactions between microgel monolayers at the air-water interface and a hard colloid in the water. Force-distance curves between the monolayer and a silica particle were measured with the Monolayer Particle Interaction Apparatus. The measurements were conducted at different temperatures and lateral compressions, i.e., different surface pressures. The force-distance approach curves display long-range repulsive forces below the volume phase transition temperature of the microgels. Temperature and lateral compression reduce the stiffness of the monolayer. The adhesion increases with temperature and decreases with a lateral compression of the monolayer. When compressed laterally, the interactions between the microgels are hardly affected by temperature, as the directly adsorbed microgel fractions are nearly insensitive to temperature. In contrast, our findings show that the temperature-dependent swelling of the microgel fractions in the aqueous phase strongly influences the interaction with the probe. This is explained by a change in the microgel monolayer from a soft to a hard repulsive interface.
Article
Aqueous colloidal microgels have been prepared, based on poly(N-isopropylacrylamide)[poly(NIPAM)], cross-linked with bisacrylamide, containing 5% w/w acrylic acid (AAc) as a comonomer. Transmission electron micrographs of the microgels show that the copolymer microgels are monodisperse spheres. The size of the microgel particles containing 5% AAc has been studied, using dynamic light scattering, as a function of pH (3–10), ionic strength (10–4–10–1M NaCl) and temperature (20–75 °C). The hydrodynamic diameter of the copolymer microgels decrease both with increasing ionic strength (at pH 6 and 25 °C) and reversibly with increasing temperature at pH values of 2.6, 3.4 and 6.5. However, under isothermal conditions in 10–3M NaCl at 25 °C, the hydrodynamic diameter increases in going from pH 3.3–9.4. In addition, the temperature-induced conformational changes in the polymer chains have been followed using high-sensitivity differential scanning calorimetry (HSDSC). A comparison is made with the behaviour of poly(NIPAM) microgel particles, not containing AAc. Explanations are offered to account for the pronounced difference in the physico-chemical properties observed.
Article
An important challenge for colloid scientists is to design injectable dispersions that provide structural support for damaged soft tissue and enable regeneration of tissue over the longer term. In this article we highlight a new area of research that aims to produce pH-responsive microgel dispersions that restore the mechanical properties of damaged, load-bearing, soft tissue. Chronic back pain due to degeneration of the intervertebral disc (IVD) is a major health problem and is the primary potential application for the work discussed. pH-Responsive microgel dispersions contain cross-linked polymer particles that swell when the pH approaches the pKa of the incorporated ionic co-monomer. The work considered here involves microgel particles containing MAA (methacrylic acid). The particles show pronounced pH-triggered swelling. The concentrated microgel dispersions change from a fluid to a gel at pH values greater than ca. 6.2, which is within the physiological pH range. The rheological properties are pH-dependent and can be adjusted using particle composition or concentration. Degenerated IVDs containing injected, gelled, microgel dispersions show improved mechanical properties. The disc height under biomechanically meaningful loads can be restored to values observed in non-degenerated IVDs. We also discuss the steps required to provide a minimally invasive injectable microgel system for restoring both the IVD mechanical properties and regenerating tissue in vivo. The approach discussed should also be suitable for other soft tissue types in the body.
Article
Poly(2-aminoethyl methacrylate) (PAMA) has a pKa of approximately 7.6 and is chemically stable in acidic or neutral aqueous solution in its protonated form. However, chemical degradation of PAMA is known to occur in alkaline media as its primary amine groups become deprotonated (He L et al. Macromolecules 2007; 40: 4429–38). In the present work, the effect of temperature, pH and polymer concentration on the rate of PAMA degradation in dilute aqueous solution has been examined. 1H NMR spectroscopy indicates that both elimination of 2-aminoethanol and formation of 2-hydroxyethyl methacrylamide repeat units occur above pH 9; elimination is observed first and occurs to a greater extent. FT-IR studies of aqueous PAMA solutions aged at pH 12 and 50°C confirm the presence of anionic carboxylate groups, which suggests that such elimination is simply due to ester hydrolysis. A control experiment suggests that methacrylamide formation occurs via internal rearrangement, rather than by amidation of the remaining ester groups by the eliminated 2-aminoethanol.
Article
In this work, poly(N-isopropylacrylamide-co-acrylic acid) (poly(NIPAAm-AA)) copolymer latex particles (microgels) were synthesized by the method of soapless emulsion polymerization. Poly(NIPAAm-AA) copolymer microgels have the property of being thermosensitive. The concentration of acrylic acid (AA) and crosslinking agent N,N′-methylenebisacrylamide were important factors to influence the lower critical solution temperature (LCST) of poly(NIPAAm-AA) microgels. The effects of AA and crosslinking agent on the swelling behavior of poly(NIPAAm-AA) microgels were also studied. The poly(NIPAAm-AA) copolymer microgels were then used as a thermosensitive drug carrier to load caffeine. The effects of concentration of AA and crosslinking agent on the control release of caffeine were investigated. How the AA content and crosslinking agent influenced the morphology and LCST of the microgels was discussed in detail. The relationship of morphology, swelling, and control release behavior of these thermosensitive microgels was established. A new scheme was proposed to interpret the control release of the microgels with different morphological structures. © 2008 Wiley Periodicals, Inc. J Polym Sci Part A: Polym Chem 46: 5734–5741, 2008
Article
Ammonium persulfate (APS), 2,2′‐azobis(amidinopropane) dihydrochloride (V50) and 4,4′‐azobis(4‐cyanovaleric acid) (ACVA) were utilized to prepare temperature‐sensitive poly(N‐isopropylacrylamide) (PNIPAM) microgels by precipitation polymerization under various reaction pH conditions. Their particle sizes and swelling ratios depended on the reaction pH due to the pH dependence on the ionization degree of the decomposed fragments originating from the initiators and their hydrophilicity‐hydrophobicity. The more hydrophobic initiator, under the reaction pH conditions used, could be partitioned to a greater extent into the microgel particles due to the hydrophobicity of PNIPAM chains at the reaction temperature, which led to a more cross‐linked structure inside the microgels resulting in their smaller swelling ratio. pH dependence of surface charge density of the microgels with amidino groups or carboxylic acid groups on their surfaces was evidenced by the variation of their zeta potentials as a function of pH.
Article
The synthesis and characterization of new water-soluble shear-responsive polymers based on N-isopropylacrylamide (NIPAM) is reported. These copolymers are synthesized using a two-steps procedure: (a) synthesis of the precursors by radical copolymerization in aqueous solvent between NIPAM and an amine-containing monomer N,N-[(dimethylamino)propyl]methacrylamide (MADAP) and (b) modification of the macromolecular backbone by reaction between MADAP units and alkylbromide. From kinetic studies performed during copolymerization, the reactivity ratios of the two monomers were determined in media of different ionic strength (rNIPAM=0.6andrMADAP=0.6−1.3). Under these conditions, it was shown that the composition and the distribution of NIPAM and MADAP units into the macromolecular backbone remain approximately constant until 80% of conversion. For all the copolymers, in the composition range 0–25mol% of MADAP units, an effective control of the molar mass is reported with the concentration of the reducing agent involved in the redox initiator. Finally, a quantitative alkylation procedure of MADAP units is described in chloroform using an inactive base and a large excess of alkylbromide.
Article
In this work, the effects of salt concentration over the swelling of ionic mesoscopic gels will be studied theoretically and verified with experiments. We will restrict ourselves to the weak screening limit where the Debye screening length is larger than the mesh size of the gel. Under this condition, direct electrostatic interactions are negligible and the swelling is driven by the osmotic pressure of the ions. The swelling response of a mesoscopic gel is strongly dependent on the ionization degree of the gel. In particular, a maximum in the size–salt concentration curve appears for a partially ionized gel, when the salt concentration equals the network charge concentration. This maximum is removed for a totally ionized polymer network. Despite these facts, a charge independent asymptotic behavior between size and salt concentration becomes apparent, at sufficiently high values of the latter. The Flory–Huggins mean-field approach together with the Donnan relations describe the observed swelling adequately, thus accounting for the essential physics of the problem. As a final remark, an experimental method for determining gel charge will be inferred from the results. © 2001 American Institute of Physics.
Article
We describe the straightforward synthesis of “clickable” multiresponsive microgels containing both carboxylic acid groups and azidohydrin or terminal alkyne groups, via a one-pot multistage polymerization approach. The clickable functional groups on the microgels were confirmed by FTIR. Additionally, we simultaneously performed “click” and acid−amine coupling reactions on microgels with fluorescent dyes containing complementary functional groups. Epifluorescence microscopy was employed to confirm the coupling of those dyes to the microgels. The orthogonality of the click reaction to other functional groups such as hydroxyl, carboxylic acid, and amino groups was confirmed, suggesting the potential utility of such microgels in applications where multifunctional colloidal particles are required.
Article
A series of novel nanogels with both thermoresponsive and hydrolytically degradable properties were synthesized by emulsion polymerization of N-isopropylacrylamide (NIPAAm) and dextran-lactate-2-hydroxyethyl methacrylate (DEXlactateHEMA), a hydrolytically degradable and cross-linkable dextran derivative, without using low molar mass surfactants. Various lengths of degradable oligolactate units and different precursor feeding ratios between NIPAAm and DEXlactateHEMA were used to synthesize the nanogels. FTIR measurements confirmed the chemical compositions and hydrolytic degradation of the synthesized nanogels. Dynamic light scattering measurements of the hydrodynamic radii of the nanogels in phosphate buffer saline (PBS, pH 7.4) against temperature and angle revealed that the nanogels were thermoresponsive with a lower critical solution temperature (LCST) of 32 °C. The size and morphology changes of the nanogels with degradation were investigated by using transmission electron microscopy, atomic force microscopy, and static light scattering techniques. AFM image analysis and Holtzer plots revealed that the nanogels became more rigid with degradation in water solutions.
Article
Poly(N-isopropylacrylamide) (poly(NIPAM)) microgel dispersion stability has been investigated in the presence of a range of electrolytes. The stability of the dispersions to flocculation is strongly related to the free energy change of formation in aqueous solution for the anions (ΔGfo) and their position in the lyotropic series. The ability of the anions to induce particle deswelling and flocculation has been investigated using photon correlation spectroscopy and turbidity-wavelength measurements, respectively. Anions with a high charge density (and high negative values for ΔGfo) have the ability to dehydrate the poly(NIPAM) chain and induce flocculation. The ability of cations to induce particle deswelling and flocculation was also investigated. The ability of the cations to induce particle deswelling did not coincide with their values for ΔGfo and their position in the lyotropic series. Time-dependent turbidity measurements of collapsed particles in the presence of NaCl were used to calculate the Hamaker constant (Ap) for poly(NIPAM). The value found was Ap = 6.2 × 10-20 J, which compares favorably to that reported for polystyrene and poly(NIPAM)−polystyrene composites. Turbidity-wavelength measurements for poly(NIPAM) particles dispersed in aqueous NaCl solutions revealed that the critical flocculation temperature of the dispersion coincided with the lower critical solution temperature.