Sunteți pe pagina 1din 12

Journal of Colloid and Interface Science 536 (2019) 536–547

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


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

Regular Article

Synthesis and characterization of poly(N-isopropylacrylamide-co-


acrylamide) mesoglobule core–silica shell nanoparticles
Ngoc-Hanh Cao-Luu a, Quoc-Thai Pham a, Zong-Han Yao c, Fu-Ming Wang b, Chorng-Shyan Chern a,⇑
a
Department of Chemical Engineering, National Taiwan University of Science and Technology, Taipei, Taiwan
b
Graduate Institute of Applied Science and Technology, National Taiwan University of Science and Technology, Taipei 106, Taiwan
c
Department of Internal Medicine, National Taiwan University Hospital, Taipei 106, Taiwan

g r a p h i c a l a b s t r a c t

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

Article history: Hypothesis: How to encapsulate poly(N-isopropylacrylamide) (PNIPAM) mesoglobule cores by silica
Received 10 September 2018 shells greatly affects the resultant nanoparticle structures. Incorporation of acrylamide (AM) unit into
Revised 26 October 2018 PNIPAM in combination with 3-glycidyloxypropyltrimethoxysilane (GLYMO, as a coupling agent)
Accepted 27 October 2018
effectively induces nucleation and growth of silica on PNIPAM core surfaces, where the –NH2 of
Available online 29 October 2018
acrylamide reacts with the epoxide of GLYMO while GLYMO further participates in subsequent sol-gel
reaction of tetraethyl orthosilicate (TEOS), thereby leading to desirable particle morphology.
Keywords:
Experiments: PNIPAM-based core–silica shell nanoparticles were prepared by sol-gel reaction of TEOS
PNIPAM based mesoglobule
Silica
and GLYMO in the presence of polymeric core particles. The major parameters investigated in a system-
Coupling agents atic fashion include acrylamide concentration and weight ratio of polymer:GLYMO:TEOS. GPC, DLS, DSC,
Hybrid core–shell nanoparticles FE-SEM, TEM, FTIR and TGA were then used to characterize polymeric cores and hybrid nanoparticles.
Findings: The particle morphology was governed primarily by the acrylamide content and the weight
ratio of PNIPAM/AM:GLYMO:TEOS, and desirable hybrid nanoparticles with narrow particle size distribu-
tion were achieved. The LCST of PNIPAM-based mesoglobules increases with increasing acrylamide
content. Encapsulation of PNIPAM-based mesoglobules with silica also reduces their thermo-
sensitivity. This is the first report of developing a novel approach to prepare PNIPAM-based mesoglobule
core–silica shell nanoparticles with controllable particle morphologies.
Ó 2018 Published by Elsevier Inc.

1. Introduction

Amongst various thermo-sensitive polymers, N-


isopropylacrylamide (PNIPAM) is the most prominent candidate
⇑ Corresponding author. for biomedical applications because of its biocompatibility and
E-mail address: cschern@mail.ntust.edu.tw (C.-S. Chern).

https://doi.org/10.1016/j.jcis.2018.10.091
0021-9797/Ó 2018 Published by Elsevier Inc.
N.-H. Cao-Luu et al. / Journal of Colloid and Interface Science 536 (2019) 536–547 537

the lower critical solution temperature (LCST) around 32 °C that is adopted to enhance the affinity between the organic and inorganic
quite close to the physiological temperature of human body [1,2]. part [43]. In this manner, the amide group of AM can react with the
PNIPAM comprising both the hydrophilic amide group and epoxide group of GLYMO, while the four ethoxysilyl groups of TEOS
hydrophobic isopropyl group is well characterized [3–5]. Below (silica precursor) and the three methoxysilyl groups of GLYMO can
LCST, the amide group of PNIPAM interacts strongly with water simultaneously undergo the sol-gel reaction [43–45] to form tiny
molecules via hydrogen bonding, thereby leading to a large silica nanoparticles right on the PNIPAM/AM mesoglobule core sur-
amount of water being contained in the expanded polymer coils faces, thereby leading to the desirable particle morphology. The
in water (as a homogeneous aqueous polymer solution) [6–11]. results obtained from this work are useful in designing thermo-
In contrast, PNIPAM chain becomes more compact due to the dis- responsive PNIPAM/AM mesoglobule core–silica shell nanoparti-
rupted hydrogen bonds between the amide group of PNIPAM and cles with controlled particle morphology and physicochemical
water molecules when the temperature is greater than LCST properties.
[12–16]. This will result in separation of liquid and solid phases
and then cause collapse of the swollen polymer coils (termed the 2. Experimental
mesoglobules) [12–16].
Nun et al. [17] reported that PNIPAM is of relatively high soft- 2.1. Materials
ness in the hydrated state, which greatly limits its mechanical sta-
bility. To resolve this issue, interpenetrating networks comprising The chemicals used in this work include N-isopropylacrylamide
PNIPAM and polymers with desirable mechanical properties or (NIPAM, 99%, Acros), acrylamide (AM, 99%, Sigma-Aldrich), potas-
the incorporation of inorganic nanoparticles into PNIPAM were sium persulfate (KPS, 99%, Acros), 3-glycidyloxypropyltrimethoxy
proposed [17]. Byun et al. [18] pointed out that inorganic silica silane (GLYMO, 98%, Evonik), tetraethyl orthosilicate (TEOS, 98%,
showing excellent physicochemical properties (e.g., mechanical Acros), potassium bromide (KBr, 99+, Acros), tetrahydrofuran
strength, chemical and thermal stability, biocompatibility, low tox- (THF, 99.9%, Sigma-Aldrich) and ethanol (95%, Uni-Onward Corp).
icity and low permeability) and low cost. Furthermore, the synthe- Deionized water (Barnsted, Nanopure Ultrapure Water System,
sis of silica can be directly carried out on the PNIPAM particle core specific conductance <0.057 mS cm 1) was used for all experiments.
surfaces to form PNIPAM core–silica shell particles, which is a All chemicals were reagent grade and used as received.
highly desirable approach [17]. For example, PNIPAM particles
filled with multiple colloidal silica with the particle size close to
2.2. Synthesis of polymeric cores
1 mm via in situ sol-gel process was achieved [18]. The thermo-
sensitive raspberry-like hybrid particles (particle size 500 nm in
The samples, PNIPAM, PNIPAM/5AM and PNIPAM/10AM, were
diameter) consisting of PNIPAM microgel cores decorated with sil-
synthesized via free radical polymerization of 1 g of NIPAM in
ica particles was reported by Dechezelles et al. [19]. Wang and
the presence of 0, 0.05 and 0.1 g of AM, respectively, with 0.04 g
Asher [20] prepared the PNIPAM core surrounded by a shell com-
of KPS as the initiator in 10 g of water. The reaction was carried
prising very small silica particles with a size of ca. 400 nm in diam-
out at 70 °C over a period of 24 h in nitrogen atmosphere. The
eter. Poly(NIPAM-co-acrylic acid) (PNIPAM/AA) core–silica shell
crude polymer solution was washed with hot water (50 °C) to
particles (ca. 250 nm in diameter) were successfully synthesized
remove residual monomer(s) and initiator (if present). The resul-
by Hu and co-workers [21]. Duan et al. [22] proposed an approach
tant polymer (PNIPAM, PNIPAM/5AM or PNIPAM/10AM) was then
to synthesize PNIPAM/silica composite microspheres by Pickering
dissolved in cold water and stored as the stock solution for further
emulsion polymerization. It is noteworthy that most of the PNI-
experiments.
PAM/silica composite particles reported in the literature so far
are rather limited in drug delivery applications because of their
micron-sized characteristic, originating from the micron-sized 2.3. Synthesis of hybrid core–shell particles
crosslinked PNIPAM core particles [17–23]. It is very difficult for
micron-sized particles (as a drug carrier) to penetrate tissues and The polymeric cores were encapsulated by silica shells via the
then enter the cells, thereby leading to inefficient delivery of drug sol-gel reaction of GLYMO (a coupling agent) and tetraethyl
molecules to target sites of action [24–29]. Uptake of nanoparticles orthosilicate (TEOS, a silica precursor). First, the aqueous polymer
is approximately 15–250 times higher than that of microparticles solution (0.5 wt%) was adjusted to pH 12 by sodium hydroxide
in the particle size range 1–10 mm [25,30,31]. Therefore, it is crucial (0.5 M) and then heated up to 50 °C. Subsequently, GLYMO was
for the dimension of PNIPAM/silica composite particles to be con- added into the polymer dispersion, immediately followed by addi-
trolled within the nano-sized range via minimizing the core parti- tion of TEOS into the reaction mixture. The reaction was carried out
cle size. In this regard, some representative methods involving at 50 °C over a period of the prescribed time (e.g., 5 h). The final
adjustment of the size of cross-linked PNIPAM core particles product was first precipitated by centrifugation (25,000 rpm,
include the addition of surfactants [32–34], electrolytes [35–38] 30 min) and washed with alcohol and then water three times.
and highly charged co-monomers [35,39]. Nevertheless, the micro- Table 1 summarizes the recipes used to prepare the hybrid
gel particle size was not optimally controlled in these studies, even organic–inorganic (core–shell) colloidal particles in this work.
under well-managed experimental conditions [35].
In this work, we report a novel approach to control the size of 2.4. Characterization
poly(NIPAM-co-acrylamide) (PNIPAM/AM) mesoglobule core–sil-
ica shell nanoparticles prepared at temperature above the LCST. The relative weight average molecular weight (Mw) of polymer
These mesoglobule cores in nanoscale form via separation of liquid was determined by gel permeation chromatography (GPC, Waters
and solid phases and subsequent collapse of the swollen polymer Styragel HR2, HR4 and HR6). A series of polystyrene standards was
coils (in the absence of surfactant) above the LCST in the dilute used to establish the calibration curveTHF was used as the eluent
solution [40,41]. This is followed by using the mesoglobule cores solvent, and a flow rate of 1 mL min 1 used for GPC measurements
as the nucleating agent for silica encapsulation to form the perfect at 40 °C. The average hydrodynamic diameter of colloidal particles
core–shell nanostructure [42]. For efficient silica encapsulation, 3- (d) and the lower critical solution temperature (LCST) were
glycidyloxypropyltrimethoxysilane (GLYMO, a coupling agent) obtained from dynamic light scattering technique (DLS, Malvern
containing one epoxide group and three methoxysilyl groups is 1000 HAS) in the temperature range 26–50 °C with 2 °C intervals
538 N.-H. Cao-Luu et al. / Journal of Colloid and Interface Science 536 (2019) 536–547

Table 1
Recipes for encapsulation of PNIPAM-based core particles by silica.

Samples PNIPAM (g) PNIPAM/5AM (g) PNIPAM/10AM (g) GLYMO (g) TEOS (g) H2O (g)
a
PNIPAM-111 0.1 – – 0.1 0.1 20
PNIPAM/5AM-101a – 0.1 – – 0.1 20
PNIPAM/5AM-111a – 0.1 – 0.1 0.1 20
PNIPAM/5AM-121 a – 0.1 – 0.2 0.1 20
PNIPAM/5AM-112 a – 0.1 – 0.1 0.2 20
PNIPAM/5AM-122 a – 0.1 – 0.2 0.2 20
PNIPAM/10AM-111 a – – 0.1 0.1 0.1 20
GLYMO/TEOS-11b – – – 0.1 0.1 20
a
The last three digits represent the weight ratio of polymer:GLYMO:TEOS.
b
The last two digits represent the weight ratio of GLYMO:TEOS.

with a concentration range 0.1–1 wt%. The thermal behavior of the –NH2 group of the AM unit is much more reactive toward
both polymer and core–shell particles with a total solids content GLYMO as compared to the secondary hydrogen of the >NH group
of 10% were investigated by Differential Scanning Calorimetry of the NIPAM unit in PNIPAM-based chain that is most likely due to
(DSC, TA Instruments Q20) from 25 to 60 °C with a scanning rate the satiric effect involved in N-isopropylacrylamide, as supported
of 2 °C min 1. The dried particle size and morphology were charac- by the literature dealing with competitive primary and secondary
terized by field emission scanning electron microscopy (FE-SEM, amines in the ring-opening reaction of epoxy resins [45,46]. It is
JEOL JSM-6500F) and transmission electron microscopy (TEM, H- crucial for the formation of silica nuclei to be induced on the sur-
7100 TEM, Hitachi). The sample was diluted in water to a total face of polymeric cores during the early stage of the sol-gel process
solids content of ca. 0.5%, dropped on the carbon tapes (SEM) or in order to achieve desirable organic core–inorganic shell particle
the polymer coated copper grids (TEM) and dried at room temper- morphology. This is followed by diffusion of hydrolyzed TEOS
ature. The number average particle diameter (dn), weight average [i.e., silicic acid (Si(OH)4)] and oligomeric silica from the bulk aque-
particle diameter (dw) and polydispersity index (PDI = dw/dn) were ous phase to the interfacial layer between the continuous aqueous
determined by SEM or TEM (at least 150 particles measured). Four- phase and the disperse phase, in which the subsequent sol-gel
ier transform infrared (FTIR) spectra were recorded on FTS-3500 reactions take place, thereby leading to the continuous growth of
instrument (Bio-Rad) in the wavenumber range 4000–400 cm 1 the silica shell. Ultimately, the hybrid nanoparticles comprising
with a total of 32 scans and a resolution of 4 cm 1 under nitrogen PNIPAM core and silica shell can be achieved.
flow to remove the moisture (if present). Thermogravimetric anal-
ysis (TGA) measurement was carried out on a Perkin-Elmer Dia- 3.2. Preparation and characterization of PNIPAM based mesoglobule
mond TG/DTA instrument with a heating rate of 10 °C min 1 cores
under an air flow rate of 200 mL min 1 in the temperature range
100–800 °C. PNIPAM based mesoglobule cores prepared by free radical poly-
merization of NIPAM in the presence of 0 wt% AM (PNIPAM), 5 wt%
AM (PNIPAM/5AM) and 10 wt% AM (PNIPAM/10AM) at 70 °C were
3. Results and discussion characterized by DLS and GPC, and the results summarized in
Table 2. Above LCST, the intra- and/or intermolecular hydrophobic
3.1. The approach used to synthesize PNIPAM based mesoglobule core– interactions of PNIPAM induce the precipitation of polymer chains
silica shell nanoparticles to form particle embryos in the continuous aqueous phase. This is
followed by the growth of these nuclei via molecular diffusion of
Formation of PNIPAM core–silica shell nanoparticles is PNIPAM from the bulk aqueous phase to the particle surfaces
schematically shown in Scheme 1, in which mesoglobule cores and subsequent deposition therein or aggregation of particles in
are encapsulated by silica nanoparticles with GLYMO as the cou- order to reduce the interfacial free energy. Table 2 shows that, at
pling agent. Note that GLYMO possessing the epoxy group that constant AM concentration, the average hydrodynamic particle
can react with the amino group of the AM unit (as a comonomer, diameter determined at 50 °C (d50) increases with increasing
if present) or it may react with the amide group of the NIPAM unit concentration and the particle size distribution is quite broad
by the ring-opening reactions. Note that the primary hydrogen of (PDI > 0.1). Yan et al. [10] indicated that the size of PNIPAM

Scheme 1. Schematic illustration of the formation of PNIPAM - silica (core - shell) nanoparticles.
N.-H. Cao-Luu et al. / Journal of Colloid and Interface Science 536 (2019) 536–547 539

Table 2
Particle size and particle size distribution and molecular weight and molecular weight distribution data for PNIPAM, PNIPAM/5AM and PNIPAM/10AM samples with different
concentrations.

Samples Concentration (wt%) d50a (nm) PDIa Mwb  104 (g mol 1


) PDIb
PNIPAM 0.1 42 ± 1 0.08–0.10 6.33 1.71
0.5 65 ± 2 0.06–0.08
1.0 98 ± 5 0.06–0.14
PNIPAM/5AM 0.1 56 ± 2 0.06–0.10 1.88 1.99
0.5 76 ± 4 0.05–0.09
1.0 119 ± 4 0.07–0.17
PNIPAM/10AM 0.1 58 ± 2 0.07–0.11 1.07 1.70
0.5 79 ± 5 0.08–0.14
1.0 136 ± 6 0.09–0.19
a
d50 is the average hydrodynamic particle diameter and PDI is the polydispersity index determined by DLS at 50 °C. PDI < 0.1 indicates narrow particle size distribution,
whereas PDI > 0.1 then indicates broad particle size distribution. All experiments were conducted three times and DLS measurements are repeated ten times for each
experiment.
b
Mw is the weight-average molecular weight and PDI (=Mw/Mn) is the polydispersity index measured by GPC, where Mn is the number-average molecular weight.

particles was primarily dependent on the PNIPAM concentration over a period of 7 h (Fig. S1C). It is noteworthy that smaller hybrid
and molecular weight. In general, the lower the PNIPAM concen- core–shell particles were produced in the run at 50 °C for 5 h in
tration, the smaller the particles formed [12]. Particle embryos comparison with that at 40 °C for 7 h. This is because the stronger
generated in the continuous aqueous phase tend to associate into hydrophobic interactions of PNIPAM/5AM at higher temperature
multi-chain aggregates, which in many instances stop growing allow these mesoglobules to favorably expel their water contents
once a certain dimension is achieved during the particle nucleation and subsequently form the tightly collapsed particle structure
process [12]. [49]. For the reaction conducted at the highest temperature
It is also interesting to note that, at constant a concentration (60 °C) over a period of ca. 1 h, the reaction mixture rapidly turned
(e.g., 0.5 wt%), the relative weight-average molecular weight to a milky white dispersion (observed at room temperature), fol-
(Mw) decreases with increasing AM concentration. In brief, free lowed by fast particle aggregation to form the network structure,
radical polymerization taking place in the heterogeneous reaction as shown in Fig. S1D.
system (e.g., emulsion polymerization) generally results in higher
polymer molecular weight as compared to the solution polymer- 3.3.2. Effects of GLYMO and TEOS concentrations
ization. For the copolymerization of NIPAM and AM using KPS as In this series of experiments, the reaction temperature and time
the initiator at 70 °C, the solution polymerization of the water- were kept constant at 50 °C and 5 h, respectively, and the weight
soluble AM with NIPAM becomes more important and, thus, percentage of PNIPAM based mesoglobule cores was also kept con-
polymer chains with lower molecular weight form. By contrast, stant at 0.5 wt%. In this manner, the effects of the weight ratio of
the molecular weight distribution (PDI in the range 1.7–2.0, char- GLYMO (as a coupling agent) to TEOS were investigated. The initial
acteristic of emulsion polymerization [47,48]) does not vary signif- compositions of PNIPAM, GLYMO and TEOS for the runs
icantly as the AM concentration is increased. In this study, PNIPAM, PNIPAM/5AM-111, PNIPAM/5AM-121, PNIPAM/5AM-112 and
PNIPAM/5AM and PNIPAM/10AM with a concentration of 0.5 wt% PNIPAM/5AM-122 can be found in Table 1. For example, the last
that balance the small enough particle size and high enough total three digits in PNIPAM/5AM-111 represent the weight ratio of PNI-
solids content were chosen for further silica encapsulation. PAM/5AM:GLYMO:TEOS = 1:1:1. For comparison, PNIPAM/5AM-
101 (in the absence of coupling agent) was also included in this
3.3. Encapsulation of PNIPAM based mesoglobule cores with silica study.
shells For the run PNIPAM/5AM-101, the reaction system of TEOS in
the presence of PNIPAM/5AM mesoglobules remained translucent
3.3.1. Effects of reaction temperature and time at room temperature. It was then postulated that the nucleation
First, PNIPAM/5AM mesoglobules encapsulated by silica shells and growth of silica nuclei occurred primarily in the continuous
via the sol-gel reaction of TEOS in the absence of coupling agent aqueous phase and encapsulation of PNIPAM/5AM mesoglobules
above its LCST (34.7 °C determined by DSC) were investigated. by silica (if present) did not contribute to the sol-gel reaction of
The variables chosen for study are the reaction temperature and TEOS to an appreciable extent. This scenario might explain why
time. It should be noted that the initial reaction mixture is trans- PNIPAM/5AM-101 appears translucent, which is a great contrast
parent at room temperature (ca. 25 °C < TLCST) and it becomes to the opaque PNIPAM/5AM-111 (see Fig. S2). Nevertheless, inde-
translucent or milky white (observed at room temperature) after pendent experiments are required to verify this speculation. This
the reaction was carried out above LCST over a period of the pre- is attributed to GLYMO possessing one reactive epoxide group
scribed time (e.g., 1, 3, 5 and 7 h). At 40 °C, it took about 5 h for and three methoxysilyl groups. The epoxide group can react with
the reaction mixture to become a stable milky dispersion the amide group of AM under basic conditions via the ring-
(observed at room temperature), indicating the successful forma- opening reaction (Scheme 2). On the other hand, the methoxysilyl
tion of hybrid organic core–inorganic shell particles. The resulting groups can be hydrolyzed to form reactive silanol groups, which
particles with dw = 132 nm and PDI = 1.09 obtained from silica can participate in the sol-gel reaction of TEOS [45]. This will then
encapsulation over a period of 7 h show characteristic of spherical promote the sol-gel reaction of TEOS taking place on the
shape and rough particle surface, as shown by SEM (Fig. S1A). As PNIPAM/5AM-111 mesoglobule cores via the bridging effect of
expected, the appearance of the reaction mixture at 50 °C over a GLYMO. Under the circumstances, these mesoglobule cores were
period of 3 h was transformed into milky white (observed at room capable of growing effectively with the aid of GLYMO during the
temperature). The values of dw and PDI for the sample prepared at sol-gelsol-gel reaction of TEOS and reached the ultimate dry
50 °C for 5 h are 92 nm and 1.06, respectively (Fig. S1B). This is fol- weight-average particle diameter of 92 nm determined by SEM
lowed by extensive agglomeration to form the network structure and TEM. Experiments with other compositions showed similar
540 N.-H. Cao-Luu et al. / Journal of Colloid and Interface Science 536 (2019) 536–547

(PNIPAM/5AM-112 vs. PNIPAM/5AM-122), the effect of doubling


the GLYMO concentration on the resultant hybrid particle size
becomes less pronounced (an increase in dw by a factor of only
ca. 29%). This is simply because the runs with higher TEOS concen-
tration provide the reaction system with more silica precursors
available for encapsulation of PNIPAM/5AM mesoglobule cores,
which compensates the effect of doubling the GLYMO concentra-
tion to some extent. It is also interesting to note that Fig. 1 and
the insets in Fig. 2 clearly show the PNIPAM encapsulated by tiny
Scheme 2. Schematic illustration of possible reactions between the epoxide group silica nanoparticles.
of GLYMO and the primary amine group of the AM unit and/or the secondary amine The significance of this result is that we have established an
group of the NIPAM unit.
effective approach for silica encapsulation of PNIPAM based
mesoglobule cores, which involves the incorporation of some
comonomer units (e.g., AM chosen for this study) into PNIPAM
silica encapsulation behavior, and some representative data shown cores in combination with an adequate coupling agent (e.g.,
in Figs. 1 (SEM) and 2 (TEM) and Table 3 (particle size data mea- GLYMO), which can simultaneously react to the amide group of
sured by SEM and TEM). the AM unit in PNIPAM and the silica precursors originating from
First, the weight-average PNIPAM/5AM core–silica shell particle TEOS. In this manner, desirable nucleation and growth of silica
diameter (dw) increases with increasing GLYMO or TEOS concen- embryos occurring primarily on the PNIPAM based mesoglobule
tration (Table 3). Taking the effect of GLYMO concentration as an core surfaces to form hybrid nanoparticles can thus be achieved.
example, at constant TEOS concentration, comparing Furthermore, this approach allows one to fine tune the particle size
PNIPAM/5AM-111 with PNIPAM/5AM-121 (or comparing and the thickness and morphology of the shell (i.e., the permeabil-
PNIPAM/5AM-112 with PNIPAM/5AM-122), dw is increased by a ity of the shell) of PNIPAM based mesoglobule core–silica shell
factor of ca. 57% {=[(146 92)/92 + (143 92)/92]/2, in which particles.
the first term in the numerator is based on the dw1 data, and the
second term based on the dw2 data in Table 3} (or 29%). This result 3.3.3. Effect of AM content
suggests that doubling the GLYMO concentration significantly The effects of AM content in PNIPAM based mesoglobule core–
increases the probability of nucleation and growth of silica silica shell particles on their physical properties are shown in
embryos surrounding the PNIPAM/5AM mesogloblue cores, Table 3 and Figs. 1A, 2A, 3 and 4. First, PNIPAM/5AM-111 shows
thereby leading to thicker silica shells and higher surface rough- the comparable particle size and particle size distribution deter-
ness (Fig. 2C and D). However, at higher TEOS concentration mined by SEM to PNIPAM-111 in the absence of AM (Table 3).

Fig. 1. Representative SEM images of PNIPAM/5AM core–silica shell particles obtained from the runs: (A) PNIPAM/5AM-111, (B) PNIPAM/5AM-121, (C) PNIPAM/5AM-112
and (D) PNIPAM/5AM-122.
N.-H. Cao-Luu et al. / Journal of Colloid and Interface Science 536 (2019) 536–547 541

Fig. 2. Representative TEM images of PNIPAM/5AM core–silica shell particles obtained from the runs: (A) PNIPAM/5AM-111, (B) PNIPAM/5AM-121, (C) PNIPAM/5AM-112
and (D) PNIPAM/5AM-122.

Table 3 ably due to the greatly enhanced nucleation and growth of silica on
Particle size and particle size distribution data for PNIPAM/5AM core–silica shell the PNIPAM-based mesoglobule surfaces containing very high con-
particles obtained from the runs with different weight ratios of PNIPAM/5AM:
centration of AM units. This will then greatly promote the sol-gel
GLYMO:TEOS.
reaction between the methoxysilyl groups of GLYMO and the
Samples dw1a (nm) PDI a
dw2 (nm) a
PDI a
ethoxysilyl groups of TEOS on the polymeric core surfaces, thereby
PNIPAM-111 90 1.13 – – increasing the probability of forming oblong nanoparticles. Never-
PNIPAM/5AM-111 92 1.06 92 1.07 theless, additional independent experiments are required to verify
PNIPAM/5AM-121 146 1.15 143 1.06
this speculation. Nevertheless, the abundant AM-rich chain
PNIPAM/5AM-112 143 1.11 153 1.10
PNIPAM/5AM-122 176 1.14 206 1.23 segments in PNIPAM/10AM mesoglobule cores might play an
important role in the evolution of such particle morphology with
a
dw1 and dw2 are the weight-average particle diameter and polydispersity index
the minimal interfacial free energy. Finally, the TEM images in
(PDI = dw/dn, where dn is the number-average particle diameter) determined by
SEM and TEM, respectively.
Figs. 2A and 4A show the contrast difference between the core
and shell phases, indicating the formation of PNIPAM based
mesoglobule core–silica shell particle morphology for PNIPAM-
These two products are characterized by relatively spherical shape 111 and PNIPAM/5AM-111. Unfortunately, the ambiguous
and raspberry morphology (Figs. 1A vs. 3A (SEM) and 2A vs. 4A boundary between the core and shell phases prevented one from
(TEM)). On the other hand, PNIPAM/10AM-111 is characterized estimating the shell thickness. Fig. 4B also illustrates the oblong
by mixed oblong (with the longitudinal length being about two and spherical particles with the core phase inside for
to three times of the latitudinal length) and spherical particle mor- PNIPAM/10AM-111.
phology with relatively smooth particle surface (Fig. 4B). It is also
interesting to note that the transition from the spherical particle
morphology to oblong shape with a higher aspect ratio for 3.4. Characterization of PNIPAM-based core–silica shell particles
PNIPAM/10AM-111 is quite similar to those surfactant systems
where the transition from spherical micelles to oblong ones were The hydrodynamic hybrid particle diameter (d) determined by
reported [50,51] though they may be attributed to different mech- DLS as a function of temperature (T) profiles is shown in Fig. 5A.
anisms. In this case, such a particle morphology transition is prob- By taking the first derivative of the d vs. T curve, the LCST was iden-
542 N.-H. Cao-Luu et al. / Journal of Colloid and Interface Science 536 (2019) 536–547

Fig. 3. Representative SEM images of PNIPAM-based core–silica shell particles obtained from the runs: (A) PNIPAM-111 and (B) PNIPAM/10AM-111.

Fig. 4. Representative TEM images of PNIPAM-based core–silica shell particles obtained from the runs: (A) PNIPAM-111 and (B) PNIPAM/10AM-111.

tified as the temperature at which point the minimum occurs in heating rate of 2 °C min 1. The temperature corresponding to the
the d d/d T vs. T curve, as shown in Fig. 5B. minimal endothermic peak in Fig. 6A and B was defined as the LCST
Fig. 5A demonstrates that, just like PNIPAM based mesoglobules, phase transition temperature [53], and the results listed in Table 4.
the hybrid particles shrink significantly with increasing tempera- The LCST data determined by DSC show the same trend as those
ture primarily due to the permeability of the silica layer [52]. This measured by DLS (Table 4). Furthermore, both incorporation of
is attributed to the shell structure of silica clusters containing AM units into PNIPAM and the encapsulation of PNIPAM-based
numerous slots that allow water molecules in PNIPAM based mesoglobules with silica tend to reduce the thermo-sensitivity of
mesoglobule cores to be expelled through the silica shells to the con- PNIPAM based core–silica shell particles. The total heat (DH) corre-
tinuous aqueous phase as the temperature is increased. Moreover, sponding to the DSC peak related to breaking of hydrogen bonds
these slots in the silica shell provide the hybrid particle with the between the polymeric repeating units and the surrounding water
buffering space for tightening the silica clusters during the phase molecules is also included in Table 4. As expected, the DH of PNI-
transition from the swelling state to the shrinking state. This is the PAM based mesoglobules is significantly higher than that of the
reason why PNIPAM based mesoglobule core–silica shell particles counterpart of PNIPAM based core–silica shell particles (e.g., PNIPA-
still can exhibit a distinct LCST phase transition behavior. The LCST M/5AM vs. PNIPAM/5AM-111). This can be attributed to the fact
determined by DLS shifts slightly to higher temperature with that AM units in the mesoglobule cores are not thermo-sensitive
increasing AM concentration due to the higher degree of hydration to an appreciable extent and the dilution effect of the inorganic silica
around the AM-rich chain segments, which can form much content. In addition, the DH data for both PNIPAM based mesoglob-
stronger hydrogen bonding with water molecules. The LCST data ules (e.g., PNIPAM/5AM vs. PNIPAM/10AM) and PNIPAM based
for PNIPAM-111, PNIPAM/5AM-111, PNIPAM/5AM-121, PNIPAM/ core–silica shell particles (e.g., PNIPAM/5AM-111 vs. PNIPAM/
5AM-112, PNIPAM/5AM-122 and PNIPAM/10AM-111 obtained 10AM-111) show that DH decreases with increasing AM content.
from the DLS technique are summarized in Table 4. It is noteworthy PNIPAM based mesoglobule core–silica shell particles of
that, at constant T, the hydrodynamic hybrid particle size (d) PNIPAM-111, PNIPAM/5AM-111, PNIPAM/5AM-121, PNIPAM/
increases with increasing AM, GLYMO and TEOS concentration, 5AM-112, PNIPAM/5AM-122 and PNIPAM/10AM-111 were further
which is consistent with the results attained from both SEM and characterized by FTIR. For comparison, the samples of GLYMO/
TEM (Table 3). TEOS-11 (the weight ratio of GLYMO:TEOS = 1:1, Fig. 7(VII)) and
DSC was also used to study the thermo-responsive property of PNIPAM/10AM (Fig. 7(VIII)) were also included in this study.
PNIPAM based mesoglobules (Fig. 6A) and PNIPAM based core–sil- First, all hybrid particles (Fig. 7(I)–(VI)) show similar character-
ica shell particles (Fig. 6B) in the temperature range 25–60 °C at a istic absorption peaks to PNIPAM/10AM in the wavenumber range
N.-H. Cao-Luu et al. / Journal of Colloid and Interface Science 536 (2019) 536–547 543

Fig. 5. (A) Hydrodynamic hybrid particle diameter (d) as a function of temperature (T) profiles and (B) d d/d T vs. T curves: ( ) PNIPAM-111, (r) PNIPAM/5AM-111, (.)
PNIPAM/10AM-111, ( ) PNIPAM/5AM-112, (d) PNIPAM/5AM-121, (j) PNIPAM/5AM-122.

Table 4 1351–4000 cm 1. The peaks at 1351 and 1458 cm 1 are attributed


Some LCST phase transition data for PNIPAM based mesoglobules and PNIPAM-based to vibration of isopropyl group, and those at 1602 and 3321 cm 1
mesoglobule core–silica shell particles determined by the DLS and DSC techniques.
due to the NAH stretching of the amide group in PNIPAM based
Samples LCSTa (°C) DHb (J g 1) chain [21]. The peak at 1730 cm 1 can be attributed to the intrinsic
DLS DSC vibrational band of carbonyl group of NIPAM and AM units [21]. As
to the triplet peaks at 2876, 2935 and 2972 cm 1, they are assigned
PNIPAM – 32.1 ± 0.1 42.9 ± 0.7
PNIPAM-111 32.0 ± 0.4 32.0 ± 0.1 18.6 ± 0.4 to CAH bands [54]. The shoulder appearing at 3082 cm 1 is
PNIPAM/5AM – 34.7 ± 0.1 34.2 ± 0.9 ascribed to the NAH stretching vibration of the AM unit [55]. On
PNIPAM/5AM-111 35.0 ± 0.3 33.4 ± 0.2 16.9 ± 1.2 the other hand, the peaks with the wavenumber smaller than
PNIPAM/5AM-121 35.3 ± 0.4 33.7 ± 0.3 16.4 ± 1.3
1351 cm 1, the characteristic bands from 1056 to 1250 cm 1 in
PNIPAM/5AM-112 35.2 ± 0.3 34.9 ± 0.3 12.1 ± 0.8
PNIPAM/5AM-122 35.0 ± 0.5 34.8 ± 0.2 10.6 ± 0.9
the GLYMO/TEOS-11 spectrum representing the SiAOASi network
PNIPAM/10AM – 37.2 ± 0.1 27.6 ± 0.5 structure are also observed in PNIPAM-111, PNIPAM/5AM-111,
PNIPAM/10AM-111 37.0 ± 0.5 35.6 ± 0.3 10.2 ± 0.4 PNIPAM/5AM-112, PNIPAM/5AM-121, PNIPAM/5AM-122 and
a
The LCST was identified by taking the first derivative of the average hydrody-
PNIPAM/10AM-111, in which the ring-opening reaction of the
namic diameter vs. T (DLS) or the minimum of the endothermic peak (DSC). The epoxide group of GLYMO by the amide group of AM unit in PNIPAM
standard error was calculated based on three runs for each sample. based chain and hydrolysis and condensation between the
b
The total heat (DH) for the phase transition was determined from the integral methoxysilyl group of GLYMO and the ethoxysilyl group of TEOS
area under the endothermic peak of the DSC curve.
to form the Si-O-Si network structure between the organic core

Fig. 6. Non-isothermal DSC profiles at a heating rate of 2 °C min 1 (a) PNIPAM, (b) PNIPAM/5AM, (c) PNIPAM/10AM (d) PNIPAM-111, (e) PNIPAM/5AM-111, (f) PNIPAM/5AM-
121, (g) PNIPAM/5AM-112, (h) PNIPAM/5AM-122, and (i) PNIPAM/10AM-111.
544 N.-H. Cao-Luu et al. / Journal of Colloid and Interface Science 536 (2019) 536–547

Fig. 8. TGA profiles for (I) PNIPAM-111, (II) PNIPAM/5AM-111, (III) PNIPAM/5AM-
Fig. 7. FTIR spectra for (I) PNIPAM-111, (II) PNIPAM/5AM-111, (III) PNIPAM/5AM-
112, (IV) PNIPAM/5AM-121, (V) PNIPAM/5AM-122, (VI) PNIPAM/10AM-111, and
112, (IV) PNIPAM/5AM-121, (V) PNIPAM/5AM-122, (VI) PNIPAM/10AM-111, (VII)
(VII) GLYMO/TEOS-11.
GLYMO/TEOS-11, and (VIII) PNIPAM/10AM.

and inorganic shell phases. In addition, there are weak peaks at 4. Conclusion
about 451, 692 and 787 cm 1, which correspond to the SiAOAC
bond mostly present in the interfacial region between the organic The way to encapsulate PNIPAM mesoglobule cores by silica
core and the inorganic shell phases [56]. shells significantly affects the resultant hybrid nanoparticle struc-
The effect of AM, GLYMO and TEOS content in PNIPAM based tures and their physicochemical properties. This work presented a
mesoglobule core–silica shell on their thermal properties mea- novel approach to prepare thermo-responsive PNIPAM-based
sured by TGA is shown in Fig. 8. The residue at 800 °C increases mesoglobule core–silica shell nanoparticles with narrow particle
gradually from 37% to 45% with the AM content being increased size distribution. The key of this unique technique is to incorporate
from 0 to 10 wt%. This result provides supporting evidence for some acrylamide (AM) units into PNIPAM chain in combination
the primary amide group of AM is much more reactive to with 3-glycidyloxypropyltrimethoxysilane (GLYMO, as a coupling
GLYMO than the secondary amide group of NIPAM. Moreover, agent), in which the –NH2 group of the AM unit in mesoglobule
the residue of hybrid samples also increases from 41% to 49% core can react with the epoxide group of GLYMO and the four
with the weight ratio of GLYMO (or TEOS) to PNIPAM/5AM being ethoxysilyl groups of TEOS (silica precursor) and the three
increased from 1 to 2. This is most likely due to the increased methoxysilyl groups of GLYMO can simultaneously undergo sol-
probability of silica particle nucleation and growth on PNIPAM gel reaction to form silica. Thus, the probability of silica particle
based mesoglobule cores when the AM concentration is nucleation and growth on PNIPAM-based mesoglobule core sur-
increased. faces is greatly increased, thereby leading to highly desirable
Finally, the morphology of PNIPAM-based mesoglobule core– hybrid organic core–inorganic shell particle morphology.
silica shell particles were confirmed by the SEM images The particle morphology including the shape, core size, shell
(Fig. 9). The samples were calcined in TGA cell with the gradu- thickness and surface roughness can be controlled primarily by
ally increased temperature from 500 to 800 °C over a period of the AM content and the weight ratio of PNIPAM/AM:GLYMO:TEOS.
30 min, followed by remaining at 800 °C for 30 min in air. For For example, the PNIPAM/5AM core (i.e., containing 5 wt% AM)–sil-
both PNIPAM-111 (Fig. 9A) and PNIPAM/5AM-111 (Fig. 9B), the ica shell particle size increases with increasing GLYMO (or TEOS)
inorganic silica shells are not significantly affected by calcina- concentration in the silica encapsulation process. For the series of
tion, as evidenced by the hollow particle structure, in which experiments with the weight ratio of PNIPAM/nAM:GLYMO:
the organic PNIPAM based mesoglobule cores can be effectively TEOS = 1:1:1 and n = 0, 5 and 10 wt%, PNIPAM/5AM-111 shows
decomposed and eventually removed. It is noteworthy that both comparable particle size (dw = 92 nm) and particle size distribution
samples with hollow particle structure show comparable particle (PDI = 1.13) to PNIPAM-111 in the absence of AM. Furthermore,
sizes to those of the original hybrid core–shell particles. How- these two products are characterized by relatively spherical shape
ever, PNIPAM/5AM-111 exhibits distinct hollow silica particles, and raspberry morphology. On the other hand, PNIPAM/10AM-111
while PNIPAM-111 particles seem to undergo aggregation to shows mixed oblong (with the longitudinal length being about two
form the silica clumps containing hollow hemispherical holes. to three times of the latitudinal length) and spherical particle mor-
In contrast, PNIPAM/10AM-111 with the highest AM content in phology with relatively smooth particle surface. The LCST of
organic mesoglobule cores does not maintain the core–shell PNIPAM-based mesoglobules increases with increasing AM con-
structure after calcination (Fig. 9C). In cases of samples with tent. This is due to the fact that the more hydrophilic AM unit in
increasing GLYMO or TEOS concentration (Fig. 9D–F), although PNIPAM-based chain makes it less sensitive to changes in temper-
they do not show hollow spherical holes due to their thicker ature. In addition, encapsulation of PNIPAM-based mesoglobules
shells, they have the equivalent particle size with original hybrid with silica reduces the thermo-sensitivity of PNIPAM-based core–
core–shell particles as well as the significant weight reduction silica shell nanoparticles.
during the calcination. It is implied that the hollow holes can It was reported that PNIPAM core–silica shell particles were
be contained inside the thick and rigid silica layers. prepared by encapsulation of silica (or precursor of silica) onto
N.-H. Cao-Luu et al. / Journal of Colloid and Interface Science 536 (2019) 536–547 545

Fig. 9. Representative SEM images of PNIPAM-based core–silica shell particles obtained from the runs: (A) PNIPAM-111, (B) PNIPAM/5AM-111, (C) PNIPAM/10AM-111, (D)
PNIPAM/5AM-121, (E) PNIPAM/5AM-112, and (F) PNIPAM/5AM-122 after calcination.

crosslinked PNIPAM microgel particles in the presence of surfac- loidal and interfacial phenomena are of great challenge to the future
tant [19–21]. The resultant hybrid particle sizes are about 2.5–5 work.
times larger than those of PNIPAM-based mesoglobule core–silica
shell nanoparticles prepared in this work. Such a characteristic
Acknowledgements
dimension of crosslinked PNIPAM microgel core–silica shell parti-
cles greatly limits their use in drug carrier application although
Financial support from Ministry of Science and Technology,
surfactant was used in an attempt to improve the colloidal stability
Taiwan is gratefully acknowledged.
and reduce the particle size.
How to fine tune the LCST of PNIPAM-based core–silica shell
nanoparticles without detriment to their thermo-sensitivity and Appendix A. Supplementary material
form well defined pore structure of the silica shell is an important
issue in biomedical applications. These subjects dealing with the Supplementary data to this article can be found online at
integration of organic and inorganic materials and the related col- https://doi.org/10.1016/j.jcis.2018.10.091.
546 N.-H. Cao-Luu et al. / Journal of Colloid and Interface Science 536 (2019) 536–547

References [25] R. Singh, J.W. Lillard, Nanoparticle-based targeted drug delivery, Exp. Mol.
Pathol. 86 (2009) 215–223, https://doi.org/10.1016/j.yexmp.2008.12.004.
[26] T. Sun, Y.S. Zhang, B. Pang, D.C. Hyun, M. Yang, Y. Xia, Engineered nanoparticles
[1] A. Gandhi, A. Paul, S.O. Sen, K.K. Sen, Studies on thermoresponsive polymers:
for drug delivery in cancer therapy, Angew. Chem. Int. Ed. 53 (2014) 12320–
phase behaviour, drug delivery and biomedical applications, Asian J. Pharm.
12364, https://doi.org/10.1002/anie.201403036.
Sci. 10 (2015) 99–107, https://doi.org/10.1016/j.ajps.2014.08.010.
[27] K. Loomis, K. McNeeley, R.V. Bellamkonda, Nanoparticles with targeting,
[2] Y. Kamachi, B.P. Bastakoti, S.M. Alshehri, N. Miyamoto, T. Nakato, Y. Yamauchi,
triggered release, and imaging functionality for cancer applications, Soft
Thermo-responsive hydrogels containing mesoporous silica toward controlled
Matter 7 (2011) 839–856, https://doi.org/10.1039/C0SM00534G.
and sustainable releases, Mater Lett. 168 (2016) 176–179, https://doi.org/
[28] A.H. Faraji, P. Wipf, Nanoparticles in cellular drug delivery, Bioorg. Med. Chem.
10.1016/j.matlet.2015.12.132.
17 (2009) 2950–2962, https://doi.org/10.5772/51384.
[3] J. Čejková, J. Hanuš, F. Štěpánek, Investigation of internal microstructure and
[29] M. Danaei, M. Dehghankhold, S. Ataei, F.H. Davarani, R. Javanmard, A. Dokhani,
thermo-responsive properties of composite PNIPAM/silica microcapsules, J.
S. Khorasani, M.R. Mozafari, Impact of particle size and polydispersity index on
Colloid Interface Sci. 346 (2010) 352–360, https://doi.org/10.1016/j.
the clinical applications of lipidic nanocarrier systems, Pharmaceutics 10
jcis.2010.02.060.
(2018) 57–74, https://doi.org/10.3390/pharmaceutics10020057.
[4] H.L. Lim, Y. Hwang, M. Kar, S. Varghese, Smart hydrogels as functional
[30] L.Y.T. Chou, K. Ming, W.C.W. Chan, Strategies for the intracellular delivery of
biomimetic systems, Biomater. Sci. 2 (2014) 603–618, https://doi.org/10.1039/
nanoparticles, Chem. Soc. Rev. 40 (2011) 233–245, https://doi.org/10.1039/
c3bm60288e.
c0cs00003e.
[5] P. Schattling, F.D. Jochuma, P. Theato, Multi-stimuli responsive polymers – the
[31] Z. Cai, Y. Wang, L.J. Zhu, Z.Q. Liu, Nanocarriers: a general strategy for
all-in-one talents, Polym. Chem. 5 (2014) 25–36, https://doi.org/10.1039/
enhancement of oral bioavailability of poorly absorbed or pre-systemically
C3PY00880K.
metabolized drugs, Curr. Drug Metab. 11 (2010) 197–207, https://doi.org/
[6] T. López-León, J.L. Ortega-Vinuesa, D. Bastos-González, A. Elaissari, Thermally
10.2174/138920010791110836.
sensitive reversible microgels formed by poly(N-Isopropylacrylamide) charged
[32] M. Kwok, Z. Li, T. Ngai, Controlling the synthesis and characterization of
chains: a Hofmeister effect study, J. Colloid Interface Sci. 426 (2014) 300–307,
micrometer-sized PNIPAM microgels with tailored morphologies, Langmuir 29
https://doi.org/10.1016/j.jcis.2014.04.020.
(2013) 9581–9591, https://doi.org/10.1021/la402062t.
[7] T.E.D. Oliveira, D. Mukherji, K. Kremer, P.A. Netz, Effects of stereochemistry and
[33] K. Tauer, D. Gau, S. Schulze, A. Völkel, R. Dimova, Thermal property changes of
copolymerization on the LCST of PNIPAm, J. Chem. Phys. (2017), https://doi.
poly(N-isopropylacrylamide) microgel particles and block copolymers, Colloid
org/10.1063/1.4974165.
Polym. Sci. 287 (2009) 299–312, https://doi.org/10.1007/s00396-008-1984-x.
[8] K. Shiraga, H. Naito, T. Suzuki, N. Kondo, Y. Ogawa, Hydration and hydrogen
[34] B. Wedel, T. Brandel, J. Bookhold, T. Hellweg, Role of anionic surfactants in the
bond network of water during the coil-to-globule transition in poly(N–
synthesis of smart microgels based on different acrylamides, ACS Omega 2
isopropylacrylamide) aqueous solution at cloud point temperature, J. Phys.
(2017) 84–90, https://doi.org/10.1021/acsomega.6b00424.
Chem. B 119 (2015) 5576–5587, https://doi.org/10.1021/acs.jpcb.5b01021.
[35] T. Still, K. Chen, A.M. Alsayed, K.B. Aptowicz, A.G. Yodh, Synthesis of
[9] Y. Kang, H. Joo, J.S. Kim, Collapse swelling transitions of a thermo- responsive,
micrometer-size poly(N-isopropylacrylamide) microgel particles with
single poly(N–isopropylacrylamide) chain in water, J. Phys. Chem. B 120
homogeneous crosslinker density and diameter control, J. Colloid Interface
(2016) 13184–13192, https://doi.org/10.1021/acs.jpcb.6b09165.
Sci. 405 (2013) 96–102, https://doi.org/10.1016/j.jcis.2013.05.042.
[10] L.J. Abbott, A.K. Tucker, M.J. Stevens, Single chain structure of a poly(N–
[36] H. Shimizu, R. Wada, M. Okabe, Preparation and characterization of
isopropylacrylamide) surfactant in water, J. Phys. Chem. B 119 (2015) 3837–
micrometer-sized poly(N-isopropylacrylamide) hydrogel particles, Polym. J.
3845, https://doi.org/10.1021/jp511398q.
41 (2009) 771–777, https://doi.org/10.1295/polymj.PJ2009039.
[11] Y. Yan, L. Huang, Q. Zhang, H. Zhou, Concentration effect on aggregation and
[37] A. Lee, H.-Y. Tsai, M.Z. Yates, Steric stabilization of thermally responsive N-
dissolution behavior of poly(N-isopropylacrylamide) in water, J. Apll. Polym.
isopropylacrylamide particles by poly(vinyl alcohol), Langmuir 26 (2010)
Sci. 10 (2015) 41669–41676, https://doi.org/10.1002/app.41669.
18055–18060, https://doi.org/10.1021/la200540c.
[12] K. Liu, P. Pan, Y. Bao, Synthesis, micellization, and thermally-induced
[38] F. Meunier, A. Elaissari, C. Pichot, Preparation and characterization of cationic
macroscopic micelle aggregation of poly(vinylchloride)-g-poly(N-
poly(n-isopropylacrylamide) copolymer latexes, Polym. Adv. Technol. 6 (1995)
isopropylacrylamide) amphiphilic copolymer, RSC Adv. 5 (2015) 94582–
489–496, https://doi.org/10.1002/pat.1995.220060710.
94590, https://doi.org/10.1039/C5RA16726D.
[39] R. Chen, X. Jin, X. Zhu, Investigation of the formation process of PNIPAM-based
[13] I. Bischofberger, V. Trappe, New aspects in the phase behaviour of poly-N-
ionic microgels, ACS Omega 2 (2017) 8788–8793, https://doi.org/10.1021/
isopropyl acrylamide: systematic temperature dependent shrinking of
acsomega.7b01624.
PNiPAM assemblies well beyond the LCST, Sci. Rep. 5 (2015), https://doi.org/
[40] Z. Guangzhao, W. Chi, Folding and formation of mesoglobules in dilute
10.1038/SREP15520.
copolymer solutions, ADV Polym Sci 195 (2006) 101–176, https://doi.org/
[14] P. Kujawa, V. Aseyev, H. Tenhu, F.M. Winnik, Temperature-sensitive properties
10.1007/12_050.
of Poly(N-isopropylacrylamide) mesoglobules formed in dilute aqueous
[41] S. Penga, C. Wu, Influence of initial chain conformation on the formation of
solutions heated above their demixing point, Macromolecules 39 (2006)
mesoglobule phase in a dilute heteropolymer solution, Polymer 44 (2003)
7686–7693, https://doi.org/10.1021/ma061604b.
1089–1093, https://doi.org/10.1016/S0032-3861(02)00860-1.
[15] B. Trzebicka, E. Haladjova, Ł. Otulakowski, N. Oleszko, W. Wałach, M. Libera, S.
[42] N. Toncheva-Moncheva, P. Dimitrov, C.B. Tsvetanov, B. Robak, B. Trzebicka, A.
Rangelov, A. Dworak, Hybrid nanoparticles obtained from mixed mesoglobules,
Dworak, S. Rangelov, Formation of mesoglobules in aqueous media from
Polymer 68 (2015) 65–73, https://doi.org/10.1016/j.polymer.2015.04.085.
thermo-sensitive poly(ethoxytriethyleneglycolacrylate), Polym. Bull. 67
[16] V. Aseyev, S. Hietala, A. Laukkanen, M. Nuopponen, O. Confortini, F.E.D. Prezb,
(2011) 1335–1346, https://doi.org/10.1007/s00289-011-0545-5.
H. Tenhu, Mesoglobules of thermoresponsive polymers in dilute aqueous
[43] X. Guillory, A. Tessier, G.-O. Gratien, P. Weiss, S. Colliec-Jouault, D. Dubreuil, J.
solutions above the LCST, Polymer 46 (2005) 7118–7131, https://doi.org/
Lebretonc, J.L. Bideaua, Glycidyl alkoxysilane reactivities towards simple
10.1016/j.polymer.2005.05.097.
nucleophiles in organic media for improved molecular structure definition in
[17] N. Nun, S. Hinrichs, M.A. Schroer, D. Sheyfer, G. Grübel, B. Fischer, Tuning the
hybrid materials, RSC Adv. 6 (2016) 74087–74099, https://doi.org/10.1039/
size of thermoresponsive poly(N-Isopropyl Acrylamide) grafted silica
C6RA01658H.
microgels, Gels (2017), https://doi.org/10.3390/GELS3030034.
[44] D.-Y. Kim, S.H. Jin, S.-G. Jeong, B. Lee, K.-K. Kang, C.-S. Lee, Microfuidic
[18] H. Byun, J. Hu, P. Pakawanit, L. Srisombat, J.H. Kim, Polymer particles filled
preparation of monodisperse polymeric microspheres coated with silica
with multiple colloidal silica via in situ sol-gel process and their thermal
nanoparticles, Sci. Rep. (2018), https://doi.org/10.1038/S41598-018-26829-Z.
property, Nanotechnology (2017), https://doi.org/10.1088/0957-4484/28/2/
[45] X. Wang, J.K. Gillham, Competitive primary amine/ epoxy and secondary amine/
025601.
epoxy reactions: effect on the isothermal time-to-vitrify, J. Appl. Polym. Sci. 43
[19] J.F. Dechezelles, V. Malik, J.J. Crassous, P. Schurtenberger, Hybrid raspberry
(1991) 2267–2277, https://doi.org/10.1002/app.1991.070431216.
microgels with tunable thermoresponsive behaviour, Soft Matter 9 (2013)
[46] C.E. Estridge, The effects of competitive primary and secondary amine
2798–2802, https://doi.org/10.1039/C3SM27433K.
reactivity on the structural evolution and properties of an epoxy thermoset
[20] L. Wang, S.A. Asher, Fabrication of silica shell photonic crystals through
resin during cure: a molecular dynamics study, Polymer 141 (2018) 12–20,
flexible core templates, Chem. Mater. 21 (2009) 4608–4613, https://doi.org/
https://doi.org/10.1016/j.polymer.2018.02.062.
10.1021/cm901666b.
[47] G.M. Saidel, S. Katz, Emulsion polymerization: a stochastic approach to the
[21] X. Hu, X. Hao, Y. Wu, J. Zhang, X. Zhang, P.C. Wang, G. Zou, X.J. Liang,
polymer size distribution, J. Polym. Sci. Part C 27 (1969) 149–169, https://doi.
Multifunctional hybrid silica nanoparticles for controlled doxorubicin loading
org/10.1002/polc.5070270112.
and release with thermal and pH dually response, J. Mater. Chem. B 1 (2013)
[48] C.S. Chern, Principles and Applications of Emulsion Polymerization, Wiley,
1109–1118, https://doi.org/10.1039/c2tb00223j.
2008, pp. 120–122. doi: 978-0-470-12431-4.
[22] L. Duan, M. Chen, S. Zhou, L. Wu, Synthesis and characterization of poly(N-
[49] E. Haladjova, N. Toncheva-Moncheva, M.D. Apostolova, B. Trzebicka, A.
isopropylacrylamide)/silica composite microspheres via inverse pickering
Dworak, P. Petrov, I. Dimitrov, S. Rangelov, C.B. Tsvetanov, Polymeric
suspension polymerization, Langmuir 25 (2009) 3467–3472, https://doi.org/
nanoparticle engineering: from temperature-responsive polymer
10.1021/la8041617.
mesoglobules to gene delivery systems, Biomacromolecules 15 (2014) 4377–
[23] W.H. Blackburn, L.A. Lyon, Size controlled synthesis of monodispersed, Core/
4395, https://doi.org/10.1021/bm501194g.
Shell Nanogels, Colloid Polym Sci. 286 (2008) 563–569, https://doi.org/
[50] B. Herzog, Micelle shape and capacity of solubilization, Progr. Colloid Polym.
10.1007/s00396-007-1805-7.
Sci. 84 (1991) 325–326, https://doi.org/10.1007/BFb0115995.
[24] A. Albanese, P.S. Tang, W.C.W. Chan, The effect of nanoparticle size, shape, and
[51] A. Abe, T. Imae, S. Ikeda, Solubilization properties of aqueous solution of
surface chemistry on biological systems, Annu. Rev. Biomed. Eng. 14 (2012) 1–
alkyltrimethylammonium halids toward a water-insoluble dye, Colloid Polym.
16, https://doi.org/10.1146/annurev-bioeng-071811-150124.
Sci. 265 (1987) 637–645, https://doi.org/10.1007/BF01412780.
N.-H. Cao-Luu et al. / Journal of Colloid and Interface Science 536 (2019) 536–547 547

[52] R. Rogoziński, P. Karasiński, C. Tyszkiewicz, Silica layers produced by the sol- quantitative grafting on mesoporous silica, J. Appl. Polym. Sci. 133 (2016)
gel method as dielectric masks in ion exchange processes, Photon. Lett. Pol. 51 44181–44189, https://doi.org/10.1002/app.44181.
(2012) 137–139, https://doi.org/10.4302/plp.2012.4.06. [55] T. Cai, Z. Yang, H. Li, H. Yang, A. Li, R. Cheng, Effect of hydrolysis degree of
[53] M. Schonhoff, A. Larsson, P.B. Welzel, D. Kuckling, Thermoreversible polymers hydrolyzed polyacrylamide grafted carboxymethyl cellulose on dye removal
adsorbed to colloidal silica: a 1HNMR and DSC study of the phase transition in efficiency, Cellulose 20 (2013) 2605–2614, https://doi.org/10.1007/s10570-
confined geometry, J. Phys. Chem. B 106 (2002) 7725–7728, https://doi.org/ 013-9987-2.
10.1021/jp015538l. [56] J.V.G. Tinio, K.T. Simfroso, A.D.M.V. Peguit, R.T. Candidato Jr., Influence of OH –
[54] S.A. Jadhav, V. Brunella, I. Miletto, G. Berlier, D. Scalarone, Synthesis of poly(N- Ion concentration pn the surface morphology of ZnO-SiO2 nanostructure, J.
isopropylacrylamide) by distillation precipitation polymerization and Nanotechnol. 1 (2015) 1–7, https://doi.org/10.1155/2015/686021.

S-ar putea să vă placă și