Soft, 2013, 2, 19-26
Published Online December 2013 (http://www.scirp.org/journal/soft)
http://dx.doi.org/10.4236/soft.2013.24005
Open Access Soft
Thermo- and pH-Responsive Hydrogels Based on
N-Isopropylacrylamide and Allylamine Copolymers
Victoria Konovalova1, Yuri Samchenko2, Ganna Pobigai1, Anatoly Burban1, Zoya Ulberg2
1Department of Chemistry, National University of Kiev-Mohyla Academy, Kiev, Ukraine
2Institute of Biocolloidal Chemistry, National Academy of Sciences of Ukraine, Kiev, Ukraine
Email: vita@ukma.kiev.ua
Received October 17, 2013; revised November 19, 2013; accepted December 5, 2013
Copyright © 2013 Victoria Konovalova et al. This is an open access article distributed under the Creative Commons Attribution
License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original work is properly cited.
ABSTRACT
The thermo- and pH-responsive hydrogels were synthesized via copolymerization of N-isopropylacrylamide and al-
lylamine hydrochloride monomers. The equilibrium swelling of the hydrogels was studied as a function of temperature
and pH in aqueous solutions. It was shown that controlled alteration of the hydrogel phase transition temperature can be
achieved by changing their composition and pH of the environment. Increase in content of hydrophilic allylamine from
10 to 60 wt% in monomer mixture causes a shift of the phase transition temperature from 35˚C to 47˚C. Hydrogels with
N-isopropylacrylamide/allylamine hydrochloride mass ratio of 3:2 show the highest pH-response. Values of average
molecular weight between polymer cross-links, c
M
, and Flory parameter, χ, were calculated using temperature de-
pendences of the equilibrium swelling of the synthesized hydrogel.
Keywords: Thermo-Responsive Hydrogels; Copolymer Hydrogels; Allylamine Hydrochloride;
N-Isopropyl-Acrylamide; The Parameter of Flory
1. Introduction
Polymer hydrogels are one of the most promising bio-
medical materials in controlled drug delivery, immuno-
assays, biotechnology, separation processes, etc. [1-3].
Stimuli-responsive hydrogels are of special interest ena-
bling a variety of approaches in the development of smart
materials. Thermo-sensitive hydrogels undergo an abrupt
structural transition from the swollen to collapsed state at
transition temperature (Tp). Important properties of thermo-
sensitive hydrogels are temperature range, at which the
network collapses, as well as the rate of the system re-
sponse to the outer stimulus.
Poly-N-isopropylacrylanide (PNIPA) is one of the
most studied temperature-sensitive polymers [3-7]. At
room temperature, polymer chains of (co)polymers based
on NIPA adopt expanded conformation with high affinity
to water due to intensive hydrogen bonding. Solutions
are homogenous and clear [8]. As the temperature in-
creases, molecular vibrations weaken H-bonds and water
molecules migrate from the polymer. Hydrophobicity of
the polymer chains increases and eventually this change
initiates phase separation due to polymer aggregation and
precipitation when temperature is higher than Tp [7]. The
phase transition is reversible and is initiated by the slight
temperature changes in both directions within Tp region.
Similar transition takes place in hydrogels, which are
cross-linked materials. Phase transition of poly-NIPA-
based hydrogels occurs in a narrow temperature range
between 32˚C and 34˚C. This temperature interval is very
close to the temperature of a human body, though not
reaching it, which limits some biomedical applications of
thermo-sensitive hydrogels. Other disadvantages of the
poly-NIPA-based hydrogels include their poor mechanic-
cal strength and weak pH sensitivity.
There are several ways to influence the transition
temperature by controlling the hydrophilic-hydrophobic
balance of a macromolecular system. This is achievable
by changes in composition of both aqueous media and
polymer molecular structure. They include variations in
the solution ionic strength [4], introduction of surfactants
[5] or changes in ratio of hydrophilic and hydrophobic
units in the hydrogel polymer structure [6,7]. These fac-
tors also influence the temperature range of the hydrogel
thermo-stimulated response [9]. Both an abrupt discrete
[10,11] and discontinuous [12] decrease in swelling de-
gree at temperature rise are described in literature. Thus
V. KONOVALOVA ET AL.
20
it was shown that continuous phase transition transforms
into discrete one with the decrease in cross-linking den-
sity of NIPA-based hydrogels [13]. Non-ionized NIPA-
based gels are characterized by continuous phase transi-
tion, while phase transformation of ionized hydrogels is
of discrete nature [14]. On the contrary, only continuous
phase transition was observed for NIPA/acrylamide (AA)
hydrogels [15]. Such a behavior was explained by the
weakening of the aggregation of dehydrated NIPA chains
at high temperatures due to the presence of highly hy-
drophilic acrylamide moieties. Several publications [9,16]
deal with mathematical modeling of the above-mentioned
process, though this problem is still unclear and needs
father investigation.
Thermosensitivity can be combined with pH-sensitiv-
ity via introduction of ionic moieties into the hydrogel
structure [17,18]. Such a combination enhances pur-
poseful controlling of hydrogels properties and thereby
their applications. Incorporation of residues containing
weak acidic groups causes the hydrogel collapse at lower
pH and, vice versa, weak basic groups in the polymer
structure lead to the phase transition at higher pH values.
Temperature and pH influence on physical and chemical
properties of NIPA co-polymers with non-ionogenic
monomers (acrylamide, acrylonitrile, methylacrylate) and
with ionogenic acidic monomer (acrylic acid) were stud-
ied in detail in [19,20].
In this study, hydrogels based on NIPA copolymer
with a basic monomer, allylamine, have been synthesized.
Contrary to vinyl monomers allyl ones are known to po-
lymerize with little yields due to high stability of the
relevant radical and prevailing side reactions. In the pre-
sent work, corresponding copolymer hydrogels were syn-
thesized using allylamine hydrochloride.
Introduction of AlAm component improves the me-
chanical strength of PNIPA hydrogels and adds pH-sen-
sivity to the system. The effect of external stimuli such as
the pH and temperature on the equilibrium swelling ra-
tios of these hydrogels was investigated. Values of aver-
age molecular weight between polymer cross-links, c
M
,
and Flory parameter, χ, were calculated using tempera-
ture dependences of the equilibrium swelling of the hy-
drogel synthesized. Such hydrogels have potential in the
development of smart both pH-and thermo-responsive
membranes, systems for targeted drug delivery, sensors,
etc.
2. Experimental
Sigma-Aldrich reagents were used to synthesize the hy-
drogels. NIPA was recrystallizated from hexane. A cross-
linking agent, N,N-methylenebisacrylamide (MBA), and
components of redox initiating system, ammonium per-
sulphate (APS) and tetramethylethylenediamine (TEMED),
were used without further purification. Allylaminehydro-
cloride (AlAmH) was synthesized via allylamine reaction
with gaseous hydrochloride acid. 30 g of allylimine were
dissolved in 300 cm3 of anhydrous diethyl ether and
cooled to 15˚C. Gaseous HCl was allowed to bubble
through the solution for one hour until white precipitate
of allylamine hydrochloride was formed.
White AlAmH crystals were filtered out and dried un-
der vacuum at room temperature. Molecular structure of
the product obtained was confirmed NMR spectroscopy.
NMR spectra of allylamine hydrochloride are given in
Figure 1 Contrary to unprotonated monomer (see spec-
tral base of given organic compounds SDBS, No. 4258H
SP-00-733)[http://rio db01.ibase.aist.go.jp /sdbs/cgi-bin /di
rect_frame_top.cgi], amino-group protons of which pos-
ses singlet with chemical shift 1.29 m.d., corresponding
allylamine hydrochloride protons are characterized by
singlet at 8.4 m.d.(Figure 1).
2.1. Synthesis of P(NIPA-co-AlAm) Hydrogels
P(NIPA-co-AlAmH) hydrogels were synthesized by free-
radical crosslinking copolymerization of NIPA and AlAmH
in aqueous solutions. APS (0.056 M) and TEMED (0.32
M) were used as the redox initiator system (see Scheme
1).
Cross-linking agent content was 0.125 wt% with re-
spect to monomer amount and a total monomer concen-
tration was 21 wt%. AlAmH concentration in the mixture
with NIPA was varied in the range between 0 to 60 wt%.
Monomer concentrations and the corresponding NIPA/
AlAm mass ratio of synthesized hydrogels are summa-
rized in Table 1. Polymerisation was performed at about
0˚C in argon atmosphere.
2.2. Characterization
NMR
NMR spectra were obtained using the spectrometer
Bruker 400 Avance (1 H 400.08 MHz) with dimethyl-
sulfoxide as a solvent. Impulse duration constituted 12.15
microseconds, the receiver amplification being 39, ac-
cumulation time 2.25 sec., and measurement tempera-
ture 293 K.
Swelling Measurements
For the swelling measurements, the hydrogels were
synthesized, washed in distillated water and then neu-
tralized by 2 M NaOH during 24 h and then washed
again up to pH = 7. Netralized hydrogels were dried and
then immersed in buffer solution at pH at 1.69, 6.86 and
9.18 that corresponding ionic strength 0.05; 0.1; and 0.03
mol/dm3 respectively. Equilibrium swelling ratio was
determined in the temperature range from 5˚C to 53˚C.
The hydrogels were weighed at different times until the
hydrated weight remained constant. Swelling ratio was
Open Access Soft
V. KONOVALOVA ET AL.
Open Access Soft
21
Figure 1. NMR spectra of allylamine hydrochloride.
NH NHO
CH2
O
CH2
+
+
crosslinking
agent
monomers
CH2O
NH
CH3
CH3
CH2
NH2
AA
NH2
O
NH
NH
O
O
NH CH3
CH3
AA
NH2
O
NHCH
3
CH3
Initiator
Scheme 1. P(NIPA-co-AlAm) hydrogels formation.
Table 1. Hydrogel component content (wt%) in feed.
Hydrogel name NIPA AlAmH MBA
P(NiPA) 21.05 - 0.125
P(NIPA-co-AlAm) 9:1 18.95 2.10 0.125
P(NIPA-co-AlAm) 4:2 16.84 4.21 0.125
P(NIPA-co-AlAm) 3:2 12.63 8.42 0.125
P(NIPA-co-AlAm) 2:3 8.42 12.63 0.125
determined as an average value of 10 parallel measure-
ments calculated from the following formula:
s
dd
Qmmm , where ms is the mass of hydrogel
swollen to the equilibrium state and md is mass of the
corresponding dried copolymer.
Thermogravimetric/Differential Thermalanalysis
(TG/DTA)
The dynamic weight loss tests were conducted on
V. KONOVALOVA ET AL.
22
thermogravimetric analyzer (TGA) Derivatograph Q-
1500 D for 50 mg samples with heating rate 10˚C/min in
the 20˚C - 300˚C temperature range. Measurements were
performed under nitrogen atmosphere with the simulta-
neous removal of evolved products. Temperature inter-
vals of weight losses were estimated from the differential
curves taking into account the fact that distinguishing the
stages on DTA curves are more efficient (overrate capa-
bilities) than the integral curves of the weight loss (TG).
Moreover the area under DTA curve is proportional to
the weight loss at a corresponding stage.
3. Results and Discussion
TGA analysis was conducted in the temperature range of
20˚C to 300˚C. As it can be seen from the DTG curves
(Figure 2), the obtained hydrogels have one basic stage
of the weight loss ranged from 50˚C to 110˚C with the
maximum rate of weight loss at approximately 80˚C.
This corresponds to the evaporation of weakly associ-
ated water. The quantity of the weakly associated water
depends on the composition of hydrogels. NIPA hy-
drogel without ionic monomer is characterized by the
weight loss of about 64%. On incorporating 10 wt% and
40 wt% of AlAmH into hydrogel structure (hydrogel
composition P(NIPA-co-AlAm) 9:1 and 2:3, respectively),
quantity of the weakly associated water sharply increases
up to 78% and 85%, respectively due to the incorporation
of highly hydrophilic domains containing amino-groups.
Further increase of AlAmH content up to 60% (hydrogel
P(NIPA-co-AlAm) 2:3) results in decreasing content of
weakly associated water to about 68%. Despite the con-
siderable rise of weakly associated water quantity in the
hydrogels, copolymers containing allylamine are charac-
terized by much better elastic and mechanical properties
as compared to NIPA homopolymers.
Figure 2. DTG curve of various hydrogels: 1—P(NiPA);
2—P(NIPA-co-AlAm) 9:1; 3—P(NIPA-co-AlAm) 3:2, 4—
P(NIPA-co-AlAm) 2:3.
Figure 3 shows that synthesized PNIPA hydrogels re-
veal predicted sharp phase transition in the temperature
range between 25˚C to 32˚C. This coincides with the
literature data [4]. The highest swelling degree reaches
up to 16 g of a liquid per 1 g of the polymer, while at
temperature above Tp swelling ratio drops to as low as 2
g/g. NIPA copolymerization with monomers of various
natures allows purposeful change of their phase transition
temperature and sufficiently widens the area of their ap-
plications. Thus, addition of hydrophilic domains leads to
the increase of Tp, while that of hydrophobic ones causes
the decrease of the relevant parameter [10]. Allylamine is
a monomer containing hydrophilic amino-groups. The
phase transition temperature of the hydrogels was de-
termined as a maximum on differential curves of hy-
drogel swelling ± 1 - 2 grad. Thus, incorporating 10 wt%
of AlAmH into the hydrogel results in shifting the tem-
perature range of the phase transition to 35˚C - 37˚C.
Further increase in AlAmH concentration up to 20% and
40% shifts temperature range of the phase transition to
40˚C - 41˚C and 47˚C - 49˚C, respectively. Swelling ra-
tios of such copolymer hydrogels are also considerably
higher than those of NIPA-based hydrogels. The highest
swelling degree (35 g/g) is reached for the hydrogels with
P(NIPA-co-AlAm)-3:2 composition. Differential curves
of hydrogel swelling (Figure 3(b)) demonstrate enhanced
intensity and discreteness of the phase transition on sub-
stituting NIPA chains by allylamine up to the ratio of 3:2.
The further increase of allylamine (P(NIPA-co-ALAm)
gives continuous phase transition. Thus P(NIPA-co-
AlAm)-3:2 composition shows continuous phase transi-
tion. This fact is explained by the decrease of the hydro-
phobic aggregation of dehydrated NIPA chains at high
temperatures when they appear separated by hydrophilic
allylamine chains. Therefore, any further increase of al-
lylamine is not expedient since it leads to the decrease of
swelling degree and the loss of hydrogel thermo-respon-
sive properties.
Swelling ratio is an important parameter that defines
solvent quantity in the hydrogel in the equilibrium state
and is the function of the polymer network structure,
cross-linking degree, hydrophilysation and functional
groups dissociation degree. The latter is dependent on
temperature, ionic strength and pH of the environment.
Therefore swelling ability of P(NIPA-co-AlAmH) hy-
drogels was studied at pH 1.68; 6.86 and 9.18. Non-io-
nogenic PNIPA hydrogel only slightly varies the swell-
ing degree depending on the media pH. Also low pH-
sensitivity was observed for P(NIPA-co-AlAm) hydrogels
9:1 and 4:2. However, as Figure 4 shows, swelling of the
P(NIPA-co-AlAm) hydrogel 3:2 depends on both tem-
perature and pH of the environment. The highest swell-
ing was achieved in the acidic range of pH. At such con-
ditions, polymer chains are in the most extended confor-
Open Access Soft
V. KONOVALOVA ET AL.
Open Access Soft
23
(a)
(a)
(b)
(b) Figure 4. Integral (a) and differential (b) temperature de-
pendences of P(NIPA-co-AlAm) hydrogel (3:2) swelling ratio
in the temperature range from 5˚C to 55˚C at different рН
values.
Figure 3. Integral (a) and differential (b) temperature de-
pendences of P(NIPA-co-AlAm) hydrogels swelling ratio in
the temperature range from 5˚C to 55˚C at рН 6.86.
mation due to electrostatic repulsion between ionized
3 groups and the best conditions for macromolecu-
lar solvation are achieved.
NH As for P(NIPA-co-AlAm) 2:3 composition, it’s also
pH-responsive in phase transition temperature diapason,
but rate of response on temperature changes is rather
lower especially in acidic environment were the swelling
of hydrogels are the most higher.
Alkaline pH narrows the temperature range of phase
transition and hydrogel swelling is independent of tem-
perature above approximately 43˚C. This can be explained
by the system stabilization due to inter-chain H-bonds
and their transition into more compact conformation. At
low temperatures (up to 25˚C) the copolymer swelling is
slightly dependent on pH. The hydrogels demonstrate
higher pH-sensitivity at temperatures higher than 40˚C.
Moreover, differential curves of the swelling temperature
dependence on pH (Figure 3(b)) show that pH also shifts
the hydrogel phase transition temperature. In the range of
neutral pH, the copolymers are characterized by the big-
gest change of the swelling degree and collapse tem-
perature 47˚C - 49˚C. The shift of pH to both alkaline
and acidic values leads to lowering of the phase transi-
tion temperature (to Tp = 41˚C - 42˚C at pH-1.68 and Tp
= 38˚C - 39˚C at pH = 9.18).
The hydrogel swelling degree was studied at various
temperatures and pH values in order to find such pa-
rameters of macromolecular matrix as the average mo-
lecular weight between crosslinks c
M
, as well as the
parameter of Flory χ that describes polymer-solvent in-
teraction.
The Flory-Rehner [21] models describe the Mc values
for neutral polymer networks, although hydrogels may be
neutral or ionic in nature. Brannon-Peppas and Peppas
derived an equation to describe this ionic contribution
term for both anionic and cationic hydrogels [22].
The next algorithm was used for calculating c
M
and
χ According to Flory-Rener swelling theory the osmotic
pressure of hydrogel is the sum of the pressures attrib-
uted to (1) polymer-solvent mixing (mix), (2) deforma-
V. KONOVALOVA ET AL.
24
tion of network chains to a more elongated state (el), and
(3) the non-uniform distribution of mobile counter ions
between the hydrogel and the external solution (ion) [21]:
.
mix el ion

  (1)
The mixing term is satisfactorily represented by a
Flory-Huggins-type expression of the form

2
22 2
ln 1
mixm mm
RTvv xv

 

(2)
where R is a gas constant, Ttemperature, v2mvolume
polymer fraction in the hydrogel defined as:

1
2
1
1w polymer
m
solvent
qp
Vp





(3)
w
mass ofswollenhydrogel
qmass ofdryhydrogel
and ρ is polymer and solvent density.
Osmotic pressure produced due to the deformation of
network chains to more elongated state:
23 13
122
2
1rm
el
c
Vpv v
RT M






 


(4)
where Φ = 4 - number of cross-linking chains,
V1 is solvent molar volume, constituting 18 cm3/mol
for water,
V2r is polymer volume fraction in the hydrogel upon its
obtaining. The last is found as follows:

1
2
1
1polymer
r
solvent
pf p
Vp





(5)
Osmotic pressure caused by the non-uniform distribu-
tion of mobile counter ions between the hydrogel and the
external solution is calculated as:
22 2
2
12
4
mc
ion
r
iv f
VRT IV


(6)
where I is solution ionic strength, that equals 0.05; 0.1;
and 0.03 mol/dm3 for pH 1.68; 6.86 and 9.18 respec-
tively.
Vr is molecular volume of the polymer was found as
follows:

1
N
iPAcAA c
r
polymer
M
fMf
V
 
where fc is mole fraction of ionic groups in allylamine.
Ionization degree was calculated by 10
a
pH
a
K
i
K
with Kа = 2 × 1010 for allylamine.
When hydrogel reaches swelling state balanced with a
pure solvent, the solvent activity in the hydrogel become
equal to the solvent activity in the pure solvent (i. e.
equals 1). Thus, chemical potential of the solvent in the
hydrogel becomes equal to zero. Using Equations (1)-(6)
the equilibrium equation can be written as:


22
21
1
22
2
23 53
122 2
ln 1
10 4
12
ac
mm
pH
ar
rm
c
KVf
vv
KIV
VVv
M







2
m
v
The above equation can be given as the dependence:
c
B
A
 and by plotting the dependence in A-B
coordinates it is possible to find Flory parameter and the
average molecular mass between the polymer crosslinks
(Figure 5).
The calculations were carry out for hydrogels at tem-
(a)
(b)
Figure 5. Determination of χ and c
M values of hydrogels
from swelling data. () P(NIPA-co-AlAm) (9:1) and ()
P(NIPA-co-AlAm) (3:2) at temperature 23˚C (а) and 38˚C (b).
Open Access Soft
V. KONOVALOVA ET AL. 25
peratures lower and higher of Tp, such as for P(NIPA-co-
AlAm) (9:1) at temperature 23˚C and 38˚C and for hy-
drogel P(NIPA-co-AlAm) (3:2) at temperature 23˚C, 38˚C
and 53˚C, and results were summarized in Table 2.
The data in Table 2 show that for the P(NIPA-co-
AlAm) hydrogel (9:1) composition, the solvent quality
change is observed at elevated temperature. It is known
that the interaction parameter value χ smaller than ~0.5
indicates strong interactions between polymer macro-
molecules and solvent, high solvophility. On the contrary
χ values exceeding 0.5 indicate materials solvophobity.
Thus, in our work, P(NIPA-co-AlAm) hydrogel (9:1) with
90% content of thermo-sensitive monomer changes χ
from 0.5185 to 0.6916 when temperature increases from
23˚C to 38˚C. Moreover, c
M
decreases by 150 times!
And the c
M
value approximates average molecular
mass of copolymer chain at 38˚C. This effect can be ex-
plained by the increase of hydrophobic-hydrophobic in-
teractions between macromolecules and additional cross-
linking of the hydrogel due to H bonds.
With the decrease of content of hydrophobic NIPA
units up to 60%, solvent quality decreases insignificantly
even at temperature of 53˚C. At the same time, there is
20-fold decrease in c
M
values, which manifest as hy-
drogel transition to the collapsed state at the temperatures
higher than Tp.
4. Conclusion
In the present work thermo- and pH-responsive hy-
drogels were synthesized via copolymerization of N-
isopropylacrylamide and allylamine hydrochloride mono-
mers. It was shown that targeted control of the hydrogel
phase transition temperature can be realized by changing
their composition and pH of the environment. Thus rise
of hydrophilic allylamine mass fraction from 10% to
60% causes shift of phase transition temperature from
35˚C to 47˚C. Hydrogels of P(NIPA-co-AlAm) (3:2)
composition are the most pH-responsive. Values of av-
erage molecular mass between c
M
polymer cross-links
and Flory parameter χ were calculated using temperature
dependences of the synthesized hydrogel swelling. It was
found that the increase of hydrophilic component (AlAH)
in the copolymer results in the growth of c
M
, while
temperature rise causes decrease of the above-mentioned
Table 2. χ and c
M values for P(NIPA-co-AlAm) hy-
drogels.
Hydrogel Temperature, ˚C χ c
M
R2
23 0.5185 50 0000.99
P(NIPA-co-AlAm) (9:1) 38 0.6916 323 0.96
23 0.5096 143 0000.97
38 0.5117 100 0000.99P(NIPA-co-AlAm) (3:2)
53 0.5983 5000 0.87
parameter. Flory parameter χ grows both with tempera-
ture and NIPA content rise in copolymer hydrogels show-
ing intensification of hydrophilic interactions. The syn-
thesized “smart” hydrogel systems can be used to de-
velop novel medications, membranes, various sensors
and probes that are capable of radical change of their
working characteristics in response to the slightest envi-
ronmental changes.
REFERENCES
[1] S. Malik, O. Boyko, N. Atkar and W. F. Young, “A
Comparative Study of MR Imaging Profile of Titanium U.
G. Spizzirri, I.Altimari, F. Puoci, O. I. Parisi, F Iemma,
“Innovative Antioxidant Thermo-Responsive Hydrogels
by Radical Grafting of Catechin on Inulin Chain,” Car-
bohydrate Polymers, Vol. 84, No. 11, 2011, pp.517-523.
http://dx.doi.org/10.1016/j.carbpol.2010.12.015
[2] I. Velthoen, J. Beek and P. Dijkstra, “Thermo-Responsive
Hydrogels Based on Highly Branched Poly(Ethylene Gly-
col)-Poly(L-Lactide) Copolymers,” Reactive and Func-
tional Polymers, Vol. 71, No. 3, 2011, pp. 245-253.
http://dx.doi.org/10.1016/j.reactfunctpolym.2010.08.007
[3] T. Caykara, “Effect of Maleic Acid Content on Network
Structure and Swelling Properties of Poly(N-Isopropy-
lacrylamide-Co-Maleic Acid) Polyelectrolyte Hydrogels,”
Journal of Applied Polymer Science, Vol. 92, No. 2, 2004,
pp. 763-769. http://dx.doi.org/10.1002/app.20032
[4] H. Inomata, S. Goto, K. Otake and S. Saito, “Effect of
Additives on Phase Transition of N-Isopropylacrylamide
Gels,” Langmuir, Vol. 8, No. 2, 1992, pp. 687-690.
http://dx.doi.org/10.1021/la00038a064
[5] M. Meewes, J. Ricka, M. Desilva, R. Nyffenegger and T.
Binkert, “Coil-Globule Transition of Poly(N-Isopropy-
lacrylamide): A Study of Surfactant Effects by Light
Scattering,” Macromolecules, Vol. 24, No. 21, 1991, pp.
5811-5816. http://dx.doi.org/10.1021/ma00021a014
[6] B. Tasdelen, N. Kayaman-Apohan and M. Baysal, “Pre-
paration, Characterization, and Drug-Release Properties
of Poly(N-Isopropylacrylamide) Microspheres Having
Poly(Itaconic Acid) Graft Chains,” Journal of Applied
Polymer Science, Vol. 97, No. 3, 2005, pp. 1115-1124.
http://dx.doi.org/10.1002/app.21777
[7] G. Chen and A. S. Hoffman, “Graft Copolymers That
Exhibit Temperature-Induced Phase Transition over a Wide
Range of pH,” Nature, Vol. 373, No. 6509, 1995, pp. 49-
52. http://dx.doi.org/10.1038/373049a0
[8] I. Hiroshi, G. Shuichi and S. Shozaburo, “Phase Transi-
tion of N-Substituted Acrylamide Gels,” Macromolecules,
Vol. 23, No. 22, 1990, pp. 4887-4888.
http://dx.doi.org/10.1021/ma00224a023
[9] W. Chi and Z. Shuiqin, “Volume Phase Transition of
Swollen Gels: Discontinuous or Continuous?” Macro-
molecules, Vol. 30, 3, 1997, pp. 574-576.
http://dx.doi.org/10.1021/ma960499a
[10] S. Sasaki and S. J. Okabe, “Effects of Ions on the Solubil-
ity Transition and the Phase-Separation of N-Isopropy-
lacrylamide in Water,” Physical Chemistry B, Vol. 115,
Open Access Soft
V. KONOVALOVA ET AL.
Open Access Soft
26
No. 44, 2011, pp. 12905-12910.
http://dx.doi.org/10.1021/jp207576z
[11] Y. Hirose, Y. Hirokawa and T. Tanaka, “Phase Transition
of Submicron Gel Beads,” Macromolecules, Vol. 20, No.
6, 1987, pp. 1342-1344.
http://dx.doi.org/10.1021/ma00172a029
[12] H. Yu and D. W. Grainger, “Amphiphilic Thermosensi-
tive N-Isopropylacrylamide Terpolymer Hydrogels Pre-
pared by Micellar Polymerization in Aqueous Media,”
Macromolecules, Vol. 27, No. 16, 1994, pp. 4554-4560.
http://dx.doi.org/10.1021/ma00094a019
[13] Y. Li and T. J. Tanaka, “Study of the Universality Class
of the Gel Network System,” Journal of Chemical Phys-
ics, Vol. 90, No. 9, 1989, pp. 5161-5166.
http://dx.doi.org/10.1063/1.456559
[14] S. J. Hirotsu, “Phase Transition of a Polymer Gel in Pure
and Mixed Solvent Media,” Journal of the Physical So-
ciety of Japan, Vol. 56, No. 1, 1987, pp. 233-242.
http://dx.doi.org/10.1143/JPSJ.56.233
[15] C. Tuncer, K. Simin and D. Goеkhan, “Thermosensitive
Poly(N-Isopropylacrylamide-Co-Acrylamide) Hydrogels:
Synthesis, Swelling and Interaction with Ionic Surfac-
tants,” European Polymer Journal, Vol. 42, No. 2, 2006,
pp. 348-355.
http://dx.doi.org/10.1016/j.eurpolymj.2005.07.006
[16] L. Hua, W. Xiaogui, Li Hua, X. Wang, Z. Wang and K.Y.
Lam, “Multiphysics Modeling of Volume Phase Transi-
tion of Ionic Hydrogels Responsive to Thermal Stimu-
lus,” Macromolecular Bioscience, Vol. 5, No. 9, 2005, pp.
904-914. http://dx.doi.org/10.1002/mabi.200500082
[17] H. Bishta, L. Wanb, G. Maob and D. Oupicky, “pH-
Controlled Association of PEG-Containing Terpolymers
of N-Isopropylacrylamide and 1-Vinylimidazole,” Poly-
mer, Vol. 46, No. 19, 2005, pp. 7945-7952.
http://dx.doi.org/10.1016/j.polymer.2005.06.080
[18] H. Dautzenberg, Y. B. Gao and M. Hahn, “Formation,
Structure, and Temperature Behavior of Polyelectrolyte
Complexes between Ionically Modified Thermosensitive
Polymers,” Langmuir, Vol. 16, No. 23, 2000, pp. 9070-
9081. http://dx.doi.org/10.1021/la000368+
[19] V. V. Konovalova, Yu. M. Samchenko, G. A. Pobigay, A.
F. Burban and Z. R. Ulberg, “Hydrogel Membranes with
pH- and Thermo-Responsive Parameters,” Proceedings of
the 15th International SymposiumArs Separatoria 2010’,
Torun, 4-7 July 2010, pp. 222-225.
[20] Yu. Samchenko, V. Konovalova, G. Pobigay, A. Burban
and Z. Ulberg, “Thermo Sensitive Copolymers Hydrogels
with Controlled Phase Transition Temperature,” Reports
of Ukrainian NAS, Vol. 8, 2011, pp. 123-129.
[21] P. J. Flory, “Principles of Polymer Chemistry,” Cornell
University Press, Ithaca, 1953.
[22] L. Brannon-Peppas and N. A. Peppas, “Equilibrium Swel-
ling Behavior of pH-Sensitive Hydrogels,” Chemical En-
gineering Science, Vol. 46, No. 3, 1991, pp. 715-722.
http://dx.doi.org/10.1016/0009-2509(91)80177-Z