You are on page 1of 12

polymers

Article
Synthesis of Poly(2-Acrylamido-2-Methylpropane
Sulfonic Acid) and its Block Copolymers with Methyl
Methacrylate and 2-Hydroxyethyl Methacrylate by
Quasiliving Radical Polymerization Catalyzed by a
Cyclometalated Ruthenium(II) Complex
Vanessa Martínez-Cornejo 1 , Joaquin Velázquez-Roblero 1,2 , Veronica Rosiles-González 1 ,
Monica Correa-Duran 1 , Alejandro Avila-Ortega 2 , Emanuel Hernández-Núñez 3 ,
Ronan Le Lagadec 4, * and Maria Ortencia González-Díaz 5, *
1 Unidad de Materiales, Centro de Investigación Científica de Yucatán, A.C., Calle 43 No. 130,
Chuburná de Hidalgo, C.P. 97205 Mérida, Yucatán, Mexico; vmartinezcornejo10@gmail.com (V.M.-C.);
joaquinwr96@gmail.com (J.V.-R.); veronica.rosiles@estudiantes.cicy.mx (V.R.-G.);
monica.correa@estudiantes.cicy.mx (M.C.-D.)
2 Facultad de Ingeniería Química, Universidad Autónoma de Yucatán, Periférico Norte Km. 33.5,
Chuburná de Hidalgo Inn, C.P. 97203 Mérida, Yucatán, Mexico; alejandro.avila@correo.uady.mx
3 CONACYT, Departamento de Recursos del Mar, Centro de Investigación y de Estudios Avanzados del IPN,
97310 Unidad Mérida, Yucatán, Mexico; emanuel.hernandez@cinvestav.mx
4 Instituto de Química, Universidad Nacional Autónoma de México, Circuito Exterior s/n,
Ciudad Universitaria, 04510 Ciudad de México, Mexico
5 CONACYT–Centro de Investigación Científica de Yucatán, A.C., Calle 43 No. 130, Chuburná de Hidalgo,
97205 Mérida, Yucatán, Mexico
* Correspondence: ronan@unam.mx (R.L.L.); maria.gonzalez@cicy.mx (M.O.G.-D.)

Received: 24 June 2020; Accepted: 20 July 2020; Published: 27 July 2020 

Abstract: The first example of quasiliving radical polymerization and copolymerization of


2-acrylamido-2-methylpropane sulfonic acid (AMPS) without previous protection of its strong acid
groups catalyzed by [Ru(o-C6 H4 -2-py)(phen)(MeCN)2 ]PF6 complex is reported. Nuclear magnetic
resonance (RMN) and gel permeation chromatography (GPC) confirmed the diblock structure of
the sulfonated copolymers. The poly(2-acryloamido-2-methylpropanesulfonic acid)-b-poly(methyl
methacrylate) (PAMPS-b-PMMA) and poly(2-acryloamido-2-methylpropanesulfonic acid)-b-poly
(2-hydroxyethylmethacrylate) (PAMPS-b-PHEMA) copolymers obtained are highly soluble in organic
solvents and present good film-forming ability. The ion exchange capacity (IEC) of the copolymer
membranes is reported. PAMPS-b-PHEMA presents the highest IEC value (3.35 mmol H+ /g),
but previous crosslinking of the membrane was necessary to prevent it from dissolving in aqueous
solution. PAMPS-b-PMMA exhibited IEC values in the range of 0.58–1.21 mmol H+ /g and it
was soluble in methanol and dichloromethane and insoluble in water. These results are well
correlated with both the increase in molar composition of PAMPS and the second block included
in the copolymer. Thus, the proper combination of PAMPS block copolymer with hydrophilic or
hydrophobic monomers will allow fine-tuning of the physical properties of the materials and may lead
to many potential applications, such as polyelectrolyte membrane fuel cells or catalytic membranes
for biodiesel production.

Keywords: poly(2-acryloamido-2-methylpropane sulfonic acid); cyclometalated ruthenium(II)


complex; radical polymerization; block copolymer; membranes

Polymers 2020, 12, 1663; doi:10.3390/polym12081663 www.mdpi.com/journal/polymers


Polymers 2020, 12, 1663 2 of 12

1. Introduction
Poly(2-acrylamido-2-methyl-1-propanesulfonic acid) (PAMPS) displays interesting properties
that may lead to many potential applications. Such properties arise from the strongly ionizable
sulfonate groups in its chemical structure, its pH-responsiveness and its swelling behavior [1–3]
AMPS-containing polymers have been successfully applied in polyelectrolyte membrane fuel
cells [1], as catalytic membranes for biodiesel production [4] and in medical applications due
to their low toxicity, hydrolytic stability and antimicrobial activity against microorganisms [5,6].
Moreover, they are used in a wide range of industrial products such as cosmetics, coatings and
adhesives, amongst others [7]. Homo- and copolymers of this monomer have been synthesized
straightforwardly by free radical polymerization [8–10]. However, it is well known that it is
nearly impossible to obtain predetermined molecular weights through the conventional radical
mechanism due to chain termination reactions [11]. On the other hand, atom transfer radical
polymerization (ATRP) has emerged as a robust technique for synthesizing well-defined polymers
with predetermined molecular weights, low dispersity and with different compositions, topologies
and functionalities [12,13]. A literature review indicates that in the polymerization of PAMPS by
the ATRP technique with copper catalysts, neutralization of the acid groups is required to prevent
protonation of the amino-based ligands, which can lead to the decomposition of the copper/ligand
catalytic system and to avoid subsequent side reactions [14–19]. In many cases, the characteristics of
a quasiliving ATRP have not been observed, despite the use of AMPS in its salt form. For example,
the synthesis of PAMPS by ATRP in DMF:water solvent (50:50 v/v) at 20 ◦ C with CuCl and different
ligands, such as 2,20 -bipyridine (bpy), N,N,N0 ,N00 ,N00 -pentamethyldiethylenetriamine (PMDETA) and
1,1,4,7,10,10-hexamethyltriethylenetetramine (HMTETA), presents a nonlinear first-order kinetic plot
and experimental molecular weight (Mn,GPC ) higher than its theoretical values, which are indicative
of a high concentration of radicals and extensive termination in the initial stage [14]. In that work,
for the ATRP synthesis of PAMPS, high catalyst loadings were necessary to reduce its deactivation
(CuCl/tris[2-(dimethylamino)ethyl]amine ligand with the addition of equimolar amounts of CuCl2
with respect to CuCl) [14,15]. A. Tolstov et al. reported the ATRP of AMPS in aqueous conditions
with activators generated by electron transfer (AGET) of sodium 2-acryloamido-2-methyl-N-propane
sulfonate and its copolymerization with 2-hydroxyethyl acrylate (HEA), using CuBr2 /HMTETA and
L-ascorbic acid as a reducing agent. However, the resulting copolymers presented higher Mn,GPC and
dispersity (Ð = Mw /Mn ) than the expected values [16]. On the other hand, the ATRP synthesis of
PAMPS was reported with in situ neutralization of AMPS with tri(n-butyl)amine (TBA) at 60 ◦ C in
dimethylformamide (DMF) with CuCl/bpy catalyst [20]. The authors reported very low conversions
(less than 15%), albeit the use of ascorbic acid as a reducing agent improved the reaction rate.
Ruthenium catalysts in ATRP have been described as more tolerant to several functional
groups with good solubility in protic and aprotic solvents, which permit quasiliving polymerization
in polar media [21–23]. In particular, series of cationic 18-electron coordinatively saturated
cyclometalated ruthenium(II) complexes have been successfully applied in homopolymerizations and
copolymerizations of hydrophobic and hydrophilic monomers [21,24–26]. The use of cyclometalated
ruthenium(II) complexes can eliminate the undesired copper–promoted side reactions. However,
the radical polymerization of AMPS using ruthenium catalyst has not been evaluated yet. Therefore,
in this study we present the ATRP synthesis of AMPS catalyzed by [Ru(o-C6 H4 -2-py)(phen)
(MeCN)2 ]PF6 complex (phen: 1,10-phenanthroline), without previous protection of the acid groups.
The copolymerization approach was carried out by sequential polymerization of a second monomer,
methyl methacrylate (MMA) and 2-hydroxyethyl methacrylate (HEMA), respectively, via direct fresh
feeding into the PAMPS prepolymer solution.
Polymers 2020, 12, 1663 3 of 12
Polymers 2020, 12, x 3 of 12

2. Materials
2. Materials and
and Methods
Methods

2.1. Materials and Reagents


2.1.
All reactions
All reactions were
were carried
carried out
out under
under inert
inert atmosphere
atmosphere using
using conventional
conventional Schlenk techniques.
2-Acrylamido-2-methylpropane sulfonic
2-Acrylamido-2-methylpropane sulfonicacid
acid(AMPS,
(AMPS, 99%) andand
99%) all reagents were supplied
all reagents by Sigma-
were supplied by
Aldrich, México.México.
Sigma-Aldrich, The 2-hydroxyethyl methacrylate
The 2-hydroxyethyl (HEMA,(HEMA,
methacrylate 99%) and methyl
99%) methacrylate
and methyl (MMA,
methacrylate
99%) monomers
(MMA, were purified
99%) monomers by passing
were purified throughthrough
by passing an inhibitor removerremover
an inhibitor column.column.
Methanol (MeOH,
Methanol
99.9%) was
(MeOH, distilled
99.9%) prior toprior
was distilled use. N,N-dimethylformamide (DMF, (DMF,
to use. N,N-dimethylformamide 99.8%),99.8%),
methanol and water
methanol were
and water
previously deoxygenated with nitrogen bubbling for 20 min before use. Ethyl 2–bromoisobutyrate
were previously deoxygenated with nitrogen bubbling for 20 min before use. Ethyl 2–bromoisobutyrate
(EBiB) and
(EBiB) and deuterated
deuterated solvents
solvents were
wereused
usedas
asreceived.
received.

2.2.
2.2. Synthesis of the Ruthenium(II) Catalyst
Catalyst
The
The[Ru(o-C
[Ru(o-C6H
6H4 -2-py)(phen)(MeCN)
4-2-py)(phen)(MeCN) 2 ]PF6 catalyst
2]PF waswas
6 catalyst synthesized according
synthesized to the to
according literature [27,28].
the literature
The chemical
[27,28]. structurestructure
The chemical of the ruthenium catalyst is
of the ruthenium presented
catalyst in Figurein1.Figure 1.
is presented

Figure1.1.Chemical
Figure Chemicalstructure
structureof
ofthe
the[Ru(o-C
[Ru(o-C6 6H
H44-2-py)(phen)(MeCN)
-2-py)(phen)(MeCN)22]PF
]PF66 catalyst.
catalyst.

2.3.
2.3. Synthesis
Synthesis of
of the
the Homopolymers
Homopolymers and
and Copolymers
Copolymers
Radical
Radicalhomopolymerization
homopolymerizationofofAMPS. AMPS. The
Thereaction
reaction was wascarried outout
carried in solution withwith
in solution a 200:1:1 mol
a 200:1:1
ratio between the monomer, initiator and the catalyst, respectively. In a typical
mol ratio between the monomer, initiator and the catalyst, respectively. In a typical PAMPS synthesis, PAMPS synthesis, a
25
a 25mLmL Schlenk
Schlenk flask was
flask charged
was charged with
with[Ru(o-C
[Ru(o-C 6H 4 -2-py)(phen)(MeCN)
6H 4-2-py)(phen)(MeCN) 2 2]PF
]PF66 (25
(25 mg,
mg, 0.037
0.037 mmol),
mmol),
AMPS
AMPS (1.56 g, 7.52 mmol) and solvent (2 mL DMF). The homogeneous solution was degassed by
(1.56 g, 7.52 mmol) and solvent (2 mL DMF). The homogeneous solution was degassed by
three
three freeze–pump–thaw
freeze–pump–thaw cyclescycles and
and EBiB
EBiB (5.5 μL, 0.037
(5.5 µL, 0.037 mmol)
mmol) was was added.
added. The The reaction
reaction mixture
mixture was
was
immersed ◦ C for 10 h. Samples were taken after certain time intervals
immersed in in an
an oil
oil bath
bath with
with stirring
stirring at
at 80
80 °C for 10 h. Samples were taken after certain time intervals
using a N -purged syringe. The polymer
using a N22-purged syringe. The polymer was poured was poured intointo ethyl
ethyl acetate,
acetate, filtered
filtered and and finally
finally purified
purified
through
through aa Florisil
Florisil column
column with
with methanol.
methanol.
Synthesis
Synthesis ofofPAMPS-b-PHEMA
PAMPS-b-PHEMA copolymer.
copolymer. TheThe
PAMPSPAMPS macroinitiator was first
macroinitiator waspolymerized under
first polymerized
the conditions described above using DMF as solvent at 80 ◦ C. The reaction was stopped after 16 h and
under the conditions described above using DMF as solvent at 80 °C. The reaction was stopped after
0.6
16 hmLandof 0.6
DMF mLwas addedwas
of DMF to decrease
added tothe high viscosity
decrease the highofviscosity
the reaction mixture.
of the reaction Then, 0.9 mL
mixture. HEMA
Then, 0.9
was added by syringe under
mL HEMA was added by syringe N 2 purge. After complete homogenization, the reaction
under N2 purge. After complete homogenization, the reaction was was carried out
at 70 ◦ C for
carried out3 h.
at The
70 purification
°C for 3 h. procedure was similarprocedure
The purification to that of PAMPS homopolymerization.
was similar to that of PAMPS Molar
composition for the block
homopolymerization. copolymerizations
Molar composition forwas controlled
the block by varying thewas
copolymerizations HEMA ratio. by varying
controlled
the HEMA ratio.
Synthesis of PAMPS-b-PMMA Copolymer
Synthesis of PAMPS-b-PMMA
A solution of AMPS (0.39 copolymer.
g, 1.88 mmol), [Ru(o-C6 H4 -2-py)(phen)(MeCN)2 ]PF6 (12.5 mg,
0.0185
A mmol)
solutionand EBiB (2.75
of AMPS (0.39 µL, 0.0185
g, 1.88 mmol)
mmol), in 0.5
[Ru(o-C mL DMF was heated at 80 ◦ C for 16 h.
6H4-2-py)(phen)(MeCN)2]PF6 (12.5 mg, 0.0185
Then
mmol) 0.2and
mLEBiB
or 0.5(2.75
mL ofμL,degassed MMAin
0.0185 mmol) was
0.5added
mL DMFby syringe underatN80
was heated 2 purge
°C for for a copolymer
16 h. Then 0.2 mLwith
or
75 and 88% mol of PMMA block, respectively. After complete homogenization,
0.5 mL of degassed MMA was added by syringe under N2 purge for a copolymer with 75 and 88% the flask was placed
in
molanofoil
PMMA 80 ◦ Crespectively.
bath atblock, for 20 h. TheAfter
reaction mixture
complete was cooled and
homogenization, thepoured into
flask was diethyl
placed ether
in an and
oil bath
the precipitate was filtered. Finally, the polymer was dissolved in dichloromethane and
at 80 °C for 20 h. The reaction mixture was cooled and poured into diethyl ether and the precipitate purified by
passing through
was filtered. a silica
Finally, thegel column
polymer was to dissolved
remove theinresidual catalyst. and purified by passing through
dichloromethane
a silica gel column to remove the residual catalyst.
Polymers 2020, 12, 1663 4 of 12

2.4. Characterization and Membrane Preparation


The conversions were determined by 1 H-NMR on a Varian NMR spectrometer (Agilent
technologies, Santa Clara, CA, USA) operating at 600 MHz using deuterium oxide (D2 O) as the
deuterated solvent. The monomer conversion (C) was calculated as reported by [29].

I
 
C = 100 × 1 − A or C = 100 × (1 − IA ) (1)
IB

where C is the monomer conversion, IA is the integrated signal between 6.13 to 6.28 ppm corresponding
to CH2 protons adjacent to the double bond in the monomer, and IB is the integrated signal at 3.44 ppm
due to CH2 protons adjacent to the –SO3 H group in the monomer and polymer (this peak was selected
as a reference IB = 1).
Molecular weight (Mn ) and dispersity (Ð) were determined by gel permeation chromatography
(GPC) on an Agilent 1100 HPLC equipped with a refractive index detector (IR) and two columns:
Zorbax PSM 60-S and PSM 1000-S (Agilent technologies, CA, USA). DMF containing LiBr (0.05%) was
used as the eluent (flow rate 0.70 mL min−1 ) at 30 ◦ C and PMMA was used as the calibration standard.

2.5. Membrane Preparation


Membranes were prepared by casting a 6% solution of PAMPS-b-PHEMA and PAMPS-b-PMMA
copolymers in deionized water and methanol, respectively.
PAMPS-b-PMMA membrane. A total of 410 mg of PAMPS-b-PMMA was dissolved in 6.8 mL of
methanol at room temperature for 5 h under continuous stirring until a homogeneous and highly
viscous solution was obtained. The solution was poured into an aluminum ring and the solvent was
evaporated slowly at 30 ◦ C for 24 h. Finally, the obtained membranes were dried in a vacuum oven at
80 ◦ C for 24 h to eliminate the residual solvent.
PAMPS-b-PHEMA membrane. A total of 410 mg of PAMPS-b-PMMA was dissolved in 6.8 mL of
deionized water at 60 ◦ C for 24 h under continuous stirring. The polymer solution was then cooled and
sulfosuccinic acid (15 wt% SSA/weight of PHEMA block) was added. In the next step, the solution with
SSA was stirred for 12 h and an additional 2 h in an ultrasonic bath. The PAMPS-b-PMMA solution
was poured onto a Teflon plate and the solvent was evaporated slowly at 50 ◦ C for 24 h. The obtained
membranes were dried in a vacuum oven at 60 ◦ C for 24 h. Finally, the membranes were crosslinked at
120 ◦ C for 1 h.

2.6. Determination of Ion Exchange Capacity (IEC)


The ion exchange capacity (IEC, mmol/g) value of the membranes was calculated using an
acid–base titration method. Membrane samples of 0.1 g were immersed in 5 mL of NaOH solution
(0.1 M) at room temperature for 24 h to be converted into their sodium salt form. Subsequently,
the remaining solution was titrated with 0.02 M HCl. The IEC value was calculated according to
Equation (2):
  (VNaOH )(CNaOH ) − (VHCl )(CHCl )
IEC mmol g−2 = (2)
Ws
where VNaOH and CNaOH are the volume and molarity of NaOH, respectively; VHCl and CHCl are the
volume and molarity of HCl consumed in the titration, respectively; and Ws is the dry weight of the
samples. The IEC was determined in duplicate.

3. Results and Discussion

3.1. Synthesis and Characterization of AMPS Homopolymer


Initially, the polymerization of AMPS was carried out using a mixture of water:DMF or
methanol:DMF (50:50, v/v), as reported in the literature (See Table 1) [14,15,30]. As shown in Figure 2,
With regard to the spontaneous polymerization of AMPS in the presence of water, the dependence
of the initial polymerization rate on the neutralization degree was reported. In absence of a
neutralizing agent (NaOH), the AMPS polymerization proceeds slowly, followed by a considerable
increase in the reaction viscosity and reaction rate [31]. Another possible explanation for the
induction period
Polymers 2020, 12, 1663could be the slow generation of the actual catalytically active system from the [Ru(o- 5 of 12
C6H4-2-py)(phen)(MeCN)2]PF6 precatalyst, possibly involving the oxidation of Ru(II) to Ru(III), as in
the generally accepted mechanism for ATRP [24]. Furthermore, polymers were obtained with high
the study of
molecular the reaction
weights kinetics
(Mn) and in water:DMF
an increase of the Ðshowed a slow
values was initiation,
observed with
(> 1.7) an induction
(Table 1), thoughperiod
at lowof
around 2 h and a considerable increase in viscosity from the fourth hour.
conversions. The polymerization of AMPS in methanol:DMF proceeded with a similar behavior to This can be attributed to
lowreaction
the efficiency in of the initiator caused by the decrease of the solubility of EBiB in aqueous media or the
water:DMF.
spontaneous polymerization capacity of this type of acid monomer in aqueous solutions [21,31,32].
WithTable
regard1.toPolymerization
the spontaneous of polymerization of AMPS in the presence of
2-acrylamido-2-methyl-1-propanesulfonic acidwater, the dependence
(AMPS) mediated byof the
initial[Ru(o-C
polymerization rate on the neutralization
6H4-2-py)(phen)(MeCN) degree
2]PF6 in different was[AMPS]
solvents. reported. In absence
o:[EBiB] of oa=neutralizing
o:[catalyst] 200:1:1. agent
(NaOH), the AMPS polymerization proceeds slowly, followed by a considerable increase in the reaction
Time Mn × 103
viscosity and reaction rate Solvent
[31]. Another possible Conversion
explanation(%) for the induction Ð period could be the
(h) g/mol
slow generation of the actual catalytically active system from the [Ru(o-C6 H4 -2-py)(phen)(MeCN)2 ]PF6
Water:DMF 6 68 88.3 1.83
precatalyst, possibly involving the oxidation of Ru(II) to Ru(III), as in the generally accepted mechanism
Methanol:DMF 8 73 75.2 1.75
for ATRP [24]. Furthermore, polymers were obtained with high molecular weights (Mn ) and an
DMF 8 61 39.3 1.51
increase of the Ð values was observed (> 1.7) (Table 1), though at low conversions. The polymerization
of AMPS in methanol:DMF proceeded with a similar behavior to the reaction in water:DMF.
On the other hand, the polymerization of AMPS in DMF catalyzed by [Ru(o-C6H4-2-
py)(phen)(MeCN) 2]PF6 was slightly slower—reaching a conversion of 61% in 8 h and 74% in 10 h—
Table 1. Polymerization of 2-acrylamido-2-methyl-1-propanesulfonic acid (AMPS) mediated by
than [Ru(o-C
the aforementioned systems. Nevertheless, it presented a pseudo-first-order kinetic (Figure 1),
6 H4 -2-py)(phen)(MeCN)2 ]PF6 in different solvents. [AMPS]o :[EBiB]o :[catalyst]o = 200:1:1.
which indicate that the concentration of active species in propagation remains constant during the
Solventand the induction
Time (h) period Conversion (%) as shown
Mn × 10in 3 g/mol Ð behavior is
reaction [14,33,34] disappeared, Figure 2. This
characteristic of a quasiliving polymerization
Water:DMF 6 [33,34],
68 which can 88.3 be attributed to1.83 the dramatic
increases Methanol:DMF
of the solubility of both 8 the catalyst and 73 EBiB initiator in 75.2DMF that leads 1.75to a higher
efficiency of the DMF
initiator and fast8generation of the 61catalytically active 39.3
species. 1.51

1.4

1.2
Water:DMF
DMF
Methanol:DMF
1.0
Ln [Mo]/[Mt]

0.8

0.6

0.4

0.2

0.0

0 1 2 3 4 5 6 7 8 9 10 11
Time (h)

Figure 2. Kinetic plot of poly(2-acrylamido-2-methyl-1-propanesulfonic acid) (PAMPS) synthesized by


Figure 2. Kinetic plot of poly(2-acrylamido-2-methyl-1-propanesulfonic acid) (PAMPS) synthesized
[Ru(o-C6 H4 -2-py)(phen)(MeCN)2 ]PF6 in different solvents.
by [Ru(o-C6H4-2-py)(phen)(MeCN)2]PF6 in different solvents.
On the other hand, the polymerization of AMPS in DMF catalyzed by [Ru(o-C6 H4 -2-py)
As can be seen in Figure 3, the Mn increased with conversion and a narrow molecular weight
(phen)(MeCN)2 ]PF6 was slightly slower—reaching a conversion of 61% in 8 h and 74% in 10 h—than
distribution is obtained (in the range of 1.22–1.55). A discrepancy was observed between the
the aforementioned systems. Nevertheless, it presented a pseudo-first-order kinetic (Figure 1),
theoretical (Mn-t) and experimental Mn values due to the difference in the hydrodynamic volume
which indicate that the concentration of active species in propagation remains constant during the
between PAMPS, which is hydrophilic, and the hydrophobic calibration standard (PMMA) [21,33,35].
reaction [14,33,34] and the induction period disappeared, as shown in Figure 2. This behavior is
characteristic of a quasiliving polymerization [33,34], which can be attributed to the dramatic increases
of the solubility of both the catalyst and EBiB initiator in DMF that leads to a higher efficiency of the
initiator and fast generation of the catalytically active species.
As can be seen in Figure 3, the Mn increased with conversion and a narrow molecular weight
distribution is obtained (in the range of 1.22–1.55). A discrepancy was observed between the theoretical
Polymers 2020, 12, 1663 6 of 12

(Mn-t ) and experimental Mn values due to the difference in the hydrodynamic volume between PAMPS,
which is
Polymers hydrophilic,
2020, 12, x and the hydrophobic calibration standard (PMMA) [21,33,35]. 6 of 12

Evolution
Figure 3.3.Evolution of molecular
of the the molecular weight
weight (Mndispersity
(Mn) and ) and dispersity
(Ð) with (Ð) with the conversion
the conversion for AMPS
for AMPS polymerization ◦ C.
polymerization in DMF in catalyzed
DMF catalyzed by [Ru(o-C
by [Ru(o-C 6 H4 -2-py)(phen)(MeCN)
6H4-2-py)(phen)(MeCN) 2]PF62 ]PF
at6 at8080 °C.
[AMPS]oo:[EBiB] :[catalyst]oo==200:1:1.
:[EBiB]oo:[catalyst] 200:1:1.

3.2. Synthesis of Block Copolymers


3.2. Synthesis of Block Copolymers
PAMPS is a water-swollen homopolymer to the point of being soluble. Thus, to be used
PAMPS is a water-swollen homopolymer to the point of being soluble. Thus, to be used for
for potential applications, such as polyelectrolyte membrane fuel cells, a catalytic membrane for
potential applications, such as polyelectrolyte membrane fuel cells, a catalytic membrane for
biodiesel production or for medical applications, it can only be used copolymerized, crosslinked,
biodiesel production or for medical applications, it can only be used copolymerized, crosslinked,
blended or in gel forms to control its swelling in aqueous solution [4,36]. In this work,
blended or in gel forms to control its swelling in aqueous solution [4,36]. In this work, PAMPS-b-
PAMPS-b-PHEMA and PAMPS-b-PMMA copolymers were synthesized in order to prove the
PHEMA and PAMPS-b-PMMA copolymers were synthesized in order to prove the effectiveness of
effectiveness of [Ru(o-C6 H4 -2-py)(phen)(MeCN)2 ]PF6 catalyst in a living polymerization using PAMPS
[Ru(o-C6H4-2-py)(phen)(MeCN) 2]PF6 catalyst in a living polymerization using PAMPS as
as macroinitiator.
macroinitiator.
3.2.1. Synthesis of PAMPS-b-PMMA Block Copolymer
3.2.1. Synthesis of PAMPS-b-PMMA Block Copolymer
Amphiphilic block copolymers are very interesting due to their applications, properties and
Amphiphilic block copolymers are very interesting due to their applications, properties and
morphological self-assembly capacity. However, the synthesis of amphiphilic block copolymers is
morphological self-assembly capacity. However, the synthesis of amphiphilic block copolymers is
not an easy task, first due to the incompatibility of each segment of different nature (hydrophilic
not an easy task, first due to the incompatibility of each segment of different nature (hydrophilic and
and hydrophobic) and second due to the difficulty of finding a common solvent for both blocks.
hydrophobic) and second due to the difficulty of finding a common solvent for both blocks. The
The synthesis is even more complicated when the hydrophilic block possesses acidic functional
synthesis is even more complicated when the hydrophilic block possesses acidic functional groups
groups [37]. Considering that the most active monomers (methacrylates) are polymerized before
[37]. Considering that the most active monomers (methacrylates) are polymerized before the least
the least active ones (acrylamides) in the copolymers synthesis [38], copolymerization was carried
active ones (acrylamides) in the copolymers synthesis [38], copolymerization was carried out using
out using previously isolated PMMA (synthesized under similar conditions described for the AMPS
previously isolated PMMA (synthesized under similar conditions described for the AMPS
homopolymerization in THF at 80 ◦ C for 4 h) as macroinitiator in DMF at 80 ◦ C. Nevertheless, the
homopolymerization in THF at 80 °C for 4 h) as macroinitiator in DMF at 80 °C. Nevertheless, the
copolymerization did not proceed under these conditions. However, the PAMPS-b-PMMA copolymer
copolymerization did not proceed under these conditions. However, the PAMPS-b-PMMA
was successfully obtained with the one-pot sequential method with the addition of MMA monomer
copolymer was successfully obtained with the one-pot sequential method with the addition of MMA
to the PAMPS macroinitiator. This can be attributed to the presence of the polar groups (–SO3 H) of
monomer to the PAMPS macroinitiator. This can be attributed to the presence of the polar groups (–
AMPS which may be acting as accelerators in the system and therefore increasing the reactivity of the
SO3H) of AMPS which may be acting as accelerators in the system and therefore increasing the
AMPS monomer compared to MMA [39–41]. The effect of acceleration on the ATRP by polar groups
reactivity of the AMPS monomer compared to MMA [39–41]. The effect of acceleration on the ATRP
(either from polar additives, solvents or monomers containing polar substituents) when using Cu
by polar groups (either from polar additives, solvents or monomers containing polar substituents)
catalyst has been attributed to the increase in polarity in the reaction medium and the generation of
when using Cu catalyst has been attributed to the increase in polarity in the reaction medium and the
more active catalytic species through the interaction of polar molecules with the metallic center [39–41].
generation of more active catalytic species through the interaction of polar molecules with the
However, so far, we do not have evidence of some kind of AMPS interaction with the ruthenium
metallic center [39–41]. However, so far, we do not have evidence of some kind of AMPS interaction
catalyst. We will be carrying out in-depth future research on this specific topic. On the other hand,
with the ruthenium catalyst. We will be carrying out in-depth future research on this specific topic.
On the other hand, Oikonomou et al. reported the synthesis of a series of amphiphilic copolymers of
poly(sodium styrene sulfonate)-b-poly(methyl methacrylate) (PSSNa-b-PMMA) by ATRP using
PSSNa macroinitiator [37], in a similar way to the synthesis of copolymers reported in this work.
The PAMPS macroinitiator was synthesized previously for 16 h to ensure high monomer
consumption (conversion of 82%). It is worth noting that a 100:1:1 molar ratio was used for the
Polymers 2020, 12, 1663 7 of 12

Oikonomou et al. reported the synthesis of a series of amphiphilic copolymers of poly(sodium styrene
sulfonate)-b-poly(methyl methacrylate) (PSSNa-b-PMMA) by ATRP using PSSNa macroinitiator [37],
in a similar way to the synthesis of copolymers reported in this work.
The PAMPS macroinitiator was synthesized previously for 16 h to ensure high monomer
consumption (conversion of 82%). It is worth noting that a 100:1:1 molar ratio was used for the
monomer/catalyst/initiator to obtain a low molecular weight macroinitiator that would facilitate
growth of the second polymer block. With this method, PAMPS-b-PMMA block copolymers with 10
and 25% molPolymers
PAMPS were
2020, 12, x synthesized. In contrast to the AMPS homopolymer, the PAMPS-b-PMMA 7 of 12
copolymers were insoluble in water and soluble in dichloromethane.
monomer/catalyst/initiator to obtain a low molecular weight macroinitiator that would facilitate 1
Detailedgrowth
structural information
of the second for the
polymer block. WithPAMPS-b-PMMA
this method, PAMPS-b-PMMA synthesized was obtained
block copolymers with 10by H-NMR.
As can be seenandin25%Figure
mol PAMPS4, the PAMPS-b-PMMA
were spectra
synthesized. In contrast exhibit
to the AMPS signals from
homopolymer, both PAMPS and PMMA
the PAMPS-b-PMMA
copolymers
polymeric blocks. Thewere insoluble signals
PAMPS in water and
at soluble in dichloromethane.
3.22, 1.92, 1.58 and 1.37 ppm correspond to methylene
Detailed structural information for the PAMPS-b-PMMA synthesized was obtained by 1H–NMR.
protons (d) Asadjacent to the acid group, polymer backbone protons (a, b) and methyl protons (c),
can be seen in Figure 4, the PAMPS-b-PMMA spectra exhibit signals from both PAMPS and
respectively PMMA
[29,42].polymeric
The assignments of PMMA
blocks. The PAMPS signalsareat shown directly
3.22, 1.92, 1.58 andin1.37theppm
spectra in accordance
correspond to with
the literaturemethylene
[24]. The protons (d) adjacent
composition oftothe
thecopolymers
acid group, polymer backbone by
was obtained 1 H-NMR
protons (a, b) and
frommethyl
the integration
protons (c), respectively [29,42]. The assignments of PMMA are shown directly in the spectra in
of the PAMPS signal at 3.08 ppm (signal d) and at 3.65 ppm (signal g) for PMMA and the factors 2 and
accordance with the literature [24]. The composition of the copolymers was obtained by 1H–NMR
3 derive from thetheinvolved
from integrationprotons in each
of the PAMPS signalsignal, according
at 3.08 ppm to the
(signal d) and relationship:
at 3.65 ppm (signal g) for PMMA
and the factors 2 and 3 derive from the involved protons in each signal, according to the relationship:
 
3(I3
3.08Ippm
3.08) ppm
Molar compositionPAMPS = (3)
Molar composition PAMPS = 3(I
3.08 ppm ) + 2(I
 3.65 ppm ) (3)
3 I3.08 ppm + 2(I3.65 ppm )
In Figure 4, we present the spectrum of two copolymers with an estimated block composition
for the PAMPS/PMMA diblock of 12/88 and 25/75 (% mol).

CD OD
3 f

bc
ae
12 % PAMPS d

25 % PAMPS

4.0 3.5 3.0 2.0 1.5 1.0


ppm

Figure 4. 1 H-NMR
Figure 4. 1(CD 3 OD)
H–NMR (CDof poly(2-acryloamido-2-methylpropanesulfonic
3OD) acid)-b-poly(methyl
of poly(2-acryloamido-2-methylpropanesulfonic acid)-b-poly(methyl
methacrylate)
methacrylate) (PAMPS-b-PMMA)
(PAMPS-b-PMMA) withwith 12 and
12 and 25% mol
25% mol ofof
PAMPS.
PAMPS.
The GPC elugrams (Figure 5) confirmed the diblock structure of the PAMPS-b-PMMA
In Figure 4, we present
copolymers. the spectrum
A significant of two in
increase is observed copolymers
molecular weightwithin an estimated
copolymers withblock
12 and composition
25% for
the PAMPS/PMMA
mol PAMPS diblock of 12/88
from 37.050 and
to 78.520 and25/75 (% mol).
from 42.400 to 68.000, respectively, with a reduction in the Ð
The GPC with respect to(Figure
elugrams the PAMPS macroinitiator,
5) confirmed theindicating
diblock the efficient of
structure formation of the copolymer. copolymers.
the PAMPS-b-PMMA
Furthermore, the curves of the copolymers were monodispersed and symmetrical, attributed to the
A significantabsence
increase is observed
of impurities or deadin chains
molecularof the weight in copolymers
macroinitiator [21, 43]. It is with
worth 12 andthat
noting 25% themol PAMPS
from 37.050 to 78.520 and
copolymers from bimodal
presented 42.400 to 68.000,
curves whenrespectively,
they originated with a reduction
from a PAMPS in the Ð with respect to
macroinitiator
synthesized in a reaction
the PAMPS macroinitiator, indicatingof lessthe
than 16 h, due
efficient to the propagation
formation of the MMA monomer
of the copolymer. with the curves
Furthermore,
unreacted AMPS monomer [11]. It is important to mention that the one-pot sequential method has
of the copolymers were monodispersed
the disadvantage that the propagationand of thesymmetrical,
second monomersattributed
could involveto the absence
a mixture of impurities or
of the second
dead chainsmonomer
of the macroinitiator
with unreacted first [21,43].
monomer, Itresulting
is worth in anoting that asthe
second block copolymers
a random copolymer presented
[11]. In bimodal
this case, such “contamination” of the second block in the PAMPS-b-PMMA
curves when they originated from a PAMPS macroinitiator synthesized in a reaction of less than 16 h, copolymerization would

due to the propagation of the MMA monomer with unreacted AMPS monomer [11]. It is important to
mention that the one-pot sequential method has the disadvantage that the propagation of the second
Polymers 2020, 12, 1663 8 of 12

monomers could involve a mixture of the second monomer with unreacted first monomer, resulting in
Polymers
a second 2020, 12, xas
block a random copolymer [11]. In this case, such “contamination” of the second block 8 of 12
in
the PAMPS-b-PMMA copolymerization would be difficult to identify. However, we did not observe
be difficult
a tailing to identify.
of the molecularHowever, we did not
mass distribution observe
towards a tailing
lower molarofmasses
the molecular
or bimodal mass distribution
curves in GPC
towards lower molar
chromatogram masses
(see Figure 5). or bimodalwe
Moreover, curves in GPC
identified thechromatogram (see Figure
characteristic signals of the5).
twoMoreover,
PAMPS and we
identified the characteristic signals of the two PAMPS and PHEMA blocks in
PHEMA blocks in RMN spectra that cannot be identified in random PAMPS-co-PMMA, in particular RMN spectra that
cannot be identified
the signal at 3.08 ppm in(signal
random PAMPS-co-PMMA,
d in in particular
Figure 4) corresponding the signal
to the protons of at 3.08 ppmclosest
methylene (signaltodthe
in
Figure
sulfonic4)acid
corresponding
group [1,10].to the protons of methylene closest to the sulfonic acid group [1,10].

Figure
Figure 5. GPC
5. GPC elugram
elugram of PAMPS-b-PMMA
of PAMPS-b-PMMA copolymers
copolymers withwith (a) 12%
(a) 12% molmol
andand (b) 25%
(b) 25% molmol of
of AMPS.
AMPS.
3.2.2. Synthesis of PAMPS-b-PHEMA Block Copolymer
3.2.2.Copolymers
Synthesis of PAMPS-b-PHEMA
with double hydrophilic Block Copolymer
blocks are also highly attractive, because most of them
are biocompatible
Copolymers with and present
double stimuli-responsive
hydrophilic blocks are also andhighly
adjustable amphiphilic
attractive, because mostproperties
of themwith
are
self-assembly capacity in solution [44,45]. Poly(2-acryloamido-2-methylpropanesulfonic
biocompatible and present stimuli-responsive and adjustable amphiphilic properties with self- acid)-b-poly
(2-hydroxyethylmethacrylate)
assembly capacity in solution [44,45]. (PAMPS-b-PHEMA) block copolymers were successfully
Poly(2-acryloamido-2-methylpropanesulfonic synthesized
acid)-b-poly(2-
in the system catalyzed by(PAMPS-b-PHEMA)
hydroxyethylmethacrylate) [Ru(o-C6 H4 -2-py)(phen)(MeCN)
block copolymers 2 ]PF6 were
usingsuccessfully
PAMPS as synthesized
macroinitiator.in
PAMPS-b-PHEMA copolymers were synthesized with molar compositions
the system catalyzed by [Ru(o-C6H4-2-py)(phen)(MeCN)2]PF6 using PAMPS as macroinitiator. of PHEMA between 28
and 36% mol. In polymerization
PAMPS-b-PHEMA copolymers were reactions with higher
synthesized with HEMA concentrations,
molar compositions of obtained
PHEMA copolymers
between 28
were insoluble in the solvents, such as DMF, DMSO and methanol, even
and 36% mol. In polymerization reactions with higher HEMA concentrations, obtained copolymers upon heating. This suggests
that some degree of crosslinking occurred during the polymerization,
were insoluble in the solvents, such as DMF, DMSO and methanol, even upon heating. This suggestspossibly due to intermolecular
interactions
that and the
some degree formation ofoccurred
of crosslinking hydrogen bondsthe
during between the different
polymerization, functional
possibly due togroups present in
intermolecular
the two blocks. In addition, it was necessary to reduce the reaction temperature to 70 ◦ C on adding the
interactions and the formation of hydrogen bonds between the different functional groups present in
second
the two charge
blocks.ofInmonomer
addition, to it minimize
was necessary the formation
to reduceofthe insoluble
reactioncrosslinked
temperature products.
to 70 °C on adding
1
the second charge of monomer to minimize the formation of insoluble crosslinkedsynthesized
Figure 6 shows the H-NMR spectrum and the GPC curves of the copolymer products. with a
molar composition percentage of 65/35 (% mol, PAMPS/PHEMA). The 1 H-NMR analysis confirmed
Figure 6 shows the H-NMR spectrum and the GPC curves of the copolymer synthesized with a
1

the chemical
molar structure
composition of the copolymer
percentage of 65/35 (% andmol,
the characteristic
PAMPS/PHEMA). signals Theof 1the two PAMPS
H-NMR analysis and PHEMA
confirmed
blocks were identified, matching those reported in the literature [21,46].
the chemical structure of the copolymer and the characteristic signals of the two PAMPS and PHEMA The molar composition
of the were
blocks copolymers
identified,wasmatching
calculated by reported
those integration of the
in the peaks [21,46].
literature at 3.43 The and molar
3.86 ppm associated
composition of
with the methylene protons adjacent to the –SO H group of PAMPS and
the copolymers was calculated by integration of the peaks at 3.43 and 3.86 ppm associated with the
3 the –OH group of PHEMA,
respectively.protons
methylene The GPCadjacent
analysis to alsothe
confirmed
–SO3H the diblock
group structureand
of PAMPS of thethePAMPS-b-PHEMA
–OH group of copolymer.PHEMA,
The molecular weight of the copolymer increased when compared
respectively. The GPC analysis also confirmed the diblock structure of the PAMPS-b-PHEMA to the PAMPS macroinitiator and
the GPC curve
copolymer. Theof molecular
the copolymer weightwas ofbroader (Ð = 1.92), this
the copolymer was due
increased the percentage
when comparedoftodead the chains
PAMPS in
the PAMPS macroinitiator as result of high molecular weight and
macroinitiator and the GPC curve of the copolymer was broader (Ð = 1.92), this was due the high percentage conversion of the
first block [11].
percentage of dead chains in the PAMPS macroinitiator as result of high molecular weight and high
percentage conversion of the first block [11].
Polymers 2020,12,
Polymers2020, 12,1663
x 99ofof12
12

Figure 1 H-NMR (D O) y GPC elugram of poly(2-acryloamido-2-methylpropanesulfonic


Figure 6.
6. 1H–NMR (D2O)2 y GPC elugram of poly(2-acryloamido-2-methylpropanesulfonic acid)-b-
acid)-b-poly(2-hydroxyethylmethacrylate)
poly(2-hydroxyethylmethacrylate) (PAMPS-b-PHEMA) catalyzed
(PAMPS-b-PHEMA) catalyzed byby [Ru(o-C 6 H4 -2-py)
[Ru(o-C 6H4-2-
(phen)(MeCN) ]PF
py)(phen)(MeCN)2]PF6.
2 6 .

In
Ingeneral,
general, the
the copolymerization
copolymerization of of PAMPS-b-PMMA
PAMPS-b-PMMAand andPAMPS-b-PHEMA
PAMPS-b-PHEMAwas wascarried
carriedout
outin
in a quasiliving polymerization without the protection and deprotection of the AMPS
a quasiliving polymerization without the protection and deprotection of the AMPS monomer, which monomer,
which demonstrated
demonstrated the efficiency
the efficiency of theofcyclometalated
the cyclometalated [Ru(o-C
[Ru(o-C 6 H4 -2-py)(phen)(MeCN)2 ]PF6 complex
6H4-2-py)(phen)(MeCN)2]PF6 complex and
and confirmed
confirmed the the living
living character
character of this
of this polymerization
polymerization system
system in DMF.
in DMF.
3.3. Ion Exchange Capacity (IEC) of Membranes
3.3. Ion Exchange Capacity (IEC) of Membranes
The PAMPS-b-PHEMA and PAMPS-b-PMMA copolymers showed excellent solubility in various
The PAMPS-b-PHEMA and PAMPS-b-PMMA copolymers showed excellent solubility in
solvents, which allowed obtaining flexible and transparent membranes. These membranes were used
various solvents, which allowed obtaining flexible and transparent membranes. These membranes
to calculate the ion exchange capacity (IEC, mmol H+ /g) or the amount of the ion-exchangeable [H+ ]
were used to calculate the ion exchange capacity (IEC, mmol H+/g) or the amount of the ion-
sulfonic groups, and the results are presented in Table 2.
exchangeable [H+] sulfonic groups, and the results are presented in Table 2.
The IEC
Table 2. Ionvalues of capacity
exchange the PAMPS-b-PMMA membranes
(IEC) of PAMPS-b-PMMA ranged from
and crosslinked 0.55 to 1.40 mmol
PAMPS-b-PHEMA H+/g, very
membranes.
close to the theoretical values. The PAMPS-b-PHEMA membrane was previously crosslinked with
sulfosuccinic acid (SSA) at 120 °C forPAMPS 1 h to preventTheoretical IEC
it from dissolving in Experimental
aqueous NaOH IEC solution.
Copolymer + g–1 ) + g–1 )
%mol (mmol H
The experimental IEC value of PAMPS-b-PHEMA was slightly higher (3.35 mmol H /g) than the (mmol H +

PAMPS-b-PMMA
theoretical value (3.10 mmol H+/g), due 12 to the contribution 0.58
of –SO3H groups of the 0.55
SSA crosslinking
25 1.21
agent. Thus, the crosslinking reaction in PAMPS-b-PHEMA membrane has a dual function: 1) To 1.40
PAMPS-b-PHEMA 65 3.13 3.35
make the membrane more stable, and 2) to increase the IEC or the number of acid sites of the
materials. The IEC values of these membranes depend strongly on the molar composition of PAMPS
Thediblock
in the IEC values of the and
copolymer PAMPS-b-PMMA membranes
the balance between ranged from 0.55 to behavior
hydrophilic-hydrophobic 1.40 mmol H+the
from /g,
very close to used
comonomer the theoretical
(MMA or values.
HEMA).The PAMPS-b-PHEMA membrane was previously crosslinked
with sulfosuccinic acid (SSA) at 120 ◦ C for 1 h to prevent it from dissolving in aqueous NaOH solution.
The experimental
Table 2. Ion IEC valuecapacity
exchange of PAMPS-b-PHEMA was slightly
(IEC) of PAMPS-b-PMMA andhigher (3.35 mmol
crosslinked H+ /g) than the
PAMPS-b-PHEMA
theoretical value (3.10 mmol H+ /g), due to the contribution of –SO3 H groups of the SSA crosslinking
membranes.
agent. Thus, the crosslinking reaction in PAMPS-b-PHEMA membrane has a dual function: 1) To make
PAMPS Theoretical IEC Experimental IEC
the membrane Copolymer
more stable, and 2) to increase the IEC or the number of acid sites +of–1the materials.
%mol (mmol H+g–1) (mmol H g )
The IEC values of these membranes depend strongly on the molar composition of PAMPS in the
PAMPS-b-PMMA 12 0.58 0.55
diblock copolymer and the balance between hydrophilic-hydrophobic behavior from the comonomer
25 1.21 1.40
used (MMA or HEMA).
PAMPS-b-PHEMA 65 3.13 3.35
4. Conclusions
4. Conclusions
The polymerization of AMPS without previous protection of its strong acid groups catalyzed by
TheHpolymerization
[Ru(o-C of AMPS without previous◦ protection of its strong acid groups catalyzed by
6 4 -2-py)(phen)(MeCN)2 ]PF6 in DMF at 80 C was investigated. The radical polymerization
[Ru(o-C6H4-2-py)(phen)(MeCN)2]PF6 in DMF at 80 °C was investigated. The radical polymerization
Polymers 2020, 12, 1663 10 of 12

was carried out successfully and its living character was confirmed by the syntheses of PAMPS-b-PMMA
and PAMPS-b-PHEMA. The well-defined copolymers were obtained by one-pot sequential method,
with the addition of the second monomer (MMA or HEMA) into a PAMPS macroinitiator.
The combination of PAMPS macroinitiator with MMA and HEMA comonomer produced an amphiphilic
block copolymer and a double hydrophilic block copolymer, respectively. A higher molar composition
of PAMPS increases the ionic exchange capacity of the membranes, which can be attributed to
the greater number of sulfonic acid groups present. Given the well-defined structure, IEC and
excellent membrane-forming ability of block copolymers, these materials could be adequate for use
in polyelectrolyte membrane fuel cells or as catalytic membranes for biodiesel production. We are
working on these main applications and the results will be reported in a forthcoming publication,
where, in particular, the relationship between the morphology of the block copolymer membranes
with diffusion and permeability properties will be discussed.

Author Contributions: Data curation and investigation, V.M.-C.; validation and investigation, J.V.-R., V.R.-G. and
M.C.-D.; methodology, E.H.-N. and A.A.-O.; formal analysis, writing—review and editing, R.L.L.; writing—original
draft preparation, supervision and funding acquisition, M.O.G.-D. All authors have read and agreed to the
published version of the manuscript.
Funding: This research was funded by Ciencia Básica CONACYT grant number 2016-286973.
Acknowledgments: The authors thank LANNBIO CINVESTAV-Merida (projects FOMIX-Yucatan 2008–108160
and CONACYT LAB-2009-01 for NMR spectroscopic analysis. V.M.-C. acknowledges the support of CONACYT
through the postdoctoral research grant 711302, M.O.G-D. acknowledges Cátedras CONACYT project No. 3139
and R.L.L. thanks the DGAPA-UNAM (PAPIIT project IN-207419).
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Ahmadian-Alam, L.; Kheirmand, M.; Mahdavi, H. Preparation, Characterization and Properties of PVDF-g
PAMPS/PMMA-Co-PAMPS/Silica Nanoparticle as a New Proton Exchange Nanocomposite Membrane.
Chem. Eng. J. 2016, 284, 1035–1048. [CrossRef]
2. Ahmadian-Alam, L.; Mahdavi, H. Preparation and Characterization of PVDF-Based Blend Membranes as
Polymer Electrolyte Membranes in Fuel Cells: Study of Factor Affecting the Proton Conductivity Behavior.
Polym. Adv. Technol. 2018, 29, 2287–2299. [CrossRef]
3. Ganguly, S.; Maity, T.; Mondal, S.; Das, P.; Das, N.C. Starch Functionalized Biodegradable Semi-IPN as a
PH-Tunable Controlled Release Platform for Memantine. Int. J. Biol. Macromol. 2017, 95, 185–198. [CrossRef]
4. Corzo-González, Z.; Loria-Bastarrachea, M.I.; Hernández-Nuñez, E.; Aguilar-Vega, M.; González-Díaz, M.O.
Preparation and Characterization of Crosslinked PVA/PAMPS Blends Catalytic Membranes for Biodiesel
Production. Polym. Bull. 2017, 74, 2741–2754. [CrossRef]
5. Benkhaled, B.T.; Hadiouch, S.; Olleik, H.; Perrier, J.; Ysacco, C.; Guillaneuf, Y.; Gigmes, D.; Maresca, M.;
Lefay, C. Elaboration of Antimicrobial Polymeric Materials by Dispersion of Well-Defined Amphiphilic
Methacrylic SG1-Based Copolymers. Polym. Chem. 2018, 9, 3127–3141. [CrossRef]
6. Muñoz-Bonilla, A.; Fernández-García, M. Poly(Ionic Liquid)s as Antimicrobial Materials. Eur. Polym. J. 2018,
105, 135–149. [CrossRef]
7. Williams, P.A. Handbook of Industrial Water Soluble Polymers, 1st ed.; Wiley-Blackwell: Iowa City, IA, USA, 2007.
8. Atta, A.M.; El-Mahdy, G.A.; Allohedan, H.A.; Abdullah, M.M.S. Synthesis and Application of Poly Ionic
Liquid-Based on 2-Acrylamido-2-Methyl Propane Sulfonic Acid as Corrosion Protective Film of Steel. Int. J.
Electrochem. Sci. 2015, 10, 6106–6119.
9. Son, Y.J.; Kim, S.J.; Kim, Y.J.; Jung, K.H. Selective Vapor Permeation Behavior of Crosslinked PAMPS
Membranes. Polymers 2020, 12, 987. [CrossRef]
10. Shen, Y.; Xi, J.; Qiu, X.; Zhu, W. A New Proton Conducting Membrane Based on Copolymer of
Methyl Methacrylate and 2-Acrylamido-2-Methyl-1-Propanesulfonic Acid for Direct Methanol Fuel Cells.
Electrochim. Acta 2007, 52, 6956–6961. [CrossRef]
11. Odian, G. Principles of Polymerization, 4th ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2004.
Polymers 2020, 12, 1663 11 of 12

12. Iván, B. Macromolecular Nomenclature Note No. 19: Terminology and Classification of Quasiliving
Polymerizations and Ideal Living Polymerizations on the Basis of the Logic of Elementary Polymerization
Reactions, and Comments on Using the Term “Controlled”. Macromol. Chem. Phys. 2000, 201, 2621–2628.
13. Matyjaszewski, K. Advanced Materials by Atom Transfer Radical Polymerization. Adv. Mater. 2018,
30, 1706441. [CrossRef]
14. Masci, G.; Giacomelli, L.; Crescenzi, V. Atom Transfer Radical Polymerization of Sodium 2-Acrylamido-2-
Methylpropanesulfonate. J. Polym. Sci. Part A Polym. Chem. 2005, 43, 4446–4454. [CrossRef]
15. Masci, G.; Diociaiuti, M.; Crescenzi, V. ATRP Synthesis and Association Properties of Thermoresponsive
Anionic Block Copolymers. J. Polym. Sci. Part A Polym. Chem. 2008, 46, 4830–4842. [CrossRef]
16. Tolstov, A.; Gromadzki, D.; Netopilík, M.; Makuška, R. Aqueous AGET ATRP of Sodium 2-Acrylamido-2-
Methyl-N-Propane Sulfonate Yielding Strong Anionic Comb Polyelectrolytes. E-Polymers 2012, 1–12.
[CrossRef]
17. Xiao, C.; Lin, J. PAMPS-Graft-Ni3 Si2 O5 (OH)4 Multiwalled Nanotubes as a Novel Nano-Sorbent for the
Effective Removal of Pb(Ii) Ions. Rsc Adv. 2020, 10, 7619–7627. [CrossRef]
18. Motlaq, V.F.; Momtazi, L.; Zhu, K.; Knudsen, K.D.; Nyström, B. Differences in Self-Assembly Features of
Thermoresponsive Anionic Triblock Copolymers Synthesized via One-Pot or Two-Pot by Atom Transfer
Radical Polymerization. J. Polym. Sci. Part. B Polym. Phys. 2019, 57, 524–534. [CrossRef]
19. Jing, X.; Huang, Z.; Lu, H.; Wang, B. Use of a Hydrophobic Associative Four-Armed Star Anionic Polymer
to Create a Saline Aqueous Solution of CO2-Switchability. J. Dispers. Sci. Technol. 2017, 38, 1698–1704.
[CrossRef]
20. McCullough, L.A.; Dufour, B.; Matyjaszewski, K. Incorporation of Poly(2-Acrylamido-2-Methyl-N-
Propanesulfonic Acid) Segments into Block and Brush Copolymers by ATRP. J. Polym. Sci. Part A
Polym. Chem. 2009, 47, 5386–5396. [CrossRef]
21. Gonzalez-Diaz, M.O.; Morales, S.L.; Lagadec, R.L.; Alexandrova, L. Homogeneous Radical Polymerization
of 2-Hydroxyethyl Methacrylate Mediated by Cyclometalated Cationic Ruthenium(II) Complexes with PF6−
and Cl− in Protic Media. J. Polym. Sci. Part A Polym. Chem. 2011, 49, 4562–4577. [CrossRef]
22. Das, K.; Dutta, M.; Das, B.; Srivastava, H.K.; Kumar, A. Efficient Pincer-Ruthenium Catalysts for Kharasch
Addition of Carbon Tetrachloride to Styrene. Adv. Synth. Catal. 2019, 361, 2965–2980. [CrossRef]
23. Nishizawa, K.; Ouchi, M.; Sawamoto, M. Design of a Hydrophilic Ruthenium Catalyst for Metal-Catalyzed
Living Radical Polymerization: Highly Active Catalysis in Water. Rsc Adv. 2016, 6, 6577–6582. [CrossRef]
24. Alfredo, N.V.; Jalapa, N.E.; Morales, S.L.; Ryabov, A.D.; Le Lagadec, R.; Alexandrova, L. Light-Driven
Living/Controlled Radical Polymerization of Hydrophobic Monomers Catalyzed by Ruthenium(II)
Metalacycles. Macromolecules 2012, 45, 8135–8146. [CrossRef]
25. Martínez Cornejo, V.; Olvera Mancilla, J.; López Morales, S.; Oviedo Fortino, J.A.; Hernández-Ortega, S.;
Alexandrova, L.; Le Lagadec, R. Synthesis and Comparative Behavior of Ruthena(II)Cycles Bearing Benzene
Ligand in the Radical Polymerization of Styrene and Vinyl Acetate. J. Organomet. Chem. 2015, 799–800,
299–310.
26. García Vargas, M.; Mendoza Aquino, G.; Aguilar Lugo, C.; López Morales, S.; Torres González, J.E.;
Le Lagadec, R.; Alexandrova, L. Living Radical Polymerization of Hydrophobic Monomers Catalyzed
by Cyclometalated Ruthenium(II) Complexes: Improved Control and Formation of Block Co-Polymers.
Eur. Polym. J. 2018, 108, 171–181. [CrossRef]
27. Barbosa, A.S.L.; Werlé, C.; Colunga, C.O.O.; Rodríguez, C.F.; Toscano, R.A.; Le Lagadec, R.; Pfeffer, M.
Further Insight into the Lability of MeCN Ligands of Cytotoxic Cycloruthenated Compounds: Evidence
for the Antisymbiotic Effect Trans to the Carbon Atom at the Ru Center. Inorg. Chem. 2015, 54, 7617–7626.
[CrossRef]
28. Ryabov, A.D.; Sukharev, V.S.; Alexandrova, L.; Le Lagadec, R.; Pfeffer, M. New Synthesis and New
Bio-Application of Cyclometalated Ruthenium(II) Complexes for Fast Mediated Electron Transfer with
Peroxidase and Glucose Oxidase. Inorg. Chem. 2001, 40, 6529–6532. [CrossRef]
29. Nadim, E.; Bouhendi, H.; Ziaee, F.; Nouri, A. Kinetic Study of the Aqueous Free-Radical Polymerization
of 2-Acrylamido-2-Methyl-1-Propanesulfonic Acid via an Online Proton Nuclear Magnetic Resonance
Technique. J. Appl. Polym. Sci. 2012, 126, 156–161. [CrossRef]
Polymers 2020, 12, 1663 12 of 12

30. Mincheva, R.; Paneva, D.; Mespouille, L.; Manolova, N.; Rashkov, I.; Dubois, P. Optimized Water-Based
ATRP of an Anionic Monomer: Comprehension and Properties Characterization. J. Polym. Sci. Part A
Polym. Chem. 2009, 47, 1108–1119. [CrossRef]
31. Kazantsev, O.A.; Igolkin, A.V.; Shirshin, K.V.; Kuznetsova, N.A.; Spirina, A.N.; Malyshev, A.P. Spontaneous
Polymerization of 2-Acrylamido-2-Methylpropanesulfonic Acid in Acidic Aqueous Solutions. Russ. J. Appl.
Chem. 2002, 75, 465–469.
32. Ouchi, M.; Terashima, T.; Sawamoto, M. Transition Metal-Catalyzed Living Radical Polymerization: Toward
Perfection in Catalyst and Precision Polymer Synthesis. Chem. Rev. 2009, 109, 4963–5050. [CrossRef]
33. Kennedy, J.P.; Ivan, B. Designed Polymers by Carbocationic Macromolecular Engineering: Theory and Practice;
Hanser Publishers: Munich, Germany; New York, NY, USA, 1992.
34. Iván, B. Macromolecular Nomenclature Note No. 19—Terminology and Classification of Quasiliving
Polymerization and Ideal Living Polymerization on the Basis of the Logic Elementary Polymerization
Reactions, and Comments on Using the Term “Controlled”. Polym. Prepr. 2000, 41, 6–13.
35. Robinson, K.L.; Khan, M.A.; De Paz Báñez, M.V.; Wang, X.S.; Armes, S.P. Controlled Polymerization of
2-Hydroxyethyl Methacrylate by ATRP at Ambient Temperature. Macromolecules 2001, 34, 3155–3158.
[CrossRef]
36. Ye, Y.S.; Rick, H.J.; Joe, B. Water Soluble Polymers as Proton Exchange Membranes for Fuel Cells. Polymers
2012, 4, 913–963. [CrossRef]
37. Oikonomou, E.K.; Bethani, A.; Bokias, G.; Kallitsis, J.K. Poly(Sodium Styrene Sulfonate)-b-Poly(Methyl
Methacrylate) Diblock Copolymers through Direct Atom Transfer Radical Polymerization: Influence of
Hydrophilic–Hydrophobic Balance on Self-Organization in Aqueous Solution. Eur. Polym. J. 2011, 47,
752–761. [CrossRef]
38. Peng, C.H.; Kong, J.; Seeliger, F.; Matyjaszewski, K. Mechanism of Halogen Exchange in ATRP. Macromolecules
2011, 44, 7546–7557. [CrossRef]
39. Datta, H.; Singha, N.K.; Bhowmick, A.K. Beneficial Effect of Nanoclay in Atom Transfer Radical Polymerization
of Ethyl Acrylate: A One Pot Preparation of Tailor-Made Polymer Nanocomposite. Macromolecules 2008, 41,
50–57. [CrossRef]
40. Wang, X.S.; Armes, S.P. Facile Atom Transfer Radical Polymerization of Methoxy-Capped Oligo(Ethylene
Glycol) Methacrylate in Aqueous Media at Ambient Temperature. Macromolecules 2000, 33, 6640–6647.
[CrossRef]
41. Reining, B.; Keul, H.; Höcker, H. Block Copolymers Comprising Poly(Ethylene Oxide) and Poly(Hydroxyethyl
Methacrylate) Blocks: Synthesis and Characterization. Polymer 2002, 43, 3139–3145. [CrossRef]
42. Szabó, Á.; Wacha, A.; Thomann, R.; Szarka, G.; Bóta, A.; Iván, B. Synthesis of Poly(Methyl Methacrylate)-
Poly(Poly(Ethylene Glycol) Methacrylate)-Polyisobutylene ABCBA Pentablock Copolymers by Combining
Quasiliving Carbocationic and Atom Transfer Radical Polymerizations and Characterization Thereof.
J. Macromol. Sci. Part A Pure Appl. Chem. 2015, 52, 252–259. [CrossRef]
43. Sumerlin, B.S.; Donovan, M.S.; Mitsukami, Y.; Lowe, A.B.; McCormick, C.L. Water-Soluble Polymers. 84.
Controlled Polymerization in Aqueous Media of Anionic Acrylamido Monomers via RAFT. Macromolecules
2001, 34, 6561–6564. [CrossRef]
44. Bucholz, T.L.; Loo, Y.-L.L. Polar Aprotic Solvents Disrupt Interblock Hydrogen Bonding and Induce
Microphase Separation in Double Hydrophilic Block Copolymers of PEGMA and PAAMPSA. Macromolecules
2008, 41, 4069–4070. [CrossRef]
45. Agut, W.; Brûlet, A.; Taton, D.; Lecommandoux, S. Thermoresponsive Micelles from Jeffamine-b-Poly
(L-Glutamic Acid) Double Hydrophilic Block Copolymers. Langmuir 2007, 23, 11526–11533. [CrossRef]
[PubMed]
46. Hajibeygi, M.; Shafiei-Navid, S.; Shabanian, M.; Vahabi, H. Novel Poly(Amide-Azomethine) Nanocomposites
Reinforced with Polyacrylic Acid-Co-2-Acrylamido-2-Methylpropanesulfonic Acid Modified LDH: Synthesis
and Properties. Appl. Clay Sci. 2018, 157, 165–176. [CrossRef]

© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like