Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

European Polymer Journal 76 (2016) 156–169

Contents lists available at ScienceDirect

European Polymer Journal


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

Macromolecular Nanotechnology

PEI-based hydrogels with different morphology and sizes:


Bulkgel, microgel, and cryogel for catalytic energy and
environmental catalytic applications
Nurettin Sahiner a,b,⇑, Sahin Demirci a
a
Faculty of Science & Arts, Chemistry Department, Canakkale Onsekiz Mart University, Terzioglu Campus, 17100 Canakkale, Turkey
b
Nanoscience and Technology Research and Application Center (NANORAC), Canakkale Onsekiz Mart University, Terzioglu Campus, 17100 Canakkale, Turkey

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

Article history: Polyethyleneimine (PEI)-based hydrogels with different morphologies as bulkgel, microgel,
Received 7 November 2015 and a cryogel form were synthesized, and used as templates for in situ Co, Ni, and Cu metal
Received in revised form 27 January 2016 nanoparticles preparation. Because of the size, porosity and morphological differences of the
Accepted 28 January 2016
prepared PEI-based hydrogels (bulkgel, microgel and cryogel), interesting and versatile
Available online 29 January 2016
applications were shown. The prepared PEI-M (M: Co, Ni) hydrogel composite systems were
used as catalysts for the hydrolysis of NaBH4 for H2 production, in the reduction reaction of
Keywords:
4-, and 2- nitrophenols (4-NP, and 2-NP) to their corresponding amino forms, 4-AP and 2-AP,
Polyethyleneimine
Bulkgel/microgel/nanogel/cryogel
respectively, and in the decolorization reaction of methylene blue (MB) and Eosin Y (EY) dyes
Hydrogel–metal composites (EY) (M: Cu). It was shown that without a metal catalyst, bare NaBH4 is not good source of H2
Catalyst sources e.g., in the hydrolysis of NaBH4 for H2 production, the self-hydrolysis of NaBH4 can
only produce 152 ± 2.5 mL H2 in 280 min whereas the same amount of NaBH4 can produce
250 mL for under 60 min in the presence of Co nanoparticles within PEI matrices. It was
found that the reduction reactions of 4-, and 2-nitrophenols to 4-, and 2- aminophenols,
and the decolorization of dyes also required metal nanoparticles catalysts within PEI hydro-
gel. The hydrogen generation rate (HGR, ml H2/(g of cat)(min)), and turnover frequency (TOF,
mol H2/(mol cat)(min)) values were calculated and compared for PEI hydrogels with differ-
ent sizes and morphologies. Higher HGR and TOF values were observed in H2 production
using PEI-Co microgel composite catalyst systems than PEI-Co bulk hydrogel and cryogel
composite catalyst systems with 424 ml H2/(g of cat)(min) and 0.98 mol H2/(mol cat)
(min) respectively. The higher TOF value was observed for the reduction reaction of 4-,
and 2-NPs to their corresponding amino forms as 0.881, and 1.323 mol H2/(mol cat)(min),
respectively catalyzed by PEI-Ni cryogel composite systems than PEI-Ni microgel and bulk-
gel composite catalyst system. Better Ea value for H2 production was calculated for PEI-Ni
microgel composite systems with 38.3 J/mol K. And the activation energy values for reduc-
tion of 4-NP to 4-AP by PEI-Cu composite cryogel system, and for reduction of 2-NP to 2-AP
by PEI-Ni composite cryogel system were calculated as 15.2 and 29.6 J/mol K, respectively.
The catalytic performances of PEI-Cu hydrogel composite system for decolorization of MB
and EY in different media such as sea water, creek water, tap water, and distilled water were

⇑ Corresponding author at: Faculty of Science & Arts, Chemistry Department, Canakkale Onsekiz Mart University, Terzioglu Campus, 17100 Canakkale,
Turkey.
E-mail address: sahiner71@gmail.com (N. Sahiner).

http://dx.doi.org/10.1016/j.eurpolymj.2016.01.046
0014-3057/Ó 2016 Elsevier Ltd. All rights reserved.
N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169 157

compared and the PEI-Cu cryogel composite provided the fastest decolorization capabilities
for both dyes. Interestingly, amongst all the media, MB and EY decolorization in seawater by
PEI-Cu cryogel was accomplished in 2 and 9 min with 96.3% and 97.3% decolorization,
respectively, with the same catalyst system.
Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction

Hydrogels are developed by the chemical and physical linking of monomers or polymer chains to generate three-
dimensional structures possessing hydrophilic and/or charge developing functional groups such as –COOH, –NH2, –CONH2,
–NH2, –SO3H, and –PO3H. Due to the ease of preparation methods, abundance of functional groups and inherent physical and
chemical properties, hydrogels are of great significance for the design of advanced materials in biomedical and environmen-
tal engineering applications due to the resemblance to living tissues with their elastic, soft, flexible, and water holding capa-
bility [1–3]. Additionally, hydrogels can be synthesized with different morphologies, sizes, and porosity such as bulk
hydrogel, microgel/nanogel, and cryogel by employing different polymerization techniques [4]. The conventional description
for pore sizes (macroporous >50 nm, mesoporous up to 2 nm, and microporous <2 nm) of solid materials may not be appli-
cable to soft and flexible hydrogels [5,6] as the pore sizes of a special type of hydrogels, cryogels, can be considered super-
porous or hyper porous with interconnected pores up to few hundred micrometers in size [7]. Moreover, as hydrogels swell
in aqueous environments, the pore size is highly changeable depending on the media and stimuli such as pH, ionic strength,
temperature, solvent polarity, saltiness and so on. Because of the functional groups that form complexes with various metal
ions, hydrogel 3-D networks, regardless of their size (bulk, micro, nano) and shape (cylinder, sphere, capsule, hollow etc), can
be used as a template and stabilizing agent for in situ metal nanoparticles and finally employed as catalysts [8–10]. In recent
decades, the use of hydrogels as soft templates with tunable physical and chemical functionalities has attracted great atten-
tion due to the versatility of the chemical functional groups, tunable porosity, versatility of sizes and morphology [8–10].
Most strikingly, many metal nanoparticles have been prepared within these different soft templates and directly employed
as catalysts in anticipation of better catalytic performances compared to bare metal nanoparticles and extended lifespans
against oxidation, inactivation and aggregation [11–13].
Hydrogen is attracting great interest as a most important energy source due to its renewable, clean, carbon free and eco-
friendly properties [14]. The most common and facile method for production is the hydrolysis of metal hydrides e.g., NaBH4,
LiAlH4, or NH3BH3 [15–17] in the presence of metal nanoparticle catalysts such as Co, Ni, Cu, Ru and so on [18–20]. Amongst
the H2 sources, NaBH4 has been intensely investigated due to its cheap and more convenient nature than other H2 sources
such LiAlH4, or NH3BH3, high H2 storage efficiency, stability and easy availability [21].
Phenolic compounds are the most common pollutants in wastewater and they pose a significant threat to human health,
and have harmful effects on organisms even at low concentrations [22,23]. Both 4-NP and 2-NP, the two often encountered
and problematic pollutants in nature, are byproducts formed by various phenolic compound producing industries [24]. Fur-
thermore, the reduction of 4-NP to 4-AP has great significance for various biomedical applications such as preparation of
antipyretic and analgesic drugs, and in the paint and photographic industries [25]. Therefore, the reduction of nitrophenolic
compounds to their aminophenolic compounds is also important and has been extensively investigated by many researchers
as a model reaction to test the catalytic performances of new catalysts [4,9,10,12,13,22,24,25]. Wastewaters discharged from
chemical industries, such as dying/painting facilities and textile industries, are highly controversial and cause significant
environmental pollution to air, soil, and water and eventually threaten human health. The discharges of organic compounds
and dyes not only introduce intensive coloring and toxicity to aquatic environments, but also cause serious health problems
to the ecosystem and living organisms [26]. Methylene blue (MB) and eosin Y (EY) are dyes extensively used in printing and
leather industries, as fluorescent pigments for histological analysis of tissues, and as photo sensitizers in semiconductors
[27–29]. The direct release of wastewater containing EY and MB can cause serious ecological problems due to their dark color
and toxicity [30]. Biological methods are often ineffective for de-colorization of EY and MB due to the stability and complex
aromatic structures of these dyes [31–34]. Therefore, metal nanoparticles are often preferred to eliminate these dyes by
degradation [28,35–37].
In this study, we report the synthesis of PEI hydrogels with different morphology and sizes such as bulk hydrogel, micro-
gel and cryogel by employing the same crosslinker with different polymerization techniques (micro-emulsion, cryopolymer-
ization). Then, metal ions (Co2+, Ni2+, and Cu2+) were loaded into the synthesized PEI hydrogels from their aqueous solutions
of sulfate salts and reduced in situ to the corresponding metal nanoparticles (Co, Ni, and Cu) via NaBH4 treatment employing
the PEI polymeric network as a template. Then these prepared PEI-M hydrogel composite systems (bulk hydrogel, microgel,
and cryogel) were used as a catalyst for the hydrolysis of NaBH4, reduction of 4- and 2-NP to their corresponding amino
forms and decolorization of MB and EY dyes. Various parameters affecting the catalyst properties of the PEI-M hydrogel com-
posite systems such as morphology, size and porosity were investigated.
158 N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169

2. Experimental

2.1. Materials

Polyethyleneimine (PEI, 50% in water, Mn: 1800 g mol 1, Sigma Aldrich) was used for synthesis of PEI-based hydrogels
with different morphology. Dioctyl sulfosuccinate sodium salt (AOT, 98%, Sigma Aldrich) as a surfactant, and gasoline as a
solvent and Glycerol diglycidyl ether (GDE, 6100%, Sigma Aldrich) was used as a crosslinker were used for synthesis of
PEI microgels, cryogels and bulkgels. Cobalt sulfate heptahydrate (CoSO47H2O, 99%, Merck), nickel sulfate heptahydrate
(NiSO47H2O, 99%, Sigma Aldrich), and copper sulfate pentahydrate (CuSO45H2O, 99%, Sigma Aldrich) were used as sources
of metal ions. The 2-nitrophenol (2-NP, 99%, Merck) and 4-nitrophenol (4-NP, 99%, Merck) were used as sources of nitro com-
pounds and methylene blue (MB, 98%, Sigma Aldrich), and Eosin Y (EY, Dye content 99%, Sigma Aldrich) were used as organic
dye sources. Sodium borohydride (NaBH4, 98%, Merck) was used as a reducing agent for metal ions to metal nanoparticles,
and for reduction of nitro compounds to their corresponding amino compounds and decolorization of organic dyes.
Technical-grade ethanol, acetone and DI water were used for washing processes of PEI-based hydrogel materials.

2.2. Synthesis of PEI hydrogels with different morphologies; microgel, cryogel, and bulkgel

The PEI bulkgels were synthesized via epoxy-amine reactions. In short, a certain amount of PEI solution (4 mL, 50% in
water) was placed in a vial, and 4 ml water was added and vortexed. To this solution, a certain amount of GDE crosslinker
(8 mol% based on the repeating unit of PEI) was added and quickly vortex mixed again, then placed into a glass tube and
placed in an autoclave (WAC 80, WiseClave) at 90 °C for 2.5 h. The synthesized PEI bulkgels were washed with water for
8 h, and wash water was changed every 1 h, then dried in oven at 50 °C. The synthesized, washed and dried PEI bulkgels were
stored in a closed zip lock bag for characterization and studies.
For PEI microgel synthesis, the reported method was followed with some modifications [38]. Briefly, a certain amount of
PEI solution (10 mL, 50% in water) was placed in 300 mL 0.1 M lecithin/gasoline solution in a 500 mL flask, and stirred at
1500 rpm for about 2 h. Then, a certain amount of GDE crosslinker (50 mol% of repeating unit of PEI) was added and stirring
continued at 1500 rpm for 2 h more. After that, the formed PEI microgels were washed in 100 mL cyclohexane, then washed
with ethanol, ethanol–water (1:1), water–acetone (1:1) and acetone respectively by centrifuging at 24,700g for 5 min each
time, and dried with a heat gun. Finally, the synthesized, washed and dried PEI microgels were stored in a closed tube for
characterization and further studies.
The PEI cryogels were synthesized by using a cryopolymerization technique [2,4,7]. For this purpose, a certain amount of
PEI solution (1 mL, 50% in water) was placed into a vial and was brought to 10 mL by adding 9 mL DI water and vortex mixed.
Then, the vial was placed into a deep freeze for 5 min, followed by addition of GDE crosslinker (10 mol% based on the repeat-
ing unit of PEI). After vortex mixing to produce a homogenous solution, the mix was placed into plastic straws (8 mm diam-
eter), then quickly placed into a deep freezer at 18 °C for 20 h to complete cryopolymerization. After that, the PEI cryogels
were cut to similar sizes of about 1 cm and 0.8 mm, washed with DI water 5 times and dried in oven at 50 °C. The synthe-
sized, washed and dried PEI cryogels were stored in a closed zip lock bag for characterization and studies.

2.3. Characterization of PEI-based hydrogels

The synthesis of PEI bulkgel, microgel, and cryogel was confirmed with Fourier transform infrared (FT-IR, Nicolet iS10,
Thermo) spectrometer in terms of functional groups. The measurements were done by using attenuated total reflectance
(ATR) apparatus and the %T was recorded between 650 and 4000 cm 1 wave numbers.
The thermogravimetric analysis of PEI-based hydrogels was done with a thermogravimetric analyzer (TGA, SII 6300 DTA/
TG, Exstar). The measurements were carried out in the presence of N2 flow (200 ml/min) by increasing temperature from 90
to 900 °C with 10 °C/min heating rate.
The morphological characterization of PEI-based hydrogels was completed via scanning electron microscopy (SEM, JSM-
5600, JEOL) with an operating voltage of 20 kV. The images were acquired after placing PEI cryogels onto carbon tape-
attached aluminum SEM stubs at room temperature after coating with gold to a few nanometer thickness in a vacuum.
The prepared metal nanoparticles inside PEI-based hydrogels were visualized via transmission electron microscopy (TEM,
JEOL 2010, Japan) in a vacuum with an operating voltage of 150 kV.

2.4. Synthesis of metal nanoparticles within PEI-based hydrogels for various catalytic applications

The use of hydrogels with different size, shape and morphology as a template for in situ metal nanoparticle preparation
has already been reported [8–10]. PEI-based hydrogels can also be used as a template for metal nanoparticles such as Co, Ni,
and Cu. For this purpose, the metal ion solutions of Co(II), Ni(II), and Cu(II) were prepared from their sulfate salts in DI water
and certain amount of PEI-based hydrogels with different morphologies such as bulkgel, microgel, and cryogel (5 g) were
separately placed into 1 L solution containing 1000 ppm metal ion (Co2+, Ni2+, and Cu2+) solution separately under constant
stirring at 750 rpm for 12 h. Then, the metal ion-loaded PEI-based hydrogels were washed with DI water and dried in an oven
N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169 159

at 50 °C. A certain amount of metal ion-loaded dry PEI-based hydrogels were placed into 50 mL 0.1 M NaBH4 solution in DI
water to reduce metal ions to their corresponding metal nanoparticle forms of Co, Ni, and Cu as PEI-M (M: Co, Ni, and Cu)
composite hydrogels. After cessation of gas evolution, the PEI-M composite hydrogels were washed with DI water one time.
The metal contents of prepared PEI-M hydrogel composites were determined with an atomic absorption spectrometer (AAS,
iCE 3000, Thermo). Thus, the PEI-M composite hydrogels were placed in 30 mL 5 M HCl solution and stirred at 500 rpm for
12 h, and after dissolution of metal nanoparticles to the corresponding solution, the metal content in these solutions were
measured with AAS.

2.4.1. Hydrogen production from hydrolysis of NaBH4 using PEI-M based hydrogel composites as catalyst
The prepared PEI-M (M: Co, and Ni) composite hydrogels with different morphologies were used for hydrolysis of NaBH4.
A certain amount of PEI-M composite hydrogels (each one containing 0.183 mmol metal nanoparticles) was placed into a
flask that contained 50 mL DI water, and 0.0965 g NaBH4 (50 mM) was added into this flask and produced H2 was recorded
with time. The H2 generation rate (HGR, ml H2/(g of cat)(min)), and turnover frequency (TOF, mol H2/(mol cat)(min)) values
were calculated. Moreover, the effect of temperature on this hydrolysis reaction was investigated to determine activation
parameters such as Ea, DH, and DS. Additionally, the re-usability studies of PEI-Co composite hydrogels for H2 production
from hydrolysis of NaBH4 were investigated at 70 °C.

2.4.2. Reduction of 2-nitrophenol and 4-nitrophenol


A certain amount of PEI-M (M: Co, Ni, Cu) (0.183 mmol) composite hydrogels was added to previously prepared 0.4 M
NaBH4 in 50 mL 0.1 M 4- and 2- NP solutions. The peaks at 400 and 414 nm wavelengths in UV–Vis spectroscopy for
4-NP and 2-NP reduction, respectively, were measured with a UV–Vis spectrophotometer (T80+, PG Instruments Limited)
at certain time intervals after 80- and 30-fold dilution for 4-NP and 2-NP respectively. The activation parameters such as
Ea, DH, and DS, and TOF values were also determined by completing the reduction reaction at different temperatures of
30, 50 and 70 °C.

2.4.3. Decolorization of Methylene blue and Eosin Y by PEI-Cu composite hydrogels with different morphology
The catalytic degradation of the organic dyes (MB, and EY) was carried out in the presence of NaBH4 (2  10 2 M, 100 mL)
by using PEI-Cu (0.025 g) composite hydrogels as catalyst. The concentrations of MB and EY were 1.6  10 4 M and
4  10 5 M, respectively. During catalytic degradation of these dyes, 0.5 mL and 1 mL samples were taken from the reaction
flask at certain time intervals, and diluted 10-fold for MB and 2-fold for EY, and the change in the intensities of UV–Vis
absorption were measured at 664 and 514 nm, respectively. The reaction conditions for both organic dyes were the same
(30 °C, 500 rpm mixing rate). The reduction reactions were carried out in four different media of DI water, tap water, sea-
water and creek water (Saricay, Canakkale, Turkey).

3. Results and discussion

3.1. Synthesis and characterization of PEI-based hydrogels with different morphology

The molecular structures of PEI and GDE, and the schematic presentation of the crosslinking reaction mechanism are
given in Fig. 1(a). Amine compounds are known to react easily with epoxies, especially primary and secondary amines
are highly reactive toward epoxides [39]. The epoxy-amine, and epoxy-hydroxyl reaction process has been the subject of
many interesting studies [40,41], due to the importance of substituent effects on the reactivity; nonetheless, several issues
are still not fully understood [42]. It is well accepted that epoxy reactions with amines follow the SN2 mechanism where the
nucleophilic nitrogen atom attacks a carbon atom of the epoxy ring, and then the hydrogen atom from the amine eventually
transfers to the epoxide oxygen to form OH. Therefore, a primary amine can react two times with two epoxy groups whereas
a secondary amine can react only once. Tertiary amines do not react with epoxides, as their nulceophilicity is low. The
branched PEI chains contain many primary, secondary, and tertiary amines in their natural structure. Therefore, branched
PEI chains crosslinked with epoxy groups of glycerol diglycidyl ether in different media and conditions for the synthesis
of PEI-based bulk hydrogel, microgel and cryogel. It is obvious that oxirane rings of epoxy groups were opened, and bonds
with amine groups of PEI chains formed. The bulk PEI hydrogel was prepared in an autoclave at 90 °C for 3.5 h after mixing
PEI and GDE crosslinker as described earlier and given in. The digital camera image of synthesized PEI bulk hydrogel is given
in Fig. 1(b), whereas the optical microscopy image of synthesized PEI microgels is illustrated in Fig. 1(c). The PEI microgels
were synthesized by employing a micro-emulsion polymerization technique. A certain amount of branched PEI chains
(10 mL, 50% in water) was reacted with GDE within the reverse micelle of lecithin for 2 h. The SEM image of PEI cryogel
is given in Fig. 1(d). As the cryogels of PEI were prepared under cryogenic conditions at 18 °C, the synthesized PEI cryogels
are superporous because ice crystals are sacrificial pore generating templates, and upon thawing at room temperature inter-
connected super pores are generated. This super porous material with amine functionality is of great significance for many
environmental and catalytic applications, as well as biological molecule separation of DNA, proteins, amino acids, and hor-
mones and so on [4,9,10,38,43].
160 N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169

H
H 2N + NH2
NH2 N

N H N
H N
N H +N N
H
N
H
N + N
N H NH OH
H NH
N
OH
NH2
O NH2
O

HO
HO
O O
NH
N
H H OH
(a)
HO N N
N H + N HO O
N H
H O O
N

n O

Water Lechitin/Gasoline Water


90 oC 3.5 h RT 3 h -18 oC 20 h

(b) (c) (d)

Fig. 1. The schematic presentation of (a) crosslinking reaction of PEI based hydrogels, and digital camera image of (b) PEI bulk hydrogel, and (c) optic
microscope image of PEI microgel, and (d) SEM images of PEI cryogel.

The FT-IR spectrum of PEI-based hydrogels with different morphology that were synthesized by GDE crosslinking are
given in Supp. Fig. 1(a). The N–H stretching at 3286 cm 1, C–H stretching at 2982 and 2856 cm 1, N–H bending at
1653 cm 1, and C–N stretching at 1066 cm 1 are attributed to the branched PEI chains. As most distinct peaks for all studied
PEI hydrogels showed ether peaks at 1113 and 1071 cm 1 from their corresponding FT-IR spectra, confirming the successful
crosslinking of PEI with GDE in the hydrogels. Moreover, the thermal stability of PEI-based hydrogels was also compared via
TG analysis, and the corresponding thermograms are illustrated in Supp. Fig. 1(b). The branched PEI chains degraded up to
99% weight loss at 900 °C, whereas the crosslinked PEI hydrogels degraded by a lesser amount than branched PEI chains e.g.,
95.8%, 98.6%, and 95.1% weight losses for PEI bulk, microgel and cryogel, respectively observed at 900 °C. Depending on the
morphology of PEI hydrogels, different thermal degradation paths were obtained as demonstrated in Supp. Fig. 1(b). In gen-
eral, from the thermograms, due to differences in the amounts of used crosslinker, PEI hydrogels showed different thermal
degradation over different temperature ranges and after about 300 °C, bulk PEI hydrogel showed sharp degradation steps.

3.2. Metal nanoparticle preparation within PEI-based hydrogels

The synthesized, washed and dried PEI-based hydrogels were used as a template and stabilizing network for metal Co, Ni
and Cu nanoparticle preparation. The schematic presentation of this procedure for PEI bulkgel, microgels, and cryogels is
N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169 161

shown in Supp. Fig. 2(a). A certain amount of synthesized PEI bulkgel, microgels, and cryogels (2 g each) were loaded with
corresponding metal ions by adsorption from metal solutions by placing them into 1000 mL 1000 ppm Co2+, Ni2+, and Cu2+
solutions previously prepared from their sulfate salts as described earlier. The metal ion-loaded and dried PEI hydrogels,
whether bulk, micro and cryo, were treated with 50 mL 0.1 M NaBH4 in DI water. As illustrated in Supp. Fig. 2(b), for each
metal ion loaded, PEI hydrogels assumed the color of the metal ion due to complex formation between the amine groups and
M2+ salts; however, upon reduction within PEI hydrogels, a dark coloration is apparent due to corresponding nanoparticles
formation. To estimate the metal content of PEI hydrogels, a certain amount (0.1 g for each) of PEI-M bulk gel, microgel and
cryogel composites were treated 30 mL 5 M HCl to dissolve the metal nanoparticles. Then the metal ions within PEI hydro-
gels were quantified with AAS measurements and corresponding values are given in Table 1. It is obvious that PEI-based
hydrogel whether bulkgel, microgel and cryogel, bind more to Cu2+ ions, followed by Ni2+ and then Co2+ ions, respectively.
The PEI-M (M: Cu, Ni and Co) composite hydrogels contain 3.34 ± 0.31, 3.17 ± 0.19, 3.40 ± 0.27 mmol Cu2+/g, 2.69 ± 0.05,
2.41 ± 0.12, 2.61 ± 0.07 mmol Ni2+/g, and 1.68 ± 0.12, 1.20 ± 0.03, 1.72 ± 0.05 mmol Co2+/g for bulkgel, microgel and cryogel,
respectively. The higher amount of metal ions amongst all the hydrogel templates was measured for bulk PEI-M composite
for all metal ions. This is reasonable, in addition to the used amount of crosslinker e.g., PEI hydrogels with different mor-
phologies were prepared by using different amounts of crosslinkers, the sacrifice of functional groups at the expense of sur-
face area and maybe porosity is the main reason for the lesser amounts of metal ion absorption by PEI microgels. On the
other hand, bulk and cryogel PEI hydrogels with different amounts of metal ion absorption ability are about the same. More-
over, the TEM images of metal nanoparticles in PEI bulkgels, microgels, and cryogels are illustrated in Fig. 2(a) –(c) respec-
tively. The metal nanoparticles within PEI hydrogels are a few tens of nanometers in sizes and they are distributed evenly
throughout the PEI particle network. There is no precise control on the metal nanoparticle size, because in addition to the
difference in size and porosity of PEI hydrogels, the functional groups such as primary, secondary and even tertiary amine
and hydroxyl groups and their distribution and position within the PEI-based hydrogel network affects the metal ion binding
tendency. Moreover, the interaction between the diverse metal ions and different functional groups may be different. Nev-
ertheless, the metal nanoparticles of Co, Ni and Cu can be readily prepared within PEI networks and homogeneously dis-
tributed in their matrices. Within the hydrogel templates, metal nanoparticles are confined and can readily perform
catalytic tasks in aquatic environments due to the hydrophilic nature of hydrogel matrices whether they are bulk, micro
or cryogels. No leakage of metal nanoparticles was observed from hydrogel matrices during the catalytic use hydrogel-M
composites [9–12,18].

3.3. The usage of PEI-M composite hydrogels as a catalyst for hydrolysis of NaBH4

The chemical hydrides such as NaBH4 or NH3BH3 are commonly used H2 sources for the production of H2. As an exother-
mic reaction, the hydrolysis of NaBH4 releases about 300 kJ mol 1, and the reaction can be carried out at ambient temper-
atures readily. Solution of NaBH4 in water hydrolyzes to sodium metaborate (NaBO2) and H2 only when in contact with
specific catalysts producing 2 mol extra H2 coming from the water in solution during hydrolysis reaction. Besides, metal
nanoparticle catalysts are required to use in order to obtain faster and controllable H2 production from hydrolysis of NaBH4
[10–12]. In this study, PEI-M based catalytic system were used for the production of H2 from hydrolysis of NaBH4. Certain
amount of PEI-Co composite were placed into a 50 mL water in a flask and the flask is connected to a water-filled cylinder
after adding 0.0965 g NaBH4. The amount of produced H2 was recorded as a function of time by monitoring replacement of
water from water filled cylinder with the produced H2 gas. The PEI hydrogels, bare bulkgel, micro and cryogel without any
metal nanoparticle within, are tested for H2 production reaction, and the self-hydrolysis of NaBH4 was done as control. For
this purpose, certain amount of NaBH4 (50 mM) was placed into 50 mL DI water and 152 ± 2.5 mL H2 produced in 285 min
(self-hydrolysis), whereas each of PEI hydrogels (bulkgel, micro and cryogel) weighing 0.1 g produced 146 ± 6, 164 ± 7, and
154 ± 3 mL H2, respectively from the hydrolysis of NaBH4 in 280 min as illustrated in Fig. 3(a). Each of PEI-M composite
hydrogels (M: Co, Ni) containing 0.183 mmol M, was placed in the same NaBH4 reaction conditions (50 mL 50 mM NaBH4)
at room temperature. The graphs of H2 production for PEI-Co, and PEI-Ni composite hydrogels are shown in Fig. 3(b), and (c),
respectively. It is clear that Co and Ni metal nanoparticle-containing PEI microgels produced the same amount of H2 faster
than PEI cryogels and bulkgels. Moreover, the H2 production was completed by Co composite hydrogels in 105, 55 and
85 min for bulk hydrogel, microgel and cryogel of PEI, respectively. The H2 production was finished at 180, 85, and
130 min under the same reaction conditions (30 °C and 1000 rpm stirring) for PEI-Ni hydrogel, microgel and cryogel com-
posites, respectively. It is obvious that the H2 generation rate of PEI-Ni composite hydrogels is slower than PEI-Co composite
hydrogels from hydrolysis of NaBH4. The HGR (ml H2/(g of cat)(min)) and TOF (mol H2/(mol of cat)(min)) values were cal-

Table 1
The metal content of PEI-M bulkgel, microgel, and cryogel composites.

Hydrogel morphology Co (mmol/g) Ni (mmol/g) Cu (mmol/g)


PEI bulkgel 1.68 ± 0.12 2.69 ± 0.05 3.34 ± 0.31
PEI microgel 1.20 ± 0.03 2.41 ± 0.12 3.17 ± 0.19
PEI cryogel 1.72 ± 0.05 2.61 ± 0.07 3.40 ± 0.27
162 N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169

(a) PEI bulk hydrogel


Co Ni Cu

1 μm 0.5 μm 0.5 μm

(b) PEI microgel

Co Ni Cu

0.2 μm 0.2 μm 0.5 μm

(c) PEI cryogel

Co Ni Cu

0.2 μm 0.1 μm 0.2 μm

Fig. 2. The TEM images of PEI-M (M: Co, Ni, and Cu) (a) bulkgel, (b) microgel, and (c) cryogel composite systems.

culated for Co and Ni metal nanoparticle-containing PEI bulkgels, microgels and cryogels and are reported in Table 2. The
highest HGR value of PEI-Co composite hydrogels was obtained from PEI microgel with 431 ml H2/(g of cat)(min), whereas
292 and 259 ml H2/(g of cat)(min) were obtained for PEI-Co cryogel and bulk hydrogel, respectively. Similarly, the highest
HGR value for PEI-Ni composite hydrogels was obtained for PEI microgel with 306 ml H2/(g of cat)(min), while PEI-Ni cryogel
and bulk hydrogel provided HGR of 179 and 137 ml H2/(g of cat)(min), respectively. Furthermore, the PEI-Co microgel TOF
value was the highest, 1.02 mol H2/(mol of cat)(min), in comparison to TOF values of PEI-Co bulkgels and cryogels that were
0.59, and 0.67 mol H2/(mol of cat)(min), respectively. The TOF values for PEI-Ni also displayed a similar trend i.e., 0.31, 0.68,
and 0.41 mol H2/(mol cat)(min) for bulk, micro and cryogel composites, respectively. It was indicated that the rate for the
hydrogen production from hydrolysis of NaBH4 depends on the hydrogel size, porosity and morphology, as well as metal
nanoparticles. The most active catalysts are PEI-M composite microgels due to their higher contact surface even if all PEI-
M composites hydrogel contain the same mmole amounts of metal nanoparticles. The effect of temperature on the hydrol-
ysis of NaBH4 catalyzed by PEI-M hydrogel composite catalyst systems was investigated, and the activation parameters such
as Ea, DH, and DS were also determined. The H2 production graphs at different temperatures (30, 50, and 70 °C) are illus-
trated in Supp. Fig. 3(a)–(c) for PEI bulk hydrogel, microgel, and cryogel, respectively. It can be clearly seen that the H2 pro-
duction speeds from NaBH4 hydrolysis catalyzed by PEI-M (M: Co, Ni) composite hydrogel increased with the increasing
temperature of the materials. With the increase in the temperature from 30 to 70 °C, the time to produce the same amount
of H2 (250 mL) with PEI-Co composite hydrogels decreased from 105 to 5.25, from 50 to 9, and from 85 to 2.8 min for bulkgel,
microgel and cryogel composite systems, respectively. A similar trend was also observed for PEI-Ni hydrogel composite
where the same amount of H2 production time decreased from 180 to 5.5, from 85 to 23, and from 130 to 5 min for PEI-
Ni bulkgel, microgel and cryogel composite systems, respectively, upon increasing the hydrolysis reaction from 30 to
70 °C. Furthermore, the activation parameters for these catalytic reactions were calculated and are presented in Table 2.
The lower Ea, and DH values were obtained for both Co and Ni metal nanoparticle-containing PEI-M composite microgel sys-
N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169 163

200

Volume of H2 (ml)
150

100
NaBH₄
50 NaBH₄+PEI bulkgel
NaBH₄+PEI microgel
NaBH₄ +PEI cryogel (a)
0
0 50 100 150 200 250 300
Time (min)

300
Co
Volume of H2 (ml)

225

150
PEI microgel
75 PEI cryogel
PEI bulkgel (b)
0
0 50 100 150
Time (min)

300
Ni
Volume of H2 (ml)

225

150
PEI microgel

75 PEI cryogel
PEI bulkgel (c)
0
0 50 100 150 200
Time (min)

Fig. 3. Hydrogen production from hydrolysis of NaBH4 using (a) PEI hydrogels without metal nanoparticles (b) PEI-Co, and (c) PEI-Ni bulkgel, microgel and
cryogel composite systems [PEI hydrogels 0.183 mmole metal, and reaction conditions: 50 mM NaBH4 30 °C, 100 rpm].

Table 2
The values for HGR, TOF and activation parameters for PEI-M bulkgel, microgel, and cryogel composites for the catalytic hydrolysis of NaBH4 for H2 production.

Values PEI bulkgel PEI microgel PEI cryogel


Co Ni Co Ni Co Ni
HGR (min 1) 259 137 424 295 292 179
TOF (min 1) 0.59 0.31 0.98 0.68 0.67 0.41
Ea (kJ/mol) 60.69 66.83 38.23 35.31 69.49 65.63
DH (kJ/mol) 56.76 62.77 34.12 31.24 65.37 61.59
DS (J/mol K) 115.95 97.46 182.71 192.97 86.75 98.04

tems with 38.23 and 34.12, and 35.31 and 31.24 kJ/mol respectively. The calculated DS values for PEI-Co composite hydro-
gels are 115.95, 182.71, and 86.75 for bulkgel, microgel and cryogel composite systems, respectively, whereas these val-
ues are 97.76, 192.97, and 98.04 J/mol K for PEI-Ni bulkgel, microgel and cryogel composite systems, respectively. The
activation energies are in the range for hydrogel-M catalyst systems reported in the literature [10,11,18]
For the industrial application of any catalytic system, many characteristics of the catalyst system should meet significant
criteria such as shelf-life, selectivity, durability, eco-friendliness, and reusability. Therefore, the reusability studies of PEI-Co
composite hydrogels were investigated at 70 °C, and corresponding graphs are illustrated in Fig. 4(a)–(c). For the reusability
studies of PEI-Co composites in the hydrolysis of NaBH4, the prepared PEI-Co composite was used for H2 production once and
164 N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169

Bulkgel Conversion Activity (a)


100

% 50

0
1 2 3 4 5
Number of usage

Microgel Conversion Activity (b)


100

% 50

0
1 2 3 4 5
Number of usage

Cryogel Conversion Activity (c)


100

% 50

0
1 2 3 4 5
Number of usage

Fig. 4. The reusability of PEI-Co (a) bulk hydrogel, (b) microgel, and (c) cryogel composite systems on the H2 production. [0.183 mmole metal, 50 mM
NaBH4, 70 °C, 1000 rpm].

then washed one time with DI water and used again for the next hydrolysis reaction. And this process was repeated 5 times
in a row. The% activity of PEI-Co composite hydrogels for the hydrolysis of NaBH4 decreased from 100% to 61, 83, and 78 for
bulkgel, microgel, and cryogel composite systems at end of 5th use, respectively, due to production of NaBO2 in reaction
causing deactivation of metal nanoparticles. On the other hand, 100% conversion at every use was obtained. The% activity
is calculated by taking the initial HGR to the HGR of each use of the catalyst composite system in the H2 generation reaction
for NaBH4 hydrolysis. The% conversion is calculated based on H2 content of NaBH4 plus the same amount of H2 that is gained
from water for this catalytic reaction. At elevated temperatures e.g., at 70 °C, PEI-Co microgel still maintains over 80% activity
which is an important surplus for this type of catalyst, as they can be readily regenerated or enhanced by further treatments
of PEI-based materials e.g., chemical modification to generate ionic liquid functional groups on PEI networks [4,38].

3.4. Reduction of nitrophenolic compounds to their corresponding aminophenolic compounds by NaBH4+PEI-M (Co, Ni, Cu)
hydrogel composite catalyst

Phenolic compounds are considered serious hazardous pollutants due to their toxic nature and potential harmful effect on
living organisms and human health [22,23]. Therefore, the reduction of 4-NP and 2-NP to their corresponding amino phenol
forms 4-AP and 2-AP was studied as a model reaction with PEI-M (M: Co, Ni, and Cu) composite hydrogels in the presence of
NaBH4 and schematic presentation is given in Fig. 5(a). PEI hydrogels without any metal nanoparticles were tested as a cat-
alyst for control experiments and the results were shown in Fig. 5(b). Bare NaBH4 bare PEI hydrogels without any metal
N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169 165

(a)
HO HO
PEI-M
M: Co, Ni, Cu
NO2 NaBH4 (Aq) NH2

1.2 1.2
(b) Co (c)
0.9 0.9 Microgel
NaBH₄ Cryogel
A
0.6
NaBH₄+PEI bulkgel A 0.6 Bulkgel
0.3 NaBH₄+PEI microgel 0.3
NaBH₄+PEI cryogel
0 0
0 50 100 150 0 25 50 75 100
Time (min) Time (min)

1.2 1.2
Ni (d) Cu (e)
0.9 0.9
Microgel Microgel
A 0.6 A 0.6
Cryogel Cryogel
0.3 Bulkgel 0.3 Bulkgel

0 0
0 5 10 15 20 0 25 50 75 100
Time (min) Time (min)
0.7 0.7
(f) Co (g)
0.525 0.525 Microgel
Cryogel
A 0.35 NaBH₄ A 0.35 Bulkgel
NaBH₄+PEI bulkgel
0.175 NaBH₄+PEI microgel 0.175
NaBH₄+PEI cryogel
0 0
0 50 100 150 0 5 10 15 20
Time (min) Time (min)

0.7
Ni (h)
0.525 Microgel
Cryogel
A 0.35 Bulkgel

0.175

0
0 2.5 5 7.5 10
Time (min)
Fig. 5. (a) The schematic presentation of 4-, 2- nitrophenols (NPs) to 4-, 2- Amino phenols (APs), and (b) the reduction reaction of 4-NP to 4-AP by using
bare NaBH4, and bare NaBH4 + PEI hydrogels without metal nanoparticles, and (c) PEI-Co, (d) PEI-Ni, and (e) PEI-Cu in 4-NP reduction reactions, and (f) the
reduction reaction of 2-NP to 2-AP by using bare NaBH4 and bare NaBH4 + PEI hydrogels without metal nanoparticles, (g) PEI-Co, and (h) PEI-Ni bulkgel,
microgel and cryogel composite systems in 2-NP reduction reactions [0.183 mmole metal in catalyst with 50 mL 50 mM NaBH4, at 30 °C, 1000 rpm mixing
rate].

nanoparticles did not showed any catalytic effect for 4-NP reduction. Besides both the effects of metal species and morpholo-
gies of hydrogel templates on reduction reaction rates were compared and corresponding graphs are given in Fig. 5(c)–(e).
Amongst the PEI-M hydrogels materials, cryogel templates have the fastest reduction reactions for all metal types e.g., Co, Ni,
and Cu. The reduction of 4-NP to 4-AP was completed by PEI-Co composite hydrogels in about 90, 41 and 13 min; by PEI-Ni
composite hydrogels in about 10, 6, and 3 min, and by PEI-Cu composite hydrogels in about 45, 15, and 11 min for PEI bulk-
166 N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169

gels, microgels and cryogels, respectively. Additionally, the% yields of the reduction reaction of 4-NP to 4-AP by using PEI-M
composite hydrogel systems were obtained as 97% at the lowest, and about 99% with details given in Table 3. Moreover, the
highest TOF values for all PEI-M (M: Co, Ni, and Cu) composite hydrogels, regardless of their morphology such as bulk gel,
microgel and cryogel, were obtained for Ni containing composites with 0.26, 0.41, and 0.88 mol 4-NP/(molcat)(min) values,
respectively, also given in Table 3. The reduction of 2-NP to 2-AP catalyzed by PEI hydrogels without metal nanoparticles,
PEI-Co and PEI-Ni composite hydrogel (bulk hydrogel, microgel, and cryogel) systems was also investigated and correspond-
ing graphs are illustrated in Fig. 5(f)–(h). The effect of self-hydrolysis of NaBH4 and in the presence of PEI hydrogels without
metal nanoparticles was shown in Fig. 5(f). It is obvious that there is no 2-NP reduction without metal nanoparticles within
hydrogel matrices. As illustrated in Fig. 5(g), the PEI-Co composite hydrogel systems completed the reduction reaction of
2-NP to 2-AP within 17.5, 10 and 7 min for bulkgel, microgels and cryogel composite systems, respectively. On the other
hand, the PEI-Ni bulkgel, microgel and cryogel composite systems completed the reduction reaction of 2-NP to 2-AP under
the same reaction conditions within 7, 5 and 2 min, respectively, illustrated in Fig. 5(h). The PEI-Ni composite hydrogel sys-
tems are faster than PEI-Co composite hydrogel systems for the reduction reaction of 2-NP to 2-AP. The yields for the lowest
reduction reaction of 2-NP to 2-AP catalyzed with PEI-M composite hydrogel systems were observed to have a minimum of
about 97% and details are given in Table 3. As can be seen, for 2-NP reduction to 2-AP the rest of the catalyst system were all
above 98%. The TOF values of PEI-Ni composite cryogel system for reduction of 2-NP to 2-AP are higher than the PEI-Co com-
posite hydrogel systems and are even better than the literature as it was found as 1.32 mol 2-NP/(mol of cat)(min) [43].
Moreover, the activation parameters for the reduction reaction of 4-NP to 4-AP via PEI-M composite hydrogel systems were
calculated and are given in Table 4. The lower Ea value for PEI-Co composite hydrogels was obtained for microgel composite
systems with 30 ± 1 kJ/mol; on the other hand, the lower Ea values for PEI-Ni and PEI-Cu composite hydrogels was obtained
for cryogel composite systems with 26 ± 2 and 15 ± 2 kJ/mol, respectively. It is clear that the morphology and sizes of PEI-M
hydrogel systems have great effect on Ea for the 4-NP reduction reaction under the same conditions, as well as the type of the
metal catalyst. The comparison of Ea values of PEI-Cu composite cryogels for the reduction of 4-NP reaction is lower than
reported in the literature [4,43]. For 4-NP reduction by PEI-M hydrogel catalyst systems it can be said that almost all PEI-
M (M: Co, Ni, Cu) cryogel composite systems provide lower Ea compared to other hydrogels.
The activation parameters such as Ea, DH, and DS for the reduction reaction of 2-NP to 2 AP were also determined and
corresponding values are given in Table 4. The Ea values of 31 ± 1, 44 ± 3, 41 ± 2 kJ/mol, DH values of 28 ± 2, 40 ± 2, 37 ± 3 kJ/-
mol, and DS values with 171 ± 9, 141 ± 7, and 142 ± 6 J/mol K were calculated for PEI-Co bulk hydrogel, microgel and
cryogel composite systems respectively. The Ea values of 41 ± 3, 46 ± 4, and 30 ± 3, DH values with 38 ± 2, 42 ± 2, and
26 ± 3 kJ/mol, and DS values with 140 ± 8, 125 ± 6, and 167 ± 6 J/mol K were calculated for PEI-Ni bulk hydrogel, micro-
gel and cryogel composite systems, respectively, for reduction reaction of 2-NP to 2-AP. PEI-Cu hydrogel composite system
was found not to be effective for the 2-NP reduction reaction. It is obvious that PEI-Ni cryogel composite system is the better
choice of catalyst for 2-NP reduction.

3.5. Decolorization of Methylene blue and Eosin Y by PEI-Cu hydrogel composite catalysts

Dyes, such as MB and EY, widely used in drug and ink industries cause serious environmental problems due to their tox-
icity and dark coloring; they also have antioxidant and antibacterial properties [27–30]. Therefore, their elimination, espe-
cially from different aquatic environments, has paramount significance and the use of metal nanocatalysts such as Cu [28,35]
for this purpose the use of metal catalyts is considered among the most practical methods. An efficient catalyst having inter-
mediate redox potential values can be used for this purpose due to its usability in electron transfer as well as electron trans-
mitting system. In this catalyst system, metal nanoparticles are responsible for the catalyzed decolorization of MB and EY in

Table 3
The effect of the morphology and the sizes of hydrogels on 4-NP and 2-NP reduction catalyzed by PEI-M bulkgel, microgel and cryogel (M = Ni, Cu, Co).
1
Compound Hydrogels type PEI-M composite Finish time (min) Yield (%) TOF (min )
4-NP Bulkgel Co 90 99 ± 0.5 0.03
Ni 10 99 ± 0.3 0.26
Cu 45 98 ± 0.9 0.06
Microgel Co 41 99 ± 0.4 0.08
Ni 6 98 ± 0.8 0.41
Cu 15 97 ± 1.1 0.16
Cryogel Co 13 99 ± 0.4 0.20
Ni 3 99 ± 0.7 0.88
Cu 11 98 ± 0.3 0.24
2-NP Bulkgel Co 17,5 99 ± 0.4 0.10
Ni 7 99 ± 0.3 0.38
Microgel Co 10 98 ± 0.4 0.26
Ni 5 98 ± 0.7 0.53
Cryogel Co 7 99 ± 0.6 0.38
Ni 2 99 ± 0.4 1.32
N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169 167

Table 4
The activation parameters of PEI-M bulkgel, microgel, and cryogel composites for reduction of 4-NP and 2-NP.

Compound Metal type Activation parameters Bulkgel Microgel Cryogel


4-NP Co Ea (kJ/mol) 39 ± 2 30 ± 1 33 ± 2
DH (kJ/mol) 36 ± 3 26± 29 ± 1
DS (J/mol K) 161 ± 9 192 ± 10 168 ± 7
Ni Ea (kJ/mol) 28 ± 3 39 ± 2 26 ± 2
DH (kJ/mol) 24 ± 1 35 ± 3 23 ± 2
DS (J/mol K) 184 ± 7 144 ± 5 180 ± 7
Cu Ea (kJ/mol) 27 ± 3 54 ± 2 15 ± 2
DH (kJ/mol) 24 ± 2 50 ± 2 12 ± 1
DS (J/mol K) 197 ± 8 99 ± 6 224 ± 11
2-NP Co Ea (kJ/mol) 31 ± 1 44 ± 3 41 ± 2
DH (kJ/mol) 28 ± 2 40 ± 2 37 ± 3
DS (J/mol K) 171 ± 9 141 ± 7 142 ± 6
Ni Ea (kJ/mol) 41 ± 3 46 ± 4 30 ± 3
DH (kJ/mol) 38 ± 2 42 ± 2 26 ± 3
DS (J/mol K) 140 ± 8 125 ± 6 167 ± 6

the presence of NaBH4, that reduces the double bonds to single bond resulted in destruction of dye structures. In this way
solution of dyes becomes colorless and their characteristic peaks at 664 and 514 nm disappear [44]. The decolorization of MB
and EY catalyzed by PEI-Cu hydrogel systems was carried out in different aquatic environments, seawater, creek water, tap
water and DI water, and monitored by UV–Vis spectroscopy. The effect of self-hydrolysis and PEI hydrogels without metal
nanoparticles were also tested as catalyst on the decolorization of MB and it was shown that in the presence on self-
hydrolysis of NaBH4 MB decolorization stopped at 75 min with 56.7% decolorization whereas in the presence of PEI hydro-
gels 61.3%, 60.2%, and 59.4% decolorization was observed for bulkgel, microgel and cryogel respectively. As can be seen from
Table 5, the decolorization reactions of MB were completed in seawater at about 6, 4, and 2 min with 97 ± 2%, 98 ± 2%, and
96 ± 2% yield 3 in the presence of 0.047 ± 0.01 mmol Cu nanoparticle PEI-Cu bulkgel, microgel and cryogel composite sys-
tems, respectively at 30 °C with 500 rpm mixing rate. The decolorization in creek water (a local creek-Saricay water) was
completed in about 7, 5, and 2 min with 95 ± 4%, 98 ± 1%, and 98 ± 2% yields. On the other hand, the MB decolorization reac-
tions catalyzed by PEI-Cu composite in DI water was completed in about 10, 8, and 4.5 min with 98 ± 1%, 98 ± 1%, and 97 ± 3%
yields, and in tap water the reactions were completed in about 8, 6, and 3 min with 96 ± 3%, 98 ± 1%, and 98 ± 2% yields for
PEI-Cu composite bulkgel, microgel, and cryogel systems, respectively. Therefore, overall these results demonstrate that PEI-
Cu composite hydrogel systems have great potential for real use in the elimination of some organic dyes such as MB in all
types of contaminated water, including seawater and creeks.
Furthermore, the decolorization of EY in seawater, creek water, DI, and tap water catalyzed by PEI-Cu composite hydrogel
systems was also investigated and corresponding results for completion time and %yield are also given in Table 5. It is obvi-
ous that the EY decolorization was finished at about 21, 16, and 9 min in seawater with 96 ± 2%, 96 ± 3%, and 97 ± 3% yields
for PEI-Cu bulkgel, microgel and cryogel composite systems, respectively. The same reaction was finished in creek water in
about 23, 16, and 11 min with 80 ± 7%, 73 ± 5%, and 82 ± 4% yields, for EY-Cu bulkgel, microgel and cryogel composite sys-
tems, respectively. On the other hand, the reaction catalyzed by PEI-Cu bulkgel, microgel and cryogel composite catalyst sys-
tems yielded 88 ± 5%, 80 ± 8%, and 80 ± 6% with the completion times of 35, 20, and 12 min, respectively in tap water,
whereas 81 ± 1%, 83 ± 7%, and 85 ± 2% yields in 51, 45, and 37 min, respectively, were attained in DI water. It is important

Table 5
The decolorization of MB and EY catalyzed by PEI-Cu bulkgel, microgel and cryogel composite systems in different aqueous media.

Hydrogel morphology Amount of Cu(II): 0.047 ± 0.01 mmol


Water type Methylene blue Eosin Y
Time (min) % Decolorization Time (min) % Decolorization
Bulkgel Sea water 6 98 ± 2 21 96 ± 2
Saricay water 7 95 ± 4 23 80 ± 7
Tap water 8 96 ± 3 35 80 ± 6
DI water 10 98 ± 1 51 81 ± 1
Microgel Sea water 4 98 ± 2 16 96 ± 3
Saricay water 5 98 ± 1 16 73 ± 5
Tap water 6 98 ± 1 20 80 ± 8
DI water 8 98 ± 1 45 83 ± 7
Cryogel Sea water 2 96 ± 3 9 97 ± 3
Saricay water 2 98 ± 2 11 82 ± 4
Tap water 3 96 ± 3 12 88 ± 5
DI water 4.5 97 ± 3 37 85 ± 2
168 N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169

to state that the decolorization of both dyes were about 75% without any PEI-Cu hydrogel composite catalyst and took
120 min for MB, and 60 min for EY, confirming the superior catalytic effectiveness of the PEI-Cu hydrogel composite catalyst
systems. The fast decolorization of both dyes in seawater with PEI-Cu cryogel composite systems may be attributed to the
existence of salts and the higher ionic strength of seawater (high saltiness) in comparison to creek water. It can be seen that
the catalytic degradation of MB was not affected much in any media, e.g., faster amounts of degradation in seawater and
creek water (about 2 min) by PEI-Cu cryogel composite system, and the degradation rate of EY was in the order of seawa-
ter > tap water > DI water > creek water. Considering the differences in chemical structure of the dyes, e.g., MB is a cationic
dye whereas EY is an anionic dye, PEI-M (Co, Ni, Cu) hydrogel (bulkgel, microgel, and cryogel) composite systems may be
very effective for degradation of other organic compounds and dyes as only Cu nanoparticle-containing hydrogel composite
system was tested for MB and EY. In addition to Co, Ni and Cu metal nanoparticles, other metal nanoparticles such as Fe, Ru,
Pt, Pd, Au and others can be readily prepared and used for different catalytic reactions. It is also pertinent that the interaction
of dyes and polymeric template may be important. For example, in degradation of MB there is no electrostatic interaction
with the positively-charged PEI matrices; however, EY could be absorbed into PEI matrices causing a reduction in catalytic
degradation of EY by Cu nanoparticles embedded within PEI matrices as observed here with lower decolorization rates. This
controversial situation could be very useful for real environmental applications where the actual dye concentration is lower
and many other contaminants may exist concurrently, so their confiscation by the oppositely charged hydrogel template
facilitates the elimination of toxic organic compounds. Therefore, PEI hydrogels with different morphologies and sizes retain
great potential for in situ metal nanoparticle preparation and catalytic reactions in different aquatic environments.

4. Conclusion

Herein, PEI-based hydrogels with different size, morphology and porosity such as bulkgel, microgel, and cryogel were
synthesized by using an epoxide crosslinker (GDE) utilizing different polymerization techniques. The synthesized PEI hydro-
gel matrices whether bulkgel, microgel or cryogel were then used as metal ion binding networks for Co2+, Ni2+, and Cu2+ and
then the metal ions entrapped within hydrogel matrices reduced to the corresponding nanoparticles via reduction reaction
by NaBH4 treatment. It was also demonstrated here that the prepared PEI-M (M: Co, Ni, Cu) hydrogel composite systems can
be used as catalysts for different catalytic reactions: e.g., H2 generation from hydrolysis of NaBH4, reduction of 4-, 2- nitro-
phenols, and decolorization of MB and EY dyes. It was found that Co nanoparticle-containing PEI bulkgel, microgel, and cryo-
gel composite systems are more effective than Ni-containing PEI bulkgel, microgel, and cryogel composite systems for H2
generation from hydrolysis of NaBH4. Moreover, the PEI-Co microgel composite systems showed better catalytic perfor-
mances than PEI-Co bulkgel and cryogel composite systems for the hydrolysis of NaBH4 with 424 ml H2/(g of cat)(min)
HGR, 0.98 mol H2/(mol of cat)(min) TOF, and 38.2 kJ/mol Ea values. In the reusability of PEI-Co hydrogel composite systems
at 70 °C, it was found that PEI-Co microgel composite achieved better catalytic performances with 17% decrease in activity
compared to PEI-Co cryogel and bulkgel where they showed 22, and 39% catalytic activity reductions, respectively at end of
5th use. On the other hand, the PEI-Ni bulkgel, microgel, and cryogel catalyst composite systems were found to be more
effective than PEI-Co and PEI-Cu hydrogel composite systems for the reduction of 4- and 2- NP to their corresponding amino
forms. Furthermore, PEI-Ni cryogel composite systems showed better performance than PEI-Ni bulkgel and microgel com-
posite systems in the reduction of 4- and 2- NP with 0.88 4-NP/(mol of cat)(min) and 1.32 mol 2-NP/(mol of cat)(min)
TOF values, respectively. Moreover, PEI-Cu hydrogel systems was found to be very effective for dye decolorization in differ-
ent aquatic environments such as seawater, creek water, DI and tap water. The PEI-Cu cryogel composite system was shown
to have better catalytic performances than PEI-Cu bulkgel and microgel composite systems for the decolorization of MB and
EY in seawater completing the decolorization within 2 min and 96% yield, and 9 min and 97% decolorization for MB and EY
respectively. The PEI-M bulk hydrogel, microgel and cryogel composite systems reported here provide immense potential as
templates for different metal nanoparticles such as Fe, Ru, Pt, Pd, Au and others in addition to Co, Ni and Cu, providing an
environment for their stabilization and offering prolonged self-life by protecting them from oxidation, aggregation and deac-
tivation within their polymeric network. PEI-M hydrogel composite systems possesses great potential as soft reactor cham-
ber for various catalytic reactions e.g., H2 production, toxic metal such as Cd(II), Pb(II), Pb(II), Co(II), Ni(II), Hg(II) etc. removal,
hazardous aromatic nitro compound such as 4-NP, 2- NP reduction, and harmful dye elimination i.e., MB, EY and so on.

Acknowledgement

Financial support from Canakkale Onsekiz Mart University (FYL-2013-146) is greatly appreciated.

Appendix A. Supplementary material

Supplementary data associated with this article can be found, in the online version, at http://dx.doi.org/10.1016/j.eur-
polymj.2016.01.046.
N. Sahiner, S. Demirci / European Polymer Journal 76 (2016) 156–169 169

References

[1] Z.Q. Tan, S. Ohara, M. Naito, H. Abe, Adv. Mater. 23 (2011) 4053–4057.
[2] N. Orakdogen, P. Karacan, O. Okay, React. Funct. Polym. 71 (2011) 782–790.
[3] J.M. Zhu, Biomaterials 3 (2010) 4639–4656.
[4] S. Yildiz, M. Sahiner, N. Sahiner, Eur. Polym. J. 70 (2015) 66–78.
[5] R. Dawson, A.I. Cooper, D.J. Adams, Prog. Polym. Sci. 37 (2012) 530–563.
[6] O. Okay, Prog. Polym. Sci. 25 (2000) 711–779.
[7] V.I. Lozinsky, I.Y. Galaev, F.M. Plieva, I.N. Savina, H. Jungvid, B. Mattiasson, Trends Biotechnol. 21 (2003) 445–451.
[8] J.S. Jenkins, M.C. Flickinger, O.D. Velev, J. Colloid Interf. Sci. 380 (2012) 192–200.
[9] N. Sahiner, Colloid Surf. A. 433 (2013) 212–218.
[10] N. Sahiner, Prog. Polym. Sci. 38 (2013) 1329–1356.
[11] N. Sahiner, A.O. Yasar, Fuel Proces. Technol. 111 (2013) 14–21.
[12] N. Sahiner, H. Ozay, O. Ozay, N. Aktas, App. Catal. B-Environ. 101 (2010) 137–143.
[13] K. Esumi, R. Isono, T. Yoshimura, Langmuir 20 (2004) 237–243.
[14] L. Schlapbach, A. Zuttel, Nature 414 (2001) 353–358.
[15] Y. Li, Y. Dai, X. Tian, Int. J. Hydrogen Energy 40 (2015) 9235–9243.
[16] Y. Wang, K. Qi, S. Wu, Z. Cao, K. Zhang, Y. Lu, H. Liu, J. Power Sources 284 (2015) 130–137.
[17] Rafi-ud-din, L. Zhang, L. Ping, Q. Xuanhui, J. Alloys Compd. 508 (2010) 119–128.
[18] S. Demirci, N. Sahiner, Fuel Proces. Technol. 127 (2014) 88–96.
[19] M. Paladini, G.M. Arzac, V. Godinho, M.C. Jimenez de Haro, A. Fernandez, Appl. Catal. B-Environ. 158–159 (2014) 400–409.
[20] X. Li, G. Fan, C. Zeng, Int. J. Hydrogen Energy 39 (2014) 14927–14934.
[21] S.C. Amendola, P. Onnerud, M.T. Kelly, P.J. Petillo, S.L. Sharp-Goldman, M. Binder, J. Power Sources 85 (2000) 186–189.
[22] Y.Y. Chu, Y. Qian, W.J. Wang, X.L. Deng, J. Hazard. Mater. 199–200 (2012) 179–185.
[23] USEPA, Federal Register, Washington, D.C. 52, 131 (1987) 25861–25962.
[24] T. Komatsu, T. Hirose, Appl. Catal. A-Gener. 276 (2004) 95–102.
[25] M.J. Vaidya, S.M. Kulkami, R.V. Chaudhari, Org. Pocess Res. Dev. 7 (2003) 202–208.
[26] J. Pierce, J. Soc. Dyers Color. 110 (1994) 131–133.
[27] T. Hannappel, B. Burfeindt, W. Storck, F. Willig, J. Phys. Chem. B 101 (1997) 6799–6802.
[28] L. Xia, H. Zhao, G. Liu, X. Hu, Y. Liu, J. Li, D. Yang, X. Wang, Colloid Surf. A 384 (2011) 358–362.
[29] A.C. Fisher, L.M. Peter, E.A. Ponomarev, A.B. Walker, K.G.U. Wijayantha, J. Phys. Chem. B 104 (2000) 949–958.
[30] M. Muruganandham, M. Swaminathan, Dyes Pigm. 63 (2004) 315–321.
[31] M.S. Lucas, J.A. Peres, Dyes Pigm. 71 (2006) 236–244.
[32] M. Kositzi, A. Antoniadis, I. Poulios, I. Kiridis, S. Malato, Sol. Energy 77 (2004) 591–600.
[33] A. Santos, P. Yustos, T. Cordero, S. Gomis, S. Rodriguez, F. Garcia-Ochoa, Catal. Today 102–103 (2005) 213–218.
[34] Y.D. Xie, F. Chen, J.J. He, J.C. Zhao, H. Wang, J. Photochem. Photobiol. A 136 (2000) 235–240.
[35] M. Basu, A.K. Sinha, M. Pradhan, S. Sarkar, A. Pal, C. Mondal, T. Pal, J. Phys. Chem. C 116 (2012) 25741–25747.
[36] A. Syoufian, K. Nakashima, J. Colloid Interf. Sci. 313 (2007) 213–218.
[37] S. Miao, Z. Liu, B. Han, H. Yang, Z. Miao, Z. Sun, J. Colloid Interf. Sci. 301 (2006) 116–122.
[38] S. Demirci, N. Sahiner, J. Mol. Liquids 194 (2014) 85–92.
[39] H.Q. Pham, M.J. Marks, in: Ulmann’s Encyclopedia of Industrial Chemistry, WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim, 2006.
[40] K. Horie, H. Hiura, M. Sawada, I. Mita, H. Kambe, J. Polym. Sci. 8 (1970) 1357–1372.
[41] Y. Zhang, S. Vyazovkin, J. Phys. Chem. B 111 (2007) 7098–7104.
[42] N. Sahiner, Colloid. Surf. A 452 (2014) 173–180.
[43] S. Demirci, N. Sahiner, Water Air Soil Poll. 226 (2015) 64–76.
[44] S. Srivastava, S.K. Sharma, R.K. Sharma, Colloid Surf. A 373 (2011) 61–65.

You might also like