You are on page 1of 7

Applied Catalysis B: Environmental 101 (2010) 137143

Contents lists available at ScienceDirect

Applied Catalysis B: Environmental


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

A soft hydrogel reactor for cobalt nanoparticle preparation and use in the
reduction of nitrophenols
Nurettin Sahiner a,b, , Hava Ozay a , Ozgur Ozay a , Nahit Aktas c
a
b
c

Canakkale Onsekiz Mart University, Faculty of Sciences and Arts, Chemistry Department, Terzioglu Campus, 17020 Canakkale, Turkey
Nanoscience and Technology Research and Application Center (NANORAC), Terzioglu Campus, 17020 Canakkale, Turkey
Yuzuncu Yil University, Faculty of Engineering, Chemical Engineering Department, 65080 Van, Turkey

a r t i c l e

i n f o

Article history:
Received 14 July 2010
Received in revised form 1 September 2010
Accepted 11 September 2010
Available online 20 October 2010
Keywords:
Hydrogel-templates
Flexible hydrogel-reactor
In situ nanoparticle formation
Hydrogel-reaction vessel
Hydrogel-nanocomposites
Co nanoparticles

a b s t r a c t
Bulk poly(2-acrylamido-2-methyl-1-propansulfonic acid) (p(AMPS)) hydrogels were prepared by irradiation of an aqueous solution of AMPS in the presence of crosslinker and photoinitiator. These p(AMPS)
hydrogel networks were utilized for in situ cobalt nanoparticle synthesis by reduction of metal ions
absorbed into the hydrogel network with a reducing agent, i.e., NaBH4 . TEM images conrmed that Co
particles are about 100 nm in size. The hydrogel network with embedded Co nanoparticles was utilized
as a catalyst in the reduction of 4-nitrophenol (4-NP) and 2-nitrophenol (2-NP) in aqueous media in the
presence of an excess amount of NaBH4 . The kinetics of the reduction reaction under different reaction
conditions was investigated to determine the activation parameters. Activation energies are 27.8 kJ mol1
and 39.3 kJ mol1 for 4-NP and 2-NP, respectively. It was found that hydrogelCo composites were 99%
active after 5 days storage.
2010 Elsevier B.V. All rights reserved.

1. Introduction
The increased use of metal nanoparticles in many advanceddesign devices in the elds of optics, microelectronics, sensors,
information storage, catalysts and nanoelectronics urge the
researcher to seek out new approaches to their synthesis [15].
There is a high demand for metal nanoparticles or nanoclusters that
are versatile in size, shape and morphology and that surmount the
main issues in the use of bare metal nanoparticles as a catalyst.
To overcome these major problems such as aggregation, oxidation and inactivation of metal nanoparticles, numerous preventive
measures have been employed. The utilization of supports such
as polymers, silicates, resins, alumina and zeolites is amongst the
some of the measures taken in the use of metal nanoparticles or
nanoclusters as catalysts [1,410].
Hydrogels as three-dimensional hydrophilic polymeric networks can provide unique environments for the preparation
and protection of metal nanoparticles. That hydrogels are threedimensional, water-swollen, crosslinked, hydrophilic polymer
chains makes them especially suitable. The polymer chains in the

Corresponding author at: Canakkale Onsekiz Mart University, Faculty of Sciences


and Arts, Chemistry Department, Terzioglu Campus, 17020 Canakkale, Turkey.
Tel.: +90 2862180010x2041; fax: +90 2862181948.
E-mail address: sahiner71@gmail.com (N. Sahiner).
0926-3373/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcatb.2010.09.022

hydrogel networks have functional groups such as SO3 H, COOH,


CONH2 , OH, and NH2 that render their hydrophilic character.
Some of these chemical functional groups as SO3 and COO
in the hydrogel network can be formed as negative charges in
aqueous environments and so absorb metal ions via electrostatic
interactions. Taking advantage of these characteristics of hydrogel networks, various metal nanoparticles or nanoclusters can be
prepared in situ by reduction of absorbed metal ions with suitable reducing agents such as NaBH4 [6]. Additionally, because of
these properties, hydrogels can even be used for the removal of
toxic metal ions from aquatic media. Hydrogels are soft and versatile materials, and are biocompatible, used in many applications
in biomedical elds as an active agent (drugs, protein, and gene)
delivery vehicle, as a template for tissue engineering, wound dressing materials and articial organs, and as actuator responding to
various stimuli [1116].
Nitro-aromatic compounds hold a signicant role in the industrial chemistry and are generally produced as a by-product from
pharmaceuticals, agrochemicals, urethanes, dyes and so on [17,18],
and are also toxic organic species. Therefore, to reduce nitroaromatic compounds to their corresponding useful amines, various
catalysts have been developed. Amongst these, metallic catalysts are silver nanoparticles [1,4,5], PtNi nanoparticles [2], gold
nanoparticles [68], hydroiodic acids [18], various nickel and iron
nanoparticles [1921], and TiO2 [22]. Wang et al. have prepared Ni
nanoparticle with different sizes and crystal structures using var-

138

N. Sahiner et al. / Applied Catalysis B: Environmental 101 (2010) 137143

ious modiers, stabilizing and reducing agents and utilized them


in the reduction of p-nitrophenol to p-aminophenol with better or
comparable results with the literature [2325].
4-Aminophenol (or p-aminophenol), traditionally obtained by
the reduction of 4-nitrophenol, is an important reactive intermediate for pharmaceutical, photographic and plastic industries. In
the reduction process generally an ironacid pair is utilized. The
biggest disadvantage of this reduction process is the generation
of large amounts of iron-oxide sludge (1.2 kg FeO sludge/kg product) causing great environmental concerns [20,21]. The utilization
of environmentally benign hydrogels in such a catalytic reduction
reaction is a reasonable goal.
Although hydrogels are useful resources, they have not been
used in simultaneous dual action settings i.e., both as a template
for metal nanoparticle synthesis and as a reactor in the reduction of various organic toxic reagents. Therefore, this investigation
will demonstrate the utilization of hydrogels in dual action settings for the rst time. One is in the synthesis of Co nanoparticles
in situ, and the other one is the utilization of these Co-containing
hydrogel composites as a reactor. For this objective, the in situ
prepared Co nanoparticles generated in a exible hydrogel matrix
were used in the reduction of 4- and 2-nitrophenols in aqueous
sodium borohydride solutions. The reaction kinetics were investigated under different reaction conditions and various parameters
such as catalyst and NaBH4 amounts, and temperature, affecting the
reduction processes of the nitrocompounds were evaluated. The
activity, repetitive usage and the shelf-life of the catalysthydrogel
system were also investigated.
2. Materials and methods

cleaning procedure, hydrogels were dried in an oven at 45 C to


a constant weight and preserved in an air tight container or further
use.
2.2. Cobalt nanoparticle preparation in p(AMPS) hydrogel
network
For in situ preparation of Co nanoparticles inside hydrogels, the
following method was employed; 50 mg of dried hydrogel was
placed into 100 mL 0.1 M aqueous solution of Co(II) prepared from
CoCl2 6H2 O to load the hydrogels with cobalt ions. The absorption
studies were performed in a water-shaker bath at room temperature for 2 days. The absorption amount was determined with
ICP-AES measurement by removing 1 mL of solution after diluting to a certain concentration before measurements. The removed
amounts in 1 mL of Co(II) ions were included in the calculation
for the next measurements. After reaching a plateau region in the
absorption studies, cobalt ion-loaded hydrogels were washed with
DI for 24 h to remove physically adsorbed cobalt ions. After cleaning, the cobalt ion-loaded hydrogels were placed into 100 mL 0.5 M
aqueous NaBH4 solution for 12 h in a water-shaker bath at room
temperature to reduce the absorbed Co(II) ions to Co nanoparticles. After reduction of cobalt ions in situ, hydrogel composites were
cleaned by washing with DI water, stored for reduction reactions of
nitro compounds and denoted as hydrogelCo(HCo) composites.
To determine the shelf-life of HCo catalyst systems, the composite materials were kept for certain time periods in nitrogen-purged
100 mL DI water and stored in darkness at ambient temperature in
a closed container.
2.3. Catalytic reduction studies

p(AMPS) hydrogels were prepared using 2-acrylamido2-methyl-1-propansulfonic


acid
(AMPS)
as
monomer,
N,N -methylenebisacrylamide as cross-linker (Bis), N,N,N ,N tetramethylenediamine (TEMED) as an accelerator, and
ammonium persulfate (APS) as redox initiator. Nitro-compounds,
2-NP (99%) and 4-NP (99%), and NaBH4 (98%) were purchased
from Merck and used without further purication. 18.2 Mohm cm
(Millipore Direct-Q3 UV) distilled water was used throughout the
experiments. All products were of analytical grade and purchased
from SigmaAldrich, Acros and Fluka Chemical Companies.
To monitor the conversion of nitro compounds, a UVVis spectrometer was employed (PG Instruments Ltd. T80+ model UVVis
spectrophotometer). The Co nanoparticle content of p(AMPS)
hydrogels was determined via Thermogravimetric Analysis (SII
TG/DTA 6300) and Inductive Coupled Plasma-Atomic Spectrometry
(Varian Liberty II AX Sequential ICP-AES) measurements. TG measurements were performed between 50 and 1200 C with 10 C/min
heating rate and 100 mL/min nitrogen ow rate.

A 50 mL aqueous solution containing 1.44 102 mol NaBH4


and 7.19 104 mol nitro compound (4-NP or 2-NP) was prepared
and 6.22 mg Co particles containing 50 mg (determined by ICPAES measurements) HCo composite catalyst was added to this
solution. As soon as the HCo composite catalyst was added, the
reduction reaction initiated. To monitor the reduction of 4-NP
and 2-NP, 0.5 mL samples were taken from the reaction medium,
diluted to certain concentration and the absorbance values at
400 nm for 4-NP, and 414 nm for 2-NP was recorded using a UVVis
spectrometer. The reduction rate constants of 4-NP and 2-NP were
calculated from the decrease in intensities of the corresponding
peaks using their previously constructed calibration curves. All
the measurements in this investigation were conducted in triplicate and average values with standard deviation were determined.
To see the effect of temperature, the reduction reactions of both
reagents were also conducted at 30, 40, 50, and 60 C. The reuse
experiments were performed sequentially after washing the HCo
catalyst composites three times with water before next use.

2.1. Hydrogel synthesis

3. Results and discussion

p(AMPS) hydrogels were prepared via radical polymerization


in the presence of a redox initiator. Briey, 0.0217 mol AMPS was
mixed with Bis (0.5 mol% of monomer) in 3.5 mL water containing 100 l TEMED. One mL aqueous initiator (APS) solution (5 wt%)
was added (1 mol% with respect to monomer) to the above hydrogel precursors and the solution was placed into plastic pipettes
of 4 mm diameter. The polymerization and crosslinking reaction
was allowed to proceed for 24 h to complete the reaction at ambient temperature. The prepared p(AMPS) hydrogels were sliced
to a cylindrical shape of 45 mm length and washed by keeping
the hydrogel pieces in water for three days. The wash water was
replaced every 12 h to get rid of reaction residues such as monomer,
polymer, crosslinker and the unreacted initiator. Following the

3.1. In situ formation and characterization of cobalt particles


In situ cobalt nanoparticle syntheses were carried out according
to the previously reported literature [9]. A schematic representation of cobalt particle synthesis in p(AMPS) hydrogel network
is depicted in Fig. 1, with corresponding digital camera images.
Fig. 1(a) and (b) illustrates Co(II) ion absorption by p(AMPS) hydrogel, and Fig. 1(c) shows Co(II) ion absorption isotherm in the same
hydrogel. As can be seen from the absorption isotherm, within
10 h all the available sites are assumed to be occupied with metal
ions. The hydrogel was in contact with metal ions for longer but
there was no further increase in the absorption capacity of p(AMPS)
hydrogel observed. Following Co(II) ion loadings and washing with

N. Sahiner et al. / Applied Catalysis B: Environmental 101 (2010) 137143

139

Fig. 1. p(AMPS) hydrogel network and its digital camera images (a) in DI water, and (b) Co(II) ion-absorbed p(AMPS) hydrogel network, and (c) Co(II) absorption isotherm.

DI water to remove unbound or excess Co(II) ions, hydrogelCo(II)


networks were treated with aqueous NaBH4 to generate Co metal
nanoparticles inside the p(AMPS) network. Fig. 2(a) demonstrates
the in situ-formed metal nanoparticle with its corresponding digital camera image, while Fig. 2(b) illustrates the TEM images of the
HCo composite system. As revealed by TEM images, the obtained
Co nanoparticles were approximately 50150 nm in size. It is also
known that one can tune the metal nanoparticles size by changing
the metal ion loading amounts, and/or introducing other functional
groups that do not have any tendency towards metal ions, and
using different amounts of crosslinker during hydrogel synthesis
as elucidated in the literature [9]. The main purpose of this investigation was the preparation of Co metal nanoparticles inside a
hydrogel network and their utilization as reaction media. Therefore, it is important to know the amount of metal nanoparticles
inside the hydrogel networks. The amount of metal ions inside
p(AMPS) hydrogel was determined by ICP-AES using two methods.

In the rst method, the metal ion absorption was determined from
the solution (Fig. 1(c)) and found to be 128.1 mg Co(II) absorbed
per gram dry p(AMPS) hydrogel. The amount of absorbed Co(II)
ion was calculated by subtracting the amount of Co(II) remaining in solution after absorption from the initial concentration of
Co(II) ions determined by ICP-AES measurements. In the second
method, the prepared p(AMPS)Co composite system was treated
with 100 mL 5 M HCl three times to dissolve Co metal particles,
and the amount of the Co(II) ion was determined as 124.3 mg/g
dry p(AMPS) hydrogel. Both methods gave good correlation. Moreover, the amount of Co metal particles was also determined with
themogravimetric measurements by comparing empty p(AMPS)
hydrogel with p(AMPS)Co hydrogel composite systems. Thermograms were taken under nitrogen ow to avoid oxidation on
heating up to 1200 C. As illustrated in Fig. 3, the amount of metal
nanoparticles in p(AMPS) hydrogel is about 22.7% (0.5) by weight.
It is possible that this weight may not be completely pure Co

140

N. Sahiner et al. / Applied Catalysis B: Environmental 101 (2010) 137143

Fig. 2. Representation of Co nanoparticle formation inside p(AMPS) hydrogel network (a) digital camera, and (b) TEM images of p(AMPS)Co composites.

100.0

p(AMPS)

80.0

nanoparticles. There could be some metal oxides which can form


during heating with TGA, and even boron-containing complexes of
Co can be formed during the reduction of Co(II) with NaBH4 .

p(AMPS)-Co

TG %

3.2. Catalytic activity


60.0

40.0

22.7 %

20.0

0.0
200

400

600

800

1000

Temp Cel
Fig. 3. TG thermogram of p(AMPS) and p(AMPS)Co composites.

Due to the important application of the reduced forms of nitroaromatic compounds in pharmacy, agriculture, biomedicine and
many other industries, novel catalyst systems that can operate in
very different environments are of great signicance [17,18]. For
instance, many important analgesic and antipyretic pharmaceuticals such as paracetemol, acetanilide and phenacetin are produced
from 4-aminophenol obtained from the reduction of 4-nitrophenol
[20,21]. In this investigation we prepared Co metal nanoparticles
inside environmentally friendly hydrogels, and demonstrated the
use of these composite systems in the reduction 4-NP and 2-NP
in the presence of an excess amount of aqueous NaBH4 (20-fold
excess of NaBH4 compared to 4-NP and 2-NP). Both reagents have
adsorption maxima at 400 and 414 nm, respectively, in the reaction
mixture and the decreases in these absorption bands were tracked
by UVVis spectrophotometers. Fig. 4(a) illustrates the reduction

N. Sahiner et al. / Applied Catalysis B: Environmental 101 (2010) 137143

141

a 0.016

30 0C
40 0C
50 0C
60 0C

C (mol/l)

0.012
0.008
0.004
0

10

15

20

Time (min)

25

-4.5

35

30 0C
40 0C
50 0C
60 0C

-5.5

ln C

30

-6.5
-7.5
-8.5

Fig. 4. (a) Reduction mechanism of 4-nitrophenol (4-NP) to 4-aminophenols and


(b) UVVis spectra of this reduction reaction of 4-NP by p(AMPS)Co composites in aqueous NaBH4 . (Reaction conditions: 50 mL [4-NP] = 1.44 102 mol/L,
[NaBH4 ] = 2.88 101 mol/L, 6.22 mg catalyst, 750 rpm, 30 C).

reaction mechanism of 4-nitrophenol to 4-aminophenol with the


prepared hydrogel composite catalytic system, whereas, Fig. 4(b)
depicts the reduction in the intensities of the absorption maximum with time for the reaction carried out at 30 C. The peak
at 400 nm is assumed to be from the 4-nitrophenolate anion due
to sodium boron hydrate alkaline solutions. As the reaction progressed, new band formation at 290 nm for 4-aminophenol was
observed as reported [2]. Before the reduction reaction, nitrogen
was purged in aqueous nitro compounds for 15 min to prevent
delaying the reactions as it is reported that the reduction of 4NP starts as all oxygen is consumed from the medium [7]. By
using 20-fold more NaBH4 than 4-NP or 2-NP, it was ensured that
the reaction rates depend only on the concentration of nitrophenol compounds and that the reactions are rst order. From the
experiments tested at 10-, 20-, 40- and 80-fold excess NaBH4 compared to nitro-compounds; it was found that 20-fold excess amount
was enough for rst degree dependence in the reaction rates. The
reduction reaction rate constants at 30 C for 4-NP when used at
20-, 40- and 80-fold NaBH4 were found to be 0.1184, 0.1184, and
0.1187 min1 , respectively. The same relationship was observed in
the reduction of 2-NP. Therefore, throughout this investigation 20fold excess NaBH4 was used for the reduction reactions of 4-NP
and 2-NP. The change in the reaction rate constants with amount
of NaBH4 is illustrated in Fig. 5. As can be clearly seen from the gure

k (min -1)

0.12
0.08
0.04

0.01

0.02

0.03

0.04

0.05

0.06

0.07

Amount of NaBH4 (mol)


Fig. 5. The change of reduction rate constant of 4-NP with amount of NaBH4 . (Reaction conditions: 50 mL [4-NP] = 1.44 102 mol/L, 6.22 mg catalyst, 750 rpm, 30 C.)

10

15

20

25

30

35

Time (min)
Fig. 6. (a) The change in the concentration of 4-NP with time in reduction reaction by p(AMPS)Co catalyst system in the presence of aqueous NaBH4 with
reduction, and (b) ln c vs time of the same graph. (Reaction conditions: 50 mL [4NP] = 1.44 102 mol/L, [NaBH4 ] = 2.88 101 mol/L, 6.22 mg catalyst, 750 rpm.)

when more than 20-fold excess of NaBH4 is used, there is no significant change in the rate constants observed. Fig. 6(a) illustrates the
decrease in the concentration of 4-NP at different temperatures,
and the constructed ln Ct graph is shown in Fig. 6(b). A good linear
correlation, as seen in Fig. 6(b), conrms that the reduction reactions comply with rst order reaction kinetics. Similar results were
also obtained for 2-NP with p(AMPS)Co catalyst systems (data are
not shown). The total turnover frequency (TOF) was calculated for
4-NP based on Fig. 6 between 30 and 60 C and found between 0.213
and 0.568 mol 4-NP/mol Co.min that is comparable to the silver
nanoshell-coated cationic polystyrene beads used as solid phase
catalyst for the reduction of 4-NP [26].
The pseudo-rst order rate constants (k) for the reactions of
4-NP and 2-NP were calculated and are listed in Table 1. The
activation parameters for both reduction reactions with the synthesized p(AMPS)Co catalyst systems were calculated using the
well-known Arrhenius (Eq. (1)) and Eyring equations (Eq. (2)) as
shown, and their individual constructed ln k vs 1/T, and ln(k/T) vs
1/T, graphs are shown in Fig. 7(a) and (b), respectively.
ln k = ln A
ln

0.16

k
= ln
T

E 
a

RT

k 
B

S #
H #

R
R

(1)

1
T

(2)

where Ea is activation energy, T is absolute temperature, kB


is the Boltzmann constant (1.381 1023 J K1 ), h is the Planck
constant (6.626 1034 J s), H# is activation enthalpy, S# is activation entropy and R is the ideal gas constant (8.314 J K1 mol1 ).
The activation energies for 4-NP and 2-NP reduction reactions
were found as Ea = 27.8 kJ mol1 and 39.3 kJ mol1 respectively, as
shown in Table 1 with the other parameters. It should be noted
that the p(AMPS)Co composite system requires lower activation
energy for 4-NP than 2-NP. Again, for 4-NP more favorable activation parameters H# = 25.1 kJ mol1 and S# = 180.2 J mol1 K1
were calculated whereas for 2-NP the activation parameters were
H# = 36.7 kJ mol1 , S# = 147.2 J mol1 K1 .

142

N. Sahiner et al. / Applied Catalysis B: Environmental 101 (2010) 137143

Table 1
The change in the reduction rate constants of 4-NP and 2-NP with different temperatures and activation parameters in catalysis with p(AMPS)Co composites.
Compound

4-Nitrophenol

2-Nitrophenol

Temperature ( C)

k (min1 )

30
40
50
60
30
40
50
60

0.12
0.16
0.23
0.31
0.06
0.11
0.16
0.25

Ea (kJ mol1 )

H# (kJ mol1 )

S# (J mol1 K1 )

27.8

25.1

180.2

39.3

36.7

147.2

Reaction conditions: 50 mL [4-NP] and [2-NP] = 1.44 102 mol/L, 6.22 mg catalyst, 750 rpm.

Table 2
The change in the conversion and activity with repetitive use of p(AMPS)Co composite catalyst system in sequential reactions.

-1

4-NP
2-NP

-1.5

ln k

Use #
-2

% conversion
% activity

-2.5
-3
2.900E-03

3.000E-03

3.100E-03

3.200E-03

3.300E-03

3.400E-03

100
100

98
94

98
91

97
83

97
78

Reaction conditions: 50 mL [4-NP] = 1.44 102 mol/L, [NaBH4 ] = 2.88 101 mol/L,
6.22 mg catalyst, 750 rpm, 30 C.

1/T

ln (k/T)

-6

4-NP
2-NP

-6.5
-7
-7.5
-8
-8.5
-9
2.900E-03

3.000E-03

3.100E-03

3.200E-03

3.300E-03

3.400E-03

1/T
Fig. 7. (a) ln k1/T graphs for 4-NP and 2-NP (Arrhenius Eq. (1)), and (b) ln(k/T)1/T
graphs (Erying Eq. (2)) by p(AMPS)Co composites. (Reaction conditions: 50 mL [4NP] = 1.44 102 mol/L, 6.22 mg catalyst, 750 rpm, 30 C.)

To determine whether the amount of catalyst affected the reduction rates or not, the amount of catalyst varied by reducing the
amount of 4-NP used with 20-fold excess NaBH4 at 30 C. The
amount of Co nanoparticles inside the used p(AMPS) hydrogels was
3.11, 6.22, 9.33 and 12.44 mg, and the plotted graphs for the reduction rate constant of 4-NP vs the amount of catalyst are illustrated in
Fig. 8. The reduction reactions were also performed in the absence
of p(AMPS)Co catalyst system. No reaction was observed as no
change at 400 nm was detected for 4-NP in the UVVis absorption
spectrum. This is in accordance with the literature which states
that without a catalyst, the reduction of 4-NP is thermodynamically

0.24

k (min-1)

0.2
0.16

unfavorable [2]. As can be expected the reduction rate constant


linearly increases with the amount of catalyst at a given temperature (30 C). To control whether the empty hydrogel affects the
reduction rate or not, the same reaction conditions were arranged
(20-fold excess NaBH4 to 4-NP) at 30 C with hydrogels containing
no Co particles, and no reaction was identied.
Additionally, to evaluate whether these p(AMPS)Co composite systems are reusable or not, the composite catalyst was used 5
times successively, rinsing with DI water between every use. As can
be seen from Table 2, the conversion did not change signicantly,
and only a 22% reduction in the activity of the catalyst was observed
at the end of the fth use. The activity is determined by the ratio
of nal reduction rate to the initial reduction rates. It can be concluded that this catalyst-vessel (p(AMPS)Co composite systems)
can be used up to 5 cycles without any signicant loss of conversion and very little loss of activity in the reduction of 4-NP. The
slight decrease in the activity can be attributed to the formation of
some boron compound and/or some oxides that prevent catalytic
performance of active sites.
In real applications, the shelf-life of a catalyst is crucial especially
for costly catalytic systems. Therefore, the prepared catalystvessels shelf-life was investigated by storing freshly prepared
p(AMPS)Co systems in DI water for 1, 5, 10 and 15 days, and
using them (each one containing 6.22 mg Co nanoparticles) in the
reduction reaction of 4-NP at 30 C with 20-fold excess of NaBH4 .
The conversion and the activity values are presented in Table 3.
As revealed in the table, even p(AMPS)Co reactor vessels stored
for 15 days have 90% conversion and almost 80% activity. This is a
considerable achievement giving these systems great potential for
sequential and long term usage for providing catalytic activities in
the reduction of toxic nitrophenols.

0.12
Table 3
The shelf-life measurements of the same p(AMPS)Co catalyst systems at different
storage time intervals.

0.08
0.04

Days

0
0

10

12

14

Amount of catalyst (mg)


Fig. 8. The change in reduction rate constant of 4-NP with the amount of
Co catalyst inside p(AMPS)Co composites. (Reaction conditions: 50 mL [4NP] = 1.44 102 mol/L, [NaBH4 ] = 2.88 101 mol/L, 750 rpm, 30 C.)

Initial (0)
% conversion
% activity

100
100

1
100
100

10
99
98

93
91

15
90
79

Reaction conditions: 50 mL [4-NP] = 1.44 102 mol/L, [NaBH4 ] = 2.88 101 mol/L,
6.22 mg catalyst, 750 rpm, 30 C.

N. Sahiner et al. / Applied Catalysis B: Environmental 101 (2010) 137143

To further point out the versatility of the presented system here,


there are many reports on metallic nanoparticles catalyst that were
somehow situated on some sort of support such as zeolites, alumina, boehmite, PMMA, and polyelectrolyte brushes. Almost all the
support materials are hard and used only for valuable metal catalysts such as Au, Pt, and Pd [2729]. The support material used
in this investigation is smart hydrogel that is soft, exible with
adjustable pore size with external stimuli as well as ability to be
applicable for the synthesis of other kinds of catalysts. In fact, our
current research is focused on the synthesis of different catalysts
in situ and their application in different organic reactions.
4. Conclusions
This investigation demonstrated that p(AMPS) network can be
used in dual action: one of which is as a template for Co nanoparticle synthesis inside the hydrogel matrix and the other one is the
utilization of the hydrogelCo system as a reaction vessel in which
4-NP and 2-NP reductions were successfully carried out with very
high conversions and activities after up to 15 days storage. The
method presented here is novel in that metal nanoparticles can
easily be prepared at room temperature inside environmentally
friendly hydrogel networks, and used as a reaction vessel in the
reduction of toxic and/or industrially important organic reagents.
The system established here is pertinent as the hydrogel network
is exible and can swell and deswell in the presence and absence
of water, respectively, making them excellent adaptable materials to prevent aggregation of metal nanoparticles or clusters while
offering a lavish space (in their swollen state) for the reactions to
take place. These hydrogel networkmetal nanoparticle systems
are promising reaction-vessels and endow great opportunities for
many specic reactions. For instance, in a common catalytic reaction, a reactor is the whole dimension of the ask. On the other
hand, in the system presented here the reaction is limited to the
hydrogel network which behaves as a reactor, providing many benets and economical advantages in practical applications.
Acknowledgements
This work is supported by the Scientic and Technological
Research Council of Turkey (Grant No.: 108T122). Also, N. Sahiner

143

greatly acknowledges the Turkish Academy of Science for nancial


support under the TUBA-GEGIP 2008 program.
References
[1] Y. Lu, Y. Mei, M. Schrinner, M. Ballauff, M.W. Mller, J. Phys. Chem. C 111 (2007)
7676.
[2] S.K. Ghosh, M. Mandal, S. Kundu, S. Nath, T. Pal, Appl. Catal. A: Gen. 268 (2004)
61.
[3] G. Sharma, Y. Mei, Y. Lu, M. Ballauff, T. Irrgang, S. Proch, R. Kepme, J. Catal. 246
(2007) 10.
[4] A.C. Patel, S. Li, C. Wang, W. Zhang, Y. Wei, Chem. Mater. 19 (2007)
1231.
[5] S. Jana, S. Pande, S. Panigrahi, S. Praharaj, S. Basu, A. Pal, T. Pal, Langmuir 22
(2006) 7091.
[6] D.M. Dotzauer, J. Dai, L. Sun, M.L. Bruening, Nano Lett. 6 (2006) 2268.
[7] S. Panigrahi, S. Basu, S. Praharaj, S. Pande, S. Pande, S. Jana, A. Pal, S.K. Ghosh, T.
Pal, J. Phys. Chem. C 111 (2007) 4596.
[8] Y. Chen, J. Qiu, X. Wang, J. Xiu, J. Catal. 242 (2006) 227.
[9] N. Sahiner, Colloid Polym. Sci. 285 (2006) 283.
[10] N. Sahiner, Colloid Polym. Sci. 285 (2007) 413.
[11] H. Liu, M. Liu, S. Jin, S. Chen, Polym. Int. 57 (2008) 1165.
[12] C. Tan, R. Lu, P. Xue, C. Bao, Y. Zhao, Mater. Chem. Phys. 112 (2008)
500.
[13] D.Q. Wu, C.C. Chu, F.A. Chen, J. Mater. Sci.: Mater. Med. 19 (2008)
3593.
[14] S.H. Jang, Y.G. Jeong, B.G. Min, W.S. Lyoo, S.C. Lee, J. Hazard. Mater. 159 (2008)
294.
[15] N. Sahiner, A.M. Alb, R. Graves, T. Mandal, G.L. McPherson, W.F. Reed, V.T. John,
Polymer 48 (2007) 704.
[16] W. Li, H. Zhao, P.R. Teasdale, R. John, Polymer 43 (2002) 4803.
[17] S.U. Sonavane, M.B. Gawande, S.S. Deshpande, A. Venkataraman, R.V. Jayaram,
Catal. Commun. 8 (2007) 1803.
[18] J.S. Dileep Kumar, M.M. Ho, T. Toyokuni, Tetrahedron Lett. 42 (2001)
5601.
[19] Y. Zheng, K. Ma, H. Wang, X. Sun, J. Jiang, C. Wang, R. Li, J. Ma, Catal. Lett. 124
(2008) 268.
[20] T. Swathi, G. Buvaneswari, Mater. Lett. 62 (2008) 3900.
[21] H. Lu, H. Yin, Y. Liu, T. Jiang, L. Yu, Catal. Commun. 10 (2008) 313.
[22] V. Brezova, A. Blazkova, I. Surina, B. Havlinova, J. Photochem. Photobiol. A:
Chem. 107 (1997) 233.
[23] A. Wang, H. Yin, H. Lu, J. Xue, M. Ren, T. Jiang, Catal. Commun. 10 (2009)
2060.
[24] A. Wang, H. Yin, H. Lu, J. Xue, M. Ren, T. Jiang, Langmuir 25 (2009)
12736.
[25] A. Wang, H. Yin, M. Ren, H. Lu, J. Xue, T. Jiang, New J. Chem. 34 (2010)
708.
[26] S. Jana, S.K. Ghosh, S. Nath, S. Pande, S. Praharaj, S. Panigrahi, S. Basu, T. Endo,
T. Pal, Appl. Catal. A: Gen. 313 (2006) 41.
[27] S. Arora, P. Kapoor, M.L. Singla, React. Kinet. Mech. Catal. 99 (2010) 157.
[28] S. Wunder, F. Polzer, Y. Lu, Y. Mei, M. Ballauff, J. Phys. Chem. C 114 (2010) 8814.
[29] K. Kuroda, T. Ishida, M. Haruta, J. Mol. Catal. A: Chem. 298 (2009) 7.

You might also like