Вы находитесь на странице: 1из 8

Journal of Colloid and Interface Science 461 (2016) 203210

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


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

Fine CuO anisotropic nanoparticles supported on mesoporous


SBA-15 for selective hydrogenation of nitroaromatics
Shweta Sareen a, Vishal Mutreja b, Satnam Singh a, Bonamali Pal a,
a
b

School of Chemistry and Biochemistry, Thapar University, Patiala 147004, Punjab, India
Department of Chemistry, Maharishi Markandeshwar University, Mullana 133207, Haryana, India

g r a p h i c a l a b s t r a c t
This paper deals with the study of co-relation of surface structural and morphological properties of CuO anisotropic nanoparticles supported on
mesoporous SBA-15, as a function of increased Cu loading for selective hydrogenation of nitroaromatics.

a r t i c l e

i n f o

Article history:
Received 9 June 2015
Revised 30 August 2015
Accepted 1 September 2015
Available online 1 September 2015
Keywords:
Anisotropic CuO nanoparticles
CuO/SBA-15 nanocomposites
Hostguest systems
Selective nitroaromatic reduction
Zeolite analogues
Mesoporous SBA-15
Supported catalyst
CuO nanostructures
Heterogeneous catalysis
Post modification

Corresponding author.
E-mail address: bpal@thapar.edu (B. Pal).
http://dx.doi.org/10.1016/j.jcis.2015.09.002
0021-9797/ 2015 Elsevier Inc. All rights reserved.

a b s t r a c t
SBA-15 modified with APTMS (3-aminopropyl trimethoxysilane) having pore diameter (8 nm) has been
synthesized and impregnated with 110 wt.% Cu using Cu(NO3)2 as a metal source followed by calcination at 350 C. As-prepared CuO/ap-SBA-15 powder showed changes in the color from white for bare
SBA-15 to light green due to formation of anisotropic CuO nanoparticles that exhibited a characteristic
plasmon absorption band at 359 and 747 nm. TEM studies showed a change in the morphology of CuO
NPs as a function of increased Cu loading. Moreover, well dispersed CuO nanospheres (56 nm) and
nanorods (aspect ratio 1120 nm) having monoclinic crystal phase were observed within the mesoporous channels of SBA-15. Elemental mapping studies confirmed uniform distribution of CuO nanoparticles on the surface of SBA-15. An increase in surface area was also observed from 694 m2 g 1 for SBA-15
to 762 m2 g 1 for 10 wt.% Cu loading probably due to the deposition of excess of CuO nanoparticles on the
outer siliceous surface. The catalytic activity also increased with Cu loading and 10 wt.% CuO/ap-SBA-15
catalyst displayed the highest catalytic activity for the reduction of m-chloronitrobenzene and
m-nitrotoluene with 83% and 100% selectivity for m-chloroaniline and m-aminotoluene respectively.
2015 Elsevier Inc. All rights reserved.

204

S. Sareen et al. / Journal of Colloid and Interface Science 461 (2016) 203210

1. Introduction
Synthesis of metal nanoparticles (NPs) of varying shapes and
sizes have received great attention in the last decade due to their
unique physico-chemical properties leading to vast applications
in the field of electronics [1], sensing [2] and catalysis [3,4] as compared to their bulk counterparts. Extensive work has been done in
the synthesis of noble metal NPs like Au [5], Ag and Pd [6] and Pt
[7]. Though these catalysts exhibit excellent catalytic activity, but
their high cost limits their applicability on a large scale, stimulating studies on non-precious metals. Compared with noble metals,
Cu and CuO being economically viable and easily available, offers
a promising material owing to its wide applications in the field
of optics, electronics and catalysis [8,9]. Cu and CuO NPs has been
used as a catalyst in many organic reactions like low temperature
CO oxidation [10], oxidation of alcohols [11], click synthesis [12]
and cross coupling reactions [13]. Most of the catalytic systems
being homogeneous suffer with the problem of separation of the
catalyst from the mixture. Moreover, the synthetic procedures rely
on elaborate processes utilizing hazardous chemicals, strong
reducing agents like NaBH4 [14] and capping agents like PVP [15]
with harsh experimental conditions. Though capping agents prevent aggregation, but also lead to inhibition of the molecular access
to the surface of the catalyst leading to decrease in the catalytic
efficiency. Hence, there is a need for the development of an ecofriendly, economical route for the synthesis of stabilized metallic
nanostructures. This can be achieved through heterogenisation of
homogeneous catalysts i.e., by immobilizing metal NPs on some
suitable support which is desirable for the stabilization and dispersion of metal NPs.
Mesoporous silica materials (MSM) such as SBA-15 are preferred to be an ideal host for the deposition of metal NPs as they
possess a high surface area and pore volume, greater hydrothermal
stability, hexagonal structure with tunable pore diameter (5
30 nm) with minimum hindrance, thus allowing easy diffusion
[1619] of the reactants. It is accepted that the unique architecture
of SBA-15 can contain NPs within mesopores with improved dispersal leading to the enhancement of their catalytic efficiency
[2022]. A great deal of research has been dedicated to the incorporation of various precious metals like Pt [23], Ag [24], Au and
Pd [25] on SBA-15 by different chemical approaches and their
use as catalyst for catalyzing various oxidationreduction reactions. In recent decades, Cu modified molecular sieves have shown
to be a good catalyst for different reactions. Ghosh et al. [26] synthesized Cu/SBA-15 nanocomposites by post modification in the
presence of NaBH4 as reducing agent for reduction of dyes. Zhang
et al. [27] prepared a series of Cu/SBA-15 nanocomposites by
evaporation induced self assembly route for hydroxylation of
phenol. Gu et al. [28] developed well dispersed Cu nanospecies
within SBA-15 channels by post modified method for cyclohexane
oxidation. However, there are a few reports on the synthesis of
supported CuO nanocomposites for catalyzing different oxidationreduction reactions. Zhong et al. [29] optimized the preparation of CuO/SBA-15 nanocomposites by varying the metal loading
and autoclaving temperature for catalyzing wet peroxide oxidation
of phenol and found that nanocomposite with lower metal loading
(4 wt.%) exhibited higher catalytic activity while higher loading
(10 wt.%) resulted in CuO aggregates within the mesopores of the
host. Chen et al. [30] compared the synthesis of CuO/SBA-15
nanocomposites by different post modified methods such as incipient wetness impregnation (IWI), deposition precipitation (DP),
grafting and homogeneous deposition precipitation (HDP), and
found that Cu/SBA-15 prepared by HDP method exhibited the
highest catalytic activity for the hydrogenolysis of dimethyl maleate to 1,4 butanediol. In contrast, the impregnation method
formed larger CuO NPs (size >90 nm) aggregated on the outer

surface beside blocking the pores. So, the major difficulties


involved with the above reported methods are uncontrolled
growth of metal NPs [31] resulting in their agglomeration and
the low amount of loading [10] that restricts the application of this
method leading to decrease in catalytic efficiency. So, synthesis of
well dispersed supported metal/metal oxide nanocomposites with
high catalytic efficiency even at higher metal loading, by a simple,
sustainable process involving green chemistry credentials is still a
challenge. Moreover, it has been found that CuO supported SBA-15
materials have been less explored in case of hydrogenation reactions particularly, nitroaromatic reduction. The hydrogenation of
nitroaromatics lead to the formation of functionalized anilines that
form important intermediates in the field of pharmaceuticals,
agrochemicals, polymers, dyes and pigments [32]. These anilines
are generated either through catalytic or non-catalytic process.
The non-catalytic process is associated with generation of hazardous waste and limitations of product isolation whereas the catalytic reduction involves the use of expensive metal based reaction
systems under harsh reaction conditions (e.g., 100 C temp., hydrogen sources such as CO/H2O, NH2NH2) with prolonged reaction
times, and poor reducibility. In this context, the present study
deals with an economical CuO catalyst for reduction of nitroaromatics with NaBH4 by a simple but efficient procedure under mild
reaction conditions.
We have previously reported [33] the synthesis of 4 wt.% of Au,
Ag and Cu loaded SBA-15 nanocomposites by post modified
method and compared their surface structural, morphological
and catalytic properties. It was found that of Au (size 5 nm), Ag
(size 11 nm) and Cu (size 13 nm) NPs, Au being the most
reactive exhibited the highest catalytic activity for the reduction
of m-dinitrobenzene. But, keeping in view the cost effectiveness,
rich abundance and environment friendly properties of CuO and
to gain further understanding of the changes in physicochemical
parameters as a function of increased metal loading, we have
prepared SBA-15 supported CuO nanocomposites (size 58 nm)
of varying wt.% of Cu by simple post modified method with small
modifications, i.e., using thermal treatment under inert atmosphere, but without employing any reducing agents and evaluated
their catalytic activity for reduction of m-chloronitrobenzene
(CNB) and m-nitrotoluene (NT) respectively.

2. Experimental section
2.1. Chemicals
Pluronic (M.W. 5800, EO20PO70EO20), tetraethoxysilane (TEOS),
copper nitrate Cu(NO3)23H2O, sodium borohydride (NaBH4) and
3-aminopropyltrimethoxysilane (APTMS) were obtained from
Sigma Aldrich. Aniline, m-chloronitrobenzene, m-chloronitroaniline, m-nitrotoluene, m-aminotoluene and other reagents were
obtained from Loba Chemie, India and were used as received without further purification. De-ionized water with ultra-pure filtration
system (Milli-Q, Millipore) was used throughout the experiments.

2.2. Synthesis of copper oxide loaded APTMS modified SBA-15


Mesoporous silica was prepared by reported method [34] using
nonionic triblock copolymer surfactant EO20PO70EO20 (P123) as
surface directing agent. In a typical synthesis, 2 g of nonionic triblock copolymer surfactant EO20PO70EO20 (P123) was dispersed
in 60 ml of 2 M HCl under stirring at 40 C until a clear solution
was obtained. Then, 0.025 mol of TEOS were added to the above
solution and the contents were stirred for 24 h at 40 C, transferred
into an autoclave and aged at 95 C for 72 h. The suspension, thus

S. Sareen et al. / Journal of Colloid and Interface Science 461 (2016) 203210

obtained was filtered, dried at 60 C and finally calcined at 550 C


for 8 h resulting in SBA-15.
Cu was impregnated within SBA-15 channels by a post grafting
method [33] with slight modifications. In a typical preparation, 1 g
of prepared SBA-15 was suspended under stirring in 50 ml of 2 wt.%
of APTMS-ethanol solution for 3 h, filtered, washed with ethanol
and dried at 60 C resulting in white powder designated as
ap-SBA-15. APTMS functions as a stabilizer by building strong
interaction between metal precursor and silica support. Then, 1 g
of ap-SBA-15 was mixed with an appropriate amount of 1 mM
Cu(NO3)2(aq.) solution under stirring for 2 h to allow chelation of
Cu2+ with APTMS. The obtained solution was filtered, rinsed with
deionized water and finally calcined at the 350 C for 3 h under
inert atmosphere to decompose the precursor, leading to the
formation of light green colored samples designated as X wt.%
CuO/ap-SBA-15 where X is amount of Cu loading.
2.3. Catalyst characterization
Different techniques viz., powder XRD, transmission electron
microscope (TEM), elemental mapping, EDX, BET, TGA, FT-IR and
Solid state UVvis absorbance spectra were used to characterize
the prepared materials. Powder X-ray diffraction (XRD) patterns
was recorded on Analytical Expert Pro diffractometer utilizing Cu
Ka radiation (k = 1. 54 ) within the 2h range of 0.55 in step size
of 0.02 and scan step time of 86.19s and within the 2h range of
1080 in a step size of 0.0130 and scan step time of 29.07s.
Hitachi (H-7500) 120 kW was used to obtain transmission electron
micrographs. HR-TEM, elemental mapping and EDX images were
recorded on FEI Tecnai F20 microscope operated at 200 kV. XPS
was carried out using KRATOS-AXIS DLD spectrometer (Kratos
Analytical, U.K.) equipped with monochromatic Al K radiation at
1486.6 eV operated at 10 kW. Solid state UVvisible absorption
spectra were obtained within the range of 400800 nm by using
Analytikjena Specord 205 spectrophotometer. Nitrogen adsorptiondesorption measurements were determined by pretreatment
of 20 mg of all samples at the 200 C under vacuum for 2 h using
BET surface area analyzer (BEL Sorp-max). TGA-50 Shimadzu
Thermogravimetric analyzer was used for thermal analyses within
the temperature range of 100800 C at a heating rate of 1 C/min
under nitrogen and ambient atmospheric pressure. IR studies were
performed on Cary 660 series Agilent FTIR spectrometer within the
range of 4004000 cm 1using KBr pellet method. Agilent Microwave Plasma Atomic Emission Spectrometer (MP-AES) was used
for performing MP-AES analysis in order to determine the amount
of Cu loading in the synthesized samples.
2.4. Catalytic activity
Catalytic activity and stability of CuO/ap-SBA-15 catalysts were
evaluated for reduction of m-CNB and m-NT by mixing catalyst
(10 mg), substrate (10 ml, 3 mM) in ethanol and sodium borohydride (1 ml, 0.3 M) with constant stirring at 70 C, respectively.
The progress of the reaction was monitored on HPLC (Agilent,
1120 compact LC using C-18 column) at wavelength k = 220 nm
with a flow rate of 1 ml/min using MeOH: H2O (70:30) as mobile
phase.
2.5. Results and discussion
2.5.1. Surface structural studies
Low angle XRD patterns for different wt.% Cu loading (Fig. 1a)
exhibited similar XRD spectra with well resolved diffraction peaks
at 1.0, 1.6 and 1.9 corresponding to 100, 110 and 200 planes of the
2D hexagonal structure [25] of p6mm symmetry characteristic of
MSM. It indicated the retention of long range mesopore ordering

205

and textural uniformity of SBA-15 even after metal impregnation.


However, a decrease in the intensity of 110 and 200 peaks was
observed with the loading of Cu on the surface of amino functionalized SBA-15 due to pore filling effect leading to decrease in electron density contrast upon introduction of CuO nanospecies within
the mesochannels of the silica host. Ungureanu et al. [35] also
reported the decrease in the intensity of peaks due to the partial
filling of SBA-15 mesopores with NiO for impregnated NiOCuO/
SBA-15 materials. Further, as Cu loading was increased from 1 to
10 wt.%, the peak corresponding to (1 0 0) plane shifted to lower
2h value (from 1.00 to 0.97) illustrating an increase in structural
parameters viz., lattice spacings (d100) and unit cell parameters
(ao) (S.I. Table 1). The shifting of the (1 0 0) peak can be attributed
to the slight disordering of the pore channels due to development
of the strain arising from the confinement of the CuO NPs within
the mesopores. This further indicated the existence of CuO NPs
within the mesoporous host. Similar findings have been reported
elsewhere [36,37]. The shifting of the peaks to lower angle resulted
in increase in various structural parameters viz., lattice spacings
(d100) and unit cell parameters (ao) (S.I. Table 1) that further illustrated different dispersion and mode of assembly of CuO NPs
within/on the surface of a mesoporous host structure. However,
when the Cu loading was increased to 5 wt.% and 10 wt.%, the
peaks begin to fade away implying slight disruption of mesoporous
structure at higher metal loading.
Wide angle XRD patterns (Fig. 1b) of varying wt.% CuO/ap-SBA15 nanocomposites exhibited a broad band at 22 [38] implying
the amorphous nature of SBA-15. Moreover, diffraction peaks at
2h = 32, 35.4, 38.3, 48.4 and 61.3 corresponding to monoclinic
phase of CuO NPs (JCPDS card no. 48-1548) was also observed.
The intensity of the peaks was directly dependent upon the metal
loaded on mesoporous silica. Presence of very small and broad
diffraction peaks in positions corresponding to CuO NPs were
indicative of small crystallite size probably resulting from the confined growth of CuO within the mesoporous channels. Gu et al. [39]
and Meer et al. [40] also reported broader diffraction peaks for
crystallites formed within the mesochannels of SBA-15. However,
the spectra depicted an increase in intensity and sharpness of the
peaks with an increase in Cu loading from 1 to 4 wt.%, illustrating
an increase in size of CuO NPs dispersed within/on the surface of
mesoporous silica. This is supported by an increase in crystallite
size i.e., 5.5 nm, 6.5 nm, 7.1 nm and 8.3 nm for 1, 2, 3 and 4 wt.%
Cu loading respectively as calculated by the Scherrer equation for
Cu (1 1 1) peak. Moreover, the (1 1 0) peak broadened in comparison
to the (1 1 1) and (2 0 0) peaks for the diffraction pattern of 4 wt.%
CuO/ap-SBA-15 signifying that with increase in Cu loading, effective amount of CuO nanospecies within the mesopores increase
gradually leading to the development of strain. As a result, crystal
deformations w.r.t. specified planes also increases leading to the
formation of anisotropic NPs (NS and NR) within mesoporous
sieves and has been further supported by TEM studies. Strzalka
et al. [41] also reported that differences in the intensity and FWHM
(full width half maxima) of the XRD peaks (wide angle) illustrate a
change in the morphology of embedded Ag NPs within the SBA-15
for Ag/SBA-15 nanocomposites. Ma et al. [42] also stated that
increase in intensity and narrowing of peaks with increased Ag
loading indicated a change in the morphology of Ag NPs for Ag/
SBA-15 materials. However, further increase in Cu loading to 5
and 10 wt.%, did not had any negative effect on the dispersion, as
it did not lead to a considerable increase in size of CuO metal
NPs, which was found to be 7.1 nm and 7.7 nm respectively. In fact,
the mesoporous channels acted as a nanoreactor restricting the
size, probably resulting in a change in morphology of CuO NPs.
However, beyond 4 wt.% Cu loading, the reflections become weak
and diffuse implying poor crystallinity [43] due to partial
disruption of SBA-15 structure as a result of increased Cu loading.

206

S. Sareen et al. / Journal of Colloid and Interface Science 461 (2016) 203210

(a)

(b) SiO2

10 wt. %

(110)
(111)
(200)

10 wt.%

5 wt. %

5 wt.%

(202)

Intensity (a.u.)

Intensity (a.u.)

4 wt. %

3 wt. %
2 wt. %

4 wt.%
(113)

3 wt.%

2 wt.%

1 wt. % CuO/ap-SBA-15
(100)

1.0

1 wt.% CuO/ap-SBA-15

(110) (200) SBA-15


1.5

2.0

2.5

2 (degree)

10

20

30

40

50

60

70

80

2 (degree)

Fig. 1. (a) Low angle and (b) wide angle XRD patterns of bare SBA-15 and various wt.% CuO/ap-SBA-15 catalysts.

Moreover, the effect of increase in calcination (S.I Fig. 1) temperature up to 550 C on the dispersion and size of 4 wt.% and 10 wt.%
CuO/ap-SBA-15 nanocomposites was also studied. It showed an
increase in the intensities and sharpness of all the peaks, indicating
the formation of large sized CuO NPs within the mesoporous
matrix.
TEM micrographs (Fig. 2a) displayed regular hexagonal
mesoporous SBA-15 structure with cylindrical channels exhibiting
narrow pore size distribution. Although, no disruption of mesoporous channels was observed yet large differences in morphology
of CuO NPs were observed as a function of increased Cu loading.
Fine CuO NPs (size 5 nm, S.I Fig. 2) were depicted as black dots
well dispersed within the light background of mesoporous support
for 1 wt.% Cu loading (Fig. 2b). With increased metal loading to
2 wt.% (Fig. 2c), uniformly distributed CuO NS (size 6 nm, S.I
Fig. 2) were noticed within mesochannels. These results are in
accordance with wide angle XRD studies. However, with increase
in Cu loading to 4 wt.%, instead of aggregation, anisotropic CuO
NPs (Fig. 2d) were observed as trapped within as well as on the
surface of siliceous host. Both spherical and rod like CuO NPs were
observed with diameter 8 nm (in consistency with a pore diameter of SBA-15) and length 80150 nm. Due to strong metalsupport interaction and restriction from channel walls, excessive
CuO NPs (with increased Cu loading) were forced to align one after
another resulting in change in morphology from spherical to rod
shape (marked in Fig. 2d). Similar changes in particle morphology
with increased metal loading have been reported elsewhere
[40,44]. Moreover, Chamber et al. [45] also reported that due to
strong metal-support interaction, Ni NPs supported on carbon
nanofibres adopt different morphologies. For 5 wt.% Cu loading,
small NR (aspect ratio 11 nm, S.I. Fig. 2) along with nanobundles (Fig. 2e) formed by the growth of some nanorods in adjacent
mesopores were also observed. However, larger CuO NR (aspect
ratio 20 nm, S.I. Fig. 2) were seen homogeneously dispersed
and extending throughout the entire mesochannels for 10 wt.%
Cu loading (Fig. 2f) due to the perfect alignment of CuO NPs forming rod like morphology. Furthermore, the lattice distance
(0.25 nm) measured from HR-TEM (S.I. Fig. 3a and b) was found
to be in agreement with (1 1 1) plane of crystalline CuO (JCPDS
card no. 48-1548). Moreover, the selective area electron diffraction
(SAED) pattern (S.I. Fig. 3c) confirmed (1 1 1), (1 1 1) and (2 0 2)
planes corresponding to the presence of CuO nanospecies. Its
corresponding EDX spectra (S.I. Fig. 3d) further established the

presence of well dispersed and crystalline CuO within mesoporous


support with 1.29 wt.% Cu loading. Elemental mapping studies
(Fig. 3) also confirmed the uniform distribution of CuO NPs on
the mesoporous host.
XPS spectrum of 10 wt.% CuO/ap-SBA-15 (Fig. 4) depicted a Cu
2p3/2 peak at binding energy of 934 eV with a satellite at
943.3 eV which is the characteristic of Cu2+ species. Similar results
have been reported by Zhang et al. [27] for Cu/SBA-15 nanocomposites. The value of Cu 2p3/2 peak was little higher than the
reported value of bulk CuO [46], attributed to the uniform dispersion of CuO nanospecies within MSM.
2.5.2. Optical studies
The solid state UVVisible absorption spectra (Fig. 5) of different wt.% Cu impregnated SBA-15 catalysts depicted a sharp band
at 359 nm due to the charge transfer between mononuclear Cu2+
and oxygen in (CuOCu)n surface species, indicating the presence
of some Cu oligomers or CuO clusters in the extraframework position. Similar results have been reported by Chen et al. [47] and
Zhang et al. [48] for impregnated Cu/SBA-15 and Cu/MCM-41 samples respectively. However, a broad band at 747 nm can be attributed to the dd transition of Cu2+ in a pseudo-octahedral ligand
oxygen environment, implying the presence of CuO NPs [49,50],
while no band was observed for bare SBA-15. An increase in the
absorbance intensities further reflect an increase in Cu loading
with color change from white for bare SBA-15 to light green indicating the presence of Cu in the samples.
The FTIR spectra represented (S.I. Fig. 4) a broad absorption
band at 3460 cm 1, sharp bands at 1629 cm 1and 973 cm 1
assigned to the stretching and bending vibrations of SiOH bonds.
In addition, characteristic bands of absorption at 1077, 790 and
463 cm 1 corresponding to asymmetric and symmetric stretching
vibrations of SiOSi bonds of the host SBA-15 matrix [17] were
also observed in all the samples. The presence of a band at
2930 cm 1 attributed to NH+3 stretching confirmed the presence
of aminopropyl groups [51] in ap-SBA-15. Similar spectra for apSBA-15 and Cu loaded samples indicated retention of structural
integrity of SBA-15 even after surface functionalisation and metal
impregnation in accordance with the XRD and TEM studies. However, slight decrease in the intensity of silanol groups is due to the
inclusion of Cu within the mesoporous sieves. Similar findings
have been reported by Gao et al. [52] for Vanadium loaded SBA15 materials.

S. Sareen et al. / Journal of Colloid and Interface Science 461 (2016) 203210

207

Fig. 2. TEM images of (a) SBA-15, (b) 1 wt.% CuO/ap-SBA-15, (c) 2 wt.% CuO/ap-SBA-15, (d) 4 wt.% CuO/ap-SBA-15, (e) 5 wt.% CuO/ap-SBA-15 and (f) 10 wt.% CuO/ap-SBA-15
catalyst.

2.5.3. Thermal analysis


The TGA pattern of SBA-15 and various wt.% CuO/ap-SBA-15 catalysts (Fig. 6) showed degradation in three different wt. loss zones.
The first wt. loss zone at 100 C (constituting 17% for SBA-15,
4 wt.% and 10 wt.% CuO/ap-SBA-15 respectively), is associated with
the loss of physically adsorbed water molecules in silica channels
and coordinated to Cu complexes [27]. The second wt. loss (8%
for SBA-15, 9% for 4 wt.% and 10% for 10 wt.% CuO/ap-SBA-15)
between 100 and 300 C may be attributed to the decomposition
of remaining surfactant (if any). The third wt. loss zone (6% for
SBA-15, 7% for 4 wt.% and 10% for 10 wt.% CuO/ap-SBA-15)
between 300 and 550 C can probably be assigned to the decomposition of the aminopropyl groups. This is further supported by the
wide angle XRD studies, where an increase in the size of CuO NPs
was observed with increase in calcination temperature to 550 C
probably due to the decomposition of APTMS. However, a small
wt. loss of 3% for SBA-15, 7% for 4 wt.% and 9% for 10 wt.%
CuO/ap-SBA-15 above 570 C is related to the combustion of remaining carbon species and dehydroxylation of SiOH groups [53].

2.5.4. Textural properties


Nitrogen adsorptiondesorption isotherm of SBA-15 exhibited a
typical type IV isotherm (S.I. Fig. 5) displaying HI hysteresis loop
with relative pressures between 0.65 and 0.8 characteristic of MSM
[54] having a narrow pore size distribution of cylindrical channels.
Similar shape of the CuO/ap-SBA-15 isotherms as that of SBA-15
further revealed that mesostructural ordering of SBA-15 was maintained even after Cu loading [44] as suggested by low angle XRD
and TEM studies. In comparison to SBA-15, the capillary condensation step for various wt.% CuO/ap-SBA-15 slightly shifted toward
lower relative pressure (from 0.65 to 0.6) signifying a decrease in
pore diameter with Cu loading. Moreover, the BET profile for
1 wt.% CuO/ap-SBA-15 indicated a decrease in volume of adsorbed
nitrogen implying a decrease in the surface area in comparison to
bare SBA-15. This decrease in surface area can be attributed to the
pore filling effect [55] by the formation of CuO NPs within the
mesopores simultaneously followed by an increase in wall thickness (S.I. Table 1). Moreover, for higher Cu loadings of 4 and
10 wt.%, respectively, forced closure of the hysteresis loop on the

208

S. Sareen et al. / Journal of Colloid and Interface Science 461 (2016) 203210

Fig. 3. STEM image and elemental mapping of 4 wt.% CuO/ap-SBA-15 catalyst.

desorption branch at p/po 0.6 was noticed. This results from the
cavitation phenomenon [56] that occurs due to the partial blockage
of SBA-15 pore openings with the increased amount of Cu impregnation and has been further supported by the increase in the surface area (more than that of SBA-15) for both 4 and 10 wt.% Cu
loaded materials possibly due to the deposition of excess of CuO
NPs (as shown in TEM studies) to the outer silica surface due to
the partial blockage of SBA-15 mesopores. Patel et al. [57] also
found that at higher Cu loading (10 wt.%) deposition of excess of
CuO NPs on the external surface resulted in an increase in the surface area for CuO/SBA-15 nanocomposites.
2.5.5. Catalytic activity
The catalytic activity of prepared CuO/ap-SBA-15 nanocomposites was evaluated for the reduction of m-substituted nitroaromatics such as CNB (Scheme 1) and NT (Scheme 2) to their respective
amines. It was observed that in the absence of CuO/ap-SBA-15 catalysts, reduction of nitroaromatics with NaBH4 and with bare SBA15 did not took place, implying that metal oxide NPs were the real
active sites. Moreover, the reaction was initiated by the addition of
CuO/ap-SBA-15 catalyst suggesting adsorption of electron donor

(BH4 ) and electron acceptor (substrate) species on the surface of


the CuO loaded SBA-15. It seems that the catalyst merely provides
the adsorption sites and is not involved in the electron transfer
process. Instead the transfer of electrons takes place from BH4 to
the nitroaromatics through the CuO nanospecies. Similar results
has also been reported for CuO/c-Al2O3 [58]. Moreover, reduction
of m-CNB gave m-CAN as the major product accompanied by side
reaction such as dechlorination leading to the formation of aniline
whereas m-NT resulted in the formation of m-AT respectively (as
confirmed by HPLC analysis S.I. Figs. 6 and 7). Zhang et al. [59]
reported that the reactions catalyzed by supported metal (Pd) catalysts often exhibit moderate CAN selectivity with dechlorination
to aniline. Linear plots of ln (Ct/Co) vs time were obtained for all
catalysts indicating pseudo first order kinetics of the reaction (S.I.
Figs. 8 and 9). It was seen that the catalytic activity increased
with the increase in Cu loading from 1 to 10 wt.% leading to
increase in the selectivity of m-CAN from 66% to 83%. This can be
attributed to the fact that at lower Cu loading, Cu exist in the form
of very fine CuO NPs (56 nm) homogeneously dispersed and
deep-seated (due to the lower surface area i.e., 512 m2 g 1 for
1 wt.% CuO/ap-SBA-15) within the mesochannels of the host,
resulting in lesser accessibility of active sites toward the reactant
molecules. But at higher Cu loading of 410 wt.%, effective amount
of CuO NPs within the mesopores also gradually increased (confirmed by MP-AES analysis, 3.1 wt.% and 8.7 wt.% for 4 and 10 wt.%
CuO/ap-SBA-15 respectively) resulting in development of the
strain leading to slight disruption of the mesoporous channels. As
a result, NPs come to lie both within as well as on the surface of
the mesoporous host favoring improved access of CuO active sites
toward the reactant that probably resulted in higher catalytic
activity. This is further supported by the BET studies where
10 wt.% CuO/ap-SBA-15 exhibited highest surface area
(762 m2 g 1) due to the presence of CuO NPs within as well as on
the outer surface facilitating better adsorption of reactant molecules on the active catalytic sites. Thus, the catalytic activity is
found to be strongly influenced by the amount of Cu loading,
distribution of CuO NPs and surface area of the CuO/ap-SBA-15
nanocomposites. Similar findings have been reported by Kalbasi
[60], for nitroaromatic reduction catalyzed by Ni incorporated poly
amidoaminepolyvinylamine/SBA-15 nanocomposites. Jiang et al.
[61] also reported the presence of the active sites on the outer surface of the support to be the contributing factors for high catalytic
activity of supported Pt and Pd catalysts for the reduction of o-CNB.

Counts/s

934 eV
943.3 eV

930

935

940

945

950

Binding energy (eV)


Fig. 4. XPS spectra of 10 wt.% CuO/ap-SBA-15 catalyst.

Fig. 5. Solid state UVVisible absorption spectra of (a) bare SBA-15, (b) 1 wt.% CuO/
ap-SBA-15, (c) 4 wt.% CuO/ap-SBA-15 and (d) 10 wt.% CuO/ap-SBA-15 catalysts
[Inset shows the color change in SBA-15 with increase in metal loading].

S. Sareen et al. / Journal of Colloid and Interface Science 461 (2016) 203210

100

SBA-15
4 wt. % CuO/ap-SBA-15
10 wt. % CuO/ap-SBA-15

Zone 1

% Weight loss

209

90

80

Zone 2
Zone 3

70

60
100

200 300 400 500

600 700 800 900

Fig. 7. Product distribution of reduction of m-chloronitrobenzene by various wt.%


CuO/ap-SBA-15 catalysts.

Temperature (C)
Fig. 6. TGA of SBA-15 and various wt.% CuO/ap-SBA-15 catalysts.

As a result, of all the prepared catalysts, 10 wt.% Cu loaded catalyst


exhibited highest catalytic activity with 83% and 100% selectivity
of m-chloroaniline (CAN) and m-aminotoluene (AT) for m-CNB
and m-NT reduction respectively (Figs. 7 and 8). The catalytic activity also showed a dependence on the nature of substituents present on the nitrobenzene [56]. The reaction rate increased with
the increase in electron withdrawing effect of the substituents (-I
effect) and varied as m-CNB (4.52  10 2) > m-NT (1.47  10 3)
for 10 wt.% CuO/ap-SBA-15. .For m-CNB, presence of one NO2
group being electron withdrawing, lowered the electron density
favoring fast conversion of nitro to amino group resulting in higher
selectivity for m-CAN. However, it was also observed that the presence of the electron releasing group, i.e., CH3 group, increases the
electron density on the nitrobenzene and reduces the rate of reduction. Moreover, although selectivity was 100%, but yield obtained
was 81% for m-NT reduction.
In order to gain insight into the reusability and stability of the
CuO/ap-SBA-15 catalyst under the reaction conditions, recycling
experiments were carried out with 10 wt.% CuO/ap-SBA-15 catalyst for the reduction of m-NT. For recyclabilty, catalyst was recovered from the reaction mixture through filtration, washed with
ethanol, dried at the 60 C for 8 h and reused in the subsequent
run. It was found that the catalyst can be reused for successive four
recycles with a small decrease in the catalytic activity to 77% for
the reduction of m-NT. In addition, the low angle XRD pattern
(S.I. Fig. 10) of the reused catalyst after the fourth run depicted
a slight shifting (0.970.95) of the (1 0 0) peak toward the lower
angle, implying a slight increase in the lattice parameters probably
due to the hydrolysis of surface silica resulting in expansion of the
pore-wall. TEM image (S.I. Fig. 10) of the recycled catalyst did not
depicted any aggregation of the CuO NPs relative to that of the
fresh catalyst. The solid state UVVisible absorption spectra of
the recycled catalysts after 1st and 4th recycle also represented
bands at 343 nm and 753 nm respectively almost similar to that
of the fresh catalyst (absorption bands at 350 nm and 753 nm).
However, with a slight decrease in the intensity which may be
due to the loss of small amount of CuO during catalyst recovery

Scheme 1. Reduction of m-chloronitrobenzene.

Fig. 8. Product distribution of reduction of m-nitrotoluene by various wt.% CuO/apSBA-15 catalysts.

Scheme 2. Reduction of m-nitrotoluene.

(S.I. Fig. 11). Moreover, the recycled catalysts retained same color
(light green) as that of the fresh catalyst illustrating no change in
the chemical composition of the catalyst. These results suggests
the high stability and reusability of the catalyst.
3. Conclusion
In conclusion, it was demonstrated that metal loading had a significant effect upon the size, morphology, dispersion ability and
catalytic activity of the metal oxide NPs present within the mesoporous host. The high surface area of the 10 wt.% CuO/ap-SBA-15
also contributed to the efficient adsorption of nitro groups on the
active sites of the CuO/ap-SBA-15 resulting in enhanced catalytic
activity. The reaction rate showed dependence on the electron
withdrawing ability of the substituents present on nitrobenzene
and was found to be maximum for m-CNB. The easy preparation
method, use of economical catalyst, benign reaction conditions,
greater selectivity of products and high recyclability of the catalyst

210

S. Sareen et al. / Journal of Colloid and Interface Science 461 (2016) 203210

make this protocol to be highly efficient and appealing for the


reduction of nitroaromatics.
Acknowledgement
The authors are thankful to Department of Science and Technology for providing financial support under women scientist scheme
(File no. SR/WOS-A/CS-09/2102).
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.jcis.2015.09.002.
References
[1] G. Schmid, B. Corain, Eur. J. Inorg. Chem. (2003) 3081.
[2] Y. Sabri, S. Ippolito, A.O. Mullane, J. Tardio, V. Bansal, S. Bhargava,
Nanotechnology 22 (2011) 305501.
[3] V. Bansal, R. Ramanathan, S.K. Bhargava, Aust. J. Chem. 64 (2011) 279.
[4] A.P.O. Mullane, S.J. Ippolito, Y.M. Sabri, V. Bansal, S.K. Bhargava, Langmuir 25
(2009) 3845.
[5] A. Mahmoud, El-Sayed A. Mostafa, J. Am. Chem. Soc. 132 (2010) 12704.
[6] H. Kochkar, M. Aouine, A. Ghorbel, G. Berhault, J. Phys. Chem. C 115 (2011)
11364.
[7] J. Zhang, W. Yao, M. Liu, Q. Xu, Q. Wu, T. Zeng, Catal. Lett. 143 (2013) 1030.
[8] G.H. Chan, J. Zhao, E.M. Hicks, G.C. Schatz, R.P.V. Duyne, Nano Lett. 7 (2007)
1947.
[9] S. Magdassi, M. Grouchko, A. Kamyshny, Mater 3 (2010) 4626.
[10] C.H. Tu, A.Q. Wang, M.Y. Zheng, X.D. Wang, T. Zhang, Appl. Catal. A 297 (2006)
40.
[11] R.J. Kalbasi, A.A. Nourbakhshand, M. Zia, J. Inorg. Organomet. Polym. 22 (2012)
536.
[12] C.O. Kappe, E. Van der Eycken, Chem. Soc. Rev. 39 (2010) 1280.
[13] J. Dulle, K. Thirunavukkarasu, M.C.M. Hazeleger, D.V. Andreeva, N.R. Shiju, G.
Rothenberg, Green Chem. 15 (2013) 1238.
[14] P. Deka, R.C. Deka, P. Bharali, New J. Chem. 38 (2014) 1789.
[15] P. Abdulkin, Y. Moglie, B.R. Knappett, D.A. Jefferson, M. Yus, F. Alonso, A.E.H.
Wheatley, Nanoscale 5 (2013) 342.
[16] A. Sayari, B.H. Han, Y. Yang, J. Am. Chem. Soc. 126 (2004) 14348.
[17] R.M.M. Aranda, J. Top Cejka, Catal 53 (2010) 141.
[18] K. Cassiers, T. Linssen, M. Mathieu, M. Benjelloun, K. Schrijnemakers, P. van der
Voort, P. Cool, E.F. Vansant, Chem. Mater. 14 (2002) 2317.
[19] A. Galarneau, H. Cambon, F. Di Renzo, R. Ryoo, M. Choi, F. Fajula, New J. Chem.
27 (2003) 73.
[20] X. Liu, A. Wang, X. Wang, C.Y. Mou, T. Zhang, Chem. Commun. (2008) 3187.
[21] X. Yang, D. Chen, S. Liao, H. Song, Y. Li, Z. Fu, Y. Su, J. Catal. 291 (2012) 36.
[22] J. Sun, D. Ma, H. Zhang, X. Liu, X. Han, X. Bao, G. Weinberg, N. Pfander, D. Su, J.
Am. Chem. Soc. 128 (2006) 15756.
[23] R.M. Rioux, H. Song, J.D. Hoefelmeyer, P. Yang, G.A. Somorjai, J. Phys. Chem. B
109 (2005) 2192.
[24] X. Zhang, Z. Qu, X. Li, Q. Zhao, Y. Wang, X. Quan, Mater. Lett. 65 (2011) 1892.
[25] S.M. El-Sheikh, A.A. Ismail, J.F. Al-Sharab, New J. Chem. 37 (2013) 2399.
[26] B.K. Ghosh, S. Hazra, B. Naik, N.N. Ghosh, Powder Technol. 269 (2015) 371.

[27] H. Zhang, C. Tang, Y. Lv, C. Sun, F. Gao, L. Dong, Y. Chen, J. Colloid Interface Sci.
380 (2012) 16.
[28] J. Gu, Y. Huang, S.P. Elangovan, Y. Li, W. Zhao, I. Toshio, Y. Yamazaki, J. Shi, J.
Phys. Chem. C 115 (2011) 21211.
[29] X. Zhong, J. Barbier Jr., D. Duprez, H. Zhang, S. Royer, Appl. Catal., B: Environ.
123 (2012) 121.
[30] L.F. Chen, P. J Guo, L.J. Zhu, M.H. Qiao, W. Shen, H.L. Xu, K.N. Fan, Appl. Catal., A:
General 356 (2009) 129.
[31] J.S. Yang, W.Y. Jung, G.D. Lee, S.S. Park, E.D. Jeong, H.G. Kim, S.S. Hong, J. Ind.
Eng. Chem. 14 (2008) 779.
[32] R.S. Downing, P.J. Kunkeler, H.V. Bekkum, Catal. Today 37 (1997) 121.
[33] S. Sareen, V. Mutreja, S. Singh, B. Pal, RSC Adv. 5 (2015) 184.
[34] D.Y. Zhao, J.L. Feng, Q.S. Huo, N. Melosh, G.H. Fredrickson, B.F. Chmelka, G.D.
Stucky, Science 279 (1998) 548.
[35] A. Ungureanu, B. Dragoi, A. Chirieac, C. Ciotonea, S. Royer, D. Duprez, A.S.
Mamede, E. Dumitriu, Appl. Mater. Interfaces 5 (2013) 3010.
[36] S. Ajitha, S. Sugunan, J. Porous Mater. 17 (2010) 341.
[37] P.B. Lihitkar, S. Violet, M. Shirolkara, J. Singh, O.N. Srivastava, R.H. Naik, S.K.
Kulkarni, Mater. Chem. Phys. 133 (2012) 850.
[38] W.H. Zhang, J.J. Shi, L.Z. Wang, D.S. Yan, Chem. Mater. 12 (2000) 1408.
[39] J. Gu, W. Fan, A. Shimijima, T. Okubo, J. Solid State Chem. 181 (2008) 957.
[40] J.V. Meer, I. Bardez, F. Bart, P.A. Albouy, G. Wallez, A. Davidson, Micropor.
Mesopor. Mat. 118 (2009) 183.
[41] M.Z. Strzakaa, S.P. Patkowska, M. Kozak, S. Pikus, Appl. Surface Sci. 266 (2013)
337.
[42] L. Ma, L. Jia, X. Guo, L. Xiang, Chin. J. Catal. 35 (2014) 108.
[43] M.A. Karakassides, A. Bourlinos, D. Petridis, L.C.G. nte, P. Labbe, J. Mater.
Chem. 10 (2000) 403.
[44] J. Taghavimoghaddam, G.P. Knowles, A.L. Chaffee, J. Mol. Catal. A: Chem. 358
(2012) 79.
[45] A. Chambers, T. Nemes, N.M. Rodriguez, R.T.K. Baker, J. Phys. Chem. B 102
(1998). 2251.
[46] A.G. Kong, W.H. Wang, X. Yang, Y.W. Hou, Y.K. Shan, Micropor. Mesopor. Mat.
118 (2009) 348.
[47] C.S. Chen, Y.T. Lai, T.W. Lai, J.H. Wu, C.H. Chen, J.F. Lee, H.M. Kao, ACS Catal. 3
(2013) 667.
[48] G. Zhang, J. Long, X. Wang, Z. Zhang, W. Dai, P. Liu, Z. Li, L. Wu, X. Fu, Langmuir
26 (2) (2010) 1362.
[49] A.L. Petre, J.B. Carbajo, R. Rosal, E.G. Calvo, J.A.P. Meln, Chem. Eng. 225 (2013)
162.
[50] M.R. Prasad, G. Kamalakar, S.J. Kulkarni, K.V. Raghavan, J. Mol. Catal. A: Chem.
180 (2002) 109.
[51] X. Wang, K.S.K. Lin, J.C.C. Chan, S. Cheng, J. Phys. Chem. B 109 (2005) 1763.
[52] F. Gao, Y. Zhang, H. Wan, Y. Kong, X. Wu, L. Dong, B. Li, Y. Chen, Micropor.
Mesopor. Mat. 110 (2008) 508.
[53] X.S. Zhao, G.Q. Lu, A.K. Whittaker, G.J. Millar, H.Y. Zhu, J. Phys. Chem. B 101
(1997) 6525.
[54] K.S.W. Sing, D.H. Everett, R.A.W. Haul, L. Moscow, R.A. Pierotti, J. Rouquerol, T.
Siemieniewska, Pure Appl. Chem. 57 (1985) 603.
[55] D.H. Lin, Y.X. Jiang, Y. Wang, S.G. Sun, J. Nanomater. (2008) 473791.
[56] P. Van Der Voort, P.I. Ravikovitch, K.P. De Jong, A.V. Neimark, A.H. Janssen, M.
Benjelloun, E. Van Bavel, P. Cool, B.M. Weckhuysen, E.F. Vasant, Chem.
Commun. (2002) 1010.
[57] A. Patel, T.E. Rufford, V. Rudolph, Z. Zhu, Catal. Today 166 (2011) 188.
[58] S.U. Nandanwar, M. Chakraborty, Chin. J. Catal. 33 (9) (2012) 1532.
[59] Y.P. Zhang, S.J. Liao, Y. Xu, Tetrahedron Lett. 35 (1994) 4599.
[60] R.J. Kalbasi, F. Zamani, RSC Adv. 4 (2014) 7444.
[61] L. Jiang, H. Gu, X. Xu, X. Yan, J. Mol. Catal. A: Chem. 310 (2009) 144.

Вам также может понравиться