Sie sind auf Seite 1von 12

Journal of Catalysis 336 (2016) 1122

Contents lists available at ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Shape-dependent catalytic activity of CuO nanostructures


Suraj Konar, Himani Kalita, Nagaprasad Puvvada, Sangeeta Tantubay, Madhusudan Kr. Mahto,
Suprakash Biswas, Amita Pathak
Department of Chemistry, Indian Institute of Technology, Kharagpur, West Bengal 721302, India

a r t i c l e

i n f o

Article history:
Received 3 November 2015
Revised 21 December 2015
Accepted 23 December 2015
Available online 27 January 2016
Keywords:
Copper oxide
Nanostructures
Catalytic activity
4-Nitrophenol
NaBH4

a b s t r a c t
Ligand-regulated growth of various morphologies of copper oxide (CuO) nanostructures has been
achieved through an aqueous-based chemical precipitation route where the Cu2+ ions were stabilized
through complexation with different organic acids (viz. acetic/citric/tartaric acid). The rod-, spherical-,
star-, and flower-shaped morphologies of the CuO nanostructures, obtained in the presence of the
different carboxylic acids, have been characterized using XRD, FTIR, HRTEM, DLS, and zeta potential measurements, while their specific surface areas and the associated band gap energies have been compared.
The catalytic performance of the different CuO nanostructures has been affirmed by monitoring the
reduction of 4-nitrophenol by NaBH4 in real time using UVvisible absorption spectroscopy. The starshaped morphologies of CuO have been found to show maximum catalytic activity toward the reduction
of 4-nitrophenol in the presence of NaBH4, which can be correlated to their higher specific surface area
and positive surface charge and to the presence of a high indexed facet.
2016 Elsevier Inc. All rights reserved.

1. Introduction
Controlled synthesis of metal oxide nanostructures with various
shapes and sizes is of great importance in the field of nanotechnology, since each of their technological applications entails specific
physicochemical properties, which can only be associated with a
certain morphology. The morphology-modulated properties
acquired by controlling the crystalline structure, usually endowed
with a higher surface-to-volume ratio and crystallographic facet
features, have stimulated researchers to pursue diverse morphologies of metal oxide nanostructures. The size- and morphologymodulated properties are particularly significant in availing their
surface-sensitive applications, such as in the field of heterogeneous
catalysis [1], in the fabrication of photoelectrochemical cells [2]
and batteries [3], in gas sensors [4], and as antibacterial agents [5].
Morphology-dependent superior catalytic performance shown
by certain metal oxide nanostructures [6] is an immense contribution in the industry; thus, their study is of great societal relevance
as well. CuO is one such transition metal oxide system, which is
reported to illustrate significant size- and morphology-dependent
catalytic activity. For example, Zhou et al. reported that the oxidation of CO [7] can be catalyzed by CuO nanobelts and nanoplatelets.
Another widely studied and industrially relevant model reaction
that is used for evaluating the catalytic efficacy of the various CuO
Corresponding author.
E-mail address: ami@chem.iitkgp.ernet.in (A. Pathak).
http://dx.doi.org/10.1016/j.jcat.2015.12.017
0021-9517/ 2016 Elsevier Inc. All rights reserved.

morphologies in aqueous medium is the synthesis of


4-aminophenol (4-AP), an important intermediate in the production
of various analgesic and antipyretic drugs, via hydrogenation of
4-nitrophenol (4-NP) using an excess of sodium borohydride
(NaBH4). It has been well documented that the reduction of
4-NP, which is a toxic chemical typically found in industrial waste
and a contaminant in agricultural waste waters, can be catalyzed
by various metal [8,9] and metal oxide [10,11] nanoparticles in
aqueous medium, and the color change associated with the reduction reaction provides a basis for monitoring the reaction kinetics.
Studying this model reaction, Pande and Pal [12] and Zhou et al.
[13] reported the catalytic behavior shown by nanoboxes and
cellulose-nanocrystal-supported nanospheres of CuO, respectively.
Tamuly et al. [14] have shown that biogenically synthesized spherical and oval CuO nanoparticles, self-assembled to form flowerlike
nanostructures, can catalyze the reduction of several nitro aryl compounds to their respective amino compounds. Mandlimath and
Gopal [15] and Nasrollahzadeh et al. [16] have also reported the
reduction of 4-NP using spherical CuO nanoparticles as catalysts.
Thus, catalytic propertied offered by different morphologies of
CuO nanostructures are reported in the literature. However, there
are only few reports that have attempted to understand the growth
mechanism of the different morphological forms of CuO nanostructures and have also tried to correlate it with their catalytic performance. Although in a recent report, Che et al. [17] have studied
leaf-, flower-, and dumbbell-shaped CuO nanostructures and have
correlated the different shapes with their catalytic performance in

12

S. Konar et al. / Journal of Catalysis 336 (2016) 1122

the reduction of 4-nitrophenol, they have stopped short of exploring


the growth mechanism of different nanostructures in detail.
The method of preparation of metal oxide nanostructures is
known to influence the size and morphology of the final products,
which in turn have a significant impact on their properties. Thus,
the adopted methodology and choice of processing parameters
(such as temperature, starting reagents and their concentration,
and pH) is expected to play an important role in determining the
catalytic activity of CuO nanostructures. Various physical and
chemical synthesis routes such as the hydrothermal method [18],
the pulsed-laser-induced method [19], the arc-discharge method
[20], and the chemical precipitation method [21] have been
reported to respectively yield nanosphere, nanoshuttle, nanorod,
and nanoflower morphologies of CuO nanostructures.
In this work, we have attempted to prepare several nanostructures of CuO through aqueous-based chemical precipitation and
have endeavored to control their morphologies by chelating the
Cu2+ ions with different organic acids (acetic acid, citric acid, and tartaric acid). Attempts have also been made to investigate the catalytic
performance of the different CuO morphologies based on the study
of the model reaction of reduction of 4-nitrophenol (4-NP) by
NaBH4.
2. Materials and methods
2.1. Materials
Copper acetate monohydrate (Cu(CH3COO)2H2O, Merck Ltd,
Mumbai, 98%), glacial acetic acid (Merck Ltd, Mumbai, 99.5%), citric
acid (Merck Ltd, Mumbai, 98%), tartaric acid (Merck Ltd, Mumbai,
99.7%), 4-nitrophenol (4-NP, Merck Ltd, Mumbai, 98%), sodium
borohydride (NaBH4, Merck Ltd, Mumbai, 97%), and sodium
hydroxide (NaOH, Merck Ltd, Mumbai, P97%) were used in this
experiment. All chemicals used were of analytical grade and no
further purification was done prior to use.

Table 1
Summary of the reagents required for the preparation of the various CuO samples.

Sample
name

Copper
acetate
(mol)

Organic
acids added

Amount of
organic acids
(mol)

Amount of
NaOHa (mol)

CuOW
CuOA
CuOC
CuOT

0.1
0.1
0.1
0.1

Acetic acid
Citric acid
Tartaric
acid

0.20
0.20
0.20

0.50
0.58
0.66
0.60

Amount of NaOH needed to maintain the solution pH at 10.

aminophenol (AP) in an excess amount of NaBH4. For the experiment, aqueous stocks of CuO, 4-NP, and NaBH4 were prepared at
concentrations of 2 g/L, 102 (M), and 0.1 (M), respectively. The
stock of NaBH4 was maintained at 4 C to restrain the evolution
of H2 gas. Aliquots of 3.0 mL of 0.1 (M) NaBH4, 20 lL of 102 (M)
4-NP, and 5 lL of CuO were taken from the respective stocks and
mixed together and their absorption spectra were monitored
between 450 and 600 nm at different time intervals. For control,
the study was also carried out in the absence of CuO.
The temperature, in the determination of apparent activation
energies (Ea) of the 4-NP reduction reaction, was varied from 12
to 37 C, while the concentrations of the respective reagents were
kept unchanged. To find the effect of NaBH4 concentration, the
experiment was carried out at various initial concentrations of
NaBH4 (0.040.16 M) while holding [4-NP] = 102 M and [CuOC]
= 2 g/L constant. To study the effect of the concentration of 4-NP
on the reduction reaction, the 4-NP concentration was varied
between 0.4  102 and 1.4  102 M while the initial concentration of NaBH4 and the dose of CuO were kept unchanged. To determine the dose of CuO required for maximum catalytic activity, the
amount of CuO was varied between 1.0 and 2.5 g/L, while the initial concentrations of 4-NP and NaBH4 were kept unaltered. All the
graphs have been plotted based on the average values obtained
from three repeat sets of the experiments.

2.2. Synthesis of CuO nanostructures


In a typical synthesis of CuO, an aliquot amount of aqueous solution of 0.1 (M) copper acetate was taken and the solution pH was
adjusted to 10 by the addition of 1.0 (M) NaOH, while the temperature of the solution mixture was held at 100 C. The solution mixture
was then cooled to room temperature and the black precipitate was
separated through centrifugation. After repeated washing with
water, the residue was dried in a vacuum oven at 80 C for 6 h to
obtain dry powders of CuO (referred as CuOW in the text). This procedure for obtaining CuO was repeated with the addition of glacial
acetic acid (A), citric acid (C), and tartaric acid (T) into three separate
sets of initial aqueous solutions of 0.1 (M) copper acetate, and the pH
of each of the solutions was adjusted to 10 by adding 1.0 (M) NaOH,
while the temperature of the reaction mixtures was maintained at
100 C. The precipitates in each set were separated on cooling and
were washed and dried in a vacuum oven at 80 C for 6 h to obtain
the dry powders of the respective CuO samples. The samples are
referred as CuOA, CuOC, and CuOT in the following text, corresponding to the carboxylic acids used in their preparation. The reagents
used for the preparation of the various CuO samples are summarized
in Table 1. All the prepared samples have been characterized in
detail and the sample preparation has been repeated to check for
consistency in the observed properties.
2.3. Catalytic activity
The morphology-dependent catalytic activity of CuO nanostructures has been investigated through the reduction of 4-NP to 4-

2.4. Characterization of nanostructures


The crystalline phase of CuO in the synthesized samples was
ascertained through powder X-ray diffraction (XRD) studies carried out on a Bruker AXS diffractometer D8, Germany, using CuKa
radiation (k = 1.5418 ) at an applied voltage of 40 kV and at a scan
rate of 3min1 in the 2h range of 2080. The presence of surface
functional groups in the synthesized CuO samples was established
through Fourier-transformed infrared (FTIR) spectroscopy carried
out between 4000 and 400 cm1 using a PerkinElmer Spectrum
RX-II Model 73713, USA. The microstructural features of the synthesized CuO samples were analyzed by transmission electron
microscopy (TEM) using an FEI-TECNAI G2 20STWIN, USA and
high-resolution transmission electron microscope (HRTEM) at an
acceleration voltage of 200 kV using a JEOL JEM-2100, Japan. Zeta
potential measurement and dynamic light scattering (DLS) studies
were undertaken to determine the surface charge and the hydrodynamic diameter of the samples using a Malvern Nano ZS instrument, UK. The isoelectric point (IEP) of CuO was determined
using aqueous dispersions of the samples at a series of pH values
ranging between 3 and 10. The optical and catalytic activity of
the samples was investigated through a Shimadzu-2450 UV
visNIR spectrophotometer, Japan. The band gap energies of the
samples were evaluated from UVvis absorption spectra using
the Tauc plot. The specific surface areas and pore radius of the
degassed (at 200 C for 3 h) CuO samples have been calculated
from the N2 adsorptiondesorption isotherms obtained on a
Quantachrome Chem BET analyzer, USA, at 77 K using the standard

S. Konar et al. / Journal of Catalysis 336 (2016) 1122

BrunauerEmmettTeller (BET) adsorption equation and the Bar


rettJoynerHalenda (BJH) model, respectively.
3. Results and discussion
3.1. Characterization of nanostructures
The XRD peaks of all the prepared samples were found to match
the JCPDS file (80-1916) for the monoclinic phase of CuO (space
group C2/c). The indexed XRD peaks of the CuOW, CuOA, CuOC,
and CuOT samples are shown in Fig. 1A. The absence of any peaks
corresponding to Cu(OH)2, Cu2O, or Cu implied the purity of CuO
phase.
The FTIR spectra of the synthesized CuO samples are shown in
Fig. 1B. In all the samples, the characteristic CuAO stretching band
along the (2 0 2) direction was positioned at 433 cm1, while CuAO
 0 2 direction was assigned to the bands 520
stretching along the 2
and 608 cm1, thus confirming the formation of monoclinic CuO
[22]. The presence of absorbed water in all the samples was affirmed
from the [m(OH)] stretching absorption band at 3392 cm1. In CuOA,
the absorption peaks positioned around 1545 and 1410 cm1 could
respectively be assigned to the antisymmetric and symmetric
stretching frequencies of the carboxylate groups [m(COO)] arising
from the adsorbed acetate anions, which possibly remained as a
bidentate ligand chelated to the Cu2+ center [23]. In CuOC, the broad
peaks at around 1580 and 1375 cm1 may respectively be ascribed
to the asymmetric and symmetric C@O stretching frequencies arising from the adsorbed citrate group, which stayed chelated to Cu2+
center [24]. Two distinct peaks at 1610 and 1365 cm1 in the CuOT
spectrum may be ascribed to the asymmetric and symmetric C@O
stretching modes of the adsorbed tartrate anions. The absence of
any peak between 1680 and 1750 cm1 in the FTIR spectra of the
CuOA, CuOC, and CuOT samples confirmed the absence of any unreacted residues of the respective acids.
The FTIR studies of the thoroughly washed and dried (at 100 C)
samples of CuOA, CuOC, and CuOT therefore indicated the presence
of the respective organic ligands adsorbed onto the surface after
the reaction, which possibly stabilized and passivated the surface
of the CuO nanoparticles. In fact, controlled nucleation and colloidal stability of nanoparticles using organic acids is quite widely
reported for capping the metal/metal oxide surfaces during their
solution-phase synthesis. For example, acetic acid has been used
as a capping agent in the preparation of ZnO [25] nanoparticles.
Similarly, citrate and tartrate anions have also been reported as

13

capping/stabilizing agents in the synthesis of Fe3O4 [26], Cu [27],


noble metals [2830], and MgO [31] nanoparticles.
TEM studies of the CuOW samples (prepared in the absence of
any organic acid) revealed the rod-shaped morphologies of the
particles (Fig. 2a) with lengths ranging between 150 and 180 nm
and diameters between 20 and 30 nm. Their respective HRTEM
images, shown in Fig. 2c, depicted lattice fringe spacing of around
0.275 nm, which was comparable to the lattice spacing along the
(1 1 0) plane in monoclinic crystals of CuO (JCPDS file value,
0.28 nm).
The CuOA samples, prepared in presence of acetic acid in the
starting solution were found to have spherical particles with diameters ranging between 10 and 15 nm (Fig. 2e and f). The lattice
fringe spacing in CuOA samples was measured to be around
0.23 nm, corresponding to the (2 0 0) crystallographic plane of
monoclinic CuO crystal (JCPDS file value, 0.229 nm) (Fig. 2g). In
CuOC samples, prepared in the presence of citric acid, the particles
were observed to be star-shaped with core sizes between 10 and
20 nm and conical tip (pentapod or hexapod) lengths ranging
between 80 and 100 nm (Fig. 2i and j). The lattice fringe spacings
were measured to be 0.23 and 0.252 nm, which in monoclinic
crystals of CuO corresponded to the interplanar spacing along the
 1 1 crystallographic planes (Fig. 2k and l respec(2 0 0) and 1
tively). Furthermore, the particles in CuOT samples, prepared using
tartaric acid as a chelating ligand, were found to have nanoflowershaped morphologies (Fig. 2n and o). The nanoflowers were all
uniform in size with quasi-spherical shapes, consisting of a solid
core of average 6080 nm and short, regular branches of length
between 30 and 50 nm (Fig. 2n and o). From Fig. 2p the branches
appeared to be produced from the core of the nanoflowers. The
measured lattice fringe spacing of 0.23 nm was in reasonable
agreement with the interlayer spacing of the (2 0 0) reflection in
monoclinic CuO crystal.
All the samples showed rings in their SAED patterns (as shown
in Fig. 2d, h, m, and q), which indicated the polycrystalline nature
of sample, where an assembly of randomly distributed and
randomly oriented nanocrystals of CuO diffracted coherently. The
visible rings could be indexed to the monoclinic phase of CuO.
To understand the formation of different morphological forms
of CuO, the process of crystallization can be proposed to occur in
two stages: nucleation and crystal growth. To begin, the solution
contained Cu(OAc)22H2O. The Cu2+ formed water-soluble complexes of Cu(OAc)2/Cu3(cit)2/Cu(tart)2 when acetic/citric/tartaric
acids were respectively added to the reaction mixture. Upon addi-

Fig. 1. (A) XRD and (B) FTIR spectra of (a) CuOW, (b) CuOA, (c) CuOC, and (d) CuOT samples obtained after vacuum-drying of the respective precipitates at 80 C for 6 h.

14

S. Konar et al. / Journal of Catalysis 336 (2016) 1122

Fig. 2. Representative TEM and HRTEM images of CuOW nanorods (a, b), CuOA nanospheres (e, f), CuOC nanostars (i, j), CuOT nanoflowers (n, o), and their respective lattice
fringes (c, g, k, l, and p) and SAED patterns (d, h, m, and q).

tion of excess of NaOH to the solution, the Cu2+ formed the coordinated octahedron complex [Cu(OH)6]4, where the four OH
groups were located in the square plane and the remaining two
OH along the axial position. The two axial OH are easily eliminated via dehydration at the solution temperature of 100 C, due
to lower binding energies compared with those located in the
plane, to form the deep blue, metastable, orthorhombic phase of
Cu(OH)2. The formation of Cu(OH)2, in the presence of the different
organic acids in the reaction medium, can be represented by the
following set of reactions:
Formation of Cu(OH)2 in the absence of any organic acid in the
precursor solution:

CuCH3 COO2  2H2 Oaq 2NaOHaq


! CuOH2s 2CH3 COONaaq 2H2 O:

Formation of Cu(OH)2 in the presence of citric acid in the precursor solution:

3CuCH3 COO2  2H2 Oaq 2H8 C6 O7aq


! Cu3 C6 H5 O7 2aq 6CH3 COOHaq 6H2 O:
Cu3 C6 H5 O7 2aq 6NaOH

aq

! 3CuOH2s
2Na3 C6 H5 O7 aq :

Formation of Cu(OH)2 in the presence of tartaric acid in the precursor solution:

CuCH3 COO2  2H2 Oaq 2H6 C4 O6aq


! CuC4 H4 O6 2aq 2CH3 COOHaq 2H2 O:

CuC4 H4 O6 2aq 2NaOHaq ! CuOH2aq 2C4 H4 O6 Naaq :

The transformation of metastable, orthorhombic Cu(OH)2 to


stable, monoclinic CuO in solution at 100 C can be justified in conformity with the explanation offered by Cudennec and Lecerf [32].
It can be proposed that in excess of NaOH, the metastable Cu(OH)2
gives rise to the oxide by a reconstructive transformation process
involving the dissolution of Cu(OH)2, via the formation of the complex anion [Cu(OH)2
4 ], followed by crystallization of CuO. Thus, the
JahnTeller-effect-stabilized square planar complex anion of Cu
(OH)2
4 can be considered as a precursor entity for the nucleation
of monoclinic CuO. In effect, the anion Cu(OH)2
4 undergoes a condensation phenomenon at the solution temperature of 100 C,
which entails loss of two hydroxyl ions and one water molecule
and leads to the formation of chains of square planar CuO4 groups,
which on complete transformation form nanocrystals of CuO.
These chains of square planar CuO4 groups can thus be rationalized
as the fundamental building blocks for the nucleation of CuO [32]
nanocrystals where, the Cu atom has square planar coordination
with the neighboring oxygen atoms. The complete transformation
of Cu(OH)2 to CuO can be therefore represented by the following
reaction:

CuOH2s 2OHaq ! CuOH4 2
aq CuOs H2 O 2OHaq :

6
It needs to be pointed out that the conversion of Cu(OH)2 to CuO is
so significantly accelerated under excess OH and at a solution
temperature of 100 C that the appearance of Cu(OH)2 during the
reaction may only be momentary or may not appear at all.

15

S. Konar et al. / Journal of Catalysis 336 (2016) 1122

Orthorhombic Cu(OH)2 are known to have a layered structure


with anisotropic growth rates along the energetically favored axis.
Thus, the orthorhombic Cu(OH)2 nanocrystals are not spherical but
display a morphology where one dimension is longer than the
others [3336]. It can be anticipated that the monoclinic CuO
nanocrystals would retain the anisotropic growth characteristic
of orthorhombic Cu(OH)2 when they were formed from complete
transformation of Cu(OH)2 via condensation [33]. This possibly
explains the formation of nanorod-shaped particles in CuOW samples, prepared in the absence of any organic acids in the starting
solution. Similar observation of CuO crystals sustaining the
anisotropic growth attributes of orthorhombic Cu(OH)2 to form
nanowires on complete transformation was reported by Lu et al.
[34] when KOH/NH3 was added to the solution of Cu2+ at 50 C.
Furthermore, the crystal growth and thus the formation of stable
morphologies in anisotropic nanocrystals [37] of noble metals is
well documented to be largely governed by surface energy (i.e.,
excess free energy per unit area) along facets that minimizes the
free energy. Drawing analogies from crystal growth in noble metals, and also as per-facet directed growth, recently reported in
crystallographic studies on free-standing CuO nanocrystals [38],
the growth of the rod-shaped particles in CuOW samples can be
attributed to anisotropic growth along the energetically favored
{0 0 1} facet rather than {1 0 0} and {1 0 1}.
From TEM studies, it is apparent that the morphologies of the
CuO change when different organic acids are added into the precursor solution while all other experimental parameters are left
unaltered. Similar observations have been reported for CuO when
prepared using NaOH and/or NH3 in the presence of various amino
acids [39] and alcohols/nonionic surfactants [40]. To comprehend
the mechanism of crystal growth in the different CuO samples, it
is therefore essential that the discussion be focused on the role
of the various organic acids added during the preparation process.
It can be envisaged that in the presence of acetic/citric/tartaric
acids in the reaction medium, their AOH and ACOOH groups can
serve as coordinating sites that can fix the position of the Cu2+ ions
in the first precipitate (i.e., orthorhombic Cu(OH)2), prior to the formation of CuO nanocrystals. It is possible that the acetate/citrate/
tartrate ions, when present in the medium, will get adsorbed onto
specific crystal planes of the Cu(OH)2 and then change its redissolution/reprecipitation process to gain crystallographic control in
the nucleation of the CuO nanocrystals. It is therefore imperative
that the number of AOH and ACOOH coordinating sites available
in the acetate/citrate/tartrate ions and hence their interaction with
the chains of square planar CuO4 groups, nucleating during the
transformation process, direct the coordination self-assembly and
oriented attachment of the nucleated CuO nanocrystals and thus
decide the final morphology of the CuO nanostructures.
Formation of spherical particles in CuOA samples, prepared on
addition of acetic acid in the precursor solution, can be attributed
to the protective ability of the monodentate acetate group along
the {0 0 1} facet [41]. The shielding ability of the acetate group
would possibly hinder the preferential growth of CuO nanocrystals
along the energetically favored {0 0 1} facet (as discussed for CuOW
samples in reference to freestanding CuO nanocrystals). This would
cause the CuO nanocrystals to grow uniformly along all three
crystallographic directions and result in the spherical morphologies of the particles in CuOA samples.
The nanostar-shaped morphologies of the particles in CuOC
samples can be credited to the capping ability of the three
carboxylic (ACOOH) and one hydroxyl (AOH) group of citric acid
(which was present in the precursor solution). Formation of similar
star-shaped and pentapod-shaped morphologies in Au nanocrystals has been reported by Xiao and Qi [42]. They have attributed
the formation of star-shaped morphologies to surfactant-assisted
growth along the {2 1 1} and {1 1 0} facets, and growth along the

{1 0 0} facet for the hexapods. Based on morphological resemblance, citric acid as a capping agent could be envisaged as promoting controlled growth of the CuOC crystals along {2 1 1} and {1 1 0}
facets and aiding in the formation of star-shaped morphologies.
Likewise, in CuOT samples, the formation of the nanoflowershaped morphologies with quasi-spherical solid cores and short,
regular branches can be ascribed to the capping capability of the
two carboxylic (ACOOH) and two hydroxyl (AOH) groups of
tartaric acid (which was present in the precursor solution). Based
on previous reports on the formation of the CuO nanoflowers
[43], tartaric acid as a capping agent could be presumed to aid crystal growth along the {0 0 1} facet, owing to its high stability, while
aiding formation of the branches through preferential aggregated
growth of the crystals along the {0 1 0} facet because of their lower
stability and higher reactivity [43] compared with {1 0 0} and
{0 0 1} facets.
Thus, the different morphologies of CuO nanostructures can be
understood to be formed in situ through transformation of the
transient deep blue precipitate of Cu(OH)2 via condensation phenomenon while the different organic acids (acetic/citric/tartaric
acids) acted as structure-directing agents (SDA). The observations
from TEM and the predicted growth directions for prepared CuO
samples have been summarized in Table 2.
From DLS studies carried out in water at pH 7, the mean hydrodynamic size of particles in the samples was found to range
between 103 and 460 nm depending on their solvation in water
(Fig. S1, Supplementary Information). The trend of hydrodynamic
particle size was found to follow the order CuOA < CuOC <
CuOW < CuOT (Fig. 3A). Although the trend of the hydrodynamic
particle size from DLS studies was similar to those obtained from
the TEM studies, the average particle sizes obtained from DLS
studies were much higher due to the solvation of the particles in
aqueous medium, which was retained even after ultrasonication
of the respective dispersions.
The surface charge, determined from zeta potential measurements, has been examined using the aqueous dispersion of the
samples under different pH conditions (Fig. 3B). High value of
the zeta potential (either + or ) is indicative of the stability of particles against aggregation or flocculation in the dispersion medium.
The pH at which the aqueous suspensions of CuOW, CuOA, CuOC,
and CuOT showed the zeta potential to be zero (i.e., isoelectric
point (IEP)) was found to be 6.6, 8.0, 9.8, and 9.4, respectively.
Thus, the surfaces of all the samples was positively charged at
pH 7, except for the rod-shaped particles of CuOW.
To determine the optical band gap energies (Eg), UVvisible
absorption studies were carried out using aqueous suspensions
of the samples (Fig. 4). All the samples showed a weak and broad
band 360 nm, possibly due to the excitation of the O2p electron
to the Cu3d level. All the samples, except CuOW, showed another
broad absorption band around 250270 nm, probably arising due
Table 2
Comparison of the morphological observations and band gap values shown by
different nanostructures of CuO.
Sample

Particle
morphology

Diameter
(nm)

Length
(nm)

Growth direction

Eg
(eV)

CuOW

Rod-shaped

2030

Along {0 0 1} facet

2.70

CuOA

Spherical

1015

150
180

3.02

CuOC

Star-shaped

1020

80100

CuOT

Flowershaped

6080

3050

Uniform along all


crystallographic
directions
Controlled growth
along {2 1 1} and
{1 1 0} facets
Flowers along {0 0 1}
and branches along
{0 1 0} facet

2.93

2.96

16

S. Konar et al. / Journal of Catalysis 336 (2016) 1122

Fig. 3. (A) Hydrodynamic particle size for CuOW, CuOA, CuOC, and CuOT samples when dispersed in water at pH 7; (B) their respective surface charge at different pH values.

to n ? p transition in the surface-adsorbed C@O groups. The Eg (in


eV) for CuO nanostructures were calculated using the Tauc
equation,

ahm

1=n

Bhm  Eg ;

where a is the absorption coefficient (in cm1), hm is the incident


photon energy (in eV), B is a constant relative to the materials,
and n 12 (for direct allowed transition) and n = 2 (for indirect transition) [44]. CuO being a wide direct band gap material, the value of
n has been taken as 12 in the Tauc equation. Moreover, in most cases,
indirect transition is not allowed in UVvisible spectra because of
weak absorption of nanomaterials [45]; therefore, the n value is
taken to be 12 (i.e., direct allowed transition). Therefore, ahm is
plotted versus hm for the Tauc plot.
The Tauc plots for the different CuO samples are depicted in
Fig. 5. The average band gap (Eg) values were estimated from the
2

intercept of the linear portion of the ahm versus hm plots on


2

the hm axis (i.e., ahm 0) as shown in Fig. 5. The Eg values for


all the samples were found be higher than that of the bulk CuO
(1.85 eV) [46] (see Table 2). In fact, the Eg values were noted to
increase (i.e., CuOA > CuOT > CuOC > CuOW) with a decrease in
2

the aspect ratio of the particle (i.e., CuOA < CuOT < CuOC < CuOW).
Blue shift in the absorbance peak and hence the increase (i.e., red
shift) in the Eg values of the CuO nanostructures may be attributed
to the quantum size effect [45,47].
Further, the specific surface areas of the degassed (at 200 C for
3 h) samples of the various CuO nanostructures have been calculated from the nitrogen adsorptiondesorption isotherms (Fig. S2,
Supplementary Information) using the BrunauerEmmettTeller
(BET) adsorption equation and the respective values have been tabulated in Table 3. The specific surface area values for CuOC, CuOT,
and CuOA samples were all found to be more than 50.00 m2 g1,
with the highest value (72.74 m2 g1) for the nanostar-shaped particles of CuOC samples and the lowest (23.50 m2 g1) for the
nanorod-shaped particles of CuOW samples. The relatively high
values of BET surface area for CuOA, CuOC, and CuOT samples
may be ascribed to the surface defects owing to the presence of
carboxylic (ACOOH) functional groups adsorbed onto the nanoparticles surface. The average pore radius values have also been
calculated using the BarrettJoynerHalenda (BJH) model and the
respective values have been given in Table 3. The spherical particles of CuOA samples revealed to have the largest average pore
radius value of 7.74 nm, while the values ranged between 1.8 to
2.0 nm for the rest of the CuO samples.

3.2. Catalytic activity

Fig. 4. UVvisible absorption spectra of various CuO samples: 1: CuOW; 2: CuOA;


3: CuOC; 4: CuOT.

The role of the CuO nanostructures as a catalyst has been


validated by studying the model reaction of reduction of 4-NP by
excess NaBH4 using UVvisible spectroscopy [13]. The reduction
of 4-NP to 4-aminophenol (4-AP) by NaBH4 is not a kinetically
favorable process due to large difference in the redox potential of
the donor and acceptor couple (i.e., E{(H3BO3)(aq.)/(BH
4 )(aq.)}
= 1.33 V and E{(4-NP)(aq.)/(4-AP)(aq.)} = 0.76 V versus NHE,
respectively) [48]. In the presence of a catalyst, the kinetic restrictions [49] on the electron transfer from donor BH()
to acceptor
4
4-NP is overcome and the reaction rate is accelerated.
In the present reduction process, the light yellow aqueous solution of 4-NP, with absorbance peak at k = 317 nm, was observed to
change to deep yellow along with a spectral shift of the absorbance
peak to 400 nm as the alkalinity of the solution increased with the
addition of excess of freshly prepared ice cold aqueous solution of
NaBH4. This marked the formation of the 4-nitrophenolate ion in
solution. Next, aliquots of aqueous dispersions of the respective
CuO samples were added into the 4-NP solution and the reduction

17

S. Konar et al. / Journal of Catalysis 336 (2016) 1122

Fig. 5. (ahm)2 vs. hm plots for the different CuO samples: (A) CuOW, (B) CuOA, (C) CuOC, (D) CuOT.

Table 3
Summary of the rate constant, Ea values, and specific surface area and pore size calculated for the different morphologies of CuO nanostructures.
Sample
CuOW
CuOA
CuOC
CuOT
a

Morphology
Nanorods
Spherical NPs
Nanostars
Nanoflowers

Rate constant (k)a (s1)


3

1.20  10
1.60  103
1.95  103
1.30  103

Activation energy (Ea)

Surface area (m2 g1)

Pore radius (nm)

TOF (molecules g1 s1)

26.24 kJ/mol
18.12 kJ/mol
15.36 kJ/mol
21.60 kJ/mol

23.50
58.53
72.74
67.79

1.92
7.74
1.84
1.89

1.25  1018
1.90  1018
2.50  1018
1.79  1018

Measured at 25 C.

was manifested by a decrease in absorbance at k = 400 nm with


fading of the yellow color of the solution and concurrent appearance of a gradually intensifying new peak at k = 295 nm due to
the formation of the reduced product, 4-AP (Fig. 6AD). Since the
absorbance peak of 4-nitrophenolate ions at k = 400 nm was much
stronger than that of 4-NP at k = 315 nm, the progress of the
reaction with time or the kinetics of the reaction was measured
in terms of the concentrations of 4-nitrophenolate ions in the
reaction mixture by monitoring the changes in absorbance at
k = 400 nm at different time intervals (Fig. S3, Supplementary
Information). The reaction between 4-NP and excess NaBH4 has
been taken as the control, since the absorbance at k = 400 nm
was found to remain unaltered with time in the absence of CuO
in the reaction mixture. It may be mentioned that the presence
of CuO nanostructures did not affect the absorbance spectra of
either the 4-nitrophenolate ions or 4-NP because of their very
low concentration [49] (3 ppm) in the system. This establishes

the role of the CuO nanostructures as catalyst in the reduction of


4-NP to 4-AP by excess NaBH4.
For the studied reduction reaction, the concentration of the
reducing agent, NaBH4, has been used in large quantities (50
times) with respect to 4-NP and/or CuO. High concentration,
increased pH, and ice cold condition of the freshly prepared aqueous solution of NaBH4 ensured the stability of the borohydride
anion in the reaction medium by dampening its decomposition/
protonolysis to liberate hydrogen gas. Yet whatever small amount
of hydrogen gas was released assisted in circumventing the aerial
oxidation of 4-AP [49] by providing a reducing atmosphere
in situ. The bubbles of hydrogen gas released during the reaction
also helped in stirring the reaction mixture [48].
A large excess of NaBH4 compared with the concentration of
4-NP also ensured that the concentration of BH()
4 ions in the reaction medium remained invariant during the reaction. In addition,
occurrence of only one isosbestic point in the UVvisible

18

S. Konar et al. / Journal of Catalysis 336 (2016) 1122

Fig. 6. Time-dependent UVvisible absorption spectra of the reduction of 4-NP by excess NaBH4 in the presence of: (A) CuOW; (B) CuOA; (C) CuOC; and (D) CuOT. Condition:
[4-NP] = 102 M, [NaBH4] = 0.1 M, and [Catalyst] = 2 g/L.

absorption spectra of the reduction of 4-NP (as is evident from


Fig. 6) signified that the reaction mixture involved only a pair of
species, 4-NP and 4-AP [50], that varied in concentration and activity of nanocatalysts. From these observations, it may be inferred
that the reduction of 4-NP followed pseudo-first-order kinetics
[49]. Thus, by taking the absorbance of the principal species (i.e.,
4-nitrophenolate ions) to be proportional to their respective concentrations, the rate constant of the reaction, in the presence of
CuOW, CuOA, CuOC, and CuOT, was determined using the equation


 ln

Ct
C0


 ln

At
A0


kt;

where At and A0 are the absorbance (at k = 400 nm) at times t and 0,
respectively, while Ct and C0 are the equivalent concentrations of
4-nitrophenolate ions at times t and 0, respectively. The apparent
rate constants (k) in the presence of each of the CuO nanostructured
samples were determined from the slopes of the respective linear
plots of [ln(At/A0)] versus time, as shown in Fig. 7. The plots shown
in Fig. 7 depict the linearity of [ln(At/A0)] with time; exponential
decay profiles of the absorbance with time support pseudo-firstorder kinetics with respect to 4-nitrophenolate ion.
Fig. 8A shows a typical Arrhenius plot for the catalytic reduction
of 4-NP in presence of CuOC in the temperature (T) range of 12
37 C. The apparent activation energies (Ea), obtained from the

slope (Ea/R) of each of the Arrhenius plots of ln k versus T1,


was found to range between 15 and 26 kJ/mol, which is typical
for surface-catalyzed reactions [48]. Thus, the catalytic reduction
of 4-NP in the presence of synthesized CuO can be deemed a typical surface-catalyzed reaction. The rate constant and Ea values calculated for the different morphologies of CuO are summarized in
Table 4. From the kinetic data, it is apparent that of the four CuO
nanostructures, the star-shaped morphologies of CuOC served as
the most efficient catalyst in the reduction of 4-NP. The catalytic
capacity of the prepared CuO nanostructures has also been compared with those reported in the literature and presented in
Table 5.
The variation of rate constants (k) with concentration of NaBH4
for the CuOC-catalyzed reduction of 4-NP is shown as an example
in Fig. 8B. From the plot, it is evident that the reaction rate
increased as the concentration of NaBH4 was increased from
0.04 M to 0.10 M NaBH4 and saturated at 0.16 M. Consequently,
the concentration of NaBH4 in all experiments was chosen to be
0.1 M; this exceeded the concentration of 4-NP significantly, and
thus made the reaction independent of the concentration of NaBH4.
The effect of the concentration of 4-NP and the quantity of CuO
nanostructures on the reduction rate of 4-NP was also studied by
varying their concentrations individually. Fig. 9A and B show the
increase in reaction rate with increased the concentrations of
4-NP and CuO, respectively. A positive, nonzero intercept observed

19

S. Konar et al. / Journal of Catalysis 336 (2016) 1122

Table 4
Summary of the rate constants and Ea values calculated for the different morphologies
of CuO nanostructures.

Sample

Morphology

Rate constant
(k; s1)a

Activation energy
(Ea; kJ/mol)

CuOW
CuOA
CuOC
CuOT

Nanorods
Spherical NPs
Nanostars
Nanoflowers

1.20  103
1.60  103
1.95  103
1.30  103

26.24
18.12
15.36
21.60

Measured at 25 C.

Number of molecules of substrate


;
Number of grams of catalyst  Time in seconds

Fig. 7. Plots of ln (At/A0) versus time for the reduction of 4-NP by NaBH4 in the
presence of CuOW, CuOA, CuOC, and CuOT nanostructures. Conditions: 3.0 mL of
0.1 M NaBH4, 20 lL of 102 M 4-NP, and 5 lL of 2 g/L nanocatalysts.

in Fig. 9B may be due to a slow uncatalyzed reaction between borohydride and 4-NP. The increase in the conversion of 4-NP with
increased amount of CuO (i.e., the catalyst) may be credited to
the proportional increase in the number of active sites with
increased catalyst amount.
To investigate the reusability of the CuO nanocatalysts, the
intensity of the absorption peak at k = 400 nm was monitored
when fresh 4-NP was again added to the reaction mixture after
completion of one catalytic cycle of reduction of 4-NP, while NaBH4
and CuO nanostructures were excluded from fresh addition. It was
apparent that the reduction of 4-NP again followed the pseudofirst-order kinetics under the maintained reaction conditions. The
recycle ability of the CuO nanostructures was further explored by
increasing the concentration of 4-NP in fresh addition up to five
times, while neither NaBH4 nor CuO nanostructures were added
afresh. The decreasing trend of the absorption peak at 400 nm
was found to be almost similar to that of the first catalytic cycle.
This signified that the synthesized CuO samples could be efficiently
reused as a catalyst for up to five cycles.
The turnover frequencies (TOF) of the different CuO nanostructures have been calculated using the expression

where 1  102 M of 4-NP has been taken as the substrate and 2 g/L
of the nanostructured CuO as the catalyst. The values of TOF for the
various CuO nanostructures are listed in Table 3. Among the four
types of morphologies, CuOC was found to show the highest TOF
value (2.50  1018 molecules g1 s1).
The catalytic performance of CuO nanomaterials appears to be
in the order CuOC > CuOA > CuOT > CuOW, while the specific surface areas follow the order CuOC > CuOT > CuOA > CuOW (as
shown in Table 3). The highest catalytic activity shown by CuOC
and the lowest for CuOW are consistent with the observed requirements. The higher catalytic activity shown by CuOC nanostars
compared with other CuO nanostructures may be attributed to
its highest specific surface area value (72.74 m2 g1), which is
expected to facilitate the adsorption of 4-NP and NaBH4 onto the
nanoparticle surface. In addition, the branched, star-shaped morphologies with sharp tips endowed with high-index facets {2 1 1}
are dominant factors for the higher catalytic activity associated
with the particles of CuOC samples. Highly branched nanocrystals
possess rougher surfaces than their more isotropic counterparts,
which can provide a larger number of active surface reaction sites
for surface-sensitive applications such as catalysis. In addition,
nanostructures with branched arms on the surface are often
endowed with high-index features. High-index facets, represented
by a set of Miller indices {h k l} with at least one index greater than
unity, are considered as a kind of open structures that possess a
high density of atomic steps, kinks, and edges with low coordination numbers, which can act as highly active sites for breaking
chemical bonds. Such observations have been well documented for
noble metals, when prepared using amino acid-based surfactants

Fig. 8. (A) Plot of ln k versus 1/T for the reduction of 4-NP (102 M) by NaBH4 (0.1 M) in the presence of CuOC (2 g/L). (B) Plot of rate constant (k) versus concentration of
NaBH4 for the reduction of 4-NP in the presence of CuOC. Conditions: [4-NP] = 102 M and [CuOC] = 2 g/L.

20

S. Konar et al. / Journal of Catalysis 336 (2016) 1122

Table 5
Comparison of the catalytic performance of the CuO samples with various morphologies reported in the literature for the reduction of 4-NP to 4-AP in excess NaBH4.
Morphology
Spherical [13]
Nanoflowers [14]
Spherical [15]
Spherical [16]
(CuOW): nanorods
(CuOA): spherical
(CuOC): nanostars
(CuOT): banoflowers

Conc. of 4-NP (mM)


2.5
5
5
2.5
2.5
2.5
2.5
2.5

Conc. of NaBH4
0.025 M
0.5 mM
0.05 M
0.25 M
0.1 M
0.1 M
0.1 M
0.1 M

Amount of catalyst (mg/L)


1.33
3
100
10
3.3
3.3
3.3
3.3

Temp. (K)
298
303
303
298
298
298
298
298

Rate constant (s1)


3

6.9  10
1.4  103
1.9  102
4.5  102
4.5  102
6.0  102
7.5  102
4.9  102

Rate constant/[catalyst] (s1 g1 L)


5.19
0.47
0.19
4.50
13.63
18.18
22.72
14.84

The bold value signifies the maximum value of Rate constant/[catalyst] among the particles.

Fig. 9. (A) Plot of rate constant versus concentration of 4-NP in the reduction of 4-NP in excess of NaBH4 and CuOC. Conditions: [NaBH4] = 0.1 M and [CuOC] = 2 g/L. (B) Plot of
rate constant versus amount of CuOC in the reduction of 4-NP. Conditions: [4-NP] = 102 M and [NaBH4] = 0.1 M.

as structure-directing agents, and they have been reported to


exhibit significantly enhanced catalytic activities compared with
those of the close-packed low-index facet {1 0 0} [9,51]. Higher
catalytic activity shown by the CuOC samples can be further linked
to their relatively high positive surface charge under the experimental condition of pH 7. Their higher positive surface charge
enabled them to adsorb larger numbers of 4-nitrophenolate anions
and BH()
ions on their surfaces, which made them more reactive
4
than the other morphologies of CuO nanostructures.

The spherical nanoparticles of CuOA have been found to have


somewhat higher catalytic activity (or reactivity) than the CuOT
nanoflowers, despite CuOT nanoparticles having a slightly higher
specific surface area (67.79 m2 g1) than CuOA nanoparticles
(58.53 m2 g1). This incongruity may be explained on the basis of
pore radius. The pore size of CuOA nanoparticles was found to have
a higher value (7.74 nm) than that of CuOT (1.89 nm). Higher pore
size leads to more adsorption sites and thus higher catalytic activity. Similar observations have been reported by Ghampson et al. in

Scheme 1. The mechanism proposed for 4-nitrophenol reduction in the presence of CuO in excess NaBH4.

S. Konar et al. / Journal of Catalysis 336 (2016) 1122

cobalt nanocatalysts, where TOF and pore size were found to have
linear dependence, whereas TOF and surface area had inverse
dependence [52].
Between CuOW and CuOT, the nanoflower-shaped CuOT
showed more reactivity toward catalytic reduction of 4-NP due
to relative high specific surface area (67.79 m2 g1) and the presence of dual facets {0 1 0} and {0 0 1}, while CuOW showed the lowest surface area (23.50 m2 g1), as well as possessing only one facet
{0 0 1}. Similar observations have also been reported by Narayanan
et al., where near-spherical Pt nanoparticles containing dual facets
showed more catalytic activity than cubic Pt nanoparticles comprising a single facet [53]. Besides, CuOW nanostructures had a
negative surface charge at pH 7, which is expected to give rise to
a lesser number of adsorption sites for the anions and hence reduce
their reactivity. Thus, the catalytic activity (and corresponding
activation energy) of synthesized CuO nanocatalysts was explained
on the basis of four imperative factors(i) effective specific surface
area, (ii) pore radius calculated by BJH method, (iii) indexed facets,
and (iv) surface charge of catalysts at experimental pH.
Based on the inferences drawn from the present study, it may be
proposed that 4-NP (i.e., the electron acceptor) on reaction with
excess NaBH4 (i.e., the electron donor and the only source of H)
results in the formation of 4-aminophenol in solution, where CuO
acted as mediator. Superior surface area, large pore radius, presence
of high-indexed facets, and positively charged surface of CuO at the
experimental pH are the key reasons for it to coadsorb the negatively
charged BH
4 ions from NaBH4 and 4-NP rapidly. After accepting the
hydrogen species via interfacial electron transfer, 4-NP gets reduced
to 4-aminophenol over the CuO surface in a reversible manner. The
present reaction is an example of a bimolecular reaction, and thus
can be explained through a LangmuirHinshelwood mechanistic
model. Similar observations have also been cited by Wunder et al.,
where metallic nanoparticles behaved as the nanocatalysts for the
reduction of 4-NP [54,55]. A schematic representation of the proposed mechanism is depicted in Scheme 1.
4. Conclusions
Ligand-regulated growth of CuO nanostructures with rod-,
sphere-, star-, and flower-shaped morphologies has been brought
about through an aqueous-based chemical precipitation method
where different organic acids (viz. acetic/citric/tartaric acid) acted
as chelating ligands for stabilization of the Cu2+ ions in the solution. The effectiveness of the prepared CuO nanostructures as
catalysts for the reduction of 4-nitrophenol to 4-aminophenol in
excess NaBH4 has been assessed. All prepared morphologies of
CuO nanostructures were found to be effective and stable catalysts
and followed pseudo-first-order kinetics. The reaction was, however, observed to be faster for the nanostar-shaped morphologies
(prepared using citric acid as the chelating agent) with
TOF  2.50  1018 molecules g1 s1; this has been attributed to
their high specific surface area and to the presence of highindexed facets.
Acknowledgments
The authors acknowledge the Central Research Facility (CRF),
IIT Kharagpur for the infrastructural support received for the
characterization of the samples. S. Konar acknowledges the fellowship support received from the Indian Institute of Technology
Kharagpur for carrying out the present research work.
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.jcat.2015.12.017.

21

References
[1] G.R. Chaudhary, P. Bansal, N. Kaur, S.K. Mehta, Recyclable CuO nanoparticles as
heterogeneous catalysts for the synthesis of xanthenes under solvent free
conditions, RSC Adv. 4 (2014) 4946249470.
[2] C.Y. Chiang, Y. Shin, S. Ehrman, Li doped CuO film electrodes for
photoelectrochemical cells, J. Electrochem. Soc. 159 (2012) B227B231.
[3] W.P. Kang, C.H. Zhao, Q. Shen, Lithium storage capability of nanocrystalline
CuO improved by its water-based interactions with sodium alginate, Int. J.
Electrochem. Soc. 7 (2012) 81948204.
[4] T. Ghodselahi, H. Zahrabi, M.H. Saani, M.A. Vesaghi, CO gas sensor properties of
Cu@CuO coreshell nanoparticles based on localized surface plasmon
resonance, J. Phys. Chem. C 115 (2011) 2212622130.
[5] M. Eshed, J. Lellouche, S. Matalon, A. Gedanken, E. Banin, Sonochemical
coatings of ZnO and CuO nanoparticles inhibit streptococcus mutans biofilm
formation on teeth model, Langmuir 28 (2012) 1228812295.
[6] N. Puvvada, P.K. Panigrahi, D. Mandal, A. Pathak, Shape dependent peroxidase
mimetic activity towards oxidation of pyrogallol by H2O2, RSC Adv. 2 (2012)
32703273.
[7] K.B. Zhou, R.P. Wang, B.Q. Xu, Y.D. Li, Synthesis, characterization and catalytic
properties of CuO nanocrystals with various shapes, Nanotechnology 17 (2006)
39393943.
[8] H.K. Kadam, S.G. Tilve, Copper(II) bromide as a procatalyst for in situ
preparation of active Cu nanoparticles for reduction of nitroarenes, RSC Adv.
2 (2012) 60576060.
[9] R. He, Y.C. Wang, X.Y. Wang, Z.T. Wang, G. Liu, W. Zhou, L.P. Wen, Q.X. Li, X.P.
Wang, X.Y. Chen, J. Zeng, J.G. Hou, Facile synthesis of pentacle goldcopper
alloy nanocrystals and their plasmonic and catalytic properties, Nat. Commun.
5 (2014) 4327.
[10] T.K. Ghorai, D. Dhak, A. Azizan, P. Pramanik, Investigation of phase formation
temperature of nano-sized solid solution of copper/cobalt molybdate and
chromium-phosphate ((MxCr1xMoxP1xO4)-Cr1) [M1 = Co, Cu], Mater. Sci.
Eng., B: Solid 121 (2005) 216223.
[11] B.H. Liu, Z.P. Li, A review: hydrogen generation from borohydride hydrolysis
reaction, J. Power Sources 187 (2009) 527534.
[12] S. Pande, T. Pal, Controlled synthesis of CuO box shaped nanoparticles and
their application in nitrophenol reduction, J. Indian Chem. Soc. 87 (2010) 405
415.
[13] Z.H. Zhou, C.H. Lu, X.D. Wu, X.X. Zhang, Cellulose nanocrystals as a novel
support for CuO nanoparticles catalysts: facile synthesis and their application
to 4-nitrophenol reduction, RSC Adv. 3 (2013) 2606626073.
[14] C. Tamuly, I. Saikia, M. Hazarika, M.R. Das, Reduction of aromatic nitro
compounds catalyzed by biogenic CuO nanoparticles, RSC Adv. 4 (2014)
5322953236.
[15] T.R. Mandlimath, B. Gopal, Catalytic activity of first row transition metal
oxides in the conversion of p-nitrophenol to p-anninophenol, J. Mol. Catal. A:
Chem. 350 (2011) 915.
[16] M. Nasrollahzadeh, M. Maham, S.M. Sajadi, Green synthesis of CuO
nanoparticles by aqueous extract of Gundelia tournefortii and evaluation of
their catalytic activity for the synthesis of N-monosubstituted ureas and
reduction of 4-nitrophenol, J. Colloid Interface Sci. 455 (2015) 245253.
[17] W. Che, Y.H. Ni, Y.X. Zhang, Y. Ma, Morphology-controllable synthesis of CuO
nanostructures and their catalytic activity for the reduction of 4-nitrophenol, J.
Phys. Chem. Solids 77 (2015) 17.
[18] E. Reitz, W.Z. Jia, M. Gentile, Y. Wang, Y. Lei, CuO nanospheres based
nonenzymatic glucose sensor, Electroanalysis 20 (2008) 24822486.
[19] X.Y. Chen, H. Cui, P. Liu, G.W. Yang, Shape-induced ultraviolet absorption of
CuO shuttlelike nanoparticles, Appl. Phys. Lett. 90 (2007) 183118.
[20] W.T. Yao, S.H. Yu, Y. Zhou, J. Jiang, Q.S. Wu, L. Zhang, J. Jiang, Formation of
uniform CuO nanorods by spontaneous aggregation: selective synthesis of
CuO, Cu2O, and Cu nanoparticles by a solidliquid phase arc discharge process,
J. Phys. Chem. B 109 (2005) 1401114016.
[21] M. Basu, A.K. Sinha, M. Pradhan, S. Sarkar, A. Pal, T. Pal, Monoclinic CuO
nanoflowers on resin support: recyclable catalyst to obtain perylene
compound, Chem. Commun. 46 (2010) 87858787.
[22] M.A. Dar, Y.S. Kim, W.B. Kim, J.M. Sohn, H.S. Shin, Structural and magnetic
properties of CuO nanoneedles synthesized by hydrothermal method, Appl.
Surf. Sci. 254 (2008) 74777481.
[23] L.F. Liao, C.F. Lien, J.L. Lin, FTIR study of adsorption and photoreactions of acetic
acid on TiO2, Phys. Chem. Chem. Phys. 3 (2001) 38313837.
[24] I.A. Mudunkotuwa, V.H. Grassian, Citric acid adsorption on TiO2 nanoparticles
in aqueous suspensions at acidic and circumneutral pH: surface coverage,
surface speciation, and its impact on nanoparticlenanoparticle interactions, J.
Am. Chem. Soc. 132 (2010) 1498614994.
[25] S.A. Oh, X.W. Sim, S. Tripathy, Laser spectroscopic study of acetate-capped
colloidal ZnO nanoparticles, in: Ninth International Symposium on Laser
Metrology Pts. 1 and 2, vol. 7155, 2008.
[26] E. Cheraghipour, A.M. Tamaddon, S. Javadpour, I.J. Bruce, PEG conjugated
citrate-capped magnetite nanoparticles for biomedical applications, J. Magn.
Magn. Mater. 328 (2013) 9195.
[27] C. Kind, A. Weber, C. Feldmann, Easy access to Cu-0 nanoparticles and porous
copper electrodes with high oxidation stability and high conductivity, J. Mater.
Chem. 22 (2012) 987993.
[28] G. Mpourmpakis, D.G. Vlachos, Growth mechanisms of metal nanoparticles via
first principles, Phys. Rev. Lett. 102 (2009) 155505.

22

S. Konar et al. / Journal of Catalysis 336 (2016) 1122

[29] J. Turkevich, P.C. Stevenson, J. Hillier, A study of the nucleation and growth
processes in the synthesis of colloidal gold, Discuss. Faraday Soc. 11 (1951)
5575.
[30] J.Y. Park, H. Lee, J.R. Renzas, Y.W. Zhang, G.A. Somorjai, Probing hot electron
flow generated on Pt nanoparticles with Au/TiO2 Schottky diodes during
catalytic CO oxidation, Nano Lett. 8 (2008) 23882392.
[31] M.S. Mastuli, N. Kamarulzaman, M.A. Nawawi, A.M. Mahat, R. Rusdi, N.
Kamarudin, Growth mechanisms of MgO nanocrystals via a solgel synthesis
using different complexing agents, Nanoscale Res. Lett. 9 (2014) 19.
[32] Y. Cudennec, A. Lecerf, The transformation of Cu(OH)2 into CuO, revisited, Solid
State Sci. 5 (2003) 14711474.
[33] X.G. Wen, W.X. Zhang, S.H. Yang, Synthesis of Cu(OH)2 and CuO nanoribbon
arrays on a copper surface, Langmuir 19 (2003) 58985903.
[34] C.H. Lu, L.M. Qi, J.H. Yang, D.Y. Zhang, N.Z. Wu, J.M. Ma, Simple template-free
solution route for the controlled synthesis of Cu(OH)2 and CuO nanostructures,
J. Phys. Chem. B 108 (2004) 1782517831.
[35] X.G. Wen, W.X. Zhang, S.H. Yang, Solution phase synthesis of Cu(OH)2
nanoribbons by coordination self-assembly using CU2S nanowires as
precursors, Nano Lett. 2 (2002) 13971401.
[36] X.J. Zhang, G.F. Wang, X.W. Liu, J.J. Wu, M. Li, J. Gu, H. Liu, B. Fang, Different
CuO nanostructures: synthesis, characterization, and applications for glucose
sensors, J. Phys. Chem. C 112 (2008) 1684516849.
[37] P.R. Sajanlal, T.S. Sreeprasad, A.K. Samal, T. Pradeep, Anisotropic
nanomaterials: structure, growth, assembly, and functions, Nano Rev. 2
(2011) 5883.
[38] D.W. Su, X.Q. Xie, S.X. Dou, G.X. Wang, CuO single crystal with exposed {0 0 1}
facets a highly efficient material for gas sensing and Li-ion battery
applications, Sci. Rep. UK 4 (2014) 5753.
[39] M.U.A. Prathap, B. Kaur, R. Srivastava, Hydrothermal synthesis of CuO
micro-/nanostructures and their applications in the oxidative degradation of
methylene blue and non-enzymatic sensing of glucose/H2O2, J. Colloid
Interface Sci. 370 (2012) 144154.
[40] R. Srivastava, M.U.A. Prathap, R. Ore, Morphologically controlled synthesis of
copper oxides and their catalytic applications in the synthesis of
propargylamine and oxidative degradation of methylene blue, Colloids Surf.,
A 392 (2011) 271282.
[41] N. Puvvada, P.K. Panigrahi, A. Pathak, Room temperature synthesis of highly
hemocompatible hydroxyapatite, study of their physical properties and
spectroscopic correlation of particle size, Nanoscale 2 (2010) 26312638.

[42] J.Y. Xiao, L.M. Qi, Surfactant-assisted, shape-controlled synthesis of gold


nanocrystals, Nanoscale 3 (2011) 13831396.
[43] S. Zaman, M.H. Asif, A. Zainelabdin, G. Amin, O. Nur, M. Willander, CuO
nanoflowers as an electrochemical pH sensor and the effect of pH on the
growth, J. Electroanal. Chem. 662 (2011) 421425.
[44] X.J. Zhang, D.G. Zhang, X.M. Ni, H.G. Zheng, Optical and electrochemical
properties of nanosized CuO via thermal decomposition of copper oxalate,
Solid State Electron. 52 (2008) 245248.
[45] S. Rehman, A. Mumtaz, S.K. Hasanain, Size effects on the magnetic and optical
properties of CuO nanoparticles, J. Nanopart. Res. 13 (2011) 24972507.
[46] A. El-Trass, H. ElShamy, I. El-Mehasseb, M. El-Kemary, CuO nanoparticles:
synthesis, characterization, optical properties and interaction with amino
acids, Appl. Surf. Sci. 258 (2012) 29973001.
[47] H. Wang, J.Z. Xu, J.J. Zhu, H.Y. Chen, Preparation of CuO nanoparticles by
microwave irradiation, J. Cryst. Growth 244 (2002) 8894.
[48] S. Saha, A. Pal, S. Kundu, S. Basu, T. Pal, Photochemical green synthesis of
calcium-alginate-stabilized Ag and Au nanoparticles and their catalytic
application to 4-nitrophenol reduction, Langmuir 26 (2010) 28852893.
[49] A. Gangula, R. Podila, M. Ramakrishna, L. Karanam, C. Janardhana, A.M. Rao,
Catalytic reduction of 4-nitrophenol using biogenic gold and silver nanoparticles
derived from breynia rhamnoides, Langmuir 27 (2011) 1526815274.
[50] G.Q. Gao, L. Lin, C.M. Fan, Q. Zhu, R.X. Wang, A.W. Xu, Highly dispersed
platinum nanoparticles generated in viologen micelles with high catalytic
activity and stability, J. Mater. Chem. A 1 (2013) 1220612212.
[51] Z.W. Quan, Y.X. Wang, J.Y. Fang, High-index faceted noble metal nanocrystals
(vol. 46, pg 191, 2013), Acc. Chem. Res. 46 (2013) 1050.
[52] I.T. Ghampson, C. Newman, L. Kong, E. Pier, K.D. Hurley, R.A. Pollock, B.R.
Walsh, B. Goundie, J. Wright, M.C. Wheeler, R.W. Meulenberg, W.J. DeSisto, B.
G. Frederick, R.N. Austin, Effects of pore diameter on particle size, phase, and
turnover frequency in mesoporous silica supported cobalt FischerTropsch
catalysts, Appl. Catal., A: Gen. 388 (2010) 5767.
[53] R. Narayanan, M.A. El-Sayed, Shape-dependent catalytic activity of platinum
nanoparticles in colloidal solution, Nano Lett. 4 (2004) 13431348.
[54] S. Wunder, Y. Lu, M. Albrecht, M. Ballauff, Catalytic activity of faceted gold
nanoparticles studied by a model reaction: evidence for substrate-induced
surface restructuring, ACS Catal. 1 (2011) 908916.
[55] S. Wunder, F. Polzer, Y. Lu, Y. Mei, M. Ballauff, Kinetic analysis of catalytic
reduction of 4-nitrophenol by metallic nanoparticles immobilized in spherical
polyelectrolyte brushes, J. Phys. Chem. C 114 (2010) 88148820.

Das könnte Ihnen auch gefallen