Anda di halaman 1dari 8

ARTICLE

pubs.acs.org/JPCC

Bifunctional AuFe3O4 Heterostructures for Magnetically Recyclable


Catalysis of Nitrophenol Reduction
Fang-hsin Lin and Ruey-an Doong*
Department of Biomedical Engineering and Environmental Sciences, National Tsing Hua University, Hsinchu, 30013, Taiwan

bS Supporting Information
ABSTRACT: The dumbbell- and ower-like AuFe3O4 heterostructures by thermal decomposition of the ironoleate
complex in the presence of Au nanoparticles (NPs) have been
successfully fabricated using dierent sizes of Au NPs as the
seeds for magnetically recyclable catalysis of p-nitrophenol and
2,4-dinitrophenol reduction. The heterostructures exhibit bifunctional properties with high magnetization and excellent catalytic activity toward nitrophenol reduction. The epitaxial linkages
in dumbbell- and ower-like heterostructures are dierent, leading
to the change in magnetic and catalytic properties of the heterostructured nanocatalysts. The pseudo-rst-order rate constants for
nitrophenol reduction are 0.630.72 min1 and 0.380.46 min1
for dumbbell- and ower-like AuFe3O4 heterostructures, respectively. In addition, the heterostructured nanocatalysts show good
separation ability and reusability which can be repeatedly applied
for nearly complete reduction of nitrophenols for at least six successive cycles. The reaction mechanism for nitrophenol reduction by
AuFe3O4 nanocatalysts is also proposed and conrmed by XPS and FTIR analyses. These unique properties make AuFe3O4
heterostructures an ideal platform to study various heterogeneous catalytic processes which can be potentially applied in a wide variety of
elds in purication, catalysis, sensing devices, and green chemistry.

1. INTRODUCTION
Nobel metal nanostructures have recently received much
attention because of their unique optical, catalytic, and electrochemical properties which make them suitable materials for
potential applications in various elds.1,2 Gold nanoparticles
(Au NPs) have been found to play an important role in several
catalytic processes including low-temperature CO oxidation,3
reductive catalysis of chlorinated or nitrogenated hydrocarbons,46
and organic synthesis.7,8 Due to the high cost and limited supply,
however, the improvement of the catalytic eciency and the
reduction of the used amounts are the top priorities for practical
applications. The deposition of Au NPs onto porous supports
such as TiO2, SiO2, and carbon is regarded as a conventional way
to solve the problem by maximizing the loading of catalysts and
to enhance the catalytic activity by well-tuning the surface functionality.912 However, the entrapment or immobilization of the
nanocatalysts on solid supports normally results in a decrease in
the catalytically active surface area and the reactivity of catalytic
species.13 Recently, the doping of Au NPs into the interior of
spherical AgC composites containing Ag NPs has been
synthesized for reduction of 4-nitrophenol in the presence of
sodium borohydride.12 The catalytic activity of bimetallic composites is highly enhanced over the monometallically doped
carbon spheres. However, these catalysts are usually recycled by
tedious and time-consuming centrifugation/redispersion cycles,
r 2011 American Chemical Society

thus hampering the recovery and reusability of catalysts in


aqueous solutions.
The magnetic nanoparticles have recently emerged as viable
alternatives to conventional materials for catalyst supports.14,15
Their insoluble and superparamagnetic natures enable troublefree separation of the nanocatalysts from the reaction mixture
using an external magnet, which eliminates the necessity of
catalyst ltration.16 Ge et al. synthesized a nanostructured
composite with a high specic surface area and magnetic separation ability for 4-nitrophenol reduction.17 A complete conversion
of 4-nitrophenol was obtained within 1 h, and the catalysts were
recycled by an external magnet and reused eight times with almost
identical reaction rate. Deng et al. also fabricated a multicomponent nanostructure composed of a magnetic-silica core, a layer of
gold nanoparticles, and a mesoporous silica shell for both 4-nitrophenol reduction and styrene epoxidation.18 Although these
materials show improved stability and recyclability, the synthesis
procedures are complicated and the magnetic property is usually
hindered by the silica shell. Dumbbell-like AuFe3O4 nanostructures, where one nanoparticle is linked to another, have
been used as the nanocatalysts for CO oxidation as well as H2O2
Received: November 16, 2010
Revised:
February 19, 2011
Published: March 17, 2011
6591

dx.doi.org/10.1021/jp110956k | J. Phys. Chem. C 2011, 115, 65916598

The Journal of Physical Chemistry C


reduction.19,20 The combination of Au NPs with magnetic Fe3O4
NPs can not only provide catalytic activity but also be reclaimed
via magnetic separation after use. Moreover, the magnetic property
was enhanced via interfacial interaction.21 It is believed that
electron transfer across the interface between these two NPs may
lead to a dramatic change in physicochemical properties, thus
oering an ideal platform to study the multifunctionality of
nanomaterials.22,23 In addition, the AuFe3O4 heterostructures
contain both magnetically, optically, and catalytically active NPs,
which show high potential applications to chemical catalysis,
drug delivery, and biomedical imaging.2426
The thermal decomposition of iron precursors including iron
pentacarbonyl, iron acetylacetonate, and the ironoleate complex at high temperature is one of the most ecient techniques to
synthesize monodisperse magnetic NPs.27,28 The size and morphology of Fe3O4 NPs can be tuned by simply adjusting the
reaction temperature ranging from 280 to 380 C.2934 In
addition, the synthesis of heterostructures that contain a noble
metal particle in the structures has recently received considerable
attention.3537 Yu et al. fabricated AuFe3O4 dumbbell structures by thermal decomposition of iron pentacarbonyl in the
presence of preformed Au nanoparticles in 1-octadecene followed by oxidation of iron nanocrystals in air at room temperature.38 The thermal decomposition of mixtures of metaloleate
complexes and metaloleylamine complexes in the presence of
1,2-hexadecanediol has also been reported.39 Although these
strategies produce well-crystallized nanostructures, the synthesis
processes are usually expensive and may contain toxic reagents,
leading to the diculty in practical application. In addition, the
fabrication of dierent morphologies of AuFe3O4 heterostructures by tuning the size of Au NPs has received less attention.
Herein, we demonstrate a facile method for the synthesis of
dierent morphologies of monodisperse AuFe3O4 heterostructures by thermal decomposition of ironoleate complex
(Fe(OL)3) in the presence of Au seeds at 310 C. The designed
heterostructures show excellent dual functions which can not
only undergo rapid catalytic reduction of nitrophenols including
p-nitrophenol and 2,4-dinitrophenol in the presence of NaBH4
but also be easily recycled using an external magnetic eld. To
our best knowledge, this is the rst report demonstrating the use
of AuFe3O4 heterostructures for magnetically recyclable catalysis of nitroaromatic compounds.

2. EXPERIMENTAL SECTION
2.1. Chemicals. Iron chloride (FeCl3 3 6H2O, 98%), oleylamine (>70%), oleic acid (90%), 1-octadecene (90%), and
tert-butylamineborane complex (97%) were purchased from
Sigma-Aldrich. Sodium oleate (95%) was purchased from TCI.
Trisodium citrate dehydrate (>99%) was purchased from Ferak.
Sodium borohydride (95%) and ethanol absolute (99.8%) were
purchased from Riedel-de Haen. p-Nitrophenol and 2,4-dinitrophenol were purchased from Fluka. Hydrogen tetrachloroaurate(III) trihydrate (HAuCl4 3 3H2O) was purchased from Alfa
Aesar. Cyclohexane (99.95%) was purchased from TEDIA. nHexane was purchased from J. T. Baker.
2.2. Synthesis of Gold NPs with Different Sizes. The Au
NPs with sizes of 45 nm were prepared by dissolving 40 mg of
HAuCl4 3 3H2O in a mixture containing 4 mL of oleylamine and
4 mL of cyclohexane in air followed by magnetic stirring at 10 C
under a gentle stream of nitrogen gas. An amount of 0.2 mmol
of tert-butylamineborane complex was dissolved in 0.4 mL of

ARTICLE

oleylamine and 0.4 mL of cyclohexane and then injected into the


precursor solution. The solution color changed to deep red
immediately after injecting the borane complex solution. The
mixture was aged for 40 min at 10 C followed by addition of
30 mL of ethanol to precipitate the Au NPs. The Au NPs were
then harvested by centrifugation and redispersed in hexane. For
preparation of 10 nm Au NPs, 40 mg of HAuCl4 3 3H2O was
dissolving in a mixture containing 4 mL of 1-octadecene and
4 mL of oleylamine in air. The resulting solution was put in an oil
bath at 120 C and reaction for 30 min under N2 atmosphere.
After reaction, the mixture was cooled to room temperature and
followed by addition of 30 mL of ethanol to precipitate the Au
NPs. The product was centrifuged and redispersed in hexane.
2.3. Synthesis of AuFe3O4 Heterostructures. A solution
containing 0.5 mmol of oleic acid, 0.5 mmol of oleylamine,
1 mmol of Fe(OL)3, 0.1 mmol of gold colloid dispersion, and
5 mL of octadecene was heated to 110 C for 20 min. The
solution was refluxed at 310 C for 30 min. After cooling to room
temperature, the particles were separated by adding absolute
ethanol, centrifugation, and redispersion into hexane.
2.4. Preparation of Water-Soluble AuFe3O4 NPs. The
AuFe3O4 NPs were washed with a mixture of hexane and
ethanol (1:2) several times to remove excess capping agent on
the surface of NPs. The heterostructured AuFe3O4 NPs were
then dried and added into an aqueous solution containing
50 mM sodium citrate. After reaction of 10 min, the AuFe3O4
heterostructures were separated by a magnet and washed with
deionized water three times. The particles were then dissolved in
deionized water.
2.5. Catalytic Reaction. The reduction of nitrophenol compound by water-soluble AuFe3O4 NPs in the presence of
NaBH4 was carried out to examine the catalytic activity and
recyclability of the AuFe3O4 nanocatalysts. Amounts of 2 mL
of deionized water, 40 L of 10 mM nitrophenol, and 0.16 mL
of 0.1 M NaBH4 solutions were added into a quartz cuvette
followed by addition of 2 mg of water-soluble AuFe3O4 NPs to
the mixture. The color of the solution changed gradually from
yellow to transparent as the reaction proceeded. UVvis spectrometry was used to record the change in absorbance at a time
interval of 2 min.
2.6. Surface Characterization. Transmission electron microscopy (TEM) images were obtained on a JEOL 2011 microscope
operated at 120 kV. High-resolution transmission electron
microscopy (HR-TEM) was carried out on a JEOL JEM-2010
microscope at 200 kV. The samples were prepared by suspension
in hexane. Wide-angle XRD patterns were recorded on a Bruker
D8 X-ray diffractometer with Ni-filtered Cu KR radiation ( =
1.5406 ) and operated at a generator voltage and an emission
current of 40 kV and 40 mA, respectively. A Hitachi U-3010
UVvis spectrometer using a 1 cm path length quartz cuvette
was used to identify the change in concentration over a wavelength range from 200 to 600 nm. Magnetic measurements were
carried out using a superconducting quantum interference device
magnetometer (SQUID MPMS5, Quantum Design Inc.) with a
maximum applied continuous field of 10 000 G at room temperature. X-ray photoelectron spectroscopy (XPS) measurements were performed by an ESCA PHI 1600 photoelectron
spectrometer using an Al KR X-ray source (1486.6 eV photon
energy). During data acquisition, the pressure in the sample
chamber was maintained below 2.5  108 Torr. The binding
energies of the photoelectrons were determined under the
assumption that Au has a binding energy of 84.0 eV. FTIR
6592

dx.doi.org/10.1021/jp110956k |J. Phys. Chem. C 2011, 115, 65916598

The Journal of Physical Chemistry C

ARTICLE

Figure 1. TEM images of (a) 5 nm Au NPs, (b) dumbbell-like AuFe3O4 heterostructures, (d) 10 nm Au NPs, and (e) ower-like AuFe3O4
heterostructures. Figures (c) and (f) are HRTEM images of dumbbell- and ower-like AuFe3O4 heterostructures, repsectively.

Figure 2. Histogram analysis of particle sizes of (a) dumbbell-like and (b) ower-like AuFe3O4 heterostructures.

spectra were obtained by a Horiba FT-720 spectrometer with


KBr method.

3. RESULTS AND DISCUSSION


3.1. Characterization of AuFe3O4 Heterostructures. The
heterostructured AuFe3O4 nanoparticles were prepared by
thermal decomposition of the ironoleate complex in the
presence of different sizes of Au NPs. The morphology of these
heterostructures is highly dependent on the size of Au seeds.
Figure 1 shows the TEM and HR-TEM images of AuFe3O4
heterostructures synthesized by using different sizes of Au seeds
ranging between 5 and 10 nm (Figure 1a, d). The nanostructured
nanoparticles show a dumbbell-like structure when small-sized
Au NPs are used as seeds (Figure 1b). The epitaxial relationship
between Au and Fe3O4 nanoparticles was further examined by
HR-TEM. The interfringe distances are measured to be 0.24 nm
for Au nanoparticles and 0.24 nm for Fe3O4 nanoparticles, which
correspond to the (111) plane of face-centered cubic (fcc) Au
and (311) plane of inverse spinel structured magnetite, respectively (Figure 1c). In addition, the line-scan analysis was used to
get information on relative locations of Au and Fe3O4 in the
heterostructures. As depicted in Figure S1 (Supporting Information), different distribution patterns of Au and Fe are observed. The Fe signals mainly locate at 1426 and 3644 nm, while

Au signals appear at 2633 and 4449 nm, which confirm that


the Au and Fe3O4 NPs in dumbbell-like heterostructures are in
an epitaxial relationship. The histogram analysis shows that the
Au and Fe3O4 NPs in dumbbell-like structures are in the range
2.85.8 and 1115 nm with mean sizes of 5 and 12 nm, respectively (Figure 2a). The sizes of dumbbell-like nanostructures
are also in the range of 1216 nm. Using large Au NPs of
713 nm as seeds, flower-like structures with sizes of 2028 nm
are formed (Figure 2b). A previous study depicted that the
crystallinity of Au seeds controlled the nucleation process, with
one iron oxide leaf nucleated per monocrystalline domain of
gold.40 In this study, large Au NPs provide large surface areas
and multiple monocrystalline domains for Fe3O4 to nucleation,
resulting in the production of flower-like heterostructures. In
addition, the d-spacings of 0.24 and 0.48 nm for Au and magnetite NPs, respectively, are observed, clearly showing the growth
of the Fe3O4(111) plane onto a Au(111) plane to form a flowerlike heterostructure. In addition, parallelogram-like AuFe3O4
heterostructures were obtained when the particle size of Au seeds
increased to 20 nm (Supporting Information, Figure S2), clearly
indicating that the morphology of AuFe3O4 heterostructures is
highly dependent on the size of Au seeds.
The crystallinity of AuFe3O4 heterostructures is characterized by XRD. Figure 3a shows the XRD patterns of dierent
morphologies of AuFe3O4 NPs. Five resolved peaks at 30.10,
6593

dx.doi.org/10.1021/jp110956k |J. Phys. Chem. C 2011, 115, 65916598

The Journal of Physical Chemistry C

ARTICLE

Figure 3. (a) XRD patterns and (b) magnetic hysteresis loops of dumbbell- and ower-like AuFe3O4 heterostructures.

Figure 4. Time-dependent UVvis spectral changes in (a) p-nitrophenol (4-NP) and (b) 2,4-dinitrophenol (2,4-DNP) catalyzed by AuFe3O4
heterostructures and concentration change in nitrophenol compounds (Ct/C0) in the presence of (c) dumbbell-like and (d) ower-like AuFe3O4
nanocrystals. Insets in Figures (c) and (d) are linear relationship of ln(Ct/C0) as a function of time for 4-NP and 2,4-DNP, respectively.

35.54, 43.09, 56.98, and 62.58 2, which can be assigned as


the fcc Fe3O4, are observed. In addition, peaks at 38.18, 44.39,
64.58, and 77.55 2 are common patterns for fcc-structured
Au. The XRD patterns of AuFe3O4 NPs match well with those
corresponding JCPDS standards of Au and Fe3O4 (JCPDS 040784; JCPDS 65-3107, clearly indicating the nature of heterodimer structures of AuFe3O4 NPs. In addition, the epitaxial
linkage in heterostructures has a signicant eect on the change
in optical properties of Au NPs.37 The pure Au NPs show a
surface plasmon resonance peak at 517 and 520 nm for 5 and
10 nm Au NPs, respectively. After conjugation with Fe3O4 NPs,
the peak is broadening and red-shifts to 567 nm in dumbbell-like
structure and 550 nm in ower-like structures (Supporting
Information, Figure S3). The relatively weak reectance of
AuFe3O4 NPs is primarily attributed to the dilution eect of
Fe3O4 on Au NPs in the heterostructures.9 Moreover, the magnetic measurement shows that AuFe3O4 NPs are superparamagnetic at room temperature (300 K) (Figure 3b). The
hysteresis loops of AuFe3O4 NPs indicate that the saturation
magnetization is 31 emu/g for dumbbell-like structures and 43
emu/g for ower-like structures at 300 K. After normalization to
the unit weight of Fe3O4, the saturation magnetizations are 41

and 51 emu/g-Fe3O4 for dumbbell- and ower-like AuFe3O4


nanostructures, respectively. It is noteworthy that saturation
magnetization obtained in this study is lower than that of bulk
magnetite (90 emu/g).41 However, these values are higher than
those reported data prepared by the similar procedure after
normalization to the unit weight of Fe3O4.28,42
3.2. Application of AuFe3O4 Heterostructures for Catalytic Reduction of Nitrophenols. The catalytic reduction of
p-nitrophenol to their corresponding daughter derivatives, paminophenol, in the presence of NaBH4 was chosen as a model
reaction to investigate the bifunctionality of AuFe3O4 heterostructures. Such a reaction catalyzed by Au catalysts has been
reported because this reaction can be rapidly and easily characterized.4345 In addition, 2,4-dinitrophenol was also selected as
another model compound for elucidating the reaction kinetics as
well as a mechanism for nitrophenol reduction. Figure 4 shows
the typical UVvis spectra and concentration change of nitrophenol compounds in the presence of different morphologies of
AuFe3O4 heterostructures and NaBH4. The original absorption peak of p-nitrophenol is centered at 317 nm and shifts to
400 nm after addition of freshly prepared NaBH4 solution,
indicating the formation of p-nitrophenolate ions (Supporting
6594

dx.doi.org/10.1021/jp110956k |J. Phys. Chem. C 2011, 115, 65916598

The Journal of Physical Chemistry C

ARTICLE

Figure 5. Catalytically recyclable reduction of (a) p-nitrophenol and (b) 2,4-nitrophenol by dumbbell-like AuFe3O4 NPs in the presence of NaBH4.
Conversion eciency of (c) p-nitrophenol in six successive cycles of reduction and (d) 2,4-nitrophenol in seven successive cycles of reduction by
AuFe3O4 and citrate-stabilized Au nanocatalysts.

Information, Figure S4).46 This peak starts to decrease when the


reduction proceeds in the presence of AuFe3O4 nanocatalysts.
Addition of NaBH4 in the absence of AuFe3O4 NPs has little
effect on the change in absorbance at 400 nm, confirming that the
reduction is mainly catalyzed by the AuFe3O4 NPs. In addition,
the absorption peak at 400 nm decreases with the concomitant
increase in peak intensity at 300 nm within 10 min after addition
of AuFe3O4 catalysts (Figure 4a).
A similar reduction pattern for 2,4-dinitrophenol is also
observed in which the peak at 357 nm decreases with the increase
in absorption at 450 nm (Figure 4b), clearly indicating the production of intermediate of 2-amino-4-nitrophenol.47 The concentration of 2-amino-4-nitrophenol at 450 nm decreases again
with the concurrent increase in peak intensity of p-aminophenol
at 300 nm. This means that 2,4-dinitrophenol undergoes the
consecutive reaction to form 2-amino-4-nitrophenol and then to
p-aminophenol in the presence of AuFe3O4 nanocatalysts and
NaBH4.
The pseudorst-order kinetics can be applied to evaluate the
rate constants for nitrophenol reduction because the concentration of NaBH4 is higher than those of nitrophenols and can be
considered as a constant during the reaction period. The concentration of p-nitrophenol and 2,4-dinitrophenol at time t is
denoted as Ct, and the initial concentration of nitrophenols at t =
0 is regarded as C0. The Ct/C0 is measured from the relative
intensity of absorbance (At/A0). The linear relationship of
ln(Ct/C0) versus time (t) indicates that the reduction of nitrophenols by AuFe3O4 heterostructures follows the pseudorstorder kinetics. The rate constants for nitrophenol reduction are
0.63 min1 for p-nitrophenol and 0.72 min1 for 2,4-dinitrophenol by using dumbbell-like AuFe3O4 nanocatalysts (Figure 4c).
In addition, the ower-like AuFe3O4 NPs are also used as
nanocatalysts for reduction of nitrophenols. The rate constants
for p-nitrophenol and 2,4-dinitrophenol reduction catalyzed by
ower-like AuFe3O4 nanocatalysts are 0.38 and 0.46 min1,
respectively (Figure 4d). The catalytic eciency as well as the rate

constants for nitrophenol reduction by both dumbbell- and owerlike AuFe3O4 are higher than those previously reported values
obtained from the catalysis of p-nitrophenol by Au-based materials
(Supporting Information, Table S1).48,49 This result clearly indicates
that AuFe3O4 heterostructures are superior nanocatalysts
which can enhance the catalytic eciency and minimize the used
amounts of catalysts for reaction, especially only when trace
amounts of Au catalysts are used (0.380.96 mg Au) for reduction. It is noteworthy that the catalytic eciency of ower-like
AuFe3O4 is lower than that of dumbbell-like structures
(Supporting Information, Figure S5), presumably due to that
the Au surfaces in ower-like structures are mainly occupied
by the Fe3O4 leaves and thus suppress the reaction rate of
nitrophenols. Therefore, the dumbbell-like AuFe3O4 heterostructures are selected as the model nanocatalysts for further
experiments.
The as-prepared AuFe3O4 heterostructures show both
catalytic and magnetic properties which can be easily recycled
by an external magnet after the catalytic reduction. Figure 5 shows
the magnetically recyclable reduction of nitrophenols in the
presence of dumbbell-like AuFe3O4 nanocatalysts. The catalysts can be successfully recycled and reused for at least six
successive cycles of reaction with stable conversion eciency of
around 100%. The hydrodynamic size of AuFe3O4 nanoparticles remains unchanged after several cycles of catalytic reactions
(Supporting Information, Figure S6), indicating the stability of
the dumbbell-like nanoparticles in aqueous solutions. In addition, the citrate-stabilized Au NPs were used as the catalysts to
reduce nitrophenol compounds for comparison. The conversion
eciency of p-nitrophenol and 2,4-dinitrophenol by citratestabilized Au NPs drops dramatically after the second cycle,
which is primarily attributed to the loss of Au NPs after periodic
centrifugation/redispersion cycles. The average particle sizes of
citrate-stabilized Au NPs before and after the centrifugation,
determined by TEM images, are 11.2 ( 0.8 and 11.7 ( 1.8 nm,
respectively, clearly indicating that the decrease in conversion
6595

dx.doi.org/10.1021/jp110956k |J. Phys. Chem. C 2011, 115, 65916598

The Journal of Physical Chemistry C

ARTICLE

Scheme 1. Possible Mechanism for Magnetically Recyclable Catalysis of Nitrophenols by AuFe3O4 Heterostructures

eciency of nitrophenols by Au NP is primarily attributed to the


loss of Au NPs after periodic centrifugation/redispersion cycles.
In addition, the same Au seeds used for synthesis of AuFe3O4
heterostructures were also employed for reduction of p-nitrophenol. The single-component Au NP catalysts were obtained by
etching Fe3O4 away from the AuFe3O4 NPs in 0.5 M H2SO4
solutions.10 The catalytic eciency of p-nitrophenol by Au seeds
is low when compared with that by dumbbell-like AuFe3O4
heterostructures, presumably due to the aggregation of Au seeds
during the etching step (Supporting Information, Figure S7).
When adding sodium citrate to the solution containing p-nitrophenol and NaBH4, little p-nitrophenol was reduced, suggesting
that citrate has no eect on nitrophenol reduction. These results
demonstrate that the AuFe3O4 NPs are superior catalysts than
Au itself and other supported Au catalysts, presumably attributed
to the electronic junction eect of Au and Fe3O4 NPs.2022 This
electronic junction eect can also be observed in reduction
catalysis of H2O2 by AuFe3O4 dumbbell-like structure.20 In
addition, the Au NPs in dumbbell-like structure were stable
against aggregation during harvest procedure, resulting in the
enhanced catalysis of nitrophenol compounds. It is obvious that
the presence of Fe3O4 NPs makes the dumbbell-like heterostructures a promising bifunctional probe for magnetically recyclable catalytic reduction. When the reduction is complete, the
AuFe3O4 nanocatalysts can be separated easily and rapidly
from the solution within 10 s by a magnet and then be redispersed into deionized water for the next cycle of catalysis (Supporting Information, Figure S8).
Although the AuFe3O4 heterostructures show superior catalytic and recycling eciencies, the rate constant for nitrophenol
reduction decreased when AuFe3O4 NPs were reused (Supporting Information, Figure S9). The decrease in rate constants for nitrophenol reduction may probably be attributed to
the generation of aminophenols after catalytic reduction and
then bound to the surface of Au NPs. Scheme 1 shows the
catalytic mechanisms for nitrophenol reduction by AuFe3O4

in the presence of NaBH4. When AuFe3O4 NPs are used for


catalytic reduction, BH4 and nitrophenols (p-nitrophenol
and 2,4-dinitrophnol) are rst diused from aqueous solution
to the Au surface, and then the bare Au NPs on heterostructures serve as catalysts to transfer electrons from BH4 to
nitrophenols, leading to the production of amino derivatives,
2-amino-4-nitrophenol and p-aminophenol.50 It is noteworthy that the amine (NH2) group in aminophenols has
a strong binding ability with Au NPs and, therefore, adsorbs
onto the surface of Au NPs, resulting in the block of reactive
sites on Au NPs. To verify the hypothesis of surface blocking
by NH2 groups, aniline is used to pretreat the AuFe3O4
nanocatalysts prior to the reduction of p-nitrophenol. The rate
constants for p-nitrophenol reduction by aniline pretreated
AuFe3O4 nanocatalysts decrease dramatically after the second cycle (Supporting Information, Figure S10), which is
similar to the results shown in Figure S9 (Supporting Information). In addition, XPS and FTIR are used to characterize
the change in surface species on AuFe3O4 heterostructures
before and after the reduction (Figure 6). The XPS of N1s
spectra show no peak before the reaction, while one predominant peak appears at 400 eV after the catalytic reduction,
which is consistent with the result of p-aminophenol, and can
be assigned as the amine (NH2) group after peak deconvolution (Figure 6a).51 This result clearly indicates the chemisorption of aminophenol onto the surface of Au NPs. In
addition, all the FTIR spectra exhibit symmetric and asymmetric stretching vibrations of the CH3 bond at 2840 and
2950 cm1, respectively (Figure 6b). The FeO bonding in
the range 570600 cm1 is also clearly observed. The
AuFe3O4 heterostructures in the organic phase show weak
CdO and CN stretching peaks at 1704 and 1439 cm1,
respectively, indicating the presence of oleylamine and oleic
acid on the surface of nanoparticles. After ligand exchange
with sodium citrate in aqueous solution, the peak of CN
stretching disappears, while the CdO peak at 1613 cm1 is
6596

dx.doi.org/10.1021/jp110956k |J. Phys. Chem. C 2011, 115, 65916598

The Journal of Physical Chemistry C

ARTICLE

Figure 6. (a) XPS N1s spectra of of dumbbell-like AuFe3O4 NPs before and after catalytic reaction. (b) FT-IR spectra of dumbbell-like AuFe3O4
NPs capped with dierent ligands.

observed. After the catalytic reduction of nitrophenols, the


CN stretch in AuFe3O4 reappears at 1421 cm1, conrming the attachment of aminophenols onto the catalyst surfaces.

4. CONCLUSIONS
In this study, we have rst demonstrated that the heterostructured AuFe3O4 nanocatalysts synthesized via thermal
decomposition of the ironoleate complex in the presence of
Au seeds have excellent bifunctional characteristics for reusability
and catalytic reduction. The size and morphology of AuFe3O4
nanocatalysts are highly dependent on the size of Au seeds. The
catalytic performance of both dumbbell- and ower-like Au
Fe3O4 NPs is excellent for nitrophenol reduction in the presence
of NaBH4. In addition, the catalytic eciency of dumbbell-like
AuFe3O4 NPs is higher than that of ower-like heterostructures because of the high surface coverage of the Au surface by
Fe3O4 nanocrystals in ower-like heterostructures. The dumbbell-like nanoparticles also show good separability and reusability
in successive cycles of reduction. The reaction mechanism of
successive reduction of nitrophenols by AuFe3O4 heterostructures has been proposed and conrmed. Results obtained in this
study open an avenue to the fabrication of highly ecient heterodimer nanocatalysts for serving as an ideal platform to study the
various heterogeneous catalytic processes.
ASSOCIATED CONTENT

bS

Supporting Information. Line-scan analysis of AuFe3O4;


TEM image of parallelogram-like heterostructures; UVvis spectra
of AuFe3O4 and p-nitrophenol; concentration change of nitrophenol compounds with time; hydrodynamic size of AuFe3O4 NPs;
pictures of magnetic separation of nanocatalysts; pseudorst-order
rate constants for nitrophenol reduction as a function of recycling
times; and pseudorst-order rate constants for nitrophenol reduction
by various Au catalysts. This material is available free of charge via the
Internet at http://pubs.acs.org.

AUTHOR INFORMATION
Corresponding Author

*E-mail: radoong@mx.nthu.edu.tw. Phone number: 886-35726785. Fax number: 886-3-5718649.

ACKNOWLEDGMENT
The authors thank the National Science Council, Taiwan, for
nancial support under Contract No. NSC 99-2113-M-007-007MY3. The authors thank Prof. Hong-Ping Lin at National ChengKung University, Tainan, for help with the HR-TEM analysis.
REFERENCES
(1) Sau, T. K.; Rogach, A. L.; Jackel, F.; Klar, T. A.; Feldmann, J. Adv.
Mater. 2010, 22, 1805.
(2) Jain, P. K.; Huang, X. H.; El-Sayed, I. H.; El-Sayed, M. A. Acc.
Chem. Res. 2008, 41, 1578.
(3) Haruta, M.; Yamada, N.; Kobayashi, T.; Iijima, S. J. Catal. 1989,
115, 301.
(4) Orlov, A.; Jeerson, D. A.; Macleod, N.; Lambert, R. M. Catal.
Lett. 2004, 92, 41.
(5) S. Praharaj, S.; Nath, S.; Ghosh, S. K.; Kundu, S.; Pal, T. Langmuir
2004, 20, 9889.
(6) Zeng, J.; Zhang, Q.; Chen, J.; Xia, Y. Nano Lett. 2010, 10, 30.
(7) Gong, J. L.; Mullins, C. B. Acc. Chem. Res. 2009, 42, 1063.
(8) Hutchings, G. J. Top. Catal. 2008, 48, 55.
(9) Xu, C.; Xie, J.; Ho, D.; Wang, C.; Kohler, N.; Walsh, E. G.;
Morgan, J. R.; Chin, Y. E.; Sun, S. Angew. Chem., Int. Ed. 2008, 47, 173.
(10) Chen, M. S.; Goodman, D. W. Science 2004, 306, 252.
(11) Comotti, M.; Li, W. C.; Splietho, B.; Schuth, F. J. Am. Chem.
Soc. 2006, 128, 917.
(12) Tang, S. C.; Vongehr, S.; Meng, X. K. J. Mater. Chem. 2010, 20, 5436.
(13) Lim, C. W.; Lee, I. S. Nano Today 2010, 5, 412.
(14) Shylesh, S.; Schunemann, V.; Thiel, W. R. Angew. Chem., Int. Ed.
2010, 49, 3428.
(15) Zhu, Y.; Stubbs, L. P.; Ho, F.; Liu, R.; Ship, C. P.; Maguire, J. A.;
Hosmane, N. S. ChemCatChem 2010, 2, 365.
(16) Polshettiwar, V.; Varma, R. S. Green Chem. 2010, 12, 743.
(17) Ge, J. P.; Huynh, T.; Hu, Y. P.; Yin, Y. D. Nano Lett. 2008,
8, 931.
(18) Deng, Y. H.; Cai, Y.; Sun, Z. K.; Zhao, D. Y. J. Am. Chem. Soc.
2010, 132, 8466.
(19) Wang, C.; Yin, H. F.; Dai, S.; Sun, S. H. Chem. Mater. 2010,
22, 3277.
(20) Lee, Y. M.; Garcia, M. A.; Huls, N. A. F.; Sun, S. H. Angew.
Chem., Int. Ed. 2010, 49, 1271.
(21) Lopes, G.; Vargas, J. M.; Sharma, S. K.; Beron, F.; Pirota, K. R.;
Knobel, M.; Rettori, C.; Zysler, R. D. J. Phys. Chem. C 2010, 114, 10148.
(22) Costi, R.; Saunders, A. E.; Banin, U. Angew. Chem., Int. Ed. 2010,
49, 4878.
6597

dx.doi.org/10.1021/jp110956k |J. Phys. Chem. C 2011, 115, 65916598

The Journal of Physical Chemistry C

ARTICLE

(23) Frey, N. A.; Phan, M. H.; Srikanth, H.; Srinath, S.; Wang, C.;
Sun, S. J. Appl. Phys. 2009, 105, 07B502.
(24) Xu, C. J.; Wang, B. D.; Sun, S. H. J. Am. Chem. Soc. 2009,
131, 4216.
(25) Jiang, J.; Gu, H. W.; Shao, H. L.; Devlin, E.; Papaefthymiou,
G. C.; Ying, J. Y. Adv. Mater. 2008, 20, 4403.
(26) Choi, J. S.; Jun, Y. W.; Yeon, S. I.; Kim, H. C.; Shin, J. S.; Cheon,
J. J. Am. Chem. Soc. 2006, 128, 15982.
(27) Sun, S. H.; Zeng, H. J. Am. Chem. Soc. 2002, 124, 8204.
(28) Park, J.; Lee, E.; Hwang, N. M.; Kang, M.; Kim, S. C.; Hwang,
Y.; Park, J. G.; Noh, H. J.; Kim, J. Y.; Park, J. H.; Hyeon, T. H. Angew.
Chem., Int. Ed. 2005, 44, 2872.
(29) Kovalenko, M. V.; Bodnarchuk, M. I.; Lechner, R. T.; Hesser,
G.; Schaer, F.; Heiss, W. J. Am. Chem. Soc. 2007, 129, 6352.
(30) Shavel, A.; Rodruez-Gonzaez, B.; Pacico, J.; Spasova, M.;
Farle, M.; Liz-Marza, L. M. Chem. Mater. 2009, 21, 1326.
(31) Zeng, H.; Rice, P. M.; Wang, S. X.; Sun, S. J. Am. Chem. Soc.
2004, 126, 11458.
(32) Sun, S.; Zeng, H.; Robinson, D. B.; Raoux, S.; Rice, P. M.;
Wang, S. X.; Li, G. J. Am. Chem. Soc. 2004, 126, 273.
(33) Jana, N. R.; Chen, Y.; Peng, X. Chem. Mater. 2004, 16, 3931.
(34) Cheon, J.; Kang, N. J.; Lee, S. M.; Lee, J. H.; Yoon, J. H.; Oh, S. J.
J. Am. Chem. Soc. 2004, 126, 1950.
(35) Gu, H.; Yang, Z.; Gao, J.; Chang, C. K.; Xu, B. J. Am. Chem. Soc.
2005, 127, 34.
(36) Jiang, J.; Gu, H.; Shao, H.; Devlin, E.; Papaefthymiou, G. C.;
Ying, J. Y. Adv. Mater. 2008, 20, 4403.
(37) Wang, C.; Xu, C.; Zeng, H.; Sun, S. Adv. Mater. 2009, 21, 1.
(38) Yu, H.; Chen, M.; Rice, P. M.; Wang, S. X.; White, R. L.; Sun, S.
Nano Lett. 2005, 5, 379.
(39) Choi, S. H.; Na, H. B.; Park, Y. I.; An, K.; Kwon, S. G.; Jang, Y.;
Park, M.; Moon, J.; Son, J. S.; Song, I. C.; Moon, W. K.; Hyeon, T. J. Am.
Chem. Soc. 2008, 130, 15573.
(40) Wei, Y.; Klajn, R.; Pinchuk, A. O.; Grzybowski, B. A. Small
2008, 4, 1635.
(41) Cornell, M. R.; Schwertmann, U. The Iron Oxides; VCH: New
York, 1996; p 117.
(42) Zhen, G.; Muir, B. W.; Moat, B. A.; Harbour, P.; Murray, K. S.;
Moubaraki, B.; Suzuki, K.; Madsen, I.; Agron-Olshina, N.; Waddington,
L.; Mulvaney, P.; Hartley, P. G. J. Phys. Chem. C 2011, 115, 327.
(43) Wu, Y. P.; Zhang, T.; Zheng, Z. H.; Ding, X. B.; Peng, Y. X.
Mater. Res. Bull. 2010, 45, 513.
(44) Dotzauer, D. M.; Dai, J. H.; Sun, L.; Bruening, M. L. Nano Lett.
2006, 6, 2268.
(45) Huang, J. F.; Vongehr, S.; Tang, S. C.; Lu, H. M.; Shen, J. C.;
Meng, X. K. Langmuir 2009, 25, 11890.
(46) Pradhan, N.; Pal, A.; Pal, T. Langmuir 2001, 17, 1800.
(47) Wessels, J. S. C. Biochim. Biophys. Acta 1965, 109, 357.
(48) Hayakawa, K.; Yoshimura, T.; Esumi, K. Langmuir 2003, 19, 5517.
(49) Rashid, M. H.; Bhattacharjee, R. R.; Kotal, A. T.; Mandal, K.
Langmuir 2006, 22, 7141.
(50) Huang, J.; Vongehr, S.; Tang, S.; Lu, H.; Shen, J.; Meng, X.
Langmuir 2009, 25, 11890.
(51) DAmours, M.; Beanger, D. J. Phys. Chem. B 2003, 107
48114817.

6598

dx.doi.org/10.1021/jp110956k |J. Phys. Chem. C 2011, 115, 65916598

Anda mungkin juga menyukai