Synthesis of Porous Fe O Nanospheres and Its Application For The Catalytic Degradation of Xylenol Orange

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

ARTICLE

pubs.acs.org/JPCC

Synthesis of Porous Fe3O4 Nanospheres and Its Application for the


Catalytic Degradation of Xylenol Orange
Maiyong Zhu and Guowang Diao*
College of Chemistry and Chemical Engineering, Yangzhou University, Yangzhou, 225002, P. R. China
bS Supporting Information
ABSTRACT: Porous magnetite (Fe3O4) nanospheres composed
of primary nanocrystals have been successfully synthesized by
solvothermal method with FeCl3 3 6H2O serving as the single iron
resource, polyvinylpyrrolidone (PVP) as the capping agent, and
sodium acetate as the precipitation agent. To understand the
formation mechanism of the porous Fe3O4 nanospheres, the
reaction conditions such as the concentration of the precursor,
capping agent, precipitation agent, the reaction temperature, and
reaction time were investigated. The characterization of the as-
prepared product was identified with transmission electronic micro-
scopy (TEM), field emission scanning electronic microscopy (FE-SEM), X-ray diffraction (XRD), Raman spectroscopy, X-ray
photoelectron spectroscopy (XPS), vibrating sample magnetometer (VSM), N2 adsorptiondesorption technique, and Fourier
transform infrared spectroscopy (FTIR). The results indicate that the porous Fe3O4 nanospheres display excellent magnetic properties at
room temperature, which allows them to be easily separated from the reaction system with the help of external magnet when they serve as
catalysts. Catalytic activity studies show that the as-prepared porous Fe3O4 nanospheres are highly effective catalysts for the degradation
of xylenol orange (XO) in aqueous solution with H2O2 as oxidant. The degradation reaction is first-order, its rate constant at room
temperature being 0.056 min1. Furthermore, the catalytic activity of Fe3O4 nanospheres decreases very slightly after seven cycles of the
catalysis experiment. Therefore, porous Fe3O4 nanospheres can serve as effective recyclable catalysts for the degradation of XO.

1. INTRODUCTION temperature, which can be a great challenge for many groups.


During the past few decades, magnetic nanoparticles with Microemulsions can also be used to synthesize nanoparticles
peculiar size and structure have attracted wide attention because with specific morphology and uniform size. Unfortunately, this
of their unique physical and chemical properties.15 Many method consumes a large amount of organic solvent; besides,
strategies have been developed to get a variety of magnetic organic surfactants are needed to control the size and shape of
nanomaterials, such as iron, cobalt, nickel, alloys, iron oxide, the product.39 Compared with other options, solvothermal is
metal ferrite, and so on.68 Of all magnetic materials, magnetite often characterized by low yield, yet with this method almost all
(Fe3O4) is the most important and most widely used one in many materials can be solubilized by heating and pressurizing the
fields. When prepared into well-defined nanoscale structures, solvent system to its critical point, allowing the synthesis of high-
Fe3O4 may be practically or potentially applied in a wide range quality nanostructured materials.40
of fields, such as biomolecular separation,9 chemical sensor,10 Hydrothermal/solvothermal has become one of the most
energy storage,11 catalysis,1214 microwave absorption,15 biome- popular methods to get inorganic nanostructured materials,
dicine/biotechnology,1623 and environmental remediation.2429 especially metals and their oxides, despite low yield.41 Higher
Generally speaking, Fe3O4 nanoparticles could be prepared with pressure and higher temperature (sometimes even above critical
different methods including coprecipitation of Fe2+ and Fe3+, point of the solvent) occurred in this method and can increase
thermal decomposition, microemulsion, and hydrothermal/ the solubility of solids and accelerate reactions between solid
solvothermal.3035 Of all these methods, coprecipitation is a species. Regarding magnetic nanomaterials, many research
preferred choice for most researchers because the reaction could groups have employed solvothermal method. Gao et al. prepared
be performed under a mild condition by using water as solvent. water-soluble magnetite nanocrystals with different sizes, that is,
However, the shape of the outcome product is hard to control.7,36 4, 12, and 60 nm, by refluxing 2-pyrrolidone solution of FeCl3.42 Li
Thermal decomposition is considered to be the best route in and his coworkers reported a generalized hydrothermal method
obtaining nanomaterials with controllable size and morphology. for synthesizing various nanocrystals by liquidsolidsolution
However, the precursors of the product are organometallic com-
pounds, such as Fe(CO)5, metal cupferronates,37 and metal Received: January 14, 2011
acetylacetonates,38 which are poisonous. In addition, thermal Revised: August 24, 2011
decomposition of organometallic compound often requires high Published: August 26, 2011

r 2011 American Chemical Society 18923 dx.doi.org/10.1021/jp200418j | J. Phys. Chem. C 2011, 115, 18923–18934
The Journal of Physical Chemistry C ARTICLE

reaction.43,44 Pinna et al. reported nonaqueous solvothermal water used in this study was deionized by milli-Q Plus system
synthesis of monocrystalline magnetite particles with sizes ranging (Millipore, France), whose electrical resistance is 18.2 MΩ.
from 12 to 25 nm.45 Huang et al. synthesized different kinds of 2.2. Syntheses of Porous Fe3O4 Nanospheres. Porous
functionalized magnetic microspheres for different bioapplications Fe3O4 nanospheres were prepared with the solvothermal. We
by hydrothermal reduction. The typical procedure is as follows: added 1.5 g FeCl3 3 6H2O, 1 g PVP, and 2 g NaAc to 30 mL of
mixture consisting of FeCl3, ethylene glycol, sodium acetate, and a ethylene glycol. The mixture was stirred vigorously for 2 h to
polymer, such as dextran, chitosan, or poly(acrylic acid) (PAA), ensure all materials dissolve completely. Then, the mixture was
was stirred vigorously to form a clear solution, sealed in a Teflon- transferred to a Teflon-lined stainless-steel autoclave and sealed
lined stainless-steel autoclave, and heated to and maintained at for heating at 200 °C for 8 h. The precipitated black products
200 °C for 830 h.46 In Yin’s research group, Fe3O4 colloidal were collected from the solution with an external magnet and
nanocrystal clusters (CNCs), with uniform sizes ranging between washed with ethanol for a few times. Finally, the black products
∼30 and ∼180 nm, have been successfully synthesized with were dried in a vacuum at 60 °C for 24 h. The sizes of the as-
diethylene glycol solvent. These CNC structures were designed prepared porous Fe3O4 nanospheres could be controlled by
with two-stage growth process in which primary nanocrystals changing the concentration of FeCl3 3 6H2O, PVP, and NaAc, or
nucleate first and then aggregate uniformly into larger secondary the reaction temperature.
structures.47,48 2.3. Characterization. Transmission electron microscopy
Although a variety of iron oxide nanocrystals with different (TEM: Tecnai-12, HRTEM: Tecnai-G2 F30 S-TWIN, Philips)
morphologies, such as nanorods,4953 fractal,54 pea-pod-like,55 and field emission scanning electron microscopy (FE-SEM;
dumbbell-like,56 cubes,57 wires,58 tubes,59 spindles,60 flakes,61 Hitachi S-4800) were used to observe the morphologies of
flower-shaped,27 and hollow spheres,62,63 have been successfully the product. X-ray powder diffraction (XRD) was performed
synthesized, to the best of the authors’ knowledge, there are few on a D8 Advance (Super speed) X-ray diffractometer (Bruker).
reports on the synthesis of porous Fe3O4 nanoparticles.52,53,64 Raman spectra were taken on a Renishaw InVia spectrometer at
Compared with the solid counterparts of the same size, porous room temperature with an argon-ion laser at an excitation
nanoparticles have their particular merits, such as high surface wavelength of 532 nm. The X-ray photoelectron spectroscopy
area, less resistance of mass transfer in catalytic system, con- (XPS) experiments were carried out on a Thermo Escalab 250
trollable pore size, and adjustable framework. system using Al Kα radiation (hν = 1486.6 eV). The test chamber
In this Article, an effective route for the synthesizing of pressure was maintained below 2  109 Torr during spectral
monodispersed porous Fe3O4 nanospheres could be adjusted acquisition. The Fourier transform infrared (FT-IR) spectra were
by controlling the reaction conditions, although solvothermal recorded with a Tensor 27 spectrometer (Bruker) using a KBr
method was reported. The average diameter of the porous Fe3O4 wafer with the wavenumber ranging 4000400 cm1. Magnetic
nanospheres was ∼250 nm. In this route, ethylene glycol properties of the products were investigated using a vibrating
solution of FeCl3 3 6H2O containing PVP and NaAc was heated sample magnetometer (VSM, EV7, ADE) with an applied field
to produce porous Fe3O4 nanospheres, where ethylene glycol between 8000 and 8000 Oe at room temperature. The surface
acted as both solvent and reducing agent, NaAc and PVP as the area and pore size distribution were measured by N2 adsorp-
precipitation and capping agents, respectively. In general, many tiondesorption technique in an automated surface area and
reports indicate that template methods can be used to get hollow porosity analyzer (ASAP 2020, Micromeritics) at 77 K after the
and porous nanostructures. However, in doing so, the presynthe- sample was dried at 200 °C for 4 h. The BrunauerEmmett
sized templates and the removal of them without destruction Teller (BET) surface area and the pore size distribution plots
of the product are required, which make the procedures were calculated by applying the linear part of the BET plot and
complex.65,66 Compared with that method, the one reported in the BarrettJoynerHalenda (BJH) model, respectively.
this study is more convenient. The effects of temperature, time, 2.4. Catalytic Activity of Porous Fe3O4 Nanospheres. To
as well as the concentration of FeCl3 3 6H2O, NaAc, and PVP on investigate the catalytic activity of the as-prepared porous Fe3O4
the crystal phase of the product and its final structure were nanospheres, the degradation of XO with H2O2 oxidant under
investigated. On the basis of the investigation, a possible me- ultrasound were chosen as a test reaction model. We mixed 10 mg
chanism for the formation of porous Fe3O4 nanospheres was porous Fe3O4 nanospheres, 20 mL of 1  104 M XO mother
suggested. The as-prepared porous Fe3O4 nanospheres show solution, and 10 mL of 30% H2O2 aqueous solution were mixed
excellent magnetic behavior at room temperature, which makes it together. The concentration of XO at different reaction time was
suitable to work as a recyclable heterogeneous catalyst because determined by UV spectroscopy (Shimadazu UV-2500 spectro-
they can be separated easily from the reaction system by an photometer). The content of total organic carbon (TOC) was
external magnet. The degradation of xylenol orange, using H2O2 analyzed by using a multi N/C 2100 analyzer (Analytik Jena AG).
as the oxidation reagent, was chosen as a model reaction to
investigate its catalytic activity. The results indicate that the 3. RESULTS AND DISCUSSION
porous Fe3O4 nanospheres synthesized in this Article exhibit
high catalytic activity for the degradation of xylenol orange by 3.1. Morphology, Structure, and Properties of Porous
H2O2. The porous Fe3O4 nanospheres could be separated Fe3O4 Nanospheres. The size and morphology of Fe3O4 nano-
conveniently from the reaction system with an external magnet spheres were characterized via TEM and FE-SEM. The TEM
field and can be used in recycling. images of magnetic Fe3O4 nanospheres in different magnifica-
tions are shown in Figure 1A,B. The diameter of those Fe3O4
2. EXPERIMENTAL SECTION nanospheres is ∼250 nm. A typical HRTEM image of a Fe3O4
nanosphere is depicted in Figure 1C, where the polycrystalline
2.1. Chemicals. All chemicals were analytical grade and nature of the product can be confirmed by electron-diffraction
purchased from Sinopharm Chemical Reagents Company. The pattern (see the inset). From Figure 1D, the lattice spacing
18924 dx.doi.org/10.1021/jp200418j |J. Phys. Chem. C 2011, 115, 18923–18934
The Journal of Physical Chemistry C ARTICLE

Figure 1. (A,B) TEM images of Fe3O4 nanospheres with different magnifications, (C) representative HRTEM image and the electron-diffraction
pattern (inset) of a Fe3O4 nanosphere, (D) HRTEM image of the boxed region of part C, and (E,F) FE-SEM images with different magnifications of
Fe3O4 nanospheres. The synthesis conditions: 1.5 g FeCl3 3 6H2O, 1 g PVP, 2 g NaAc, 30 mL ethylene glycol, 200 °C, and 8 h.

distance is calculated to be 0.29 nm, in good agreement with the It is widely accepted that the XRD pattern of Fe3O4 nano-
(220) spacing of the Fe3O4 structure. Figure 1E,F shows the FE- structures is very similar to that of γ-Fe2O3.67 To confirm further
SEM images of the nanospheres, which clearly reveal that each the crystal phase of our product, laser micro-raman spectra were
magnetic microsphere consists of many magnetic grains. Un- recorded. Figure 2B shows the Raman spectra of the as-prepared
fortunately, the size distribution of them is very wide. porous Fe3O4 nanospheres and the commercial Fe3O4 powder.
Figure 2A shows the XRD patterns of the porous Fe3O4 In both curves, the positions of main peaks are almost identical.
nanospheres. The diffraction peaks locating at 2θ = 30.1, 35.8, No peaks at either 1430 or 1580 cm1, typical for γ-Fe2O3,46
43.1, 53.8, 57.3, and 63.0° can be indexed to (220), (311), (400), were observed, further confirming that there is no γ-Fe2O3 in the
(422), (511), and (440) planes of Fe3O4 in a face-centered as-prepared porous Fe3O4 nanospheres. XPS as in Figure 2CE
cubic (fcc) Fe3O4 (JCPDS card no. 19-629), respectively. The also comes to the same conclusion. Figure 2C shows XPS of the
broadened peak centered at a small angle (2θ < 30°) indicates porous Fe3O4 nanospheres and the inset the expanded spectra of
the presence of amorphous materials. The crystal size deter- N1s. Figure 2 D shows the high-resolution XPS spectra of Fe2p,
mined by Debye-Scherre equation from XRD is 11.8 nm, and Figure 2E shows the high-resolution XPS spectra of O1s. It is
which further confirms that the porous Fe3O4 nanospheres clear that in the as-prepared porous Fe3O4 nanospheres, some
were composed of the primary grains with a diameter of PVP exists as N 1s and C 1s are observed from Figure 2C. Two
ca. 12 nm. distinct peaks at 712.3 and 726.2 eV appear in Figure 2D. The
18925 dx.doi.org/10.1021/jp200418j |J. Phys. Chem. C 2011, 115, 18923–18934
The Journal of Physical Chemistry C ARTICLE

Figure 2. (A) XRD patterns of Fe3O4, (B) Raman spectra of Fe3O4 obtained by this work (a) and commercial (b), (C) XPS of Fe3O4 and the inset is
expanded spectra of N1s, (D) high-resolution XPS spectra of Fe2p, (E) high-resolution XPS spectra of O1s, (F) magnetic hysteresis and enlarged partial
(inset) curves of Fe3O4, (G) typical BET isotherm of Fe3O4, and (H) BJH isotherm and enlarged partial (inset) curves of the Fe3O4. All Fe3O4 are
prepared under same conditions as those described in Figure 1.

former is attributed to Fe2p3/2, and the latter to Fe2p1/2. In that the porous Fe3O4 nanospheres display weak ferromagnet-
Figure 2E, The O1s centered at 529.9 eV belongs to O2. The ism. The experimental phenomena can be explained by the fact
above results confirm the formation of Fe3O4 in the system. that many primary Fe3O4 particles assemble together to form
It is well known that the magnetic properties of the material larger Fe3O4 nanoparticles as in Figure 1, making the porous
are related to its size, structure, and morphology. Figure 2F is the Fe3O4 nanospheres look like a block magnet and exhibit higher
hysteresis loop of the as-prepared porous Fe3O4 nanospheres saturation magnetization than individual primary particles. Nor-
measured at room temperature. It can be found that the mally, when the diameter is smaller than 20 nm, the super-
magnetization saturation value (Ms) of the Fe3O4 nanospheres paramagnetic property of Fe3O4 can be observed.69 However,
is 79.8 emu/g, which is close to that of the bulk sample of Fe3O4, the magnetic property of the porous Fe3O4 nanospheres con-
92 emu/g,68 and the coercivity value (Hc) is 35.5 Oe, indicating sisting of small Fe3O4 grains is different from that of the
18926 dx.doi.org/10.1021/jp200418j |J. Phys. Chem. C 2011, 115, 18923–18934
The Journal of Physical Chemistry C ARTICLE

Figure 3. TEM images of Fe3O4 prepared with different initial concentrations of FeCl3 3 6H2O: (a) 0.25, (b) 0.5, (c) 0.75, (d) 1.0, (e) 1.25, and (f) 1.5 g.
Other conditions: 0.5 g PVP, 1.0 g NaAc, 30 mL of ethylene glycol, 200 °C, and reaction time: 8 h.

Table 1. Magnetic Parameters of Fe3O4 Nanoparticles Pre-


pared at Different Amount of FeCl3 3 6H2Oa
sample a b c d e f

FeCl3 3 6H2O (g) 0.25 0.5 0.75 1.0 1.25 1.5


Ms (emu/g) 77.8 80.4 82.1 82.2 83.2 87.9
Hc (Oe) 68.3 85.6 72.3 72.3 85.6 73.0
a
Other conditions: 0.5g PVP, 1.0g NaAc, 30 mL of ethylene glycol,
200 °C, reaction time: 8 h.

Figure 4. Magnetic hysteresis curves and enlarged partial curves (inset) of crystal, and self-assembly of the primary crystals to form larger
Fe3O4 prepared under the same conditions as those described in Figure 3. nanoshperes. The initial concentration of precursor is a most
important factor that affects the formation of nanocrystals.
Figure 3 shows the TEM of Fe3O4 prepared under different
individual particles. This interesting phenomenon may be related initial concentrations of FeCl3 under the given conditions. As is
to the interaction among the small particles, which is subject to displayed in Figure 3 A, no regular porous Fe3O4 nanosphere was
further study. observed at low concentration of FeCl3. The concentration of
The surface area and porosity of Fe3O4 nanospheres were FeCl3 determines the regularity and the morphology of the
determined by measuring the adsorption and desorption iso- product. The higher the concentration of FeCl3, the more regular
therms of N2. As is shown Figure 2G, the BET surface area of the morphology of porous Fe3O4 nanospheres. The diameter of
Fe3O4 nanospheres was calculated to be 47.7 m2/g. Figure 2H Fe3O4 nanospheres increases with the concentration of FeCl3.
displays the distribution of the pores in Fe3O4 nanospheres. A However, the particles are polydispersed except for those pre-
main peak centered at ∼4 nm. Most of the pores range from 3 to pared, whereas the initial concentration of FeCl3 is >0.0417
10 nm. The result shows that Fe3O4 nanospheres are a type of g 3 mL1, where uniform, regular, and monodispered Fe3O4
mesoporous materials. Such porous structure provides efficient nanospheres will be obtained, whose diameter of the nanopar-
transport pathways to the interior cavities, which would enhance ticles is ∼250 nm. The above phenomenon can be explained by
the catalytic activity of catalyst.70 the theory of Von Weimarn rules.71 At the low concentration
3.2. Mechanism for the Formation of Porous Fe3O4 Nano- of FeCl3, although the rates of both nucleation and growth are
spheres. In our reaction system, the formation of Fe3O4 nano- slow, plenty of precursors can be nucleated with the help of
spheres includes nucleation, growth of nucleus to form primary precipitation agent. Therefore, there is not enough FeCl3 for the
18927 dx.doi.org/10.1021/jp200418j |J. Phys. Chem. C 2011, 115, 18923–18934
The Journal of Physical Chemistry C ARTICLE

Figure 5. TEM images of Fe3O4 obtained at different temperature of (A) 200, (B) 220, (C) 230, (D) 235, (E) 240, and (F) 250 °C. Other conditions:
1.5 g FeCl3 3 6H2O, 1 g PVP, 2 g NaAc, 30 mL of ethylene glycol, and reaction time: 8 h.

self-assembly step in the system, and thus no porous Fe3O4 NaAc is widely used as the precipitation agent in the preparing
nanospheres are observed in Figure 3A. When the concentration nanostructured materials of metals, alloys, and metal oxides.
of precursors slightly increases, some of them could not be Figure 6 shows Fe3O4 prepared in the mixture of 1.5 g FeCl3 3
nucleated at the initial period for the limitation of precipitation 6H2O, 1 g PVP, and 30 mL of ethylene glycol at 220 °C for 8 h
agent. Therefore, some porous Fe3O4 nanospheres can be with different amount of NaAc. The results show that in the
obtained (Figure 3BD). As for further increasing the precursor above reaction system NaAc is necessary for the formation of
concentration, the rate of nucleation will not increase. However, Fe3O4 particles. Moreover, in our systems, when no NaAc was
the growth and self-assembly steps will take longer, so large-sized used, Fe3O4 could not be formed. When the concentration of
porous Fe3O4 nanospheres can form in the system as in NaAc in the system is low, porous nanospheres with small size
Figure 3E,F. The magnetic properties of all samples were shown could be obtained, as in Figure 6A. It is observed in Figure 6B,C
in Figure 4. It is observed that all samples exhibit weak ferro- that a higher concentration of NaAc brings about larger porous
magnetic behavior at room temperature. Their Ms values increase Fe3O4 nanospheres. However, too high concentration of NaAc
with the concentration of FeCl3, which might be related to size. will lead to worse uniformity of the prepared Fe3O4 nanospheres.
In another word, the larger the size of Fe3O4 nanospheres, the The reason might be that when the concentration of NaAc is low,
higher the value of Ms. Table 1 lists the values of the main the nucleation rate of Fe3O4 nanocrystals remains slow and even,
magnetic parameters (Ms and Hc) of the as-prepared samples. but the self-assembly step for the crystals has no limitation. The
Temperature is another important parameter in the formation. higher concentration is a different case. It is well known that the
Figure 5 shows TEM images of Fe3O4 prepared at 200, 220, 230, nucleation rate is much faster with higher concentration of
235, 240, and 250 °C. Figure 5AD displays the fact that precipitation agent, which may in turn prevent part of crystals
temperature has a strong impact on size and morphology. The from assembly because there is not enough space for them to
higher the temperature is, the smaller the particles are. Below grow. As a result, the uniformity of the obtained porous Fe3O4
235 °C, Fe3O4 nanospheres formed of primary particles can be nanospheres was not good when the concentration of NaAc is
obtained. However, when the temperature is >235 °C, the mor- too high, seen in Figure 6 D.
phologies of Fe3O4 changed significantly. From Figure 5E,F, the The impact of capping agent PVP on the morphology of
shapes of Fe3O4 obtained at 240 and 250 °C are not regular, the size Fe3O4 nanospheres was also studied. The TEM images of Fe3O4
was <50 nm, and the separate sphere-shaped particles and polygon- nanospheres prepared in 1.5 g FeCl3.6H2O, 2 g NaAc, and 30 mL
shaped grains occur, compared with those obtained between 200 of ethylene glycol at 220 °C for 8 h with different amount of PVP
and 235 °C. As for the size, the product at 250 °C is smaller than that were shown in Figure 7. The role of PVP was found to be very
at 240 °C. This reason might be that the capping agent (PVP) could critical. When no PVP was used, the product was particle in
not be effectively adsorbed on the surface of Fe3O4 nanoparticles at morphology without identified size or structure, as is shown in
higher temperature, stopping the process of self-assembly. Figure 7A. When 0.5 g PVP was added, porous Fe3O4
18928 dx.doi.org/10.1021/jp200418j |J. Phys. Chem. C 2011, 115, 18923–18934
The Journal of Physical Chemistry C ARTICLE

Figure 6. TEM images of Fe3O4 obtained at different amount of NaAc: (A) 0.5, (B) 1, (C) 1.5, and (D) 2.0 g. Other conditions: 1.5 g FeCl3 3 6H2O,
1 g PVP, 30 mL of ethylene glycol at 220 °C, and reaction time: 8 h.

Figure 7. TEM images of Fe3O4 obtained at different amount of PVP: (A) 0.00, (B) 0.5, (C) 1, (D) 1.5, and (E) 2.0 g. Other conditions: 1.5 g
FeCl3 3 6H2O, 2 g NaAc, 30 mL of ethylene glycol at 220 °C, and reaction time: 8 h.

18929 dx.doi.org/10.1021/jp200418j |J. Phys. Chem. C 2011, 115, 18923–18934


The Journal of Physical Chemistry C ARTICLE

nanospheres were obtained. However, the shape of the product was Therefore, interestingly, the uniformity of porous Fe3O4 nano-
not very regular, seen in Figure 7 B. With the increase in PVP, the spheres was the function of the amount of PVP. When the PVP is
size of the prepared nanospheres decreased slightly, and the shape 1.5 g, the uniformity of the product was most satisfying. Therefore,
of the product became more regular, as is shown in Figure 7BE. it could be believed that the surfactant PVP played a role in the
system. On one hand, when no or low-concentrated PVP was
applied, there lacked the capping effect for the effective coverage or
passivation of Fe3O4, generating particles with no identified size
and structure. On the other hand, the capping effect of PVP with
high concentration was so strong that some of the nanocrystals
were prevented from growing, leading to irregular nanospheres.
To understand the interaction between the capping agent and
Fe3O4 nanocrystals, we conducted FT-IR spectra. Figure 8
presents the spectra of pure capping agent (curve a) and the
as-prepared porous Fe3O4 nanospheres (curve b). It is noted that
the spectrum of the obtained porous Fe3O4 nanospheres is
somewhat similar to that of the pure capping agent. The
absorption peaks centered around 29003000 cm1 in curve a
are attributed to CH stretching, and the absorption peak at
1637.4 cm1 can be assigned to the CdO stretching. The strong
Figure 8. FTIR spectra of PVP (a) and PVP-stabilized Fe3O4 nano- absorption peak at 584.4 cm1 in Figure 8 b corresponds to the
spheres (b). FeO vibrations. The deformation vibration absorption at
1427.2 cm1 for CH in PVP moved to 1425.3 cm1 in the
Scheme 1. Model of the Interaction between Fe3O4 Nano- porous Fe3O4 nanospheres. Meanwhile, compared with the pure
crystals and PVP PVP, the stretching vibration of CdO in Fe3O4 moved from
1637.4 to 1625.9 cm1. The shifts of the CdO and CH of PVP
vibration are due to the fact that in the system this capping agent
coordinates with the Fe3O4 nanocrystals. The possible interac-
tion between the obtained Fe3O4 nanocrystals and PVP mol-
ecules can be diagrammatically shown in Scheme 1.
To understand further the mechanism for the formation of
Fe3O4 nanospheres, time-dependent experiments were carried out.

Figure 9. TEM images of Fe3O4 at different reaction time: (a) 3, (b) 4, (c) 5, (d) 6, (e) 7, and (f) 8 h. Other conditions: 1.0 g FeCl3 3 6H2O, 0.5 g PVP,
2 g NaAc, and reaction temperature: 200 °C.

18930 dx.doi.org/10.1021/jp200418j |J. Phys. Chem. C 2011, 115, 18923–18934


The Journal of Physical Chemistry C ARTICLE

Figure 10. Magnetic hysteresis curves and enlarged partial curves


(inset) of Fe3O4 shown in Figure 9.

Table 2. Magnetic Parameters of Fe3O4 Nanoparticles Ob- Figure 11. UV spectra of XO oxidized by H2O2 under ultrasound
tained at Different Reaction Timea measured at (a) initial time and (b) 45 min. 10 mL 1  104 mol/L
10 mL 30% H2O2. The photo in the inset demonstrates no obvious
sample a b c d e f
degradation of XO without the catalysts. Reaction temperature: room
reaction time (h) 3 4 5 6 7 8 temperature.
Ms (emu/g) 66.1 81.2 81.6 81.6 80 79.1
Hc (Oe) 47.8 64.1 72.5 39.7 67.8 50.2
a
Other conditions: 1.0 g FeCl3 3 6H2O, 0.5 g PVP, 2.0g NaAc dispersed
in 30 mL of ethylene glycol, at 200 °C.

Scheme 2. Possible Formation Mechanism of Fe3O4


Nanospheres

Figure 12. UV spectra of XO oxidized by H2O2 under ultrasound with


Fe3O4 as catalyst measured at different reaction time. We used 10 mg
catalyst. The photo in the inset demonstrates the convenient separation
of the catalysts applying an external magnetic field. The other reaction
As shown in Figure 9A, after 3 h of reaction, Fe3O4 nanospheres condition was the same as that described in Figure 11.
with the diameter of 150 nm are obtained. The diameter of the
sample collected at 4 h later (Figure 9B) showed the same size with could be formulated in Scheme 2. First, the solid precursor
those of 3 h but with fine hierarchical nanoarchitectures. As the FeCl3 3 6H2O was dissolved in ethylene glycol. Second, part of
reaction proceeded (Figure 9 C), the diameter of the porous the dissolved species was reduced by the ethylene glycol. Then,
nanospheres continues to grow. The corresponding TEM images nanocrystal nucleations of Fe3O4 were formed. At last, Fe3O4
of the Fe3O4 samples obtained at 200 °C for 6 h (Figure 9D) nanocrystals grew larger and assembled to porous nanospheres
showed that many nanospheres are derived from the small by self-assembling with PVP as capping agent. Here ethylene
nanocrystals, which is in agreement with the morphology shown glycol served not only as a solvent but also as a reducing agent.72
in Figure 1. Unfortunately, failure occurred to obtain the primary As for anhydrous sodium acetate, it initialized the nucleation of
Fe3O4 nanocrystals, which may be due to the quick growing of the Fe3O4 nanocrystals.
primary Fe3O4 nanocrystals. However, as shown in Figure 9E,F, no 3.3. Catalytic Activity of the Porous Fe3O4 Nanospheres.
further change with increase in reaction time indicated the The porous Fe3O4 nanospheres could be used in catalytic
completion of the growth of nanocrystal. In our system, product degradation of organic dyes. Chemical oxidation degradation
at 8 h presents the best morphology. The magnetic property of XO by H2O2 was used as a model reaction to study the
evolution of the products is shown in Figure 10. It is clear that catalytic property of the porous Fe3O4 nanospheres. According
reaction time directly affects the final magnetic property of the to Figure 11, without the porous Fe3O4 nanospheres as a catalyst,
products. However, there is no linear relationship between the the absorption peak at 434 nm for XO decreased a little when
reaction time and the magnetic property. The values of Ms and Hc H2O2 was added under ultrasound for 45 min, which indicates
for the samples are listed in Table 2. The results reveal that all that the degradation of XO is very slow, even with excessive
products show weak ferromagnetic property. H2O2. The photo in the inset of Figure 11 can further demon-
According to above experimental results, an illustration of the strates that no obvious degradation occurred without the cata-
predicted possible formation mechanism of Fe3O4 nanospheres lysts. As shown in Figure 12, however, when trace porous Fe3O4

18931 dx.doi.org/10.1021/jp200418j |J. Phys. Chem. C 2011, 115, 18923–18934


The Journal of Physical Chemistry C ARTICLE

Figure 13. (A) EDS of Fe3O4 nanospheres recycled after catalyzing degradation of XO seven times. (B) FTIR spectra of the porous Fe3O4 nanospheres
after treating XO (a) and XO (b). (C) XPS spectra of the used porous Fe3O4 nanosphere catalyst. The inset is expanded spectra of S2p.

Table 3. Measurement of Residual Xylenol Orange under Different Systems


treat method only H2O2 only prepared Fe3O4 prepared Fe3O4 +H2O2 commercial Fe3O4 + H2O2

C/C0 0.97 0.94 0 0.90


TOC/TOC0 95% 93% 6% 88%

nanospheres were added to the solution, the intensity of absorp- discussion, it was concluded that the degradation of XO occurred
tion peak of XO at 434 nm decreased gradually and disappeared in the presence of H2O2 with the catalysis of the porous Fe3O4
completely after 45 min under ultrasound, which clearly shows nanospheres.
that the porous Fe3O4 nanospheres are an efficient catalyst for As a comparison, a series of experiments was carried out. The
the degradation of XO by H2O2. Furthermore, it can be clearly experimental results are listed in Table 3, where the solution that
demonstrated by the photography in the inset of Figure 12 that contained XO was treated by several methods, for example: (1)
the catalysts can be conveniently separated by applying an only H2O2, (2) only Fe3O4 prepared in this Article, (3) H2O2 +
external magnetic field. Fe3O4 prepared in this Article, and (4) H2O2 + commercial
To confirm that the decrease in XO in the solution is due to the Fe3O4. The decrease in XO in solution was determined by
catalysis of Fe3O4 nanospheres, instead of adsorption by porous measuring the UV spectrum and TOC of the reaction system.
Fe3O4 nanospheres, EDS were taken to study the surface It can be found from Figure S1 of the Supporting Information
element of Fe3O4. According to Figure 13A, it is clear that that the commercial Fe3O4 shows poor catalytic performance
no sulfur element on the surface of Fe3O4 nanospheres was because the UV absorption at 434 nm decreases slightly after
observed, which shows that no XO was absorbed by Fe3O4 ultrasound irradiation for 45 min in the presence of H2O2. As
nanospheres. Figure 13B shows the FT-IR Fe3O4 catalysts displayed in Table 3, concentration of both XO and TOC in the
recovered from the catalytic system; also, the characteristic peaks reaction system decreased near to zero while the oxidant H2O2
of XO could not be observed. Furthermore, XPS, a more- and the as-prepared Fe3O4 were added. However, the decrease in
sensitive surface analysis technique, was used to examine quanti- XO was minimal by other methods, further confirming that the
tatively if the sulfur is present on the surface of the used porous decrease in XO in our reaction system is due to the oxidation of
Fe3O4. As is seen in Figure 13C, the signal of sulfur is too weak to XO by H2O2 with the as-prepared Fe3O4 catalyst. The very low
be observed in the whole spectra. From the expanded spectra of catalytic activity of the commercial Fe3O4 may be mainly due to
S2p (Figure 13C inset), the percentage of sulfur was determined its large size, low porosity, and low surface area. Figure S2 of the
to be 0.16%, which indicates that the adsorption of XO on the Supporting Information shows the TEM image and BET iso-
surface of catalyst is negligible. On the basis of the above therm of the commercial Fe3O4. From Figure S2A of the

18932 dx.doi.org/10.1021/jp200418j |J. Phys. Chem. C 2011, 115, 18923–18934


The Journal of Physical Chemistry C ARTICLE

repeated seven times, and the results are shown in Figure 15.
The catalytic performance of the porous Fe3O4 nanospheres
decreased slowly with the increase in cycle number. The percen-
tage degradation remained steady (>95%) in the first four cycles;
then, the degradation of XO decreased from the fifth cycle to
the seventh cycle. At the end of seven cycles, the percentage
degradation remained steady at >75% of the initial value. These
results show that the porous Fe3O4 nanospheres can be recycled
for catalyzing degradation of XO with H2O2 solution. Therefore,
the as-prepared Fe3O4 nanospheres can be used as ideal catalysts
in practical applications.

4. CONCLUSIONS
In conclusion, porous Fe3O4 nanospheres have been prepared
Figure 14. First-order kinetic plot of XO degradation using H2O2 as with solvothermal method. The products exhibit weak ferromag-
oxidizing agent under ultrasound in with Fe3O4 nanospheres as catalyst.
netic properties at room temperature. The quality of the product
is determined by the experimental variables, such as the con-
centration of the precursor, capping agent, and precipitation
agent as well as the reaction temperature and reaction time. The
disadvantage of the method reported here is the wide size
distribution of the product. The as-prepared porous Fe3O4
nanospheres show highly catalytic degradation ability toward
degradation of organic dyes in aqueous solution. Although the
catalytic activity decreases slightly with increasing the cycles of
tests, such nanostructured catalyst is expected to find application
in industrial applications, where separation and recycling are
crucially required in terms of cost and environmental protection.

’ ASSOCIATED CONTENT

bS Supporting Information. UV spectra of XO catalyzed by


the commercial Fe3O4; TEM image and BET isotherm of the
Figure 15. Catalytic activity of Fe3O4 nanospheres in different cycling commercial Fe3O4. This material is available free of charge via the
numbers. Internet at http://pubs.acs.org.

Supporting Information, the size of the commercial Fe3O4 can be ’ AUTHOR INFORMATION
determined to be >1 μm, which is very large compared with that
prepared in this work. Meanwhile, the porosity of the materials is Corresponding Author
another most important factor influencing the catalytic activity. *Tel: +86-514-87975436. Fax: +86-514-87975244. E-mail:
Figure S2B of the Supporting Information shows the N2 adsorp- gwdiao@yzu.edu.cn.
tiondesorption isotherm of the commercial Fe3O4, determining
that the BET surface area of the commercial Fe3O4 is 1.41 m2/g, ’ ACKNOWLEDGMENT
whereas the value for the prepared Fe3O4 in this study was
determined to be 47.7 m2/g (seen in Figure 2G). We acknowledge the financial support from the National
The degradation kinetic of XO was investigated by adding Natural Science Foundation of China (grant no. 20973151,
excess H2O2. Therefore, the degradation can be considered to be 20901065), the Natural Science Key Foundation of Educational
a pseudo-first-order reaction. The concentration, ct, of XO at Committee of Jiangsu Province of China (grant no. 07KJA
different reaction time, t, could be described as follows 15015), the Specialized Research Fund for the Doctoral Program
of Higher Education (SRFDP, 20093250110001), the Founda-
ct ¼ c0 expðktÞ tion of Jiangsu Provincial Key Program of Physical Chemistry in
Yangzhou University, and the Foundation of Jiangsu Key La-
where c0 is the initial concentration of XO and k is the pseudo- boratory of Fine Petrochemical Technology. We also acknowl-
first-order rate constant. Figure 14 shows the logarithmic plot of edge the Foundation of the Educational Committee of Jiangsu
the concentration of XO as a function of degradation time. A Provincial General Universities Graduate Student Scientific
well-behaved linear straight line shows that the degradation of Research Invention Plan.
XO is a pseudo-first-order reaction. The rate constant k can be
calculated from the slope of the straight line. The value of k is ’ REFERENCES
0.056 min1, which is much larger than the literature reported (1) David, H.; Gracias, J. T.; Tricia, L. B.; Carey, H.; George, M. W.
0.013 min1 using Pd/TiO2 as the catalyst.73 Science 2000, 289, 1170–1172.
For practical recyclable catalysts, keeping high catalytic activity (2) Park, S.; Lim, J. H.; Chung, S. W.; Mirkin, C. A. Science 2004,
in each cycle is necessary. The effect of recycling times of the 303, 348–351.
porous Fe3O4 nanospheres on the catalytic performance was (3) Service, R. F. Science 2005, 309, 95.

18933 dx.doi.org/10.1021/jp200418j |J. Phys. Chem. C 2011, 115, 18923–18934


The Journal of Physical Chemistry C ARTICLE

(4) Zhu, M.; Diao, G. Nanoscale 2011, 3, 2748–2767. (39) Moumen, N.; Pileni, M. P. J. Phys. Chem. 1996, 100,
(5) Guardia, P.; Labarta, A.; Batlle, X. J. Phys. Chem. C 2011, 1867–1873.
115, 390–396. (40) Kumar, S.; Nann, T. Small 2006, 2, 316–329.
(6) Li, Q.; Li, H.; Pol, V. G.; Bruckental, I.; Koltypin, Y.; Calderon- (41) Byrappa, K.; Adschiri, T. Prog. Cryst. Growth Charact. Mater.
Moreno, J.; Nowike, I.; Gedanken, A. New J. Chem. 2003, 27, 2007, 53, 117–166.
1194–1199. (42) Li, Z.; Sun, Q.; Gao, M. Angew. Chem., Int. Ed. 2005, 44
(7) Hyeon, T. Chem. Commun. 2003, 927–934. 123–126.
(8) Chen, X.; Unruh, K. M.; Ni, C.; Ali, B.; Sun, Z.; Lu, Q.; Deitzel, J.; (43) Wang, X.; Zhuang, J.; Peng, Q.; Li, Y. Nature 2005, 437,
Xiao, J. Q. J. Phys. Chem. C 2011, 115, 373–378. 121–124.
(9) Li, Y. C.; Lin, Y. S.; Tsai, P. J.; Chen, C. T.; Chen, W. Y.; Chen, (44) Deng, H.; Li, X.; Peng, Q.; Wang, X.; Chen, J.; Li, Y. Angew.
Y. C. Anal. Chem. 2007, 79, 7519–7525. Chem., Int. Ed. 2005, 44, 2782–2785.
(10) Chen, J.; Xu, L.; Li, W.; Gou, X. Adv. Mater. 2005, 17, 582–586. (45) Pinna, N.; Grancharov, S.; Beato, P.; Bonville, P.; Antoniette,
(11) Frey, N. A.; Peng, S.; Cheng, K.; Sun, S. Chem. Commun. 2009, M.; Niederberger, M. Chem. Mater. 2010, 22, 3183–3193.
38, 2532–2542. (46) Huang, X.; Zhuang, J.; Chen, D.; Liu, H.; Tang, F.; Yan, X.;
(12) Gao, L.; Zhuang, J.; Leng, N.; Zhang, J.; Zhang, Y.; Gu, N.; Meng, X.; Zhang, L.; Ren, J. Langmuir 2009, 25, 11657–11663.
Wang, T.; Feng, J.; Yang, D.; Perrett, S.; Yan, X. Nat. Nanotechnol. 2007, (47) Ge, J.; Yin, Y. J. Mater. Chem. 2008, 18, 5041–5045.
2, 577–583. (48) Ge, J.; He, L.; Goebl, J.; Yin, Y. J. Am. Chem. Soc. 2009,
(13) Perez, J. M. Nat. Nanotechnol. 2007, 2, 535–536. 131, 3484–3486.
(14) Zeng, T.; Chen, W. W.; Cirtiu, C. M.; Moores, A.; Song, G.; Li, (49) Nath, S.; Kaittanis, C.; Ramachandran, V.; Dalal, N. S.; Perez,
C. J. Green Chem. 2010, 12, 570–573. J. M. Chem. Mater. 2009, 21, 1761–1767.
(15) Sun, G.; Dong, B.; Cao, M.; Wei, B.; Hu, C. Chem. Mater. 2011, (50) Vayssieres, L.; Sathe, C.; Butorin, S.; Shuh, D.; Nordgren, J.;
23, 1587–1593. Guo, J. Adv. Mater. 2005, 17, 2320–2323.
(16) Valero, E.; Tambalo, S.; Marzola, P.; Ortega-Mu~ noz, M.; (51) Woo, K.; Lee, H. J.; Ahn, J. P.; Park, Y. S. Adv. Mater. 2003,
Lopez-Jaramillo, F. J.; Santoyo-Gonzalez, F.; Lopez, J. D.; Delgado, 15, 1761–1764.
J. J.; Calvino, J. J.; Cuesta, R.; Domínguez-Vera, J. M.; Galvez, N. J. Am. (52) Chen, Y. J.; Zhang, F.; Zhao, G. G.; Fang, X. Y.; Jin, H. B.; Gao,
Chem. Soc. 2011, 133, 4889–4895. P.; Zhu, C. L.; Cao, M. S.; Xiao, G. J. Phys. Chem. C 2010,
(17) Lewin, M.; Carlesso, N.; Tung, C. H.; Tang, X. W.; Cory, D.; 114, 9239–9244.
Scadden, D. T.; Weissleder, R. Nat. Biotechnol. 2000, 18, 410–414. (53) Chen, Y. J.; Gao, P.; Wang, R. X.; Zhu, C. L.; Wang, L. J.; Cao,
(18) Yoon, T. J.; Kim, J. S.; Kim, B. G.; Yu, K. N.; Cho, M. H.; Lee, M. S.; Jin, H. B. J. Phys. Chem. C 2009, 113, 10061–10064.
J. K. Angew. Chem., Int. Ed. 2005, 44, 1068–1071. (54) Zou, G.; Xiong, K.; Jiang, C.; Li, H.; Li, T.; Du, Jin.; Qian, Y.
(19) Ceyhan, B.; Alhorn, P.; Lang, C.; Sch€uler, D.; Niemeyer, C. J. Phys. Chem. B. 2005, 109, 18356–18360.
Small 2006, 2, 1251–1255. (55) Ye, M.; Zorba, S.; He, L.; Hu, Y.; Maxwell, R. T.; Farah, C.;
(20) Mykhaylyk, O.; Antequera, Y. S.; Vlaskou, D.; Plank., C. Nat. Zhang, Q.; Yin, Y. J. Mater. Chem. 2010, 20, 7965–7969.
Protoc. 2007, 2, 2391–2411. (56) Wang, C.; Xu, C.; Zeng, H.; Sun, S. Adv. Mater. 2009, 21, 1–8.
(21) Park, H.; Yang, J.; Seo, S.; Kim, K.; Suh, J.; Kim, D.; Haam, S.; (57) Hamada, S.; Matijevic, E. J. Colloid Interface Sci. 1981, 84
Yoo, K. H. Small 2008, 4, 192–196. 274–277.
(22) Shi, D. Adv. Funct. Mater. 2009, 19, 3356–3373. (58) Wen, X.; Wang, S.; Ding, Y.; Wang, Z. L.; Yang, S. J. Phys. Chem.
(23) Qiao, R.; Yang, C.; Gao, M. J. Mater. Chem. 2009, 19, B. 2005, 109, 215–220.
6274–6293. (59) Jia, C. J.; Sun, L. D.; Yan, Z. G.; You, L. P.; Luo, F.; Han, X. D.;
(24) Shariati, S.; Faraji, M.; Yamini, Y.; Rajabi, A. A. Desalination Pang, Y. C.; Zhang, Z.; Yan, C. H. Angew. Chem., Int. Ed. 2005, 44
2011, 270, 160–165. 4328–4333.
(25) Elliott, D. W.; Zhang, W. X. Environ. Sci. Technol. 2001, 35, (60) Ozaki, M.; Kratohvil, S.; Matijevic, Egon. J. Colloid Interface Sci.
4922–4926. 1984, 102, 146–151.
(26) Takafuji, M.; Ide, S.; Ihara, H.; Xu, Z. Chem. Mater. 2004, 16, (61) Zhu, Y. W.; Yu, T.; Sow, C. H.; Liu, Y. J.; Wee, A. T. S.; Xu, X. J.;
1977–1983. Lim, C. T.; Thong, J. T. L. Appl. Phys. Lett. 2005, 87, 0231031–023103-3.
(27) Zhong, L. S.; Hu, J. S.; Liang, H. P.; Cao, A. M.; Song, W. G.; (62) Xia, H.; Foo, P.; Yi, J. Chem. Mater. 2009, 21, 2442–2451.
Wan, L. J. Adv. Mater. 2006, 18, 2426–2431. (63) Wang, L.; Bao, J.; Wang, L.; Zhang, F.; Li, Y. Chem.—Eur.
(28) Dong, J.; Xu, Z.; Kuznicki, S. M. Environ. Sci. Technol. 2009, J. 2006, 12, 6341–6347.
43, 3266–3271. (64) Liu, R.; Zhao, Y.; Huang, R.; Zhao, Y.; Zhou, H. Eur. J. Inorg.
(29) Wang, C.; Tao, S.; Wei, W.; Meng, C.; Liu, F.; Han, M. J. Mater. Chem. 2010, 4499–4505.
Chem. 2010, 20, 4635–4641. (65) Gu, J.; Li, S.; J, M.; Wang, E. J. Cryst. Growth 2011, 320, 46–52.
(30) Lu, A. H.; Salabas, E. L.; Sch€uth, F. Angew. Chem., Int. Ed. 2007, (66) Huang, Z.; Tang, F. J. Colloid Interface Sci. 2005, 281, 432–436.
46, 1222–1244. (67) Pinna, N.; Grancharov, S.; Beato, P.; Bonville, P.; Antonietti,
(31) Jeong, U.; Teng, X.; Wang, Y.; Yang, H.; Xia, Y. Adv. Mater. M.; Niederberger, M. Chem. Mater. 2005, 17, 3044–3049.
2007, 19, 33–60. (68) Liu, J.; Wei, J.; Li, S. Mater. Lett. 2007, 61, 1529–1532.
(32) Zeng, H.; Sun, S. Adv. Funct. Mater. 2008, 18, 391–400. (69) Schwertmann, U.; Cornell., R. M. The Iron Oxides: Structure,
(33) Lauren, S.; Forge, D.; Port, M.; Roch, A.; Robic, C.; Elst, L. V.; Properties, Reactions, Occurrences and Uses; Wiley-VCH: Weinheim,
Muller, R. N. Chem. Rev. 2008, 108, 2064–2110. Germany, 2003.
(34) Teja, A. S.; Koh, P. Y. Prog. Cryst. Growth Charact. Mater. 2009, (70) Li, H.; Bian, Z.; Zhu, J.; Zhang, D.; Li, G.; Huo, Y.; Li, H.; Lu, Y.
55, 22–45. J. Am. Chem. Soc. 2007, 129, 8406–8407.
(35) Hao, R.; Xing, R.; Xu, Z.; Hou, Y.; Gao, S.; Sun, S. Adv. Mater. (71) Barlow, D. A.; Baird, J. K.; Su, C. H. J. Cryst. Growth. 2004,
2010, 22, 2729–2742. 264, 417–423.
(36) Shylesh, S.; Sch€unemann, V.; Thiel, W. R. Angew. Chem., Int. Ed. (72) Bee, A.; Massart, R.; Neveu, S. J. Magn. Magn. Mater. 1995,
2010, 49, 3428–3459. 149, 6–9.
(37) Rockenberger, J.; Scher, E. C.; Alivisatos, A. P. J. Am. Chem. Soc. (73) Iliev, V.; Tomova, D.; Bilyarska, L.; Petrov, L. Catal. Commun.
1999, 121, 11595–11596. 2004, 5, 759–763.
(38) Sun, S.; Zeng, H.; Robinson, D. B.; Raoux, S.; Rice, P. M.;
Wang, S. X.; Li, G. J. Am. Chem. Soc. 2004, 126, 273–279.

18934 dx.doi.org/10.1021/jp200418j |J. Phys. Chem. C 2011, 115, 18923–18934

You might also like