Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

pubs.acs.

org/IECR Article

From Core−Shell to Yolk−Shell: Improved Catalytic Performance


toward CoFe2O4@ Hollow@ Mesoporous TiO2 toward Selective
Oxidation of Styrene
Liang Liu, Wei He, Zheng Fang, Zhao Yang,* Kai Guo, and Zhixiang Wang
Cite This: Ind. Eng. Chem. Res. 2020, 59, 19938−19951 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
Downloaded via KING MONGKUTS UNIV TECH THONBURI on April 27, 2024 at 17:13:26 (UTC).
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Developing catalysts with structural characteristics,


reusability, cost-effectiveness, environmental friendliness, and
preferable catalytic performance is challenging for the selective
oxidation of alkenes. In this article, we report the synthesis of
magnetically separable CoFe2O4@ hollow@ mesoporous TiO2
with yolk−shell structure, which could be used as a selectivity
variable and stable catalyst for the selective oxidation of styrene,
yielding different oxidation products. With oxygen as the oxidant,
epoxidation of styrene occurred, affording an epoxy group. With
hydrogen peroxide as the oxidant, oxidative cleavage of double
bonds in the styrene and further oxidation were conducted,
affording a carboxylic acid group. Compared with pristine
CoFe2O4, TiO2, and CoFe2O4/TiO2 core−shell structure, the
CoFe2O4@ hollow@ mesoporous TiO2 with yolk−shell structure was found to be a more efficient catalyst for the oxidation of
styrene. Meanwhile, the kinetic analysis and activation energies of different catalysts were investigated to illustrate the relationship
between the performance and the structure of catalysts. Active radical scavenging experiments were conducted to figure out the
dominant radicals in the oxidation reaction. The corresponding possible reaction mechanisms of the oxidation reaction were also
proposed.

■ INTRODUCTION
Selective oxidation of styrene to corresponding epoxides,
catalysts,14−16 hydrotalcites,11 and Ga-Co-HMS-X,17 have
been developed to catalyze the oxidation of styrene due to
mild reaction conditions, easy separation, and reusability.2
aldehydes, or carboxylic acids is of great interest for both
Although high activity (>98%) and selectivity (>68%) could be
industrial production and academic research due to that the
achieved, most of the catalysts involved precious metals. In
oxidation products, such as tyrene oxide (SO), benzaldehyde
addition, peroxides were used to initiate the reaction, which
(BzA), and benzoic acid (BA), are key organic intermediates
was considered hazardous and environmentally unfriendly.18,19
for pharmaceuticals and fine chemicals.1−3 Notably, SO and
On the other hand, these oxidation reactions are limited by
BA are considered as starting materials for various industrial
uneconomical, harsh conditions or complicated preparation
products, such as plasticizer, epoxy resins, and perfumes.2,4
procedures. Thus, it is necessary to develop efficient
Traditionally, SO is synthesized through cyclization process or
heterogeneous catalysts in the oxidation of styrene.
epoxidation process.2,5 However, basic or acidic catalysts are
Spinel ferrites, MFe2O4 (M = Fe, Mn, Co, Ni, Cu, Zn), have
necessary in these processes, resulting in the generation of
recently attracted tremendous attention in magnetic resonance
enormous waste water. Similarly, BA is mainly produced by
imaging,20 controlled drug release,21 superparamagnetic
direct oxidation of toluene in liquid−gas reaction system with materials,22 electrocatalytic energy conversion,23,24 biosens-
high temperature and pressure. More efficient green oxidants ing,25 and organic pollutant degradation26−28 owing to their
such as hydrogen peroxide, molecular oxygen, and tert-butyl environmental compatibility and stability as well as easy
hydroperoxide6−8 have been applied in this liquid−gas
oxidation reaction to realize the oxidation process at relatively
low temperatures and pressures. Homogeneous and heteroge- Received: August 7, 2020
neous catalysts have also been designed in the oxidation of Revised: October 18, 2020
styrene.6,9−11 However, homogeneous catalysts have received Accepted: October 19, 2020
less attention because of the difficulty in separating catalysts Published: October 28, 2020
from the reaction mixture. Various heterogeneous catalysts,
such as Au supported by metal oxides,12,13 Ag-doped

© 2020 American Chemical Society https://dx.doi.org/10.1021/acs.iecr.0c03884


19938 Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

synthesis.29 Especially, they have been widely used as was proposed for a better understanding of the reaction
heterogeneous catalysts in many organic reactions due to the process.
synergistic intermetallic interactions and magnetic recovery
properties.30,31 Ding et al.32 developed spinel CoMn2O4
hollow spheres with a high selectivity (100%) and conversion
■ MATERIALS AND METHODS
Materials. Absolute ethanol (C 2 H 6 O), 1,4-dioxane
(41.6%) for the oxidation of 5-hydroxymethylfurfural. In (C4H8O2), concentrated ammonia aqueous (NH3·H2O), and
addition, these catalysts could be separated from a liquid cetyltrimethylammonium bromide (C19H42BrN) were ob-
reaction by simply placing a magnet on the outside of the flask. tained from Adamas-Beta (Shanghai, China). Ferric chloride
Among all of these spinel ferrites, CoFe2O4, was used as an (FeCl 3 ·6H 2 O), cobalt chloride (CoCl 2 ·6H 2 O), poly-
efficient catalyst for many oxidation reactions. Nano-CoFe2O4 (vinylpyrrolidone) (PVP), and tetraethyl orthosilicate
catalysts were fabricated through the bio-gel template method. (TEOS) were purchased from Macklin (Shanghai, China).
The results revealed that the maximum conversion of Sodium hydroxide (NaOH) was supplied by Sinopharm
cyclohexane oxidation reached 16.43% and the selectivity for Chemical Reagent (Shanghai, China). Styrene was purchased
KA oil was 90.3%.33 Alarcon et al.34 found that Au from J&K Scientific (Beijing, China). All chemicals used in this
nanoparticles deposited at the surface of magnetic CoFe2O4 study were of analytical grade and used without further
presented high performance and enhancement in the oxidation purification. Ultrapure water (18.2 MΩ·cm) from a Millipore
of dimethylphenylsilane. CoFe2O4 nanoparticles with an system (ELGA, U.K.) was employed throughout the experi-
average particle size of 4 nm presented high conversion ments.
(80%) in the selective oxidation of 2-propanol.35 Thus, Catalyst Preparation. Synthesis of CoFe2O4. Spinel ferrite
CoFe2O4 has displayed excellent catalytic activity toward CoFe2O4 was prepared by a facile coprecipitation method.43 In
oxidation reaction. a typical synthesis, 2.5 mmol of CoCl2·6H2O and 5 mmol of
Recently, CoFe2O4 displayed certain catalytic activity FeCl3·6H2O were dissolved in 40 mL of DI water. Then, 0.75 g
towards the oxidation of alkene, yielding epoxide products.36,37 of poly(vinylpyrrolidone) (PVP) K30 was added to the above
However, relatively expensive reagents, such as tert-butyl solution under magnetic stirring and maintained at 90 °C.
hydroperoxide, were applied in the oxidation process. In Afterward, the pH of the solution was adjusted to 12 by
addition, single oxidation product was obtained in these dropwise addition of 2 M NaOH and kept for 2 h under
oxidation processes of alkene. Double bonds with high reaction stirring conditions before cooling down to room temperature
activity in the alkene could react with various substrates, naturally. The resulting black precipitate was collected by a
affording numerous products involved in pharmaceuticals, magnet and washed with DI water and ethanol repeatedly.
materials, and chemical industry. Thus, development of novel Finally, the products were dispersed in 100 mL of deionized
catalysts based on CoFe2O4 toward the oxidation of alkene was water for further use.
necessary for the selective and accurate synthesis of various Synthesis of CoFe2O4/SiO2 Core−Shell Structure. The
oxidation products. Complex catalysts including CoFe2O4 and interlayer of SiO2 was prepared through a modified Stöber
other active components have been identified as more active method.44 In a typical process, 20 mL of the as-prepared
catalysts. On the one hand, catalytic performance would be CoFe2O4 suspension was added in 60 mL of absolute ethanol.
enhanced due to the synergistic effects in the ferrites.38,39 On Then, the suspension was dispersed under ultrasound for 15
the other hand, the active components with inherent catalytic min. Meanwhile, 1 mL of tetraethyl silicate solution was added
activity would promote the uniform dispersion of active sites, dropwise. After that, the mixture solution was mechanically
avoiding catalyst deactivation caused by aggregation of active stirred at 50 °C. Then, 2 mL of concentrated ammonia
sites. In our previous study, metal oxides, such as TiO2, have aqueous solution was added dropwise. After stirred for 12 h,
displayed excellent catalytic activity toward oxidation reaction the precipitate was washed alternately with absolute ethanol
due to the existence of Ti4+. Metal-alkylperoxy intermediate and deionized water, and then dried under vacuum at 60 °C
would be formed to promote the oxidation of alkene.40 Thus, for further use.
the integration of CoFe2O4 and TiO2 might promote the Synthesis of CoFe2O4/SiO2/Mesoporous TiO2 Core−Shell
oxidation of alkene. More importantly, the constitution form of Structure. The obtained CoFe2O4/SiO2 composite and 0.116
two different components played a vital role in the perform- g of cetyltrimethylammonium bromide were dispersed in 75
ance of the resulting hybrid compound.41,42 However, there mL of absolute ethanol under ultrasound. After forming a
has been no study concerning the relationship between homogeneous dispersion, 0.635 mL of titanium tetraisopropa-
constitution form of complex catalyst and catalytic activity in nolate was introduced under mechanical stirring and
the oxidation of alkene. maintained at 50 °C. Then, 0.85 mL of concentrated ammonia
In this study, two different complex catalysts including aqueous solution was added dropwise. After stirred for 24 h,
yolk−shell structure and core−shell structure based on the precipitate was washed alternately with absolute ethanol
CoFe2O4 were prepared through the integration of a facile and DI water and then dried under vacuum at 60 °C for 12 h.
coprecipitation method and a sol−gel method. A comparison Finally, the precipitate was calcined in air atmosphere at 600
between the yolk−shell structure and the normal core−shell °C for 2 h.
structure was made to reveal the role of internal structure in Synthesis of CoFe2O4@ Hollow@ Mesoporous TiO2 Yolk−
the catalytic activity. Different oxidants (O2 and hydrogen Shell Structure. A CoFe2O4@ hollow@ mesoporous TiO2
peroxide) were applied in the study to investigate the selective yolk−shell structure was prepared by etching silica inter-
oxidation of styrene. The reaction parameters including mediate layer.44 Typically, the obtained CoFe2O4/SiO2/
reaction temperature, catalyst amount, molar ratio, reaction mesoporous TiO2 composite was dispersed in NaOH solution
time, solvent, and the amount of urea were investigated in (2 M, 60 mL). The reaction mixture was stirred in 60 °C for 5
detail. In addition, the kinetic analysis and activation energies h. Then, the resulting participate was washed with deionized
of different catalysts were studied. The reaction mechanism water after filtration. Finally, the CoFe2O4@ hollow@
19939 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Scheme 1. Overall Procedure for the Synthesis of CoFe2O4@ Hollow@ Mesoporous TiO2 Yolk−Shell Structure and CoFe2O4/
TiO2 Core−Shell Structure

mesoporous TiO2 yolk−shell catalyst was obtained by drying When hydrogen peroxide was selected as the oxidant, a 15
at 60 °C for 12 h. mL glass flask was used. In a typical reaction, certain amounts
Synthesis of CoFe2O4/TiO2 Core−Shell Structure. For a of styrene, solvent, and catalyst were added into the flask under
better comparison, a CoFe2O4/TiO2 core−shell structure stirring. After heating the reaction mixture to the desired
without silica intermediate layer was prepared as well under temperature, a certain amount of hydrogen peroxide (30%)
similar conditions. Typically, 20 mL of the as-prepared was added dropwise into the reaction mixture to initiate the
CoFe2O4 suspension was dried under vacuum at 60 °C reaction.
overnight. Then, the dried black powder and 0.116 g of For the recycling test, the used catalyst was magnetically
cetyltrimethylammonium bromide were dispersed in absolute separated and washed with absolute ethanol three times. Then,
ethanol (75 mL) under ultrasound. After that, 0.635 mL of the obtained catalyst was dried at 100 °C for 4 h and applied in
titanium tetraisopropanolate was added into the solution under the next run. After the reaction, 0.1 mL of the suspension was
mechanical stirring and maintained at 50 °C. Then, extracted and analyzed. The content of SO and the conversion
concentrated ammonia aqueous solution (0.850 mL) was of styrene were determined by a gas chromatograph equipped
added dropwise. The reaction was conducted for 24 h. After with an HP-5 column.
the reaction, the precipitate was washed alternately with Liquid chromatography equipped with C18 column was
absolute ethanol and DI water, and then dried under vacuum used to quantify the content of BA. The styrene conversion
at 60 °C for 12 h. Finally, the precipitate was calcined in air and the selectivities of SO and BA were calculated by the
atmosphere at 600 °C for 2 h. In addition, the overall following equations:
procedure for the synthesis of the CoFe2O4@ hollow@
styrene conversion (%)
mesoporous TiO2 yolk−shell structure and CoFe2O4/TiO2
core−shell structure is presented in Scheme 1. = (converted styrene in molar/initial styrene in molar)
Catalyst Characterization. X-ray powder diffraction
× 100% (1)
(XRD) analysis was conducted on an Ultima IV diffractometer
with Cu Kα-radiation in the 2θ range of 10−80° with a scan
SO selectivity (%) = (SO in the product in molar
rate of 2°/min. Fourier transform infrared (FTIR) spectros-
copy was carried out on a Bruker Tensor II. Scanning electron /converted styrene in molar) × 100%
microscopy (SEM) images and energy-dispersive X-ray (2)
spectrum (EDX) mapping were obtained from a Quanta
FEG 250 equipped with an energy-dispersive X-ray spec- BA selectivity (%) = (BA in the product in molar
trometer. Transmission electron microscopy (TEM) and high-
resolution TEM (HRTEM) images were recorded using an /converted styrene in molar) × 100%
HT7800 microscope equipped with an energy-diffusive X-ray (3)
spectroscopy (EDS) attachment. X-ray photoelectron spec-
troscopy (XPS) measurements were characterized on a
Thermo Scientific K-Alpha+ spectrometer with an Al Kα
■ RESULTS AND DISCUSSION
Catalyst Characterization. The XRD patterns of
microfocus monochromatic source. The N2 adsorption− CoFe 2 O 4 , CoFe 2 O 4 /TiO 2 core−shell structure, and
desorption isotherms were collected on a Quantachrome CoFe2O4@ hollow@ mesoporous TiO2 yolk−shell catalysts
Autosorb-iQ3 at 77 K. are displayed in Figure 1. The characteristic diffraction peaks
Catalytic Test. When molecular oxygen was selected as the observed at 18.2, 30.1, 35.5, 43.5, 53.9, 57.2, and 62.7°
oxidant, a 25 mL Schlenk tube was used. In a typical reaction, a corresponded to the (111), (220), (311), (400), (422), (511),
certain amount of catalyst was added into the tube. Afterward, and (440) planes of the spinel CoFe2O4 (PDF card no. 22-
the system was evacuated and then filled with molecular 1086). The peaks at 25.3, 37.8, 48.1, 54.0, 55.1, 62.7, 68.8,
oxygen three times. Especially, a balloon filled with molecular 70.3, and 75.1° could be assigned to the (101), (004), (200),
oxygen was inserted on the stopper to meet the consumption (105), (211), (204), (116), (220), and (215) planes of the
of molecular oxygen in the system. After adding the mixture of anatase TiO2, respectively (PDF card no. 84-1286). As
styrene and solvent into the tube, the reaction mixture was expected, the characteristic peaks of pure CoFe2O4 and TiO2
heated to the desired temperature. were clearly displayed in CoFe2O4@ hollow@ mesoporous
19940 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

The TEM image combined with EDS elemental mapping


allowed a direct view of the surface features and elemental
distribution in the CoFe2O4@ hollow@ mesoporous TiO2
(Figure 3). The HRTEM image in Figure 3b suggested a clear

Figure 1. XRD patterns of CoFe2O4, CoFe2O4/TiO2 core−shell


structure, and CoFe2O4@ hollow@ mesoporous TiO2 yolk−shell
structure.

TiO2 yolk−shell catalysts, indicating that the TiO2 shell had no


effect on the lattice structure of CoFe2O4 core. The strong and
sharp peaks of pristine CoFe2O4 catalyst suggested high
crystallinity. However, when coupled with CoFe2O4 directly,
the diffraction peaks of TiO2 became weak. It could be
attributed to the unexpected reaction between CoFe2O4 and
TiO2 in calcination process with high temperature, which
might result in the generation of Fe2TiO5, CoTi2O5, and
Fe2CoTi3O10.45,46 Thus, it was necessary to introduce an inert
silicon interlayer before the calcination process to prevent the
interaction between CoFe2O4 and TiO2.
The representative SEM image of the CoFe2O4@ hollow@ Figure 3. TEM image (a), HRTEM image (b), and the corresponding
mesoporous TiO2 catalyst was obtained to investigate the EDS mapping images (c−g) of CoFe2O4@ hollow@ mesoporous
TiO2.
surface morphology of the samples (Figure 2a). The elemental
distribution in the catalyst was evaluated by EDS elemental
mapping. The even distribution of Co, Fe, Ti, and O elements
in the catalyst is displayed in Figure 2c−f. In addition, the lattice fringe, further confirming a high crystallinity of the
distribution of different elements in Figure 2b demonstrated catalyst. The lattice fringe spacing of 0.25 nm in the spherical
the overlap of different elements, indicating the formation of particles was assigned to the (311) plane of CoFe2O4.47 In
the core−shell structure. EDX scanning of CoFe2O4@ addition, an interplanar distance of 0.35 nm was exhibited by
hollow@ mesoporous TiO2 (Figure S1) also confirmed the the TiO2 layer, which was consistent with the (101) plane of
elementary composition with 7.95 wt % Co, 15.13 wt % Fe, anatase TiO2.48 The EDS mapping images (Figure 3c−g)
30.0 wt % Ti, and 31.85 wt % O (Table S1). showed the elemental composition and distribution of Co, Fe,

Figure 2. SEM image of the CoFe2O4@ hollow@ mesoporous TiO2 (a) and EDS elemental mapping (b−f) of the CoFe2O4@ hollow@
mesoporous TiO2 samples: Ti (c), Fe (d), O (e), and Co (f).

19941 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 4. FTIR spectrum (a), XPS survey (b), high-resolution Co 2p spectrum (c), Fe 2p spectrum (d), O 1s spectrum (e), and Ti 2p spectrum (f)
of CoFe2O4@ hollow@ mesoporous TiO2.

Ti, and O in the CoFe2O4@ hollow@ mesoporous TiO2, removal of the silicon interlayer. This phenomenon further
which also confirmed the yolk−shell structure. demonstrated the existence of yolk−shell structure.
The FTIR spectra of CoFe2O4, CoFe2O4/SiO2/mesoporous XPS analysis was carried out to further investigate the
TiO2, and CoFe2O4@ hollow@ mesoporous TiO2 in the range surface elemental composition and the chemical state of
of 500−4000 cm−1 are given in Figure 4a. A broad stretching individual elements in the CoFe2O4@ hollow@ mesoporous
band at 3000−3500 cm−1 in these three curves represented the TiO2 catalyst with yolk−shell structure. As shown in Figure 4b,
stretching vibration of hydroxyl groups (−OH) and hydrogen the XPS survey confirmed the presence of the elements Co, Fe,
bonds owing to the ambient water from the surrounding O, and Ti in the CoFe2O4@ hollow@ mesoporous TiO2,
atmosphere.49,50 The characteristic peak of CoFe2O4 was which was consistent with the results of EDS analysis. The
detected in the CoFe2O4/SiO2/mesoporous TiO2 and high-resolution spectra of Co 2p, Fe 2p, O 1s, and Ti 2p are
CoFe2O4@ hollow@ mesoporous TiO2 samples. In addition, displayed in Figure 4c−f, respectively. In the Co 2p spectrum
a new weak band at 793.7 cm−1 was observed in these two (Figure 4c), two main signals of Co 2p3/2 and Co 2p1/2 were
observed at 779.2 eV (with a shakeup satellite peak at 785.7
samples. It was generally believed that the infrared spectrum at
eV) and 794.9 eV (with a shakeup satellite peak at 802.0 eV),
400−900 cm−1 was attributed to Ti−O vibrations.51 Mean-
respectively.53,54 Figure 4d shows the high-resolution XPS
while, from the curves of these three samples, two specific spectrum of Fe 2p, in which two spin−orbit doublets were
peaks of spinel ferrites could be detected in the range 500−750 attributed to Fe 2p3/2 and Fe 2p1/2. Especially, the fitted two
cm−1. The low band at about 535 cm−1 and the high band at peaks at 709.6 and 723.1 eV could be assigned to the
about 580 cm−1 were attributed to the vibration of metal− contributions from Fe3+ in the octahedral sites. In addition, the
oxygen bond at tetrahedral site and octahedral site, two peaks at 711.8 and 725.2 eV belonged to the contributions
respectively.52 In addition, the characteristic peak of Si−O from Fe3+ in the tetrahedral sites,55 and the other two peaks
was observed at 1065.6 cm−1 in the curve of CoFe2O4@SiO2@ were attributed to shakeup satellites.56 In the O 1s spectrum
mesoporous TiO2 sample,44 indicating the existence of SiO2. (Figure 4e), two peaks were fitted: one at 529.5 eV could be
Especially, the sharp peak at 1065.6 cm−1 only existed in the assigned to surface lattice oxygen in the metal−oxygen bond,
CoFe2O4@SiO2@ mesoporous TiO2 sample. No correspond- and the other one at 531.5 eV could be assigned to the
ing peak of Si−O was observed in the CoFe2O4@ hollow@ adsorbed with the intrinsic oxygen vacancies on the sur-
mesoporous TiO2 sample, further proving the successful face.57,58 With respect to the Ti 2p spectrum (Figure 4f), two
19942 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 5. Nitrogen adsorption−desorption isotherms (a) and the corresponding BJH pore size distributions (b) of the CoFe2O4@ hollow@
mesoporous TiO2 sample.

Scheme 2. Schematic Illustration of the Selectivity Preparation of BA (a) and SO (b)a

a
Reaction conditions: (a) 3 mmol of styrene, 8 mg of CoFe2O4@ hollow@ mesoporous TiO2, 2 mL of H2O, styrene/H2O2 = 1:3, 90 °C, 12 h; (b)
3 mmol of styrene, 8 mg of CoFe2O4/mesoporous TiO2, 2 mL of 1,4-dioxane, molecular oxygen, 90 °C, 8 h.

Figure 6. Effects of reaction temperature (a) and catalyst amount (b) on the catalytic performance of the CoFe2O4@ hollow@ mesoporous TiO2
yolk−shell catalyst. Reaction conditions: (a) 3 mmol of styrene, 8 mg of CoFe2O4@ hollow@ mesoporous TiO2, 2 mL of 1,4-dioxane, molecular
oxygen, 8 h; (b) 3 mmol of styrene, 2 mL of 1,4-dioxane, molecular oxygen, 90 °C, 8 h.

obvious fitted peaks were observed at 458.3 and 464.0 eV, Selective Oxidation of Styrene. Initially, the CoFe2O4@
which corresponded to Ti 2p3/2 and Ti 2p1/2 of TiO2.59,60 hollow@ mesoporous TiO2 yolk−shell catalyst was tested in
To further investigate the structure of the as-prepared the oxidation of styrene with either molecular oxygen or
CoFe2O 4@ hollow@ mesoporous TiO2 sample, the N 2 hydrogen peroxide as the oxidant (Scheme 2). Interestingly,
adsorption−desorption isotherms were carried out. As shown SO was found to be the main product with a styrene
in Figure 5a, the CoFe2O4@ hollow@ mesoporous TiO2 conversion of 81.3% and a SO selectivity of 67.3% in the
catalyst presented a type IV isotherm with a narrow hysteresis presence of molecular oxygen. When hydrogen peroxide was
loop, which was the characteristic curve in mesoporous used as the oxidant, 96.3% styrene conversion and 46.6%
materials.36 The surface area obtained by the Brunauer− selectivity for BA could be achieved.
Emmett−Teller method61 was 141.7 m2 g−1 for the CoFe2O4@ Molecular Oxygen Oxidation Route. To investigate the
effect of reaction temperature on catalytic performance,
hollow@ mesoporous TiO2, which was 2.2 times higher than
different temperatures in the range of 70−100 °C were
that of pure CoFe2O4 (65.5 m2 g−1) in the previous study.36
evaluated. The results are summarized in Figure 6a. Usually,
The high surface area of the CoFe2O4@ hollow@ mesoporous increased conversion was observed with higher reaction
TiO2 sample might be attributed to the novel yolk−shell temperatures.64 However, the conversion of styrene exhibited
structure. On the one hand, high surface area could facilitate a volcano-type relationship with respect to the reaction
the diffusion of reactants toward active sites. On the other temperature. On the one hand, high temperature would
hand, high surface area would be beneficial to the uniform promote the proceeding of side reaction, such as ring-opening
distribution of catalytically active sites.2,62 Furthermore, the reaction of epoxides.65 On the other hand, high temperature
mean pore size obtained by the Barrett−Joyner−Halenda was also unfavorable for the dissolution of molecular oxygen
(BJH) method63 from the desorption branches of the into the reaction mixture. For the SO selectivity versus reaction
isotherms (Figure 5b) was 3.1 nm for the CoFe2O4@ hollow@ temperature, it displayed the same trend with styrene
mesoporous TiO2. conversion. More side reactions with the generation of more
19943 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

unwanted byproducts could occur with high temperatures.1,66 Table 1. Effect of the Addition of Urea on Catalytic
Taking both the catalytic selectivity and activity into Performancea
consideration, 90 °C was selected as the optimal reaction
selectivity (%)
temperature for further investigations.
The amount of catalyst from 0 to 12 mg was used to study entry urea (g) conversion (%) SO BzA BA others
the effect of catalyst amount on the SO selectivity and styrene 1 0 97.7 65.8 15.1 10.5 8.6
conversion. When the catalyst amount was increased from 0 to 2 0.025 96.3 66.7 19.9 10.2 3.2
8 mg, the conversion of styrene could be remarkably increased 3 0.05 98.7 65.9 17.3 13.1 3.7
from 13.7 to 81.3% due to more accessible active sites for the 4 0.075 87.2 70.2 21.4 8.4
oxidation reaction (Figure 6b). And the conversion showed a 5b 0.075 89.7 70.4 19.4 10.2
slight increase when the catalyst amount was further increased 6c 0.075 96.8 71.7 13.4 12.3 2.6
to 12 mg. However, the variation trend in the conversion of a
Reaction conditions: 3 mmol of styrene, 8 mg of CoFe2O4@
styrene was not similar to that in the selectivity for SO. The hollow@ mesoporous TiO2, 2 mL of 1,4-dioxane, molecular oxygen,
plotting of SO selectivity with respect to the catalyst amount 90 °C, 12 h. bThe reaction time was 18 h. cThe reaction time was 24
exhibited a volcano-shaped curve. The highest selectivity was h.
obtained when the amount of catalyst was 8 mg. This
phenomenon might be attributed to the increased magnetic
agglomeration of the catalyst with the increase in the amount Hydrogen Peroxide Oxidation Route. The influence of
of catalysts,64 leading to reduced interaction between reactants reaction temperature on the oxidation of styrene was evaluated
and active sites. Therefore, 8 mg of catalyst was chosen as the in the range of 70−100 °C. The results are listed in Figure 7a.
optimum amount of catalyst for the following research. It could be seen that high temperature did not favor high
Different solvents might interact with or modify the catalyst selectivity for BA. The selectivity for BA increased quickly from
surface, thus playing a vital role in the accessibility of active 0 to 46.6% when the temperature was increased from 70 to 90
sites.7 For this reason, various types of solvents were used to °C. With higher temperatures, the selectivity decreased to
study the effect of different solvents on the oxidation of 41.2%. With an increase in reaction temperature from 70 to 90
styrene. The results are illustrated in Table S2. Among all of °C, the conversion of styrene could be remarkably increased
these solvents, 1,4-dioxane was found to give the best catalytic from 12.9 to 96.3%. Subsequently, no obvious increase was
performance with a styrene conversion of 97.7% and a SO detected with a further increase in the temperature. Taking
selectivity of 65.8%. Other solvents, such as acetonitrile, both the catalytic activity and selectivity into consideration, 90
ethanol, tetrahydrofuran, and dimethylformamide, showed low °C was selected as the optimal reaction temperature for further
conversion of styrene (<35%) and selectivity for SO (<50%). research.
To our surprise, the oxidation of styrene was suppressed when Similar to the molecular oxygen route, the amount of
some solvents were applied in the reaction (entries 2−4). catalyst from 0 to 12 mg was evaluated. As expected, the
Considering the relatively low boiling point of these solvents, conversion of styrene increased rapidly when the amount of
we performed the same experiments at a lower temperature catalyst ranged from 0 to 6 mg. Then, the conversion remained
(60 °C, entries 7−9). The results showed lower conversion. As constant even if the catalyst increased from 6 to 12 mg.
a kind of ether solvent, 1,4-dioxane could undergo autoxidation Interestingly, the BA selectivity exhibited a volcano-type
to produce unstable and dangerous hydroperoxide and relationship with respect to the catalyst amount. The highest
peroxides products.67 Therefore, 1,4-dioxane was considered selectivity was obtained when the amount of catalyst was 8 mg
to be the suitable solvent in the molecular oxygen oxidation (Figure 7b). The variation trend for the selectivity was similar
route. to the molecular oxygen route. The magnetic agglomeration
Liu et al.36 found that the basicity of acetonitrile might phenomenon might be the cause for this phenomenon.64 Thus,
promote the improvement in the SO selectivity, which inspired 8 mg was selected as the optimum amount of catalyst in the
researchers to introduce urea into the reaction system. To following investigations.
further increase the SO selectivity, various amounts of urea Table S3 shows the effect of different solvents on the
were added to the reaction mixture. The results are conversion of styrene and products distribution over the
summarized in Table 1. As observed, the addition of urea CoFe2O4@ hollow@ mesoporous TiO2 catalyst. H2O was
reduced the conversion of styrene inevitably (entry 4) due to found to give the highest conversion of styrene and selectivity
the stabilization effect of urea on the peroxides. Thus, extended for BA owing to the fact that the solvent with a high dielectric
reaction time was necessary to obtain high conversion (entries constant favored the oxidation of styrene.64 Other solvents
5 and 6). However, a slight increase in the SO selectivity was displayed low selectivity for BA (<25%) with high conversion
observed with the addition of urea. After further increasing the (>85%).
reaction time to 24 h, the yield of SO could reach 69.4% with a The effect of molar ratio of styrene to H2O2 on the catalytic
styrene conversion of 96.8% and SO selectivity of 71.7% (entry performance was investigated. The amount of styrene
6). remained constant when the amount of H2O2 was altered.
After systematically investigating the effects of reaction The selectivity for BA increased from 44 to 68.7% when the
parameters on the catalytic performance in the oxidation styrene/H2O2 molar ratio was altered from 1:2 to 1:4 (Figure
reaction of styrene, the optimal reaction conditions were listed 7c). However, the selectivity for BA dropped to 61.0% when
as follows: reaction temperature, 90 °C; reaction time, 24 h; 8 the molar ratio of styrene/H2O2 was further decreased to 1:5.
mg of CoFe2O4@ hollow@ mesoporous TiO2 catalyst; 0.075 g In addition, the conversion was almost 100%. When excessive
of urea; 1,4-dioxane as the solvent. Under these reaction H2O2 was added as the oxidant, large occupation of the active
conditions, the conversion of 96.8% was obtained with a sites by H2O2 would take place, resulting in few accessible
selectivity of 71.7%. active sites for styrene.36 Therefore, the ratio of 1:4 was
19944 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 7. Effects of reaction temperature (a), catalyst amount (b), and molar ratio of styrene/H2O2 (c) on the catalytic performance of the
CoFe2O4@ hollow@ mesoporous TiO2 catalyst. Reaction conditions: (a) 3 mmol of styrene, 8 mg of CoFe2O4@ hollow@ mesoporous TiO2, 2
mL of H2O, styrene/H2O2 = 1:3, 12 h; (b) 3 mmol of styrene, 2 mL of H2O, styrene/H2O2 = 1:3, 90 °C, 12 h; (c) 3 mmol of styrene, CoFe2O4@
hollow@ mesoporous TiO2, 2 mL of H2O, 90 °C, 24 h.

Figure 8. Recycling performance of the CoFe2O4@ hollow@ mesoporous TiO2 yolk−shell catalyst: (a) molecular oxygen oxidation route, reaction
conditions: 3 mmol of styrene, 8 mg of catalyst, 2 mL of 1,4-dioxane, molecular oxygen, 90 °C, 12 h; (b) hydrogen peroxide oxidation route,
reaction conditions: 3 mmol of styrene, 8 mg of CoFe2O4@ hollow@ mesoporous TiO2, 2 mL of H2O, styrene/H2O2 = 1:4, 90 °C, 24 h.

Figure 9. Evaluation of the catalytic performance of the CoFe2O4, CoFe2O4/TiO2 core−shell structure, TiO2, and CoFe2O4@ hollow@
mesoporous TiO2 yolk−shell structure. Reaction conditions: 3 mmol of styrene, 8 mg of catalyst, 2 mL of 1, 4-dioxane, molecular oxygen, 90 °C, 12
h.

selected as the optimal molar ratio in hydrogen peroxide Recycling Performance of the Catalyst. Catalyst
oxidation route. recyclability is an important parameter for heterogeneous
To further investigate the influence of reaction time on the catalysis.68 The CoFe2O4@ hollow@ mesoporous TiO2 with
oxidation of styrene with hydrogen peroxide as the oxidant, the yolk−shell structure could be easily separated from the
reaction time was increased from 12 to 30 h. As shown in reaction mixture by applying a magnetic field to the surface
Table S3, the selectivity for BA increased from 46.0 to 52.6% of the tube. The magnetic separation made the recovery of the
when the reaction time was increased from 12 to 24 h, which CoFe2O4@ hollow@ mesoporous TiO2 catalyst easier. As
was attributed to further oxidation of benzaldehyde. And the shown in Figure 8a, both the catalytic activity and SO
selectivity changed slightly after five consecutive runs with the
BA selectivity further showed a slight decrease when the
catalyst from the same batch, suggesting the good stability and
reaction time was further increased to 30 h. Thus, 24 h was
recyclable applicability of the CoFe2O4@ hollow@ mesopo-
selected as the optimum reaction time. rous TiO2 catalyst.
Last but not least, the effect of silicate residue in the etched The recyclability of the catalyst was also tested with
sample was also a factor that could not be ignored. Different hydrogen peroxide as the oxidant. Using the same recycle
amounts of sodium silicate were selected to add into the method in molecular oxygen oxidation route, the CoFe2O4@
reaction system. The results are summarized in Table S4. hollow@ mesoporous TiO2 catalyst with yolk−shell structure
Regardless of the oxygen oxidation route or the hydrogen was subjected to five consecutive runs. As shown in Figure 8b,
peroxide oxidation route, the silicate residue in the etched the conversion of styrene and the selectivity for BA changed
sample might have little effect on the reaction efficiency. slightly after five runs. The decrease in the selectivity of BA
19945 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

could be attributed mainly to unavoidable loss of the catalyst Table 2. Effect of Different Catalysts on Catalytic
during the process of collection. Performance
The TEM image (Figure S2) of the spent catalyst displayed
selectivity (%)
no obvious changes with those of fresh catalyst. All of these
results demonstrated that the CoFe2O4@ hollow@ meso- entry time (h) conversion (%) BA BzA SO
porous TiO2 catalyst with yolk−shell structure displayed good 1a 24 94.6 52.6 3.1
stability and recyclable applicability for the oxidation of styrene 2b 24 96.2 45.5 4.2
under optimal conditions. 3c 24 95.8 35.5 14.0
4d 24 97.0 36.4 6.9
Comparison between Core−Shell Structure and a
Yolk−Shell Structure. Reaction Efficiency. The Reaction conditions: 3 mmol of styrene, 8 mg of CoFe2O4@
hollow@ mesoporous TiO2, 2 mL of H2O, styrene/H2O2 = 1:3, 90
CoFe2O4@ hollow@ mesoporous TiO2 yolk−shell catalyst
°C. bThe same reaction conditions with (a), but using 8 mg of
was tested in the oxidation of styrene with molecular oxygen as CoFe2O4 as the catalyst. cThe same reaction conditions with (a), but
the oxidation. For comparison, CoFe2O4 nanoparticles, using 8 mg of CoFe2O4/TiO2 as the catalyst. dThe same reaction
CoFe2O4/TiO2 core−shell catalysts, and TiO2 were also tested conditions with (a), but using 8 mg of TiO2 as the catalyst.

ij C yz
−lnjjj zzz = kt
under the same reaction conditions. SO was found to be the

j C0 z
main product. Figure 9a shows the plot of styrene conversion
k {
versus reaction time. For all of the four catalysts, the
(4)
conversion of styrene increased with extended reaction time.
Especially, the CoFe2O4@ hollow@ mesoporous TiO2 catalyst where t is the reaction time; k is the apparent rate constant;
with yolk−shell structure exhibited the highest initial activity. and C0 and C are the initial concentration of styrene and the
The conversion of styrene quickly reached 58.3% in the first 4 concentration of styrene at the time t, respectively.36 From the
h. Then, the conversion increased continuously to nearly 100% slopes of these fitted lines, the k values for the reaction
within a total reaction time of 16 h. Interestingly, the performed at 70, 80, 90, and 100 °C could be calculated to be
conversion of styrene catalyzed by CoFe2O4@ hollow@ 0.1873, 0.2968, 0.3487, and 0.2002 h−1, respectively (molec-
mesoporous TiO2 displayed the highest level compared with ular oxygen oxidation route). The k value at 100 °C decreased
other three catalysts. However, the selectivity for SO in Figure significantly since the temperature was close to the boiling
9b did not show the same trend as the conversion of styrene. point of 1,4-dioxane, making it difficult for molecular oxygen to
For CoFe2O4@ hollow@ mesoporous TiO2 catalyst with interact with 1,4-dioxane. For the hydrogen peroxide oxidation
yolk−shell structure, the selectivity for SO remained constant route, the k values could be calculated to be 0.110, 0.2385, and
(60−65%). However, the selectivity for SO increased rapidly in 1.0639 h−1 with respect to 80, 90, and 100 °C, respectively.
the first 4 h and then decreases when other three catalysts were Due to the low conversion of styrene at 80 °C in the hydrogen
applied in the oxidation reaction. The decrease in the peroxide oxidation route, reaction time from 6 to 10 h was
selectivity for SO was mainly due to the overoxidation of SO chosen to investigate the apparent rate constant (Figure 10c,d,
inset). All of the k values are collected in Table S5.
and the hydrolysis of SO into 1-phenyl-1,2-ethanediol.36,66
Further, the apparent activation energy (Ea) could be
This phenomenon indicated that CoFe 2O4@ hollow@ calculated by the Arrhenius equation
mesoporous TiO2 displayed better catalytic activity toward
the epoxidation. The yolk−shell structure was beneficial to the Ea
formation of epoxides. On the contrary, the catalysts in the ln k = ln A −
RT (5)
control experiments showed catalytic activity toward epox-
idation reaction and ring-opening reaction, which was the where R is the molar gas constant, T is the reaction
common phenomenon in the design of catalysts for temperature, and A is the preexponential factor. From the
epoxidation reaction. When TiO2, CoFe2O4, and CoFe2O4/ slope of the fitted line with respect to ln k versus 1000/T
TiO2 were used as the catalyst, process optimization was (Figure 10e), the Ea value for the oxidation reaction of styrene
necessary for the high SO selectivity. Correspondingly, the using molecular oxygen oxidation route could be calculated to
yield of SO increased with extended reaction time and reached be 32.3 kJ/mol, which was lower than the previously reported
the platform after 12 h (Figure 9c). Taking the reaction time data (42.4 and 36.4 kJ/mol, respectively).2,36 And the Ea value
into consideration, the high yield of SO for CoFe2O4@ for the oxidation reaction of styrene using hydrogen peroxide
hollow@ mesoporous TiO2 yolk−shell catalyst was 64.3% with oxidation route is 39.9 kJ/mol (Figure 10e).
a styrene conversion of 97.7% and SO selectivity of 65.8%. In addition, the apparent activation energies of CoFe2O4/
TiO2 were also calculated to be 67.9 and 66.7 kJ/mol,
For the hydrogen peroxide route, CoFe2O4, CoFe2O4/TiO2,
respectively (Figure 10f). An obvious decrease in the activation
and TiO2 (8 mg) were also tested under the same reaction
energies was observed when the CoFe2O4@ hollow@
conditions (Table 2, entries 2−4), which also showed lower mesoporous TiO2 yolk−shell structure was formed (Table
catalytic activity compared to CoFe2O4@ hollow@ meso- 3). On the one hand, CoFe2O4@ hollow@ mesoporous TiO2
porous TiO2. yolk−shell structure with a large specific surface area was
Reaction Kinetics. Kinetics studies of the CoFe2O4@ obtained through etching of silicon layer, promoting the
hollow@ mesoporous TiO2 catalyst with yolk−shell structure interaction between catalysts and reactants. On the other hand,
at various temperatures showed that pseudo-first-order kinetics the interaction between active oxygen species and metal
could be applied for the evaluation of the catalytic rate (Figure cations would be realized in the CoFe2O4@ hollow@
10a,c). The kinetic equation for the reaction could be mesoporous TiO2, resulting in the selective oxidation of
expressed as follows styrene and the improvement in the catalytic activity.
19946 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 10. ln (C/C0) versus reaction time and Arrhenius plot of the styrene oxidation reactions performed at different temperatures: molecular
oxygen oxidation route (a, b, e, f); hydrogen peroxide oxidation route (c, d, e, f).

Table 3. Apparent Activation Energies of Different Catalysts peroxy radical. The hydroperoxide was formed when the
Using Different Oxidants peroxy radical abstracted a hydrogen atom from another 1,4-
dioxane molecule. Subsequently, the metal (Co, Fe, and Ti)
oxidation Ea
entry route catalysts (kJ/mol) sites coordinated with the activated hydroperoxide molecules,
1 molecular CoFe2O4/TiO2 67.9
yielding a metal-peroxo species (I) and 1,4-dioxane. The
2 oxygen CoFe2O4@ hollow@ mesoporous TiO2 32.3
complex (I) would attack the CC bonds of the styrene
route molecules to give an intermediate (II). Migratory insertion of
3 hydrogen CoFe2O4/TiO2 66.7 the intermediate (II) would afford metal-peroxo metallocycles
4 peroxide CoFe2O4@ hollow@ mesoporous TiO2 39.9
route (III). Four-membered cyclic peroxide radical (IV) was
generated along with the simultaneous release of metal species.
Reaction Mechanism. Depending on the reaction data and The obtained peroxide intermediate (IV) further reacted with
previous studies,2,5,36,62 mechanistic research of the oxidation another styrene to form styrene oxide. In addition, thermal
reaction was investigated. Superoxide radical scavenger decomposition of unstable peroxide intermediate (IV) also
molecules, benzoquinone (BQ) and 4-hydroxy-2,2,6,6-tetra- could occur to afford benzaldehyde.
methyl-piperidinooxy (TEMPOL), were used to check if For the hydrogen peroxide oxidation route, the mechanism
superoxide radical (O·− 2 ) was involved in the reaction. As
proposed here was similar to the mechanism proposed by
expected, introduction of radical scavengers suppressed the Ghosh et al.15 First, active Ti4+ centers interacted with H2O2,
reaction obviously whether it was molecular oxygen oxidation affording a metal-peroxo intermediate (named as Intermediate-
route or hydrogen peroxide oxidation route (Table S6). These I). Then, two types of oxametallacycle intermediates (named
results suggested that a superoxide radical intermediate was Intermediate-IIa and Intermediate-IIb) were formed through
involved in the reaction pathway. the reaction between styrene and Intermediate-I. With the
The possible reaction pathways for the two different transfer of oxygen to the double bond of styrene, styrene oxide
oxidation routes catalyzed by CoFe2O4@ hollow@ mesopo- was formed. In addition, benzaldehyde would be formed
rous TiO2 catalyst with yolk−shell structure were proposed through the breakage of C−C bond of Intermediate-IIb.
and are illustrated in Scheme 3. In the molecular oxygen Further oxidation of benzaldehyde would promote the
oxidation route, an active hydrogen atom in 1,4-dioxane was formation of benzoic acid.
abstracted, affording a 1,4-dioxane radical. Then, the resulting Comparison Our Method with Other Reported
1,4-dioxane radical reacted with molecular oxygen to yield the Methods for the Oxidation of Styrene. To give a
19947 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Scheme 3. Schematic Illustration of the Possible Reaction Mechanism for the Styrene Oxidation Reaction with the CoFe2O4@
Hollow@ Mesoporous TiO2 Yolk−Shell Catalysta

a
(a) Molecular oxygen oxidation route; (b) hydrogen peroxide oxidation route.

Table 4. Catalytic Oxidation of Styrene into SO or BA Using Various Methods


yield
entry catalyst oxidant T (°C) styrene conversion (%) SO (%) BA (%) references
1 Pd@Co TBHP 50 100 75 69
2 Co-Y-ZrO2 air 120 >99 68 70
3 MnFe-based hydrotalcite H2O2 50 19.6 10.6 71
4 CoFe2O4 TBHP 80 96.4 82.7 36
5 CoFe2O4@ hollow@ mesoporous TiO2 oxygen balloon 90 96.8 69.4 11.9 this work
6 nano-HKUST-1 TBHP 80 99 75 72
7 [3-ampH]4[{Zn(3-ampy)(H2O)4}Mo7O24]·4H2O H2O2 80 100 3.8 69 73
8 CoFe2O4@ hollow@ mesoporous TiO2 H2O2 90 94.5 68.7 this work

reasonable evaluation of our developed method for the obtained by adjusting oxidant and the reaction solvent when
synthesis of SO and BA from styrene, the developed method the same catalyst was used.
in this work was compared with other recently reported
methods. As shown in Table 4, various green oxidants, such as
TBHP, H2O2, and air, have been reported for the catalytic
■ CONCLUSIONS
In conclusion, a novel catalyst CoFe2O4@ hollow@ meso-
oxidation of styrene into SO or BA. Although higher styrene porous TiO2 catalyst with yolk−shell structure has been
conversion and SO yield could be obtained (entry 1), Pd-based developed and applied in the selective oxidation of styrene.
catalysts may not be applied on a large scale in industry. Other With molecular oxygen as the oxidant, styrene oxide was
non-noble catalysts also had some distinct drawbacks such as selectively obtained. With hydrogen peroxide as the oxidant,
the need for high temperature (entry 2) and low catalytic benzoic acid was selectively synthesized. Under the optimum
activity (entry 3). Owing to the stronger oxidation of TBHP, reaction conditions, the yield of SO could reach 69.4% with a
styrene conversion of 96.8% and a SO selectivity of 71.7%.
pure CoFe2O4 catalyst gave a yield of 82.7% for SO (entry 4).
Furthermore, the yield of BA reached 64.9% with a styrene
In this study, molecular oxygen or hydrogen peroxide at conversion of 94.5% and a BA selectivity of 68.7%. In addition,
ambient pressure, accompanying medium reaction temper- the yolk−shell structure displayed higher reaction conversion
ature, was used (entries 5 and 8), avoiding noble metal and and selectivity compared with the normal core−shell structure.
high temperature. Furthermore, the used catalyst could be Through calculation, yolk−shell CoFe2O4@ hollow@ meso-
quickly separated from the reaction system with an external porous TiO2 displayed lower Ea values in the oxidation of
magnetic field. Importantly, SO and BA could be selectively styrene compared with core−shell CoFe2O4/TiO2. Interest-
19948 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research


pubs.acs.org/IECR Article

ingly, the Ea values of yolk−shell CoFe2O4@ hollow@ ABBREVIATIONS


mesoporous TiO2 was approximately half of the Ea values of SO styrene oxide
core−shell CoFe2O4/TiO2 in two oxidation routes. The BzA benzaldehyde
addition of superoxide radical scavenger molecules revealed BA benzoic acid


that superoxide radical was involved in the oxidation reaction.
Corresponding mechanisms in these two routes were
proposed. With the attractive catalytic activity and product REFERENCES
selectivity, magnetic separation and reusability of the catalyst (1) Masunga, N.; Tito, G. S.; Meijboom, R. Catalytic evaluation of
would promote broad application in the chemical trans- mesoporous metal oxides for liquid phase oxidation of styrene. Appl.
formation. Catal., A 2018, 552, 154−167.


(2) Weerakkody, C.; Biswas, S.; Song, W.; He, J.; Wasalathanthri, N.;
Dissanayake, S.; Kriz, D. A.; Dutta, B.; Suib, S. L. Controllable
ASSOCIATED CONTENT synthesis of mesoporous cobalt oxide for peroxide free catalytic
* Supporting Information
sı epoxidation of alkenes under aerobic conditions. Appl. Catal., B 2018,
The Supporting Information is available free of charge at 221, 681−690.
https://pubs.acs.org/doi/10.1021/acs.iecr.0c03884. (3) Chen, L.; Cui, W.; Li, J.; Wang, H.; Dong, Xa.; Chen, P.; Zhou,
Y.; Dong, F. , The high selectivity for benzoic acid formation on
EDX information of CoFe2O4@ hollow@ mesoporous
Ca2Sb2O7 enables efficient and stable toluene mineralization. App.
TiO2, TEM image of the used CoFe2O4@ hollow@ Catal., B 2020, 271, No. 118948.
mesoporous TiO2, effect of different solvents on catalytic (4) Zhou, Y.; Sekar, B. S.; Wu, S.; Li, Z. Benzoic acid production via
performance, effect of sodium silicate on catalytic cascade biotransformation and coupled fermentation-biotransforma-
performance, a list of the apparent rate constants of tion. Biotechnol. Bioeng. 2020, 117, 2340−2350.
different catalysts, and the effect of radical scavenger on (5) Sebastian, J.; Jinka, K.; Jasra, R. Effect of alkali and alkaline earth
catalytic performance (PDF) metal ions on the catalytic epoxidation of styrene with molecular


oxygen using cobalt(II)-exchanged zeolite X. J. Catal. 2006, 244,
208−218.
AUTHOR INFORMATION (6) Pathan, S.; Patel, A. Transition-metal-substituted phosphomo-
Corresponding Author lybdates: Catalytic and kinetic study for liquid-phase oxidation of
Zhao Yang − College of Engineering, China Pharmaceutical styrene. Ind. Eng. Chem. Res. 2013, 52, 11913−11919.
University, Nanjing 211198, P. R. China; orcid.org/0000- (7) Sharma, A. S.; Kaur, H. Alloying of AuNPs with palladium: A
0003-1283-1947; Phone: +86-18013895668; promising tool for tuning of selectivity for epoxide in oxidation of
Email: yzcpu@cpu.edu.cn styrene using molecular oxygen. Appl. Catal., A 2017, 546, 136−148.
(8) Ren, J.; Liu, X.; Gao, R.; Dai, W.-L. Morphology and crystal-
Authors plane effects of Zr-doped CeO2 nanocrystals on the epoxidation of
Liang Liu − College of Engineering, China Pharmaceutical styrene with tert-butylhydroperoxide as the oxidant. J. Energy Chem.
University, Nanjing 211198, P. R. China 2017, 26, 681−687.
Wei He − College of Biotechnology and Pharmaceutical (9) Wang, H.; Xiao, X.; Liu, S.; Chiang, C. L.; Kuai, X.; Peng, C. K.;
Lin, Y. C.; Meng, X.; Zhao, J.; Choi, J.; Lin, Y. G.; Lee, J. M.; Gao, L.
Engineering, Nanjing Tech University, Nanjing 211816, P. R. Structural and electronic optimization of MoS2 edges for hydrogen
China evolution. J. Am. Chem. Soc. 2019, 141, 18578−18584.
Zheng Fang − College of Biotechnology and Pharmaceutical (10) Prabhu, P.; Jose, V.; Lee, J. M. Heterostructured catalysts for
Engineering, Nanjing Tech University, Nanjing 211816, P. R. electrocatalytic and photocatalytic carbon dioxide reduction. Adv.
China Funct. Mater. 2020, 30, No. 1910768.
Kai Guo − College of Biotechnology and Pharmaceutical (11) Li, H.; Hu, R.; Yang, P.; He, Y.; Feng, J.; Li, D. Synthesis of
Engineering and State Key Laboratory of Materials-Oriented efficient Ce modified CuO/CoAl-HT catalysts for styrene epox-
Chemical Engineering, Nanjing Tech University, Nanjing idation. Catal. Lett. 2018, 148, 1589−1596.
211816, P. R. China; orcid.org/0000-0002-0013-3263 (12) Choudhary, V. R.; Dumbre, D. K.; Patil, N. S.; Uphade, B. S.;
Zhixiang Wang − College of Engineering, China Pharmaceutical Bhargava, S. K. Epoxidation of styrene by t-butyl hydroperoxide over
gold nanoparticles supported on Yb2O3: Effect of gold deposition
University, Nanjing 211198, P. R. China method, gold loading, and calcination temperature of the catalyst on
Complete contact information is available at: its surface properties and catalytic performance. J. Catal. 2013, 300,
https://pubs.acs.org/10.1021/acs.iecr.0c03884 217−224.
(13) Jin, Y.; Wang, P.; Yin, D.; Liu, J.; Qiu, H.; Yu, N. Gold
Author Contributions nanoparticles stabilized in a novel periodic mesoporous organosilica
L.L. and W.H. contributed equally to this work. The of SBA-15 for styrene epoxidation. Microporous Mesoporous Mater.
2008, 111, 569−576.
manuscript was written through contributions of all authors.
(14) Ghosh, B. K.; Moitra, D.; Chandel, M.; Patra, M. K.; Vadera, S.
All authors have given approval to the final version of the R.; Ghosh, N. N. CuO nanoparticle immobilised mesoporous TiO2-
manuscript. cobalt ferrite nanocatalyst: A versatile, magnetically separable and
Notes reusable catalyst. Catal. Lett. 2017, 147, 1061−1076.
The authors declare no competing financial interest. (15) Ghosh, B. K.; Moitra, D.; Chandel, M.; Lulla, H.; Ghosh, N. N.

■ ACKNOWLEDGMENTS
This research was supported by The National Natural Science
Ag nanoparticle immobilized mesoporous TiO2-cobalt ferrite nano-
catalyst: A highly active, versatile, magnetically separable and reusable
catalyst. Mater. Res. Bull. 2017, 94, 361−370.
(16) Pinnaduwage, D. S.; Zhou, L.; Gao, W.; Friend, C. M. Chlorine
Foundation of China (21776130 and 21908094); the State promotion of styrene epoxidation on Au(111). J. Am. Chem. Soc.
Key Laboratory of Materials-Oriented Chemical Engineering 2006, 129, 1872−1873.
(KL19-01); and the Top-notch Academic Programs Project of (17) Rahman, S.; Santra, C.; Kumar, R.; Bahadur, J.; Sultana, A.;
Jiangsu Higher Education Institutions. Schweins, R.; Sen, D.; Maity, S.; Mazumdar, S.; Chowdhury, B.

19949 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Highly active Ga promoted Co-HMS-X catalyst towards styrene (34) Saire-Saire, S.; Barbosa, E. C. M.; Garcia, D.; Andrade, L. H.;
epoxidation reaction using molecular O2. Appl. Catal., A 2014, 482, Garcia-Segura, S.; Camargo, P. H. C.; Alarcon, H. Green synthesis of
61−68. Au decorated CoFe2O4 nanoparticles for catalytic reduction of 4-
(18) Chaudhary, V.; Sharma, S. Synthesis and catalytic activity of nitrophenol and dimethylphenylsilane oxidation. RSC Adv. 2019, 9,
SBA-15 supported catalysts for styrene oxidation. Chin. J. Chem. Eng. 22116−22123.
2018, 26, 1300−1306. (35) Anke, S.; Falk, T.; Bendt, G.; Sinev, I.; Hävecker, M.; Antoni,
(19) Brégeault, J.-M. Transition-metal complexes for liquid-phase H.; Zegkinoglou, I.; Jeon, H.; Knop-Gericke, A.; Schlögl, R.; Roldan
catalytic oxidation: some aspects of industrial reactions and of Cuenya, B.; Schulz, S.; Muhler, M. On the reversible deactivation of
emerging technologies. Dalton Trans. 2003, 3289−3302. cobalt ferrite spinel nanoparticles applied in selective 2-propanol
(20) Piché, D.; Tavernaro, I.; Fleddermann, J.; Lozano, J. G.; oxidation. J. Catal. 2020, 382, 57−68.
Varambhia, A.; Maguire, M. L.; Koch, M.; Ukai, T.; Hernandez (36) Liu, J.; Meng, R.; Li, J.; Jian, P.; Wang, L.; Jian, R. Achieving
Rodriguez, A. J.; Jones, L.; Dillon, F.; Reyes Molina, I.; Mitzutani, M.; high-performance for catalytic epoxidation of styrene with uniform
Gonzalez Dalmau, E. R.; Maekawa, T.; Nellist, P. D.; Kraegeloh, A.; magnetically separable CoFe2O4 nanoparticles. Appl. Catal., B 2019,
Grobert, N. Targeted T1 magnetic resonance imaging contrast 254, 214−222.
enhancement with extraordinarily small CoFe2O4 nanoparticles. ACS (37) Liu, J.; Meng, R.; Wang, H.; Jian, P. Boosting styrene
Appl. Mater. Interfaces 2019, 11, 6724−6740. epoxidation via CoMn2O4 microspheres with unique porous yolk-shell
(21) Georgiadou, V.; Makris, G.; Papagiannopoulou, D.; Vourlias, architecture and synergistic intermetallic interaction. J. Colloid
G.; Dendrinou-Samara, C. Octadecylamine-mediated versatile coating Interface Sci. 2020, 579, 221−232.
of CoFe2O4 NPs for the sustained release of anti-inflammatory drug (38) Yang, Z.; Pedireddy, S.; Lee, H. K.; Liu, Y.; Tjiu, W. W.; Phang,
naproxen and in vivo target selectivity. ACS Appl. Mater. Interfaces I. Y.; Ling, X. Y. Manipulating the d-band electronic structure of
2016, 8, 9345−9360. platinum-functionalized nanoporous gold bowls: Synergistic inter-
(22) Meidanchi, A.; Akhavan, O. Superparamagnetic zinc ferrite metallic interactions enhance catalysis. Chem. Mater. 2016, 28, 5080−
spinel-graphene nanostructures for fast wastewater purification. 5086.
Carbon 2014, 69, 230−238. (39) Kim, J. S.; Kim, B.; Kim, H.; Kang, K. Recent progress on
(23) Fu, G.; Lee, J. M. Ternary metal sulfides for electrocatalytic multimetal oxide catalysts for the oxygen evolution reaction. Adv.
energy conversion. J. Mater. Chem. A 2019, 7, 9386−9405. Energy Mater. 2018, 8, No. 1702774.
(24) Masa, J.; Weide, P.; Peeters, D.; Sinev, I.; Xia, W.; Sun, Z.; (40) He, W.; Zhu, G.; Gao, Y.; Wu, H.; Fang, Z.; Guo, K. Green
Somsen, C.; Muhler, M.; Schuhmann, W. Amorphous cobalt boride plasticizers derived from epoxidized soybean oil for poly (vinyl
(Co2B) as a highly efficient nonprecious catalyst for electrochemical chloride): Continuous synthesis and evaluation in PVC films. Chem.
water splitting: Oxygen and hydrogen evolution. Adv. Energy Mater. Eng. J. 2020, 380, No. 122532.
2016, 6, No. 1502313. (41) Wan, X.; Liu, J.; Wang, D.; Li, Y.; Wang, H.; Pan, R.; Zhang, E.;
(25) Guo, Y.; Tao, Y.; Ma, X.; Jin, J.; Wen, S.; Ji, W.; Song, W.; Zhang, X.; Li, X.; Zhang, J. From core-shell to yolk-shell: Keeping the
Zhao, B.; Ozaki, Y. A dual colorimetric and SERS detection of Hg2+ intimately contacted interface for plasmonic metal@semiconductor
based on the stimulus of intrinsic oxidase-like catalytic activity of Ag- nanorods toward enhanced near-infrared photoelectrochemical
CoFe2O4/reduced graphene oxide nanocomposites. Chem. Eng. J. performance. Nano Res. 2020, 13, 1162−1170.
2018, 350, 120−130. (42) Hu, J.; Li, R.; Han, J.; Sun, J.; Wang, Y.; Yu, L.; Guo, R. Yolk-
(26) Behera, A.; Kandi, D.; Sahoo, S.; Parida, K. Construction of shell or yolk-in-shell nanocatalysts? A proof-of-concept study. J. Mater.
isoenergetic band alignment between CdS QDs and CaFe2O4@ Chem. A 2020, 8, 10217−10225.
ZnFe2O4 heterojunction: A promising ternary hybrid toward (43) Chen, L.; Zuo, X.; Zhou, L.; Huang, Y.; Yang, S.; Cai, T.; Ding,
norfloxacin degradation and H2 energy production. J. Phys. Chem. C D. Efficient heterogeneous activation of peroxymonosulfate by facilely
2019, 123, 17112−17126. prepared Co/Fe bimetallic oxides: Kinetics and mechanism. Chem.
(27) Li, X.; Liu, X.; Lin, C.; Zhang, H.; Zhou, Z.; Fan, G.; Ma, J. Eng. J. 2018, 345, 364−374.
Cobalt ferrite nanoparticles supported on drinking water treatment (44) Huang, J.; Jing, H. X.; Li, N.; Li, L. X.; Jiao, W. Z. Fabrication
residuals: An efficient magnetic heterogeneous catalyst to activate of magnetically recyclable SnO2-TiO2/CoFe2O4 hollow core-shell
peroxymonosulfate for the degradation of atrazine. Chem. Eng. J. 2019, photocatalyst: Improving photocatalytic efficiency under visible light
367, 208−218. irradiation. J. Solid State Chem. 2019, 271, 103−109.
(28) Liu, P.; Ren, Y.; Ma, W.; Ma, J.; Du, Y. Degradation of shale gas (45) Chi, Y.; Yuan, Q.; Li, Y.; Zhao, L.; Li, N.; Li, X.; Yan, W.
produced water by magnetic porous MFe2O4 (M = Cu, Ni, Co and Magnetically separable Fe3O4@SiO2@TiO2-Ag microspheres with
Zn) heterogeneous catalyzed ozone. Chem. Eng. J. 2018, 345, 98−106. well-designed nanostructure and enhanced photocatalytic activity. J.
(29) Li, R.; Cai, M.; Xie, Z.; Zhang, Q.; Zeng, Y.; Liu, H.; Liu, G.; Hazard. Mater. 2013, 262, 404−411.
Lv, W. Construction of heterostructured CuFe2O4/g-C3N4 nano- (46) Gao, Y.; Chen, B.; Li, H.; Ma, Y. Preparation and
composite as an efficient visible light photocatalyst with peroxydi- characterization of a magnetically separated photocatalyst and its
sulfate for the organic oxidation. Appl. Catal., B 2019, 244, 974−982. catalytic properties. Mater. Chem. Phys. 2003, 80, 348−355.
(30) Sheykhan, M.; Mohammadnejad, H.; Akbari, J.; Heydari, A. (47) Yu, Y.; Yan, L.; Cheng, J.; Jing, C. Mechanistic insights into
Superparamagnetic magnesium ferrite nanoparticles: a magnetically TiO2 thickness in Fe3O4 @TiO2-GO composites for enrofloxacin
reusable and clean heterogeneous catalyst. Tetrahedron Lett. 2012, 53, photodegradation. Chem. Eng. J. 2017, 325, 647−654.
2959−2964. (48) Liu, G.; Yang, H. G.; Pan, J.; Yang, Y. Q.; Lu, G. Q.; Cheng, H.
(31) Zhang, X.; Junhui, Y.; Jing, Y.; Ting, C.; Bei, X.; Zhe, L.; M. Titanium dioxide crystals with tailored facets. Chem. Rev. 2014,
Kunfeng, Z.; Ling, Y.; Dannong, H. Excellent low-temperature 114, 9559−9612.
catalytic performance of nanosheet Co-Mn oxides for total benzene (49) Abbas, T.; Tahir, M.; Saidina Amin, N. A. Enhanced metal−
oxidation. Appl. Catal., A 2018, 566, 104−112. support interaction in Ni/Co3O4/TiO2 nanorods toward stable and
(32) Ding, L.; Yang, W.; Chen, L.; Cheng, H.; Qi, Z. Fabrication of dynamic hydrogen production from phenol steam reforming. Ind. Eng.
spinel CoMn2O4 hollow spheres for highly selective aerobic oxidation Chem. Res. 2019, 58, 517−530.
of 5-hydroxymethylfurfural to 2,5-diformylfuran. Catal. Today 2020, (50) Cai, K.; Shen, W.; Ren, B.; He, J.; Wu, S.; Wang, W. A phytic
347, 39−47. acid modified CoFe2O4 magnetic adsorbent with controllable
(33) Tan, X.; Wang, X.; Liu, Q.; Zhou, J.; Zhang, P.; Zheng, S.; morphology, excellent selective adsorption for dyes and ultra-strong
Miao, S. Bio-gel template synthesis of CoFe2O4 nano-catalysts and adsorption ability for metal ions. Chem. Eng. J. 2017, 330, 936−946.
application in aerobic oxidation of cyclohexane. Int. J. Hydrogen Energy (51) Ibrahim, I.; Athanasekou, C.; Manolis, G.; Kaltzoglou, A.;
2017, 42, 19001−19009. Nasikas, N. K.; Katsaros, F.; Devlin, E.; Kontos, A. G.; Falaras, P.

19950 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Photocatalysis as an advanced reduction process (ARP): The hydrogenation reduction of p-nitrophenol. J. Colloid Interface Sci.
reduction of 4-nitrophenol using titania nanotubes-ferrite nano- 2017, 497, 102−107.
composites. J. Hazard. Mater. 2019, 372, 37−44. (69) Jain, R.; Gopinath, C. S. New strategy toward a dual functional
(52) Zhu, T.; Chang, S.; Song, Y.-F.; Lahoubi, M.; Wang, W. PVP- nanocatalyst at ambient conditions: Influence of the Pd-Co interface
encapsulated CoFe2O4/rGO composites with controllable electro- in the catalytic activity of Pd@Co core-shell nanoparticles. ACS Appl.
magnetic wave absorption performance. Chem. Eng. J. 2019, 373, Mater. Interfaces 2018, 10, 41268−41278.
755−766. (70) Tyagi, B.; Shaik, B.; Bajaj, H. C. Microwave-assisted
(53) Yan, K.-L.; Qin, J.-F.; Lin, J.-H.; Dong, B.; Chi, J.-Q.; Liu, Z.-Z.; epoxidation of styrene with molecular O2 over sulfated Co-Y-ZrO2
Dai, F.-N.; Chai, Y.-M.; Liu, C.-G. Probing the active sites of Co3O4 solid catalyst. Catal. Commun. 2009, 11, 114−117.
for the acidic oxygen evolution reaction by modulating the Co2+/Co3+ (71) de Brito, S. N. A.; Pinheiro, L. G.; Filho, J. M.; Oliveira, A. C.
ratio. J. Mater. Chem. A 2018, 6, 5678−5686. Studies on styrene selective oxidation over iron-based catalysts:
(54) Zhou, L.; Ji, L.; Ma, P. C.; Shao, Y.; Zhang, H.; Gao, W.; Li, Y. Reaction parameters effects. Fuel 2015, 150, 305−317.
Development of carbon nanotubes/CoFe2O4 magnetic hybrid (72) Zhang, J.; Xiao, D.; Tan, H.; Liu, W. Highly selective synthesis
material for removal of tetrabromobisphenol A and Pb(II). J. Hazard. of 2-tert-butoxy-1-arylethanones via copper(I)-catalyzed oxidation/
Mater. 2014, 265, 104−114. tert-butoxylation of aryl olefins with TBHP. J. Org. Chem. 2020, 85,
(55) Zhou, Z.; Zhang, Y.; Wang, Z.; Wei, W.; Tang, W.; Shi, J.; 3929−3935.
Xiong, R. Electronic structure studies of the spinel CoFe2O4 by X-ray (73) Amanchi, S. R.; Patel, A.; Das, S. K. Polyoxometalate
photoelectron spectroscopy. Appl. Surf. Sci. 2008, 254, 6972−6975. coordinated transition metal complexes as catalysts: Oxidation of
(56) Huang, Q.; Zhou, P.; Yang, H.; Zhu, L.; Wu, H. In situ styrene to benzaldehyde/benzoic acid. J. Chem. Sci. 2014, 126, 1641−
generation of inverse spinel CoFe2O4 nanoparticles onto nitrogen- 1645.
doped activated carbon for an effective cathode electrocatalyst of
microbial fuel cells. Chem. Eng. J. 2017, 325, 466−473.
(57) Li, J.; Xu, M.; Yao, G.; Lai, B. Enhancement of the degradation
of atrazine through CoFe2O4 activated peroxymonosulfate (PMS)
process: Kinetic, degradation intermediates, and toxicity evaluation.
Chem. Eng. J. 2018, 348, 1012−1024.
(58) Gong, Y.; Wu, Y.; Xu, Y.; Li, L.; Li, C.; Liu, X.; Niu, L. All-solid-
state Z-scheme CdTe/TiO2 heterostructure photocatalysts with
enhanced visible-light photocatalytic degradation of antibiotic waste
water. Chem. Eng. J. 2018, 350, 257−267.
(59) Bajorowicz, B.; Nadolna, J.; Lisowski, W.; Klimczuk, T.;
Zaleska-Medynska, A. The effects of bifunctional linker and reflux
time on the surface properties and photocatalytic activity of CdTe
quantum dots decorated KTaO3 composite photocatalysts. Appl.
Catal., B 2017, 203, 452−464.
(60) Challagulla, S.; Nagarjuna, R.; Ganesan, R.; Roy, S. Acrylate-
based polymerizable sol−gel synthesis of magnetically recoverable
TiO2 supported Fe3O4 for Cr(VI) photoreduction in aerobic
atmosphere. ACS Sustainable Chem. Eng. 2016, 4, 974−982.
(61) Liu, J.; Fang, S.; Jian, R.; Wu, F.; Jian, P. Silylated Pd/Ti-MCM-
41 catalyst for the selective production of propylene oxide from the
oxidation of propylene with cumene hydroperoxide. Powder Technol.
2018, 329, 19−24.
(62) Liu, J.; Chen, T.; Jian, P.; Wang, L.; Yan, X. Hollow urchin-like
NiO/NiCo2O4 heterostructures as highly efficient catalysts for
selective oxidation of styrene. J. Colloid Interface Sci. 2018, 526,
295−301.
(63) Liu, J.; Wang, Z.; Yan, X.; Jian, P. Metallic cobalt nanoparticles
imbedded into ordered mesoporous carbon: A non-precious metal
catalyst with excellent hydrogenation performance. J. Colloid Interface
Sci. 2017, 505, 789−795.
(64) Tong, J.; Li, W.; Bo, L.; Wang, H.; Hu, Y.; Zhang, Z.;
Mahboob, A. Selective oxidation of styrene catalyzed by cerium-
doped cobalt ferrite nanocrystals with greatly enhanced catalytic
performance. J. Catal. 2016, 344, 474−481.
(65) Zhan, W.; Guo, Y.; Wang, Y.; Liu, X.; Guo, Y.; Wang, Y.;
Zhang, Z.; Lu, G. Synthesis of lanthanum-doped MCM-48 molecular
sieves and its catalytic performance for the oxidation of styrene. J.
Phys. Chem. B 2007, 111, 12103−12110.
(66) Pardeshi, S. K.; Pawar, R. Y. SrFe2O4 complex oxide an effective
and environmentally benign catalyst for selective oxidation of styrene.
J. Mol. Catal., A 2011, 334, 35−43.
(67) Hu, E.; Chen, Y.; Zhang, Z.; Chen, J.-Y.; Huang, Z. Ab initio
calculation and kinetic modeling study of diethyl ether ignition with
application toward a skeletal mechanism for CI engine modeling. Fuel
2017, 209, 509−520.
(68) Liu, J.; Yan, X.; Wang, L.; Kong, L.; Jian, P. Three-dimensional
nitrogen-doped graphene foam as metal-free catalyst for the

19951 https://dx.doi.org/10.1021/acs.iecr.0c03884
Ind. Eng. Chem. Res. 2020, 59, 19938−19951

You might also like