[go: up one dir, main page]

AppliedSurfaceScience6762024160971 (1)

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

Applied Surface Science 676 (2024) 160971

Contents lists available at ScienceDirect

Applied Surface Science


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

Full Length Article

Influence of iron doping in mesoporous ceria on the physicochemical


properties and catalytic activity in styrene oxidation
Kamma Yogendra a, b , Palli Sitaramulu a , Silligandla Nazeer a, b , Palnati Manoj Kumar a ,
Benjaram Mahipal Reddy c , Tumula Venkateshwar Rao a, b, *
a
Department of Catalysis and Fine Chemicals, CSIR-Indian Institute of Chemical Technology, Uppal Road, Hyderabad 500007, India
b
Academy of Scientific and Innovative Research (AcSIR), Ghaziabad 201002, India
c
Department of Chemistry and Materials Center for Sustainable Energy & Environment, Birla Institute of Technology & Science (BITS) Pilani, Hyderabad Campus,
Hyderabad 500078, India

A R T I C L E I N F O A B S T R A C T

Keywords: A series of nanosized Fe-Ce binary oxides (2, 5, 10, 20, and 30 wt% Fe) and pure oxides of Fe and Ce were
Styrene oxidation prepared by the template-assisted coprecipitation method and evaluated for styrene oxidation. The synthesised
Ceria-iron solid solutions binary oxides were investigated by Powder X-ray diffraction (PXRD), N2-physisorption, X-ray fluorescence (XRF),
Iron-Ceria binary oxide
Field emission scanning electron microscopy (FE-SEM), Highresolution transmission electron microscopy (HR-
Iron doped ceria
TBHP
TEM), Raman spectroscopy, Fourier transform infrared spectroscopy (FT-IR), Ultraviolet–Visible diffuse reflec-
Styrene tance spectroscopy (UV–Vis DRS), X-ray photoelectron spectroscopies (XPS), and Hydrogen-temperature pro-
Styrene oxide grammed reduction (H2-TPR). In particular, the influence of quantity of doped iron on the structure, phase
composition, surface morphology, reducibility, and the catalytic performance of ceria is deeply investigated. All
the synthesized Fe-doped ceria catalysts exhibited the formation of solid solutions, but a little α-Fe2O3 segregated
on the surface of ceria at 20 and 30 wt% Fe doped ceria. Among all Fe-Ce binary oxides, the 10 wt% Fe-doped
ceria catalyst demonstrated superior efficiency due to its optimal Fe content which achieved the high surface
area, lattice oxygens, reducibility, and low crystallinity, as confirmed by BET (Brunauer-Emmett-Teller) mea-
surements, XPS, H2-TPR and powder XRD analysis, respectively. To optimize the reaction conditions, effect of
several experimental variables, such as reaction temperature, duration, catalyst quantity, styrene/tert-Butyl
hydroperoxide (TBHP) mole ratio, and solvent polarity were investigated. Furthermore, the catalyst stability and
consistent activity were confirmed by testing six reusability cycles.

1. Introduction are used to synthesize various chemicals like β-amino alcohols, alcohols,
1,2-diols, carbonyl compounds, etc [5]. Particularly, styrene oxide is
The scope of selective oxidation of organic molecules is vast, ranging used as a raw material in the production of various commercial prod-
from small-scale pharmaceutical and fine chemical synthesis to large- ucts, including epoxy resins, and is a key intermediate in the synthesis of
scale commodity chemical manufacturing. Selective oxidation pro- phenylethyl alcohol and its derivatives, which are widely utilized in the
cesses are complicated and difficult to manage or stop at a particular production of surface coatings, cosmetics, and textiles [6]. The oxidative
stage when compared to other chemical procedures. Therefore, the se- transformation of styrene is challenging as it contains terminal double
lective oxidation is a potential field of research for obvious reasons [1]. bond. Traditionally, styrene is epoxidized in two different ways: one is
The selective oxidative transformation of olefins is considered as an the dehydrohalogenation of styrene halohydrin with base [7], and the
important reaction, as the oxygenated products of the olefins are key other is the epoxidation of styrene with stoichiometric peracids, which
intermediates to produce fine chemicals, drugs, fragrances, pharma- are environmentally hazardous and expensive [8]. To overcome envi-
ceuticals, and agrochemicals [2–4]. Among olefin oxidation products, ronmental constraints, greener oxidants like TBHP, H2O2, and molecular
oxiranes are the key intermediates in synthetic organic chemistry that oxygen (O2) were employed in styrene epoxidation. TBHP and H2O2 are

* Corresponding author at: Department of Catalysis & Fine Chemicals, CSIR-Indian Institute of Chemical Technology (CSIR-IICT), Hyderabad, Telangana 500007,
India.
E-mail address: tvrao@iict.res.in (T.V. Rao).

https://doi.org/10.1016/j.apsusc.2024.160971
Received 28 May 2024; Received in revised form 20 July 2024; Accepted 11 August 2024
Available online 12 August 2024
0169-4332/© 2024 Elsevier B.V. All rights are reserved, including those for text and data mining, AI training, and similar technologies.
K. Yogendra et al. Applied Surface Science 676 (2024) 160971

Fig. 1. Powder XRD patterns of pure CeO2, 2FC, 5FC, 10FC, 20FC, 30FC, and Fe2O3 (a) and enlarged most intense reflection peak (111 plane) (b).

eco-friendly oxidants since t-butanol and water are their byproducts, environmental chemistry. Ceria has fluorite-structure and can easily
respectively, which are safe and easily separatable. O₂ is the preferred form oxygen vacancies (Ov) or defects in the lattice because of the
oxidant because it is easily accessible and advantageous for both envi- reduction of Ce4+ to Ce3+. These Ov can be filled by surrounding oxygen,
ronmental and economic reasons [5]. According to the literature, TBHP reverting cerium to its Ce4+ state. Because of its effortless redox
is the most widely used oxidant for styrene epoxidation because it yields behaviour (Ce4+ ↔ Ce3+) ceria acts as an excellent oxygen buffer. The
fewer side products compared to H2O2 and O2 [9]. Ov facilitates the diffusion of lattice oxygens, providing pathways for
Many systems were developed for epoxidation of styrene in the oxygen atoms to move through the material [27,28]. This mobility is
presence of homogenous catalysts. However, various difficulties, such as crucial for the catalytic activity of ceria, allowing continuous supply and
separation and reuse of catalyst, benzaldehyde as a major product, and removal of oxygen during catalytic processes, thus maintaining the
high process costs, resulted in industrially unviable [1]. As a result, material’s activity and stability. In catalytic reactions, it follows the
various heterogeneous catalytic systems were explored in the oxidative Mars-van Krevelen mechanism: a reactant molecule reacts with a lattice
transformation of styrene due to their ease of catalyst separation and oxygen, creating an Ov, which is then replenished by surrounding ox-
recyclability. Accordingly, the styrene oxidation with H2O2 in organic ygen, restoring the lattice structure, and enabling continuous catalytic
solvents over the Ti-containing silica-based heterogeneous catalyst was cycles [29]. The lattice-defect CeO2 fluorite structure is dynamic,
investigated and found to produce selectively phenylacetaldehyde [1]. changing spontaneously or in response to external variables like tem-
The use of transition-metal (Mn to Zn) ferrite nanoparticles [10], Fe2O3 perature, oxygen partial pressure, and doping of other ions [30]. The
coated platinum nanowires [11], cerium-doped cobalt ferrite [12], Mg- redox characteristics (Ce4+ ↔ Ce3+) and oxygen storage and release
Cu ferrite [13], V-MCM-48 [14], and nanosized MgxFe3− xO4 [15] cata- capability of CeO2 are significantly improved by the incorporation of
lysts was investigated for selective oxidation of styrene to benzaldehyde. transition elements or inner transition elements. Among them, due to
In contrast, nanocrystalline Ag/WO3 [16], Au nanoparticles on layered their various oxidation states and superior redox characteristics
double hydroxide [17], Au nanoparticles on dendrimer@resin [18], Au compared to inner transition elements, transition elements such as Mn,
nanoparticles supported on titania [19], robust core–shell microspheres Fe, Co, and Cu are prominent [31].
containing Au nanoparticles [20], Pd and Pt nanoparticles [21], Ag on Over the last few years, doped ceria has been employed as a catalyst
mesoporous silica [22], copper-containing hexagonal mesoporous silica for oxidation of vanillin alcohol [32], benzylamine [33], NO [34], CO
[23], CuO on CoAl-HT [24], Ni-based metasilicate [25], and Cu-Mn [35], diesel soot [36,37], formaldehyde [38], benzyl alcohol [39], eth-
binary oxides-based mesoporous silica [26] catalysts selectively pro- ylbenzene [40,41], cyclohexene [42], and styrene [4,5]. In recent times,
duce styrene epoxide. Most of the reported work on the styrene oxida- co-doped ceria [31] has also been employed in many catalytic oxidation
tion to styrene oxide product was based on noble metal catalysts. It is processes. As a result of its economic benefits, environmental friendli-
required to replace them with cheaper, abundant, non-noble metal- ness, abundance, and excellent redox behaviour, iron is considered one
based catalysts and use alternative, clean, cheap, and accessible oxidants of the most attractive dopants of ceria for various applications [43].
to develop economical and reliable technologies for the oxidation of Iron-cerium mixed oxide catalysts exhibit high activity in the photo-
styrene in the journey towards sustainable processes. catalytic degradation of organic pollutants [44]. Iron-doped ceria also
Cerium (IV) oxide (ceria) is an abundant rare earth metal oxide and shows a high adsorption ability for organic dyes, As5+, and Cr6+ from
an important material due to numerous applications in the fields of wastewaters [45], and superior catalytic activity for oxidation of CO
medicinal chemistry, industrial chemistry, biotechnology, and [46], vanillyl alcohol [47], and diesel soot [48].

2
K. Yogendra et al. Applied Surface Science 676 (2024) 160971

Table 1
XRF, Powder XRD and N2-Physisorption data of samples.
S. No Sample Composition (%)a Surface area (m2g− 1)b Pore size (nm)c Pore volume (cm3g− 1)c Crystallite size (nm)d

Fe Ce

1 Fe2O3 − − 24 19.7 0.123 −


2 CeO2 − − 50 11.1 0.096 13.15
3 2FC 2.4 97.6 61 8.1 0.12 7.98
4 5FC 5.6 94.4 84 8.3 0.20 7.41
5 10FC 10.3 89.7 97 11.2 0.25 7.22
6 20FC 17.4 82.6 81 12.6 0.23 7.45
7 30FC 27.2 72.8 75 13.6 0.22 7.5
a
From XRF Analysis, b From BET Analysis, c From BJH Analysis, and dFrom P-XRD Analysis.

Fig. 2. N2-physisorption isotherms (a) and Raman spectra (b) of pure CeO2, 2FC, 5FC, 10FC, 20FC, 30FC, and Fe2O3.

The present work has been carried out in the above context and cubic fluorite structure of CeO2 with Fm3m space group [49]. Powder
aimed to study the effect of Fe doping in the CeO2 lattice on its physi- XRD patterns of Fe-Ce binary oxides are similar to the diffraction pat-
cochemical properties and catalytic activity for styrene oxidation. terns of pure CeO2, suggesting that the structure has been preserved. In
Accordingly, we have synthesized a series of iron-doped ceria nano- the case of Fe-Ce binary oxides, the primary (1 1 1) cubic fluorite lattice
catalysts with various amounts of iron (2, 5, 10, 15, 20, and 30 wt%), reflection peak clearly showed noticeable shift to a wider-angle side
including pure ceria and iron oxide for comparison. A systematic char- (Fig. 1 (b)) due to the replacement of larger Ce4+ (0.97 Å) ions by
acterization of the synthesized materials has been carried out using smaller Fe3+ (0.64 Å) ions in the CeO2 lattice leading to contraction of
different techniques, and catalytic efficacy was examined for styrene CeO2 lattice. In addition, the diffraction peaks become systematically
oxidation with TBHP to study the relationship between the catalyst broader and less intense as iron loading increases, which indicates a
structure and activity. decrease in degree of crystallization with increases of iron loading in the
Fe-Ce binary oxides [43,50,51]. The average crystallite size (D) of Fe-Ce
2. Experimental binary oxides was calculated by using the Scherrer equation and listed in
Table 1. The average crystallite sizes of Fe-Ce binary oxides were smaller
Catalyst preparation method, catalyst characterisation, and catalytic than that of pure CeO2. These are decreased with increase in iron content
activity studies were represented in supplementary information from 2 to 10 wt%, and increased, when iron content increased above 10
(Experimental section S1-S3). % (20FC, 30FC). Interestingly, the individual α-Fe2O3 characteristic
peaks were not present in the 2FC, 5FC, and 10FC binary oxides due to
3. Results and discussion iron species incorporation into the CeO2 lattice, forming ceria-iron solid
solutions (Ce-O-Fe). In the case of 20FC and 30FC binary oxides, the
3.1. Catalyst characterization studies characteristic α-Fe2O3 diffraction peaks were present due to iron species
segregated in the form of α-Fe2O3 on the CeO2 surface.
Powder XRD profiles of α-Fe2O3, pure CeO2, and Fe-Ce binary oxides N2-physisorption isotherms of Fe2O3, CeO2, and Fe-Ce binary oxides
loaded with different wt.% Fe are presented in Fig. 1 (a). Diffraction loaded with different wt.% Fe are presented in Fig. 2 (a). All synthesized
peaks of pure CeO2 could be seen at 2θ values of 28.2, 32.8, 47.3, 56.0, oxides display type IV isotherms with well-defined H3-type hysteresis
58.8, 69.2, 76.6, and 78.8 which are assigned to 111, 200, 220, 311, loops positioned in 0.6 to 0.98 P/P0 range, which denotes the meso-
222, 400, 331 and 420 lattice planes, respectively. These are related to porous nature of the materials [52]. Fe-Ce binary oxides showed higher

3
K. Yogendra et al. Applied Surface Science 676 (2024) 160971

Fig. 3. FT-IR spectra (a) and UV–Vis DRS (b) of pure CeO2, 2FC, 5FC, 10FC, 20FC, 30FC, and Fe2O3.

BET surface areas than the individual oxides and are listed in Table 1. The N2-physisorption studies supports the results of powder XRD. Binary
From N2-physisorption studies, it was observed that N2-absorption and oxides with a large surface area are ideal for catalytic activity due to
surface area increased with increase in iron content up to 10 wt% and their numerous active sites. N2-physisorption studies reveal that the
with further increases in iron content resulted in decrease N2-absorption 10FC catalyst has a higher surface area, which improves its efficiency in
and surface area, with no significant changes in the hysteresis loops. This the catalytic oxidation of styrene. Elemental composition of synthesised
likely occurs because segregated Fe2O3 species blocked the pores of the Fe-Ce binary oxides was analysed by XRF analysis and finding are listed
ceria. Among the synthesised oxides, 10FC exhibited the highest surface in Table 1.
area (97 m2/g) and the smallest crystallite size (7.22 nm). Raman spectroscopy was employed to characterize the electronic
Generally, a decrease in particle size results in an increased surface and phonon structures of CeO2 loaded with different wt.% Fe. Fig. 2 (b)
area of the material. Therefore, the higher surface area of the 10FC presents the Raman spectra of Fe2O3, CeO2, and Fe-Ce binary oxides
sample is most likely due to the presence of smaller CeO2 crystallites. loaded with different wt% Fe. Oxygen ions around the cerium ion of the

Fig. 4. TEM image (a), HR-TEM images (b and c), SEAD pattern (d), and EDX (e) of 10FC.

4
K. Yogendra et al. Applied Surface Science 676 (2024) 160971

Fig. 5. HAADF-STEM and EDS elemental mapping (a-f) of 10FC.

cubic fluorite phase with the Fm3m space group were identified by the CeO2, and Fe-Ce binary oxides loaded with different wt.% Fe are showed
existence of F2g band at 460 cm− 1 in pure CeO2 spectra that support the in Fig. 3 (b). The UV–Vis DRS of CeO2 revealed two weak absorption
powder XRD findings [53,54]. With increase of iron loading in binary bands at 240 and 280 nm and one strong absorption band at 345 nm,
oxides, the F2g band broadened and shifted to a lower wavenumber side which are associated to the charge transfer transitions (O2– → Ce3+, O2–
due to size and strain effect, and increased concentrations of oxygen → Ce4+) and interband transitions, respectively. The spectra revealed
defects in sublattice of CeO2 [55]. The substitution of Ce4+ with Fe3+ is that pure ceria did not exhibit any absorption features in the visible
recognized for inducing the reduction of neighbouring Ce4+ ions to Ce3+ range. Fe2O3 spectra showed four absorption bands, the absorption
ions. Consequently, this process generates one Ov for every two Fe3+ characteristics at less than 300 nm are attributed to the iron (III) ion in
ions, ensuring charge balance within the CeO2 lattice. XPS studies the octahedral and tetrahedral environments of charge transfer transi-
revealed the presence of both Ce3+ and Ce4+ ions in the samples, indi- tions, and the bands at ~ 370, 540, and 650 nm are assigned to α-Fe2O3.
cating that the synthesized Fe-Ce binary oxides possess Ov arising from [48].
substitution of Ce4+ with Fe3+. Additionally, all Fe-Ce binary oxides, Synthesized Fe-Ce binary oxides, 2FC, 5FC, and 10FC displayed ab-
exhibited a weak band at 595 cm− 1 which could be attributed to the sorption features only in the UV region that corresponds to CeO2 because
existence of Ov in the CeO2. Among the various Fe-Ce binary oxides, of iron species incorporation into the ceria lattice, which is a charac-
2FC, 5FC, and 10FC samples exhibit bands corresponding to CeO2 only teristic of solid solution formation. In the case of 20FC and 30FC binary
due to the creation of ceria-iron solid solutions (Ce-O-Fe), the 20FC and oxides absorption bands were seen in the UV and visible regions that
30FC binary oxides show bands of both CeO2 and α-Fe2O3, due to correspond to CeO2 and α-Fe2O3 because of iron species incorporation
presence of some iron as α-Fe2O3 apart from ceria iron solid solution, into the CeO2 lattice as well as iron species segregated on the CeO2. It is
which is aligned with the XRD results. associated with powder XRD, Raman, and FT-IR studies. In the Fe-Ce
The FT-IR spectra of α-Fe2O3, CeO2, and Fe-Ce binary oxides loaded binary oxide spectra, progressive absorption bands are slightly shifted
with different wt.% Fe was showed in Fig. 3 (a). Pure CeO2 and Fe-Ce to the higher wavelength side (red shift) and widened, which are
binary oxides display a strong absorption band at 512 cm− 1 due to the attributed to increased electron transfer and formation of defects in the
cerium oxygen bond (Ce-O) stretching vibrations. In addition, the 20FC CeO2 lattice with increased quantity of Fe doping. This phenomenon
and 30FC binary oxides showed absorption bands at ~ 464 and 538 arises from the lower-energy unoccupied orbits of Fe compared to Ce 4f,
cm− 1 due to Fe-O-Fe deformation and Fe-O stretching vibrations, reducing the energy required for the O 2p to unoccupied orbitals tran-
respectively, which are in line with powder XRD and Raman results sition. When considered alongside the H2-TPR studies, it suggests that
[43]. In addition, the broad band at the wavelength of 3443 cm− 1 electron transfer within the CeO2 lattice becomes more efficient due to
attributed to the stretching vibration mode of hydroxy groups of the strong interaction between iron and cerium, probable due to the
adsorbed water on surface [56]. creation of Fe-O-Ce solid solution phase. The evident findings indicate a
The UV–Vis DRS technique was employed to understand the metal- red shift has taken place, suggesting the existence of oxygen defects
–oxygen (M− O) surrounding environment. The UV–Vis DRS of Fe2O3, impacting the catalytic efficiency of oxides.

5
K. Yogendra et al. Applied Surface Science 676 (2024) 160971

Fig. 6. Ce 3d XP spectra of pure CeO2, 2FC, 5FC, 10FC, 20FC, and 30FC (a), Fe 2p XP spectra of 2FC, 5FC, 10FC, 20FC, 30FC and Fe2O3 (b), and O 1 s XP spectra of
pure CeO2, 2FC, 5FC, 10FC, 20FC, and 30FC (c).

respectively, which are ascribed to tetravalent cerium ion (Ce4+:


Table 2
3d104f0). Whereas another set of peaks at 903 and 884 eV assigned by u’
XPS results of Ce 3d, O 1s, and Fe 2p of synthesised samples.
and v’, respectively, are ascribed to trivalent cerium ion (Ce3+: 3d104f1).
S. No Sample Ceþ3/Ce Oα/Oα þ Oβ Feþ2/Fe The ceria 3d XP spectra suggested that both tri and tetra valent cerium
1 Fe2O3 − − − existed (Ce3+ and Ce4+) in these samples [27]. The XPS studies supports
2 CeO2 0.337 0.582 − the results of Raman studies. The ratio of area of the Ce3+ to the total
3 2FC 0.345 0.647 − area of Ce3+ and Ce4+are presented in Table 2. It shows that the relative
4 5FC 0.362 0.655
intensity of the Ce3+ related peaks (u’ and v’) raised consistently as iron

5 10FC 0.391 0.699 −
6 20FC 0.485 0.652 0.036 loading increased in the binary oxides, indicating that increase in Ce3+
7 30FC 0.558 0.452 0.078 in binary oxides with increase of iron content. This can be attributed to
the doping of iron into the CeO2 matrix facilitates the reduction of Ce4+
to Ce3+ [58].
To explore the surface morphology of ceria (a), iron oxide (b), and
Fe 2p XP spectra (Fig. 6 (b)) of Fe2O3 and Fe-Ce binary oxides
various wt.% Fe loaded Fe-Ce binary oxides, 2FC (c), 5FC (d), 10FC (e, f,
exhibited the Fe 2p3/2, Fe 2p1/2 peaks at 711, 724 and two satellite peaks
and g), 20FC (h) and 30FC (i), the FE-SEM images were acquired and
at 718, and 733 eV which designate the presence of the Fe3+ ion. Due to
displayed in Figure S1. Since, all Fe-Ce binary oxides were synthesized
the extremely low quantity of Fe in the ceri in 2FC showed two satellite
by following the same procedure, FE-SEM micrographs all the samples
peaks, 5FC and 10FC contain only Fe3+ ion-related peaks in the Fe 2p XP
demonstrate no significant morphological changes.
spectra. The remaining samples, 20FC and 30FC, another peak was seen
The HR-TEM characterization was applied to gain a more in-depth
at 709 eV, which corresponded to the Fe2+ ion. The Fe2+ formation must
understanding of the 10FC sample microstructure, and resulting im-
result from the CeO2-Fe2O3 interaction. At the CeO2-Fe2O3 interface the
ages are depicted in Figs. 4 and 5. TEM images reveal that iron-doped
Fe-O species should be in very close to Ce atoms, it is quite feasible that
ceria nanoparticles in hexagonal morphology are smaller than 8 nm
existence of Fe2+ and Ce3+ ions lead to coupling of the Ce4+-Ce3+ and
and evenly spread (Fig. 4 (a)). The TEM image (Fig. 4 (b and c)) exhibits
Fe+3-Fe2+ through interfacial redox process: xFe2O3 + (2 − y)CeO2− x →
lattice fringes with an inter-planar spacing of 0.29 nm assigned to the
xFe2O3− y + (2 − y)CeO2 [28,59]. Estimated quantity of Fe2+ by the area
(1 1 1) plane of ceria. The selected area electron diffraction (SAED)
ratio of the Fe2+ to that of total Fe (Fe2+ and Fe3+) is shown in Table 2
image (Fig. 4 (d)) reveals clearly the diffraction rings that validate the
which is showing more Fe2+ in 30FC than 20FC.
polycrystalline structure of 10FC sample. All SAED patterns’ concentric
Fig. 6 (c) shows the O 1 s XP spectra of ceria and Fe-Ce binary oxides.
diffraction rings can be assigned to (1 1 1), (2 0 0), (2 2 0), (3 1 1), (4 0 0),
This spectrum exhibited two distinct peaks; one peak at ~ 529.2 eV is
(3 3 1), and (4 2 0) planes of the ceria cubic fluorite structure [57]. These
related to lattice oxygen (Oα), another peak at ~ 531.5 eV is related to
findings are well aligned with the XRD data that show the cubic fluorite-
surface oxygens (Oβ). In Fe-Ce binary oxides, the lattice oxygen species
structure of ceria nanoparticles in 10FC. EDS analysis graph showing
peak (Oα) shifts to higher binding energy, suggesting change in chemical
composition of elements is presented in Fig. 4 (e). Fig. 5 (a-f) shows the
environment of ceria lattice oxygen species due to the doping of Fe3+
HAADF-STEM and elemental mapping imagery of 10FC. The analysis
ions [60]. The ratio of lattice oxygen species (Oα) to the of the total
here reveals iron-doped ceria, with red, yellow, and green colour dots
oxygen species (Oα + Oβ) calculated by considering peaks areas and
indicating cerium, iron, and oxygen alignment.
displayed in Table 2. It is observed that increasing the quantity of iron
The relative proportion of Ce, Fe, and O and their oxidation states on
doping from 2 to 10 wt% in the ceria lattice increased the lattice oxy-
the surface of Fe2O3, CeO2, and Fe-Ce binary oxides was studied using
gens. Further increase in the iron content increased the surface oxygens,
XPS. The Ce 3d XP spectra (Fig. 6 (a)) of ceria and Fe-Ce binary oxides
which may result from segregation of Fe2O3 on the CeO2 surface. Lattice
showed two characteristic spin–orbit multiplets, i.e., Ce 3d5/2 and 3d3/2,
oxygens are crucial for ceria redox reactions. In this process, a reactant
and deconvoluted into eight components at the binding energies of ~
molecule reacts with a lattice oxygen to form an oxidized product and
916, 907, 900, 898, 888, and 882 eV assigned by u“’, u”, u, v“’, v”, and v,
leaves behind an Ov. The Ov can be replenished by oxygen from oxidant

6
K. Yogendra et al. Applied Surface Science 676 (2024) 160971

Fig. 7. H2-TPR patterns of pure CeO2, 2FC, 5FC, 10FC, 20FC, 30FC, and Fe2O3 (a), The relationship between catalytic activity and physicochemical properties of
various wt% Fe-doped ceria catalysts (b).

through surface, thereby restoring the lattice oxygen, and maintaining 2FC and 5FC binary oxides, showing increased reducibility of surface
the catalytic cycle [29]. and lattice oxygens in CeO2 and iron oxides. This is attributed to the
The XPS investigation revealed that in the Fe-Ce binary oxides 2FC, smaller crystallite size, improved lattice oxygen mobility, and strong
5FC, and 10FC the Ce3+ ion-related peaks rise consistently as iron ceria-iron interaction in the Ce-O-Fe solid solutions. When the wt% Fe
loading increases. This can be ascribed to the creation of a solid solution, was increased to 20 (20FC), a new minor peak appeared at 530 ◦ C. The
where the doping of iron into CeO2 matrix facilitates the reduction of low-temperature peak at 460 ◦ C likely overlaps with the surface CeO2
Ce4+ to Ce3+. In 20FC and 30FC, both Ce3+ and Fe2+ ions-related peaks reduction and the phase transition of Fe2O3 to Fe3O4, while the 530 ◦ C
intensity increased with increase in iron loading, suggesting interaction peak corresponds to the reduction of Fe3O4 to FeO. With further Fe
of CeO2 with Fe2O3 formed on the surface of ceria iron solid solution. At addition (30FC), the intensity of the low-temperature peak increases and
the CeO2-Fe2O3 interface the Fe-O species should be in very close to Ce shifts to higher temperatures, indicating that Fe2O3 aggregation be-
atoms, it is quite feasible that the Ce4+-Ce3+ and Fe+3-Fe2+ through comes more significant at higher Fe levels [52].
interfacial redox process: xFe2O3 + (2 − y)CeO2− x → xFe2O3− y + (2 − y) This suggests that segregation of Fe2O3 in 20FC and 30FC samples
CeO2 which results in formation of Fe2+ and Ce3+ ions. These findings becomes too severe, impeding the reduction of surface Ce4+ in redox
indicate that the formation of the ceria-iron solid solutions and CeO2- equilibrium. Segregation of Fe2O3 when doping more than 10 % Fe, is
Fe2O3 interface species is dependent on the amount of iron doped in also evidenced by powder XRD patterns, Raman, FT-IR, and UV–vis
ceria and is in consistent with the powder XRD data. spectra. This excessive agglomeration might be one of the reasons for
H2-TPR was used to measure the reducibility of prepared catalysts. It decrease in surface area and decline in catalytic efficiency of 20FC and
is widely recognized that a decrease in reduction temperature tends to 30FC catalysts. From TPR analysis, we can conclude that the introduc-
signify greater catalyst activity. Fig. 7 (a) display the H2-TPR pattern of tion of iron promotes the reducibility of ceria. This is evidenced by the
ceria, Fe2O3, and Fe-Ce binary oxides with varying iron content. Pure increased reducibility of the binary oxides up to 10 % Fe, which is
CeO2 has reduction peaks at 570 and 860 ◦ C, related to the reduction of related to the decrease in crystallite size, enhancement of lattice oxygen
surface and oxygen lattice oxygen, respectively. Pure Fe2O3 exhibits at concentration and mobility, and strong ceria-iron interaction in the Ce-
485, 680, and 860 ◦ C, indicating a three-step reduction process: Fe2O3 O-Fe solid solutions.
(hematite) → Fe3O4 (magnetite), Fe3O4 (magnetite) → FeO (wurtzite), In the process of styrene oxidation, Fe-Ce binary oxide catalysts
and (wurtzite) FeO→Fe (metallic iron) [61]. Fe2O3 demonstrates a primarily facilitate the transfer of oxygen to styrene. The activity of Fe-
notably lower reduction temperature compared to CeO2, suggesting its Ce binary oxide catalysts is closely linked to their surface properties, as
superior reduction capability, possibly a crucial dopant contributing to these factors highly impact the creation of active oxygen species avail-
enhanced the styrene oxidation with Fe-Ce binary oxides. The 2FC ex- able for oxygen exchange. The relationship between catalytic efficiency
hibits the four low temperature reduction peaks at 339, 436, 730, and and the physicochemical properties of various wt% Fe-doped ceria
816 ◦ C assigned to reduction of surface iron-substituted cerium site, catalysts is depicted in Fig. 7 (b). Substitution of large Ce4+ ions with
surface oxygens of CeO2, bulk iron-substituted cerium cites, and bulk smaller Fe3+ ions in Ce-O-Fe solid solutions, reduce crystallite size,
oxygens of CeO2, respectively. The 5FC and 10FC catalysts exhibited thereby increase surface area and provide more active sites for catalytic
similar H2-TPR patterns to 2FC catalyst, except peak complexity. reactions. It also facilitates the formation of oxygen vacancies and en-
Reduction peaks of 10FC catalyst moved to a lower temperature than hances mobility of lattice oxygen which increases oxygen storage and
releasing capacity, redox properties, and reducibility of the material.
From the characterization studies, it was observed that increasing the
quantity of iron doping from 2 to 10 wt% in the ceria lattice, increased
the Ce-O-Fe sites in solid solution without forming Fe2O3. However,
further increasing the iron content led to the segregation of Fe2O3 on the
surface, resulting in a decrease in surface area. The 10 wt% iron-doped
ceria exhibits high surface area, abundant lattice oxygen, oxygen va-
cancies, and enhanced reducibility all of which directly impact styrene
Scheme 1. Styrene oxidation. oxidation.

7
K. Yogendra et al. Applied Surface Science 676 (2024) 160971

Table 3 oxides, 2FC had the higher SANCA (0.101), while the remaining binary
The catalyst performance for styrene oxidation. oxides showed similar activity (0.078–0.092). However, the 10FC
Entry Catalyst Conv. of Styrene Sel. of SO Sel. of BA Others catalyst exhibited high conversion of styrene due to its high surface area.
(%) (%) (%) (%) In this styrene oxidation reaction, it is observed that the selectivity to
1 − 12 55 45 − styrene oxide found to be high at higher styrene conversions. It is un-
2 Fe2O3 32 60 14 26 derstood that styrene is converted to styrene oxide with TBHP in pres-
3 CeO2 39 78 22 − ence of ceria-iron solid solution catalyst’s active sites. Benzaldehyde and
4 2FC 62 60 27 13 other products are formed by further reaction of formed styrene oxide
5 5FC 74 67 24 9
6 10FC 89 82 12 6
with TBHP. At higher styrene conversions, more TBHP is consumed and
7 20FC 75 78 11 11 less TBHP available for forming further oxidation products. This is evi-
8 30FC 68 72 9 19 denced by formation of higher side products when higher TBHP to
Reaction conditions: Styrene (10 mmol), TBHP (20 mmol), acetonitrile (10
substrate ratio used (Fig. 8 (C)).
mL), catalyst (100 mg), temperature (80 ◦ C), and time (8 h). styrene oxide (SO), The effect of experimental variables such as temperature and time,
Benzaldehyde (BA). quantity of catalyst and TBHP, and solvent polarity was explored on the
reaction in presence of highly active 10FC catalyst. The impact of tem-
3.2. Catalytic activity perature on the reaction was studied from 50 to 90 ◦ C, and the findings
are displayed in Fig. 8 (a). Styrene conversion is low at 50 ◦ C temper-
The catalytic activity of synthesised Fe-Ce binary oxides, pure ceria, ature and increased with increasing it to 90 ◦ C. A gradual slight decrease
and iron oxide, was tested for the of styrene oxidation at 80 ◦ C for 8 h in the selectivity is observed while raising the temperature up to 80 ◦ C,
with TBHP in acetonitrile solvent (Scheme 1) and results are presented and further raise to 90 ◦ C showed drastic decrease in the selectivity. As a
in Table 3. Without catalyst, low styrene conversion (12 %) was noticed, result, other reaction parameters were evaluated at 80 ◦ C considered as
which shows the need of a catalyst to enhance the reaction rate. Pure suitable temperature for the reaction. The effect of reaction time on
ceria and iron oxides exhibited reasonable catalytic activity, converting styrene oxidation was studied up to 10 h, and results are displayed in
32 and 39 % of styrene, respectively. Remarkably, higher styrene con- Fig. 8 (b). With rise in reaction period from 2 to 8 h, styrene conversion
versions were observed with Fe-Ce binary oxide catalysts, which in- increased gradually and after that did not show much improvement in
dicates the Fe dopant role in the efficiency of catalyst. With the increase the conversion. Styrene oxide selectivity slightly decreased up to 8 h,
of iron from 2 to 10 wt% (2FC, 5FC, and 10FC) in catalysts, styrene and after that, a more significant decrease in selectivity was observed.
conversion increased from 62 to 89 % and styrene oxide selectivity The effect of styrene to TBHP ratio on the reaction is optimized by
increased from 60 to 82 %. With the further increase in the iron content keeping the styrene quantity constant, and results are displayed in Fig. 8
(20FC and 30FC) the conversion and selectivity decreased gradually. (c). With increasing the TBHP styrene conversion increased, 1:2 styrene
The surface area normalized catalytic activity (SANCA: styrene con- to TBHP mole ratio was found to be suitable for high styrene conversion
version/surface area (mmol m− 2 g) of synthesised catalysts are listed in with better styrene oxide selectivity. Further increasing the TBHP, not
Table S1. For Fe2O3 SANCA was higher (0.133). Among Fe-Ce binary much improvement in conversion observed, but selectivity to styrene

Fig. 8. Effect of reaction temperature (a), effect of reaction time (b), effect of mole ratio of styrene to TBHP (c), and effect of the quantity of 10FC catalyst (d) on
styrene oxidation.

8
K. Yogendra et al. Applied Surface Science 676 (2024) 160971

Fig. 9. Effect of the solvent on styrene oxidation (a), and reusability of 10FC catalyst on styrene oxidation (b).

Table 4
Formation of styrene oxides from different substituted styrenes.
Entry Substrate Con. of substrate (%) Main product Sel. of Major product (%)

1 82 83

2 48 74

3 59 70

4 96 84

5 32 100

Reaction conditions: Substrate (10 mmol), TBHP (20 mmol), acetonitrile (10 mL), catalyst (100 mg), temperature (80 ◦ C), and time (8 h).

oxide decreased. The effect of the quantity of 10FC catalyst (0 to 150 (44 and 50%), whereas the polar organic solvents, DCE, 1,4-dioxane,
mg) on reaction was studied, and findings are displayed in Fig. 8 (d). acetone, and acetonitrile, demonstrated excellent conversion and
Very low styrene conversion (12 %) and product selectivity (55 %) were product selectivity. In investigated solvents, conversion and selectivity
noticed in the blank reaction (absence of catalyst). Conversion and is in the order: Hexane < Toluene < DCE<1,4-Dioxane < Acetone <
selectivity increased drastically in the presence of 50 mg catalyst. With Acetonitrile which is in line with the polarity order. Among them,
an increase in the catalyst quantity from 50 to 100 mg, the conversion acetonitrile showed the highest styrene conversion (89%) and styrene
considerably increased. However, with a further increase to 150 mg, oxide selectivity (82%).
there was no additional increase in conversion. Styrene oxide selectivity The 10FC catalyst that exhibited better activity was also investigated
gradually decreased with the increasing amount of catalyst. It is evident for reusability test, and findings are depicted in Fig. 9 (b). After the
form this study that 100 mg catalyst is suitable for the reaction under reaction, the catalyst was recovered from the reaction contents by a
experimental condition used. syringe filter, washed with ethanol to remove adhering molecules, dried
The polarity of solvents shows a significant impact on different 5 h at 120 ◦ C, and reused. The catalytic activity did not vary consider-
chemical processes. Thus, different solvents were used for the reaction ably up to 6 cycles, indicating a minor drop in the activity caused mainly
and outcome is shown in Fig. 9 (a). Non-polar organic solvents, hexane, by polar molecules sticking to the catalyst’s surface.
and toluene, showed less conversion (53, 60 %) and product selectivity We investigated the catalytic oxidation of different substituted

9
K. Yogendra et al. Applied Surface Science 676 (2024) 160971

Fig. 10. Plausible mechanism.

Table 5
Comparison of the catalytic efficiency of 10FC with other systems for styrene oxidation.
Entry Catalyst T (◦ C) Ti (h) Oxidant Solvent Conv. of STR (%) Se. of SO (%) Ref.

1 nanocrystalline Ag/WO3 75 12 H2O2 ACN 75 55 [16]


2 Gold Nanoparticles 80 8 TBHP Benzene 100 92 [17]
3 AuNPs/dendrimers @ resin 80 24 O2 Dioxane 95 48 [18]
4 Ti/Au Nanoparticles 70 15 TBHP Toluene 41 39 [19]
5 SiO2-AuNPs-SiO2 80 10 TBHP Toluene 74 67 [20]
6 Pd and Pt nanoparticles 80 24 TBHP (DMF: ACN=1:9) 25.4 75 [21]
7 Ag-SiO2 80 24 TBHP Toluene 77 47 [22]
8 Cu- hexagonal silicas 120 12 TBHP (ACN: DMF=9:1) 99 84 [23]
9 CuO on CoAl-HT 85 6 TBHP ACN 99.5 72 [24]
10 10 wt% Fe doped ceria 80 8 TBHP ACN 89 82 [Present work]

T: Temperature (◦ C), Ti: Time (h), TBHP: t-Butylhydroperoxide, ACN: Acetonitrile, DMF: N, N-Dimethylformamide, Con.: Conversion, Sel.: Selectivity, STR: Styrene,
SO: Styrene oxide.

styrenes by 10FC catalyst under optimum conditions, and findings are Earlier we reported cerium-bismuth mixed oxide catalysts for styrene
tabulated in Table 4. Epoxides were the main product. Additionally, we oxidation [5]. Among various % Bi doped ceria catalysts, the
observed that electron-donating groups promote the epoxidation reac- Ce0.7Bi0.3O2 (30 wt% Bi doped ceria) showed better properties like
ion, whereas electron-withdrawing groups retard it. Particularly in the reducibility, oxygen vaccines etc. and showed better catalytic activity
case of stilbene (entry 5), 32% conversion and 100 selectivity for for styrene oxidation. In the quest to study properties and catalytic ac-
epoxide were observed. 4-Vinyl nitro styrene does not participate in the tivity of other binary oxides of ceria, here we studied Fe-Ce binary ox-
catalytic oxidation with TBHP in the presence of 10FC catalyst because ides as catalysts for styrene oxidation. Among various amounts of iron
–NO2 is a strong electron withdrawing group. doped ceria binary oxides, the 10 wt% Fe-doped ceria catalyst demon-
Based on previous literature and current research, we propose a strated better properties and catalytic activity for styrene oxidation
plausible mechanism, as depicted in Fig. 10. Characterization studies which is comparable with that of Ce0.7Bi0.3O2. Both catalysts possibly
reveal that the Ce-O-Fe solid solutions contains oxygen vacancy (Ov) follow similar pathway in their reaction. In ceria bismuth mixed oxides,
defects, Ce3+, Ce4+, and Fe3+. According to several previous studies, up to 50 % bismuth content also solid solution formation is seen without
styrene reacts with lattice oxygen to form a styrene oxide through bismuth oxide phase. Whereas in ceria iron binary oxides, 20 % and
interaction with the styrene ring and leaves behind an oxygen vacancy. above iron doped ceria showed Fe2O3 phase which shows ceria can take
The oxygen vacancy can be replenished by oxygen from TBHP, thereby different amounts of various metal ions in to lattice.
restoring the lattice oxygen and maintaining the catalytic cycle [62].
The catalytic performance of 10FC was compared with other cata- 4. Conclusions
lytic systems documented in the literature for styrene oxidation, as
presented in Table 5. Most previous studies have focused on the styrene A series of mesoporous and nanosized Fe-Ce binary oxides and pure
oxidation to styrene oxide using noble metal catalysts. Our prepared oxides of Fe and Ce were synthesized by a facile template-assisted
10FC catalyst is a cheaper, abundant, non-noble metal-based alternative coprecipitation method, characterised by several studies, and evalu-
that uses a clean, inexpensive, and accessible oxidant for the oxidation of ated for styrene oxidation. All Fe-doped ceria catalysts (2, 5, 10, 20, and
styrene, achieving better conversion and selectivity. 30 wt% Fe) showed the formation of solid solutions. When the Fe

10
K. Yogendra et al. Applied Surface Science 676 (2024) 160971

content is increased to more than 10 wt%, there is slight segregation of [7] D. Pan, Q. Xu, Z. Dong, S. Chen, F. Yu, X. Yan, B. Fan, R. Li, Facile synthesis of
highly ordered mesoporous cobalt-alumina catalysts and their application in liquid
Fe3+ in the form of α-Fe2O3 in the samples (20FC and 30FC), as evi-
phase selective oxidation of styrene, RSC Adv. 5 (119) (2015) 98377–98390.
denced by XRD, Raman, FT-IR, UV–Vis DRS, and XPS analyses. The Fe [8] X. Liu, J. Ding, X. Lin, R. Gao, Z. Li, W.L. Dai, Zr - doped CeO2 nanorods as versatile
doping to ceria significantly improved the reducibility, and outperform catalyst in the epoxidation of styrene with tert - butyl hydroperoxide as the
pure Fe and Ce oxides in terms of their catalytic performance for the oxidant, Appl. Catal. A: Gen. 503 (2015) 117–123.
[9] N. Masunga, G.S. Tito, R. Meijboom, Catalytic evaluation of mesoporous metal
selective oxidation of styrene with TBHP as an oxidant. Among Fe-Ce oxides for liquid phase oxidation of styrene, Appl. Catal. A: Gen. 552 (2018)
binary oxides, the 10 wt% doped (10FC) catalyst exhibited higher effi- 154–167.
ciency because the 10 wt.% Fe appears to have an optimal iron content [10] N.M.R. Martins, A.J.L. Pombeiro, L.M. Martins, A green methodology for the
selective catalytic oxidation of styrene by magnetic metal - transition ferrite
to achieve the highest surface area, reducibility, lattice oxygens, and low nanoparticles, Catal. Commun. 116 (2018) 10–15.
crystallinity as evidenced by BET measurements, H2-TPR, XPS, and [11] H. Hong, L. Hu, M. Li, J. Zheng, X. Sun, X. Lu, X. Cao, J. Lu, H. Gu, Preparation of
powder XRD, respectively. Pt@Fe2O3 nanowires and their catalysis of selective oxidation of olefins and
alcohols, Chem. Eur. J. 17 (31) (2011) 8726–8730.
[12] J. Tong, W. Li, L. Bo, H. Wang, Y. Hu, Z. Zhang, A. Mahboob, Selective oxidation of
CRediT authorship contribution statement styrene catalysed by cerium - doped cobalt ferrite nanocrystals with greatly
enhanced catalytic performance, J. Catal. 344 (2016) 474–481.
[13] J. Tong, X. Cai, H. Wang, Q. Zhang, Mater. Improvement of catalytic activity in
Kamma Yogendra: Writing – original draft, Validation, Methodol- selective oxidation of styrene with H2O2 over spinel Mg-Cu ferrite hollow spheres
ogy, Investigation, Formal analysis, Data curation, Conceptualization. in water, Res. Bull. 55 (2014) 205–211.
Palli Sitaramulu: Writing – review & editing. Silligandla Nazeer: [14] H. Wang, W. Qian, J. Chen, Y. Wu, X. Xu, J. Wang, Y. Kong, Spherical V - MCM -
48: the synthesis, characterization and catalytic performance in styrene oxidation,
Formal analysis. Palnati Manoj Kumar: Formal analysis. Benjaram RSC Adv. 4 (92) (2014) 50832–50839.
Mahipal Reddy: Formal analysis. Tumula Venkateshwar Rao: [15] N. Ma, Y. Yue, W. Hua, Z. Gao, Selective oxidation of styrene over nanosized spinel
Writing – review & editing, Software. - type MgxFe3− xO4 complex oxide catalysts, Appl. Catal. A: Gen. 251 (1) (2003)
39–47.
[16] S. Ghosh, S.S. Acharyya, M. Kumar, R. Bal, One - pot preparation of nanocrystalline
Declaration of competing interest Ag-WO3 catalyst for the selective oxidation of styrene, RSC Adv. 5 (47) (2015)
37610–37616.
[17] F. Zhang, X. Zhao, C. Feng, B. Li, T. Chen, W. Lu, X. Lei, S. Xu, Crystal - face -
The authors declare that they have no known competing financial selective supporting of gold nanoparticles on layered double hydroxide as efficient
interests or personal relationships that could have appeared to influence catalyst for epoxidation of styrene, ACS Catal. 1 (4) (2011) 232–237.
[18] A.S. Sharma, D. Shah, H. Kaur, Gold nanoparticles supported on dendrimer@ resin
the work reported in this paper.
for the efficient oxidation of styrene using elemental oxygen, SC Adv. 5 (53) (2015)
42935–42941.
Data availability [19] M. Nemanashi, R. Meijboom, Dendrimer derived titania - supported Au
nanoparticles as potential catalysts in styrene oxidation, Catal. Lett. 143 (2013)
324–332.
Data will be made available on request. [20] S. Das, T. Asefa, Core-shell-shell microsphere catalysts containing Au nanoparticles
for styrene epoxidation, Top. Catal. 55 (2012) 587–594.
[21] J.H. Noh, R. Patala, R. Meijboom, Catalytic evaluation of dendrimer and reverse
Acknowledgments
microemulsion template Pd and Pt nanoparticles for the selective oxidation of
styrene using TBHP, Appl. Catal. A: Gen 514 (2016) 253–266.
The authors K. Y. wish to acknowledge the Council of Scientific and [22] X. Huang, W. Dong, G. Wang, M. Yang, L. Tan, Y. Feng, X. Zhang, Synthesis of
Industrial Research (CSIR), New Delhi, India, and P. S. and S. N. thank confined Ag nanowires within mesoporous silica via double solvent technique and
their catalytic properties, Colloid Interface Sci. 359 (1) (2011) 40–46.
the University Grants Commission (UGC), New Delhi, India for research [23] X. Lu, Y. Yuan, Copper - containing hexagonal mesoporous silicas for styrene
fellowships as well as Academy of Scientific & Innovative Research epoxidation using tert - butylhydroperoxide, Appl. Catal. A: Gen. 365 (2) (2009)
(AcSIR), and CSIR-IICT for the doctoral research support. The authors 180–186.
[24] R. Hu, P. Yang, Y. Pan, Y. Li, Y. He, J. Feng, D. Li, Synthesis of a highly dispersed
sincerely appreciate considerable help received from Dr. Chandrashekar CuO catalyst on CoAl - HT for the epoxidation of styrene, Dalton Trans. 46 (39)
Pendem for Raman and N2-physisorption, Dr. Pratyay Basak for FE-SEM, (2017) 13463–13471.
Dr. B. Sreedhar for HR-TEM, and Dr. Manorama for XPS studies. The [25] Fu. Yang, X. Bo Shao, S.G. Liu, Hu. Xu, M. Xu, Y. Wang, S. Zhou, Yan Kong,
Nanosheet-like Ni-based metasilicate towards the regulated catalytic activity in
support from the Director, CSIR-IICT is also duly acknowledged styrene oxidation via introducing heteroatom metal, Appl. Surf. Sci. 471 (2019)
(Manuscript Communication Number: IICT/Pubs./2024/054). 822–883.
[26] R. Wang, X. Liuc, S. Fu Yang, S.Z. Gao, Yan Kong, Neighboring Cu toward Mn site
in confined mesopore to trigger strong interplay for boosting catalytic epoxidation
Appendix A. Supplementary data of styrene, Appl. Surf. Sci. 537 (2021) 14810.
[27] H. Zhang, F. Gu, Q. Liu, J. Gao, L. Jia, T. Zhu, Y. Chen, Z. Zhong, F. Su, MnOx-CeO2
Supplementary data to this article can be found online at https://doi. supported on a three - dimensional and networked SBA - 15 monolith for NO x -
assisted soot combustion, RSC Adv. 4 (29) (2014) 14879–14889.
org/10.1016/j.apsusc.2024.160971. [28] H. Li, K. Li, H. Wang, X. Zhu, Y. Wei, D. Yan, X. Cheng, K. Zhai, Soot combustion
over Ce1 - xFexO2 - δ and CeO2/Fe2O3 catalysts: roles of solid solution and
References interfacial interactions in the mixed oxides, Appl. Surf. Sci. 390 (2016) 513–525.
[29] Ha. Gu1, T. Yokoya1, L. Kang, S. Marlow, X. Su, M. Gong, J. Yan, Y. Ren, Z. Wang,
X. Guan, L. Liu, Z. Yao, L. Keenan, H. Asakura, F.R. Wang, Oxygen vacancy
[1] V. Hulea, E. Dumitriu, Styrene oxidation with H2O2 over Ti - containing molecular
formation as the rate-determining step in the Mars-van Krevelen mechanism,
sieves with MFI, BEA and MCM - 41 topologies, Appl. Catal. A: Gen. 277 (1–2)
ChemRxiv. (2024), https://doi.org/10.26434/chemrxiv-2024-nbzhr.
(2004) 99–106.
[30] C. Sun, H. Li, L. Chen, Nanostructured ceria - based materials: synthesis, properties,
[2] M. Liu, X. Dong, Z. Guo, A. Yuan, S. Gao, Fu Yang, Enabling tandem oxidation of
and applications, Energy Environ. Sci. 5 (9) (2012) 8475–8505.
benzene to benzenediol over integrated neighbouring V-Cu oxides in mesoporous
[31] B. Govinda Rao, D. Jampaiah, P. Venkataswamy, B.M. Reddy, Enhanced catalytic
silica, Chin. J. Chem. Eng. 55 (2023) 236–245.
performance of manganese and cobalt Co-doped CeO2 catalysts for diesel soot
[3] K. Yogendra, P. Sitaramulu, S. Nazeer, P.M. Kumar, B.M. Reddy, T.V. Rao, SiO2
oxidation, ChemistrySelect 1 (21) (2016) 6681–6691.
supported Co3O4 catalyst for selective oxidation of cyclohexene with molecular
[32] S. Nazeer, P. Sitaramulu, K. Yogendra, P.M. Kumar, B.M. Reddy, T. Venkateshwar
oxygen, Biomass Convers. Biorefinery (2023) 1–15, https://doi.org/10.1007/
Rao, Mesoporous copper - cerium mixed oxide catalysts for aerobic oxidation of
s13399-023-05102-y.
vanillyl alcohol, Catalysts 13 (7) (2023) 1058.
[4] P.R.G.N. Reddy, B.G. Rao, T.V. Rao, B.M. Reddy, Mesoporous Ce-Zr mixed oxides
[33] A. Rangaswamy, P. Venkataswamy, D. Devaiah, S. Ramana, B.M. Reddy, Structural
for selective oxidation of styrene in liquid phase, Appl. Petrochem. Res. 10 (2020)
characteristics and catalytic performance of nanostructured Mn - doped CeO2 solid
67–76.
solutions towards oxidation of benzylamine by molecular O2, Mater. Res. Bull. 88
[5] P. Sitaramulu, K. Yogendra, S. Nazeer, R. Kishore, B.M. Reddy, T.V. Rao, Selective
(2017) 136–147.
oxidation of styrene over nanostructured cerium-bismuth mixed oxide catalysts,
[34] D. Shang, Q. Zhong, W. Cai, Influence of the preparation method on the catalytic
New J. Chem. 47 (16) (2023) 7556–7565.
activity of Co/Zr1− xCexO2 for NO oxidation, J Mol Catal. A Chem. 399 (2015)
[6] S.N. Antonio de Brito, L.G. Pinheiro, M. Josue Filho, A.C. Oliveira, Studies on
18–24.
styrene selective oxidation over iron - based catalysts: reaction parameters effects,
Fuel 150 (2015) 305–317.

11
K. Yogendra et al. Applied Surface Science 676 (2024) 160971

[35] P. Venkataswamy, K.N. Rao, D. Jampaiah, B.M. Reddy, Nanostructured manganese [50] L. Tang, D. Yamaguchi, N. Burke, D. Trimm, K. Chiang, Methane decomposition
doped ceria solid solutions for CO oxidation at lower temperatures, Appl. Catal. B: over ceria modified iron catalysts, Catal. Commun. 11 (15) (2010) 1215–1219.
Environ. 162 (2015) 122–132. [51] D. Tang, G. Zhang, Efficient removal of fluoride by hierarchical Ce-Fe bimetal
[36] D. Mukherjee, B.G. Rao, B.M. Reddy, CO and soot oxidation activity of doped ceria: oxides adsorbent: thermodynamics, kinetics and mechanism, Chem. Eng. J 283
Influence of dopants, Appl. Catal. B: Environ. 197 (2016) 105–115. (2016) 721–729.
[37] A. Rangaswamy, P. Sudarsanam, B.M. Reddy, Rare earth metal doped CeO2 - based [52] D. Jampaiah, P. Venkataswamy, K.M. Tur, S.J. Ippolito, S.K. Bhargava, B.M. Reddy,
catalytic materials for diesel soot oxidation at lower temperatures, J. Rare Earths. Effect of MnOx loading on structural, surface, and catalytic properties of CeO2-
33 (11) (2015) 1162–1169. MnOx mixed oxides prepared by sol-gel method, fur Anorg. Allg. Chem. 641 (6)
[38] L. Ma, D. Wang, J. Li, B. Bai, L. Fu, Y. Li, Ag/CeO2 nanospheres: Efficient catalysts (2015) 1141–1149.
for formaldehyde oxidation, Appl. Catal. B: Environ. 148 (2014) 36–43. [53] O.H. Laguna, M.A. Centeno, M. Boutonnet, J.A. Odriozola, Fe - doped ceria solids
[39] P. Sudarsanam, B. Mallesham, D.N. Durgasri, B.M. Reddy, Physicochemical and synthesized by the microemulsion method for CO oxidation reactions, Appl. Catal.
catalytic properties of nanosized Au/CeO2 catalysts for eco - friendly oxidation of B: Environ. 106 (3–4) (2011) 621–629.
benzyl alcohol, J. Ind. Eng. Chem. 20 (5) (2014) 3115–3121. [54] K. Kumari, R.N. Aljawfi, Y.S. Katharria, S. Dwivedi, K.H. Chae, R. Kumar,
[40] S. Palli, Y. Kamma, N. Silligandla, B.M. Reddy, V.R. Tumula, Aerobic oxidation of A. Alshoaibi, P.A. Alvi, S. Dalela, S. Kumar, Study the contribution of surface
ethylbenzene to acetophenone over mesoporous ceria-cobalt mixed oxide catalyst, defects on the structural, electronic structural, magnetic, and photocatalyst
Res. Chem. Intermed. 1–20 (2022). properties of Fe: CeO2 nanoparticles, J. Electron Spectrosc. Relat. Phenom. 235
[41] N. Silligandla, S. Palli, Y. Kamma, S. Dalal, M.K. Palnati, S. Bojja, B.M. Reddy, V. (2019) 29–39.
R. Tumula, Mesoporous ceria-supported vanadia catalyst for selective aerobic [55] M.B. Radović, Z. Dohcevic-Mitrovic, N.M. Paunović, M. Šćepanović, B. Matović, Z.
oxidation of ethylbenzene, New J. Chem. 48 (2024) 10391–10400. V. Popović, Effect of Fe2+ (Fe3+) doping on structural properties of CeO2
[42] P. Sitaramulu, S. Nazeer, K. Yogendra, A.S. Prasad, P. Chandrashekar, T.V. Rao, nanocrystals, Acta Phys. Pol. A. 116 (1) (2009) 84–87.
SiO2 supported Ce-Co mixed oxide catalyzed selective allylic oxidation of [56] R. Fu Yang, S. Wang, X. Zhou, Y. Wang, S.G. Kong, Mesopore-encaged V-Mn oxides:
cyclohexene, New J. Chem. 48 (2024) 1932–1942. Progressive insertion approach triggering reconstructed active sites to enhance
[43] T. Tsoncheva, C. Rosmini, M. Dimitrov, G. Issa, J. Henych, Z. Němečková, catalytic oxidative desulfuration, Chin. J. Chem. Eng. 5 (2022) 182.
D. Kovacheva, Formation of catalytic active sites in hydrothermally obtained [57] A.A. Ansari, J. Labis, M. Alam, S.M. Ramay, N. Ahmad, A. Mahmood,
binary ceria-iron oxides: composition and preparation effects, ACS Appl. Mater. Physicochemical and redox characteristics of Fe ion-doped CeO2 Nanoparticles,
Interfaces 13 (1) (2020) 1838–1852. J. Chin. Chem. Soc. 62 (10) (2015) 925–932.
[44] W.A. Aboutaleb, R.A. El-Salamony, Effect of Fe2O3 - CeO2 nanocomposite synthesis [58] L. Katta, P. Sudarsanam, G. Thrimurthulu, B.M. Reddy, Doped nanosized ceria solid
method on the Congo red dye photodegradation under visible light irradiation, solutions for low temperature soot oxidation: Zirconium versus lanthanum
Mater. Chem. Phys. 236 (2019) 121724. promoters, Appl. Catal. B: Environ. 101 (2010) 101–108.
[45] Z. Wen, Y. Zhang, G. Cheng, Y. Wang, R. Chen, Simultaneous removal of As (V)/Cr [59] M. Lykaki, S. Stefa, S.A.C. Carabineiro, P.K. Pandis, V.N. Stathopoulos,
(VI) and acid orange 7 (AO7) by nanosized ordered magnetic mesoporous Fe - Ce M. Konsolakis, Facet-dependent reactivity of Fe2O3/CeO2 nanocomposites: effect of
bimetal oxides: behavior and mechanism, Chemosphere 218 (2019) 1002–1013. ceria morphology on CO oxidation, Catalysts 9 (4) (2019) 371.
[46] P. Sudarsanam, B. Mallesham, D.N. Durgasri, B.M. Reddy, Physicochemical [60] B. Li, A. Raj, E. Croiset, J.Z. Wen, Reactive Fe-O-Ce sites in ceria catalysts for soot
characterization and catalytic CO oxidation performance of nanocrystalline Ce-Fe oxidation, Catalysts 9 (10) (2019) 815.
mixed oxides, RSC Adv. 4 (22) (2014) 11322–11330. [61] H. Liu, Y. Wu, L. Liu, B. Chu, Z. Qin, G. Jin, Z. Tong, L. Dong, B. Li, Three-
[47] S. Palli, K. Yogendra, N. Silligandla, M.R. Benjaram, V.R. Tumula, Oxidation of dimensionally ordered macroporous Fe-doped ceria catalyst with enhanced activity
vanillyl alcohol to vanillin over nanostructured cerium-iron mixed oxide catalyst at a wide operating temperature window for selective catalytic reduction of NOx,
with molecular oxygen, Res. Chem. Intermed. 48 (11) (2022) 4579–4599. Appl. Surf. Sci. 498 (2019) 143780.
[48] P. Venkataswamy, D. Jampaiah, K.N. Rao, B.M. Reddy, Nanostructured [62] Q. Gazu1, M. Shozi, P. Mpungose, Oxidation of styrene to benzaldehyde and
Ce0.7Mn0.3O2− δ and Ce0.7 Fe0.3O2− δ solid solutions for diesel soot oxidation, Appl. styrene oxide over nickel and copper ceria solution combustion catalysts, MATEC
Catal. A: Gen. 488 (2014) 1–10. Web of Conf. 374 (2023) 01004.
[49] S.K. Meher, G.R. Rao, Tuning, via counter anions, the morphology and catalytic
activity of CeO2 prepared under mild conditions, Colloid Interface Sci. 373 (1)
(2012) 46–56.

12

You might also like