Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of the Energy Institute 112 (2024) 101458

Contents lists available at ScienceDirect

Journal of the Energy Institute


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

The facilitating effect of sulfide treatment coupled sol-gel method on


NH3-SCR activity of Fe–Mn/TiO2 catalysts
Junqiang Xu a, Tao Zheng a, Xianlin Zou a, Hong Shen a, Fang Guo a, *, Qiang Zhang a,
Minghua Duan b
a
School of Chemistry & Chemical Engineering, Chongqing University of Technology, Chongqing, 400054, China
b
SPIC Yuanda Environmental Protection of Catalyst Co., Ltd, Chongqing, 401336, China

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

Handling Editor: Paul Williams The 5Fe–3Mn–S/TiO2-SG catalysts were synthesized by sulfide treatment coupled sol-gel method to develop new
NH3-SCR catalysts with wide temperature window, good low-temperature activity and environment-friendly.
Keywords: The evolution of catalyst surface chemical state, morphology, structure, acidity and interaction forces were
NH3-SCR systematically investigated by XPS, BET, XRD, SEM, H2-TPR, NH3-TPD, UV, IR and steady-state kinetic tests. The
Fe–Mn/TiO2 catalyst
influence of sulfide treatment coupled sol-gel method on NO conversion was explored. XRD, BET and SEM
Sulfide treatment
characterization revealed that sulfide treatment resulted in smaller catalyst particles and higher dispersion,
DFT
which increased the specific surface area. XPS, NH3-TPD and steady-state kinetic tests results demonstrated that
the sulfide treatment catalysts had a higher surface adsorbed oxygen (Oα) content, which could produce more
oxygen vacancies on the catalyst surface, thus increasing the number of acid sites. H2-TPR and UV results
demonstrated that the 5Fe–3Mn–S/TiO2-SG catalysts exhibited the strongest contact between the carrier and the
active component. The 5Fe–3Mn–S/TiO2-SG catalyst obtained more than 80% NO conversion in the operational
temperature range of 220–420 ◦ C, according to the NH3-SCR reaction data. The temperature window was
expanded by 120 ◦ C compared to conventional Fe–Mn/TiO2 catalysts, and 98% conversion was achieved at
300 ◦ C. Meanwhile, DFT calculations of the SO2- 4 -MnFe2O4(311) crystalline surface of the 5Fe–3Mn–S/TiO2-SG
catalysts revealed that it could form stronger chemisorption for both NH3 and NO, which promoted the NH3-SCR
reaction more effectively. The DFT calculation results and the experimental findings agreed well.

1. Introduction environmental friendliness, Fe-based catalysts have received a lot of


attention. However, the narrow working temperature window and low
NO is a significant class of atmospheric pollutants that have a activity of single Fe-based catalysts cannot meet the requirements of
negative impact on human health, plant and animal growth, and cause practical applications. Therefore, elemental doping and other methods
serious environmental issues like haze and photochemical smog [1,2]. are usually used to improve the catalytic activity and N2 selectivity of
One of the best solutions for reducing NO emissions is NH3-SCR, which iron-based oxide catalysts. As an illustration, Sun et al. [4] enhanced the
has been used extensively in industry. Currently, V2O5-WO3/TiO2 is the surface acidity of Fe2O3 catalyst by Sm doping, which strengthened the
main high-efficiency SCR catalyst on the market. It’s a pity, this catalytic adsorption of NH3 while lowering the activation energy of NH3,
system does have significant drawbacks, though, including a very small increasing denitrification efficiency by an 11 times factor. Chen et al. [5]
operating temperature range of 300–400 ◦ C, toxic and challenging va­ doped Ti4+ onto the octahedral active site of γ-Fe2O3, which reduced the
nadium material recovery, low N2 selectivity at high temperatures, and entropy of the crystal and thus improved the thermal stability. Current
astronomical cost of use [3]. Therefore, the development of green and literatures [6–8] about modified Fe-based catalysts have focused on
new NH3-SCR catalysts with wide temperature window and good low metal element doping. However, relatively few studies have been con­
temperature activity has received important attention in denitrification. ducted on the effect of sulfate on the SCR performance of denitrification
Due to their strong medium and high temperature activity and catalysts, and most of them focus on the introduction of sulfate to

* Corresponding author. School of Chemistry & Chemical Engineering, Chongqing University of Technology No.66 Hongguang Rd., Banan, Chongqing, 400054,
China.
E-mail address: guofang@cqut.edu.cn (F. Guo).

https://doi.org/10.1016/j.joei.2023.101458
Received 18 August 2023; Received in revised form 2 November 2023; Accepted 3 November 2023
Available online 10 November 2023
1743-9671/© 2023 Published by Elsevier Ltd on behalf of Energy Institute.
J. Xu et al. Journal of the Energy Institute 112 (2024) 101458

enhance the sulfur resistance of the catalysts. Yu et al. [9] introduced using a BSD-C200 automated chemisorption analyzer (BEISHIDE, Bei­
CuSO4 as a sulfur source to make the catalyst exhibit higher sulfur jing, China). A 60 mg sample was placed in a helium atmosphere with
resistance. Qi et al. [10] effectively enhanced the SO2 resistance by 300 ◦ C for 60 min, cooled and then passed through a H2/Ar mixture at
ferrying an atomic layer deposition (ALD) film of Fe2O3 over the surface 10 ◦ C/min to test the H2-TPR. NH3 programmed temperature desorption
of the catalyst, followed by sulfation. Wang et al. [11] Sulfate modifi­ analysis was performed using a BSD-C200 Automatic Chemisorption
cation could significantly broaden the SCR reaction temperature win­ Analyzer (BEISHIDE, Beijing, China) to measure the surface acidity and
dow while improving N2 selectivity. ammonia adsorption capacity of the catalyst. The SCR catalyst (60 mg)
According to the above reports, the modification of the surface was incubated at 300 ◦ C for 60 min in a helium (He) atmosphere, cooled
structure and NH3-SCR performance of Fe–Mn/TiO2 catalysts by sulfide to 50 ◦ C, and then maintained or 60 min with a 5% NH3/He mixture.
treatment is not clear. In this study, Fe–Mn–S/TiO2-SG catalysts were Finally, NH3-TPD was tested by heating the catalyst (10 ◦ C/min) from
synthesized by sulfide treatment coupled sol-gel method, and the XPS, 50 ◦ C to 800 ◦ C in a helium (He) atmosphere.
BET, XRD and H2-TPR techniques were used to thoroughly examine the The UV–vis of the catalyst was tested in a UV–Vis photometer (model
Fe–Mn–S/TiO2-SG catalysts. By combining the activity assessment of UV-27001) using BaSO4 as a reference sample. UV–visible spectral
NH3-SCR with density generalization theory (DFT) calculations, the conversion using Kubelka-Munk function. The functional groups of the
impact of sulfide treatment on the functionality of NH3-SCR was further catalysts were detected by IRTracer-100 (Shimadzu, Japan) type FTIR
studied. spectrometer for the region of 4000-400 cm− 1.

2. Experimental
2.3. Experimental conditions

2.1. Preparation of the catalyst


The NO conversion of the catalyst was measured by a flue gas
analyzer (Vario Plus, MRU GmbH, Germany). Firstly, 200 mg of catalyst
2.1.1. Impregnation process
was compressed and crushed and sieved to 20–40 mesh. Then it was
By using the impregnation process, 5Fe–3Mn–S/TiO2 catalysts were
installed in a fixed-bed reactor (XIANKE, Shandong, China) waiting for
created. Deionized water was used to dissolve the appropriate amounts
testing, and the testing temperature range was 150oC–450 ◦ C. The re­
of (NH4)2SO4 (99%, AR), Mn(NO3)2⋅4H2O (98%, AR), and Fe
action gases composed of 400 ppm of NO, 400 ppm of NH3, and 3% of
(NO3)3⋅9H2O (98.5%, AR) at a molar ratio of S:Mn:Fe = 1:3:5. The
O2, while an appropriate amount of AR was passed to maintain equi­
combined solution was then given 2.0 g of TiO2. After impregnation, the
librium. The overall volumetric flow rate was 100 mL min− 1, or 30,000
solution was continuously magnetically stirred at 80 ◦ C for 4 h. The
h− 1 of GHSV. Equation (1) is used to determine the conversion rate of
samples were dried in a 110 ◦ C oven for 12 h, and then kept at 500 ◦ C
NO.
with a 5 ◦ C/min gradient for 5 h in an air-conditioned muffle furnace.
The same process was used to create catalysts with various sulfur con­ [NO]in − [NO]out
NO conversion(%) = ×100% (1)
centrations, identified as 5Fe–3Mn-xS/TiO2, where x stands for the [NO]in
molar proportion of sulfur to iron. For activity assessment, all catalysts
were crushed and sieved to a 20–40 mesh size. The catalyst without the
2.4. Kinetic studies
addition of (NH4)2SO4 was named as 5Fe–3Mn/TiO2.
The following circumstances guided the steady-state kinetic studies:
2.1.2. Sol-gel process
400 ppm NH3 and NO, 3% O2, total gas flow rate of 50–200 mL/min,
By using the sol-gel process, the 5Fe–3Mn–S/TiO2-SG catalyst was
using 200 mg 0.18–0.85 mm particle size, in denitrification reaction
created. Firstly, 10 mL of C16H36O4Ti (98.5%, AR) was added to 10 mL
efficiency of less than 20%, so it could be regarded as a differential
of anhydrous ethanol and dispersed to prepare A solution. Then,
system. Therefore, the denitrification kinetics of Fe–Mn–S/TiO2 catalyst
appropriate amounts of (NH4)2SO4 (99%, AR), Mn(NO3)2⋅4H2O (98%,
could be studied under this reaction condition. Equation (2) was used to
AR), Fe(NO3)3⋅9H2O (98.5%, AR) were dissolved in a mixture of 5 mL of
determine the number of reaction stages, and the activation energy
anhydrous ethanol, 5 mL of acetic acid and 1 mL of water to prepare
required for the SCR denitrification reaction was calculated from
solution B, where the molar ratio of S:Mn:Fe was 1:3:5. Drop by drop,
Arrhenius equation (3):
solution B was poured into solution A while being sufficiently stirred to
create a sol, and the gel was formed after aging for 3 h. The catalyst was x
RNO = − kCNO y
CNH 3
COz 2 (2)
dried at 110 ◦ C for 12 h and calcined at 500 ◦ C for 5 h. The 5Fe–3Mn–S/
TiO2-SG catalyst was obtained. k = k0 e− Ea /(Rg T )
(3)

2.2. Catalysts characterization


2.5. Computing techniques
X-rays were obtained with Cu-Kα radiation on a Shimadzu (MPD)
instrument in Japan, and the diffraction pattern was measured at a scan The Material Studio 2019 software’s CASTEP module served as the
rate of 2o/min from 10o to 80o. foundation for all DFT computations, a fundamental quantum me­
The pore structures of the catalysts were determined by a low tem­ chanics program designed for use in solid materials science. PW91 under
perature liquid nitrogen adsorption and desorption apparatus (JW- the GGA was used in the calculation, using a 1 × 1 × 1 Gamma grid to
BK132, JWGB SCI. &TECH, China). Additionally, the samples’ specific sample the Brillouin zone. The plane wave truncation energy is set to
surface area and pore size were determined using the BET and BJH 500 eV, and the convergence criterion for the atomic forces was set to
methods, respectively. 0.05 eV/Å. In order to define the adsorption energy:
The SEM images of all samples were acquired by field emission Eads = E(adsorbate + surface)-E(adsorbate)-E(surface)
scanning electron microscope ((JEOL 7800F, Japan).
Qualitative and semi-quantitative valence distribution of each where E(adsorbate + surface) stands for the total energy of reactants
element on the catalyst surface by X-ray photoelectron spectroscopy (Al- adsorbed onto the model surface, and E(adsorbate) and E(surface)
Kα radiation, hν = 1486.46 eV). Utilizing the C 1s peak’s (284.8 eV) denote the energy of reactants in the free state and the pure model
calibration of elemental binding energies in catalysts. surface, respectively, where the larger negative value of Eads indicates
H2-programmed temperature reduction analysis was performed the stronger adsorption performance of the adsorbate.

2
J. Xu et al. Journal of the Energy Institute 112 (2024) 101458

Fig. 1. Catalytic performance of NH3-SCR with sulfide treatment catalysts: (a) (b) NO conversion, (c) Arrhenius plots for NO reduction.

Table 1
Structural parameters of the catalyst.
catalysts SBET Vp Dp TiO2 crystallite
(m2•g− 1)a (cm3•g− 1)b (nm)c size (nm)

5Fe–3Mn/TiO2 47.8 0.35 17.4 20.7


5Fe–3Mn–S/ 51.7 0.36 15.8 16.5
TiO2
5Fe–3Mn–S/ 52.5 0.38 18.2 16.2
TiO2-SG
a
BET surface areas.
b
BJH desorption pore volume.
c
Average Pore Size.

all catalysts above 300 ◦ C displayed an increasing trend, indicating that


the addition of ammonium sulfate has a positive contribution to the
reaction. When the ratio of sulfate content to active component content
was 0.25, 0.5, and 1, the catalysts showed a significant increase in high-
temperature catalytic activity above 300 ◦ C. However, the catalyst’s
denitrification activity below 300 ◦ C drastically diminished with the
Fig. 2. XRD patterns of the samples. addition of excessive ammonium sulfate. such as the ratio of 1.5, 2.5,
and 3.5, and the addition of too much surface ammonium sulfate sup­
3. Results pressed the catalyst’s low-temperature performance. The NO conversion
of Fe–Mn/TiO2 catalyst at 400 ◦ C was 21%, and when the ratio of
3.1. NH3-SCR activity of the catalysts ammonium sulfate content was 1, the NO conversion increased to 84%,
while the low-temperature activity was basically maintained. Therefore,
Fig. 1(a) illustrates the NO conversions utilizing various catalysts as a comparing from the activity diagram, the best denitrification perfor­
function of temperature at 150–450 ◦ C. With the increase of ammonium mance of Fe–Mn/TiO2 catalyst was achieved when the ratio of ammo­
sulfate((NH4)2SO4) content, the high-temperature catalytic activity of nium sulfate to active component was 1. The highest denitrification

3
J. Xu et al. Journal of the Energy Institute 112 (2024) 101458

was widened by 30 ◦ C.
To clarify the differences in the intrinsic activity on Fe, Fe–Mn/TiO2
and Fe–Mn–S/TiO2 catalysts, the NH3-SCR kinetic experiments were
carried out in the dynamic zone (NO conversion below 20%) [12,13].
The final kinetic model Equation (4) was obtained after measuring three
feed gas reaction stages. The calculated results were presented in Fig. 1
(c). The comparison revealed that sulfide treatment slightly increased
the activation energy (Ea) of the catalyst and shifted the operating
temperature to the high temperature range. It indicated that the surface
acidity of the sulfide treatment catalyst increased significantly, which
would negatively affect the redox performance of the catalyst and
properly suppress the low temperature activity, which agreed with the
information from the activity diagram.

(4)
36435.32
rNO = − kCNO = 39340.11e− RT CNO0 (1 − X)

3.2. Catalyst crystallization, pore size and morphology

Fig. 3. N2 adsorption-desorption isotherms of the samples. The XRD patterns of all samples were shown in Fig. 2. All the iden­
tified peaks were observed to be consistent with anatase TiO2 [14]. The
activity of 97% was achieved at a reaction temperature of 300 ◦ C. More strong diffraction peaks of 5Fe–3Mn/TiO2 catalyst at 25.3◦ , 37.9◦ , 48.1◦ ,
than 80% conversion of NO was achieved in the temperature range of 54.0◦ , 55.1◦ , 62.7◦ , 68.9◦ and 75.2◦ attachments belonged to anatase
220–420 ◦ C. After determining the optimal ammonium sulfate content, (101), (004), (200), (105), (211), (204), (116) and (215) diffractions,
the catalyst was optimized by the sol-gel method. As shown in Fig. 1(b), respectively [JCPDS No. 21–1272]. The characteristic diffraction peaks
the temperature window of 5Fe–3Mn–S/TiO2 catalyst T80 = 220–420 ◦ C of rutile TiO2 appeared at 27.45◦ , 36.09◦ and 41.23◦ [JCPDS No.
21–1276] [15]. Faint MnFe2O4 crystalline form appeared at 34.88◦

Fig. 4. SEM images of (a) (b) (c) 5Fe–3Mn/TiO2. (d) (e) (f) 5Fe–3Mn–S/TiO2. (g) (h) (i) 5Fe–3Mn–S/TiO2-SG at different magnifications.

4
J. Xu et al. Journal of the Energy Institute 112 (2024) 101458

Fig. 5. XPS spectra of (a) Ti 2p, (b) O 1s, (c)Fe 2p, (d) Mn 2p, (e) S 2p of the catalysts.

(311), 17.99◦ (111) [JCPDS No. 38–0430]. In contrast, the peak in­ catalytic performance at low temperatures and that rutile TiO2 only
tensity of the 5Fe–3Mn–S/TiO2 catalyst showed a significant decrease. contributes a modest percentage to the catalytic activity [16].
5Fe–3Mn–S/TiO2-SG catalysts were not observed in MnFe2O4 crystalline In addition, the Scherrer equation was used to calculate the crys­
form, while no more distinctive peaks were produced, indicating that tallites’ sizes, which relied on the characteristic peak at 25◦ for the (101)
the sulfide treatment catalysts possessed better surface dispersion. reflection. Table 1 revealed that Sulfide treatment reduced the size of
Related research has revealed that anatase TiO2 can greatly boost anatase grains. This was because sulfide treatment damaged the lattice

5
J. Xu et al. Journal of the Energy Institute 112 (2024) 101458

Table 2 Table 3
Summarize the atomic ratio (%) of each type in different samples. Acidity of the all catalysts determined by NH3-TPD.
Catalysts Oα/ Fe2+/(Fe3++ Mn4+/(Mn4++ Catalysts Surface acidity/μmol NH3 g−cat1
(Oα+Oβ) Fe2+) Mn2+)
Weak Medium strong and strong Total
5Fe–3Mn/TiO2 9.8 55.8 41.5
5Fe–3Mn/TiO2 33.6 80.5 114.1
5Fe–3Mn–S/TiO2 15.5 56.4 49.7
5Fe–3Mn–S/TiO2 39.4 191.8 231.2
5Fe–3Mn–S/TiO2- 20.7 59.6 52.3
5Fe–3Mn–S/TiO2-SG 43.9 131.7 175.6
SG

Ti 2p3/2 of the 5Fe–3Mn–S/TiO2-SG catalyst was slightly shifted toward


of TiO2 and prevented the formation of TiO2 particles. Numerous
research had demonstrated that the stronger the catalytic activity, the the lower binding energy, indicating that sulfide treatment and sol-gel
method optimization resulted in higher electron cloud density of the
better the active component dispersion, and the smaller the catalyst
grain size [17,18]. Therefore, the 5Fe–3Mn–S/TiO2-SG catalyst pre­ catalyst titanium species.
The catalyst’s O 1s energy spectrum was shown in Fig. 5(b). The
pared by sulfide treatment coupled sol-gel method possessed better
NH3-SCR activity. characteristic peak at 529.8 eV was attributed to Oβ, and the charac­
teristic peak at 531.3 eV was attributed to Oα [25]. It was observed that
The N2 adsorption/desorption isotherm of the catalyst was displayed
in Fig. 3. The essential parameters of the samples are listed in Table 1. the electron binding energy of O 1s in sulfide treatment catalysts shifted
to lower binding energy, indicated that sulfide treatment increased the
All three catalysts were discovered to contain type IV isotherms, a
characteristic mesoporous structure, with either H2-type or H1:H2-type electron cloud density of Oβ and Oα. Table 2 showed that the
Oα/(Oα+Oβ) ratio of the 5Fe–3Mn–S/TiO2-SG catalyst was 20.7, which
hysteresis loops, demonstrating that the catalysts possess a full TiO2
mesoporous structure [19,20]. Table 1 showed that the 5Fe–3Mn–­ was higher than the other samples. Studies demonstrated that Oα mi­
grates faster and possesses lower binding energy than Oβ, which could
S/TiO2-SG catalyst had a BET specific surface area of 52.5 m2 g− 1, a
desorption pore volume of 0.38 cm3 g− 1, and a pore width of 18.2 nm. In accelerate the redox reaction in SCR reaction [26], It is suggested that
both sulfide treatment and sol-gel optimization could promote the for­
general, the specific surface area of catalysts was related to the number
of active sites and the dispersion of components. The larger the specific mation of oxygen vacancies, which was a significant factor in the
surface area, the better the catalytic performance of catalyst NH3-SCR
[21,22]. The results of the BET and the activity test were in agreement.
The catalysts’ shape and particle size were assessed, and the findings
were depicted in Fig. 4. Both 5Fe–3Mn/TiO2 catalysts and sulfide
treatment catalysts were composed of irregular particle aggregates,
which contained massive pore structures. The surface morphology of the
5Fe–3Mn–S/TiO2 catalysts did not change significantly compared to the
5Fe–3Mn/TiO2 catalysts, which still showed crystalline particles. The
5Fe–3Mn–S/TiO2-SG catalysts showed significant morphological
changes, showing a larger pore structure on the scale of 1 μm and 500
nm. In Fig. 4(i), the 5Fe–3Mn–S/TiO2-SG catalyst had smaller crystallite
size and more compact position on the 100 nm scale, resulting in larger
pores.

3.3. Surface chemical states of the catalysts

As can be seen in Fig. 5, XPS was used to examine the surface


chemical states of all samples. Fig. 5 (a) showed the Ti 2p energy spectra
of the catalysts. All catalysts exhibited three characteristic peaks around
458.6 eV, 464.4 eV and 471.8 eV corresponding to Ti 2p3/2, Ti 2p1/2
and satellite peaks, which were attributed to the Ti4+ state [23,24]. The Fig. 7. UV–Vis diffuse reflectance spectra of catalysts.

Fig. 6. H2-TPR (a), NH3-TPD (b) diagrams of the samples.

6
J. Xu et al. Journal of the Energy Institute 112 (2024) 101458

Table 4
Castep Optimize the pattern parameters of the geometric configuration.
Unit Cell symmetry group Lattice constant(Å) lattice angle(degree)

MnFe2O4 FD-3M a=8.16 α=90


b=8.16 β=90
c=8.16 γ=90

to identify the type of sulfur species present in the catalyst after sulfide
treatment, the chemical state of S was examined by XPS. The energy
spectrum peaked at 168.7 eV, with S6+ being responsible for 168.8 eV
[31]. Considering that the sulfate introduced during the preparation
process was thermally stable after calcination, the presence of sulfur
species in the catalyst was considered to be in the form of sulfate SO2− 4
[32], which agreed with the results of the FTIR analysis. The result was
an increase in the catalyst’s acidity, making NH3 easier to adsorb and
activate in the SCR reaction [31].

3.4. Surface acidity and redox properties of the catalysts


Fig. 8. FTIR spectra of industrial catalysts.

The redox properties of catalysts have important impact on the


catalyst’s outstanding catalytic activity. denitrification reaction because it could act as a mediator of the reaction
Fig. 5(c) showed the Fe 2p energy spectrum of the catalyst. The two [33]. Fig. 6(a) showed the H2-TPR curves of the three catalysts. The
pairs of peaks were Fe3+ (about 712.9 eV and 726.3 eV) and Fe2+ (about reduction peaks of the 5Fe–3Mn/TiO2 catalysts at 310 ◦ C and 375 ◦ C
710.2 eV and 723.3 eV), respectively [27]. The ratio of were attributed to the reduction of Fe2O3 to Fe3O4 [34], while the
Fe2+/(Fe3++Fe2+) of the samples was calculated from the peak regions hydrogen consumption peaks at 720–800 ◦ C were attributed to the
and presented in Table 2. The results demonstrated that both sulfide reduction of Ti4+→Ti3+ [35,36]. By comparison, it was not difficult to
treatment and sol-gel method optimization promoted the formation of find that the reduction peaks of the 5Fe–3Mn–S/TiO2 and 5Fe–3Mn–­
Fe2+. It was shown that the intermediate valence of Fe2+ facilitated the S/TiO2-SG catalysts were significantly enhanced compared to the con­
formation of unsaturated bonds and enhanced electron mobility, which ventional 5Fe–3Mn/TiO2 catalysts, and the increase in the intensity of
improved the NH3-SCR activity of the catalyst [27,28]. In summary, the the H2 depletion peak could be attributed to the increase in the number
5Fe–3Mn–S/TiO2-SG catalyst revealed the best catalytic activity, and of surface oxygen species due to the increase in the lattice structure
the catalyst’s NH3-SCR performance were constant. defects on the catalyst surface. These results indicated that sulfide
The XPS energy spectrum of Mn 2p was shown in Fig. 5(d). It could treatment had accelerated the transfer and transmission of electrons and
be found that two valence states of Mn were present mainly in the promoted the formation of activation molecules, thus reduced the acti­
catalyst, located near 641.4 eV, 652.5 eV (Mn2+) and 643.9 eV, 652.2 eV vation energy of the reaction and accelerating it towards thermody­
(Mn4+), respectively [29]. Based on the peak areas, a significant increase namic equilibrium. There were strong reduction peaks at 424 ◦ C and
in Mn4+ content was found for the sulfide treatment catalysts. The cat­ 418 ◦ C, which were thought to be a superimposed reduction of sulfate
alyst’s redox capacity might be improved by more Mn4+, which would ions and Fe species [37–39]. Also the reduction peaks shift to higher
increase conversion efficiency in the low temperature range [30], which temperatures suggested that sulfide treatment enhanced the interaction
was evident from the NO conversion results. forces between the active ingredient and the carrier. In conclusion,
Fig. 5(e) displayed the S 2p energy spectrum of the catalyst. In order redox properties were the main influencing factor for SCR reactions at
low temperatures and one of the key reasons for the increased intensity
of the reduction peak of the catalyst due to sulfide treatment.
The surface acidity has a significant impact on the adsorption and
activation of NH3 on the catalyst [40]. Therefore, the NH3-TPD method
was used to investigate the surface acidity of the catalyst (Fig. 6(b)). Low
temperature NH3 desorption peaks belonging to weak acidic sites
desorption were observed for all samples at around 105 ◦ C due to the
breakage of weak hydrogen bonds between NH3 and acidic groups
adsorbed on the catalyst surface [41,42]. The high temperature NH3
desorption peaks at 250–500 ◦ C were attributed to medium to strong,
strong acid sites [43,44]. The platforms corresponding to the attachment
at 600 ◦ C could be labeled as Lewis acid sites [45–47]. Table 3 contains
the numbers of both acid sites determined from the NH3-TPD data,
which was evident that there were much more sulfide treatment cata­
lytic acid sites. It was determined that the weak acidic sites made it

Table 5
Adsorption energy of NH3 and NO at different crystal surfaces.
Crystal Surface Eads/eV

NH3 NO

MnFe2O4(111) − 0.06 − 0.70


MnFe2O4(311) − 0.73 − 2.37
SO2-
4 -MnFe2O4(311) − 1.76 − 3.86
Fig. 9. Optimized geometric configuration of MnFe2O4.

7
J. Xu et al. Journal of the Energy Institute 112 (2024) 101458

Fig. 10. The structure of NH3 and NO adsorption on Fe atom of MnFe2O4(311) (a) (b) and MnFe2O4 (111) (c) (d) surface.

results could explain that the 5Fe–3Mn–S/TiO2-SG catalysts had


exhibited better high-temperature SCR activity.

3.5. Surface interactions

To elucidate the interaction between the S component and the


catalyst, the samples were examined using UV–Vis spectroscopy, and the
findings were shown in Fig. 7. The absorption edge of the unloaded
carrier TiO2 was located at 354 nm [52]. The TiO2 absorption edge of
5Fe–3Mn/TiO2 shifted to 396 nm. Compared with 5Fe–3Mn–S/TiO2
catalyst, the TiO2 absorption edge of the catalyst moved toward the
long-wave direction and the electron energy level on the catalyst surface
became higher. This indicated that electrons move more easily from the
valence band to the conduction band, thus forming electron-hole pairs,
which may account for the better low-temperature activity of the
5Fe–3Mn/TiO2 catalyst compared to the 5Fe–3Mn–S/TiO2 catalyst. It
had been demonstrated that the peaks between 300 and 400 nm were
oligomeric Fe oxide clusters, while the absorption edge of Fe2O3 parti­
cles was above 400 nm, showing that improving the preparation process
may also help the catalyst’s active ingredients be distributed more
evenly.

3.6. FTIR characterization


Fig. 11. The structure of NH3 and NO adsorption on Fe atom of SO2-
4-
MnFe2O4(311) surface.
Infrared spectroscopy could effectively detect the formation of sul­
fides. Fig. 8 showed the IR spectra of each catalyst. The stretching and
easier for NH3 to desorb at low temperatures [48]. The comparison in bending vibrations of the surface-OH group were found at 3426 and
the figure indicated that the desorption peak of the Lewis acid sites of the 1627 cm− 1 [53], the Ti–O stretching vibration at 690 cm− 1, and the
5Fe–3Mn–S/TiO2-SG catalyst was significantly stronger than that of bending vibration of the metal oxide (M − OH) hydroxyl group at 1127
5Fe–3Mn–S/TiO2. The homogeneous dispersion of active ingredients, cm− 1 [54]. The characteristic peaks appearing at 981 and 985 cm− 1 for
which is essential to the high temperature denitrification process, may the sulfide treatment catalyst were correspond to SO2−4 [32], while the
also be promoted by the strong acid site [49–51]. Therefore, these 5Fe–3Mn/TiO2 catalyst did not had a distinct characteristic absorption

8
J. Xu et al. Journal of the Energy Institute 112 (2024) 101458

Fig. 12. Partial density of states on the surfaces of MnFe2O4(311) and SO2-
4 -MnFe2O4(311) for (a) N of NH3 molecule, (b) N of NO molecule, (c) Fe on surfaces, (d) N
and Fe when NH3 absorbed, (e)N and Fe when NO absorbed.

Fig. 13. Possible mechanism of NH3-SCR on 5Fe–3Mn–S/TiO2-SG.

peak here, indicating the introduction of SO2−


4 in the sulfide treatment and NO. The adsorption energy Eads values for NH3 and NO on the
catalyst material. MnFe2O4(311) crystal surfaces were − 0.73ev and − 2.37eV, and on the
MnFe2O4(111) crystal surfaces were − 0.06eV and − 0.70eV as seen in
table 5. The Eads value of NO was smaller than the Eads value of NH3,
3.7. DFT results which showed that NO adsorption on the catalyst was significantly
stronger than NH3 adsorption. The Eads values of MnFe2O4(311) crystal
The original crystal model MnFe2O4 used in this study was obtained surfaces became smaller, which indicated that the adsorption capacity of
from the CCDC database. The model was geometrically optimized using MnFe2O4(311) crystal surfaces was stronger and more favorable for the
CASTEP to completely relax all atomic and lattice parameters in the adsorption behavior of SCR reaction to occur. Therefore, MnFe2O4(311)
model. Fig. 9 showed the optimized MnFe2O4 crystal model, and the crystal surfaces were used in the subsequent study.
relevant lattice parameters were listed in Table 4. The crystal faces of The above characterization analysis revealed that the physico­
MnFe2O4(311) and MnFe2O4(111) were cut on the crystal model that chemical properties of the sulfide treatment catalyst surface were
reached the steady state after optimization. In addition, a vacuum layer changed. Sulfur species in the form of SO2− 4 were produced, which
of 15 Å was established to prevent the interaction forces between the interacted electronically mainly with Fe and Mn in the catalyst. The
layers SO2-
4 -MnFe2O4(311) crystal model was developed as shown in Fig. 11.
Fig. 10 showed the structures of (a) (b) MnFe2O4(311) crystalline The adsorption energy Eads values of NH3 and NO on the SO2- 4-
surface and (c) (d) MnFe2O4(111) crystalline surface adsorbed with NH3

9
J. Xu et al. Journal of the Energy Institute 112 (2024) 101458

MnFe2O4(311) crystal surface were − 1.76ev and − 3.86eV. Compared analysis demonstrated that SO2-
4 -MnFe2O4(311) crystal surface was more
with the MnFe2O4(311) crystalline surface, the adsorption energy Eads favorable for the adsorption of NH3 and NO, yielding a catalyst with
value decreased. It indicated that sulfide treatment enhanced the good NH3-SCR activity.
adsorption ability and surface acidity of NH3 and NO, and improved the
reaction rate, causing the catalyst’s operating temperature to rise into Declaration of competing interest
the high temperature zone. Also, it agreed with the catalyst’s activity
data. The authors declare that they have no known competing financial
As a result of the introduction of PDOS, the orbital modifications and interests or personal relationships that could have appeared to influence
variations of adsorbed molecules and active sites were further investi­ the work reported in this paper.
gated as shown in Fig. 12. It was discovered that both 2p and 2s orbitals
of N moved to the deep energy level after adsorption, reducing the Acknowledgments
system energy. At the same time, the 3d orbitals of Fe and 2p of N
produce overlap, indicating that both NH3 and NO could be chemisorbed The authors thank the National Natural Science Foundation of China
on the MnFe2O4(311) crystal surface and SO2- 4 -MnFe2O4(311) crystal (No.21902017), Key project of science and technology research program
surface. Only compared with the adsorption of NH3, the adsorption of of Chongqing Education Commission of China (No KJZD-M202101101),
NO was more stable because the orbitals of 2p of N and 3d of Fe in NO the Project of fundamental research and frontier exploration of
overlapped more and completely overlapped under the action of 1.8 eV, Chongqing (cstc2019jcyj-msxmX0052), Youth project of science and
and the p and d orbitals of N and Fe atoms were hybridized, which has technology research program of Chongqing Education Commission of
leaded to stronger electronic interactions. China (KJQN20211107), Foundation of technological innovation and
By comparing the PDOS maps of the two models of MnFe2O4(311) application development of Chongqing (cstc2021jscx-msxmX0308), and
crystal surface and SO2- 4 -MnFe2O4(311) crystal surface, the peak split­ Scientific Research Foundation of Chongqing University of Technology
ting of the d-orbit PDOS of the Fe atoms near the Fermi energy level of (2020ZDZ022).
the SO2-4 -MnFe2O4(311) crystal surface was found. The electron inter­
action orbitals in the core of Fe atom were divided from two peaks to Appendix A. Supplementary data
three peaks. Furthermore, the peak positions were moved to higher
energy levels, resulting in enhanced acidity of the catalyst. The results Supplementary data to this article can be found online at https://doi.
revealed that the Fe atoms on the crystalline surface of SO2- 4- org/10.1016/j.joei.2023.101458.
MnFe2O4(311) were more favorable for the adsorption of NH3 and NO,
which resulted in a better NH3-SCR activity of the catalyst. References
Therefore, it could be determined from the above characterization
and DFT that the sulfide treatment improved the catalyst acidity, [1] I. Song, H. Lee, S.W. Jeon, I.A.M. Ibrahim, J. Kim, Y. Byun, D.J. Koh, J.W. Han, D.
causing the catalyst’s high temperature NH3-SCR activity to significantly H. Kim, Simple physical mixing of zeolite prevents sulfur deactivation of vanadia
catalysts for NO(x) removal, Nat. Commun. 12 (1) (2021) 901, https://doi.org/
rise. Fig. 13 exhibits the probable SCR mechanism for this catalyst. 10.1038/s41467-021-21228-x.
Firstly, the redox cycle (Fe3+ + Mn2+ ↔ Fe2+ + Mn4+) increased the [2] T. Karl, C. Lamprecht, M. Graus, A. Cede, M. Tiefengraber, J. Vila-Guerau de
mobility of oxygen, thus generated more oxygen vacancies. The unsat­ Arellano, D. Gurarie, D. Lenschow, High urban NOx triggers a substantial chemical
downward flux of ozone, Sci. Adv. 9(3) eadd2365,http://doi.org/10.1126/sciadv.
urated coordination between Fe3+ and O2− resulted in the formation of add2365.
acid sites on Fe2O3, so that NH3 was adsorbed mainly on Fe2O3. How­ [3] Y. Yu, X. Yi, J. Zhang, Z. Tong, C. Chen, M. Ma, C. He, J. Wang, J. Chen, B. Chen,
ever, Sulfide treatment led to the disruption of the unsaturated coordi­ Application of ReOx/TiO2 catalysts with excellent SO2 tolerance for the selective
catalytic reduction of NOx by NH3, Catal. Sci. Technol. 11 (15) (2021) 5125–5134,
nation sites on Fe2O3, new acid sites were created by SO2− 4 on the https://doi.org/10.1039/d1cy00467k.
surface, which pushed NH3 adsorption to SO2− 4 on the catalyst surface. [4] C. Sun, W. Chen, X. Jia, A. Liu, F. Gao, S. Feng, L. Dong, Comprehensive
understanding of the superior performance of Sm-modified Fe2O3 catalysts with
regard to NO conversion and H2O/SO2 resistance in the NH3-SCR reaction, Chin. J.
4. Conclusion Catal. 42 (3) (2021) 417–430, https://doi.org/10.1016/s1872-2067(20)63666-x.
[5] J. Chen, W. Qu, Y. Chen, X. Liu, X. Jiang, H. Wang, Y. Zong, Z. Ma, X. Tang,
The 5Fe–3Mn–S/TiO2-SG catalysts prepared by sulfide treatment Simultaneously enhancing stability and activity of maghemite via site-specific Ti
(IV) doping for NO emission control, ChemCatChem 10 (20) (2018) 4683–4688,
coupled sol-gel method were characterized by wide temperature win­
https://doi.org/10.1002/cctc.201801169.
dow, good low-temperature activity and environment-friendly. The [6] S. Li, X. Chen, F. Wang, Z. Xie, Z. Hao, L. Liu, B. Shen, Promotion effect of Ni
Fe–Mn/TiO2 catalysts’ NH3-SCR performance was improved, and the doping on the oxygen resistance property of Fe/CeO2 catalyst for CO-SCR reaction:
active temperature window was expanded, thanks to the proper quantity activity test and mechanism investigation, J. Hazard Mater. 431 (2022), 128622,
https://doi.org/10.1016/j.jhazmat.2022.128622.
of sulfide treatment, but an excess of S may result in a rise in operating [7] L. Li, J. Ji, W. Tan, W. Song, X. Wang, X. Wei, K. Guo, W. Zhang, C. Tang, L. Dong,
temperature into the high temperature region. The catalysts had the best Enhancing low-temperature NH3-SCR performance of Fe–Mn/CeO2 catalyst by
NH3-SCR activity when the S:Fe molar ratio was 1:5. Compared with the Al2O3 modification, J. Rare Earths 40 (9) (2022) 1454–1461, https://doi.org/
10.1016/j.jre.2021.08.012.
5Fe–3Mn/TiO2 catalyst, the 5Fe–3Mn–S/TiO2-SG catalyst was more [8] Z. Chen, Q. Liu, L. Guo, S. Zhang, L. Pang, Y. Guo, T. Li, The promoting mechanism
uniformly distributed and possessed a larger specific surface area, while of in situ Zr doping on the hydrothermal stability of Fe-SSZ-13 catalyst for NH3-SCR
the interaction between the active component and the carrier in the reaction, Appl. Catal., B 286 (2021), https://doi.org/10.1016/j.
apcatb.2020.119816.
sample was stronger and was able to achieve 80% NO removal over a [9] Y. Yu, J. Miao, J. Wang, C. He, J. Chen, Facile synthesis of CuSO4/TiO2 catalysts
wider temperature window (220–420 ◦ C) and 97% conversion at 300 ◦ C. with superior activity and SO2 tolerance for NH3-SCR: physicochemical properties
In addition, the results of steady-state kinetic tests revealed that sulfide and reaction mechanism, Catal. Sci. Technol. 7 (7) (2017) 1590–1601, https://doi.
org/10.1039/c6cy02626e.
treatment slightly increased the activation energy (Ea) and shifted the [10] X. Qi, L. Han, J. Deng, T. Lan, F. Wang, L. Shi, D. Zhang, SO2-Tolerant catalytic
active temperature of the catalyst to a higher temperature, which reduction of NOX via tailoring electron transfer between surface iron sulfate and
significantly increased the surface acidity of the catalyst and slightly subsurface ceria, Environ. Sci. Technol. 56 (9) (2022) 5840–5848, https://doi.org/
10.1021/acs.est.2c00944.
suppressed the low-temperature activity. The (311) crystal surface
[11] H. Wang, Z. Qu, L. Liu, S. Dong, Y. Qiao, Promotion of NH3-SCR activity by sulfate-
adsorption of NH3 and NO on MnFe2O4 and SO2- 4 -MnFe2O4 was calcu­ modification over mesoporous Fe doped CeO2 catalyst: structure and mechanism,
lated by DFT, and the chemisorption of NO was found to be stronger. The J. Hazard Mater. 414 (2021), https://doi.org/10.1016/j.jhazmat.2021.125565.
adsorption energies of NH3 and NO on the MnFe2O4(311) crystal surface [12] Q. Yan, Y. Gao, Y. Li, M.A. Vasiliades, S. Chen, C. Zhang, R. Gui, Q. Wang, T. Zhu,
A.M. Efstathiou, Promotional effect of Ce doping in Cu4Al1OX – LDO catalyst for
and SO2-4 -MnFe2O4(311) crystal surface were calculated by DFT, and it low-T practical NH3-SCR: steady-state and transient kinetics studies, Appl. Catal., B
was found that the chemisorption of NO was stronger. Meanwhile, PDOS 255 (2019), https://doi.org/10.1016/j.apcatb.2019.117749.

10
J. Xu et al. Journal of the Energy Institute 112 (2024) 101458

[13] E. Muñoz, P. Marín, S. Ordóñez, F.V. Díez, The role of reaction kinetics and mass [34] Z. Liu, H. Su, B. Chen, J. Li, S.I. Woo, Activity enhancement of WO3 modified Fe2O3
transfer in the selective catalytic reduction of NO with NH3 in monolithic reactors, catalyst for the selective catalytic reduction of NO by NH3, Chem. Eng. J. 299
J. Chem. Technol. Biotechnol. 90 (7) (2015) 1299–1307, https://doi.org/10.1002/ (2016) 255–262, https://doi.org/10.1016/j.cej.2016.04.100.
jctb.4437. [35] S. Deng, T. Meng, B. Xu, F. Gao, Y. Ding, L. Yu, Y. Fan, Advanced MnOx/TiO2
[14] K.N. Rao, B.M. Reddy, S.-E. Park, Novel CeO2 promoted TiO2–ZrO2 nano-oxide catalyst with preferentially exposed anatase {001} facet for low-temperature SCR
catalysts for oxidative dehydrogenation of p-diethylbenzene utilizing CO2 as soft of NO, ACS Catal. 6 (9) (2016) 5807–5815, https://doi.org/10.1021/
oxidant, Appl. Catal., B 100 (3–4) (2010) 472–480, https://doi.org/10.1016/j. acscatal.6b01121.
apcatb.2010.08.024. [36] Y. Zeng, Y. Wang, S. Zhang, Q. Zhong, A study on the NH3-SCR performance and
[15] W. Liu, Y. Long, Y. Zhou, S. Liu, X. Tong, Y. Yin, X. Li, K. Hu, J. Hu, Excellent low reaction mechanism of a cost-effective and environment-friendly black TiO2
temperature NH3-SCR and NH3-SCO performance over Ag-Mn/Ce-Ti catalyst: catalyst, Phys. Chem. Chem. Phys. 20 (35) (2018) 22744–22752, https://doi.org/
evaluation and characterization, Mol. Catal. 528 (2022), https://doi.org/10.1016/ 10.1039/c8cp02270d.
j.mcat.2022.112510. [37] D. Fang, K. Qi, F. Li, F. He, J. Xie, Excellent sulfur tolerance performance over Fe-
[16] Z. Wang, M. Jiao, Z. Chen, H. He, L. Liu, Effects of montmorillonite and anatase SO4/TiO2 catalysts for NH3-SCR: influence of sulfation and Fe-based sulfates,
TiO2 support on CeO2 catalysts during NH3-SCR reaction, Microporous Mesoporous J. Environ. Chem. Eng. 10 (1) (2022), https://doi.org/10.1016/j.
Mater. 320 (2021), https://doi.org/10.1016/j.micromeso.2021.111072. jece.2021.107038.
[17] S.K. Perumal, N. Kaisare, S.K. Kummari, P. Aghalayam, Low-temperature NH3-SCR [38] H.-m. Wang, Y.-p. Ma, X.-y. Chen, S.-y. Xu, J.-d. Chen, Q.-l. Zhang, B. Zhao,
of NO over robust RuNi/Al-SBA-15 catalysts: effect of Ru loading, J. Environ. P. Ning, Promoting effect of SO2- 4 functionalization on the performance of Fe2O3
Chem. Eng. 10 (5) (2022), https://doi.org/10.1016/j.jece.2022.108288. catalyst in the selective catalytic reduction of NO with NH3, J. Fuel Chem. Technol.
[18] J. Zhu, J. Li, B. Chu, S. Liu, S. Fu, Q. Qin, L. Dong, B. Li, Excitation of catalytic 48 (5) (2020) 584–593, https://doi.org/10.1016/s1872-5813(20)30025-6.
performance on MOFs derivative carrier by residual carbon for low-temperature [39] C. Liu, Y. Bi, J. Li, Activity enhancement of sulphated Fe2O3 supported on
NH3-SCR reaction, Mol. Catal. 535 (2023), https://doi.org/10.1016/j. TiO2–ZrO2 for the selective catalytic reduction of NO by NH3, Appl. Surf. Sci. 528
mcat.2022.112859. (2020), https://doi.org/10.1016/j.apsusc.2020.146695.
[19] J. Shen, N. Gao, Y. Shan, M. Liu, J. Liu, Y. Xu, S. Shen, Y. Chen, Catalytic ozone [40] J. Zhou, X. Wang, X. He, J. Wang, K. Gui, H.R. Thomas, The effect of SO2 and Ca
oxidation toluene over supported manganese cobalt composite: influence of Co-pretreatment on the catalytic activity of Mn–Ce/TiO2 catalysts for selective
catalyst support, Environ. Sci. Pollut. Res. Int. 28 (45) (2021) 64778–64792, catalytic reduction of NO with NH3, Catal. Lett. 150 (11) (2020) 3287–3295,
https://doi.org/10.1007/s11356-021-15428-7. https://doi.org/10.1007/s10562-020-03229-5.
[20] J. Arfaoui, A. Ghorbel, C. Petitto, G. Delahay, Novel V2O5-CeO2-TiO2-SO2− 4 [41] C. Yu, B. Huang, L. Dong, F. Chen, Y. Yang, Y. Fan, Y. Yang, X. Liu, X. Wang, Effect
nanostructured aerogel catalyst for the low temperature selective catalytic of Pr/Ce addition on the catalytic performance and SO2 resistance of highly
reduction of NO by NH3 in excess O2, Appl. Catal., B 224 (2018) 264–275, https:// dispersed MnO x/SAPO-34 catalyst for NH3 -SCR at low temperature, Chem. Eng. J.
doi.org/10.1016/j.apcatb.2017.10.059. 316 (2017) 1059–1068, https://doi.org/10.1016/j.cej.2017.02.024.
[21] W. Zhao, Q. Zhong, T. Zhang, Y. Pan, Characterization study on the promoting [42] J. Yang, S. Ren, B. Su, Y. zhou, G. Hu, L. Jiang, J. Cao, W. Liu, L. Yao, M. Kong,
effect of F-doping V2O5/TiO2 SCR catalysts, RSC Adv. 2 (20) (2012), https://doi. J. Yang, Q. Liu, Insight into N2O formation over different crystal phases of MnO2
org/10.1039/c2ra20987j. during low-temperature NH3–SCR of NO, Catal. Lett. 151 (10) (2021) 2964–2971,
[22] W.J. Zhou, Y.H. Leng, D.M. Hou, H.D. Li, L.G. Li, G.Q. Li, H. Liu, S.W. Chen, Phase https://doi.org/10.1007/s10562-021-03541-8.
transformation and enhanced photocatalytic activity of S-doped Ag2O/TiO2 [43] S.S.R. Putluru, L. Schill, A.D. Jensen, B. Siret, F. Tabaries, R. Fehrmann, Mn/TiO2
heterostructured nanobelts, Nanoscale 6 (9) (2014) 4698–4704, https://doi.org/ and Mn–Fe/TiO2 catalysts synthesized by deposition precipitation—promising for
10.1039/c3nr06565k. selective catalytic reduction of NO with NH3 at low temperatures, Appl. Catal., B
[23] Y. Li, Z. Lian, J. Lin, M. Wang, W. Shan, TiO2-modified CeVO4 catalyst for the 165 (2015) 628–635, https://doi.org/10.1016/j.apcatb.2014.10.060.
selective catalytic reduction of NOx with NH3, Catal. Sci. Technol. 12 (15) (2022) [44] D.W. Kwon, K.B. Nam, S.C. Hong, Influence of tungsten on the activity of a Mn/Ce/
4884–4892, https://doi.org/10.1039/d2cy00848c. W/Ti catalyst for the selective catalytic reduction of NO with NH3 at low
[24] D.-j. Yan, T. Guo, Y. Yu, Z.-h. Chen, Lead poisoning and regeneration of Mn-Ce/ temperatures, Appl. Catal., A 497 (2015) 160–166, https://doi.org/10.1016/j.
TiO2 catalysts for NH3-SCR of NO at low temperature, J. Fuel Chem. Technol. 49 apcata.2015.01.013.
(1) (2021) 113–120, https://doi.org/10.1016/s1872-5813(21)60003-8. [45] L. Chen, S. Ren, L. Liu, B. Su, J. Yang, Z. Chen, M. Wang, Q. Liu, Catalytic
[25] T. Tang, J. Xu, X. Sheng, Y. Zhang, Q. Zhang, F. Guo, Radio-frequency (RF) plasma performance over Mn-Ce catalysts for NH3-SCR of NO at low temperature: different
synthesis of H-ZSM-5 supported MnOx catalysts with high catalytic activity for zeolite supports, J. Environ. Chem. Eng. 10 (2) (2022), https://doi.org/10.1016/j.
deNOx-SCR by propylene, Vacuum 195 (2022), https://doi.org/10.1016/j. jece.2022.107167.
vacuum.2021.110704. [46] B. Wang, M. Wang, L. Han, Y. Hou, W. Bao, C. Zhang, G. Feng, L. Chang, Z. Huang,
[26] Q. Li, X. Li, W. Li, L. Zhong, C. Zhang, Q. Fang, G. Chen, Effect of preferential J. Wang, Improved activity and SO2 resistance by Sm-modulated redox of
exposure of anatase TiO2 {0 0 1} facets on the performance of Mn-Ce/TiO2 MnCeSmTiOx mesoporous amorphous oxides for low-temperature NH3-SCR of NO,
catalysts for low-temperature selective catalytic reduction of NOx with NH3, Chem. ACS Catal. 10 (16) (2020) 9034–9045, https://doi.org/10.1021/acscatal.0c02567.
Eng. J. 369 (2019) 26–34, https://doi.org/10.1016/j.cej.2019.03.054. [47] L. Jiang, Q. Liu, G. Ran, M. Kong, S. Ren, J. Yang, J. Li, V2O5-modified Mn-Ce/AC
[27] Y. Wei, P. Liang, Y. Li, Y. Zhao, X. Min, P. Tao, J. Hu, T. Sun, Modification of Mn-Fe catalyst with high SO2 tolerance for low-temperature NH3-SCR of NO, Chem. Eng.
mixed oxide catalysts for low-temperature NH3-SCR of NO from marine diesel J. 370 (2019) 810–821, https://doi.org/10.1016/j.cej.2019.03.225.
exhausts, J. Environ. Chem. Eng. 10 (3) (2022), https://doi.org/10.1016/j. [48] F. Bertinchamps, C. Grégoire, E.M. Gaigneaux, Systematic investigation of
jece.2022.107772. supported transition metal oxide based formulations for the catalytic oxidative
[28] S. Zhan, M. Qiu, S. Yang, D. Zhu, H. Yu, Y. Li, Facile preparation of MnO2 doped elimination of (chloro)-aromatics, Appl. Catal., B 66 (1–2) (2006) 1–9, https://doi.
Fe2O3 hollow nanofibers for low temperature SCR of NO with NH3, J. Mater. Chem. org/10.1016/j.apcatb.2006.02.011.
A 2 (48) (2014) 20486–20493, https://doi.org/10.1039/c4ta04807e. [49] S. Zhang, C. Zhang, Q. Wang, W.-S. Ahn, Co- and Mn-coimpregnated ZSM-5
[29] J. Xu, T. Tang, X. Sheng, Y. Zhang, Q. Zhang, F. Guo, Excellent activity caused by prepared from recycled industrial solid wastes for low-temperature NH3-SCR, Ind.
dielectric barrier discharge (DBD) plasma activation for selective catalytic Eng. Chem. Res. 58 (51) (2019) 22857–22865, https://doi.org/10.1021/acs.
reduction with propylene (C3H6-SCR): insight into the low temperature catalytic iecr.9b04383.
behavior of Mn/ZSM-5 catalysts, J. Environ. Chem. Eng. 10 (2) (2022), https://doi. [50] C. Yu, B. Huang, L. Dong, F. Chen, X. Liu, In situ FT-IR study of highly dispersed
org/10.1016/j.jece.2021.107009. MnO x/SAPO-34 catalyst for low-temperature selective catalytic reduction of NO x
[30] J. Yang, S. Ren, B. Su, M. Wang, L. Chen, Q. Liu, Understanding the dual-acting of by NH3, Catal, Today Off. 281 (2017) 610–620, https://doi.org/10.1016/j.
iron and sulfur dioxide over Mn-Fe/AC catalysts for low-temperature, SCR of NO, cattod.2016.06.025.
Mol. Catal. 519 (2022), https://doi.org/10.1016/j.mcat.2022.112150. [51] X. Zhou, X. Huang, A. Xie, S. Luo, C. Yao, X. Li, S. Zuo, V2O5-decorated Mn-Fe/
[31] S. Liu, X. Chen, A visible light response TiO2 photocatalyst realized by cationic S- attapulgite catalyst with high SO2 tolerance for SCR of NOx with NH3 at low
doping and its application for phenol degradation, J. Hazard Mater. 152 (1) (2008) temperature, Chem. Eng. J. 326 (2017) 1074–1085, https://doi.org/10.1016/j.
48–55, https://doi.org/10.1016/j.jhazmat.2007.06.062. cej.2017.06.015.
[32] M. Sakti La Ore, K. Wijaya, W. Trisunaryanti, W.D. Saputri, E. Heraldy, N. [52] C. Song, W. Yu, B. Zhao, H. Zhang, C. Tang, K. Sun, X. Wu, L. Dong, Y. Chen,
W. Yuwana, P.L. Hariani, A. Budiman, S. Sudiono, The synthesis of SO4/ZrO2 and Efficient fabrication and photocatalytic properties of TiO2 hollow spheres, Catal.
Zr/CaO catalysts via hydrothermal treatment and their application for conversion Commun. 10 (5) (2009) 650–654, https://doi.org/10.1016/j.catcom.2008.11.010.
of low-grade coconut oil into biodiesel, J. Environ. Chem. Eng. 8 (5) (2020), [53] W. Zhao, Q. Zhong, Y. Pan, R. Zhang, Systematic effects of S-doping on the activity
https://doi.org/10.1016/j.jece.2020.104205. of V2O5/TiO2 catalyst for low-temperature NH3-SCR, Chem. Eng. J. 228 (2013)
[33] M.-y. Li, R.-t. Guo, C.-x. Hu, P. Sun, W.-g. Pan, S.-m. Liu, X. Sun, S.-w. Liu, J. Liu, 815–823, https://doi.org/10.1016/j.cej.2013.05.056.
The enhanced resistance to K deactivation of Ce/TiO2 catalyst for NH3-SCR [54] C.E. Lund Myhre, D.H. Christensen, F.M. Nicolaisen, C.J. Nielsen, Spectroscopic
reaction by the modification with P, Appl. Surf. Sci. 436 (2018) 814–822, https:// study of aqueous H2SO4 at different temperatures and compositions: variations in
doi.org/10.1016/j.apsusc.2017.12.087. dissociation and optical properties, J. Phys. Chem. A 107 (12) (2003) 1979–1991,
https://doi.org/10.1021/jp026576n.

11

You might also like