You are on page 1of 12

Chinese Journal of Catalysis 42 (2021) 2242–2253

available at www.sciencedirect.com

journal homepage: www.sciencedirect.com/journal/chinese‐journal‐of‐catalysis

Article
Effect of the modification of alumina supports with chloride on the
structure and catalytic performance of Ag/Al2O3 catalysts for the
selective catalytic reduction of NOx with propene and H2/propene
Jia Wang a, Rui You a,*, Kun Qian a, Yang Pan b, Jiuzhong Yang b, Weixin Huang a,c,#
a Hefei National Laboratory for Physical Sciences at the Microscale Key Laboratory of Surface and Interface Chemistry and Energy Catalysis of Anhui
Higher Education Institutes and Department of Chemical Physics, University of Science and Technology of China, Hefei 230026, Anhui, China
b National Synchrotron Radiation Laboratory, University of Science and Technology of China, Hefei 230029, Anhui, China

c Dalian National Laboratory for Clean Energy, Dalian 116023, Liaoning, China

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

Article history: The effect of the modification of an alumina support with chloride on the structure and the catalytic
Received 17 December 2020 performance of Ag/Al2O3 catalysts (SA) was investigated for the selective catalytic reduction (SCR)
Accepted 7 February 2021 of NO using C3H6 or H2/C3H6 as reductants. The Ag/Al2O3 catalyst and Cl–‐modified Ag/Al2O3 cata‐
Available online 10 September 2021 lysts (SA‐Cl) were prepared by a conventional impregnation method and characterized by X‐ray
diffraction, Brunauer‐Emmett‐Teller isotherm analysis, electron probe microanalysis, transmission
Keywords: electron microscopy, UV‐Vis diffuse reflectance spectroscopy, X‐ray photoelectron spectroscopy,
In situ characterization and hydrogen temperature‐programmed reduction. The catalytic activities in the C3H6‐SCR and
Reaction mechanism H2/C3H6‐SCR reactions were evaluated, and the reaction mechanism was studied using in situ dif‐
Structure‐activity relation fuse reflectance infrared Fourier transform spectroscopy and synchrotron vacuum ultraviolet pho‐
DRIFTS toionization mass spectroscopy (SVUV‐PIMS). We found that Cl modification of the alumi‐
Photoionization mass spectroscopy na‐supported Ag/Al2O3 catalysts facilitated the formation of oxidized silver species (Agnᵟ+) that cat‐
alyze the moderate‐temperature oxidation of hydrocarbons into partial oxidation products (mainly
acetate species) capable of participating in the SCR reaction. The low‐temperature promoting effect
of H2 on the C3H6‐SCR ("hydrogen effect") was found to originate from the enhanced decomposition
of strongly adsorbed nitrates on the catalyst surface and the conversion of these adsorbed species
to –NCO and –CN species. This "H2 effect" occurs in the presence of Agnᵟ+ species rather than the
metallic Ag0 species. A gaseous intermediate, acrylonitrile (CH2CHCN), was also identified in the
H2/C3H6‐SCR reaction using SVUV‐PIMS. These findings provide novel insights in the struc‐
ture‐activity relationship and reaction mechanisms of the SA‐catalyzed HC‐SCR reaction of NO.
© 2021, Dalian Institute of Chemical Physics, Chinese Academy of Sciences.
Published by Elsevier B.V. All rights reserved.

1. Introduction vehicles and industry is involved in the formation of acid rain


and photochemical smog and is a serious environmental prob‐
Air pollution by nitrogen oxides (NOx) emitted from both lem [1,2]. The use of diesel and lean‐burn gasoline engines can

* Corresponding author. Tel: +86‐551‐63600435; Fax: +86‐551‐63600437; E‐mail: yourui@ustc.edu.cn


# Corresponding author. E‐mail: huangwx@ustc.edu.cn

This work was supported by the National Natural Science Foundation of China (21703227), the University of Science and Technology of China
(KY2060000176), the Chinese academy of Sciences, and the Changjiang Scholars Program of Ministry of Education of China.
DOI: 10.1016/S1872‐2067(21)63904‐9 | http://www.sciencedirect.com/journal/chinese‐journal‐of‐catalysis | Chin. J. Catal., Vol. 42, No. 12, December 2021
Jia Wang et al. / Chinese Journal of Catalysis 42 (2021) 2242–2253 2243

improve fuel economy and lower CO2 emissions [3], but the and/or Agnδ+) are proposed to be the active species for NOx
resulting oxygen‐rich exhausts decrease the efficiency of con‐ reduction by HC‐SCR, whereas the Agn0 metallic clusters are
ventional three‐way catalysts in catalyzing the reduction of NOx responsible for the nonselective oxidation of hydrocarbons.
[4]. Therefore, additional catalytic technologies have been de‐ Previous research has found that the composition of the sup‐
veloped for NOx removal from diesel exhaust. The selective ported silver species is strongly linked to the silver loading on
catalytic reduction (SCR) of NOx with various reductants has the support. When the Ag loading is low, Ag mainly exists in the
been extensively studied as a highly promising NOx abatement form of Ag+ or Agnδ+, whereas catalysts having a high Ag load
technology [5–8]. However, additional reductants are usually usually contain more Agn0 nanoparticles. Therefore, numerous
needed because the unburned hydrocarbons (HCs) and CO in studies have attempted to establish the structure‐activity rela‐
the exhaust systems of diesel and lean‐burn gasoline engines tionship between Ag species and catalyst activity in the HC‐SCR
are not sufficient to reduce NOx alone. reaction by simply adjusting the Ag loadings and, thus, the re‐
V2O5/TiO2 catalysts promoted with WO3 or MoO3 are typical sulting Ag species [45–47].
and efficient commercial catalysts for the SCR of NOx in the In this work, we report that the modification of an Ag/Al2O3
presence of NH3 as a reductant (NH3‐SCR) [9]. However, catalyst with chloride facilitates the formation of oxidized silver
NH3‐SCR technologies rely on the presence of a urea distribu‐ species and promotes catalytic activity for the C3H6‐SCR reac‐
tion infrastructure, urea tank fill compliance, and delivery sys‐ tion. Further, based on in situ diffuse reflectance infrared Fou‐
tems for liquid urea solutions [10,11]. Therefore, the use of rier transform spectroscopy (DRIFTS) and synchrotron vacuum
reductants other than ammonia to eliminate NOx under oxidiz‐ ultraviolet photoionization mass spectroscopy (SVUV‐PIMS),
ing conditions using the SCR has been investigated. In particu‐ we reveal the reaction mechanism of the C3H6‐SCR reaction and
lar, hydrocarbon‐based SCR (HC‐SCR) has attracted much at‐ the mechanism by which H2 enhances the low‐temperature
tention over the past two decades, and numerous catalysts, activity of Ag/Al2O3 catalysts in the HC‐SCR reaction. We also
including zeolite‐based oxides, platinum group metals (PGMs), found the "H2 effect" occurs for the Agnδ+ species rather than
and base metal/oxide catalysts, have been investigated the metallic Ag0 species.
[12–16]. Held et al. [17] and Iwamoto et al. [18] first reported
that Cu‐ion‐exchanged ZSM‐5 was capable of catalyzing NOx 2. Experimental
reduction to N2 with low concentrations of hydrocarbons in the
engine exhaust streams in an oxidizing atmosphere. Later, zeo‐ 2.1. Catalyst preparation
lite‐based catalysts were found to be particularly effective for
the SCR of NO with methane (e.g., Co or Ga/ZSM‐5) or propene Alumina‐supported silver catalysts (denoted SA) having
(e.g., Cu/ZSM‐5). However, the hydrothermal resistance of different Ag loadings (0 wt%, 2 wt%, and 4 wt%) were pre‐
most zeolite‐based materials is usually unsatisfactory [19], and pared by impregnating γ‐Al2O3 with aqueous silver nitrate so‐
practical NOx reduction catalysts must remain active and stable lution, followed by drying at 110 °C and subsequent calcination
at the high temperatures typically encountered, e.g., 673 K, as at 600 °C in air for 3 h. The catalysts are denoted xSA, where x
well as in the presence of SO2 and water vapor. Thus, PGM cat‐ is the weight percent of Ag. The Cl‐modified γ‐Al2O3 support
alysts were developed by Hamada et al. [20] and Obuchi et al. was prepared by impregnating γ‐Al2O3 with aqueous ammo‐
[21] for the low‐temperature (300 °C) catalysis of NOx reduc‐ nium chloride (Al:Cl molar ratio of 1:1), followed by drying at
tion by HCs, but these catalysts are inefficient at moderate and 110 °C and subsequent calcination at 600 °C in air for 3 h. The
high temperatures. In addition, when using PGM catalysts, NOx Ag catalysts supported on Cl‐modified γ‐Al2O3 were prepared
species are often partially reduced to N2O, a very potent similarly to SA and are denoted xSA‐Cl, where x is the weight
greenhouse gas [22]. percent of Ag.
Being relatively durable and inexpensive, supported silver
catalysts are considered promising practical HC‐SCR catalysts 2.2. Structural characterization
for NOx reduction. Ag/Al2O3 has been reported to show rela‐
tively high activity and selectivity for NOx reduction to N2, as The specific surface areas and pore size distributions were
well as moderate resistance to H2O and SO2 [23–26]. However, measured at –196 °C using the Brunauer‐Emmett‐Teller (BET)
the light‐off temperature of Ag/Al2O3 catalysts is strongly de‐ method on a Micromeritics ASAP2000 instrument with an au‐
pendent on the nature of the reducing agent [27]. For instance, tomated gas sorption system. Prior to the measurements, the
low‐temperature activity can be achieved by using oxygenated catalysts were degassed at 300 °C for 3 h. Powder X‐ray diffrac‐
hydrocarbons, particularly ethanol [28–30], and higher hydro‐ tion (XRD) patterns were recorded on a Philips X’Pert Pro Su‐
carbons as reductants [31,32]. In addition, the addition of H2 to per diffractometer with Cu Kα radiation ( = 0.15406 nm) op‐
the reaction feed remarkably improves the NOx reduction ac‐ erating at 40 kV and 50 mA. Electron probe X‐ray microanalysis
tivity at low temperatures, for example, yielding a light‐off (EPMA) was carried out using a JEOL JXA8530 F system oper‐
temperature reduction of 215 °C [33–39]. ating at 15 kV and 50 nA; pure elements were used as stand‐
Different Ag species, such as isolated Ag+ cations, oxidized ards. The compositions of each equilibrium phase were meas‐
silver clusters (Agnδ+), and metallic silver clusters (Agn0), have ured at three different positions, and the average value was
been observed in both as‐prepared Ag/Al2O3 catalysts and used to establish the isothermal section. Transmission electron
HC‐SCR catalysts [40–44], and the oxidized silver species (Ag+ microscopy (TEM) and high‐resolution transmission electron
2244 Jia Wang et al. / Chinese Journal of Catalysis 42 (2021) 2242–2253

microscopy (HRTEM) images were acquired with a JEOL ature controller. The catalyst was sieved into particles with
JEM‐2100F instrument at an acceleration voltage of 120 kV. sizes of 0.250–0.375 mm diameter. Typically, 50 mg catalyst
Hydrogen temperature programmed reduction (H2‐TPR) was pretreated at 500 °C for 1 h in a flow of 10% O2/Ar (50
was performed in a flow of 5% H2/Ar (30 mL/min, 50–700 °C) mL/min) and then cooled to 200 °C. Subsequently, the reaction
on Micromeritics AutoChem II 2920. Before measurement, gas was switched to that of C3H6‐SCR (1000 ppm NO, 1500 ppm
about 0.10 g catalyst was pretreated in Ar at 120 °C for 1 h and C3H6, and 10% O2 in Ar) or H2/C3H6‐SCR (1000 ppm H2, 1000
then cooled to 50 °C in Ar. Ultraviolet‐visible diffuse reflectance ppm NO, 1500 ppm C3H6, and 10% O2 in Ar) at a total flow rate
spectra (UV‐Vis DRS) of the SA and SA‐Cl catalysts before cal‐ of 50 mL/min. The temperature was increased stepwise from
cination at 500 °C for 1 h were collected at room temperature 200 to 600 °C in steps of 50 °C and kept for about 1 h at each
(RT) in ambient air from 200 to 800 nm on a DUV‐3700 deep temperature to achieve a steady state. The concentrations of
ultraviolet (DUV)‐Vis‐near infrared (NIR) spectrophotometer NO, NO2, NH3, N2O, CO, and CO2 were measured using a FTIR
equipped with a diffuse reflectance cell. Calcined γ‐Al2O3 or the NOx analyzer (MCA14m). The NOx and C3H6 conversions were
Cl–‐modified γ‐Al2O3 were used as reference samples. X‐ray calculated using the equations below and are shown in Fig. 1.
photoelectron spectra (XPS) were obtained using an ESCALAB NOx conversion (%) = ([NOx]in − [NOx]out)/[NOx]in × 100%
250 high‐performance electron spectrometer using a mono‐ C3H6 conversion (%) = ([CO]out + [CO2]out)/([C3H6]in × 3) ×
chromatized Al Kα excitation source (hν = 1486.6 eV). The 100%
binding energy of adventitious carbon (C 1s) was set to 284.8
eV and used to correct for the likely charging of samples. 3. Results and discussion
In situ DRIFTS spectra were recorded on a Nicolet 6700
Fourier transform infrared (FTIR) spectrometer equipped with 3.1. C3H6‐SCR
an in situ high‐temperature DRIFTS reaction cell connected to a
conventional flow reaction system. About 50 mg catalyst was As shown in Fig. 1(a), the pure γ‐Al2O3 support showed only
loaded on the sample stage of the reaction cell. The samples moderate C3H6‐SCR activity, having a maximum NOx conversion
were first pretreated at 500 °C for 1 h in a flow of 10% O2/Ar of 63% at 450 °C, whereas the 2SA catalyst exhibited enhanced
(50 mL/min) and then cooled to the desired temperature in Ar catalytic activity, having a maximum NOx conversion of 93% at
to measure the background spectra. Then, various gas mixtures 400 °C. However, further increasing the Ag loading to 4 wt%
were fed into the reaction cell at a flow rate of 50 mL/min and (4SA) resulted in a reduction in the catalytic activity, and the
different temperatures, and the steady‐state DRIFTS spectra highest NOx conversion was only 61%. Meanwhile, the NOx
were recorded. All DRIFTS spectra were measured using 256 conversions of the SA catalysts in both the low‐temperature (<
scans at a resolution of 4 cm−1 using a Mercu‐ 350 °C) and high‐temperature (> 500 °C) regimes were quite
ry‐cadmium‐telluride/A (MCT/A) detector. poor. However, modification of the γ‐Al2O3 support with Cl–
SVUV‐PIMS [82–84] measurements were performed at the was found to result in an improvement in the catalytic perfor‐
combustion end station of beamline 03U at the National Syn‐ mance of SA catalysts for the C3H6‐SCR reaction. The NOx con‐
chrotron Radiation Laboratory (Hefei, China). A quartz catalytic versions of SA‐Cl catalysts were generally higher than those of
reactor was connected to the SVUV‐PIMS spectrometer, and the corresponding SA catalysts, and the enhancement was more
catalyst bed (2 × 2 mm) was placed 20–30 mm from the sam‐ obvious for the catalysts with higher Ag loadings. For example,
pling nozzle of the spectrometer. During the measurements, the the NOx conversion at 500 °C increased from 61% for the 4SA
reaction gas, which consisted of a mixture having a catalyst to 85% for the 4SA‐Cl catalyst. As shown in Fig. 1(b), all
NO:C3H6:O2:H2 ratio of 1 1.5:100:1 (or 0) with Ar balance, was catalysts showed similar C3H6 conversions except that the 4SA
fed into a mixing chamber at a total flow rate of 200 mL/min. catalyst exhibited slightly higher C3H6 conversions. Byproducts
Then, the gas mixture was pumped into the catalyst bed, re‐ other than N2, such as N2O and NH3, were not detected, demon‐
sulting in a total pressure of 1.5 Torr. After the catalytic reac‐ strating the very high N2 selectivity of the HC‐SCR reaction cat‐
tion reached the steady state at the desired temperature, the alyzed by Ag/Al2O3 catalysts. This is the most attractive ad‐
composition of the effluent gas was analyzed using an online vantage of the supported Ag catalysts for the HC‐SCR reaction.
homemade time‐of‐flight mass spectrometer (TOF‐MS) with a
mass resolution of 2000. The sample gas was photoionized Al2O3 -Al2O3
(a) 100 (b)100
with photons having energies of 10.6 or 11.8 eV. The m/z val‐ -Al2O3-Cl -Al2O3-Cl
ues of the TOF‐MS measurements were calibrated based on the
C3H6 conversion (%)

2SA
NOx conversion (%)

80 80 2SA
2SA-Cl 2SA-Cl
flight time, and the measurement time for each mass spectrum 60
4SA
60 4SA
4SA-Cl 4SA-Cl
was 180 s.
40 40

2.3. Catalytic activity tests 20 20

0 0
Steady‐state catalytic C3H6‐SCR and H2/C3H6‐SCR experi‐ 200 300 400 500 600 200 300 400 500 600
ments over SA and SA‐Cl catalysts were carried out in a o
Temperature ( C)
o
Temperature ( C)
fixed‐bed flow reactor system consisting of a quartz reactor Fig. 1. NOx (a) and C3H6 (b) conversion as a function of reaction tem‐
tube having an 8‐mm inner diameter and fitted with a temper‐ perature during the C3H6‐SCR of NO catalyzed by SA and SA‐Cl catalysts.
Jia Wang et al. / Chinese Journal of Catalysis 42 (2021) 2242–2253 2245

1.0
4SA-Cl Ag
+ 2SA
+ 2SA-Cl
Agn
4SA
0.8 4SA
4SA-Cl
2SA-Cl
Intensity (a.u.)

Absorbance (a.u.)
2SA 0.6
0
-Al2O3-Cl Agn
-Al2O3 0.4
 Ag3O4  Ag2O2
 Ag6O2  Ag2O
0.2

20 30 40 50 60 70 80

2  0.0
200 300 400 500 600 700 800
Fig. 2. XRD patterns of various catalysts after calcination at 600 °C for 3 Wavelength (nm)
h.
Fig. 4. UV‐Vis DRS spectra of the SA and SA‐Cl catalysts.
The NOx reduction activity of Ag/Al2O3 catalysts for the HC‐SCR
reaction is very sensitive to the silver species [40–44], and oxi‐ Ag2O2 (PDF#51‐0945), Ag3O4 (PDF#40‐1054), and Ag6O2
dized silver species (Ag+ or Agnδ+) are believed to be the active (PDF#74‐0878). Thus, the supported silver species in the SA
species catalyzing the reduction of NO by HC‐SCR, whereas and SA‐Cl catalysts experience different chemical environments
metallic Agn0 clusters are believed to be responsible for nonse‐ and form silver oxides with different structures. The diffraction
lective oxidation of hydrocarbons. Therefore, the modification peaks of silver oxides in the SA‐Cl catalysts are stronger and
of the γ‐Al2O3 support with chloride likely affects the Ag specia‐ narrower than those in the corresponding SA catalyst samples,
tion in the Ag/Al2O3 catalysts and, thus, the SCR activity. suggesting that silver oxides supported on Cl‐modified Al2O3
are larger and have better crystallinity than those on the un‐
3.2. Catalyst structure modified Al2O3.
Fig. 3 shows scanning tunneling electron microscopy
The BET surface areas of all catalysts were found to range (STEM), particle size distributions and HR‐TEM images of the
from 185 to 200 m2/g. In addition, the elemental compositions catalysts and the particle size distributions. On the basis of the
of the catalysts were analyzed by EPMA. The Ag contents (mass lattice distances and the angles between crystal planes of the
percentage) of all silver‐loaded catalysts were 2.12% (2SA), observed two‐dimensional lattice stripes in the HRTEM images,
3.939% (4SA), 2.005% (2SA‐Cl), and 4.087% (4SA‐Cl), and the the observed nanoparticles are silver oxides, including Ag6O2,
Cl loadings (mass percentage) were 0.125% (Al2O3‐Cl), 0.102% Ag2O2, and Ag3O4. In addition, the size distributions indicate
(2SA‐Cl), and 0.188% (4SA‐Cl). Fig. 2 shows XRD patterns of that the supported silver oxide particles are larger on the
the catalysts after calcination at 600 °C for 3 h. Reflections cor‐ Cl‐modified Al2O3 than those on the bare Al2O3. These obser‐
responding to the γ‐Al2O3 phase (PDF#46‐1131) were ob‐ vations are consistent with the above XRD results.
served in the XRD patterns of all catalysts. Additional diffrac‐ Next, the silver species in the SA and SA‐Cl catalysts were
tion peaks at 2 of 27.7°, 32.3°, and 37.2 were observed in the characterized by UV‐Vis DRS, H2‐TPR, and XPS measurements.
XRD patterns of the SA and SA‐Cl catalysts, and these can be As shown in Fig. 4, the UV‐Vis DRS spectra of both SA and SA‐Cl
attributed to a mixed phase of two or more silver oxides, main‐ catalysts show broad absorption bands below 500 nm; specifi‐
ly Ag3O4 and Ag2O2, based on a comparison with standard XRD cally, these bands were observed at 200–230, 250–260, and >
patterns for silver oxides, including Ag2O (PDF#41‐1104), 390 nm and arise from Ag+ ions, Agnδ+ clusters, and metallic

Fig. 3. STEM (left), particle size distributions (middle) and HRTEM images (right) of the 2SA, 4SA, 2SA‐Cl, and 4SA‐Cl catalysts. Scale bar in the STEM
images is 100 nm.
2246 Jia Wang et al. / Chinese Journal of Catalysis 42 (2021) 2242–2253

Agn0 clusters, respectively [48,49]. In comparison with the


bands in the UV‐Vis DRS spectra of the corresponding SA cata‐ -Al2O3

lysts, those in the UV‐Vis DRS spectra of the SA‐Cl catalysts are -Al2O3-Cl

narrow and indicate significantly enhanced Agnδ+ content at the 2SA


2SA-Cl
expense of Agn0. This demonstrates that more silver species 4SA

Intensity (a.u.)
exist as Agnδ+ in the SA‐Cl catalysts than in the corresponding 4SA-Cl
SA catalysts.
Fig. 5 shows the H2‐TPR profiles of the Al2O3 supports and
the SA and SA‐Cl catalysts. The Al2O3 supports show peaks cor‐
responding to H2 consumption above 500 °C; thus, the H2 con‐
sumption peaks at lower temperatures could be attributed to
the reduction of oxidized silver species. Various supported
100 200 300 400 500 600 700
silver oxide species have been reported to show different sta‐ o
Temperature ( C)
bilities and, thus, different reduction temperatures [50,51]. The
observed H2 consumption peaks at 50–100, 150–350, and > Fig. 5. H2‐TPR profiles of Al2O3 supports and SA and SA‐Cl catalysts.
380 °C can be attributed to the reduction of large silver oxide
particles, highly dispersed Agnδ+ clusters, and isolated Ag+ ions, silver oxide species to metallic silver, whereas the larger Ag+
respectively. The 2SA catalyst exhibited H2 consumption peaks ratio in the SA‐Cl catalysts than in the corresponding SA cata‐
corresponding to large silver oxide particles, highly dispersed lysts is consistent with the presence of more Agnδ+ clusters and
Agnδ+ clusters, and isolated Ag+ ions with similar intensities. more stable silver oxide particles. Therefore, the XRD,
However, as the silver loading was increased to 4%, the H2 STEM/HRTEM, UV‐Vis DRS, H2‐TPR, and XPS results all
consumption peaks corresponding to the highly dispersed demonstrate that the Cl– modification of the γ‐Al2O3 support
Agnδ+ clusters and isolated Ag+ ions did not change, whereas affects the speciation of silver on the γ‐Al2O3 support and facil‐
those corresponding to the large silver oxide particles in‐ itate the formation of oxidized sliver species, leading to larger
creased significantly. The H2 consumption peaks of the highly silver oxide particles and more Agnδ+ clusters, possibly because
dispersed Agnδ+ clusters dominated in the H2‐TPR profiles of the Cl– modification enhanced the affinity of Ag for O2.
SA‐Cl catalysts. In particular, as the silver loading increased to
4%, the H2 consumption peaks of highly dispersed Agnδ+ clus‐ 3.3. In situ DRIFTS studies
ters increased significantly, whereas those of large silver oxide
particles and isolated Ag+ ions did not change. The H2 con‐ The mechanism of the C3H6‐SCR catalyzed by SA and SA‐Cl
sumption peaks of highly dispersed Agnδ+ clusters in the SA‐Cl catalysts was explored using in situ DRIFTS measurements.
catalysts are much larger than those of corresponding SA cata‐ First, in situ DRIFTS spectra were obtained during the
lysts and do not shift. The H2 consumption peaks correspond‐ steady‐state NO + O2 reaction over the different catalysts at
ing to isolated Ag+ ions in the SA‐Cl catalysts were similar to
those of the SA catalysts. In contrast, the H2 consumption peaks Ag 3d5/2
(a) Ag 3d5/2 +
Ag : 61.6% (b) +
Ag : 33.5%
of large silver oxide particles in SA‐Cl catalysts are stronger 2
210
than that in 2SA but weaker than that in 4SA and shifted to 210
2
Intensity (a.u.)

Ag 3d3/2 Ag 3d3/2
Intensity (a.u.)

higher temperatures, likely because of the increased particle


size. These H2‐TPR results demonstrate that the SA‐Cl catalysts Ag
0

contain more Agnδ+ clusters than the corresponding SA cata‐


+
Ag
0
Ag
lysts but similar amounts of isolated Ag+ ions. Thus, Cl– modifi‐ Ag
+

cation of γ‐Al2O3 support facilitates the formation of oxidized


sliver species. 364 366 368 370 372 374 376 378 364 366 368 370 372 374 376 378
The surface oxidation states of silver species in the SA and Binding energy (eV) Binding energy (eV)
SA‐Cl catalysts were further characterized by XPS. As shown in (c) Ag 3d5/2 +
Ag : 64.6% (d) Ag 3d5/2 +
Ag : 58.7%
Fig. 6, the Ag 3d XPS spectra of all catalysts could be deconvo‐ 2
210
luted into two Ag 3d5/2 components having binding energies at Ag 3d3/2 Ag 3d3/2
Intensity (a.u.)

2
Intensity (a.u.)

210
367.5–367.8 and 368.3–368.6 eV, which can be assigned to the
Ag+ and Ag0 species, respectively. The ratios of Ag+ components
+
Ag
were 61.6%, 33.5%, 64.6%, and 58.7% in 2SA, 4SA, 2SA‐Cl, and Ag
+
Ag
0
Ag
0

4SA‐Cl, respectively. These results demonstrate that the oxi‐


dized silver species dominate in the 2SA and SA‐Cl catalysts,
whereas the metallic Ag0 particles dominate in the 4SA catalyst. 364 366 368 370 372 374 376 378 364 366 368 370 372 374 376 378
Binding energy (eV) Binding energy (eV)
As demonstrated in the H2‐TPR profiles, the silver oxide is
more stable in the SA‐Cl samples than that in the SA samples. Fig. 6. Ag 3d XPS spectra (dashed lines) of 2SA (a), 4SA (b), 2SA‐Cl (c),
Therefore, the large reduction in the surface Ag+ ratio from 2SA and 4SA‐Cl (d) catalysts and fitted Ag 3d5/2 peaks consisting of Ag+ and
to 4SA could also be attributed to the conversion of unstable Ag0 components. The ratio of Ag+ is indicated in each panel.
Jia Wang et al. / Chinese Journal of Catalysis 42 (2021) 2242–2253 2247

Table 1
(a) (b) Assignment of observed vibrational bands in the in situ DFIFTS spectra.
0.1 0.1 1302
1558 Wavenumber (cm–1) Surface species Ref.

Kubelka-Munk units
Kubelka-Munk units

1558
1303 1540 1582 1590 1303, 1540, 1558 Monodentate nitrates [51–54,56–58]
1614 1582–1590 Bidentate nitrates [51,52,57,59]
1614 1540 1614 Bridge‐bound nitrates [53,58,59]
1298–1300, 1526–1535 Carbonates [60,61]
1376, 1392, 1588–1592,
Formate [60,61]
2906, 3000
1454, 1580–1585, 2942 Acetate [5,6,61,77]
1200 1300 1400 1500 1600 1700 1200 1300 1400 1500 1600 1700
1 1 1458, 1575 Carboxylate [63,64]
wavenumber (cm ) wavenumber (cm ) 1454–1458, 1561–1575,
Acrylate [60]
(c) (d) 1303 1630–1643
0.11303 0.1 2152 Cyanide (CN–) [60,74]
Kubelka-Munk units

1560
Kubelka-Munk units

1560 2230 Isocyanate [15,16,60,74]


2340 Gas‐phase CO2 [79]
1590 1590
1540 1614 1540 1614
served vibrational bands in the spectra of the SA and SA‐Cl
catalysts were the same, and the overall band intensities did
not change significantly, although the relative intensities of the
1200 1300 1400 1500 1600 1700 1200 1300 1400 1500 1600 1700
bands corresponding to nitrate species varied slightly. There‐
wavenumber (cm )
1
wavenumber (cm )
1
fore, we propose that the formation of nitrate species occurred
mainly on the Al2O3 support of the SA and SA‐Cl catalysts, alt‐
Fig. 7. In situ DRIFTS spectra of γ‐Al2O3 (a), 2SA (b), (c) 4SA, and (d)
4SA‐Cl catalysts at the steady‐state during the NO + O2 reaction (1000 hough the minor formation of nitrate species on the silver sur‐
ppm NO, 10% O2, Ar in balance, total flow: 50 mL/min) at 200 °C (black face cannot be ruled out. In addition, the coverage of nitrate
line), 300 °C (red line), and 400 °C (blue line). species on the catalyst surfaces decreased as the temperature
increased from 200 to 400 °C. On the γ‐Al2O3 support, the cov‐
erage of nitrate species followed the order monodentate > bi‐
temperatures from 200 to 400 °C (Fig. 7). Bands were observed
dentate > bridge‐bound. In addition, the increase in silver
at 1303, 1540, 1558, 1582–1590, and 1614 cm–1, and these can
loading seemed to decrease the stability of bidentate nitrate
be assigned to nitrate species on the catalyst surfaces (Table 1)
species. These observations suggest that Cl– modification does
[52–59]. In particular, vibrational bands corresponding to ni‐
not exert much influence on the activation of NO and subse‐
trate species appeared in the spectra of the γ‐Al2O3 support,
quent formation of nitrate species on the SA catalysts.
indicating the occurrence of NO + O2 reaction. Compared to the
Next, in situ DRIFTS spectra were acquired during the
bands observed in the spectra of the γ‐Al2O3 support, the ob‐
steady‐state C3H6 + O2 reaction over various catalysts at tem‐

(a) 0.1 (b) 3000 (c) (d)


0.002 2942 0.1 0.002 3000
Kubelka-Munk units

Kubelka-Munk units
Kubelka-Munk units

Kubelka-Munk units

2906 2906 2942

1588
1458 1580
1574 1592
1561
1392 1535
1459 1640
13001376

1200 1300 1400 1500 1600 1700 2850 2900 2950 3000 3050 1200 1300 1400 1500 1600 1700 2850 2900 2950 3000 3050
1 1 1
wavenumber (cm ) wavenumber (cm )
1
wavenumber (cm ) wavenumber (cm )

(e) 1458 (f) (g) (h)


0.1 1575 0.002 0.1 0.002
Kubelka-Munk units

Kubelka-Munk units
Kubelka-Munk units

Kubelka-Munk units

2942 3000
2906

1454 1568
1585
1298 1630
1526 1374 1643

1200 1300 1400 1500 1600 1700 2850 2900 2950 3000 3050 1200 1300 1400 1500 1600 1700 2850 2900 2950 3000 3050
1 1 1 1
wavenumber (cm ) wavenumber (cm ) wavenumber (cm ) wavenumber (cm )
Fig. 8. In situ DRIFTS spectra of γ‐Al2O3 (a,b), 2SA (c,d), 4SA (e,f), and 4SA‐Cl (g,h) catalysts obtained during the steady‐state C3H6 + O2 reaction (1500
ppm C3H6, 10% O2, Ar in balance, total flow: 50 mL/min) at 200 °C (black line), 300 °C (red line), and 400 °C (blue line).
2248 Jia Wang et al. / Chinese Journal of Catalysis 42 (2021) 2242–2253

peratures from 200 to 400 °C (Fig. 8). Various oxygenate in‐


termediates were also observed to form (Table 1) (a) 4Ag/Al
1458
(b) 4Ag/Al
- -4
[5,6,60,61,63,64,77], including carbonates, formates, acetates, 1578 - 5E CO2
0.1

Kubelka-Munk units

Kubelka-Munk units
2942
carboxylates, and acrylates. Unlike the nitrate species formed 1640
by NO + O2 reaction, whose coverage decreased with increase 0min 0min
1min
in temperature from 200 to 400 °C, the overall coverage of ox‐ 3min 1min
3min
5min
5min
ygenate intermediates increased, suggesting that the oxidation 10min
20min 10min
30min 20min
of C3H6 is favored at high temperatures. The dominant oxygen‐ -
40min 30min
40min
ad-NO3
ate intermediates vary with both the reaction temperature and
-

1200 1300 1400 1500 1600 1700 2200 2400 2800 3000
the catalyst. On γ‐Al2O3, formate species were dominant at 300 1
Wavenumber (cm )
1
Wavenumber (cm )
and 400 °C, whereas acetate species also appeared at 400 °C. In
(c) 4Ag/Al-Cl (d) 4Ag/Al-Cl
the spectrum of the 2SA catalyst, formate, acetate, and car‐ -4 2942
5E
bonates species were found to be dominant, and the coverage 0.1

Kubelka-Munk units
Kubelka-Munk units
1460 1575 - -
CO2
of formate species decreased with increase in temperature -NCO
0min
from 300 to 400 °C, whereas the coverage of acetate species 0min
1min 1min
increased. Further, minor acrylate species appeared at 400 °C. 3min
5min 3min
5min
10min 10min
In the spectrum of the 4SA catalyst, carboxylate species were 20min
30min
20min
30min
found to be dominant, whereas acrylate and carbonates species ad-NO3
-
40min 40min
-

were also formed. On the basis of the spectral data, the varia‐ 1200 1300 1400 1500 1600 1700 2200 2400 2800 3000
tion in the surface intermediates on the 4SA‐Cl catalyst was Wavenumber (cm )
1 1
Wavenumber (cm )
similar to that on the 2SA catalyst. Both formate and acetate Fig. 9. Time‐resolved in situ DRIFTS spectra of (a,b) 4SA and (c,d)
species were found to be the dominant species, although minor 4SA‐Cl catalysts with surface intermediates formed under steady‐state
acrylate species were also observed at 400 °C. In addition, the C3H6 + O2 reaction exposed to NO + O2 flow at 500 °C.
coverage of formate species decreased with increase in tem‐
perature from 300 to 400 °C, whereas the coverage of acetate reaction. These results directly prove the acetate species
species increased. Further, the acetate/formate ratio at 400 °C formed by partial C3H6 oxidation on the silver surface as the
is larger for the 4SA‐Cl catalyst than for the 2SA catalyst. active oxygenate species for the HC‐SCR reaction. Therefore,
For the HC‐SCR reaction mechanism, it is proposed that NO more of the acetate intermediate formed on the 2SA and 4SA‐Cl
and hydrocarbons are activated into nitrate and oxygenated catalysts than on the γ‐Al2O3 support and 4SA catalyst during
species, respectively, on the catalyst surfaces, and the acetates the C3H6‐SCR reaction, and the 2SA and 4SA‐Cl catalysts exhib‐
are the active oxygenated species that react with nitrates, thus ited higher catalytic activity. Thus, by correlating the catalyst
forming the key –NCO intermediates capable of reacting with structures and surface intermediates of the SA and SA‐Cl cata‐
gaseous NO (or NO + O2) to generate N2 + CO2 (or CO) [66–74]. lysts, we can conclude that the oxidized sliver species facilitate
Thus, we next explored the reactivity of the oxygenate species the partial oxidation of C3H6 into the active acetate species.
formed on 4SA and 4SA‐Cl catalysts during the C3H6 + O2 reac‐ Further, the increase in the silver loading from 2% to 4% in the
tion at 500 °C toward NO + O2 using time‐resolved in situ SA catalysts mainly results in the formation of metallic Ag0 par‐
DRIFTS measurements. As shown in Fig. 9, after exposure to ticles and unstable large silver oxide particles, which results in
C3H6 + O2 followed by Ar purging, the 4SA catalyst contained a loss in the catalytic performance in the C3H6‐SCR reaction. In
more carboxylate species and fewer acetate species than the contrast, the increase of silver loading from 2% to 4% in the
4SA‐Cl catalyst. A band corresponding to gaseous CO2 also ap‐ SA‐Cl catalysts results in the formation of stable oxidized sliver
peared as a result of the C3H6 combustion reaction. After the species and does not result in a loss in catalytic performance.
introduction of NO + O2, the spectrum of the 4SA catalyst
showed the loss of the acetate species after 1 min, and both 3.4. H2/C3H6‐SCR
carboxylate species and gaseous CO2 species disappeared after
10 min. In contrast, the nitrate species began to appear after 5 During the HC‐SCR reaction, Ag/Al2O3 catalysts usually ex‐
min; for the 4SA‐Cl catalyst, both acetate species and gaseous hibit high efficiency for NOx conversion at temperatures above
CO2 species disappeared after 1 min, whereas the carboxylate 350 °C, which is higher than the typical temperature of diesel
species disappeared after 30 min, and the –NCO species engine exhaust [5]. The addition of H2 to the reaction feed has
(2200–2270 cm–1) appeared after 1 min, although this band been observed to promote the HC‐SCR conversion of NO into N2
then weakened and disappeared after 5 min. In addition, the significantly at temperatures as low as 200 °C in the presence
band corresponding to nitrate species appeared at 20 min. of Ag/Al2O3 catalysts [39,68,76–80]. Therefore, we also studied
Therefore, the carboxylate and acetate species exhibit different the "hydrogen effect" on our SA and SA‐Cl catalysts. Fig. 10
reactivities toward NO + O2. The carboxylate species mainly compares the NO conversions over γ‐Al2O3, 2SA, 4SA, and
reacts with O2 in the NO + O2 atmosphere to produce CO2, 4SA‐Cl in the C3H6‐SCR and H2/C3H6‐SCR reactions. No "hydro‐
whereas the acetate species reacts with NO + O2 (or surface gen effect" was observed over γ‐Al2O3. For the 2SA catalyst, the
nitrate species formed by NO + O2) to generate the –NCO spe‐ addition of 1000 ppm H2 to the C3H6‐SCR reaction gas resulted
cies, which are the key surface intermediates for the HC‐SCR in an obvious enhancement in the NO conversion at tempera‐
Jia Wang et al. / Chinese Journal of Catalysis 42 (2021) 2242–2253 2249

Al (b)100 HC-SCR (a) HC-SCR (b) HC-SCR


(a) 100 HC-SCR 2Ag/Al
H2/HC-SCR 0.1 H2/HC-SCR H2/HC-SCR
H2/HC-SCR

Kubelka-Munk units
Kubelka-Munk units
NOx conversion (%)

-3
80 80 510

NOx conversion (%)


60 60

40 40
2152 2230
20 20

0 0 1200 1300 1400 1500 1600 1700 1800 2200 2400 2800 3000
200 300 400 500 600 200 300 400 500 600 1
wavenumber (cm )
1
o o
wavenumber (cm )
Temperature ( C) Temperature ( C)
(c) HC-SCR (d) HC-SCR
4Ag/Al HC-SCR (d) 100 HC-SCR
(c) 100 H2/HC-SCR
4Ag/Al-Cl
H2/HC-SCR
H2/HC-SCR H2/HC-SCR
0.1

Kubelka-Munk units
Kubelka-Munk units
-3
80 80 510
NOx conversion (%)

NOx conversion (%)

60 60

40 40

20 20

0 0 1200 1300 1400 1500 1600 1700 1800 2200 2400 2800 3000
200 300 400 500 600 200 300 400 500 600 wavenumber (cm )
1
wavenumber (cm )
1

o o
Temperature ( C) Temperature ( C)
(e) HC-SCR (f) HC-SCR
Fig. 10. NOx conversion over γ‐Al2O3 (a), 2SA (b), 4SA (c), and 4SA‐Cl (d H2/HC-SCR H2/HC-SCR
0.1

Kubelka-Munk units
catalysts in the SCR process with and without H2. (1000 ppm NO, 1500 Kubelka-Munk units 510
-3

ppm C3H6, 0 or 1000 ppm H2, 10% O2, Ar in balance, total flow: 50
mL/min).

2152
tures lower than 400 °C, which is characteristic of the "hydro‐ 2230

gen effect." In contrast, this effect was significantly lower for 1200 1300 1400 1500 1600 1700 1800 2200 2400 2800 3000
the 4SA catalyst but was observed for the 4SA‐Cl catalyst. These 1
wavenumber (cm ) wavenumber (cm )
1

results suggest that the "hydrogen effect" only occurs on the Fig. 11. In situ DRIFTS spectra of 2SA (a,b), 4SA (c,d), and 4SA‐Cl (e,f)
silver surface and is closely related to the nature of the Ag spe‐ catalysts in the SCR process with and without H2 at 300 °C (1000 ppm
cies. The dominant silver species is oxidized Agnδ+ in the 2SA NO, 1500 ppm C3H6, 0 or 1000 ppm H2, 10% O2, Ar in balance, total flow
and 4SA‐Cl catalysts but metallic Ag0 particles in the 4SA cata‐ 50 mL/min).
lyst. Thus, oxidized Agnδ+ species exhibit an obvious "hydrogen
effect," whereas metallic Ag0 particles do not. This is likely be‐ spectively. However, the spectra of the 4SA catalyst did not
cause H2 dissociation is facile on the surface of metallic Ag0 show these features. These results indicate that the “hydrogen
particles, and the dissociated hydrogen subsequently under‐ effect” on our SA and SA‐Cl catalysts mainly results from the
goes oxidation with O2 to produce H2O. No catalysts exhibited increased decomposition of strongly adsorbed nitrate species
the "hydrogen effect" at temperatures higher than 400 °C, on the catalyst surface and the formation and conversion of
which is attributed to the direct reaction of H2 with O2 at these acetate species to the key –NCO and –CN species. Moreover, the
temperatures [39,68,76–80]. latter only occurred on oxidized Agnδ+ clusters, which further
The likely origin of the “hydrogen effect” on the HC‐SCR re‐ proves that oxidized Agnδ+ clusters are the active silver species
action over silver catalysts is proposed as follows in the Ag/Al2O3 catalysts for the HC‐SCR reaction.
[39,68,76–80]: (1) H2 reduces the isolated Ag+ ions and in‐
creases the amount of Agnδ+ clusters; (2) H2 promotes the de‐ 3.5. SVUV‐PIMS measurements
composition of strongly adsorbed nitrate species on the cata‐
lyst surface, thus exposing active sites available for the activa‐ In the HC‐SCR reactions, –CN and –NCO species are consid‐
tion of hydrocarbons; and (3) H2 promotes the transformation ered to be key intermediates on the catalyst surface and are
of surface intermediates to isocyanate and cyanide species, proposed to react with gaseous NO + O2 directly to generate N2.
which subsequently react with gaseous NO to form N2. Fig. 11 This suggests the involvement of gas‐phase reactions during
compares the in situ DRIFTS spectra of various catalysts in the the HC‐SCR reactions, but there is a lack of direct experimental
C3H6‐SCR reaction at 300 °C in the presence and absence of H2. evidence for this. SVUV‐PIMS is a sensitive technique to detect
The coverage of nitrate species on all catalyst surfaces de‐ reactive gas‐phase intermediates during catalytic reactions,
creased after the introduction of H2. For the 2SA and 4SA‐Cl having the unique ability to identify isomers and radicals be‐
catalysts, H2 addition also enhanced the formation of acetate cause of its high energy resolution and minimal fragmentation
species; in addition, the vibrational bands corresponding to the interference [82–85]. Therefore, we used online SVUV‐PIMS to
–NCO and –CN species appeared at 2232 and 2148 cm–1, re‐ analyze the gas‐phase components of C3H6‐SCR reactions cata‐
2250 Jia Wang et al. / Chinese Journal of Catalysis 42 (2021) 2242–2253

mation in the HC‐SCR reaction.


h=11.8 eV 42 h=11.8 eV
(a) 300 (b) 300 4. Conclusions
30 42
Ion intensity (a.u.)

Ion intensity (a.u.)


30
200 200
58
58
In summary, we have successfully demonstrated that the Cl–
modification of the γ‐Al2O3 support facilitates the formation of
100 100
17
oxidized silver species in Ag/Al2O3 catalysts active in catalyzing
5 10 25 5
5 10 25 553
56 the C3H6‐SCR reaction, particularly at high silver loadings. The
0 0
oxidized Agnδ+ species activate C3H6 to surface acetate species,
10 20 30 40 50 60 70 10 20 30 40 50 60 70
m/z m/z which can react with NO + O2 to generate the –NCO species that
are the key surface intermediates for the HC‐SCR reaction. In
42
(c) 300 h=11.8 eV (c)
(d) 2000 h=10.6eV 58 contrast, the metallic Ag0 species activate C3H6 to surface car‐
boxylate species by reaction with O2 in NO + O2 to produce CO2.
Ion intensity (a.u.)

The enhancing effect of H2 on the catalytic performance of the


Ion intensity (a.u.)

200 42
30
1000 30 Ag/Al2O3 catalysts at low reaction temperatures only occurs for
58
100
53
17 56 oxidized Agnδ+ species but not for metallic Ag0 particles, result‐
17
5 56 50 10 25 50 50 ing from the enhanced decomposition of strongly adsorbed
5 10 25 0
0 nitrate on the surface and conversion of adsorbed species to
10 20 30 40 50 60 70 10 20 30 40 50 60 70
key –NCO and –CN intermediates. A gaseous intermediate, ac‐
m/z m/z
rylonitrile, was observed, and this provides direct experimental
Fig. 12. SVUV‐PIMS spectra of the gas‐phase components in the (a) evidence for the occurrence of reactions involving C–N bond
C3H6‐SCRS and (b) H2/C3H6‐SCR reactions at 300 °C catalyzed by 4SA
ionized at a photon energy of 11.8 eV and in the H2/C3H6‐SCR reactions
formation in the HC‐SCR reaction. These results provide novel
at 300 °C catalyzed by 4SA‐Cl ionized at photon energies of (c) 11.8 and insights into the structure‐activity relationship and reaction
(d) 10.6 eV. mechanisms of the Ag/Al2O3‐catalyzed HC‐SCR reaction of NO.

References
lyzed by 4SA and 4SA‐Cl catalysts at 300 °C. Fig. 12 shows the
SVUV‐PIMS spectra of the catalytic reaction. To obtain these [1] L. Chen, J. H. Li, M. F. Ge, Environ. Sci. Technol., 2010, 24,
spectra, the samples were ionized at photon energies of 11.8 9590−9596.
and 10.6 eV. The SVUV‐PIMS spectrum of the C3H6‐SCR reaction [2] K. James, M. Jisue, Z. L. Wu, Chin. J. Catal., 2020, 6, 901–914.
catalyzed by 4SA and ionized with photons at 11.8 eV contains [3] M. V. Twigg, Appl. Catal. B, 2007, 70, 2–15.
major signals at m/z = 30, 42, and 58 (Fig. 12(a)) and weak [4] J. E. Parks, V. Prikhodko, W. P. Partridge, SAE, 2010, 01, 2267.
[5] R. Burch, J. P. Breen, F. C. Meunier, Appl. Catal. B, 2002, 39,
signals at m/z = 17, 53, and 56 (Fig. 12(b)). In the blank ex‐
283−303.
periments without the catalysts, no signals other than those
[6] H. He, Y. B. Yu, Catal. Today, 2005, 100, 37−47.
corresponding to the reactants were observed. In the [7] Z. M. Liu, S. I. Woo, Catal. Rev. Sci. Eng., 2006, 48, 43−89.
SVUV‐PIMS spectrum of the 4SA‐Cl‐catalyzed H2/C3H6‐SCR [8] P. Granger, V. I. Parvulescu, Chem. Rev., 2011, 111, 3155−3207.
reaction (Fig. 12(c)), the signal at m/z = 53 increased compared [9] P. Forzatti, I. Nova, E. Tronconi, Angew. Chem. Int. Ed., 2009, 48,
to that in the spectrum of the reaction catalyzed by 4SA. In the 8366–8368.
SVUV‐PIMS spectrum of the H2/C3H6‐SCR reaction catalyzed by [10] E. Seker, E. Gulari, R. Hammerle, C. Lambert, J. Leerat, S. Osuwan,
4SA‐Cl and ionized with photons at 10.6 eV (Fig. 12(d)), signals Appl. Catal. A, 2002, 226, 183−192.
at m/z = 17, 30, 42, 56, and 58 were observed but the signal at [11] S. J. Schmieg, B. K. Cho, S. H. Oh, Appl. Catal. B, 2004, 49,113−125.
m/z = 53 was not. The signals at m/z = 17, 30, 42, 56, and 58 [12] V. Zuzaniuk, F. C. Meunier, J. R. H. Ross, J. Catal., 2001, 202,
observed in the SVUV‐PIMS spectra of samples ionized at both 340–353.
[13] S. Roy, M. S. Hegde, G. Madras, Appl. Energy, 2009, 86, 2283−2297.
11.8 and 10.6 eV can be assigned to NH3 (ionization energy of
[14] F. Poignant, J. L. Freysz, M. Daturi, J. Saussey, Catal. Today, 2001,
10.2 eV), NO (ionization energy of 9.23 eV), C3H6 (ionization
70, 197−211.
energy of 9.74 eV), acrylic aldehyde (CH2=CHCHO, ionization [15] K. Sato, T. Yoshinari, Y. Kintaichi, M. Haneda, H. Hamade, Appl.
energy of 10.11 eV), and allyl alcohol (CH2=CHCH2OH, ioniza‐ Catal. B, 2003, 44, 67−78.
tion energy of 10.22 eV), respectively. The signal at m/z = 53 [16] F. C. Meunier, J. P. Breen, V. Zuzaniuk, M. Olsson, J. R. H. Ross, J.
observed in the SVUV‐PIMS spectrum of the sample ionized at Catal., 1999, 187, 493−505.
11.8 eV but not in that of the sample ionized at 10.6 eV can be [17] W. Held, A. König, T. Richter, L. Puppe, SAE, 1990, 9004961.
assigned to acrylonitrile (CH2CHCN, m/z = 53.0626, ionization [18] M. Iwamoto, H. Yahiro, S. Shundo, Y. Yoshihiro, N. Mizuno, Appl.
energy of 11.1 eV) rather than other species such as C4H5 (m/z Catal., 1991, 70, L15−L20.
= 53.0825, ionization energy of 7.95 eV). It has been reported [19] S. Matsumoto, K. Yokota, H. Poi, M. Kimura, K. Sekizawa, S.
that −NCO species can react with H2O to yield NH3 and CO2 Kasahara, Catal. Today, 1994, 22, 127−146.
[20] H. Hamada, Y. Kiutaichi, M. Sasaki, T. Ito, M. Tabata, Appl. Catal.,
during HC‐SCR [86,87], and the observation of gaseous
1991, 75, L1−L8.
CH2CHCN intermediate provides direct experimental evidence
[21] A. Obuchi, A. Ohi, M. Nakamura, A. Ogata, K. Mizuno, H. Ohuchi,
for the occurrence of reactions involving the C–N bond for‐ Appl. Catal. B, 1993, 2, 71−80.
Jia Wang et al. / Chinese Journal of Catalysis 42 (2021) 2242–2253 2251

Graphical Abstract
Chin. J. Catal., 2021, 42: 2242–2253 doi: 10.1016/S1872‐2067(21)63904‐9
Effect of the modification of alumina supports with chloride on the structure and catalytic performance of Ag/Al2O3 catalysts for
the selective catalytic reduction of NOx with propene and H2/propene
Jia Wang, Rui You *, Kun Qian, Yang Pan, Jiuzhong Yang, Weixin Huang *
University of Science and Technology of China; Dalian National Laboratory for Clean Energy

Chloride modification of alumina‐supported Ag/Al2O3 catalysts stabilizes Agnδ+ clusters that are beneficial for the selective catalytic re‐
duction of NOx with propene and hydrogen‐assisted selective catalytic reduction NO with propene.

[22] R. Burch, T. C. Watling, Appl. Catal. B, 1997, 11, 207−216. 109, 4805−4807.
[23] F. C. Meunier, J. R. H. Ross, Appl. Catal. B, 2000, 24, 23−32. [43] K. I. Shimizu, M. Tsuzuki, K. Kato, S. Yokota, K. Okumura, A.
[24] K. A. Bethke, H. H. Kung, J. Catal., 1997, 172, 93−102. Satsuma, J. Phys. Chem. C, 2007, 111, 950−959.
[25] H. He, R. D. Zhang, Y. B. Yu, J. Liu, Chin. J. Catal., 2003, 24, [44] S. T. Korhonen, A. M. Beale, M. A. Newton, B. M. Weckhuysen, J.
788−794. Phys. Chem. C, 2011, 115, 885−896.
[26] C. E. Stere, W. Adress, R. Burch, S. Chansai, A. Goguet, W. G. Gra‐ [45] T. Furusawa, K. Seshan, J. A. Lercher, L. Lefferts, K. Aika, Appl.
ham, F. D. Rosa, V. Palma, C. Hardacre, ACS Catal., 2014, 4, Catal. B, 2002, 37, 205−216.
666−673. [46] M. Richter, U. Bentrup, R. Eckelt, M. Schneider, M.‐M. Pohl, R.
[27] L.‐E. Lindfors, K. Eränen, F. Klingsledt, D. Yu. Murzin, Top. Catal., Fricke, Appl. Catal. B, 2004, 51, 261−274.
2004, 28, 185−189. [47] N. Bogdanchikova, F. C. Meunier, M. Avalos‐Borja, J. P Breen, A.
[28] M. K. Kim, P. S. Kim, J. H. Baik, I. S. Nam, B. K. Cho, S. H. Oh, Appl. Pestryakov, Appl. Catal. B, 2002, 36, 287−297.
Catal. B, 2011, 105, 1−14. [48] Z. Li, M. Flytzani‐Stephanopolos, J. Catal., 1999, 182, 313−327.
[29] F. Gunnarsson, J. A. Pihl, T. J. Toops, M. Skoglundh, H. Harelind, [49] A. Keshavaraja, X. She, M. Flytzani‐Stephanopolos, Appl. Catal. B,
Appl. Catal. B, 2017, 202, 42−50. 2000, 27, L1−L9.
[30] T. Miyadera, Appl. Catal. B, 1993, 2, 199−205. [50] D. G. Barton, S. L. Soled, G. D. Meitzner, G. A. Fuentes, E. Iglesia, J.
[31] K. Shimizu, J. Shibata, A. Satsuma, T. Hattori, Chem. Lett., 1999, Catal., 1999, 181, 57−72.
1079−1080. [51] D. E. Doronkin, S. Fogel, P. Gabrielsson, J.‐D. Grunwaldt, S. Dahl,
[32] K. Shimizu, A. Satsuma, T. Hattori, Appl. Catal. B, 2000, 25, Appl. Catal. B, 2014, 148−149, 62−69.
239−247. [52] K‐I. Shimizu, H. Kawabata, A. Satsuma, T. Hattori, J. Phys. Chem. B,
[33] R. Burch, J. P. Breen, C. J. Hill, B. Krutzsch, B. Konrad, E. Jobson, L. 1999, 103, 5240−5245.
Cider, K. Eraenen, F. Klingstedt, L. E. Lindfors, Top. Catal., 2004, [53] G. M. Underwood, T. M. Miller, V. H. Grassian, J. Phys. Chem. A,
30−31, 19−25. 1999, 103, 6184−6190.
[34] R. Burch, M. D. Coleman, Appl. Catal. B, 1999, 23, 115−121. [54] B. J. Adelman, T. Beutel, G. D. Lei, W. M. H. Sachtler, J. Catal., 1996,
[35] A. Ueda, T. Nakao, M. Azuma, T. Kobayashi, Catal. Today, 1998, 45, 158, 327−335.
135−138. [55] T. E. Hoost, K. A. Laframboise, K. Otto, Catal. Lett., 1996, 37,
[36] B. Frank, G. Emig, A. Renken, Appl. Catal. B, 1998, 19, 45−57. 153−156.
[37] M. Machida, D. Kurogi, T. Kijima, Chem. Mater., 2000, 12, [56] K. Hadjiivanov, D. Klissurski, G. Ramis, G. Busca, Appl. Catal. B,
3165−3170. 1996, 7, 251−267.
[38] M. Machida, S. Ikeda, D. Kurogi, T. Kijima, Appl. Catal. B, 2001, 35, [57] T. Tabata, H. Ohtsuka, M. Kokitsu, O. Okada, Bull Chem. Soc. Jpn.,
107−116. 1995, 68, 1905−1914.
[39] S. Satokawa, Chem. Lett., 2000, 294−295. [58] N. D. Parkyns, Proc. 5th Int. Conf. Catal., 1972, 12, 255.
[40] X. She, M. F. Stephanopoulos, J. Catal., 2006, 237, 79−93. [59] J. Valyon, W. K. Hall, J. Phys. Chem., 1993, 97, 1204−1212.
[41] H. He, Y. Li, X. L. Zhang, Y. B. Yu, C. B. Zhang, Appl. Catal. A, 2010, [60] J. A. Anderson, C.H. Rochester, J. Chem. Soc. Faraday Trans. 1, 1989,
375, 258−264. 85, 1117−1128.
[42] J. P. Breen, R. Burch, C. Hardacre, C. J. Hill, J. Phys. Chem. B, 2005, [61] K. Shimizu, J. Shibata, A. Satsuma, T. Hattori, Phys. Chem. Chem.
2252 Jia Wang et al. / Chinese Journal of Catalysis 42 (2021) 2242–2253

Phys., 2001, 3, 880−884. Phys., 2001, 3, 880−884.


[62] M. Haneda, N. Bion, M. Daturi, J. Saussey, J. C. Lavalley, D. Duprez, [75] K. Shimizu, H. Kawabata, A. Satsuma, T. Hattori, Appl. Catal. B,
H. Hamada, J. Catal., 2002, 206, 114−124. 1998, 19, L87−L92.
[63] A. Satsuma, T. Enjoji, K. Shimizu, K. Sato, H. Yoshida, T. Hattori, J. [76] K. Shimizu, K. Sawabe, A. Satsuma, Catal. Sci. Technol., 2011, 1,
Chem. Soc. Faraday Trans., 1998, 94, 301−307. 331−341.
[64] C. Morterra, G. Magnacca, Catal. Today., 1996, 27, 497−532. [77] P. Sazama, L. Capek, H. Drobna, Z. Sobalik, J. Dedecek, K. Arve, B.
[65] N. B. Colthup, L. H. Daly, S. E. Wiberley, Introduction to Infrared Wichterlova, J. Catal., 2005, 232, 302−317.
and Raman Spectroscopy, Academic Press, Boston, MA, 1990, [78] Y. B. Yu, H. He, X. L. Zhang, H. Deng, Catal. Sci. Technol., 2014, 4,
448–449. 1239−1245.
[66] K. Shimizu, J. Shibata, H. Yoshida, A. Satsuma, T. Hattori, Appl. [79] J. P. Breen, R. Burch, Top. Catal., 2006, 39, 53−58.
Catal. B, 2001, 30, 151−162. [80] P. Sazama, B. Wichterlova, Chem. Commun., 2005, 4810−4811.
[67] N. Bion, J. Saussey, M. Haneda, M. Daturi, J. Catal., 2003, 217, [81] T. Chaieb, L. Delannoy, G. Costentin, C. Louis, S. Casale, R. L.
47−58. Chantry, Z. Y. Li, C. Thomas, Appl. Catal. B, 2014, 156–157,
[68] S. Satokawa, J. Shibata, K. Shimizu, A. Satsuma, T. Hattori, Appl. 192–201.
Catal. B, 2003, 42, 179−186. [82] R. You, W. X. Huang, ChemCatChem, 2020, 12, 675–688.
[69] A. Iglesias‐Juez, A. B. Hungría, A. Martínez‐Arias, A. Fuerte, M. [83] C. A. Taatjes, N. Hansen, A. Mcllroy, J. A. Miller, J. P. Senosiain, S. J.
Fernández‐García, J. A. Anderson, J. C. Conesa, J. Soria, J. Catal., Klippenstein, F. Qi, L. Sheng, Y. Zhang, T. A. Cool, J. Wang, P. R.
2003, 217, 310−323. Westmoreland, M. E. Law, T. Kasper, K. Kohse‐Höinghaus, Science,
[70] J. Shibata, K. Shimizu, S. Satokawa, A. Satsuma, T. Hattori, Phys. 2005, 308, 1887–1889.
Chem. Chem. Phys., 2003, 5, 2154−2160. [84] L. Luo, X. Tang, W. Wang, Y. Wang, S. Sun, F. Qi, W. Huang, Sci. Rep.,
[71] T. Furusawa, L. Lefferts, K. Seshan, K. Aika, Appl. Catal. B, 2003, 42, 2013, 3, 1625–1632.
25−34. [85] R. You, X. Y. Zhang, L. F. Luo, Y. Pan, H. B. Pan, J. Z. Yang, L. H. Wu, X.
[72] H. He, J. Wang, Q. Feng, Y. Yu, K. Yoshida, Appl. Catal. B, 2003, 46, S. Zheng, Y. K. Jin, W. X. Huang, J. Catal., 2017, 348, 189–199
365−370. [86] M. L. Unland, J. Phys. Chem., 1973, 77, 1952−1956.
[73] Y. Yu, H. He, Q. Feng, J. Phys. Chem. B, 2003, 107, 13090−13092. [87] S. Tamm, H. H. Ingelsten, A. E. C. Palmqvist, J. Catal., 2008, 255,
[74] K. Shimizu, J. Shibata, A. Satsuma, T. Hattori, Phys. Chem. Chem. 304−312.

Cl–改性对Ag/Al2O3催化剂结构及其催化C3H6-SCR和H2/C3H6-SCR
反应性能的影响
王 嘉a, 尤 瑞a,*, 千 坤a, 潘 洋b, 杨玖重b, 黄伟新a,c,#
a
中国科学技术大学化学物理系, 中国科学院能量转换材料重点实验室, 安徽省教育厅表界面化学与能源催化重点实验室,
合肥微尺度物质科学国家研究中心, 安徽合肥230026
b
中国科学技术大学国家同步辐射实验室, 安徽合肥230029
c
洁净能源国家实验室, 辽宁大连116023

摘要: 以烯烃为还原剂的NOx选择性催化还原(HC-SCR)是重要的环境催化反应之一. Ag/Al2O3催化剂(SA)因在HC-SCR反


应中表现高活性、高N2选择性及中等H2O和SO2耐受性等优点, 而被广泛研究. SA催化剂中存在不同的Ag物种, 包括孤立
Ag+离子, 带部分正电荷Agnδ+团簇和金属态Agn0团簇. 文献研究结果表明, SA催化剂中Agnδ+团簇是催化HC-SCR反应的活
性Ag物种, 而Ag物种类型与Ag的负载量密切相关. 因此文献中研究SA催化HC-SCR反应的结构-性能关系主要是通过改变
Ag负载量来开展的, 最优Ag负载量约为1%~2%.
本文以Cl–改性的γ-Al2O3作为载体, 采用传统的浸渍法制备了Ag/Al2O3-Cl催化剂(SA-Cl), 通过XRD、TEM、H2-TPR、
UV-Vis DRS以及XPS对催化剂进行了结构表征, 并结合C3H6-SCR和H2/C3H6-SCR活性测试, 建立催化剂结构-催化性能关
系; 同时利用原位傅里叶变换红外光谱(DRIFTS)和在线同步辐射单光子电离质谱(SVUV-PIMS)研究了SA催化HC-SCR的
反应机理.
结构表征结果表明, 在SA催化剂中, Ag负载量的提高主要是增加了Agn0物种, 而在SA-Cl催化剂中, Ag负载量的提高主
要是增加了Agnδ+物种, 因此Cl‒改性能促进SA催化剂中Agnδ+物种的形成. 活性测试结果表明, 在相同Ag负载量下, SA-Cl催
化剂表现出比SA催化剂更好的HC-SCR催化性能. Cl‒改性对SA催化剂中Ag物种的调控作用和HC-SCR催化性能的促进作
用随Ag负载量的增加变得更为明显. 原位DRIFTS结果表明, γ-Al2O3载体(Al位点)是NO氧化形成硝酸盐物种的主要活性位
点; Agnδ+物种催化丙烯适度氧化主要生成乙酸盐类物种, 可以还原表面硝酸盐物种; 而Agn0催化丙烯过度氧化主要生成羧
酸盐类物种, 进而生成CO2, 不能还原表面硝酸盐物种. 由此可见, Agnδ+是催化HC-SCR反应的活性Ag物种, 而Agn0是催化
烃类完全氧化反应的活性Ag物种; Cl–改性能有效促进Agnδ+的形成, 进而提高HC-SCR催化反应活性. 在线SVUV-PIMS结
果检测到H2/C3H6-SCR反应中存在气态中间物种丙烯腈(CH2=CHCN). –CN和–NCO物种被认为是HC-SCR反应的关键中间
物种, 能够直接与气相NO+O2反应生成N2. 因此, CH2=CHCN的存在说明HC-SCR反应涉及到气相反应机理.
Jia Wang et al. / Chinese Journal of Catalysis 42 (2021) 2242–2253 2253

基于SA和SA-Cl催化剂, 进一步研究了H2 对C3H6-SCR低温活性的促进作用. 结果表明, H2 的促进作用是通过作用于


Agnδ+物种, 而不是通过Agn0物种来实现的; H2的引入有利于低温下强吸附硝酸盐物种的脱附或分解以及中间体向–NCO和
–CN物种的转化, 从而提高HC-SCR低温催化活性.
综上, 基于Cl– 改性的Ag/Al2O3-Cl催化剂, 本文成功证实了Agnδ+物种是催化HC-SCR反应的活性Ag物种, 并结合原位
DRIFTS在线SVUV-PIMS谱分别鉴定了催化反应表面中间物种和气相中间物种. 这些结果加深了对SA催化HC-SCR反应
构-效关系和反应机理的基础理解.
关键词: 原位表征; 反应机理; 构效关系; 漫反射红外傅里叶变换光谱; 光电离质谱

收稿日期: 2020-12-17. 接受日期: 2021-02-07. 上网时间: 2021-09-10.


*通讯联系人. 电话: (0551)63600435; 传真: (0551)63600437; 电子信箱: yourui@ustc.edu.cn
#
通讯联系人. 电子信箱: huangwx@ustc.edu.cn
基金来源: 国家自然科学基金(21703227); 中国科学技术大学(KY2060000176); 中国科学院; 教育部长江学者奖励计划.
本文的电子版全文由Elsevier出版社在ScienceDirect上出版(http://www.sciencedirect.com/journal/chinese-journal-of-catalysis).

You might also like