You are on page 1of 10

Applied Surface Science 356 (2015) 181–190

Contents lists available at ScienceDirect

Applied Surface Science


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

Promotional effect of tungsten-doped CeO2 /TiO2 for selective


catalytic reduction of NOx with ammonia
Dong Wook Kwon, Sung Chang Hong ∗
Department of Environmental Energy Engineering, Graduate School of Kyonggi University, 94-6 San, Iui-dong, Youngtong-ku, Suwon-si, Gyeonggi-do
443-760, Republic of Korea

a r t i c l e i n f o a b s t r a c t

Article history: We examined the effects that the physicochemical properties of Ce/Me/Ti catalysts had on the selective
Received 16 May 2015 catalytic reduction (SCR) activity after various metals (W, Mo, and La) were added to non-vanadium-
Received in revised form 29 July 2015 based catalysts in order to improve NH3 –SCR activity. We studied the properties of the catalysts through
Accepted 9 August 2015
the use of physiochemical techniques, including Brunauer–Emmett–Teller (BET) surface area analysis,
Available online 11 August 2015
X-ray diffraction (XRD), H2 temperature-programmed reduction (H2 -TPR), X-ray photoelectron spec-
troscopy (XPS) transmission infrared spectroscopy (IR), and inductively coupled plasma optic emission
Keywords:
spectroscopy (ICP). The catalytic activity tests of the Ce/Ti catalysts with various ceria loadings revealed
Cerium oxide
SO2 resistance
that the Ce/Ti with 10 wt.% ceria (10Ce/Ti) exhibited excellent activity. Thus, various metals were added
NH3 –SCR to the 10Ce/Ti. The tungsten-doped 10Ce/Ti catalyst exhibited the highest activity (10Ce/W/Ti: Ce was
Ce/W/Ti deposited after tungsten had been deposited on TiO2 ). We investigated the correlation between the cat-
Tungsten oxide alyst’s Ce valence state and its activity. Different Ce3+ ratios were observed when various metals were
added to Ce/Ti. The highest Ce3+ ratio was observed in 10Ce/W/Ti at 0.3027, and the catalyst efficiency
had a positive correlation with higher Ce3+ ratios. The SCR activity was found to increase as the Ce3+ ratio
increased when tungsten was added to 10Ce/W/Ti. Furthermore, in the case of 10Ce/W/Ti, it seemed that
the Brønsted acid sites were more abundant relative to those on 10Ce/Ti. Upon the injection of SO2 in the
SCR reaction, 10Ce/Ti was rapidly deactivated. However, the 10Ce/W/Ti catalyst exhibited an excellent
resistance to SO2 -induced deactivation relative to 10Ce/Ti. Thus, the addition of tungsten to Ce/Ti resulted
in excellent NOx conversion and SO2 resistance.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction to o SO3 , a narrow operation temperature window (300–400 ◦ C),


and the formation of N2 O at high temperatures [5,6]. Therefore,
The NOx that is emitted during the combustion processes of recent efforts have been made to develop an environmentally
fossil fuel causes photochemical smog, acid rain, the greenhouse friendly SCR catalyst to replace V-based catalysts.
effect, and the depletion of the ozone layer [1,2]. Thus, the selec- In particular, ceria is a potential substitute for vanadium due
tive catalytic reduction (SCR) of NOx to N2 , with NH3 as a reductant to its oxygen storage capacity and high redox ability when shif-
on VOx/TiO2 , has been known to be the best available control tech- ting between Ce4+ and Ce3+ [7,8]. Surface acid sites are well known
nology (BACT) to remove NOx from the atmosphere. It has been to promote absorption and to increase NH3 –SCR activity [9], so
widely used in a large number of commercial combustion processes various studies have been conducted to develop SCR catalysts
as result of its economic feasibility and efficiency [3]. using ceria, and several catalyst systems have been developed,
The SCR catalysts that are most commonly used are those in including CeO2 /TiO2 [10,11], CeO2 /Al2 O3 [12], CeO2 /WO3 [13,14],
which WO3 is added to VOx/TiO2 (anatase). These vanadium-based MnOx CeO2 [15], CeO2 /ZrO2 [16], CeO2 /TiO2 SiO2 [17], Ge(or
catalysts exhibit an excellent efficiency in a temperature range Mn)/CeO2 WO3 [18], and Fe Ce Mn/ZSM-5 catalysts [19]. Gao
from 300 to 400 ◦ C [4]. However, these vanadium-based catalysts et al. [20,21] demonstrated that the Ce Cu Ti mixed oxide catalyst
also suffer from several drawbacks, including the harmful effects of exhibited excellent SCR activity at a low temperature. According
vanadium on the environment, the high likelihood of SO2 oxidizing to Shan et al. [10], the CeO2 /TiO2 catalyst exhibited a higher cat-
alytic activity than did V2 O5 WO3 /TiO2 and Fe-ZSM-5 catalysts.
They claimed that the high NH3 –SCR activity was mainly a result of
∗ Corresponding author. the synergistic effects between CeO2 and TiO2 . In addition, they
E-mail address: schong@kgu.ac.kr (S.C. Hong). examined the promotional effects of a tungsten-doped Ce/TiO2

http://dx.doi.org/10.1016/j.apsusc.2015.08.073
0169-4332/© 2015 Elsevier B.V. All rights reserved.
182 D.W. Kwon, S.C. Hong / Applied Surface Science 356 (2015) 181–190

catalyst. The excellent catalytic performance of the Ce W Ti cata- set to 750 ppm NO, 48 ppm NO2 , 800 ppm NH3 , 3 vol.% O2 , 6 vol.%
lysts was associated with the highly dispersed ceria and primitive H2 O, and 200 ppm SO2 (when used) in Ar. The reactant gases were
W species on TiO2 [22]. Liu et al. [23] reported that the addition fed to the reactor using a mass flow controller (MKS Co.). Approxi-
of molybdenum to a Ce/Ti catalyst improved the NH3 –SCR activ- mately 0.15 g of the catalyst (40–50 mesh, 0.3 cc) were used for each
ity. Thus, the above studies have revealed that the addition of a test. Under ambient conditions, the total flow rate was 600 cc/min,
promoter to a Ce/Ti catalyst could improve the NH3 –SCR catalytic and the gas hourly space velocity (GHSV) was 120,000 h−1 . The
activity. Although the catalyst modified by the addition of a pro- composition of the feed gases and the effluent streams were con-
moter shows excellent SCR activity, further studies are required to tinuously monitored on-line using non-dispersive infrared gas
investigate the correlation between the activity and the catalytic analyzers: ZKJ-2 (Fuji Electric Co.) for NO/NO2 ; Ultramat 6 (Siemens
characteristics of the metal added to the Ce/Ti system. Co.) for N2 O; and a pulsed fluorescence gas analyzer 43 C-High
Accordingly, we prepared catalysts in this study by adding Level (Thermo Co.) for SO2 . Before the gas flowed into the analyzers,
various metals to Ce/Ti to improve the NH3 –SCR activity of the moisture was removed with a moisture trap inside the chiller.
non-vanadium-based catalysts. We examined the effects of the The NOx conversion and the rate constant were calculated from
physicochemical properties of the metal-doped Ce/Me/TiO2 on the gas concentrations at a steady state according to the following
the SCR activity, and we investigated the correlation between the equations:
catalyst’s Ce valence state and its activity. We also studied the
properties of the catalyst through various physiochemical analy- NOxconversion(%)
ses, which included Brunauer–Emmett–Teller (BET) surface area = (([NO + NO2 ]in − [NO + NO2 ]out )/([NO + NO2 ]in )) × 100 (1)
analysis, X-ray diffraction (XRD), H2 temperature-programmed
reduction (H2 -TPR), X-ray photoelectron spectroscopy (XPS) trans-
mission infrared spectroscopy (IR), and inductively coupled plasma
optic emission spectroscopy (ICP). NH3 conversion(%) = (([NH3 ]in − [NH3 ]out )/([NH3 ]in )) × 100 (2)

2.3. Catalyst characterization


2. Experimental
The BET surface areas of the catalysts were measured using ASAP
2.1. Catalyst preparation
2010C (Micromeritics Co.). The specific surface area was estimated
using the BET method, and the pore size of the distribution was cal-
The Ce/Ti system was prepared via wet impregnation using
culated using the Barrett–Joyner–Halenda method that calculates
cerium(III) nitrate [10 wt.% Ce; Ce(NO3 )3 ·6H2 O, Aldrich Chemical
the pore size based on the adsorption layer thickness in relation to
Co.] by using ceria (5, 10, 20, and 50 wt.% Ce) and anatase-type TiO2
the pressure and the average radius of the meniscus, as obtained
powder (DT51, Cristal Global Co.). A calculated amount of TiO2 was
using the Kelvin method. Each sample was then analyzed after
gradually added to the ceria solution while stirring. The mixture
degassing in a vacuum at 110 ◦ C for 3–5 h. The XRD measurements
was agitated in a slurry state for more than 1 h; then the moisture
were carried out using Cu K␣ ( = 1.5056 Å) radiation, and the cat-
was removed at 70 ◦ C using a rotary vacuum evaporator (Eyela Co.,
alysts were measured at the 2 range from 10◦ to 80◦ with a step
N–N series); and the mixture was dried overnight at 110 ◦ C. Finally,
size of 0.1◦ and a time step of 1.0 s with a PANalytical X’Pert Pro
the mixture was calcined in air for 5 h at 500 ◦ C.
MRD. The TPR of H2 was measured using 10% H2 /Ar and 0.3 g of the
The Ce/Me/Ti catalyst was prepared via wet impregna-
catalyst, at a total flow rate of 50 cc/min. Prior to the H2 -TPR mea-
tion using TiO2 , cerium(III) nitrate, ammonium metatungstate
surements, the catalyst was pretreated under a flow of 5% O2 /Ar at
hydrate [(NH4 )6 H2 W12 O40 , Aldrich Co.], ammonium molybdate
400 ◦ C for 0.5 h, followed by cooling to 50 ◦ C. The catalyst was then
[(NH4 )6 Mo7 O24 ·4H2 O, Aldrich Co.], and lanthanum(III) nitrate
placed in hydrogen, and the hydrogen consumption was monitored
hexahydrate (LaN3 O9 ·6H2 O, Aldrich Co.). The calculated amount of
with an Autochem 2920 (Micromeritics) instrument while the tem-
the various metals (10 wt.% WO3 , 10 wt.% MoO3 , and 10 wt.% La2 O3 )
perature increased to 800 ◦ C at a rate of 10 ◦ C/min. An ESCALAB
was immersed in distilled H2 O. After mixing the metal solution with
210 (VG Scientific) was used to conduct the XPS analysis with
TiO2 , the moisture was removed at 70 ◦ C using a rotary vacuum
monochromatic Al K␣ (1486.6 eV) as the excitation source. After
evaporator, and the mixture was then dried overnight at 110 ◦ C.
the complete removal of moisture from the catalysts by drying at
The supports were calcined in air for 8 h at 600 ◦ C, and calculated
100 ◦ C for 24 h, the analysis was carried out without surface sputter-
amounts of cerium(III) nitrate (10 wt.% Ce) were dissolved in dis-
ing and etching so that the degree of vacuum in the XPS equipment
tilled water. After impregnation of the ceria solution into the W/Ti,
could be maintained at 10−6 Pa. The spectra were analyzed using
Mo/Ti, or La/Ti supports, the moisture was removed at 70 ◦ C using
the XPS PEAK software (version 4.1), and the intensities were esti-
a rotary vacuum evaporator, and the mixture was dried overnight
mated from the integration of each peak, subtraction of the Shirley
at 110 ◦ C. The catalysts were then calcined in air for 5 h at 500 ◦ C.
background, and fitting of the experimental curve to a combina-
The 10Ce W/Ti catalyst was prepared according to the fol-
tion of Lorentzian and Gaussian curves of various proportions. All
lowing process. Cerium nitrate (10 wt.% Ce) and ammonium
of the binding energies were referenced to the C 1s line at 284.6 eV,
paratungstate (10 wt.% WO3 ) were mixed in distilled H2 O. The
and the binding energy values were measured with a precision of
calculated amount of the TiO2 support was impregnated in this
±0.3 eV. This study used an in situ DRIFTs analysis performed with
solution by stirring for 1 h, and then the moisture was removed at
an FT-IR (Nicolet iS10, Thermo Co.). A DR (diffuse reflectance) 400
70 ◦ C using a rotary vacuum evaporator. The mixture was then dried
accessory was used for the analysis of the solid reflectance. The CaF2
overnight at 110 ◦ C, and was subsequently calcined in air for 5 h at
window was used as a plate for the DR measurement, and spectra
500 ◦ C. W/10Ce/Ti was prepared through impregnation of 10Ce/Ti
were collected using a MCT (mercury cadmium telluride) detec-
and ammonium paratungstate using the method described above.
tor. All of the catalysts used for this analysis were ground using
a rod in the sample pan of the in situ chamber with an installed
2.2. Catalytic activity tests temperature controller. To prevent the influence of moisture and
impurities, the sample was preprocessed with Ar at a rate of 100
The SCR activity tests were carried out in a fixed-bed quartz reac- cc/min at 400 ◦ C for 1 h. Then, the vacuum state was maintained
tor with an inner diameter of 6 mm. The reaction conditions were using a vacuum pump. The spectra of the catalyst was collected by
D.W. Kwon, S.C. Hong / Applied Surface Science 356 (2015) 181–190 183

measuring the single-beam spectrum of the preprocessed sample


as background, and all analyses were performed by auto scanning
at a resolution of 4 cm−1 . The elemental (sulfur) analyses were per-
formed via inductively coupled plasma optic emission spectroscopy
(ICP). The transmission IR spectra were recorded with a Nicolet
Nexus spectrometer (Model Magna IR 550 II) equipped with a liq-
uid nitrogen cooled MCT detector that collected 100 scans with a
spectral resolution of 4 cm−1 . The samples were then mixed with
KBr and were pressed into pellets to conduct the measurements.

3. Results and discussion

3.1. Catalytic activity and characteristics for NO reduction by


NH3 over Ce/Ti catalysts

The experiments were performed using Ce/Ti catalysts to exam-


ine the influence of the ceria loading on the SCR activity. As
shown in Fig. 1a, the NOx conversion activities were determined
on 5Ce/Ti, 10Ce/Ti, 20Ce/Ti, and 50Ce/Ti at 250–400 ◦ C with a GHSV
of 120,000 h−1 . The 10Ce/Ti catalyst exhibited an activity of 91% or
higher at 300 ◦ C and showed excellent activity over the entire tem-
perature range. However, if 10 wt.% ceria or more was deposited,
the activity decreased. The 50Ce/Ti catalyst exhibited very low
activity at 300 ◦ C or higher, and N2 O was not found in the SCR
reactions of all catalysts. As shown in Fig. 1b, the concentration of
NO2 emitted from the SCR reactions increased as the ceria loading
increased, and Fig. 1c shows the NH3 conversions for the SCR reac-
tions. These conversions are proportional to the NOx conversions
because the SCR reactions are held at a molar ratio of 1.0 NH3 /NOx.
Therefore, the NH3 conversion curves are related to the tempera-
ture range in which the NOx conversion decreases. The surface area
for the 5Ce/Ti, 10Ce/Ti, 20Ce/Ti, and 50Ce/Ti samples are collated in
Table 1 with values of 86.6, 87.9, 55.5, and 44.8 m2 /g, respectively.
10Ce/Ti had the highest surface are. However, the specific surface
area, total pore volume, and average pore diameter decreased with
tungsten loadings that exceeded 10 wt.%.
The variation in the catalyst structure with different ceria load-
ings was analyzed through the use of XRD. Fig. 2 shows the XRD
patterns of the 5Ce/Ti, 10Ce/Ti, 20Ce/Ti, and 50Ce/Ti catalysts. For
cubic CeO2 and anatase TiO2 , main peaks were observed at 28.6◦
and 25.3◦ , respectively, and the intensity of the anatase TiO2 peak
decreased as the ceria loading increased. For the 5Ce/Ti, 10Ce/Ti,
and 20Ce/Ti catalysts, cubic CeO2 peaks were not observed in the
XRD results, which indicates that the ceria oxides were well dis-
persed on the TiO2 support structure. When the 50Ce/Ti catalyst
was analyzed, cubic CeO2 peaks were observed at 28.6, 33.1, and
47.5. When ceria was not evenly distributed on the support, crys-
talline ceria formed, and the catalytic activity decreased [24,25].
According to Gao et al. [7], when ceria exists in an amorphous state
in Ce/Ti, the amount of chemisorbed oxygen and oxygen species
that are bound to the catalyst increase, and excellent catalytic activ-
ity can be observed. Therefore, when 20 wt.% ceria or more was
deposited, Ce did not exist in an amorphous state. Thus, crystalline
Fig. 1. Effect of reaction temperature on (a) NOx conversion, (b) NH3 conversion,
Ce could be said to have formed, and the formation of crystalline and (c) outlet NO2 concentration of 5Ce/Ti, 10Ce/Ti, 20Ce/Ti, and 50Ce/Ti cata-
ceria on the surface of the catalyst was thought to decrease the SCR lysts (NO: 750 ppm, NO2 : 48 ppm, NH3 /NOx: 1.0, O2 : 8 vol.%, H2 O: 6 vol.%, GHSV:
activity. 120,000 h−1 ).
An H2 -TPR analysis was carried out to investigate the presence of
reducible species in TiO2 , 5Ce/Ti, 10Ce/Ti, 20Ce/Ti, and 50Ce/Ti. The
TiO2 and ceria peaks in Ce/Ti catalysts were distinguished by per- for the Ce/Ti catalysts at 410, 480, and 515 ◦ C. The first peak at
forming H2 -TPR analyses of TiO2 and Ce/Ti. Fig. 3 shows the results 410 ◦ C could be attributed to the reduction in the surface oxygen
of the H2 -TPR analysis of TiO2 , 5Ce/Ti, 10Ce/Ti, 20Ce/Ti, and 50Ce/Ti. of stoichiometric ceria (Ce4+ O Ce4+ ) [26], and the second peak
All samples exhibited a major reduction in the peaks in the 200 at 480 ◦ C was attributed to the reduction of non-stoichiometric
to 600 ◦ C temperature range. For the TiO2 support, the maximum ceria (Ce3+ O Ce4+ ) [27]. In addition, the peak observed at 700 ◦ C
temperature of the hydrogen consumption was 490 ◦ C. When Ce in the 50Ce/Ti catalyst corresponded to that of bulk CeO2 [28].
was impregnated into the TiO2 support, reduction peaks emerged For mixed oxide Ce/Ti catalysts, the reduction peak areas of the
184 D.W. Kwon, S.C. Hong / Applied Surface Science 356 (2015) 181–190

Table 1
Surface area, total pore volume, average pore diameter, and H2 consumption for
Ce/Ti catalysts.

Catalyst SBET Total pore volume Average pore H2


(m2 /g) (p/p0 = 0.304) diameter (nm) consumption
(cm3 /g) (␮mol/g)

5Ce/Ti 86.6 0.043 1.983 493


10Ce/Ti 87.9 0.044 1.992 725
20Ce/Ti 55.5 0.026 1.847 1042
50Ce/Ti 44.8 0.019 1.429 1443

Ce/Ti catalysts increased as the ceria loading increased. As shown in


Table 1, the H2 consumption values of 5Ce/Ti, 10Ce/Ti, 20Ce/Ti, and
50Ce/Ti were 493, 725, 1042, and 1443 ␮mol/g, respectively. The H2
consumption increased as the ceria loading in the Ce/Ti catalysts
increased, and as a result, the removal observed for oxygen from
the 50Ce/Ti catalyst was the largest. The interaction between Ce and
TiO2 significantly influenced the reducibility of the catalyst and is
largely dependent on the ceria loading. However, when compared Fig. 4. Effect of reaction temperature on NOx conversion of 10Ce/Ti, 10Ce/W/Ti,
with the catalytic activity, the reducibility of Ce/Ti catalysts did not 10Ce/Mo/Ti, and 10Ce/La/Ti catalysts (NO: 750 ppm, NO2 : 48 ppm, NH3 /NOx: 1.0,
exhibit a direct correlation. O2 : 8 vol.%, H2 O: 6 vol.%, GHSV: 120,000 h−1 ).

Fig. 5. Effect of reaction temperature on NOx conversion of 10Ce/W/Ti, 10Ce W/Ti,


Fig. 2. XRD pattern of 5Ce/Ti, 10Ce/Ti, 20Ce/Ti, and 50Ce/Ti catalysts. and W/10Ce/Ti catalysts (NO: 750 ppm, NO2 : 48 ppm, NH3 /NOx: 1.0, O2 : 8 vol.%,
H2 O: 6 vol.%, GHSV: 120,000 h−1 ).

3.2. Catalytic activity of catalysts doped with different transition


metals

The influence that the metal added to the catalyst had on the
NOx conversion was evaluated, and the results are presented in
Fig. 4, which the NOx conversion from 250 to 400 ◦ C with a GHSV of
120,000 h−1 using 10Ce/Ti, 10Ce/W/Ti, 10Ce/Mo/Ti, and 10Ce/La/Ti
catalysts. The highest activity was achieved when 10 wt.% WO3 was
added to 10Ce/Ti. In addition, we observed the catalytic activity
of the 10Ce/W/Ti, 10Ce W/Ti, and W/10Ce/Ti catalysts to exam-
ine the effect of the different methods to add tungsten. As shown
in Fig. 5, 10Ce/W/Ti had the highest activity over all temperature
ranges. The surface areas of the 10Ce/Ti, 10Ce/W/Ti, 10Ce/Mo/Ti,
10Ce/La/Ti, 10Ce W/Ti, and W/10Ce/Ti catalysts were 87.9, 80.1,
79.6, 89.9, 76.5, and 74.7 m2 /g, respectively (Table 2), and differ-
ent surface areas were obtained as metals were added to 10Ce/Ti.
The 10Ce/La/Ti catalyst had the largest surface area at 89.9 m2 /g. As
the metals were added to 10Ce/Ti, the surface areas of all catalysts
Fig. 3. H2 -TPR profiles of TiO2 , 5Ce/Ti, 10Ce/Ti, 20Ce/Ti, and 50Ce/Ti catalysts (exper-
decreased, except for that of 10Ce/La/Ti.
imental conditions: 30 mg catalyst, pretreatment at 400 ◦ C for 30 min with 5% O2 /Ar
at 50 cm3 /min, 10% H2 /Ar reduction with a heating rate of 10 ◦ C/min, and a total flow An XPS analysis offers an effective method to analyze the
rate of 50 cm3 /min). surface of the catalysts. Fig. 6 shows the deconvoluted O 1s
D.W. Kwon, S.C. Hong / Applied Surface Science 356 (2015) 181–190 185

Table 2
Surface area, O␣ ratio, and Ce3+ ratio of various catalysts.

Catalyst SBET (m2 /g) O␣ ratioa Ce3+ ratiob

10Ce/Ti 87.9 0.2604 0.2678


10Ce/W/Ti 80.1 0.4204 0.3027
10Ce/Mo/Ti 79.6 0.3032 0.2762
10Ce/La/Ti 89.9 0.2450 0.2524
10Ce W/Ti 76.5 0.3723 0.2725
W/10Ce/Ti 74.7 0.3541 0.2514
a
Surface-adsorbed oxygen ratio calculated from XPS data.
b
Ce3+ ratio calculated from XPS data.

Fig. 7. Correlation between O␣ ratio and NOx conversion over various catalysts.

oxidation of NO to NO2 in an SCR reaction, resulting in an improve-


ment to a “fast SCR” reaction [30,31]. Fig. 7 shows the correlation
between the O␣ ratio and the catalytic activity, and the 10Ce/W/Ti
catalyst with the highest O␣ ratio exhibited excellent catalytic
activity. However, the increase in the O␣ ratio was not associated
with the catalytic activity for all catalysts.
The valence state of the active metals that are involved is
an important factor in the NH3 –SCR reaction, which for Ce was
of either Ce3+ or Ce4+ . Thus, we examined the influence of the
addition of various metals on the Ce valence state in Mn/Ce/W/Ti.
Fig. 8 shows the Ce 3d peaks in the XPS profiles of the 10Ce/Ti,
10Ce/W/Ti, 10Ce/Mo/Ti, 10Ce/La/Ti, 10Ce W/Ti, and W/10Ce/Ti
catalysts. The Ce 3d peak was fitted into sub-bands by searching
for the ideal combination of Gaussian bands, and the sub-bands
marked u1 and v1 represent the 3d10 4f1 initial electronic state
corresponding to Ce3+ while the sub-bands marked u, u2 , u3 , v, v2 ,
and v3 represent the 3d10 4f0 state of Ce4+ [22]. The Ce4+ ratios of
all samples were calculated as Ce4+ /(Ce3+ + Ce4+ ) and are shown
in Table 2. The Ce3+ ratios could be ranked as follows: 10Ce/W/Ti
(0.3027) > 10Ce/Mo/Ti (0.2762) > 10Ce W/Ti (0.2725) > 10Ce/Ti
(0.2678) > 10Ce/La/Ti (0.2524) > W/10Ce/Ti (0.2514), and the high-
est Ce3+ ratio of 0.3027 was observed for 10Ce/W/Ti. Fig. 9 shows
Fig. 6. O 1s spectra of (a) 10Ce/Ti, (b) 10Ce/W/Ti, (c) 10Ce/Mo/Ti, (d) 10Ce/La/Ti, (e)
the correlation between the Ce3+ ratio and the catalyst activity. An
10Ce W/Ti, and (f) W/10Ce/Ti catalysts from XPS analysis. excellent SCR activity was achieved at higher Ce3+ ratios. According
to Chen et al. [8], Ce3+ may improve the SCR reaction by promoting
the oxidation of NO to NO2 , and NO oxidation experiments were
spectra of the 10Ce/Ti, 10Ce/W/Ti, 10Ce/Mo/Ti, 10Ce/La/Ti, 10Ce- carried out to investigate the influence that the tungsten in 10Ce/Ti
W/Ti, and W/10Ce/Ti catalysts. The O 1s peak was fitted into had on the catalyst activity (Fig. 10). With the addition of tungsten
sub-bands by searching for an ideal combination of Gaussian to 10Ce/Ti, the conversion of NO to NO2 increased in a reaction
bands. The two bands at 531.0–531.6 and 532.8–533.0 eV are from 250 to 400 ◦ C. Particularly, the NO2 production increased
respectively assigned to surface-adsorbed oxygen (O␣ ), such when 10Ce/W/Ti was used at 300 ◦ C or less. For this reason, SCR
as O2 2− and O− , which are related to chemisorbed water catalysts with the capacity for oxidation of NO to NO2 exhibit
(expressed as O ␣ ) [29]. The sub-bands from 529.5 to 530.0 eV an excellent activity at low temperatures due to the influence of
are attributable to the lattice O2− oxygen (expressed as O␤ ) “fast SCR” [32]. Shan et al. [22,33] claimed that the introduction
[27]. The O␣ ratio, which is calculated as O␣ /(O␣ + O ␣ + O␤ ), of W species could increase the Ce3+ ratio, number of acid sites,
varied as metals were added to 10Ce/Ti (Table 2). The and number of oxygen vacancies of the catalyst, which are all
O␣ ratios in Ce/Me/Ti decreased in the order of 10Ce/W/Ti features that improve the low-temperature activity of the catalyst.
(0.4204) > 10Ce W/Ti (0.3723) > W/10Ce/Ti (0.3541) > 10Ce/Mo/Ti Thus, the SCR activity was observed to increase as the Ce3+ ratio
(0.3032) > 10Ce/Ti (0.2604) > 10Ce/La/Ti (0.2450). Wu et al. [30] increased when tungsten was added to 10Ce/W/Ti.
mentioned that surface-chemisorbed oxygen (O␣ ) is the most The NH3 adsorption capacity of the 10Ce/Ti and 10Ce/W/Ti
active oxygen species, and that it is highly active in an oxidation catalysts was compared in order to investigate the cause for
reaction due to its high mobility. The surface-adsorbed oxygen which the activity improved after the addition of tungsten. The
is considered to be more reactive in oxidation reactions since adsorption characteristics of the catalysts were measured by con-
its mobility is higher than that of lattice oxygen, and several ducting a DRIFTs analysis, and the acid site of the catalysts in
researchers have insisted that a high O␣ ratio is beneficial for the the SCR reaction was examined by performing an NH3 adsorption
186 D.W. Kwon, S.C. Hong / Applied Surface Science 356 (2015) 181–190

analysis using a DRIFT spectrometer at 250 ◦ C (Fig. 11a). The 1670 cm−1 [34]. For 10Ce/Ti, the Brønsted acid sites showed a very
1000 ppm NH3 was injected for 30 min, and the adsorbed species small peak, and for 10Ce/W/Ti, it seemed that the Brønsted acid
were examined. In these experiments, peaks corresponding to –OH, sites were more abundant relative to those for 10Ce/Ti. The largest
caused by the adsorption of NH4 + onto –OH, were observed at negative peak of the adsorption of NH4 + onto –OH was observed
3674 cm−1 . The adsorption of NH3 on Lewis acid sites was observed at 3674 cm−1 for 10Ce/W/Ti, and furthermore, bands that did not
at 1605, 3170, 3256, and 3364 cm−1 [34], and the 10Ce/Ti and belong to Lewis or Brønsted acid sites also appeared at 1560 cm−1 .
10Ce/W/Ti exhibited peaks corresponding to Lewis acid sites and Ramis et al. [36] proposed that bands at 1550 and 1570 cm−1 may
to –OH. In the case of the 10Ce/W/Ti impregnated with tungsten, be related to the intermediate of the oxidation of ammonia. Accord-
a negative peak was observed at 2020 cm−1 , which corresponds ingly, bands at 1560 cm−1 might be attributed to amid (–NH2 )
to W O [35]. In addition, the adsorption of the NH4 + ion on species that can be assigned as an intermediate of the oxidation of
the Brønsted acid sites was simultaneously observed at 1430 and ammonia. The 10Ce/Ti catalyst was purged with NH3 for 0.5 h, and
NO + O2 /N2 was then introduced into the IR at 200 ◦ C, and the spec-
tra were recorded as a function of time (Fig. 11b). After NO + O2 had
been injected, all of the ammonia species decreased. In the case of
10Ce/Ti, the peaks of the ammonia species decreased after 15 min.
At the same time, new bands were detected at 1622 and 1580 cm−1
that could be attributed to the NOx species [37]. When compared
to 10Ce/Ti, the peaks assigned to the adsorbed ammonia species
decreased more quickly on for the 10Ce/W/Ti catalyst (Fig. 11c).
When the catalyst was purged with NO + O2 for only 5 min, all of
the peaks resulting from ammonia species decreased. At the same
time, new bands were detected at 1610 and 1580 cm−1 that could
be attributed to NOx species [38]. On the basis of Fig. 11b and c,
it could be concluded that NOx readily reacted with the adsorbed
ammonia species, especially for the 10Ce/W/Ti catalyst. Accord-
ing to Chen et al. [39], the Ce/Ti and Ce/W/Ti catalysts represents
the E–L mechanism. The adsorbed ammonia was first transformed
to amide species, then formed nitrosamine (NH2 NO) after reac-
ting with gaseous NO, and then decomposed to nitrogen. Since
Ce4+ NH3 is converted to Ce3+ NH4 + by WO3 , the Brønsted acid
sites of Ce/W/Ti increased more than for Ce/Ti [8]. The addition of
tungsten in the 10Ce/Ti caused the peak corresponding to NH4 +
adsorbed on the Brønsted acid sites to increase, which suggests
that NH4 + adsorbs on the Brønsted acid sites, and this observation
is strongly correlated to the catalytic activity that was also shown
in previous studies [40].
An H2 -TPR analysis was performed to investigate the presence
of the reducible species in the addition of tungsten in the 10Ce/Ti.
The H2 -TPR analyses of the 10Ce/Ti and 10Ce/W/Ti were per-
formed to distinguish the peaks of ceria and tungsten in 10Ce/W/Ti.
Fig. 12 shows the results of the H2 -TPR analysis of 10Ce/Ti and
10Ce/W/Ti. When tungsten was impregnated into 10Ce/Ti, the
reduction peaks for 10Ce/W/Ti emerged at approximately 410, 480,

Fig. 8. Ce 3d spectra of (a) 10Ce/Ti, (b) 10Ce/W/Ti, (c) 10Ce/Mo/Ti, (d) 10Ce/La/Ti,
(e) 10Ce W/Ti, and (f) W/10Ce/Ti catalysts from XPS analysis.

Fig. 10. Effect of reaction temperature on NO to NO2 conversion of 10Ce/Ti and


10Ce/W/Ti catalysts (NO: 750 ppm, NO2 : 48 ppm, O2 : 8 vol.%, H2 O: 6 vol.%, GHSV:
Fig. 9. Correlation between Ce3+ ratio and NOx conversion over various catalysts. 120,000 h−1 ).
D.W. Kwon, S.C. Hong / Applied Surface Science 356 (2015) 181–190 187

Fig. 12. H2 -TPR profiles of 10Ce/Ti and 10Ce/W/Ti catalysts (experimental condi-
tions: 30 mg catalyst, pretreatment at 400 ◦ C for 30 min with 5% O2 /Ar at 50 cm3 /min,
10% H2 /Ar reduction with a heating rate of 10 ◦ C/min, and a total flow rate of
50 cm3 /min).

515, 550–650, and 750–800 ◦ C. The peak at 410, 480, 515 ◦ C was
attributed to the reduction in the surface oxygen of stoichiometric
ceria (Ce4+ O Ce4+ ) and non-stoichiometric ceria (Ce3+ O Ce4+ )
[26,27]. The area of the peak that was attributed to WOx increased
at ∼575 and from 757 to 790 ◦ C [41,42]. With the addition of tung-
sten to 10Ce/Ti, the area of the peak at 410 ◦ C corresponding to
stoichiometric ceria (Ce4+ O Ce4+ ) decreased. In contrast, the area
of the peak at 480 and 515 ◦ C corresponding to the reduction in
the non-stoichiometric ceria (Ce3+ O Ce4+ ) increased. This means
that the Ce3+ species of the 10Ce/W/Ti catalyst was present in large
amounts.
Fig. 13 shows the catalytic performance of the 10Ce/Ti and
10Ce/W/Ti catalysts in NH3 –SCR with and without H2 O. Under
dry conditions, the catalytic activity of 10Ce/Ti and 10Ce/W/Ti
decreased at 350 ◦ C or more while it shows a tendency to increase
at 300 ◦ C or less. When moisture is present at a low temperature,
the SCR activity is known to decrease. The ammonia that reacts
with NO due to competitive adsorption of water and ammonia in
the active site is reduced on the surface of the catalyst [43]. In con-
trast, the catalytic activity and the N2 selectivity increase since the
oxidation of ammonia is inhibited [44]. In addition, the outlet NO2
concentration of 10Ce/Ti under dry conditions increases relative to
the condition where water is present (Fig. 13a). However, the out-
let NO2 concentration of the 10Ce/W/Ti catalyst does not exhibit
a significant difference depending on the presence or absence of
water (Fig. 13b). When tungsten is added to 10Ce/Ti, it is expected
to inhibit the oxidation of the ammonia at high temperatures.

3.3. Effect of GHSV and O2 concentration on NO reduction by NH3


over Ce/W/Ti catalyst

The space velocity is a very important parameter for useful


applications, and the influence of the GHSV on the catalytic activ-
Fig. 11. DRIFT spectra of (a) 10Ce/Ti and 10Ce/W/Ti exposed to 1,000 ppm NH3 for ity was studied by varying the GHSV over the 10Ce/W/Ti catalyst.
30 min at 250 ◦ C, (b) taken at 250 ◦ C upon passing 1,000 ppm NO + 3% O2 over the The activity of the 10Ce/W/Ti catalyst was measured over a wide
NH3 presorbed on 10Ce/Ti for 0, 1, 2, 5, 10, 15, and 20 min, (c) taken at 250 ◦ C upon GHSV range of 60,000–240,000 h−1 , and the results are shown in
passing 1,000 ppm NO + 3% O2 over the NH3 presorbed on 10Ce/W/Ti for 0, 1, 2, 5,
Fig. 14. The GHSV of 60,000 h−1 exhibited excellent catalytic NOx
10, 15, and 20 min.
conversion activity of 74% or higher at 270–400 ◦ C. For a GHSV of
120,000 h−1 , 98% NO conversion was obtained at 350 ◦ C. Even with
a GHSV as high as 240,000 h−1 , the maximum NO conversion was
of 90% at 350 ◦ C. These results show that the 10Ce/W/Ti catalyst is
highly effective for NOx removal within a wide GHSV range.
188 D.W. Kwon, S.C. Hong / Applied Surface Science 356 (2015) 181–190

Fig. 15. Effect of O2 concentration on NOx conversion of 10Ce/W/Ti catalyst (NO:


750 ppm, NO2 : 48 ppm, NH3 /NOx: 1.0, O2 : 8 vol.%, H2 O: 6 vol.%, GHSV: 120,000 h−1 ).

Fig. 13. Effect of reaction temperature on NOx conversion according to the with and
without H2 O of (a) 10Ce/Ti and (b) 10Ce/W/Ti catalysts (NO: 750 ppm, NO2 : 48 ppm,
O2 : 8 vol.%, H2 O: 6 vol.%, GHSV: 120,000 h−1 ).

Fig. 16. Relative SCR activity in the presence of SO2 of NO by NH3 over 10Ce/Ti
and 10Ce/W/Ti catalysts at 270 ◦ C (NO: 750 ppm, NO2 : 48 ppm, NH3 /NOx: 1.0, O2 :
8 vol.%, H2 O: 6 vol.%, SO2 : 200 ppm, GHSV: 120,000 h−1 ).

Oxygen plays an important role in the SCR reaction. In order


to investigate the effect of the gaseous oxygen concentration on
the catalytic activity in the SCR reaction of the 10Ce/W/Ti catalyst,
an experiment was performed with an O2 concentration of 3–15%,
and the results are shown in Fig. 15. The increase in the O2 concen-
tration increases the catalytic activity at 300 ◦ C or less. The partial
oxygen pressure was related to the diffusion of oxygen, and when
you consider that the lattice oxygen is consumed in the SCR reac-
tion and that gaseous O2 fills the lattice oxygen, the redox capacity
increases since the transfer of oxygen is facilitated [45].

3.4. SO2 resistance study

The influence that the addition of SO2 had on the activity of


Ce/Ti and 10Ce/W/Ti was monitored as a function of the time on
the stream during SCR of NO with NH3 . The reaction conditions
were set to 750 ppm NO, 48 ppm NO2 , NH3 /NOx = 1.0, 3 vol.% O2 ,
6 vol.% H2 O, 200 ppm SO2 , GHSV = 120,000 h−1 , and a total flow rate
Fig. 14. Effect of space velocity on NOx conversion of 10Ce/W/Ti catalyst of 600 cc/min at 270 ◦ C. Fig. 16 shows the NOx conversion when SO2
(NO: 750 ppm, NO2 : 48 ppm, NH3 /NOx: 1.0, O2 : 8 vol.%, H2 O: 6 vol.%, GHSV:
was added as a function of the reaction time. The 10Ce/Ti catalyst
60,000–240,000 h−1 ).
showed more signs of deactivation when SO2 was added than did
the 10Ce/W/Ti catalyst. The activity of 10Ce/Ti rapidly decreased
D.W. Kwon, S.C. Hong / Applied Surface Science 356 (2015) 181–190 189

to the SCR activity of the catalyst. Due to the addition of tungsten


in the 10Ce/Ti, the peak that corresponded to NH4 + adsorbed on
the Brønsted acid sites tended to increase. After NO + O2 had been
injected, as compared to 10Ce/Ti, the peaks assigned to adsorbed
ammonia species decreased more quickly on the 10Ce/W/Ti cata-
lyst. An excellent activity could be achieved by increasing the Ce3+
ratios. The 10Ce/W/Ti catalyst exhibited an excellent SO2 resistance
relative to 10Ce/Ti, and thus, the addition of tungsten to 10Ce/Ti
effectively increased the Ce3+ ratios and could be associated with
excellent NOx conversion and SO2 resistance.

References

[1] P. Granger, V.I. Parvulescu, Catalytic NOx abatement systems for mobile
sources: from three-way to lean burn after-treatment technologies, Chem.
Rev. 111 (2011) 3155–3207.
[2] Z. Liu, S.I. Woo, Recent advances in catalytic DeNOx science and technology,
Catal. Rev. Sci. Eng. 48 (2006) 43–89.
[3] P. Forzatti, Present status and perspectives in de-NOx SCR catalysis, Appl.
Catal. A: Gen. 222 (2001) 221–236.
[4] P. Forzatti, L. Lietti, Recent advances in De-NOxing catalysis for stationary
applications, Heterogeneous Chem. Rev. 3 (1996) 33–51.
Fig. 17. Transmission IR spectra of (a) 10Ce/Ti and (b) 10Ce/W/Ti samples after 5 h [5] T. Gu, Y. Liu, X. Weng, H. Wang, Z. Wu, The enhanced performance of ceria
deactivation with SO2 . with surface sulfation for selective catalytic reduction of NO by NH3 , Catal.
Commun. 12 (2010) 310–313.
[6] M. Yates, J.A. Martin, M. Martin-Luengo, S. Suarez, J. Blanco, N2 O formation in
immediately after the addition of SO2 , and in contrast, the activity the ammonia oxidation and in the SCR process with V2 O5 –WO3 catalysts,
of 10Ce/W/Ti decreased gradually. The activity of 10Ce/Ti was less Catal. Today 107–108 (2005) 120–125.
[7] X. Gao, Y. Jiang, Y. Zhong, Z. Luo, K. Cen, The activity and characterization of
than 38%, and the activity of 10Ce/W/Ti was 66% at a reaction time
CeO2 –TiO2 catalysts prepared by the sol–gel method for selective catalytic
of 5 h. The decline in the catalyst activity upon the addition of SO2 reduction of NO with NH3 , J. Hazard. Mater. 174 (2010) 734–739.
was a result of the blocking of the active sites, which is caused [8] L. Chen, J.H. Li, M.F. Ge, Enhanced activity of tungsten modified CeO2 /TiO2 for
by the formation of metal sulfates and ammonium sulfates [46]. selective catalytic reduction of NOx with ammonia, Catal. Today 153 (2010)
77–83.
The sulfated species that formed on the surface of the Ce-based [9] R. Qu, X. Gao, K. Cen, J. Li, Relationship between structure and performance of
composite oxide catalysts are stable and resulted in irreversible a novel cerium-niobium binary oxide catalyst for selective catalytic reduction
damage. of NO with NH3 , Appl. Catal. B: Environ. 142–143 (2013) 290–297.
[10] W. Shan, F. Liu, H. He, X. Shi, C. Zhang, An environmentally-benign CeO2 –TiO2
The transmission IR spectra of the 10Ce/Ti and 10Ce/W/Ti sam- catalyst for the selective catalytic reduction of NOx with NH3 in simulated
ples after 5 h of deactivation with SO2 were determined (Fig. 17). diesel exhaust, Catal. Today 184 (2012) 160–165.
For samples exposed to SO2 , several bands were detected at 982, [11] W.Q. Xu, Y.B. Yu, C.B. Zhang, H. He, Selective catalytic reduction of NO by NH3
over a Ce/TiO2 catalyst, Catal. Commun. 9 (2008) 1453–1457.
1128, 1196, 1414, and 1631 cm−1 , and at a wide band from 3400 [12] Y.S. Shen, S.M. Zhu, T. Qiu, S.B. Shen, A novel catalyst of CeO2 /Al2 O3 for
to 3500 cm−1 . According to the literature [47,48], strong bands at selective catalytic reduction of NO by NH3 , Catal. Commun. 11 (2009) 20–23.
1196 and 1128 cm−1 are caused by bulk sulfate on the CeO2 cat- [13] Y. Peng, J. Li, L. Chen, J. Chen, J. Han, H. Zhang, W. Han, Alkali metal poisoning
of a CeO2 –WO3 catalyst used in the selective catalytic reduction of NOx with
alyst, and bands at 982 and 1414 cm−1 are due to surface sulfate
NH3 : an experimental and theoretical study, Environ. Sci. Technol. 46 (2012)
species with only one S O double bond and three S O bonds with 2864–2869.
O atoms linked to the surface. The broad bands between 3400 and [14] C. Liang, L. Junhua, G. Maofa, M. Lei, C. Huazhen, Mechanism of selective
3500 cm−1 and the band at 1631 cm−1 correspond to H2 O [49]. It catalytic reduction of NOx with NH3 over CeO2 –WO3 catalysts, Chin. J. Catal.
32 (2011) 836–841.
can thus be concluded that both the surface and bulk sulfate species [15] G. Qi, R.T. Yang, R. Chang, MnOx–CeO2 mixed oxides prepared by
formed on the samples exposed to SO2 , and in particular, the peak of co-precipitation for selective catalytic reduction of NO with NH3 at low
the 10Ce/W/Ti catalyst exposed to SO2 was smaller than the 10Ce/Ti temperatures, Appl. Catal. B: Environ. 51 (2004) 93–106.
[16] Z. Wang, Z. Qu, X. Quan, H. Wang, Selective catalytic oxidation of ammonia to
catalyst exposed to SO2 . In addition, the amount of sulfur of the nitrogen over ceria–zirconia mixed oxides, Appl. Catal. A: Gen. 411–412
10Ce/Ti and 10Ce/W/Ti after 5 h deactivation with SO2 was mea- (2012) 131–138.
sured via inductively coupled plasma optic emission spectroscopy [17] C. Liu, L. Chen, J. Li, L. Ma, H. Arandiyan, Y. Du, J. Xu, J. Hao, Enhancement of
activity and sulfur resistance of CeO2 supported on TiO2 –SiO2 for the selective
(ICP). The sulfur for the 10Ce/Ti and 10Ce/W/Ti was 1.9 and 1.4 wt.%, catalytic reduction of NO by NH3 , Environ. Sci. Technol. 46 (2012)
respectively. The results of the IR pattern and ICP analysis of the cat- 6182–6189.
alysts exposed to SO2 for the same time revealed fewer sulfur from [18] H. Chang, J. Li, J. Yuan, L. Chen, Y. Dai, H. Arandiyan, J. Xu, J. Hao, Ge, Mn-doped
CeO2 –WO3 catalysts for NH3 –SCR of NOx: effects of SO2 and H2 regeneration,
10Ce/W/Ti catalyst than from the 10Ce/Ti catalyst. The 10Ce/W/Ti Catal. Today 201 (2013) 139–144.
catalyst showed a stronger resistance to SO2 deactivation than the [19] G. Zhou, B. Zhong, W. Wang, X. Guan, B. Huang, D. Ye, H. Wu, In situ DRIFTS
10Ce/Ti catalyst, and thus, it was determined that the addition of study of NO reduction by NH over Fe–Ce–Mn/ZSM-5 catalysts, Catal. Today
175 (2011) 157–163.
tungsten to 10Ce/W/Ti inhibited the formation of sulfate species.
[20] X. Gao, X. Du, L. Cui, Y. Fu, Z. Luo, K. Cen, A Ce–Cu–Ti oxide catalyst for the
selective catalytic reduction of NO with NH3 , Catal. Commun. 12 (2010)
4. Conclusions 255–258.
[21] X. Du, X. Gao, L. Cui, Z. Zheng, P. Ji, Z. Luo, K. Cen, Experimental and theoretical
studies on the influence of water vapor on the performance of a Ce–Cu–Ti
We examined the effects that the physicochemical properties oxide SCR catalyst, Appl. Surf. Sci. 270 (2013) 370–376.
of the Ce/Me/Ti catalysts had on the SCR activity after the addition [22] W. Shan, F. Liu, H. He, X. Shi, C. Zhang, A superior Ce–W–Ti mixed oxide
catalyst for the selective catalytic reduction of NOx with NH3 , Appl. Catal. B:
of various metals (W, Mo, and La) to the catalysts in order to pro- Environ. 115–116 (2012) 100–106.
mote the NH3 –SCR activity of non-vanadium-based catalysts. The [23] Z. Liu, S. Zhang, J. Li, L. Ma, Promoting effect of MoO3 on the NOx reduction by
activity in the SCR reaction significantly increased for 10Ce/W/Ti. NH3 over CeO2 /TiO2 catalyst studied with in situ DRIFTS, Appl. Catal. B:
Environ. 144 (2014) 90–95.
Different surface areas and O␣ ratios could be observed as metals [24] L. Zhang, F. Liu, Y. Yu, Y. Liu, C. Zhang, H. He, Effects of adding CeO2 to
were added to Ce/Ti, and this did not have a direct correlation with Ag/Al2 O3 catalyst for ammonia oxidation at low temperature, Chin. J. Catal. 32
the catalytic activity. The Ce3+ ratio of 10Ce/Me/Ti was proportional (2011) 727–735.
190 D.W. Kwon, S.C. Hong / Applied Surface Science 356 (2015) 181–190

[25] I. Atribak, B. Azambre, A. Bueno Lopez, A. Garcia-Garcia, Effect of NOx [37] Z.C. Si, D. Weng, X.D. Wu, J. Li, G. Li, Structure, acidity and activity of
adsorption/desorption over ceria–zirconia catalysts on the catalytic CuOx/WOx–ZrO2 catalyst for selective catalytic reduction of NO by NH3 , J.
combustion of model soot, Appl. Catal. B: Environ. 92 (2009) 126–137. Catal. 271 (2010) 43–51.
[26] S. Damyanova, C.A. Perez, M. Schmal, J.M.C. Bueno, Characterization of [38] J. Laane, J.R. Ohlsen, Characterization of nitrogen oxides by vibrational
ceria-coated alumina carrier, Appl. Catal. A: Gen. 234 (2002) spectroscopy, Prog. Inorg. Chem. 27 (1980) 465–513.
271–280. [39] L. Chen, J. Li, M. Ge, DRIFT study on cerium–tungsten/titiania catalyst for
[27] B. Murugan, A.V. Ramaswamy, Chemical states and redox properties of selective catalytic reduction of NOx with NH3 , Environ. Sci. Technol. 44 (2010)
Mn/CeO2 –TiO2 nanocomposites prepared by solution combustion route, J. 9590–9596.
Phys. Chem. C 112 (2008) 20429–20442. [40] F. Tang, B. Xu, H. Shi, J. Qiu, Y. Fan, The poisoning effect of Na+ and Ca2+ ions
[28] P. Fornasiero, R. Dimonte, G.R. Rao, J. Kaspar, S. Meriani, A. Trovarelli, M. doped on the V2 O5 /TiO2 catalysts for selective catalytic reduction of NO by
Graziani, Rh-loaded CeO2 –ZrO2 solid-solutions as highly efficient oxygen NH3 , Appl. Catal. B: Environ. 94 (2010) 71–76.
exchangers: dependence of the reduction behavior and the oxygen storage [41] J. Engweiler, J. Harf, A. Baiker, WOx/TiO2 catalysts prepared by grafting of
capacity on the structural-properties, J. Catal 151 (1995) 168–177. tungsten alkoxides: morphological properties and catalytic behavior in the
[29] J. Fang, X. Bi, D. Si, Z. Jiang, W. Huang, Spectroscopic studies of interfacial selective reduction of NO by NH3 , J. Catal. 159 (1996) 259–269.
structures of CeO2 –TiO2 mixed oxides, Appl. Surf. Sci. 253 (2007) [42] D.W. Kwon, K.B. Nam, S.C. Hong, Influence of tungsten on the activity of a
8952–8961. Mn/Ce/W/Ti catalyst for the selective catalytic reduction of NO with NH3 at
[30] Z. Wu, R. Jin, Y. Liu, H. Wang, Ceria modified MnOx/TiO2 as a superior catalyst low temperatures, Appl. Catal. A: Gen. 497 (2015) 160–166.
for NO reduction with NH3 at low-temperature, Catal. Commun. 9 (2008) [43] F. Liu, H. He, Selective catalytic reduction of NO with NH3 over manganese
2217–2220. substituted iron titanate catalyst: reaction mechanism and H2 O/SO2
[31] L. Chen, J. Li, M. Ge, Promotional effect of Ce-doped V2 O5 –WO3 /TiO2 with low inhibition, Catal. Today 153 (2010) 70–76.
vanadium loadings for selective catalytic reduction of NOx by NH3 , J. Phys. [44] V. Turco, L. Lisi, R. Pirone, P. Ciambelli, Effect of water on the kinetics of nitric
Chem. C 113 (2009) 21177–21184. oxide reduction over high-surface-area V2 O5 /TiO2 catalyst, Appl. Catal. B:
[32] W. Shan, F. Liu, H. He, X. Shi, C. Zhang, The remarkable improvement of a Environ. 3 (1994) 133–149.
Ce–Ti based catalyst for NOx abatement, prepared by a homogeneous [45] L. Lietti, G. Ramis, F. Berti, G. Toledo, D. Robba, G. Busca, P. Forzatti, Chemical,
precipitation method, ChemCatChem 3 (2011) 1286–1289. structural and mechanistic aspects on NOx SCR over commercial and model
[33] W.P. Shan, F.D. Liu, H. He, X.Y. Shi, C.B. Zhang, Novel cerium–tungsten mixed oxide catalysts, Catal. Today 42 (1998) 101–116.
oxide catalyst for the selective catalytic reduction of NOx with NH3 , Chem. [46] J.H. Li, H.Z. Chang, L. Ma, J.M. Hao, R.T. Yang, Low-temperature selective
Commun. 47 (2011) 8046–8048. catalytic reduction of NOx with NH3 over metal oxide and zeolite catalysts – a
[34] N.Y. Topsoe, J.A. Dumesic, H. Topsoe, Vanadia–titania catalysts for selective review, Catal. Today 175 (2011) 147–156.
catalytic reduction of nitric-oxide by ammonia. I.I. Studies of active sites and [47] M. Waqif, P. Bazin, O. Saur, J.C. Lavalley, G. Blanchard, O. Touret, Study of ceria
formulation of catalytic cycles, J. Catal. 151 (1995) 241–252. sulfation, Appl. Catal. B: Environ. 11 (1997) 193–205.
[35] G. Ramis, G. Busca, F. Bregani, On the effect of dopants and additives on the [48] D.W. Kwon, K.B. Nam, S.C. Hong, The role of ceria on the activity and SO2
state of surface vanadyl centers of vanadia-titania catalysts, Catal. Lett. 18 resistance of catalysts for the selective catalytic reduction of NOx by NH3 ,
(1993) 299–303. Appl. Catal. B: Environ. 166–167 (2015) 37–44.
[36] G. Ramis, M.A. Larrubia, An FT-IR study of the adsorption and oxidation of [49] S. Lucas, E. Champion, D. Bregiroux, D. Bernache-Assollant, F. Audubert, Rare
N-containing compounds over Fe2 O3 /Al2 O3 SCR catalysts, J. Mol. Catal. A 215 earth phosphate powders RePO4 ·nH2 O (Re = La, Ce or Y). Part I. Synthesis and
(2004) 161–167. characterization, J. Solid State Chem. 177 (2004) 1302–1311.

You might also like