Sunteți pe pagina 1din 9

Journal of Molecular Catalysis A: Chemical 424 (2016) 232240

Contents lists available at ScienceDirect

Journal of Molecular Catalysis A: Chemical


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

Performance of Fe-ZSM-5 for selective catalytic reduction of NOx with


NH3 : Effect of the atmosphere during the preparation of catalysts
Shuangshuang Lai, Yong She, Wangcheng Zhan , Yun Guo, Yanglong Guo, Li Wang,
Guanzhong Lu
Key Laboratory for Advanced Materials and Research Institute of Industrial Catalysis, East China University of Science and Technology, Shanghai, 200237,
PR China

a r t i c l e

i n f o

Article history:
Received 20 March 2015
Received in revised form 26 August 2016
Accepted 27 August 2016
Available online 28 August 2016
Keywords:
Fe-ZSM-5
Selective catalytic reduction
Nitrogen oxide
Preparation of catalyst
Atmosphere effect

a b s t r a c t
Fe-ZSM-5 catalysts for selective catalytic reduction (SCR) of NOx with ammonia were prepared by ion
exchange under different atmospheres and characterized by X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), Temperature-programmed desorption of NH3 and NO (NH3 -TPD and NO-TPD),
Ultravioletvisible (UVvis) spectroscopy and in situ diffuse reectance Fourier transform infrared spectroscopy (DRIFT). The results show that the atmosphere during ion exchange and calcination signicantly
affects the cation exchange capacity and the nature of the Fe species in Fe-ZSM-5 catalysts. An air atmosphere was conducive to improving the ability of ZSM-5 zeolites for Fe ionic exchange. However, isolated
Fe3+ sites were dominant in Fe-ZSM-5 catalyst (Fe-Z(N2 )) prepared under a N2 atmosphere, and small
oligomeric Fex Oy clusters and Fex Oy nano-particles were markedly increased in Fe-ZSM-5 catalyst prepared in air, resulting in the higher SCR activity and wide temperature window of Fe-Z(N2 ). Moreover,
Fe-Z(N2 ) catalyst possessed more Brnsted acid sites and could adsorb more NH3 species, promoting its
SCR activity. In situ DRIFT studies showed that NH3 can be adsorbed on both Brnsted and Lewis acid
sites on Fe-ZSM-5 catalyst, and this adsorbed NH3 can react with gaseous NOx . Additionally, NOx can be
adsorbed on Fe-ZSM-5 catalyst, mainly as a bidentate nitrate, but these nitrate species cannot be conrmed to react with gaseous NH3 because of the overlap with the strong absorption band of adsorbed NH3
species. Based on the IR results and weak adsorption of NOx over Fe-ZSM-5 catalysts, it was concluded
that NH3 -SCR over this Fe-ZSM-5 catalyst predominantly follows the Eley-Rideal reaction mechanism.
2016 Elsevier B.V. All rights reserved.

1. Introduction
Nitrogen oxides (NOx ) and particulate matter (PM) emitted from
diesel vehicles are major atmospheric pollutants. Selective catalytic
reduction of NOx with NH3 (NH3 -SCR) is one of the most promising
techniques for NOx elimination from diesel exhaust under oxygenrich conditions. A great variety of catalysts have been developed
for this process, such as V2 O5 /TiO2 [1,2], zeolite-based catalysts
[3,4], mesoporous materials-based catalysts [5,6] and oxide catalysts [79]. Among them, V2 O5 /TiO2 catalysts possess excellent
H2 O/SO2 durability and are the most widely used in industry. However, V2 O5 /TiO2 catalysts exhibit some inevitable disadvantages,
such as a narrow temperature window and volatility of the active
component (V2 O5 ) at high temperatures. The latter is likely harmful for the ecology and human health. Therefore, the development

Corresponding authors.
E-mail addresses: zhanwc@eust.edu.cn (W. Zhan), gzhlu@ecust.edu.cn (G. Lu).
http://dx.doi.org/10.1016/j.molcata.2016.08.026
1381-1169/ 2016 Elsevier B.V. All rights reserved.

of novel, highly efcient, stable and environmentally friendly NH3 SCR catalysts has attracted extensive attention.
Among different NH3 -SCR catalysts, zeolite-based catalysts
generally have high activity and selectivity over a wider temperature range, as well as good thermal stability, which has aroused
wide interest in researchers developing zeolite-based catalysts. At
present, the zeolites ZSM-5 [3,10,11], [12,13], USY [14,15] and
SSZ-13 [1618] are mainly used as matrices for NH3 -SCR catalysts. Transition metals and rare earth metals are usually used
as the active components of zeolite-based SCR catalysts, e.g., Cu
[11,16,19], Fe [3,10,20], Mn [21] and Ce [22]. On the whole, Cu- and
Fe-modied ZSM-5 are the most widely studied among zeolitebased SCR catalysts. In general, Cu-modied ZSM-5 exhibits high
SCR activity, particularly at low reaction temperatures, while Femodied ZSM-5 exhibits high SCR activity at high temperatures
with a wide temperature window.
Several studies have focused on the effect of different preparation methods, such as chemical vapour deposition [23], hydrothermal synthesis [24], impregnation [25,26] and liquid/solid-state ion

S. Lai et al. / Journal of Molecular Catalysis A: Chemical 424 (2016) 232240

exchange [3,10,27], on the SCR activity of Fe-modied ZSM-5 catalysts. It was proposed that preparation methods have a marked
effect on the location of Fe in the ZSM-5 matrix, leading to different activities for NH3 -SCR [28]. Among the preparation methods
mentioned above, liquid ion exchange is most commonly used
for preparing metal-exchanged zeolites [3,26,29,30]. However, it is
well known that synthesis conditions generally have a signicant
inuence on the structure, physico-chemical properties and reactive activity of Fe-modied ZSM-5 catalysts. For example, thermal
treatment signicantly affected the catalytic activity of Fe-ZSM-5
catalysts prepared by chemical vapour deposition [3133]. Fe-ZSM5 catalysts exhibited a slightly higher activity when a slow heating
rate (0.5 C/min) was applied during calcination [31]. When the
thermal treatment was carried out in helium rather than in oxygen,
partial auto-reduction of iron from Fe3+ to Fe2+ occurred together
with the formation of Fe3 O4 clusters [32]. In the liquid ion exchange
method, only Fe precursors have been investigated to study their
effect on the activity and hydrothermal stability of NH3 -SCR in
Fe-ZSM-5 [34,35]. Other synthesis conditions, such as the atmosphere during ion exchange and calcination processes, have not
attracted attention. To the best of our knowledge, related work
has only reported by Brandenberger and co-authors. They synthesized Fe-ZSM-5 samples by liquid ion exchange of NH4 -ZSM-5 with
FeCl2 4H2 O under nitrogen instead of air, followed by calcination of
the obtained sample under nitrogen [36]. Unfortunately, although
a more detailed picture of the role of Brnsted acidity in the SCR of
NO by ammonia in Fe-ZSM-5 catalysts has been obtained, the role
of the N2 atmosphere in the preparation of Fe-ZSM-5 samples has
not been described. As mentioned above, although some effort has
been made to study the effect of the calcination process on the SCR
activity of Fe-modied ZSM-5 catalysts, it is very necessary to thoroughly study the nature of the Fe species and NH3 -SCR activity of
Fe-ZSM-5 catalyst as a function of preparation parameters of liquid
ion exchange.
Herein, Fe-ZSM-5 zeolite catalysts were prepared by liquid ion
exchange under atmospheres of air and N2 , respectively, and their
physicochemical and catalytic properties for the SCR reaction with
ammonia were investigated. It is interesting that different atmospheres during ion exchange and calcination processes can affect
the degree of ionic exchange, the nature of the Fe species and the
SCR activity of Fe-ZSM-5 catalysts. At the same time, the NH3 -SCR
reaction pathway over prepared Fe-ZSM-5 catalysts was explored
and discussed here on the basis of DRIFTS studies.

233

RIGAKU). The Fe content in the samples was analysed using an


inductively coupled plasma optical emission spectrometer (ICPOES, Thermo Elemental, IRIS 1000). The N2 adsorption-desorption
isotherms were measured on a Quantachrome NOVA1200 surface
area analyser at 196 C. Prior to the measurements, all samples
were degassed at 180 C until a stable vacuum of approximately
5 mTorr was reached. The specic surface area of the sample
was calculated using the Brunauer-Emmett-Teller (BET) method.
Ultravioletvisible (UVvis) spectra were recorded on a Varian Cary
500 UVvis-NIR spectrophotometer at 200800 nm with BaSO4 as
a reference, and the spectra were converted with the KubelkaMunk (K-M) function F(R) for comparison. X-ray photoelectron
spectroscopy (XPS) was conducted on a Thermo Fisher ESCALAB
250Xi spectrometer with Mg K radiation at room temperature
under 5.0 1010 mbar. Fe 2p binding energy was calibrated with
a C1s band at 285.4 eV from carbon impurity. Peak tting was performed using XPSPEAK 4.1 with a linear background and 20:80
Lorentzian:Gaussian convolution product shapes.
Temperature-programmed desorption of NH3 adsorbed on the
catalyst (NH3 -TPD) was carried out in a conventional ow system
equipped with a thermal conductivity detector (TCD). After 50 mg
of the sample was pre-treated at 600 C for 30 min and cooled to
room temperature in N2 , it was allowed to adsorb NH3 for 1 h from
the mixed gas 10% NH3 /N2 . After purging the sample in N2 for 1.5 h
at 90 C, NH3 -TPD was carried out at 90600 C in a ow of N2
(50 mL/min). The rate of heating was 10 C/min.
NO-TPD was carried out in a conventional ow system equipped
with a NOx analyser (Thermo Fisher Model 42i-HL NO-NOx chemiluminescence analyser) as the detector. 50 mg of the sample
was pre-treated in Ar (450 mL/min) at 500 C for 1 h and cooled
to room temperature. Then, the sample was exposed to a ow of
500 ppm NO/Ar (300 mL/min) for 1 h. After purging the sample in
Ar (300 mL/min) for 1 h, NO-TPD was carried out at RT 500 C in a
ow of Ar (300 mL/min). The rate of heating was 10 C/min.
In situ diffuse reectance infrared Fourier transform spectroscopy (DRIFTS) was conducted on a Nicolet NEXUS 670-FTIR
spectrometer equipped with a MCT detector. The diffuse
reectance FT-IR experiments were carried out in situ in a hightemperature IR cell. Prior to each experiment, the sample was
pre-treated in situ in the IR cell at 550 C in a ow of Ar (50 mL/min)
for 60 min and then cooled to 50 C. The background spectrum was
recorded in owing Ar and was subtracted from the sample spectrum obtained. The IR spectra were recorded by accumulating 128
scans at a resolution of 4 cm1 .

2. Experimental
2.1. Preparation of catalysts
Fe-ZSM-5 samples were prepared by ion exchange method. Typically, 3 g of NH4 -ZSM-5 (purchased from Nankai University, Si/Al
atomic ratio = 25) was added to 300 mL of 0.05 M FeSO4 7H2 O aqueous solution. The pH value of the synthesis solution was adjusted
to approximately 3.2 using dilute H2 SO4 , and the suspension was
stirred under continuous ow of N2 at 80 C for a certain time.
Then, the solid was ltered and washed with deionized water, dried
overnight at 120 C and nally calcined at 600 C for 5 h in N2 . The
obtained samples were labelled as Fe-Z(N2 -x), where x is the time of
ion exchange. The samples exposed to the air atmosphere during
ion exchange and calcination were labelled as Fe-Z(Air-x). Additionally, if NH4 -ZSM-5 was calcined at 600 C for 5 h in air, the
obtained sample was labelled as H-ZSM-5.

2.3. Catalytic activity testing


The SCR activity of the catalyst was measured in a xed-bed
quartz reactor. Fe-ZSM-5 catalysts were powdered and sieved to
2040 mesh. 200 mg of Fe-ZSM-5 catalysts was used in each run.
The reactant gas consisted of 500 ppm NO, 500 ppm NH3 , 5%O2 and
balance Ar. The total ow rate was 300 mL/min (under ambient condition) and the gas hourly space velocity (GHSV) was 58,000 h1 .
The concentrations of NO and NO2 were continually monitored by
a chemiluminescent NO/NOx analyser (Thermo-Scientic, Model
42i-HL). NOx conversion (XNOx ) was calculated as follows:
XNOx = (1-([NO] + [NO2 ])out /([NO] + [NO2 ])in ) 100%

3. Results and discussion

2.2. Catalyst characterization

3.1. Catalytic performance

Powder X-ray diffraction (XRD) patterns of samples were


acquired on a D/MAX 2550 VB/PC X-ray diffractometer (Japan,

Fig. 1 shows the NOx conversions of the NH3 -SCR reaction as


a function of reaction temperature over different Fe-ZSM-5 cata-

234

S. Lai et al. / Journal of Molecular Catalysis A: Chemical 424 (2016) 232240

Fe-Z(N2-60)

80
60
ZSM-5
Fe-Z(N2-12)
Fe-Z(N2-24)
Fe-Z(N2-36)
Fe-Z(N2-60)
Fe-Z(air-2)
Fe-Z(air-6)

40
20

Fe-Z(N2-36)

Intensity (a.u.)

NOx Conversion (%)

100

Fe-Z(Air-6)
Fe-Z(Air-2)
ZSM-5

0
100

200

300

400

500

600

700

10

20

30

Temperature ( C)

40

50

60

2 Theta ( )
o

Fig. 1. Catalytic performance of Fe-ZSM-5 and H-ZSM-5 catalysts for NH3 -SCR.
Fig. 2. XRD patterns of ZSM-5, Fe-Z(N2 -60), Fe-Z(N2 -36), Fe-Z(Air-6) and Fe-Z(Air-2)
catalysts.

3.2. Structural characterization of catalysts


3.2.1. ICP
The Fe contents of the samples were determined by inductively
coupled plasma (ICP) analysis and are shown in Table 1. The Fe content in the Fe-Z(N2 -x) samples increased with the increase in ion
exchange time. For example, under an atmosphere of N2 during ion
exchange, the Fe content of the Fe-Z(N2 -x) samples increased from
0.9 wt% to 1.9 wt% when the time for ion exchange was increased
from 12 h to 60 h. Similarly, under an atmosphere of air during ion
exchange, the Fe contents of the Fe-Z(Air-2) and Fe-Z(Air-6) samples were 1.4 wt% and 1.8 wt%, respectively, after 2 h and 6 h of ion
exchange. Conversely, to load similar Fe contents, the time of ion

724.3

710.6
718.6

Fe-Z(N2-36)

Intensity (a.u.)

lysts. The NOx conversion over the H-ZSM-5 catalyst was very low
(no more than 50%) at <300 C. Only when the reaction temperature was raised >525 C did the NOx conversion over the H-ZSM-5
catalyst increase to 95%. Compared with the H-ZSM-5 catalyst, the
SCR activity was markedly improved after introducing Fe into the
H-ZSM-5 catalyst. As shown in Fig. 1, the NOx conversion over
the Fe-Z(N2 -12) catalyst was >95% over a wide temperature window (320575 C). Furthermore, the SCR activities of the Fe-Z(N2 -x)
catalysts at low temperature increased with the increase in Fe content, while its activities at high temperature decreased slightly. In
other words, more than 90% NOx conversion was still obtained over
a wide temperature window over different Fe-Z(N2 -x) catalysts.
The SCR activity of the Fe-Z(Air-x) catalyst was lower than that of
the Fe-Z(N2 -x) catalyst with equivalent Fe contents, especially at
high reaction temperatures. Over the Fe-Z(Air-2) catalyst, the NOx
conversion sharply decreased at >455 C, showing a narrow temperature window. When the Fe content was continually increased
to 1.8 wt%, there was a further decrease in the SCR activity of the
Fe-Z(Air-6) catalyst at high reaction temperatures.
The turnover frequency (TOF) values of different catalysts for
the SCR reaction were estimated at 100 C on the basis of their BET
surface area (Table 2), and the GHSV was changed to keep the NOx
conversion less than 20% over all catalysts. The TOF values were
4.4 107 , 3.6 107 , 1.5 107 and 1.8 107 mol h1 m2 g1 for
Fe-Z(N2 -60), Fe-Z(N2 -36), Fe-Z(Air-6) and Fe-Z(Air-2) catalysts,
respectively, which clearly shows that a N2 atmosphere during the
ion exchange process can effectively improve the catalytic performance of Fe-ZSM-5 catalyst. Since an equivalent Fe content was
incorporated in the corresponding catalysts, such as Fe-Z(N2 -36)
and Fe-Z(Air-2), the difference of NH3 -SCR activity between FeZ(N2 -x) and Fe-Z(Air-x) catalysts could be attributed to the different
nature of Fe species in Fe-ZSM-5 catalysts. This is discussed in further detail below.

Fe-Z(Air-2)
735

730

725

720

715

710

705

700

Binding energy (eV)


Fig. 3. XPS spectra of Fe-Z(N2 -36) and Fe-Z(Air-2) catalysts.

exchange under an air atmosphere is much shorter than that under


a N2 atmosphere, indicating that the capacity for Fe cation exchange
under an air atmosphere is better than that under a N2 atmosphere.
3.2.2. XRD
The XRD patterns of Fe-Z(N2 -60), Fe-Z(N2 -36), Fe-Z(Air-6) and
Fe-Z(Air-2) catalysts are shown in Fig. 2. All Fe-ZSM-5 catalysts
exhibited typical diffraction peaks of the ZSM-5 zeolite, indicating that the structure of the ZSM-5 zeolite was retained after Fe
exchange. No diffraction peaks belonging to Fe2 O3 were observed at
2 = 33.1, 35.6, 40.9 and 49.5 in the XRD patterns of all the catalysts,
indicating that the Fe species are highly dispersed in all Fe-ZSM-5
catalysts.
3.2.3. XPS
Fig. 3 shows the XPS spectra of Fe 2p3/2 and Fe 2p1/2 for FeZ(N2 -36) and Fe-Z(Air-2) catalysts. Two peaks centred at 710.6 and
724.3 eV were observed in the spectra of both samples, which were
assigned to Fe 2p3/2 and 2p1/2 , respectively [37], and a satellite peak
of Fe 2p3/2 was located at 718.6 eV [38]. The peak at BE = 710.6 eV
of Fe 2p3/2 was close to the binding energy of Fe 2p3/2 of Fe2 O3
(BE = 710.4711 eV) [39,40], indicating that Fe in Fe-ZSM-5 catalysts was mainly present as the +3 form.
The deconvolution of Fe 2p3/2 peak was very difcult because
there is a large difference in the BE values of FeO and/or Fe2 O3
obtained by different researchers, and the FWHM of every peak
at deconvolution was not unambiguous. Even the deconvolution of
the Fe 2p3/2 peak for both FeO and Fe2 O3 consisted of several peaks,

S. Lai et al. / Journal of Molecular Catalysis A: Chemical 424 (2016) 232240

235

Table 1
Percentage of the area of the sub-bands and wt% Fe of the corresponding species by numerical analysis of UVvis spectra (Fig. 4) for Fe-ZSM-5 catalysts.
Catalyst

Total Fe content d (wt%)

Fe-Z(N2 -12)
Fe-Z(N2 -24)
Fe-Z(N2 -36)
Fe-Z(N2 -60)
Fe-Z(Air-2)
Fe-Z(Air-6)

0.9
1.1
1.3
1.9
1.4
1.8

a
b
c
d

Fe1 a

Fe2 b

Fe3 c

I1 (%)

(wt%)

I2 (%)

(wt%)

I3 (%)

(wt%)

78.2
58.8
57.4
35.8

1.02
1.12
0.80
0.65

17.8
25.9
24.1
28.0

0.23
0.49
0.34
0.50

4.0
15.3
18.5
36.2

0.05
0.29
0.26
0.65

Isolated Fe ion in tetrahedral and octahedral coordination.


Small oligomeric Fex Oy clusters.
Fex Oy nano-particles on zeolite surface.
detected by ICP.

Table 2
BET surface area (SBET ) and the total area of the two desorption peaks (Area) of
NH3 -TPD curves (Fig. 5A) for different ZSM-5 samples.
SBET (m2 /g)

Area

Area/SBET

ZSM-5
Fe Z(N2 -60)
Fe-Z(N2 -36)
Fe-Z(Air-6)
Fe-Z(Air-2)

334
276
275
265
272

/
2834
2664
2261
1731

/
10.3
9.7
8.5
6.4

Fe-Z(Air-6)

Fe-Z(N2-60)

F(R)

Sample

Fe-Z(Air-2)

and there was a small overlap between the deconvolution peaks


for FeO and Fe2 O3 [39,41]. To approximately compare the difference in FeO/Fe2 O3 (mol) between the Fe-Z(N2 -36) and Fe-Z(Air-2)
catalysts, the Fe 2p3/2 peaks in Fig. 3 were tted with a linear background and 20:80 Lorentzian:Gaussian convolution product shapes
[38,39,42]. The peaks at 707.5 and 714.6 eV were assigned to FeO
and Fe(OH)O, respectively, while the other ve peaks were due to
the overlap of FeO, Fe2 O3 and Fe(OH)O. Therefore, these results
clearly reveal that a small quantity of Fe2+ also existed on the surface of the Fe-Z(N2 -36) and Fe-Z(Air-2) catalysts. Conversely, the
area of every deconvolution peak for the Fe-Z(N2 -36) catalyst was
close to that of the corresponding peaks for the Fe-Z(Air-2) catalyst,
indicating that the ratio of Fe2+ /Fe3+ on Fe-Z(N2 -36) was close to
that on Fe-Z(Air-2). This is mainly due to the easy oxidation of FeO
to Fe2 O3 even at low temperature. Therefore, in the preparation
process of the Fe-Z(N2 -36) catalyst, which included steps such as
ltering under an air atmosphere, Fe2+ ions on the surface of the
Fe-Z(N2 -36) catalyst were easily oxidized to Fe3+ .
3.2.4. UVvis spectroscopy
The UVvis spectra of Fe-Z(N2 -60), Fe-Z(N2 -36), Fe-Z(Air-6) and
Fe-Z(Air-2) catalysts are shown in Fig. 4. It has been reported that
UVvis spectra can distinguish between different Fe species on FeZSM-5 catalysts to isolated Fe3+ species (bands below 300 nm),
small oligomeric Fex Oy clusters (between 300 and 400 nm) and
Fex Oy nano-particles on the zeolite surface (above 400 nm) [31,43].
The percentage areas of the sub-bands and wt% Fe of the corresponding species, obtained by numerical analysis of UVvis spectra
for Fe-ZSM-5 catalysts (Fig. 4), are shown in Table 1. I1 , I2 and I3
represent the area (%) of sub-bands at 300 nm, = 300400 nm
and > 400 nm, respectively. The results indicate that I1 for the FeZ(N2 -x) catalyst was higher than that for the Fe-Z(Air-x) catalyst
with equivalent Fe contents, whereas I2 and I3 presented a contrary result, especially I3 . For example, I1 , I2 and I3 are 78.2%, 17.8%
and 4.0% for the Fe-Z(N2 -36) catalyst with 1.3 wt% Fe content, while
they are 57.5%, 24.1% and 18.5% for the Fe-Z(Air-2) catalyst with
1.4 wt% Fe content. However, with increasing Fe content of catalysts
(1.8 wt% and 1.9 wt%), I1 decreased markedly, while I3 dramatically
increased and I2 increased slightly, regardless of the atmosphere
during the preparation process. The results mentioned above sug-

Fe-Z(N2-36)

200

300

400

500

600

700

Wavelength (nm)
Fig. 4. UVvis spectra of Fe-Z(N2 -60), Fe-Z(N2 -36), Fe-Z(Air-6) and Fe-Z(Air-2) catalysts.

gest that the atmosphere during the preparation process has a great
inuence on the nature of Fe species in Fe-ZSM-5 catalysts.
Brandenberger et al. [44] suggested that different Fe species on
Fe-ZSM-5 catalyst may act as active sites at different temperatures:
at <300 C, the SCR activity of Fe-ZSM-5 catalyst is primarily caused
by isolated iron sites; at >300 C, small oligomeric iron species are
involved in the SCR reaction; at 500 C, Fex Oy nano-particles catalyse the SCR reaction. Furthermore, surface iron particles not only
contributed to the SCR reaction but also caused nonselective oxidation of NH3 at 350 C, while oligomeric iron species also exhibited
catalytic activity for NH3 oxidation up to 500 C. As shown in Fig. 4
and Table 1, the Fe-Z(N2 -36) catalyst possessed more isolated Fe3+
sites than the Fe-Z(Air-2) catalyst, leading to its higher SCR activity
at <300 C. On the contrary, compared with the Fe-Z(N2 -36) catalyst, more Fex Oy oligomers and nano-particles on the Fe-Z(Air-2)
catalyst brought on the more serious nonselective oxidation of NH3 ,
resulting in the lower SCR activity of Fe-Z(Air-2) at high temperature (450 C). When the Fe content was increased from 1.3 wt%
for the Fe-Z(N2 -36) catalyst to 1.9 wt% for the Fe-Z(N2 -60) catalyst,
the amount of isolated Fe3+ sites and Fex Oy oligomers increased
from 1.02 wt% and 0.23 wt% to 1.12 wt% and 0.49 wt%, respectively,
leading to further increase in the SCR activity of the Fe-Z(N2 -60)
catalyst at low temperature compared with the Fe-Z(N2 -36) catalyst. At the same time, because of the increase in the amount of
Fex Oy nano-particles from 0.05 wt% to 0.29 wt%, the SCR activity
of the Fe-Z(N2 -60) catalyst at high temperature decreased slightly
compared with the Fe-Z(N2 -36) catalyst. On the contrary, for the FeZ(Air-x) catalysts, when the Fe content was increased from 1.4 wt%
for the Fe-Z(Air-2) catalyst to 1.8 wt% for the Fe-Z(Air-6) catalyst,

236

S. Lai et al. / Journal of Molecular Catalysis A: Chemical 424 (2016) 232240

A
Fe-Z(Air-6)

Fe-Z(N2-36)

0.01

Fe-Z(Air-2)

Aborbance

Intensity (a.u.)

Fe-Z(N2-60)

Fe-Z(N2-36)
3660

Fe-Z(Air-2)
3610

100

200

300

400

500

600

700

3800

3750

3700

Temperature ( C)
o

3650

3600

3550

3500

-1

Wavenumber (cm )

Fig. 5. (A) NH3 -TPD curves of Fe-Z(N2 -60), Fe-Z(N2 -36), Fe-Z(Air-6) and Fe-Z(Air-2) catalysts and (B) their DRIFT spectra after being exposed under 500 ppm NH3 /Ar
(50 mL/min) at 50 C for 90 min.

3.3. The absorption properties for NH3 and NO


Fig. 5A shows the NH3 -TPD curves of Fe-Z(N2 -60), Fe-Z(N2 36), Fe-Z(Air-6) and Fe-Z(Air-2) catalysts, in which there are two
NH3 desorption peaks. The peak located at 200 C can be ascribed
to physisorbed NH3 , and the peak at 400 C can be ascribed to
NH3 strongly adsorbed on Brnsted hydroxyl groups and strong
Lewis iron sites [45,46]. The ratios between the total area of the
two desorption peaks and the BET surface areas for different catalysts (Area/SBET ) were calculated, and the results are shown in
Table 2. It was found that Area/SBET for the Fe-Z(N2 -x) catalysts
were larger than those for the Fe-Z(Air-x) catalysts with equivalent
Fe content, indicating that Fe-Z(N2 -x) catalysts possess a higher
adsorption capacity for NH3 . Furthermore, for both Fe-Z(N2 -x) and
Fe-Z(Air-x) catalysts, Area/SBET increased with the increase in the
Fe content. The ratio between the areas of the desorption peaks
at approximately 400 C and 200 C for the Fe-Z(N2 -x) catalyst
was remarkably higher than that for the Fe-Z(Air-x) catalysts with
equivalent Fe content, indicating that the Fe-Z(N2 -x) catalysts possessed more Brnsted acid sites and Lewis iron sites that can adsorb
NH3 . This high ability of Fe-Z(N2 -x) catalysts for NH3 adsorption can
prompt their SCR activity. The detailed relationship between NH3
species adsorbed on the catalyst surface and SCR activity will be
discussed in the DRIFTS measurements section.
The relative concentration of Brnsted acid sites over Fe-ZSM-5
catalysts was also measured by FTIR spectroscopy. The sample was
pre-treated as follows: it was dried in an IR cell under a ow of
Ar at 550 C for 1 h and then cooled to 50 C; after recording the
background spectrum in Ar, the sample was exposed to a ow of
500 ppm NH3 /Ar (50 mL/min) at 50 C for 90 min, and the DRIFT
spectra of Fe-Z(N2 -36) and Fe-Z(Air-2) catalysts after NH3 adsorption are shown in Fig. 5B. When NH3 was fed into the IR cell, the
OH- groups on the catalyst surface were consumed by the interaction of NH3 with OH-, thereby creating negative peaks at the
corresponding wavenumbers. The negative peak at 3610 cm1 was
assigned to the OH stretch of the Brnsted acidic sites, and the negative peak at 3660 cm1 arose from the OH groups associated with
extra-framework Al [47,48]. Compared with the Fe-Z(Air-2) cata-

800
NOx Concentration (ppm)

the amount of isolated Fe3+ sites sharply decreased from 0.80 wt%
to 0.65 wt%, accompanied by the increase in the amount of Fex Oy
oligomers from 0.34 wt% to 0.50 wt%, leading to the same SCR activity for Fe-Z(Air-2) and Fe-Z(Air-6) catalysts at low temperatures.
However, the amount of Fex Oy nano-particles rapidly increased
from 0.26 wt% to 0.65 wt%, leading to a marked decline in the SCR
activity at high temperature over Fe-Z(Air-6) catalysts. The above
results indicate that to obtain a high SCR activity, Fe-ZSM-5 catalyst
should be designed and prepared to enhance the concentration of
isolated Fe3+ species.

600

400

Fe-Z(N2-60)

200

Fe-Z(N2-36)
Fe-Z(Air-6)
Fe-Z(Air-2)

0
0

100

200

300

400

500

Temperature ( C)
Fig. 6. TPD curves of NO adsorbed over Fe-Z(N2 -60), Fe-Z(N2 -36), Fe-Z(Air-6) and
Fe-Z(Air-2) catalysts.

lyst, the peak at 3610 cm1 for the Fe-Z(N2 -36) catalyst showed a
much higher intensity, and the peak at 3660 cm1 showed similar
intensities for the two catalysts. Since the relative concentration of
Brnsted acidic sites can be determined using the intensities of the
peaks at 3610 cm1 [48], it was clear that the Fe-Z(N2 -36) catalysts
possessed more Brnsted acid sites than the Fe-Z(Air-2) catalyst.
The NO-TPD proles of Fe-Z(N2 -60), Fe-Z(N2 -36), Fe-Z(Air-6)
and Fe-Z(Air-2) catalysts are shown in Fig. 6. As shown in Fig. 6, all
the catalysts exhibited two NO desorption peaks at approximately
126 C and 325 C. The former can be ascribed to physisorbed NO
species, and the latter can be ascribed to the decomposition of
nitrate species with higher thermal stability. Although the positions
of the two desorption peaks were very close for different catalysts,
the areas of the desorption peaks were very different. The areas
of both desorption peaks were larger for the Fe-Z(N2 -x) catalysts
than for the Fe-Z(Air-x) catalysts with equivalent BET surface areas,
indicating that the Fe-Z(N2 -x) catalysts possessed a higher capacity for NO adsorption. In other words, more nitrate species could
be adsorbed over the surface of the Fe-Z(N2 -x) catalysts, which was
benecial to the SCR activity of Fe-Z(N2 -x) catalysts. However, with
increasing content of Fe in the Fe-Z(Air-x) catalysts, the areas of
both desorption peaks in the NO-TPD curves of the Fe-Z(Air-6) and
Fe-Z(Air-2) catalysts hardly changed. On the contrary, the areas of
both desorption peaks in the NO-TPD curves of the Fe-Z(N2 -60)
catalyst were slightly higher than those of the Fe-Z(N2 -36) catalyst, which promoted the SCR activity of the Fe-Z(N2 -60) catalyst
at low temperatures, compared with the Fe-Z(N2 -36) catalyst.

S. Lai et al. / Journal of Molecular Catalysis A: Chemical 424 (2016) 232240

237

Fe-Z(N2-36)

Fe-Z(Air-2)
1165

1165

0.02

Absorbance

0.01

1592

1740

1947

1268
1463

70 min

Absorbance

1517

1947

1740

1592 1517
1463

1268
70 min

60

60

40

40

20

20

10

10

2000

1800

1600

1400

-1

1200

1000

2000

1800

1600

1400

-1

1200

1000

Wavenumber (cm )

Wavenumber (cm )

Fig. 7. In situ DRIFT spectra of Fe-Z(Air-2) and Fe-Z(N2 -36) catalysts exposed in a ow of 500 ppm NH3 /Ar (50 mL/min) at 150 C for different times.

Fe-Z(Air-2)

Fe-Z(N2-36)
1165

1165

0.01
1668 15621517
1463

Absorbance

Absorbance

1592

0.02
1268

30 min

10

30 min

1668 1562
1517
1268
1592
1463

10
5

NH3

2000

NH3

1800

1600

1400

-1

1200

1000

Wavenumber (cm )

2000

1800

1600

1400

-1

1200

1000

Wavenumber (cm )

Fig. 8. In situ DRIFT spectra of Fe-Z(Air-2) and Fe-Z(N2 -36) catalysts after adsorbing NH3 , followed by introduction of 500 ppm NO + 5 vol.% O2 /Ar at 150 C for different times.

3.4. In situ DRIFTS absorption spectroscopy


3.4.1. NH3 adsorption and the surface reaction of NO + O2
Fig. 7 shows in situ DRIFTS spectra of the Fe-Z(Air-2) and FeZ(N2 -36) catalysts during the adsorption of NH3 at 150 C. Prior to
NH3 adsorption, the catalyst was pre-treated at 550 C in Ar for 1 h
and then cooled to 150 C. The pre-treated catalyst was exposed
to a ow of 500 ppm NH3 /Ar (50 mL/min) at 150 C, and IR spectra
were taken at different times. As shown in Fig. 7, after being exposed
to NH3 at 150 C, many bands appeared at 1947, 1740, 1592, 1517,
1463, 1268 and 1165 cm1 . The bands at 1947, 1740 and 1463 cm1
were attributed to NH4 + species chemisorbed on the Brnsted acid
sites, and the bands at 1592, 1268 and 1165 cm1 were assigned
to coordinated NH3 bound to Lewis acid sites [4951]. The band at
1517 cm1 might be attributed to amide (-NH2 ) species [52]. The
intensities of all bands increased signicantly in the initial 60 min
and then remained stable, and the types of NH3 species adsorbed

on the Fe-Z(Air-2) and Fe-Z(N2 -36) catalysts were identical. However, the intensities of the corresponding bands for the Fe-Z(N2 -36)
catalyst were stronger than those for the Fe-Z(Air-2) catalyst, indicating that the Fe-Z(N2 -36) catalyst possessed a higher adsorption
capacity for NH3 , in accordance with the NH3 -TPD results.
At the end of the experiment corresponding to Fig. 7, the IR
cell was purged with Ar for 30 min at 150 C. Then, the mixed gas
NO + O2 /Ar was introduced into the IR cell at 150 C, and in situ
DRIFT spectra were recorded at different times. As shown in Fig. 8,
after NO + O2 was introduced into the cell, the intensities of the
bands at 1740, 1592, 1268 and 1165 cm1 rst decreased and
then became stable, indicating that some of the NH4 + species on
Brnsted acid sites and coordinated NH3 bound to Lewis acid sites
participated in the SCR process. The band assigned to amide (-NH2 )
species at 1517 cm1 was unchanged after 30 min, while many new
bands were detected at 1668 and 1562 cm1 , which were attributed
to NOx species. It is noticeable from Fig. 8 that the intensities of the

238

S. Lai et al. / Journal of Molecular Catalysis A: Chemical 424 (2016) 232240

Fe-Z(Air-2)

Fe-Z(N2-36)
1630

0.005

1688

Absorbance

60 min

50
40

50
40

30

30

20

20

10

10

2000

1800

1630
1570

0.005

1688

60 min

Absorbance

1574

1600

1400

1200

-1

2000

1800

1600

1400

1200

Wavenumber (cm-1)

Wavenumber (cm )

Fig. 9. In situ DRIFT spectra of Fe-Z(N2 -36) and Fe-Z(Air-2) catalysts in a ow of 500 ppm NO + 5% O2 /Ar (50 mL/min) at 150 C for different times.

Fe-Z(N2-36)
0.02

1574

Fe-Z(Air-2)

1169
1268

1630

1630

1947

50

1570

1268

60min

1947

Absorbance

60min

Absorbance

1169

0.01

50

30

30

20

20

4
NO+O2

NO+O2

2000

1800

1600

1400

-1

1200

1000

Wavenumber (cm )

2000

1800

1600

1400

1200

1000

-1

Wavenumber (cm )

Fig. 10. In situ DRIFT spectra of (A) Fe-Z(N2 -36) and (B) Fe-Z(Air-2) catalysts pre-treated by exposure to 500 ppm NO + 5 vol.% O2 /Ar, followed by exposure to 500 ppm NH3
at 150 C for different times.

corresponding bands decreased more markedly for the Fe-Z(N2 -36)


catalyst compared with the Fe-Z(Air-2) catalyst due to the higher
SCR activity of the Fe-Z(N2 -36) catalyst at low temperatures (Fig. 1).
3.4.2. NO + O2 adsorption and the reaction with NH3
In situ DRIFTS spectra for the co-adsorption of NO and O2 are
shown in Fig. 9. Prior to adsorption, the catalyst was pre-treated at
550 C in Ar for 1 h and then cooled to 150 C. The pre-treated catalysts were exposed to a ow of 500 ppm NO + 5%O2 /Ar (50 mL/min)
at 150 C, and IR spectra were taken at different times. As shown
in Fig. 9, the intensities of bands were lower due to the weak
adsorption of NOx on the surface of the Fe-Z(N2 -36) and Fe-Z(Air2) catalysts. There were two relatively strong bands at 1574 and
1630 cm1 , which can be attributed to bidentate nitrate [51,53] and
the asymmetric stretching frequency of gaseous NO2 , respectively
[49,54,55]. The intensities of these bands increased gradually with
increasing time, and the relative intensity of the band at 1574 cm1

for the Fe-Z(N2 -36) catalyst was remarkably stronger than that for
the Fe-Z(Air-2) catalyst. It was reported that the amount of bidentate nitrate on the catalyst can affect its activity in the SCR reaction
[56]. Therefore, the higher amount of bidentate nitrate on the surface of the Fe-Z(N2 -36) catalyst may possibly be responsible for its
higher SCR activity, compared with the Fe-Z(Air-2) catalyst.
At the end of the experiment corresponding to Fig. 9, the IR cell
was purged with Ar for 30 min at 150 C. After this, the mixed gas
of 500 ppm NH3 /Ar was introduced into the IR cell at 150 C, and
in situ DRIFT spectra were recorded at different times. As shown
in Fig. 10, after NH3 was introduced into the system, the intensity of the band at 1630 cm1 (assigned to gaseous NO2 ) decreased
quickly, but the change in the band at 1574 cm1 (corresponding to bidentate nitrate) was unclear because of overlap with the
band from the strongly adsorbed NH3 species. These results reveal
that the gaseous NO2 can react with adsorbed NH3 species, but it
was uncertain whether nitrate species adsorbed on the surface of

S. Lai et al. / Journal of Molecular Catalysis A: Chemical 424 (2016) 232240

Fe-ZSM-5 catalysts were active in the SCR reaction. However, the


intensities of the bands at 1169, 1268, 1516, 1593 and 1947 cm1
increased markedly due to the adsorption of NH3 on the catalyst
surface. Further, the intensities of these bands for the Fe-Z(N2 -36)
catalyst were stronger than those for Fe-Z(Air-2) catalyst, coinciding with the in situ DRIFTS of NH3 adsorption (Fig. 7).
As shown in Figs. 7 and 8, NH3 was adsorbed on the Brnsted
and Lewis acid sites on Fe-ZSM-5 catalysts. Although there was
no difference in the types of NH3 species adsorbed or the acidic
sites between Fe-Z(Air-2) and Fe-Z(N2 -36), the Fe-Z(N2 -36) catalyst possessed a higher adsorption capacity for NH3 . When the
catalysts pre-treated with NH3 were exposed to the mixed gas of
NO and O2 , part of both the NH4 + species on Brnsted acid sites
and coordinated NH3 bound to Lewis acid sites participated in
the SCR process. These results indicate that the SCR reaction over
the Fe-Z(N2 -36) and Fe-Z(Air-2) catalysts followed the Eley-Rideal
mechanism. The NO species can be adsorbed on the surface of FeZ(N2 -36) and Fe-Z(Air-2) catalysts in the form of bidentate nitrate,
but it is still unclear whether these nitrate species on the surface
of Fe-ZSM-5 catalysts are reactive in the SCR reaction. Therefore,
the Langmuir-Hinshelwood mechanism could not be dened over
the Fe-Z(N2 -36) and Fe-Z(Air-2) catalysts. However, it may be suggested that the Eley-Rideal mechanism is dominant in the NH3 -SCR
reaction over Fe-Z(N2 -36) and Fe-Z(Air-2) catalysts because both
Fe-Z(N2 -36) and Fe-Z(Air-2) catalysts exhibit a weak adsorption
capacity for NOx (Fig. 9).

21333003), Science and Technology Commission of Shanghai


Municipality (16ZR1407900) and Fundamental Research Funds for
the Central Universities (WJ1514020).
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]

4. Conclusions
In summary, the atmosphere during ion exchange and calcination remarkably affects the cation exchange capacity and the nature
of the Fe species in Fe-ZSM-5 catalysts. An air atmosphere can
improve the ability of ZSM-5 zeolites for Fe ionic exchange. There
are three types of Fe species in Fe-ZSM-5 catalysts, i.e., isolated Fe3+
sites, small oligomeric Fex Oy clusters and Fex Oy nano-particles, and
their distribution depends on the atmosphere during the preparation process. Isolated Fe3+ sites and oligomeric Fex Oy are dominant
in the Fe-Z(N2 -x) catalysts, whereas the amount of Fex Oy nanoparticles is markedly increased in the Fe-Z(Air-x) catalysts.
Since the isolated Fe3+ sites possess high SCR activity at low temperatures and Fex Oy particles show high activity for non-selective
oxidation of NH3 , Fe-Z(N2 -x) catalysts exhibit a higher SCR activity
at low temperatures and a wide temperature window. Additionally, Fe-Z(N2 -x) catalysts possess more Brnsted acid sites than
Fe-Z(Air-x) catalysts. These can adsorb more NH3 species on the
catalyst surface, thus promoting its SCR activity. Similarly, the FeZ(N2 -x) catalyst has a higher adsorption capacity for nitrate species,
which is also benecial to its SCR activity.
In situ DRIFTS was applied to investigate NH3 /NOx adsorption
and the surface reaction of NH3 with NOx . The results show that
NH3 can be adsorbed on both Brnsted acid sites and Lewis acid
sites on the Fe-Z(N2 -36) and Fe-Z(Air-2) catalysts and participates
in the SCR process. Additionally, NOx can be adsorbed on Fe-ZSM5 catalyst mainly as a bidentate nitrate, but these nitrate species
cannot be conrmed to react with gaseous NH3 . Based on the weak
adsorption of NOx over the Fe-Z(N2 -36) and Fe-Z(Air-2) catalysts, it
was proposed that NH3 -SCR over the prepared Fe-ZSM-5 catalysts
mainly follows the Eley-Rideal reaction mechanism.
Acknowledgements

[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]

This project was nancially supported by the National Basic


Research Program of China (2013CB933200), the National Key
Research and Development Program of China (2016YFC0204300),
the National Natural Science Foundation of China (21577034,

239

[45]
[46]
[47]

M.D. Amiridis, R.V. Duevel, I.E. Wachs, Appl. Catal. B 20 (1999) 111122.
S.L. Zhang, Q. Zhong, J. Mol. Catal. A 373 (2013) 108113.
R.Q. Long, R.T. Yang, J. Catal. 188 (1999) 332339.
Y. Cao, S. Zou, L. Lan, Z.Z. Yang, H.D. Xu, T. Lin, M.C. Gong, Y.Q. Chen, J. Mol.
Catal. A 398 (2015) 304311.
H.L. Zhang, C.J. Tang, C.Z. Sun, L. Qi, F. Gao, L. Dong, Y. Chen, Micropor.
Mesopor. Mater. 151 (2012) 4455.
R. Zukerman, L. Vradman, L. Titelman, C. Weidenthaler, M.V. Landau, M.
Herskowitz, Micropor. Mesopor. Mater. 116 (2008) 237245.
L. Xu, X.S. Li, M. Crocker, Z.S. Zhang, A.M. Zhu, C. Shi, J. Mol. Catal. A 378 (2013)
8290.
W.P. Shan, F.D. Liu, H. He, X.Y. Shi, C.B. Zhang, Chem. Commun. 47 (2011)
80468048.
D.M. Meng, W.C. Zhan, Y. Guo, Y.L. Guo, L. Wang, G.Z. Lu, ACS Catal. 5 (2015)
59735983.
R.Q. Long, R.T. Yang, J. Am. Chem. Soc. 121 (1999) 55955596.
L. Olsson, H. Sjvall, R.J. Blint, Appl. Catal. B 81 (2008) 203217.
N. Wilken, K. Kamasamudram, N.W. Currier, J. Li, A. Yezerets, L. Olsson, Catal.
Today 151 (2010) 237243.
A.V. Kucherov, D.E. Doronkin, A.Y. Stakheev, A.L. Kustov, M. Grill, J. Mol. Catal.
A 325 (2010) 7378.
G. Qi, R.T. Yang, R. Chang, Catal. Lett. 87 (2003) 6771.
J. Prez-Ramrez, J.M. Garca-Corts, F. Kapteijn, G. Mul, J.A. Moulijn, C.
Salinas-Martnez de Lecea, Appl. Catal. B 29 (2001) 285298.
J.H. Kwak, D. Tran, J. Szanyi, C.H.F. Peden, J.H. Lee, Catal. lett. 142 (2012)
295301.
L. Ren, L. Zhu, C. Yang, Y. Chen, Q. Sun, H. Zhang, C. Li, F. Nawaz, X. Meng, F.
Xiao, Chem. Commun. 47 (2011) 97899791.
U. Deka, A. Juhin, E.A. Eilertsen, H. Emerich, M.A. Green, S.T. Korhonen, B.M.
Weckhuysen, A.M. Beale, J. Phys. Chem. C 116 (2012) 48094818.
I.M. Saaid, A.R. Mohamed, S. Bhatia, J. Mol. Catal. A 189 (2002) 241250.
P. Boron, L. Chmielarz, J. Gurgul, K. Latka, B. Gil, B. Marszalek, S. Dzwigaj,
Micropor. Mesopor. Mater. 203 (2015) 7385.
M. Richter, A. Trunschke, U. Bentrup, K.W. Brzezinka, E. Schreier, M.
Schneider, M.M. Pohl, R. Fricke, J. Catal. 206 (2002) 98113.
K. Krishna, G.B.F. Seijger, C.M. van den Bleek, M. Makkee, H.P.A. Calis, Top.
Catal. 30 (2004) 115121.
S. Brandenberger, O. Krcher, A. Tissler, R. Althoff, Ind. Eng. Chem. Res. 50
(2011) 43084319.
J. Prez-Ramrez, G. Mul, F. Kapteijn, J.A. Moulijn, A.R. Overweg, A. Domnech,
A. Ribera, I.W.C.E. Arends, J. Catal. 207 (2002) 113126.
L. Zhu, L. Zhang, H.X. Qu, Q. Zhong, J. Mol. Catal. A 409 (2015) 207215.
A. Shishkin, P.A. Carlsson, H. Harelind, M. Skoglundh, Topic in Catal. 56 (2013)
567575.
R.Q. Long, R.T. Yang, Catal. Lett. 74 (2001) 201205.
S. Brandenberger, O. Krocher, A. Tissler, R. Althoff, Catal. Rev. Sci. Eng. 50
(2008) 492531.
J.A.Z. Pieterse, S. Booneveld, R.W. van den Brink, Appl. Catal. B 51 (2004)
215228.
L. Ma, J.H. Li, Y.S. Cheng, C.K. Lambert, L.X. Fu, Environ. Sci. Technol. 46 (2012)
17471754.
M.S. Kumar, M. Schwidder, W. Grnert, A. Brckner, J. Catal. 227 (2004)
384397.
P. Marturano, L. Drozdova, G.D. Pirngruber, A. Kogelbauer, R. Prins, Phys.
Chem. Chem. Phys. 3 (2001) 55855595.
A.A. Battiston, J.H. Bitter, D.C. Koningsberger, Catal. Lett. 66 (2000) 7579.
S. Brandenberger, O. Krcher, A. Tissler, R. Althoff, Ind. Eng. Chem. Res. 50
(2011) 43084319.
M. Rivallan, G. Berlier, G. Ricchiardi, A. Zecchina, M.T. Nechita, U. Olsbye, Appl.
Catal. B 84 (2008) 204213.
S. Brandenberger, O. Krchera, A. Wokaun, A. Tissler, R. Althoff, J. Catal. 268
(2009) 297306.
Handbook of X-Ray Photoelectron Spectroscopy, in: W.M. Wagner, L.E. Riggs,
J.F. Davis, G.E. Moulder (Eds.), Perkin-Elmer Corp., Eden Prairie, 1979.
H. Den Daas, M. Passacantando, L. Lozzi, S. Santucci, P. Picozzi, Surf. Sci. 317
(1994) 295302.
N.S. McIntyre, D.G. Zetaruk, Anal. Chem. 49 (1977) 15211529.
H. Seyama, M. Soma, J. Electron Spectrosc. Relat. Phenom. 42 (1987) 97101.
C.L. Corkhill, D.J. Vaughan, Appl. Geochem. 24 (2009) 23422361.
P. Mills, J.L. Sullivan, J. Phys. D: Appl. Phys. 16 (1983) 723.
S. Brandenberger, O. Krcher, A. Tissler, R. Althoff, Appl. Catal. A 373 (2010)
168175.
S. Brandenberger, O. Krcher, A. Tissler, R. Althoff, Appl. Catal. B 95 (2010)
348357.
L.J. Lobree, I.C. Hwang, J.A. Reimer, J. Catal. 186 (1999) 242253.
K. Krishna, M. Makkee, Catal. Today 114 (2006) 2330.
M.S. Kumar, M. Schwidder, W. Grnert, U. Bentrup, A. Brckner, J. Catal. 239
(2006) 173186.

240

S. Lai et al. / Journal of Molecular Catalysis A: Chemical 424 (2016) 232240

[48] S. Brandenberger, O. Krcher, M. Casapu, A. Tissler, R. Althoff, Appl. Catal. B


101 (2011) 649659.
[49] L. Chen, J. Li, M. Ge, Environ. Sci. Technol. 44 (2010) 95909596.
[50] R.Q. Long, R.T. Yang, J. Catal. 194 (2000) 8090.
[51] B. Guan, H. Lin, L. Zhu, Z. Huang, Phys. Chem. C 115 (2011) 1285012863.

[52] W.S. Kijlstra, D.S. Brands, H.I. Smit, E.K. Poels, A. Bliek, J. Catal. 171 (1997)
219230.
[53] A. Boix, R. Mariscal, J.L.G. Fierro, Catal. Lett. 68 (2000) 169174.
[54] H.Y. Chen, T. Voskoboinikov, W.M.H. Sachtler, J. Catal. 180 (1998) 171183.
[55] R.Q. Long, R.T. Yang, J. Catal. 190 (2000) 2231.
[56] Z.C. Si, D. Weng, X.D. Wu, J. Li, G. Li, J. Catal. 271 (2010) 4351.

S-ar putea să vă placă și