Вы находитесь на странице: 1из 13

2368 Phys. Chem. Chem. Phys.

, 2013, 15, 2368--2380 This journal is c the Owner Societies 2013


Cite this: Phys. Chem. Chem. Phys., 2013,
15, 2368
Characterization of Cu-SSZ-13 NH
3
SCR catalysts: an
in situ FTIR study
Janos Szanyi,* Ja Hun Kwak, Haiyang Zhu and Charles H. F. Peden
The adsorption of CO and NO over Cu-SSZ-13 zeolite catalysts, highly active in the selective catalytic
reduction of NO
x
with NH
3
, was investigated by FTIR spectroscopy, and the results obtained were
compared to those collected from other Cu-ion exchanged zeolites (Y,FAU and ZSM-5). Under low CO
pressures and at room temperature (295 K), CO forms monocarbonyls exclusively on the Cu
+
ions, while
in the presence of gas phase CO dicarbonyls on Cu
+
and adsorbed CO on Cu
2+
centers form, as well. At
low (cryogenic) sample temperatures, tricarbonyl formation on Cu
+
sites was also observed. The
adsorption of NO produces IR bands that can be assigned to nitrosyls bound to both Cu
+
and Cu
2+
centers, and NO
+
species located in charge compensating cationic positions of the chabasite framework.
On the reduced Cu-SSZ-13 samples the formation of N
2
O was also detected. The assignment of the
adsorbed NO
x
species was aided by adsorption experiments with isotopically labeled
15
NO. The
movement of Cu ions from the sterically hindered six member ring position to the more accessible
cavity positions as a result of their interaction with adsorbates (NO and H
2
O) was clearly evidenced.
Comparisons of the spectroscopy data obtained in the static transmission IR system to those collected in
the flow-through diuse reflectance cell points out that care must be taken when general conclusions
are drawn about the adsorptive and reactive properties of metal cation centers based on a set of data
collected under well defined, specific experimental conditions.
Introduction
The growing demand for fuel ecient vehicles has resulted in
the widespread application of lean burn combustion engines.
However, the removal (reduction) of NO
x
from internal combustion
engines operating under net oxidizing conditions poses a great
challenge, since the traditional three-way catalysts are completely
inecient for this process in the presence of excess oxygen. The
need for an ecient lean NO
x
reduction catalyst has prompted
considerable research eorts over the past twenty years to
develop new catalyst materials and catalytic processes. Zeolite-
based catalysts (in particular Cu- and Fe-ZSM-5) have shown
promising activity for both the direct decomposition of NO to N
2
and in the selective catalytic reduction (SCR) of NO with either
NH
3
or hydrocarbons.
127
Other metal-exchanged zeolites (e.g.,
Cu-Y, Cu-beta) have also displayed high initial activities for the
selective catalytic reduction of NO
x
with NH
3
; however, all of
these materials lacked the high temperature hydrothermal
stability that is required in automotive exhaust control catalysis.
Therefore, new strategies and catalyst systems have been
developed and tested (e.g., non-thermal plasma-assisted NO
x
reduction, NO
x
storage/reduction) to meet the ever tightening
environmental regulation limits on emitted NO
x
. Recently, a new,
small pore copper ion exchanged zeolite catalyst (Cu-SSZ-13) with
a chabasite (CHA) structure has been shown to exhibit both
remarkable activity and very high hydrothermal stability in the
NH
3
SCR process.
2832
Along with research aimed at assessing the properties of the
Cu-SSZ-13 catalysts under practical engine exhaust operating
conditions, the fundamental chemical and physical properties
of this very interesting catalytic material have recently been the
focus of considerable study. Initial reports have suggested that
Cu ions are located only in the six-membered rings of the
double six-member prisms of this structure.
29,33,34
However,
recent studies have revealed that Cu ions could occupy sites
near the six-member rings, and in the larger cavities of the CHA
framework. The results of H
2
temperature programmed
reduction experiments also suggested the facile movement
of copper cations in the presence of strongly interacting
adsorbates (e.g., H
2
O), significantly modifying the reducibilities
of Cu ions.
35
Institute for Integrated Catalysis, Pacific Northwest National Laboratory, Richland,
WA 99352, USA. E-mail: janos.szanyi@pnnl.gov
Electronic supplementary information (ESI) available. See DOI: 10.1039/
c2cp43467a
Received 1st October 2012,
Accepted 11th December 2012
DOI: 10.1039/c2cp43467a
www.rsc.org/pccp
PCCP
PAPER
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.
View Article Online
View Journal | View Issue
This journal is c the Owner Societies 2013 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 2369
Infrared (IR) spectroscopy is one of the most commonly used
spectroscopic techniques for the characterization of zeolite-
based catalysts. IR spectra of probe molecules are routinely
used to characterize the adsorption and reactivity properties of
both Bronsted (protons) and Lewis (metal cations) acidic
sites.
3640
The two most commonly used adsorbates for the
characterization of Cu-ions in cationic positions are CO and NO
due to their specific interaction with Cu
+
and Cu
2+
cations,
respectively. Carbon monoxide interacts selectively with Cu
+
ions, forming monocarbonyls at room temperature and under
low CO pressure conditions. At high CO pressures, the for-
mation of di-carbonyls is observed. Under these conditions
(around 300 K sample temperature and moderate pressure),
CO does not adsorb on Cu
2+
sites. On the other hand, NO
interacts with both Cu
+
and Cu
2+
sites, and besides the
adsorbed Cu
n+
NO(x) complexes, decomposition of adsorbed
NO has also been observed over certain zeolite structures
(e.g., ZSM-5). The results of these adsorption studies, in turn,
are frequently used for understanding the catalytic properties of
these materials, and more importantly to unravel the reaction
mechanisms of the complex NH
3
SCR process.
Here we report the results of our in situ IR spectroscopy
studies obtained under both static and flow conditions,
applying transmission and diuse reflectance spectroscopy,
respectively. The results obtained clearly indicate that care
must be taken when conclusion about the adsorption properties
of cationic centers is drawn based on a certain set of experimental
procedures. Here we find that in spite of many similarities,
significant discrepancies exist between data sets obtained in
the static and flow-through IR cells. Data presented provide
further evidence for the facile cation movement of the charge
compensating ions in the zeolite framework upon their inter-
action with strongly adsorbing molecules.
Experimental
The Cu-SSZ-13 catalysts used throughout this study (the same
catalysts we have used in our prior catalytic studies
30,31,35
) were
prepared by an aqueous ion exchange procedure using a
Cu(NO
3
)
2
solution. Briefly, the parent SSZ-13 zeolite was synthe-
sized in-house following the procedure described by Fickel and
Lobo.
33
The Si/2Al (B12) and the Cu/Al ratios (B0.4) were
determined by elemental analysis (ICP), while the crystallinity
of the zeolite was confirmed by XRD. (Since the parent zeolite
was prepared in Na form, most of the residual cations after the
aqueous copper ion exchange procedure were Na
+
. However, IR
spectra always showed the presence of a small number of
protonic sites originating from the ion exchange method used.)
The in situ static transmission IR experiments were con-
ducted in a home-built cell housed in the sample compartment
of a Bruker Vertex 80 spectrometer, equipped with an MCT
detector and operated at 4 cm
1
resolution. Each spectrum
reported is obtained by averaging 256 scans. Prior to spectrum
collection, a background with the activated (annealed, reduced
or oxidized) sample in the IR beam was collected. The powder
sample was pressed onto a tungsten mesh which, in turn, was
mounted onto a copper heating assembly attached to a ceramic
feedthrough. The sample could be resistively heated, and the
sample temperature was monitored by a thermocouple spot
welded onto the top center of the W grid. The sample could also
be cooled with liquid nitrogen to approximately 100 K. The IR
cell is attached to both a gas manifold and a mass spectro-
meter. The cold finger on the glass bulb containing CO was
cooled with liquid nitrogen to eliminate any contamination
originating from metal carbonyls, while NO was cleaned with
multiple freezepumpthaw cycles. The adsorption of NO was
also studied in a DRIFT flow reactor system (Thermo Fisher
Scientific), using a Mattson Research Series 1000 spectrometer
(equipped with an MCT detector, operated at 4 cm
1
resolution,
and averaging 256 scans for each spectrum). The flow rates of
the gases (NO and He) were controlled by mass flow controllers
(Brooks Instruments).
Results and discussion
1. CO adsorption
The adsorption of CO on a Cu-SSZ-13 sample reduced in the
presence of B2 Torr of CO at 573 K and then annealed in
vacuum at 773 K for 2 h was investigated at 295 K sample
temperature, and selected IR spectra from the series collected
during the stepwise CO exposure are displayed in Fig. 1(a).
(Similar experiments were carried out on annealed and oxidized
Cu-SSZ-13 samples, and the spectra (shown in Fig. S1 and S2, ESI)
are essentially identical to those shown in Fig. 1 for the reduced
sample). At low CO coverages (from the 2nd (green) through the
8th (red) spectrum) one highly symmetric feature centered at
2154 cm
1
with a low intensity shoulder at B2135 cm
1
can be
observed. This band has been routinely assigned to the n
CQO
stretching vibration of Cu
+
-adsorbed COin other zeolite structures.
As will be discussed in more detail below, the shoulder feature at
B2135 cm
1
can be assigned to the n
CQO
stretching vibration of
a similar Cu
+
-adsorbed CO species that is located in a more
constrained environment within the zeolite.
35
With further increase in the amount of CO introduced into
the IR cell, the intensity of this main band increases, and levels
o as the peak becomes broader. With the broadening of this
feature, a new IR band appears on the high frequency side of
the 2154 cm
1
peak (centered at 2178 cm
1
), and becomes
more intense with additional amounts of CO. The development
of this new IR feature coincides with a red shift of the main IR
feature. Fig. 1(b) shows the variations of the intensities of both
the 2154 and 2178 cm
1
IR features, together with the observed
redshift in the position of the 2154 cm
1
band as a function of
total CO pressure (i.e., amount of CO) introduced into the IR
cell. The appearance of the 2178 cm
1
IR band and the
simultaneous redshift of the 2154 cm
1
feature clearly suggest
the formation of di-carbonyl species bound to Cu
+
cations. The
formation of these di-carbonyl species, however, happens only
in the presence of excess gas phase in the IR cell. As long as all
of the CO introduced into the IR cell is consumed by adsorption
on the Cu
+
sites, we see only the large intensity gain of the
2154 cm
1
band. However, as soon as CO is present in the cell
Paper PCCP
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online
2370 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 This journal is c the Owner Societies 2013
in measurable amounts (>10
3
Torr), Cu
+
-bound dicarbonyls are
present on the sample. With the appearance of the 2178 cm
1
IR
band of Cu
+
(CO)
2
species, another low intensity IR feature
develops at 2220 cm
1
that can be assigned to the stretching
vibration of Cu
2+
-bound CO. The presence of Cu
2+
ions in our
sample is not surprising, since the reduction was carried out with
CO and the max reduction temperature was 573 K. (Comparison
of the series of IR spectra obtained from the reduced (Fig. 1)
and annealed (Fig. S1, ESI) Cu-SSZ-13 samples revealed that
annealing under vacuum (o1 10
7
Torr) resulted in a very
high extent of reduction of the Cu
2+
species that are exclusively
present in the catalyst after preparation. The series of IR spectra
collected after annealing the sample to increasingly higher
temperatures (from 473 K to 773 K, Fig. S3, ESI) clearly shows
the ecient reduction of Cu
2+
ions by annealing in vacuum at
increasingly higher temperatures. The IR spectra collected after
773 K annealing and after 573 K CO reduction +773 K annealing
are essentially identical.) Both the Cu
+
(CO)
2
and Cu
2+
CO species
are very unstable. After a short (1 min) evacuation (P o10
6
Torr),
both of these IR features completely disappear, while the
2154 cm
1
band retains most of its intensity.
The IR spectra of Cu
+
-bound CO on Cu-SSZ-13 (annealed,
reduced, oxidized) are very similar to those we and others
have measured on both Cu-ZSM-5
41,42
and Cu-Y [ref. 43, and
references therein] zeolites. The series of IR spectra obtained
after CO adsorption on a 773 K-annealed Cu-ZSM-5 sample is
displayed in Fig. S4A (ESI). At low CO pressures there is only
one, highly symmetric absorption feature, centered at 2158 cm
1
.
As the intensity of this IR band saturates, its position red shifts,
and a new IR feature develops at 2178 cm
1
. Similarly to what we
have seen for Cu-SSZ-13, these latter changes are due to the
formation of dicarbonyl species on Cu
+
ions. At the same time,
two new IR bands with low intensities appear at 2222 and
2203 cm
1
, characteristic of CO adsorbed onto Cu
2+
ions. Both
the Cu
+
-bound dicarbonyls and Cu
2+
-adsorbed CO appear only in
the presence of excess CO (i.e., in the presence of gas phase CO).
After a brief evacuation all of these low-intensity features
disappear. (Fig. S4B in the ESI shows a series of IR spectra
recorded after CO exposure and subsequent evacuation of
Cu-ZSM-5 samples annealed at 773 K, reduced at 473, 573,
673, and 773 K. The presence of one, highly symmetric absorp-
tion feature suggests that the environment around all the Cu
+
ions in the ZSM-5 structure is very similar.) The series of IR
spectra of adsorbed CO on Cu-Y annealed at 773 K (Fig. S5,
ESI) shows distinctly dierent characteristics from both
Cu-SSZ-13 and Cu-ZSM-5. In this experiment a given amount
of CO (10 Torr) was introduced into the IR cell (T = 295 K) and
the spectra were collected as a function of exposure time. After
30 s CO exposure time, the dominant IR feature is centered at
2144 cm
1
, with a shoulder on the high frequency side of this
peak. With exposure time, however, the intensity of the IR band
centered at 2159 cm
1
gradually increased and that of the
2144 cm
1
band decreased. These changes in the IR spectra
of Cu
+
-bound CO in Cu-Y have been observed previously, and
were attributed to the migration of Cu
+
ions in the FAU frame-
work as a result of their interaction with CO molecules [ref. 43,
and references therein]. This latter observation clearly demon-
strates that adsorbed CO is a good probe molecule to identify
Cu
+
ions in dierent chemical environments.
From the results of CO adsorption on the three Cu-form
zeolites (SSZ-13, ZSM-5 and Cu-Y) studied here, we can
conclude that the n
CQO
stretching vibration of Cu
+
-adsorbed
CO is in the range of 21542160 cm
1
for CO adsorbed onto Cu
+
ions that are located in zeolite cavities and pores. On the other
hand, the IR bands positioned in the 21352145 cm
1
region
represent CO molecules adsorbed onto Cu
+
ions located in
constrained environments (e.g., in six-membered rings in both
the Y and ZZS-13 frameworks). This conclusion is also sup-
ported by the observation that Cu
+
-adsorbed dicarbonyl species
were only observed for Cu
+
ions located in the large cavities
and channels of these zeolites, but not on cations sitting in
more constrained framework sites. Although the IR spectra of
adsorbed CO in Cu-ZZS-13 are similar to those obtained from
Cu-Y in that there are two types of mono-carbonyl species in
both structures, there is a remarkable dierence between the
Fig. 1 (a) A series of selected IR spectra collected from reduced (with CO at
573 K) and annealed (at 773 K) Cu-SSZ-13 at 300 K sample temperature as a
function of the amount of CO introduced into the IR cell. (b) Variations in both
the integrated IR intensity and peak positions of the Cu
+
-bound CO species with
the cumulative CO pressure.
PCCP Paper
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online
This journal is c the Owner Societies 2013 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 2371
two sets of IR data. In the Cu-Y zeolite, Cu
+
ions move from one
specific cation site to a more favorable site to accommodate
the adsorption of CO as evidenced by the interconversion of the
2144 cm
1
band into the feature centered at 2159 cm
1
. On the
other hand, in Cu-SSZ-13 the intensity ratio of the 2155 cm
1
/
2135 cm
1
bands remain constant during CO adsorption,
suggesting that the adsorption of CO does not result in move-
ment of cations among the dierent cationic positions in the
CHA structure. (Another alternative explanation of the results
presented above for CO adsorption on Cu-SSZ-13 is based on
the so-called heterogeneous dual cation site (DCS) adsorption
of CO in small pore zeolites. In the IR spectra of adsorbed CO
on Cu ion exchanged K- and Cs-FER zeolites Bulanek et al.
observed the presence of two IR features centered at 2156 cm
1
and 2138 cm
1
.
44
They associated the first band (2156 cm
1
)
with CO adsorbed exclusively on Cu ions (no interaction with
other cations), and the latter one with CO adsorbed also on Cu
+
sites with its C atom, but interacting with a close-by K or Cs ion
through its O atom. The similarities in the positions of these IR
bands to those we have observed in our Cu-SSZ-13 sample may
question our assignment of the vibrational features to CO
molecules adsorbed onto two types of Cu
+
ions located in
dierent cationic positions; however, we firmly believe that
our assignment is correct based on the following observations.
First, the IR signature for the heterogeneous DCS adsorption
was only observed when cations larger than Na
+
(i.e., K
+
and Cs
+
)
coexisted with the Cu
+
ions in the FER structure.
44
In our Cu-SSZ-13
sample the residual cations present in less than 20% abundance
are Na
+
and H
+
, which were shown not to initiate heterogeneous
DCS adsorption of CO. Our prior studies on H
2
temperature
programmed reduction (H
2
-TPR) of a series of Cu-SSZ-13
samples (including the one studied here) have shown unambigu-
ously the presence of two types of Cu ions located in dierent
chemical environments
35
in the SSZ-13 framework. The IR data
presented here and the H
2
-TPR and IR data presented earlier on
Cu-SSZ-13 samples at dierent levels of Cu ion exchange are both
consistent with the assignment of the two IR absorption
features to CO adsorbed onto two types of Cu
+
sites in dierent
chemical environments, i.e., dierent cationic positions in the
zeolite framework.)
2. NO adsorption
A: Transmission FTIR studies. First the adsorption of NO
was investigated on a Cu-SSZ-13 zeolite at 295 K sample
temperature using the same transmission FTIR experimental
setup we used to collect the data presented above for the
adsorption of CO. Prior to NO adsorption, the sample was
calcined at 773 K for 2 h at a base pressure P o1 10
7
Torr.
Two series of selected IR spectra recorded from the 773
K-annealed samples after exposure to
14
NO and
15
NO are
displayed in panels a and b of Fig. 2, respectively. Each
spectrum was collected after the introduction of NO aliquots
into the IR cell, and equilibration. The experiments with
15
N-labelled NO were conducted in order to aid in making
assignments of the IR features in the rather complex spectra
recorded after NO exposure.
The IR spectra obtained after
14
NO exposure can be divided
into three ranges: 17501850 cm
1
, 18502000 cm
1
, and
20502300 cm
1
. IR bands centered at 1808, 1900 and
2170 cm
1
appeared immediately after the introduction of
the first small NO aliquot (0.017 Torr). With the introduction
of the second NO dose, the intensities of all of these bands
increased, and new features developed as well. The initially
asymmetric-shaped band at B1900 cm
1
split into two well-
separated features centered at 1901 and 1909 cm
1
, while a new
peak appeared at 1948 cm
1
. Subsequent NO additions resulted
in intensity gains of all of these IR bands, along with the
appearance of two new features at higher NO pressures; notably,
a very weak shoulder at B1930 cm
1
, and a peak at 2249 cm
1
.
This latter peak can easily be assigned to adsorbed N
2
O.
In order to help the assignment of the observed NO
x
-related
IR bands, we conducted a similar set of experiments using
15
N
labeled NO on the annealed Cu-SSZ-13 catalyst. The series of
selected IR spectra obtained in these experiments are shown in
panel b of Fig. 2. The IR bands after the introduction of the first
aliquot of
15
NO are centered at 1777, 1868, and 2129 cm
1
, red
Fig. 2 Series of selected IR spectra obtained after exposure of annealed Cu-SSZ-
13 samples to
14
NO (a) and
15
NO (b) at 300 K. The samples were annealed in
vacuum at 773 K for 2 h prior to IR measurements.
Paper PCCP
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online
2372 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 This journal is c the Owner Societies 2013
shifted with respect to the features observed after the adsorption
of
14
NO (Dn = 31, 32, and 39 cm
1
, respectively). With increasing
amounts of
15
NO introduced into the IR cell, the development of
the IR bands followed the same trends as we have discussed above
for the adsorption of
14
NO, and the final spectra observed at
similar NO pressures were identical in their main characteristics
for the two isotopically dierent adsorbates.
Using a simple harmonic oscillator approximation for the
diatomic NO molecule, we can calculate the expected vibra-
tional frequencies for the isotopically labeled adsorbed
15
NO.
The thus-calculated vibrational frequencies were identical to
those observed in the IR spectra experimentally, except the one
that is located at 2249 cm
1
after the adsorption of
14
NO, and at
2177 cm
1
for
15
NO. This latter band represents adsorbed N
2
O,
as has been reported previously.
45
All of the other IR features
represent vibrations of the NO bond in the diatomic NO
molecule, and the peak positions are dependent on the adsorp-
tion environment and the charge state of the NO molecule. The
assignment of the IR features can be guided by previously
published results for NO adsorption on other Cu-ion exchanged
zeolites, especially on ZSM-5, Y,FAU and b. Since the ion
exchange was carried out with a Cu(NO
3
)
2
aqueous solution,
and the sample was annealed at 773 K prior to NO adsorption
measurements, one would expect the presence of Cu
2+
ions
exclusively in cationic zeolite exchange positions. However,
some of the Cu
2+
ions in high vacuum can undergo autoreduc-
tion to Cu
+
; therefore, some of the IR bands observed probably
represent Cu
+
NO species (i.e., adsorbed NO on Cu
+
sites).
The autoreduction of Cu
2+
ions in cationic positions of the
ZSM-5 structure has been proposed and extensively studied by
Hall et al.,
5,6,11
although their interpretation of the chemistry
that leads to the formation of Cu
+
ions in high vacuum has
recently been questioned.
46
The EPR study of Occhiuzzi et al.
46
seems to strongly suggest that the reduction of Cu
2+
ions in
Cu-ZSM-5, as a result of annealing in high vacuum, is caused by
the presence of hydrocarbon impurities in the as-prepared
zeolites as residues from the synthesis process that uses
organic templates. In our sample, however, we did not observe
the complete reduction of Cu
2+
ions to Cu
+
in high vacuum at
elevated temperatures and, also, the sample used throughout
this study was calcined at 773 K in an oxygen flow prior to its
use in the IR experiments. (In addition, when this sample was
oxidized at 773 K in the DRIFT cell and then purged with He, no
reduction of the Cu
2+
ions was observed (see below).) Since NO
is able to interact at room temperature with both Cu
2+
and Cu
+
ions (although with dierent strengths of adsorption), the IR
bands can represent NO adsorbed onto either of these two Cu
ions, as well as onto residual protons present in the parent zeolite.
The lowest frequency band centered at around 1810 cm
1
can
be assigned to n
NO
stretching vibrations of the Cu
+
-adsorbed
NO molecules. The IR band for Cu
+
-bound NO on Cu-ZSM-5 is
positioned at 1812 cm
1
(Fig. S6, ESI), on Cu-Y,FAU the peak
for this species is at 1815 cm
1
(Fig. S7, ESI), and on Cu-b at
1813 cm
1
(Fig. S8, ESI). This assignment is also supported by
data collected from samples oxidized at 673 K (Fig. 3), and
reduced at 673 K prior to NO adsorption (Fig. 4). Over the
oxidized sample, the relative intensity of the 1810 cm
1
band
was much lower, while on the reduced sample, somewhat higher
than that observed on the annealed catalyst. In contrast to
Cu-ZSM-5, however, at 300 K sample temperature the formation
of di-nitrosyl species was not observed in Cu-SSZ-13. (Dinitrosyls,
however, can form over this material, but in order for this to
happen the sample needs to be cooled to cryogenic tempera-
tures. In fact, the spectra in Fig. S9 (ESI) clearly show the facile
formation of dinitrosyl species at 200 K sample temperature over
an oxidized Cu-SSZ-13 sample (peaks in the 17201830 cm
1
spectral region).) Note also the presence of a shoulder on the low
frequency side of the 1810 cm
1
IR feature (at B1785 cm
1
) that
suggests the presence of two dierent types of Cu
+
NO com-
plexes following NO adsorption. We have explained this observa-
tion by the presence of Cu ions in two dierent cationic
positions of the CHA framework.
35
The only IR band that still needs to be assigned is the broad,
symmetric feature centered at B2165 cm
1
. Comparison of the
Fig. 3 Selected IR spectra collected upon stepwise NO adsorption on an
oxidized (673 K in O
2
) Cu-SSZ-13 sample at 300 K.
Fig. 4 A series of IR spectra obtained as a function of time from a reduced (with
CO at 673 K) Cu-SSZ-13 sample after its exposure to 10 Torr NO at 300 K.
PCCP Paper
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online
This journal is c the Owner Societies 2013 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 2373
peak position of this band observed with
14
N- and
15
N-labeled
NO (Fig. 2a and b, respectively) clearly suggests that it repre-
sents a diatomic species containing a N and an O atom. Its peak
position, however, is much more blue shifted than those
observed for nitrosyl species bound to either Cu
+
or Cu
2+
ions.
This suggests that the NO species which is responsible for this
IR band is probably an ionic NO
+
species. The weakly Lewis
acidic NO molecule contains an unpaired electron in an anti-
bonding molecular orbital. By donating this electron, the NO
bond becomes stronger, which is manifested in a blue shift of
the n
NO
vibrational frequency. The Lewis acidic Cu
2+
ions in
the Cu-SSZ-13 framework are rather easy to reduce: as we have
discussed above, even evacuation at 300 K resulted in partial
reduction. Therefore, it is reasonable to assume that NO (weak
base) donates an electron to a Cu
2+
ion (strong acid) forming
Cu
+
and NO
+
. This keeps the charge balance of the system
intact, as the two negative charges in the aluminosilicate
framework, balanced originally by a Cu
2+
ion, are neutralized
by a Cu
+
NO
+
ion pair. (The other possible way to create NO
+
ions is through disproportionation of NO
2
:
2NO
2
2NO
+
NO
3

.
For this process to occur, the presence of NO
2
is required.
However, in our system neither O
2
nor NO
2
is added during the
NO adsorption experiments. Furthermore, if this reaction was
to proceed on our sample, we should observe the formation of
nitrate ions as well, which we could not.)
Additional evidence for the formation of NO
+
ions by the
reduction of Cu
2+
ions was obtained from experiments carried
out with oxidized and reduced Cu-SSZ-13 samples. The inte-
grated intensity of the IR peak representing NO
+
was always the
highest on samples with the highest degree of oxidation as is
evident from a comparison of the spectral series obtained from
the oxidized sample (Fig. 3) to those collected from both the
annealed and reduced Cu-SSZ-13 samples (Fig. 2 and 4). (Prior
to NO adsorption, this sample was oxidized at 673 K for 1 h, and
then at 773 K for 10 min. After the 2nd oxidation step, the
sample was cooled down in O
2
to 300 K, and the cell was then
evacuated prior to NO introduction.) The most intense IR
features in the spectra collected from the oxidized sample are
those of Cu
2+
NO, and NO
+
. The intensity of the Cu
+
NO
band is the weakest among all the samples studied here, but
still significant. In contrast, when a CO-reduced (at 673 K)
Cu-SSZ-13 sample was exposed to NO, the highest intensity IR
band is the one centered at 1810 cm
1
, while the intensity of
the 2165 cm
1
band (NO
+
) is very low throughout the NO
adsorption experiment.
Selected IR spectra as a function of
15
NO exposure of an
oxidized and a reduced (at 673 K in H
2
) Cu-SSZ-13 are shown in
panels a and b of Fig. 5, respectively. Comparison of these two
sets of IR spectra clearly show that only very small amounts of
NO
+
species are formed on the H
2
-reduced sample, while on the
oxidized one the intensity of the IR band representing
15
NO
+
(2124 cm
1
) is very high, and it increases slowly with the
amount of
15
NO introduced into the IR cell. Using the data in
Fig. 5a and b, the integrated intensities of peaks associated
with Cu
+

15
NO (B1775 cm
1
, triangles), Cu
2+

15
NO(I)
Fig. 5 Selected IR spectra collected during the
15
NO exposure of oxidized (panel a)
and reduced (panel b) Cu-SSZ-13 samples at 300 K. The samples were exposed to
10 Torr of
15
NO and the time evolution of the IR features was followed. The
integrated IR intensities as a function of
15
NO exposure time for both the
oxidized and reduced Cu-SSZ-13 samples are displayed in panel c.
Paper PCCP
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online
2374 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 This journal is c the Owner Societies 2013
(B18621868 cm
1
, circles), and Cu
2+

15
NO(II) (B1913 cm
1
,
squares) are plotted as a function of
15
NO adsorption time
in Fig. 5c for both the pre-oxidized (closed symbols) and pre-
reduced (open symbols) samples. The broken arrows indicate
changes in the integrated intensities of these IR bands as a
result of oxidation of the catalyst prior to
15
NO adsorption. The
largest change is observed for the Cu
+
NO band, as its
intensity drops about 35% as a result of oxidation. On the
other hand, the intensities of the two Cu
2+
NO bands increase
as a result of reduction. Also of note in Fig. 5c is that the
intensity variations with
15
NO adsorption amount (i.e., time) of
the two types of Cu
2+
NO bands follow dierent trends. While
the intensity of the 18621888 cm
1
band (we denote as
Cu
2+
NO(I): adsorbed NO on Cu
2+
sites in a certain cationic
position) gradually increases with time on both the reduced
and oxidized Cu-SSZ-13 samples, the other Cu
2+
NO band
(we denote it Cu
2+
NO(II)) centered at 1913 cm
1
actually loses
intensity as the amount of NO introduced into the IR cell
increases. Simultaneously, the intensity of the NO
+
band
(B2127 cm
1
) on the H
2
-reduced sample is very low over the
entire dosed NO amount region (in fact we could not integrate
the intensity of this feature due to its overlap with the IR band
arising from adsorbed N
2
O.) These observations also support
our hypothesis that the NO
+
species are formed during the
reduction of Cu
2+
ions. In fact, the results also seem to suggest
that the Cu
2+
ions responsible for the formation of NO
+
are the
ones that are associated with the NO feature at 1913 cm
1
(the one we have previously assigned to be located in the
six member ring of the CHA framework
35
). Note also the
presence of an IR band originating from the adsorption of
N
2
O (B2180 cm
1
). The appearance of this IR feature suggests
that, under our experimental conditions, decomposition of NO
can also take place, particularly for the reduced sample
(the intensity of the IR band representing this adsorbed species
is always much higher on the reduced catalyst than on the
oxidized one).
We next discuss possible origins of the NO
+
species assigned
to the broad features at B2165 cm
1
or B2124 cm
1
in
14
NO
and
15
NO, respectively. A similar IR feature has been observed
in Cu-ZSM-5 at 2135 cm
1
after its exposure to NO + O
2
or
NO
2
,
47
and was assigned to NO
2
+
species in charge compensat-
ing cationic positions. This assignment, however, was later
corrected by Hadjiivanov et al.
48
who unambiguously proved
that the IR band at 2135 cm
1
in (NO + O
2
)-exposed, proton-
containing ZSM-5 samples is due to NO
+
species in cationic
positions. It has been suggested that this species formed by the
reaction of NO
+
species (produced by the disproportionation of
either N
2
O
3
(NO
+
NO
2

) or N
2
O
4
(NO
+
NO
3

)) with protonic sites,


a conclusion seemingly substantiated by the observation of
water production in this process. In our system, however, the
formation of water was not observed. In order to verify
the formation of NO
+
species from the reaction of NO
2
with
the Cu-SSZ-13 material, we conducted an experiment in which
an annealed sample was exposed to NO
2
in increasing quan-
tities. The resulting series of IR spectra is displayed in Fig. 6.
The presence of the NO
+
species (2163 cm
1
), together with
NO
3

(15501630 cm
1
) and NO
2

(1295 cm
1
), is evident in
the spectra even at the lowest NO
2
exposures. As the intensity of
the IR band representing NO
+
species increases with increasing
amounts of NO
2
introduced, a shoulder on the high frequency
side of this band (at B2217 cm
1
) develops and gains intensity.
The appearance of similar features was also observed for
Cu-ZSM-5,
47
and assigned to the formation of NO
+
NO
2
adducts
in the presence of excess NO
2
in the gas phase (this feature
loses its intensity as the gas phase NO
2
is evacuated from the IR
cell). For the case of NO adsorption, this shoulder/new feature
was never observed in our experiment, since no NO
2
was
present in the system.
All of the adsorption experiments with NO and NO
2
point
to two possible mechanisms for NO
+
formation during the
exposure of Cu-SSZ-13 to NO. As we have suggested above,
NO
+
could form by the reaction of NO with a Cu
2+
ion in a
cationic position:
Cu
2+
+ NO -Cu
+
+ NO
+
.
One may also suggest that NO
+
in fact forms only from NO
2
,
even though no NO
2
was introduced into our IR cell. However,
we have shown that this catalyst is active in the decomposition
of NO to produce N
2
O. In this case it is possible that the O atom
originating from this reaction
2NO -N
2
O + O
is picked up by another NO molecule to produce NO
2
NO + O -NO
2
and, in turn, this NO
2
can now participate in the formation of
NO
+
as we have discussed above. Arguing against this latter
mechanism are the results of the
15
NO adsorption experiments
of Fig. 5. If the formation of NO
+
was a consequence of the N
2
O
production, we should have observed much larger amounts of
NO
+
on the reduced sample which achieved the highest levels
of N
2
O formation. This was clearly not the case. Instead, the
Fig. 6 A series of selected IR spectra obtained during the stepwise NO
2
exposure of an annealed Cu-SSZ-13 sample at 300 K.
PCCP Paper
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online
This journal is c the Owner Societies 2013 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 2375
greater the extent that the catalyst was oxidized, the larger
amount of NO
+
was observed. This fact strongly suggests
that the main pathway to NO
+
formation is the reduction of
Cu
2+
by NO.
B: Diuse reflectance FTIR studies. As we have discussed
in the previous section, autoreduction of the Cu-SSZ-13 sample
was always observed in the static IR experiments, as
evidenced by the presence of Cu
+
NO bands for reduced,
annealed and oxidized samples. In part, this may be influenced
by the reducing environment of high vacuum. In order to
exclude (or minimize) this possible route to reduction of the
sample, we have conducted NO adsorption experiments in a
DRIFT cell under continuous flow conditions. The sample was
first annealed, oxidized, or reduced in a constant flow of He,
O
2
/He, or H
2
/He, respectively, at 773 K, cooled to 295 K in the
respective reactant flow, and then purged with He for 2 h prior
to NO adsorption at 295 K. These NO adsorption experiments
were conducted with a 350 ppm NO in He flow, and IR spectra
were collected as a function of exposure time. The three sets of
IR spectra obtained in these experiments are displayed in
panels a, b, and c of Fig. 7 for the annealed, oxidized and
reduced samples, respectively. The IR features observed in
these spectra are the same as those obtained for the samples
treated in the static system. Some dierences in relative inten-
sities for the various species at saturation are expected when
comparing transmission and diuse reflectance FTIR spectra.
However, the variation in peak intensities with NO coverage is
markedly dierent in the spectra collected from samples trea-
ted under static from those recorded in the flow experiments.
The integrated intensities of the NO vibrations of the four
adsorbed species discussed above are displayed in Fig. 7df for
the annealed, oxidized, and reduced samples, respectively.
Upon exposure of the 773 K-annealed sample to NO (panel a
of Fig. 7) for 3 min, IR bands characteristic of NO
+
(2164 cm
1
),
Cu
2+
NO(II) (1948 + 1930 cm
1
) and Cu
+
NO (1809 cm
1
)
appear with the highest intensity, but the signature of
Cu
2+
NO(I) species (18801920 cm
1
) is also present with
low intensity. With increasing NO exposure time, the intensi-
ties of the 2164, 1948 + 1930 and 1809 cm
1
bands gradually
and simultaneously increase, and the latter two bands reach
their maximum intensities after 14 min NO exposure time. At
even longer exposure times, the intensities of the 1948 + 1930
and 1809 cm
1
bands gradually decrease, while those of bands
at 18901910 cm
1
increase significantly (Fig. 7d). The intensity
of the 2164 cm
1
band continues to increase as well. By the
time the sample was exposed to NO for 2 h no Cu
+
-adsorbed NO
species were detected by IR, and the intensity of the 1948 cm
1
band dropped to a low level as well. As the Cu
+
NO species
disappeared from the IR spectrum, a shoulder on the high
frequency side of the 2164 cm
1
band appeared, signaling the
formation of NO + NO
2
adducts. This is exactly the time interval
where we observed the onset of nitrate formation in the
15001700 cm
1
spectral region, and the consumption of the
residual Bronsted acid sites (spectra are not shown for brevity).
After the completion of the experiment (6 h NO exposure), the
two most intense absorption features are the ones that repre-
sent NO
+
and Cu
2+
NO(I) species. In summary, during the
interaction of NO with a 773 K-annealed Cu-SSZ-13 sample,
Fig. 7 Series of DRIFT spectra collected fromannealed (panel a), oxidized (panel b) and reduced (panel c) Cu-SSZ-13 samples at 300 K. The variations of the integrated
IR peak intensities as a function of cumulative NO pressure are shown in panels d, e and f for the annealed, oxidized, and reduced samples, respectively. (IR spectra
were collected at time-on-stream values of 0, 3, 8, 14, 20, 30, 60, 120, and 240 min (spectra a -i) on all three samples. The last spectrum (j) on the annealed sample
was taken at t = 360 min, while on the reduced sample at t = 420 min.)
Paper PCCP
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online
2376 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 This journal is c the Owner Societies 2013
initially Cu
+
NO and Cu
2+
NO(II) species form, and their
concentrations pass through maxima as Cu
2+
NO becomes
the primary Cu-adsorbed NO
x
species. The concentration of the
NO
+
species reaches saturation as NO
+
NO
2
adducts start to
form. The appearance of nitrate features (and the fact that
NO
+
NO
2
adducts form) at very long NO exposure times suggests
that some adsorbed oxygen stayed in the sample after the
annealing process which, in turn, participates in the formation
of NO
2
.
The series of IR spectra recorded during the exposure of the
H
2
-reduced Cu-SSZ-13 sample (Fig. 7c) exhibits a very dierent
trend from that just described for the annealed catalyst. By
reducing the sample at high temperatures (773 K) with H
2
, we
expect the removal of all the adsorbed oxygen from the sample,
and the presence of Cu
+
ions in cationic positions exclusively
(along with some residual proton sites). In our prior H
2
TPR
study, we have shown that all Cu
2+
ions are reduced to Cu
+
ions
under the conditions used in this experiment, without any
reduction to metallic Cu.
35
Thus, as expected, after 3 min of
NO exposure the most intense IR band is the one representing
Cu
+
-adsorbed NO (1809 cm
1
), while a very weak feature
appears at 1947 cm
1
(Cu
2+
NO(II)) and an even weaker peak
at 2231 cm
1
that we assign to adsorbed N
2
O. The intensity
of the Cu
+
NO band increases very fast, and reaches its
maximum at a NO exposure time of 14 min. In this time
interval the intensity of the N
2
O-related band also increases,
while no evidence for the formation of NO
+
is observed. The IR
feature of NO
+
appears first in the spectrum after 20 min NO
exposure and its intensity grows rapidly with time. As the
intensity of this feature increases, that of the N
2
O band
(2231 cm
1
) decreases and completely disappears after 120 min
time-on-stream.
The two Cu
2+
-related NO features follow dierent trends
with NO exposure time: the intensity of the 1947 + 1922 cm
1
band (Cu
2+
NO(II)) increases up to 120 min, then drops
dramatically at longer exposure times. The drop in the intensity
of this feature coincides with the large increase in the intensity
of the 1895 + 1905 cm
1
band (Cu
2+
NO(I)). After 420 min NO
exposure, the two major bands observed in the IR spectrum are
the ones that represent NO
+
and Cu
2+
NO(I). Still, even after
7 h time-on-stream, a trace of the Cu
+
NO species is present,
while the intensity of the Cu
2+
NO(II) species is comparable to
that we have seen for the annealed sample after 240 min NO
exposure time. It is worth mentioning that during the entire NO
exposure time no IR features representing nitrate species were
observed. This further confirms our earlier claim that the
formation of nitrates in the annealed sample most likely
originates from residual oxygen in the sample, even after high
temperature annealing in an inert atmosphere.
One of the most interesting observations on this reduced
sample is the initial formation of a large amount of N
2
O. This
is evidently connected to the exclusive presence of Cu
+
ions in
the sample, since we have only seen the formation of a trace
amount of N
2
O over the annealed Cu-SSZ-13 catalyst where
most of the Cu ions were present as Cu
2+
. The formation of N
2
O
was also observed in the static transmission IR experiments
when NO was introduced onto a Cu-SSZ-13 sample that was
reduced in H
2
(see Fig. 4), and to a lesser extent in the
CO-reduced catalyst. These results seem to suggest that the
N
2
O seen in the IR spectra of reduced Cu-SSZ-13 samples
represents Cu
+
-bound N
2
O species. The integrated intensity of
this Cu
+
N
2
O band, however, goes through a maximum,
following a very similar pattern to the intensity of the Cu
+
NO
species. Note also that the IR feature of NO
+
species is
completely absent from the spectra at short NO exposure times
(up to 14 min), only appearing after the intensities of the
Cu
+
NO and Cu
+
N
2
O bands reach their maxima, and when
Cu
2+
NO bands develop.
The results just described again raise the question about the
origin of NO
+
in this system. As we have mentioned previously,
NO
+
can form by NO reducing Cu
2+
ions to Cu
+
, and then
occupying a charge balancing cationic position. This is consistent
with the lack of evidence for NO
+
until the production of Cu
2+
ions
in appreciable (although unknown) amounts in the H
2
-reduced
sample. In particular, no NO
+
is formed when the sample contains
Cu
+
exclusively. However, there might be another explanation for
the formation of NO
+
species in this system. It may originate from
the formation and subsequent disproportionation of NO
2
. NO
2
may form by the reaction between NO and the oxygen atom
produced during the decomposition of NO to N
2
O. This reaction
path, however, cannot be the major route to NO
+
formation, since
it would require the formation of NO
+
and NO
3

in equal
quantities. Instead, IR bands characteristic of nitrate species were
completely absent during the exposure of the H
2
-reduced
sample to NO.
A series of IR spectra obtained during the exposure of an
oxidized Cu-SSZ-13 sample to NO at 295 K is shown in Fig. 7b.
After 3 min of NO exposure the main IR band is the one that
represents Cu
2+
NO(I) species (centered at 1905 cm
1
), while
the intensities of both the NO
+
(2164 cm
1
) and Cu
2+
NO(II)
(1948 cm
1
) are very low. Note the complete absence of the IR
peak of Cu
+
-bound NO at 1810 cm
1
, confirming the presence
of Cu
2+
ions only after the applied oxidation treatment. (We
should point out that under static conditions, i.e., where the
sample was always evacuated to o1 10
7
Torr following
oxidation, there was always a certain extent of autoreduction
that resulted in the immediate presence of Cu
+
NO species
even on the most extensively oxidized sample.) With increasing
time-on-stream, the transient formation of small amount of
Cu
+
NO species (1810 cm
1
) is seen. At t = 20 min, the
intensity of the band representing this species reaches its
maximum, and then disappears rapidly. The intensity of the
IR band representing the Cu
2+
NO(II) species (1948 cm
1
) follows
a similar trend, although it does not completely diminish even after
the completion of the adsorption experiment (i.e., after 240 min).
The two IR features present after 240 min time-on-stream belong to
NO
+
and Cu
2+
NO(I). Note that a shoulder develops on the high
frequency side of the 2164 cm
1
IR band, signaling the formation of
NO + NO
2
adducts (see above). The appearance of the IR signature
of this adduct coincides with the development of IR features of
nitrate species in the 14001650 cm
1
spectral region (not shown).
The formation of the NO
3

species can only take place in the


PCCP Paper
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online
This journal is c the Owner Societies 2013 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 2377
presence of NO
2
that, in turn, requires the availability of oxygen
in the system to oxidize NO to NO
2
. In this case, since the
sample was oxidized extensively prior to its exposure to NO, it is
proposed that some oxygen remains on the catalyst, and can
participate in NO oxidation. The other oxygen source, as we
have discussed above, is the NO decomposition reaction. In this
sample, however, there is no Cu
+
present initially; therefore, the
contribution of this reaction to the production of adsorbed
oxygen is believed to be negligible.
Comparison of the spectral series obtained in the static and
flow IR experiments indicates that care must be taken when
conclusions are drawn about the nature and formation
mechanisms of adsorbed NO
x
species from data collected
under well-defined experimental conditions. The most important
dierence between the two sets of experimental data (i.e., static
and flow cell) presented here is the eect of high vacuum on the
Cu-SSZ-13 sample: even on the sample that was oxidized multiple
times, cooled to room temperature in O
2
and then evacuated, a
significant fraction of the Cu
2+
ions became reduced. On the other
hand, the flow system which included purging the sample with
He after high temperature oxidation did not result in any
reduction, and the sample only contained Cu
2+
ions. Furthermore,
the results of the NO adsorption experiments on both the
annealed and oxidized samples in the flowsystemseemto suggest
the retention of a non-negligible amount of oxygen following
catalyst preparation. This oxygen, in turn, is able to oxidize NO to
NO
2
, leading to the formation of both nitrate and NO + NO
2
species. No NO
+
NO
2
adduct or NO
3

formation was observed on


the fully reduced sample. Evacuation of the oxidized sample, on
the other hand, presented a strong enough reducing environment
that all the residual oxygen was removed from the catalyst
at 295 K.
3. NO adsorption on H
2
O-exposed Cu-SSZ-13
In the NH
3
SCR process, the reduction of NO is always carried
out in the presence of water in large excess. Water can strongly
interact with cations thus influencing their adsorption and
catalytic properties. In our recent study we have shown that
the addition of water to the H
2
/He stream used in H
2
tempera-
ture programmed reduction (TPR) experiments caused the
reduction temperature to shift by about 100 K lower. We
attributed this shift to the movement of Cu
2+
ions from the
six member rings of the CHA structure toward the large cavities
where H
2
can more readily access these ions. The movement
of cations can also be caused by other strongly interacting
adsorbates (e.g., CO in Cu-Y,FAU
43
). In order to substantiate the
movement of Cu ions toward sterically more accessible cationic
positions, we examined the adsorption of
15
NO on an annealed
Cu-SSZ-13 that was exposed to H
2
O just prior to
15
NO introduc-
tion into the IR cell. The resulting series of IR spectra, collected
in the static IR cell in transmission mode, are displayed in
Fig. 8a. As we have discussed in Section 2A (Fig. 2b), upon
adsorption of
15
NO on a dry sample, IR features representing
15
NO
+
(2124 cm
1
), Cu
2+

15
NO(II) (1917 + 1895 cm
1
) and
Cu
+

15
NO (1775 cm
1
) adsorbed species are present with
the highest intensities, while the intensity of the IR band of
Cu
2+

15
NO(I) (B18601880 cm
1
) is rather low. The series of
IR spectra obtained from the H
2
O-exposed Cu-SSZ-13 sample
exhibits very dierent variations of the intensities of absorption
features as a function of
15
NO exposure from those observed on
the dry sample. The most intense IR band (1873 + 1887 cm
1
)
after the addition of the last
15
NO dose represents adsorbed
15
NO onto Cu
2+
sites located in the large cavities of the CHA
structure. The variation of the relative integrated intensities of
the Cu
2+

15
NO(II) and Cu
2+

15
NO(I) IR bands as a function of
the cumulative
15
NO pressure introduced into the cell is
displayed in panel b of Fig. 8 for both the dry (solid symbols)
and H
2
O pre-dosed (open symbols) annealed Cu-SSZ-13 sam-
ples. The arrows indicate that in the presence of water in the
zeolite there are more NO adsorbed to Cu
2+
sites in positions
located in the large cavities than in the six-member rings. These
observations provide further support to our prior proposal of
Fig. 8 A series of IR spectra (panel a) obtained during the stepwise
15
NO
adsorption of an annealed Cu-SSZ-13 sample at 300 K. Prior to
15
NO adsorption,
the sample was exposed to H
2
O. Panel b shows the variations of integrated IR
peak intensities of the two Cu
2+
-bound adsorbed NO species as a function of
cumulative
15
NO pressure. (Data for the H
2
O-free annealed Cu-SSZ-13 was
obtained from the series shown in Fig. 2b.)
Paper PCCP
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online
2378 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 This journal is c the Owner Societies 2013
Cu ion migration as a result of their interaction with strongly
adsorbing molecules.
The introduction of water into the Cu-SSZ-13 zeolite prior to
15
NO adsorption results in shifts of the vibrational frequencies
of some of the Cu
2+
-bound NO features. The position of the IR
feature representing NO adsorbed onto Cu
2+
in the six member
rings (1913 cm
1
) does not shift at all, while the bands of NO
bound to Cu
2+
ions in the large cavity (1873 and 1887 cm
1
) are
blue shifted by 510 cm
1
in comparison to that in the dry
sample. The variation of the NO frequency with the co-adsorption
of water is not surprising, as the electronic environment around
the Cu
2+
adsorption centers changes as water adsorbs onto
these cations. The observation that the frequencies of neither the
six member ring-located Cu
2+
-bound NO (1913 cm
1
) nor the
Cu
+
-bound NO (2125 cm
1
) change in the presence of water
may be due to the weak interaction between NO (and perhaps
water as well) and these adsorption centers. The series of
selected IR spectra in Fig. S10 (ESI) shows that NO adsorbed
onto either Cu
2+
ions located in the six member rings or onto
Cu
+
sites is weakly held, and can be readily removed by room
temperature evacuation. On the other hand, most of the
intensity of the NO species adsorbed onto Cu
2+
sites in the
large cavities is retained upon room temperature evacuation, as
the interaction of NO with these cations is much stronger than
with those located in the six member rings.
4. The interaction of O
2
with adsorbed NOon annealed Cu-SSZ-13
As we have discussed above for the adsorption of NO on annealed
Cu-SSZ-13, the stabilities of the adsorbed NO
x
species varied
dramatically. Notably, a short evacuation at room temperature
resulted in the complete removal of NO adsorbed onto either Cu
+
or Cu
2+
(II), while the thermal stabilities of Cu
2+
(I)-adsorbed NO and
NO
+
species were much higher. Among all of the NO
x
species
observed spectroscopically upon the adsorption of NO, NO
+
species
were the most thermally stable. Oxygen has been shown to interact
readily with adsorbed NO over other zeolites, resulting in both the
oxidation of Cu
+
to Cu
2+
, and the formation and further reaction of
NO
2
. For example, in a solid state ion exchanged Cu-ZSM-5 sample,
the addition of O
2
after NO exposure resulted in the fast consump-
tion of Cu
+
-bound NO (and oxidation of Cu
+
to Cu
2+
), and the
formation of Cu
2+
NOspecies, together with the intensification of
all the nitrate-related bands in the 14001700 cm
1
spectral region.
In this experiment, we first introduced a given amount of NO onto
the annealed sample, and followed the variation of the IR peak
intensities for 60 min, and then, without evacuating the IR cell,
introduced O
2
into the system and collected IR spectra as a
function of time. The bottom spectrum in Fig. 9 was collected after
the sample was exposed to NO for 60 min. This spectrum is very
similar to those we have shown above for the adsorption of NO on
Cu-SSZ-13 samples. Dramatic changes in the IR spectra are
observed after the introduction of the same amount of O
2
as the
initial amount of NO. During the first 40 s after O
2
addition (first
black spectrum from the bottom) the intensity of the Cu
+
-bound
NO (B1810 cm
1
) dropped by about 75%. The intensity of the
Cu
2+
NO(II) species (1948 + 1935 cm
1
) also decreased, though
much less than that of the Cu
+
NO. At the same time, the
intensities of both the NO
+
and Cu
2+
NO(I) species increased.
As the intensity of the NO
+
species increased in the entire time
interval studied, that of the Cu
2+
NO(II) species continuously
decreased. The intensity of the Cu
2+
NO(I) species increased
initially as O
2
was introduced into the IR cell, then it leveled o
before decreasing slightly at long exposure times. These observa-
tions can be explained by the oxidation of Cu
+
ions to Cu
2+
, the
movement of Cu ions from the six member ring into the large
cavities, and the formation of NO
3

ions. The fast oxidation of the


Cu
+
ion as O
2
is introduced is evidenced by the rapid decrease and
disappearance of the 1809 cm
1
band characteristic of the NO
vibration of Cu
+
-bound NO molecules. This process contributes to
the increase in the intensity of the 18931914 cm
1
band, repre-
senting NO molecules adsorbed onto Cu
2+
ions located in the large
cavities of the CHA structure. At the same time, some of the Cu
2+
ions that originally occupy cationic sites in the six member rings
also move out from their positions and reposition into the large
cavities, providing additional intensity to the Cu
2+
NO(I)
feature. Eventually, however, the intensity of this peak starts slowly
decreasing as the supply of Cu
2+
(I) sites diminish, and some of
these Cu-ions bind to nitrates originating from the disproportiona-
tion of NO
2
molecules. The available Cu
2+
can also react with NO
and, thus, contribute to the continuous intensity gain of the NO
+
ions via a mechanism discussed above. The results show that least
stable adsorbed NO species (Cu
+
NO and Cu
2+
NO(II)) are the
most susceptible toward reaction with O
2
. The Cu
2+
NO(I) species
do not seem to be directly aected by the presence of O
2
; however,
as NO
2
forms due to the NO+ O
2
reaction, the concentration of this
species may decrease due to the formation of Cu-nitrates.
Conclusions
In the work presented here, in situ IR spectroscopy (both
transmission and diuse reflectance) was used to characterize
the adsorption properties of Cu ions present in Cu-SSZ-13
zeolite catalysts. Adsorption of both CO and NO showed that,
Fig. 9 A series of selected IR spectra recorded during the NO + O
2
reaction over
an annealed Cu-SSZ-13 sample at 300 K.
PCCP Paper
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online
This journal is c the Owner Societies 2013 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 2379
under the conditions our transmission FTIR studies were conducted
(batch, static experiments, the sample always evacuated in high
vacuum after annealing, reduction and oxidation), both Cu
+
and
Cu
2+
ions were present. The Cu
+
/Cu
2+
cation ratio depended on the
pretreatment conditions: it was the highest after reduction with H
2
and the lowest following oxidation with O
2
. The adsorption of CO at
295 K resulted primarily in the formation of Cu
+
-bound mono-
carbonyls; Cu
+
-dicarbonyls and Cu
2+
CO species were formed only
in the presence of gas phase CO over the catalyst. The thermal
stabilities of these latter species were low, as they were completely
removed by a short evacuation. The adsorption of NOon the Cu-SSZ-
13 samples pretreated in the static IR cell showed the formation of
three types of species: Cu
+
-bound mono-nitrosyls, Cu
2+
-adsorbed NO
(two dierent types), and NO
+
in charge compensating cation
positions. The assignment of the adsorbed species was aided by
experiments carried out with isotopically labeled
15
NO molecules.
Under continuous flow conditions (in the DRIFT cell), Cu-SSZ-13
samples containing exclusively Cu
2+
or Cu
+
sites could be prepared
by oxidation (with O
2
) or reduction (by H
2
), respectively. NO adsorp-
tion experiments on these catalysts containing only one type of Cu
ions helped to clarify the chemistry occurring during NOexposure of
these materials. On the fully reduced sample (Cu
+
-SSZ-13), NO
decomposition to N
2
O + O takes place initially on Cu
+
sites, while
no NO
+
formation is observed. The Cu
2+
ions formed in this process
locate primarily in the six member rings of the CHA framework. The
formation of NO
+
only begins when Cu
2+
sites in the six-member
ring position reach high concentration. Prolonged interaction of NO
with the fully reduced sample ultimately leads to the complete
disappearance of the Cu
+
species, and to the migration of Cu
2+
ions
from the six member rings to the large cavities of the CHA structure.
On the fully oxidized sample (Cu
2+
-SSZ-13), a very small amount of
Cu
+
ions form transiently, and most of the Cu
2+
ions binding NO
molecules are present in the large cavities of the framework.
Addition of H
2
O onto the annealed sample prior to NO adsorption
also results inthe movement of Cu
2+
ions fromthe six-member rings
to large cavities, as evidenced by the very high initial intensity of the
Cu
2+
NO IR band representing this adsorbed species.
Acknowledgements
Financial support was provided by the US Department of Energy
(DOE), Oce of Energy Eciency and Renewable Energy, Vehicle
Technologies Program. This work was performed in the Environ-
mental Molecular Sciences Laboratory (EMSL) at Pacific Northwest
National Laboratory (PNNL). The EMSL is a national scientific user
facility supported by the US DOE, Oce of Biological and Environ-
mental Research. PNNL is a multi-program national laboratory
operated for the US DOE by Battelle.
References
1 M. Iwamoto, H. Furukawa, Y. Mine, F. Uemura, S. Mikuriya
and S. Kagawa, J. Chem. Soc., Chem. Commun., 1986,
12721273.
2 A. V. Kucherov, A. A. Slinkin, S. S. Goryashenko and
K. I. Slovetskaja, J. Catal., 1989, 118, 459465.
3 M. Iwamoto, H. Yahiro, N. Mizuno, K. Tanda, Y. Mine and
S. Kagawa, J. Phys. Chem., 1991, 95, 37273730.
4 M. Iwamoto, H. Yahiro, N. Mizuno, W. Zang, Y. Mine,
H. Furukawa and S. Kagawa, J. Phys. Chem., 1992, 96, 93609366.
5 Y. Li and W. K. Hall, J. Catal., 1991, 129, 202215 and
J. Phys. Chem., 1990, 94, 61456148.
6 W. K. Hall and J. Valyon, Catal. Lett., 1992, 15, 311315.
7 G. Spoto, S. Bordiga, D. Scarano and A. Zecchina, Catal.
Lett., 1992, 13, 3944.
8 J. Sa rka ny, J. L. DItri and W. M. H. Sachtler, Catal. Lett.,
1992, 16, 241249.
9 D. J. Parrillo, D. Dolenec, R. J. Gorte and R. W. McCabe,
J. Catal., 1993, 142, 708718.
10 G. Spoto, A. Zecchina, S. Bordiga, G. Ricchiardi, G. Martra,
G. Leofanti and G. Petrini, Appl. Catal., B, 1994, 3, 151172.
11 H.-J. Jang, W. K. Hall and J. L. dItri, J. Phys. Chem., 1996,
100, 94169420.
12 M. Shelef, Chem. Rev., 1995, 95, 209225.
13 P. Ciambelli, P. Corbo, M. Gambino, G. Minelli, G. Moretti
and P. Porta, Catal. Today, 1995, 26, 3339.
14 T. Sano, N. Yamashita, Y. Iwami, K. Takeda and
Y. Kawakami, Zeolites, 1996, 16, 258264.
15 T. Beutel, J. Sa rka ny, G.-D. Lei, J. Y. Yan and W. M. H.
Sachtler, J. Phys. Chem., 1996, 100, 845851.
16 G. Moretti, G. Minelli, P. Porta, P. Ciambelli, P. Corbo,
M. Gambino, F. Migliardini and S. Iacoponi, Stud. Surf.
Sci. Catal., 1997, 105, 15251532.
17 V. I. Pa rvulescu, P. Grange and B. Delmon, Catal. Today,
1998, 46, 233316.
18 L. Olsson, H. Sjovall and R. J. Blint, Appl. Catal., B, 2009, 87,
200210.
19 P. Vanelderen, R. G. Hadt, P. J. Smeets, E. I. Solomon,
R. A. Schoonheydt and B. F. Sels, J. Catal., 2011, 284,
157164.
20 G. Fierro, G. Ferraris and G. Moretti, Appl. Catal., B, 2009,
91, 499506.
21 G. Moretti, G. Ferraris, G. Fierro, M. Lo Jacono, S. Morpurgo
and M. Faticanti, J. Catal., 2005, 232, 476487.
22 M. H. Groothaert, J. A. Van Bokhoven, A. A. Battiston,
B. M. Weckhuysen and R. A. Schoonheydt, J. Am. Chem.
Soc., 2003, 125, 76297640.
23 M. H. Groothaert, K. Lievens, H. Leeman, B. M. Weckhuysen
and R. A. Schoonheydt, J. Catal., 2003, 220, 500512.
24 M. H. Groothaert, K. Lievens, J. A. van Bokhoven,
A. A. Battiston, B. M. Weckhuysen, K. Pierloot and
R. A. Schoonheydt, ChemPhysChem, 2003, 4, 626630.
25 M. Shelef, Catal. Lett., 1992, 15, 305310.
26 A. V. Kucherov, J. L. Gerlock, H.-W. Jen and M. Shelef,
J. Phys. Chem., 1994, 98, 48924894.
27 D.-J. Liu and H. Robota, Catal. Lett., 1993, 21, 291301.
28 J. H. Kwak, R. G. Tonkyn, D. H. Kim, J. Szanyi and C. H. F.
Peden, J. Catal., 2012, 275, 187190.
29 D. W. Fickel, E. DAddio, J. A. Lauterbach and R. F. Lobo,
Appl. Catal., B, 2011, 102, 441448.
30 J. H. Kwak, D. Tran, J. Szanyi, C. H. F. Peden and J. H. Lee,
Catal. Lett., 2012, 142, 295301.
Paper PCCP
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online
2380 Phys. Chem. Chem. Phys., 2013, 15, 2368--2380 This journal is c the Owner Societies 2013
31 J. H. Kwak, D. Tran, S. D. Burton, J. Szanyi, J. H. Lee and
C. H. F. Peden, J. Catal., 2012, 287, 203209.
32 J.-S. McEwen, T. Anggara, W. F. Schneider, V. F. Kispersky,
J. T. Miller, W. N. Delgass and F. H. Ribeiro, Catal. Today,
2012, 184, 129144.
33 W. Fickel and R. F. Lobo, J. Phys. Chem. C, 2010, 114, 1633.
34 S. T. Korhonen, D. W. Fickel, R. F. Lobo, B. M. Weckhuysen
and A. M. Beale, Chem. Commun., 2011, 47, 800.
35 J. H. Kwak, H. Zhu, J. H. Lee, C. H. F. Peden and J. Szanyi,
Chem. Commun., 2012, 48, 47584760.
36 Z. Nour, H. Petitjean and D. Bartomieu, J. Phys. Chem. C,
2010, 114, 17802.
37 J. Sarkany and W. M. H. Sachtler, Zeolites, 1994, 14, 7.
38 K. Hadjiivanov and G. Vayssilov, Adv. Catal., 2002, 47, 307.
39 K. Hadjiivanov, Catal. Rev. Sci. Eng., 2000, 42, 71.
40 K. Hadjiivanov and H. Kno zinger, Surf. Sci., 2009, 603,
16291636.
41 K. Hadjiivanov and H. Knozinger, J. Catal., 2000, 191,
480485.
42 C. Lamberti, S. Bordiga, M. Salvalaggio, G. Spoto,
A. Zecchina, F. Geobaldo, G. Vlaic and M. Bellatreccia,
J. Phys. Chem. B, 1997, 101, 344.
43 G. T. Turnes Palomino, S. Bordiga, A. Zecchina, G. L. Marra
and C. Lamberti, J. Phys. Chem. B, 2000, 104, 86418651.
44 R. Bulanek, K. Frolich, P. Cicmanec, D. Nachtigallova,
A. Pulido and P. Nachtigall, J. Phys. Chem. C, 2011,
115, 13312.
45 X. Zhang, Q. Shen, C. He, C. Ma, J. Cheng, L. Li and Z. Hao,
ACS Catal., 2012, 2, 512520.
46 M. Occhiuzzi, G. Fierro, G. Ferraris and G. Moretti, Chem.
Mater., 2012, 24, 20222031.
47 J. Szanyi and M. Paett, J. Catal., 1996, 164, 232245.
48 K. Hadjiivanov, J. Saussey, J. L. Freysz and J. C. Lavalley,
Catal. Lett., 1998, 52, 103108.
PCCP Paper
P
u
b
l
i
s
h
e
d

o
n

1
1

D
e
c
e
m
b
e
r

2
0
1
2
.

D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V
E
R
S
I
T


D
E
G
L
I

S
T
U
D
I

D
I

T
O
R
I
N
O

o
n

2
1
/
0
1
/
2
0
1
4

1
4
:
1
3
:
3
2
.

View Article Online

Вам также может понравиться