You are on page 1of 7

Microporous and Mesoporous Materials 116 (2008) 586–592

Contents lists available at ScienceDirect

Microporous and Mesoporous Materials


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

Facile synthesis of Ag–OMS-2 nanorods and their catalytic applications


in CO oxidation
Junli Chen, Juan Li, Huaju Li, Xiumin Huang, Wenjie Shen *
State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese Academy of Sciences, 457 Zhongshan Road, Dalian 116023, China

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

Article history: Ag–OMS-2 nanorods were synthesized by a simple solid-state reaction of AgMnO4 with manganese ace-
Received 3 April 2008 tate. The nanorods had a diameter of about 8 nm and lengths of 50–150 nm with the co-existence of
Received in revised form 17 May 2008 micropores and mesopores. Structural analysis revealed that the nanorods were amorphous but exhibited
Accepted 22 May 2008
lattice fringes of Ag–hollandite structure. The silver species presented as Ag+ and the average oxidation
Available online 3 July 2008
state of manganese was 3.7. Most importantly, the Ag–OMS-2 nanorods showed quite high catalytic per-
formance for CO oxidation with 100% CO conversion at 90 °C, mainly because of the significantly
Keywords:
enhanced oxygen activation and CO adsorption with the presence of Ag+ in the tunnels.
Ag–OMS-2
Nanorods
Ó 2008 Elsevier Inc. All rights reserved.
CO oxidation

1. Introduction 100–200 nm with a relatively higher surface area of 35 m2/g. This


Ag–hollandite acted as excellent SO2 absorbent at low temperature
Manganese oxide octahedral molecular sieves (OMS) having (150 °C) and was also highly active for CO and NO oxidations. Gac
one-dimensional framework and variable tunnel size have been [6] has recently synthesized Ag–OMS-2 by adding silver nitrate to
extensively studied because of their interesting properties in het- the initial solution for a hydrothermal process, through which sil-
erogeneous catalysis and rechargeable battery. The typical K– ver partially replaced potassium in the K–OMS-2 tunnels. The ob-
OMS-2 consists of 2  2 edge-shared double MnO6 octahedral tained material had a high surface area of 293 m2/g, but the
chains, which are corner-connected to form one-dimensional tun- tunnel structure was slightly collapsed. Hu et al. [7] doped K–
nel having a size of 0.46 nm  0.46 nm with the presence of K+ for OMS-2 with Ag+ by a reflux method, and the resulting Ag–OMS-2
charge balance [1–3]. In order to improve the electronic and cata- had nanorod-shaped morphology with a diameter of 20 nm and
lytic properties of this hollandite-type manganese oxide, alkali lengths of 500–600 nm and exhibited significantly enhanced activ-
and/or transition metal cations were used to replace K+ in the tun- ity for CO oxidation. However, these conventional synthetic ap-
nel mainly by post-ion-exchange. For example, substitution of K+ proaches could replace K+ only partially, which would limit the
by Ag+ in the K–OMS-2 materials resulted in the formation of effective utilization of silver species as active sites when used for
Ag–hollandite (Ag–OMS-2), which was more promising as environ- catalytic reactions.
mentally benign and efficient oxidation catalyst [4–7]. However, We have recently synthesized pure Ag–hollandite nanofibers by
the structure of Ag–hollandite was slightly modified in which the a hydrothermal process, and the Ag–OMS-2 materials were highly
silver cation did not occupy the center of the cubic cages formed active and selective for ethanol oxidation to acetaldehyde [8]. In
by MnO6 octahedra, but located on the common face of the cube the present work, we synthesized pure Ag–OMS-2 nanorods with
which coordinated with four oxygen anions [4]. a diameter about 8 nm and a length up to 150 nm by a simple so-
Chang and Jansen [4] originally synthesized Ag–hollandite lid-state reaction of AgMnO4 with manganese acetate. The Ag–
materials by a solid-state reaction of AgMnO4 and Ag2O at 970 °C OMS-2 nanorods having both mesopores and micropores were
under a very high pressure of oxygen for several days. The needle- highly active for CO oxidation at low temperatures.
like crystal had very large dimensions of 0.1 mm  0.4 mm 
1.0 mm, and thereby very low surface area. Li and King [5] synthe- 2. Experimental
sized Ag–hollandite by ion-exchange of K–OMS-2 with AgNO3, and
the resulting material had a width of about 30 nm and lengths of 2.1. Synthesis of Ag–Mn oxides

* Corresponding author. Tel.: +86 411 84379085; fax: +86 411 84694447. AgMnO4 was prepared by the reaction of KMnO4 with AgNO3 in
E-mail address: shen98@dicp.ac.cn (W. Shen). aqueous solution as described elsewhere [8]. Ag–Mn oxides were

1387-1811/$ - see front matter Ó 2008 Elsevier Inc. All rights reserved.
doi:10.1016/j.micromeso.2008.05.029
J. Chen et al. / Microporous and Mesoporous Materials 116 (2008) 586–592 587

synthesized by a solid-state reaction route; 2.27 g of AgMnO4 and (40–60 mesh) were loaded and sandwiched by two layers of quartz
3.68 g of Mn(Ac)2  4H2O, equaling to a stoichiometric molar ratio wool in the reactor. The feed gas consisted of 1.0 vol% CO/2.0 vol%
of 2:3, were ground homogeneously in a mortar. The mixed pow- O2/He (30 mL/min). The effluents from the reactor were analyzed
ders were then transferred to a capped glass bottle and maintained by an on-line gas chromatograph (Varian 3800) equipped with a
at 100 °C for 4 h. The resulting black powder was washed with TDX-01 and Porapak Q packed column which connected to a TCD.
water thoroughly and dried at 100 °C overnight. The solid obtained
was designed as Ag–Mn-100. The Ag–Mn-100 sample was further 3. Results and discussion
calcined at 300 and 800 °C for 4 h in air, and the samples were de-
noted as Ag–Mn-300 and Ag–Mn-800, respectively. 3.1. Structure and morphology

2.2. Characterization Fig. 1 shows the XRD patterns of the Ag–Mn oxides. The Ag–Mn-
100 and Ag–Mn-300 samples showed diffraction peaks at d-spacings
The X-ray powder diffraction (XRD) patterns were recorded of 3.04 and 2.41 Å, representing the crystal structure of Ag–OMS-2.
with a D/Max-2500/PC powder diffractometer (Rigaku, Japan) However, these materials are poorly crystallized when compared
operated at 40 kV and 100 mA, using nickel-filtered Cu Ka with the well-defined hollandite structure of K–OMS-2 that was
(k = 0.15418 nm) radiation source. produced by a solid-state reaction of KMnO4 and manganese acetate
The high-resolution transmission electron microscopy (HRTEM) tetrahydrate [9–10]. It was reported that the hollandite structure of
images were taken on a Philips Fei Tecnai G220 microscope operat- K–OMS-2 disappeared entirely when NaMnO4 or Mg(MnO4)2 was
ing at 300 kV. The samples were ultrasonically dispersed into eth- used to replace KMnO4, and only c-MnO2 was formed [9]. Since K+
anol, and a drop of the suspension was deposited on a standard (0.28 nm) is able to satisfy the ionic contact distance to the nearby
copper grid and then dried in air. The field emission scanning elec- oxygen atoms by occupying the special position (0, 0, 0) in the unit
tron microscopy (FESEM) images were taken using a Philips Fei cell of K–hollandite, substitution of K+ by small cations like Na+
Quanta 200F instrument. The powder sample was dispersed on a (0.19 nm) and Mg2+ (0.13 nm) would cause tunnel distortion, and fi-
conductive adhesive tape. nally destroy the hollandite structure [11]. When Ag+ (0.24 nm) is
The N2 adsorption–desorption isotherms were recorded at used to replace K+, however, the hollandite structure can be main-
196 °C on a Micromeritics ASAP 2010 instrument. Before the tained because the ionic radius of Ag+ is close to that of K+ and the
measurements, the samples were degassed at 300 °C for 8 h. The Ag+ could locate on the position (0, 0, 1/2). Although the silver spe-
specific surface areas (SBET) were calculated by Braunauer–Em- cies did not occupy the centers of the cubic cages formed by MnO6
mett–Teller method, and the size distributions of micropores and octahedra, they located on the common faces of the cubes which
mesopores were determined by the H–K and BJH methods, coordinated with four oxygen anions [4]. Consequently, the hollan-
respectively. dite structure of OMS-2 was maintained in the Ag–Mn-100 and Ag–
The Fourier transform infrared (FT-IR) spectroscopy analysis Mn-300 samples, but not well-defined. For the Ag–Mn-800 sample,
was performed with a spectrometer (Brucker Vector 22) in the there was no diffraction peaks of hollandite structure, while charac-
range of 1000–400 cm 1 with a resolution of 4 cm 1. The samples teristic diffraction peaks of Mn2O3 (JCPDS 24-0508) and Ag (JCPDS
were grounded to fine powders and dispersed in KBr to compress 65-2871) were observed. This indicates that the high temperature
into pellets. calcination caused entire collapse of the OMS-2 structure.
The X-ray photoelectron spectroscopy (XPS) was recorded on a Fig. 2 shows the FESEM and TEM images of the Ag–Mn oxides.
KROTAS AMICAS spectrometer (Shimadzu, Japan) using Mg Ka The Ag–Mn-100 and Ag–Mn-300 samples exhibited nanorod-
(k = 1253.6 eV) radiation source. The powder samples were pressed shaped morphology with a diameter of about 8 nm and lengths
into thin discs and mounted on a sample rod placed in the analysis of 50–150 nm. HRTEM images of the Ag–Mn-300 sample clearly
chamber where the spectra of Ag 3d, Ag MVV, Mn 2p and Mn 3s showed spacings of two lattice fringes of 0.30 and 0.23 nm, corre-
levels were recorded. Charge effect was corrected by adjusting sponding to the {1 3 0} and {3 3 0} planes, respectively. However,
the binding energy of C1s to 285.0 eV. the FFT profiles of the selected areas exhibited featureless streaks,
The temperature-programmed reduction (TPR) measurements implying structural distortion [12–13]. FESEM images of the Ag–
were conducted in a conventional setup equipped with a thermal Mn-800 sample showed spherical Mn2O3 particles of 50–100 nm
conductivity detector (TCD). In a typical run, 20 mg of samples and very large silver particles. The EDS spectra of selected areas
were loaded and gradually heated to 300 °C (10 °C /min) with N2 indicated that large Ag and spherical Mn2O3 particles separated
flowing (40 mL/min) and maintained at this temperature for and presented as two single phases.
0.5 h. After cooling to room temperature and introducing the
reduction agent of a 5 vol% H2/N2 mixture (40 mL/min), the sam-
ples were heated to 900 °C at a rate of 10 °C /min.
The temperature-programmed desorption (TPD) measurements * * Mn O 2 3

were carried out in a fixed-bed quartz tube reactor equipped with a # # Ag


quadrupole mass spectrometer (Omnistar, Balzers). Fifty milligram
Intensity (a.u.)

of samples were loaded and heated to 300 °C with a He flow and Ag-Mn-800

kept at this temperature for 0.5 h. After cooling to room tempera- #


* # #
ture, CO adsorption was done by passing a 5 vol% CO/He mixture * * * *
over the sample for 1 h, and the sample was purged with He for * * **
2 h. The system was then heated to 900 °C at a rate of 10 °C /min
Ag-Mn-300
by flowing a He stream. The desorbed species were monitored by
the mass spectrometer. Ag-Mn-100

2.3. Catalytic evaluation


10 20 30 40 50 60 70 80
2 theta (degree)
CO oxidation was conducted in a continuous-flow fixed-bed
quartz reactor under atmospheric pressure; 200 mg of catalyst Fig. 1. XRD patterns of the Ag–Mn oxides.
588 J. Chen et al. / Microporous and Mesoporous Materials 116 (2008) 586–592

Fig. 2. SEM and TEM images of the Ag–Mn-100 (a, b), Ag–Mn-300 (c–f) and Ag–Mn-800 (g, h) samples.

3.2. Texture properties cating the non-porous nature of this material. The Ag–OMS-2
nanorods showed Type I adsorption isotherms at low P/P0, suggest-
Fig. 3a shows the N2 adsorption–desorption isotherms of the ing the existence of micropores [14]. The hysteretic loops at high P/
samples. There was no detectable micropore filling at low P/P0 P0 were typical Type H3 adsorption isotherms, indicating the pres-
and/or hysteretic loop at high P/P0 for the Ag–Mn-800 sample, indi- ence of slit-shaped pores, probably due to the aggregation of plate-
J. Chen et al. / Microporous and Mesoporous Materials 116 (2008) 586–592 589

Table 1
a Textural properties of the Ag–Mn oxides
Volume adsorbed (cm3/g (STP))

Sample SBET Smicro Sext Vmicro Vtotal


(m2/g) (m2/g) (m2/g) (10 3 cm3/g) (10 3
cm3/g)
Ag–Mn-100 158 18 140 8.4 474
Ag–Mn-300 156 13 143 5.7 483
Ag-Mn-100 Ag–Mn-800 3.3 – 3.3 – 18

Ag-Mn-300
tunnels presented in the Ag–Mn-100 sample disappeared, proba-
bly being reconstructed into 2  2 tunnels during the calcination
Ag-Mn-800
at 300 °C. The mesopores had a mean diameter of 9.8 nm. Hence,
the Ag–OMS-2 nanorods had both mesoporous and microporous
0.0 0.2 0.4 0.6 0.8 1.0
structures.
Relative pressure (P/P0)
Table 1 summarizes the textural properties of the Ag–Mn oxi-
des. The BET surface area of the Ag–OMS-2 nanorods was 156–
0.025 0.004
158 m2/g, which was mainly contributed by the external surface.
b The micropore volume only contributed to 1.8% (Ag–Mn-100)
0.020 and 1.2% (Ag–Mn-300) of the total pore volume, mainly due to
0.003
the unique structure of OMS-2 materials. Although the hollan-
dV/dD (cm3/g)

dV/dD (cm3/g)
dite-type manganese oxides has a crystallite pore opening of about
0.015
0.46 nm, the effective pore opening might be in a range of 0.265–
0.002
0.333 nm due to the presence of cations in the tunnels [18–19].
0.010 Therefore, it is highly possible that only small fraction of nitrogen
with dynamic diameter of 0.364 nm could entry into the tunnels of
0.001 the Ag–OMS-2 nanorods, resulting in less-estimated contribution
0.005
of the micropores. For the Ag–Mn-800 sample without either
micropores or mesopores, the BET surface was only 3.3 m2/g, as
0.000 0.000 expected.
0.4 0.6 0.8 1.0 1.2 10 20 30 40 50 60
Pore diameter (nm)

0.025 0.004
c Ag-Mn-800
611
579
0.020
0.003 528
dV/dD (cm3/g)
dV/dD (cm3/g)

0.015 670
0.002
0.010

0.001
0.005
Absorbance (a.u.)

0.000 0.000
0.4 0.6 0.8 1.0 1.2 10 20 30 40 50 60
Pore diameter (nm)

Fig. 3. N2 adsorption–desorption isotherms of the Ag–Mn oxides (a) and pore size
distributions of Ag–Mn-100 (b) and Ag–Mn-300 (c) nanorods. Ag-Mn-300

like particles [15]. Fig. 3b shows the pore size distribution of the
Ag–Mn-100 sample. There were two types of micropores with 525 556
mean sizes of 0.50 and 0.59 nm, respectively. The former is close
433 458
to the size (0.46 nm) of 2  2 tunnel in OMS-2 structure [16], while 607
Ag-Mn-100
the later represents the 2  3 tunnels [17]. In other words, both
2  2 and 2  3 tunnels were formed in the Ag–Mn-100 sample, 420
implying that the calcination temperature of 100 °C is not suffi- 744
cient enough to produce well-ordered 2  2 tunnels. At high P/P0,
the average pore size of 9.6 nm indicated the existence of mesopor-
ous structure, which could be due to the interparticle voids of the 400 500 600 700 800
aggregated nanorods [9]. Fig. 3c shows the pore size distribution of Wavenumber (cm-1)
the Ag–Mn-300 sample. The micropores had a mean size of
0.51 nm, representing the well-ordered 2  2 tunnels. The 2  3 Fig. 4. FT-IR spectra of the Ag–Mn oxides.
590 J. Chen et al. / Microporous and Mesoporous Materials 116 (2008) 586–592

3.3. FT-IR and XPS surface analysis Mn 3s in the Ag–OMS-2 nanorods was 4.7 eV, from which the
AOS of manganese was estimated to be 3.7. That is, Mn4+ is domi-
Fig. 4 shows the FT-IR spectra of the Ag–Mn oxides. For typical nantly present in the Ag–OMS-2 nanorods.
OMS-2 materials, the absorption bands of Mn–O lattice vibration in Fig. 5c shows the core-level XP spectra of Ag 3d in the Ag–Mn
the framework usually appear in 400–800 cm-1 [20]. IR spectrum oxides. For the Ag–OMS-2 nanorods, the bonding energies of Ag
of K–OMS-2-2 consists of absorption bands at 314, 392, 463, 530, 3d5/2 and Ag 3d3/2 were 367.8 eV and 373.8 eV, respectively, char-
579 and 710 cm-1 [20,21], but the position and the intensity might acteristics of Ag+ [26]. While for the Ag–Mn-800 sample, the bind-
be slightly varied when K+ is substituted by Ag+. Gac [6] attributed ing energy of Ag 3d5/2 shifted to 368.3 eV, corresponding to
the bands at 714, 600, 520, 460 and 410 cm 1 to Ag–K–OMS-2 metallic silver [27]. Fig. 5d shows the Ag MVV spectra, from which
structure. Here, the IR spectra of the Ag–OMS-2 nanorods had se- the Auger parameters were estimated to be 724.4 eV for the Ag–
ven bands at 420, 433, 458, 525, 556, 607 and 744 cm 1. The bands OMS-2 nanorods and 726.0 eV for the Ag–Mn-800 sample, respec-
at 458, 525 and 607 cm 1 corresponded to the hollandite structure tively. Because the Auger parameters of Ag0 and Ag+ are 726.0 eV
in the Ag–OMS-2 nanorods, and the bands at 433 and 744 cm 1 re- and 724.5 eV [28], respectively, it is clear that the silver species
ferred to the romanechite structure (2  3) [21], which is more in the Ag–OMS-2 nanorods is Ag+, while Ag0 is solely present in
pronounced in the Ag–Mn-100 sample. The IR spectrum of the the Ag–Mn-800 sample. This is in good agreement with the XRD
Ag–Mn-800 sample simply showed absorption bands at 528, 579, result, where metallic silver phase was detected in the Ag–Mn-
611 and 670 cm 1, being characteristic of a-Mn2O3 bixbyite [21]. 800 sample.
Fig. 5a displays the XP spectra of Mn 2p in the Ag–Mn oxides.
The binding energies of Mn 2p3/2 and Mn 2p1/2 in the Ag–OMS-2 3.4. H2-TPR
nanorods were 642.3 eV and 654.0 eV, respectively, which could
be simply attributed to a mixture of Mn4+ and Mn3+ species [22]. Fig. 6 shows the H2-TPR profiles of the Ag–Mn oxides. Generally,
While the binding energies of Mn 2p3/2 and Mn 2p1/2 in the Ag– the presence of silver in Ag–OMS-2 materials could significantly
Mn-800 sample was 641.5 eV and 653.3 eV, indicating the solely promote the reducibility of manganese oxide through a hydrogen
presence of Mn3+ species [23]. Since Mn2+, Mn3+ and Mn4+ have spillover effect, and the reduction temperature is shifted to much
essentially the same binding energy [24], it is difficult to identify lower region [29,30]. The Ag–OMS-2 nanorods exhibited hydrogen
the oxidation state of manganese only by the binding energy of consumptions at 109 and 140 °C. The former is probably due to a
Mn 2p. The XPS of Mn 3s is more promising for verifying manga- combined reduction of Ag+ to Ag (AgO + H2 ? Ag + H2O) and
nese oxidation state because the magnitude of the Mn 3s splitting MnO2 to Mn3O4 (3MnO2 + 2H2 ? Mn3O4 + 2H2O), and the later cor-
decreased monotonously with increasing the average oxidation responds to the further reduction of Mn3O4 to MnO
state (AOS) of manganese [25]. As shown in Fig. 5b, the binding en- (Mn3O4 + H2 ? 3MnO + H2O) [30]. It is also apparent that the
ergy difference (DE3s) between the main peak and its satellite of hydrogen consumption at 109 °C for the Ag–Mn-100 sample was

3d 5/2
a 2p1/2 2p3/2 Mn 2p c 3d 3/2
368.3 eV
Ag 3d

653.3 eV 374.4 eV
641.5 eV Ag-Mn-800
Intensity (a.u.)

Ag-Mn-800
Intensity (a.u.)

654.0 eV 367.8 eV
642.3 eV 373.8eV
Ag-Mn-300

Ag-Mn-100 Ag-Mn-300

Ag-Mn-100

660 655 650 645 640 635 380 375 370 365 360
Binding Energy (eV) Bind energy (eV)

Ag MVV
b Mn 3s
d 351.7 eV
4.7 eV M4VV
Intensity (a.u.)

350.6 eV
Intensity (a.u.)

Ag-Mn-800

Ag-Mn-300

Ag-Mn-300

Ag-Mn-100 Ag-Mn-100

92 90 88 86 84 82 80 78 340 345 350 355 360 365


Binding Energy (eV) Kinetic Energy (eV)

Fig. 5. XP spectra of Mn 2p (a), Mn 3s (b), Ag 3d (c), and Ag MVV (d) in the Ag–Mn oxides.
J. Chen et al. / Microporous and Mesoporous Materials 116 (2008) 586–592 591

a 665
Ag-Mn-800 Ag-Mn-800
Intensity (a.u.)

Ag-Mn-300
Ag-Mn-300
87
163
469
535

Intensity (a.u.)
Ag-Mn-100

Ag-Mn-100
b Ag-Mn-800

100 200 300 400 500 Ag-Mn-300


Temperature (ºC)
Ag-Mn-100
615
Fig. 6. H2-TPR profiles of the Ag–Mn oxides.

550

larger than that of the Ag–Mn-300 sample, while the hydrogen Ag-Mn-100-He
consumption at 140 °C showed an opposite trend. This phenome-
non indicates that the reduction of the Ag–Mn-100 sample is easier 100 200 300 400 500 600 700 800 900
than the Ag–Mn-300 sample, probably because of the co-existence Temperature (ºC)
of 2  2 and 2  3 tunnels. The total hydrogen consumption
Fig. 7. CO-TPD profiles of the Ag–Mn oxides: (a) CO2 desorption and (b) O2
amounts estimated from the reduction peaks further confirmed desorption.
this stepwise reduction of Ag–OMS-2 nanorods. Gac [6] found that
Ag+ species could promote the reduction of manganese oxide,
strongly depending on the location of silver species. Silver species,
which located in the channels of OMS-2, promoted the reduction of Desorption of CO2 at high temperature region (450–550 °C) for
manganese oxide more significantly than the silver species depos- the Ag–OMS-2 nanorods could be attributed to the surface reaction
ited on the outer surface. Li and King [5] also found that hydrogen of adsorbed CO with manganese oxides. The desorption at 470 °C is
could react with Ag–OMS-2 at a temperature as low as 100 °C, but possibly caused by the reaction of CO with oxygen that produced
the reduction of K–OMS-2-2 could occur only above 250 °C. Appar- from the phase transformation of MnO2 to Mn2O3, while the
ently, the Ag–OMS-2 nanorods lowered considerably the reduction desorption peak at 535 °C corresponds to the further phase trans-
temperature of manganese oxide, in which Ag+ was initially re- formation of Mn2O3 to Mn3O4 [32]. For the Ag–Mn-800 sample,
duced to Ag and the spill over of hydrogen on silver facilitated there was only one major desorption peak at 665 °C, simply due
the reduction of manganese oxides. On the other hand, the Ag– to the reaction of adsorbed CO with lattice oxygen produced by
Mn-800 only showed a broad reduction peak at 350–470 °C. Be- the phase transfer of Mn2O3 to Mn3O4 [32–33].
cause metallic Ag was readily formed and separated from Mn2O3 Fig. 7b shows the desorption of oxygen during the CO-TPD
particles, the hydrogen consumption was simply due to the reduc- experiments, which occurred just after the desorption of CO2. For
tion of Mn2O3. the Ag–OMS-2 nanorods, oxygen desorption was observed at
575 °C, indicating further oxygen release from the manganese oxi-
3.5. CO-TPD des. For comparison, a TPD measurement of the Ag–Mn-100 sam-
ple without CO adsorption was conducted, and oxygen desorption
Fig. 7a shows the CO-TPD profiles of the Ag–Mn oxides. The was detected at 615 °C with a shoulder peak at 550 °C. The
desorption of CO2 is resulted from the surface reaction of CO ad- amounts of oxygen desorbed are about five times larger than that
sorbed with the Ag–Mn oxides. Four distinct CO2 desorption peaks of CO-TPD experiment. This phenomenon suggested that consider-
at 87, 163, 469 and 535 °C were observed for the Ag–OMS-2 nano- able amounts of lattice oxygen participated in the surface reaction
rods. The main desorptions at low temperatures (80–170 °C) could of adsorbed CO. For the Ag–Mn-800 sample, oxygen desorption
be attributed to the adsorption of CO on Ag+ species. It was readily occurred at 790 °C, corresponding to the phase transformation of
confirmed that doping of K–OMS-2 with Ag+ caused synergistic ef- Mn2O3 to Mn3O4 [32].
fect between Ag+ and manganese oxide, creating new sites for CO
adsorption [7]. Ag+ species might exist on the surface of manga- 3.6. Catalytic performance for CO oxidation
nese oxides in addition to occupying the tunnel of OMS-2 [4].
Ravikumar et al. [31] proposed that Ag+ was present in the tunnel Fig. 8 shows the reaction results of CO oxidation over the Ag–
structure with two types of coordination. One is that silver located Mn oxides. The conversion of CO on the Ag–OMS-2 nanorods was
in off-center position in the tunnels of OMS-2, and another type is already 20% at room temperature, and increased rapidly to 100%
that silver occupied the tunnel sites (at the center). These two at 90 °C. But the Ag–Mn-800 sample was almost inactive for CO
types of Ag+ species provide different sites for CO adsorption and oxidation below 165 °C, and the conversion of CO was only 40%
thereby give CO2 desorption at different temperature. Therefore, even at 190 °C. Fig. 9 illustrates the catalytic stability of the Ag–
the desorption of CO2 at 87 °C is assigned to CO adsorbed on silver Mn-300 sample. At 90 °C, the conversion of CO was almost 100%
species located in the surface of manganese oxide, and desorption for the initial 15 h, and then it tended to decrease gradually to
peak at 163 °C is attributed to CO adsorbed Ag+ species in the OMS- 87% at 60 h. When tested at 40 °C, however, the CO conversion de-
2 tunnel. The peak area at 87 °C was only a quarter of that at creased significantly from 53% at the initial stage to 33% at 45 h.
163 °C, implying that Ag+ predominantly occupied the tunnel sites This fast deactivation is probably due to the accumulation of car-
in the Ag–OMS-2 nanorods. bonates at such a low reaction temperature.
592 J. Chen et al. / Microporous and Mesoporous Materials 116 (2008) 586–592

4. Conclusion
100
Ag–OMS-2 nanorods having a diameter of about 8 nm and
80 lengths of 50–150 nm were synthesized by a facile solid-state-
CO conversion (%)

reaction route. The nanorods were amorphous but exhibited lattice


60 Ag-Mn-100
fringes of Ag–hollandite. The Ag–OMS-2 nanorods had mesoporous
Ag-Mn-300
Ag-Mn-800 and microporous structures simultaneously with a higher surface
40 area of 158 m2/g. The tunnel structure and nanorod morphology
favored easy activation of oxygen molecule and effective adsorp-
20 tion of CO, giving high activity for CO oxidation at low
temperatures.
0
References
20 40 60 80 100 120 140 160 180 200
Temperature (ºC) [1] Y.F. Shen, R.P. Serger, R.N. Deguzman, S.L. Suib, L. Mccurdy, D.I. Potter, C.L.
O’Young, Science 260 (1993) 511.
Fig. 8. CO conversions over the Ag–Mn oxides; 1% CO/0.5% O2/He, [2] A. Dyer, M. Pillinger, J. Newton, R. Harjula, T. Möller, S. Amin, Chem. Mater. 12
36,000 ml g 1
h 1. (2000) 3798.
[3] Y.C. Son, V.D. Makwana, A.R. Howell, S.L. Suib, Angew. Chem. Int. Ed. 40 (2001)
4280.
[4] F.M. Chang, M. Jansen, Angew. Chem. Int. Ed. 23 (1984) 906.
[5] L.Y. Li, D.L. King, Chem. Mater. 17 (2005) 4335.
100 [6] W. Gac, Appl. Catal. B 75 (2007) 107.
90 ºC
[7] R.R. Hu, C. Yi, L.Y. Xie, D.Z. Wang, Chin. J. Catal. 28 (2007) 463.
90 [8] J.L. Chen, X.F. Tang, J.L. Liu, E.S. Zhan, J. Li, X.M. Huang, W.J. Shen, Chem. Mater.
19 (2007) 4292.
CO conversion (%)

80 [9] Y.S. Ding, X.F. Shen, S. Sithambaram, S. Gomez, R. Kumar, V.M.B. Crisostomo,
S.L. Suib, M. Aindow, Chem. Mater. 17 (2005) 5382.
70
[10] Q.W. Li, G.A. Luo, J. Li, X.J. Xia, Mater. Proc. Tech. (2003) 13725.
60 [11] J.E. Post, R.B. Von Dreele, P.R. Buseck, Acta. Crystallogr. 38 (1982) 1056.
[12] W.S. Seo, K.J. Koumoto, Am. Ceram. Soc. 79 (1996) 1777.
50 [13] T.D. Xiao, X. Bokhimi, M. Benaissa, R. Pérez, P.R. Strutt, M.J. Yacamán, Acta
Mater. 45 (1997) 1685.
40 ºC [14] S.J. Gregg, K.S.W. Sing, In Adsorption, Surface Area and Porosity, second ed.,
40
Academic Press Inc., London, 1982.
30 [15] K.S.W. Sing, D.H. Evertt, R.A.W. Haul, L. Moscou, R.A. Pierotti, J. Rouquerol, T.
Siemieniewska, Pure Appl. Chem. 57 (1985) 603.
20 [16] J. Liu, V. Makwana, J. Cai, X.F. Shen, S.L. Suib, M. Aindow, J. Phys. Chem. B 107
0 10 20 30 40 50 60 (2003) 9185.
Time on stream (h) [17] X.F. Shen, Y.S. Ding, J. Liu, K. Laubernds, R.P. Zerger, M. Polverejan, Y.C. Son, M.
Aindow, S.L. Suib, Chem. Mater. 16 (2004) 5327.
[18] Z.M. Wang, S. Tezuka, H. Kanoh, Chem. Mater. 13 (2001) 530.
Fig. 9. Stability of the Ag–Mn-300 nanorods for CO oxidation.
[19] C.L. O’Young, R.A. Sawicki, S.L. Suib, Microporous Mater. 11 (1997) 1.
[20] L. Kang, M. Zhang, Z.H. Liu, K. Ooi, Spectrochim. Acta. Part A 67 (2007) 864.
[21] C.M. Julien, M. Massot, C. Poinsignon, Spectrochim. Acta A 60 (2004) 689.
The unique structure of Ag–OMS-2 nanorods might be respon- [22] J.H. Ge, L.H. Zhou, F. Yang, B. Tang, L.Z. Wu, C.H. Tung, J. Phys. Chem. B 110
(2006) 17854.
sible for the higher activity of CO oxidation. It was previously pro-
[23] M. Oku, K. Hirokawa, S. Ikeda, J. Electron. Spectrosc. Rel. Phenom. 7 (1975) 465.
posed that the Ag–O–Mn bridge in Ag–OMS-2 materials facilitated [24] V.D. Castro, G. Polzonetti, J. Electron. Spectrosc. Rel. Phenom. 48 (1989) 117.
the electron transfer between Ag and MnOx and caused the [25] V.R. Galakhov, M. Demeter, S. Bartkowski, M. Neumann, N.A. Ovechkina, E.Z.
bridged oxygen more mobile and active, resulting in high activity Kurmaev, Y.M. Lobachevskaya, J. Mitchell, D.L. Ederer, Phys. Rev. B 65 (2002)
113102.
for CO oxidation [6,34]. Hu et al. [7] also found that doping K– [26] X.Y. Gao, S.Y. Wang, J. Li, Y.X. Zheng, R.J. Zhang, P. Zhou, M.Y. Yang, L.Y. Chen,
OMS-2-2 with Ag+ increased the diffusion of lattice oxygen species Thin Solid Films 455–456 (2004) 438.
and enhanced CO adsorption, favoring CO oxidation. For the Ag– [27] F. Zhang, N. Guan, Y. Li, X. Zhang, J. Chen, H. Zeng, Langmuir 19 (2003) 8230.
[28] C.D. Wagner, Discuss. Faraday Soc. 60 (1975) 291.
OMS-2 nanorods, the facile reduction feature implies the easy acti- [29] A. Machocki, T. Ioannides, B. Stasinska, W. Gac, G. Avgouropoulos, D. Delimaris,
vation and transformation of oxygen species between silver in the W. Grzegorczyk, S. Pasieczna, J. Catal. 227 (2004) 282.
tunnel and manganese oxide. Meanwhile, the Ag+ species around [30] R. Lin, W.P. Liu, Y.J. Zhong, M.F. Luo, Appl. Catal. A 220 (2001) 165.
[31] R. Ravikumar, D.W. Fuerstenau, G.A. Waychunas, Mat. Res. Soc. Symp. Proc.
the tunnel provide effective site for CO adsorption, as demon- 524 (1998) 353.
strated by the CO-TPD measurements. Additionally, the nanorod- [32] X. Chen, Y.F. Shen, S.L. Suib, C.L. O’Young, Chem. Mater. 14 (2002) 940.
shaped morphology also gives much higher surface area, which [33] J.K. Chang, Y.L. Chen, W.T. Tsai, J. Power Sources 135 (2004) 344.
[34] G.G. Xia, Y.G. Yin, W.S. Willis, J.Y. Wang, S.L. Suib, J. Catal. 185 (1999) 91.
in turn, providing more reactive sites for CO oxidation.

You might also like