Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

RSC Advances

View Article Online


PAPER View Journal | View Issue

Covalent triazine-based framework as an efficient


catalyst support for ammonia decomposition
Cite this: RSC Adv., 2015, 5, 3605
Fei Chang,ab Jianping Guo,ab Guotao Wu,a Lin Liu,a Miao Zhang,ab Teng He,a
Peikun Wang,ab Pei Yuab and Ping Chen*ac
Published on 05 December 2014. Downloaded on 3/14/2024 2:51:12 AM.

The covalent triazine-based framework (CTF), a new type of nitrogen-containing microporous polymer,
was employed as a catalyst support for ammonia decomposition. Either in terms of NH3 conversion rate
or turnover frequency, Ru/CTF-1 has a highly enhanced performance compared to Ru/CNTs, which rank
as one of the best un-promoted catalysts reported so far. The compositional and structural information
of Ru/CTF-1 and Ru/CNTs catalysts have been characterized by ICP, N2 physisorption, XRD, TEM, XPS,
and NH3-TPD techniques. Ru particles on CTF-1 and CNTs are ca. 3 nm in diameter and have a similar
degree of dispersion. However, the binding energy of Ru 3p electrons is ca. 0.6 eV less for Ru/CTF-1
than that for Ru/CNTs showing significant increase in electron density in the former, which is likely due
Received 21st October 2014
Accepted 2nd December 2014
to the interaction between the nitrogen-rich groups of CTF-1 and the Ru nanoparticles. Moreover, the
presence of CTF-1 enhances the chemisorption of NH3, which, together with the increased electron
DOI: 10.1039/c4ra12816h
density of Ru, may facilitate the competitive chemisorption of NH3 and recombinative desorption of
www.rsc.org/advances adsorbed nitrogen via lowered activation energy and thus, enable faster reaction rate.

strong anchoring points for the metal and behave as electron


1. Introduction donors, the superior catalytic behaviors of nitrogen-containing
There is increasing interest in employing ammonia as an energy carbon materials have been identied in different catalytic
carrier.1 Ammonia decomposition to COx-free hydrogen, an reactions.17–20
ideal fuel for PEM (proton exchange membrane) fuel cells,2,3 can Covalent triazine-based frameworks (CTFs), representing a
be effectively catalyzed by the group VIII transition metals, new type of nitrogen-containing microporous polymers, were
among which the supported ruthenium catalysts have the rst synthesized by Thomas et al. through the reversible ion-
highest activity.4,5 Au et al. investigated the effects of a number othermal trimerization of aromatic nitriles in molten ZnCl2.21
of supports and demonstrated the overall performance is in the CTFs have relatively high chemical and thermal stabilities due
order of CNTs > MgO > TiO2 z AC z Al2O3.6 Similar results to the covalently bonded structure. Previous work indicates that
were also reported by Zhu et al.7 Multi-walled Carbon Nano- CTFs are more stable than N-doped CNTs under hydrophilic
tubes (MWCNTs) have a surface area greater than 100 m2 g1 and hydrophobic conditions.22 More importantly, CTFs possess
and have relatively good graphitic degree.8 However, there are some interesting properties such as large surface area, basicity
electron-withdrawing groups (–COOH, –OH) on the surface of and high degree of graphitization which make them attractive
CNTs, which may affect the electronic structure of Ru.9–12 By candidates for gas capture/storage and catalysis.23–30 As the
using Morse potential method, Bell et al. demonstrated that the content of nitrogen functional groups in CTFs overweighs that
rate limiting step for ammonia decomposition on Ru surface is of N-doped CNTs, Pd supported on CTFs has a higher activity
the recombinative desorption of adsorbed nitrogen atoms,13 and stability in the alcohol oxidation reaction than on N-doped
which can be promoted by the electropositive elements such as CNTs.22 Due to the plenty of nitrogen base sites, CTFs have been
alkalis.7 On the other hand, increasing the charge density of used as an efficient metal-free base catalyst directly to synthe-
CNTs or CNFs (Carbon Nanobers) by doping surface nitrogen size organic carbonates.26 By a simple combination of CTFs and
groups also leads to a signicantly improved activity in Pt precursor, a highly active solid catalyst for methane selective
ammonia decomposition.14–16 As nitrogen atoms can serve as oxidation has also been demonstrated.25
Previous reports on using CTFs as support are mainly for
a
Dalian National Laboratory for Clean Energy, Dalian Institute of Chemical Physics,
oxidative reaction. Herein, we investigate the performance of
Chinese Academy of Sciences, Dalian-116023, P. R. China CTF-1, a crystalline triazine-based organic framework with
b
University of Chinese Academy of Sciences, Beijing-100049, P. R. China hexagonal packing of pores,21 in a reductive process, i.e.,
c
State Key Laboratory of Catalysis, Dalian Institute of Chemical Physics, Chinese ammonia decomposition. The catalytic activity of Ru/CTF-1 was
Academy of Sciences, Dalian-116023, P. R. China. E-mail: pchen@dicp.ac.cn; Fax: measured and compared with the benchmark Ru/CNTs. Our
+86-411-8437-9583

This journal is © The Royal Society of Chemistry 2015 RSC Adv., 2015, 5, 3605–3610 | 3605
View Article Online

RSC Advances Paper

results show that, either in terms of NH3 conversion rate or The X-ray photoelectron spectroscopy (XPS) measurement
turnover frequency, the performance of Ru/CTF-1 is signi- was performed using an Escalab 250 Xi X-ray photoelectron
cantly superior to that of Ru/CNTs under the same conditions. spectrometer (Thermo Scientic) with nonmonochromatic AlKa
The electron density of the Ru on CTF-1 is greater than that on radiation (photon energy, 1486.6 eV). Small amount of Argon
CNTs, which facilitates the recombinative desorption of adsor- was injected into the sample by a weak sputtering under a very
bed nitrogen atoms. The stronger NH3 adsorption energy on Ru/ low ion current prior to the measurement. The binding energies
CTF-1 surface may enhance the interaction of reactant with the were calibrated with reference to the Ar 2p (242.0 eV).
ruthenium active sites and improve the ammonia decomposi- The temperature programmed desorption of ammonia (NH3-
tion activity. TPD) measurement was carried out on a home-made system
combining a quartz reactor and a mass spectrometer to record
the signal of NH3. The catalyst samples were reduced at 673 K in
2. Experimental pure H2 for 2 h, and then the gas ow was switched to Ar for
Catalyst preparation purging at 773 K for 2 h. Aer that the samples were cooled
down to room temperature, where NH3 adsorption was per-
CTF was prepared according to the procedure reported in
Published on 05 December 2014. Downloaded on 3/14/2024 2:51:12 AM.

formed. Physically adsorbed NH3 was removed by Ar at 373 K for


literature.21 2,6-Dicyanopyridine (DCP) and ZnCl2 in a molar
1 h. Then the temperature programmed desorption was carried
ratio of 1/5 were ball milled at 60 rpm for 0.5 h. Then the
out by heating the samples from 373 K to 723 K at a rate of
mixture was transferred into an ampoule under Ar atmosphere.
5 K min1.
The ampoule was sealed, evacuated and heated at 673 K for 40
h. Aer cooling down to room temperature, the ampoule was
opened and the reaction mixture was grounded and washed 3. Results and discussion
with diluted HCl solution and water for several times to remove
Ru/CTF-1 and Ru/CNTs catalysts with different Ru loadings
ZnCl2. Finally the black powder product was dried in Ar ow at
were prepared by traditional wetness incipient impregnation
573 K for 3 h.
using RuCl3 as precursor and acetone as solvent, where the
CNTs were pre-treated in aqueous HNO3 solution and
actual Ru loading was measured by ICP-OES (Table 1). The XRD
calcined in an Ar ow at 573 K for 3 h.
characterizations of the Ru/CTF-1 catalyst samples only show
Ru/CTF-1 and Ru/CNTs catalysts were prepared by tradi-
the sheetlike structure of CTF-1 with (100) and (001) diffractions
tional wetness incipient impregnation with RuCl3 as precursor
at ca. 7.0 and 25.1 respectively (Fig. 1d–f),21 evidencing the
and acetone as solvent. Aer drying in air for 12 h, catalysts were
high dispersion of Ru on CTF-1, probably resulting from the
calcined in an Ar ow at 573 K for 3 h.
strong metal–support interaction. The 2 wt% and 1 wt% Ru/
CNTs catalysts also show no diffraction related to ruthenium
Catalytic test metal. The diffraction peaks at 26.5 and 42.5 are attributed to
The catalytic activity was evaluated on a continuous ow quartz CNTs (Fig. 1a–c), respectively, consistent with the results
reactor. Catalyst sample of 30 mg was placed in the central reported by Au et al.6
section of the reactor and was tested under a ow of pure NH3 Transmission electron microscope (TEM) observation
(gas ow rate ¼ 30 mL min1, 15 mL min1, 10 mL min1) or a together with statistical analyses by counting 100 particles show
ow of diluted NH3 (5 vol% NH3/Ar, gas ow rate ¼ that the particle size distributions of Ru supported on CTF-1
30 mL min1). Prior to the measurement, Ru/CTF-1 and Ru/ and CNTs are similar, between 1 nm and 5 nm (Fig. 2, insets),
CNTs catalysts were reduced in pure NH3 at 673 K for 2 h and and the average Ru particle size of Ru/CTF-1 catalysts is slightly
773 K for 3 h, respectively. The effluent gasses were analyzed by smaller than that of Ru/CNTs catalysts, but all of them are
a gas chromatograph equipped with a thermal conductivity around 3 nm (Table 1) showing that Ru can be well dispersed on
detector (TCD) and a Porapak Q column, using hydrogen as the CTF-1 as well as on CNT. Because of the micropore size of CTF-1
carrier gas. is about 1.5 nm, which is smaller than the average Ru particle
size, we believe the majority of Ru particles are located outside
the micropores. However, the largely shrunken surface area of
Catalyst characterization Ru/CTF-1 (see Table 1) compared with the pristine CTF-1 (SBET
The actual Ru loadings of the catalysts were determined by ¼ 734.7 m2 g1) indicates the blockage of the channels by Ru
inductively coupled plasma spectrometry (ICP-OES, optima particles, anchoring at the openings of the CTF-1 channels. For
7300DV, Perkin-Elmer, USA). the Ru/CNTs catalysts, Ru particles are essentially located on
X-ray diffraction (XRD) patterns were recorded on a PAN- the external surface of CNTs. The degree of dispersion was
alytical X'pert diffractometer with monochromatized Cu-Ka calculated according to the spherical model proposed by
radiation (l ¼ 0.154 nm) at a setting of 40 kV and 40 mA. Anderson (see Table 1),31 which shows that the dispersions of
Transmission electron microscope (TEM) images of the Ru in 1 wt% Ru/CNTs and 2 wt% Ru/CNTs catalysts are ca.
samples were taken on a JEM-2100 at 200 kV. 41.2%. With various functional groups (such as –COOH, –OH)
The specic surface area was measured on Autosorb-1 bonded to the surface, CNTs can effectively anchor and disperse
system (Quantachrome, USA) by N2 adsorption isotherm Ru nanoparticles.7 Dispersions of the 1 wt% Ru/CTF-1 and 2
through BET method. wt% Ru/CTF-1 catalysts are 44.5% and 44.0% respectively,

3606 | RSC Adv., 2015, 5, 3605–3610 This journal is © The Royal Society of Chemistry 2015
View Article Online

Paper RSC Advances

Table 1 Properties and activities of CTF-1 and CNTs supported Ru catalysts

Catalyst Ru loadinga [%] SBETb [m2 g1] Particle sizec [nm] Dispersiond [%] TOFe [s1] Eaf [kJ mol1]

1 wt% Ru/CTF-1 0.82 654.6 2.9 45.5 0.58 69.1


2 wt% Ru/CTF-1 1.81 614.1 3.0 44.0 0.40 67.1
1 wt% Ru/CNTs 0.76 80.3 3.2 41.2 0.19 88.5
2 wt% Ru/CNTs 1.95 67.6 3.2 41.2 0.21 87.3
a
Ru actual loading was determined by ICP-OES. b Specic surface area was determined by N2 physisorption, using BET method. c Statistical results
of TEM images. d Dispersion was calculated employing spherical model proposed by Anderson.31 e TOF values were measured under 5 vol% NH3/Ar
ow at 673 K. f Activation energy was measured under pure NH3 ow at 623–723 K.

N atoms in CTF-1 are essentially in sp2 hybridization. Their


lone electron pairs are effective Lewis bases which can affect the
Published on 05 December 2014. Downloaded on 3/14/2024 2:51:12 AM.

electronic structure of Ru attached to them. As can be seen in


Fig. 3, the electron binding energy of Ru 3p3/2 of pre-reduced
2 wt% Ru/CNTs is at ca. 461.6 eV, evidencing the metallic
state of Ru in the catalyst. The electron binding energy of Ru
3p3/2 in 2 wt% Ru/CTF-1 is, interestingly, 0.6 eV smaller than
that in 2 wt% Ru/CNTs. In other words, the electron density of
Ru on CTF-1 is greater than that on CNTs.
The chemisorption of NH3 on catalysts was investigated by
temperature-programmed desorption (TPD) technique. Fig. 4
shows the TPD proles of the supported Ru catalysts and the
Fig. 1 XRD patterns of (a) CNTs; (b) 1 wt% Ru/CNTs; (c) 2 wt% Ru/ supports themselves upon absorbing NH3 at ambient temper-
CNTs; (d) CTF-1; (e) 1 wt% Ru/CTF-1 and (f) 2 wt% Ru/CTF-1. ature followed by pre-heat treatment at 373 K. For neat CNTs
and CTF-1 supports, little NH3 can be detected during the
testing history. However, for CNTs and CTF-1 supported Ru
which are slightly higher than those of the Ru/CNTs catalysts. catalysts, NH3 desorption can be observed in the temperature
Chan-Thaw et al. reported that Pd NPs (nanoparticles) can be range of 400–500 K and 400–650 K, respectively. The NH3
xed on CTF-1 through the interaction with its nitrogen desorption peak temperature of Ru/CTF-1 is at ca. 503 K, which
groups.22 Here we also attribute the high dispersion of Ru on is about 55 K higher than that of Ru/CNTs revealing stronger
CTF-1 to such a metal–support interaction. interaction between NH3 and Ru/CTF-1. Such a stronger
chemisorption may facilitate the competitive chemisorption of
NH3 molecules onto the active sites leading to an increased

Fig. 2 TEM images of Ru-based catalysts. (a) 1 wt% Ru/CTF-1; (b) 2 wt% Ru/CTF-1; (c) 1 wt% Ru/CNTs and (d) 2 wt% Ru/CNTs. Insets: Ru particle
size distributions.

This journal is © The Royal Society of Chemistry 2015 RSC Adv., 2015, 5, 3605–3610 | 3607
View Article Online

RSC Advances Paper

ammonia (5 vol% NH3/Ar) ow was used as the reactant to


uncover the intrinsic activities of the CTF-1 and CNTs supported
Ru catalysts. Fig. 5a illustrates the NH3 conversion as a function
of reaction temperature. The NH3 conversion of 2 wt% Ru/CNTs
and 1 wt% Ru/CNTs is negligible at temperatures below 623 K.
However, much higher conversion rates were observed on the
CTF-1 supported catalysts. At 623 K, NH3 conversion rates over 1
wt% and 2 wt% Ru/CTF-1 is ca. 8.0 and 3.3 times higher than
those of 1 wt% and 2 wt% Ru/CNTs, respectively. In higher
temperature range, the catalytic activities of Ru/CTF-1 catalysts
keep on overweighing those of the Ru/CNTs catalysts. It should
be noted that the 1 wt% Ru/CTF-1 catalyst is even superior to
Fig. 3 XPS spectra of the ruthenium 3p3/2 region of (a) 2 wt% Ru/CTF-1; 2 wt% Ru/CNTs under the same reaction condition. The turn-
(b) 2 wt% Ru/CNTs. over frequency (TOF) values (see Table 1) calculated from the
Published on 05 December 2014. Downloaded on 3/14/2024 2:51:12 AM.

hydrogen formation rate (mol gcat1 s1) per surface exposed Ru


atom number per gram catalyst (mol gcat1) show that at 673 K,
the TOF over 2 wt% Ru/CTF-1 is ca. 2.0 times higher than that
over 2 wt% Ru/CNTs. For the 1 wt% Ru based catalysts, it is 3.1
times.
Under a ow of pure NH3, the ammonia conversion rates
over the Ru/CTF-1 are 1.5–2.5 times higher than those over Ru/
CNTs in the temperature range of 598–723 K (see Fig. 5b). The
turnover frequency values as a function of reaction temperature
are shown in Fig. 6. Ruthenium loading has little inuence
upon TOF for the same kind of catalysts. The apparent activa-
tion energy of Ru/CTF-1 obtained from the Arrhenius plots is
about 20 kJ mol1 lower than that of Ru/CNTs (Fig. 7).
For Ru based catalysts, the B5 sites (ve member sites on the
steps of Ru (0001) plane) are regarded generally as the active
Fig. 4 NH3-TPD profiles of (a) CNTs; (b) CTF-1; (c) 2 wt% Ru/CNTs and sites for ammonia synthesis and decomposition.32,33 The
(d) 2 wt% Ru/CTF-1. amounts of B5 sites are structural sensitive and depend
signicantly on the size and dispersion of Ru particles. Previous
investigations by Kowalczyk et al. showed that the maximum
chance of reaction. Furthermore, the amount of desorbed NH3 concentration of B5 sites was achieved when the Ru particle size
from Ru/CTF-1 is obviously more than that of Ru/CNTs, which is in the range of 1.8–2.5 nm.34 In our work, the Ru particle sizes
indicates more suitable sites for the strong adsorption of NH3 of Ru/CTF-1 and Ru/CNTs catalysts are near the optimal range.
on the surface of Ru/CTF-1. Considering the similar Ru particle sizes and dispersions of all
To minimize the effect of heat and mass transport during the the CTF-1 and CNTs supported catalysts, the differences in
endothermic ammonia decomposition reaction, a diluted

Fig. 5 Ammonia decomposition reaction activity test with 2 wt% Ru/CTF-1 (-); 2 wt% Ru/CNTs (B); 1 wt% Ru/CTF-1 (:) and 1 wt% Ru/CNTs
(P). (a) Reaction conditions: catalyst (30 mg), 5 vol% NH3/Ar (flow rate ¼ 30 mL min1); (b) reaction conditions: catalyst (30 mg), pure NH3 (for 2
wt% Ru/CTF-1 and 2 wt% Ru/CNTs, flow rate ¼ 15 mL min1, GHSV ¼ 30 000 mL h1 gcat1; for 1 wt% Ru/CTF-1 and 1 wt% Ru/CNTs, flow rate ¼
10 mL min1, GHSV ¼ 20 000 mL h1 gcat1).

3608 | RSC Adv., 2015, 5, 3605–3610 This journal is © The Royal Society of Chemistry 2015
View Article Online

Paper RSC Advances

electrons. The increase in electron density on Ru surface, as


discussed above, may lead to the decrease in the kinetic barrier
for the recombinative removal of surface nitrogen atoms, which
is also in accordance with the decrease in Ea measurements (i.e.,
Ea of Ru/CTF-1 is ca. 20 kJ mol1 lower than that of Ru/CNTs).
NH3 and H2 have reaction orders of 0.69 and 1.6, respec-
tively, over well dispersed Ru at 663 K.36 The negative reaction
order for H2 shows its inhibition to ammonia decomposition,
which can be attributed to the hindrance of surface sites by the
strongly chemisorbed hydrogen.37 Our NH3-TPD proles show
that NH3 adsorption energy on Ru/CTF-1 is stronger than that
on Ru/CNTs, which may, to a certain extent, enhance the chance
Fig. 6 Turnover frequency of 1 wt% Ru/CNTs (P); 1 wt% Ru/CTF-1 of NH3 to interact with the ruthenium active sites and facilitate
(C); 2 wt% Ru/CNTs (O) and 2 wt% Ru/CTF-1 (-) under pure NH3 the activation of N–H bond, thus further improve the ammonia
Published on 05 December 2014. Downloaded on 3/14/2024 2:51:12 AM.

flow at 598–723 K. decomposition activity.

4. Conclusions
CTF-1, a kind of novel nitrogen-containing microporous poly-
mer, is a superior support for Ru catalyst. The Ru/CTF-1 shows
much better performance in ammonia decomposition than Ru/
CNTs, which ranks one of the best un-promoted catalyst
reported so far. Besides the high dispersion of Ru nanoparticles
over CTF-1, we ascribe the improved performance to the rich
nitrogen groups on CTF-1, which can anchor Ru nanoparticles
and transfer electrons to Ru. The enhanced electron density of
Ru surface is benecial for the recombinative desorption of
adsorbed nitrogen atoms. Nitrogen groups on CTF-1 play an
Fig. 7 Arrhenius plots of 1 wt% Ru/CNTs (P); 1 wt% Ru/CTF-1 (:); 2 important role as an electron donor to Ru and improve the
wt% Ru/CNTs (C) and 2 wt% Ru/CTF-1 (,). The temperature range catalytic activity.
(648–723 K) used for the determination of apparent activation ener-
gies was based on NH3 conversion values far away from the equilib-
rium values. The calculated average values and deviation errors of the
Arrhenius plots were listed.
Acknowledgements
The authors would like to acknowledge the nancial supports
from Project of National Natural Science Funds for Distin-
activity (NH3 conversion rate and TOF) may not be ascribed to guished Young Scholars (51225206), 973 Project
the particle size effect. (2010CB631304) and the National Natural Science foundation of
Bell et al. demonstrated that the rate limiting step of China (21133004 and 21273187).
ammonia decomposition over Ru catalysts at relatively lower
temperature is the recombinative desorption of adsorbed
nitrogen atoms.13 At higher temperatures, the cleavage of N–H References
bond will be the rate-limiting step as proposed by Tsai and
1 A. Klerke, C. H. Christensen, J. K. Norskov and T. Vegge, J.
Weinberg.35 Bradford et al. suggested that both cleavage of N–H
Mater. Chem., 2008, 18, 2304–2310.
bond and recombinative desorption of adsorbed nitrogen
2 F. Hayashi, Y. Toda, Y. Kanie, M. Kitano, Y. Inoue,
atoms are kinetically slow on well-dispersed Ru.36 Support or
T. Yokoyama, M. Hara and H. Hosono, Chem. Sci., 2013, 4,
additive which can facilitate electron feedback to the anti-
3124–3130.
bonding orbital of the transient metal–nitrogen bond should
3 E. Garcia-Bordeje, S. Armenise and L. Roldan, Catal. Rev.: Sci.
be benecial for these steps.2 Compared with activated carbon,
Eng., 2014, 56, 220–237.
MgO, Al2O3 and TiO2, CNTs is the best support for ammonia
4 F. Schuth, R. Palkovits, R. Schlogl and D. S. Su, Energy
decomposition because of its high metal dispersion and good
Environ. Sci., 2012, 5, 6278–6289.
degree of graphitization, which is benecial for the electron
5 X. Z. Duan, X. G. Zhou and D. Chen, Catalysis, 2013, 25, 118–
transfer between support and Ru.11 Such an electron transfer
140.
process can be enhanced upon using CTF-1 as support due to
6 S. F. Yin, B. Q. Xu, W. X. Zhu, C. F. Ng, X. P. Zhou and
the fact that the surface of CTF-1 is rich in nitrogen groups,
C. T. Au, Catal. Today, 2004, 93–5, 27–38.
leading to the increase in electron donation to Ru, which is
7 L. Li, Z. H. Zhu, Z. F. Yan, G. Q. Lu and L. Rintoul, Appl.
evidenced by the downshi of binding energy of Ru 3p
Catal., A, 2007, 320, 166–172.

This journal is © The Royal Society of Chemistry 2015 RSC Adv., 2015, 5, 3605–3610 | 3609
View Article Online

RSC Advances Paper

8 P. Serp, M. Corrias and P. Kalck, Appl. Catal., A, 2003, 253, 23 C. E. Chan-Thaw, A. Villa, G. M. Veith, K. Kailasam,
337–358. L. A. Adamczyk, R. R. Unocic, L. Prati and A. Thomas,
9 Z. H. Zhong and K. Aika, J. Catal., 1998, 173, 535–539. Chem.–Asian J., 2012, 7, 387–393.
10 Z. H. Zhong and K. Aika, Inorg. Chim. Acta, 1998, 280, 183– 24 X. M. Liu, H. Li, Y. W. Zhang, B. Xu, A. Sigen, H. Xia and
188. Y. Mu, Polym. Chem., 2013, 4, 2445–2448.
11 S. F. Yin, B. Q. Xu, X. P. Zhou and C. T. Au, Appl. Catal., A, 25 R. Palkovits, M. Antonietti, P. Kuhn, A. Thomas and
2004, 277, 1–9. F. Schuth, Angew. Chem., Int. Ed., 2009, 48, 6909–6912.
12 K. Aika, A. Ohya, A. Ozaki, Y. Inoue and I. Yasumori, J. Catal., 26 J. Roeser, K. Kailasam and A. Thomas, ChemSusChem, 2012,
1985, 92, 305–311. 5, 1793–1799.
13 E. Shustorovich and A. T. Bell, Surf. Sci., 1991, 259, L791– 27 X. W. Chen, F. Yuan, Q. F. Gu and X. B. Yu, J. Mater. Chem. A,
L796. 2013, 1, 11705–11710.
14 S. Armenise, L. Roldan, Y. Marco, A. Monzon and E. Garcia- 28 J. L. Liu, E. M. Zong, H. Y. Fu, S. R. Zheng, Z. Y. Xu and
Bordeje, J. Phys. Chem. C, 2012, 116, 26385–26395. D. Q. Zhu, J. Colloid Interface Sci., 2012, 372, 99–107.
15 J. L. Chen, Z. H. Zhu, S. B. Wang, Q. Ma, V. Rudolph and 29 A. Modak, J. Mondal, M. Sasidharan and A. Bhaumik, Green
Published on 05 December 2014. Downloaded on 3/14/2024 2:51:12 AM.

G. Q. Lu, Chem. Eng. J., 2010, 156, 404–410. Chem., 2011, 13, 1317–1331.
16 F. R. Garcia-Garcia, J. Alvarez-Rodriguez, I. Rodriguez-Ramos 30 A. Thomas, A. Fischer, F. Goettmann, M. Antonietti,
and A. Guerrero-Ruiz, Carbon, 2010, 48, 267–276. J. O. Muller, R. Schlogl and J. M. Carlsson, J. Mater. Chem.,
17 R. Czerw, M. Terrones, J. C. Charlier, X. Blase, B. Foley, 2008, 18, 4893–4908.
R. Kamalakaran, N. Grobert, H. Terrones, D. Tekleab, 31 J. R. Anderson, Structure of metallic catalysts, Academic Press,
P. M. Ajayan, W. Blau, M. Ruhle and D. L. Carroll, Nano New York, 1975.
Lett., 2001, 1, 457–460. 32 C. J. H. Jacobsen, S. Dahl, P. L. Hansen, E. Tornqvist,
18 K. P. Gong, F. Du, Z. H. Xia, M. Durstock and L. M. Dai, L. Jensen, H. Topsoe, D. V. Prip, P. B. Moenshaug and
Science, 2009, 323, 760–764. I. Chorkendorff, J. Mol. Catal. A: Chem., 2000, 163, 19–26.
19 C. H. Hsu, H. M. Wu and P. L. Kuo, Chem. Commun., 2010, 33 S. Dahl, J. Sehested, C. J. H. Jacobsen, E. Tornqvist and
46, 7628–7630. I. Chorkendorff, J. Catal., 2000, 192, 391–399.
20 M. P. Woods, E. J. Biddinger, P. H. Matter, B. Mirkelamoglu 34 W. Rarog-Pilecka, E. Miskiewicz, D. Szmigiel and
and U. S. Ozkan, Catal. Lett., 2010, 136, 1–8. Z. Kowalczyk, J. Catal., 2005, 231, 11–19.
21 P. Kuhn, M. Antonietti and A. Thomas, Angew. Chem., Int. 35 W. Tsai and W. H. Weinberg, J. Phys. Chem. C, 1987, 91,
Ed., 2008, 47, 3450–3453. 5302–5307.
22 C. E. Chan-Thaw, A. Villa, L. Prati and A. Thomas, 36 M. C. J. Bradford, P. E. Fanning and M. A. Vannice, J. Catal.,
Chem.–Eur. J., 2011, 17, 1052–1057. 1997, 172, 479–484.
37 W. Tsai, J. J. Vajo and W. H. Weinberg, J. Phys. Chem. C, 1985,
89, 4926–4932.

3610 | RSC Adv., 2015, 5, 3605–3610 This journal is © The Royal Society of Chemistry 2015

You might also like