You are on page 1of 10

Applied Catalysis A: General 390 (2010) 3544

Contents lists available at ScienceDirect

Applied Catalysis A: General


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

Methanation of CO, CO2 and selective methanation of CO, in mixtures of CO and


CO2 , over ruthenium carbon nanobers catalysts
Vicente Jimnez a, , Paula Snchez a , Paraskevi Panagiotopoulou b , Jos Lus Valverde a , Amaya Romero a
a
Facultad de Ciencias Qumicas/Escuela Tcnica Agrcola, Departamento de Ingeniera Qumica, Universidad de Castilla-La Mancha, 13071 Ciudad Real, Spain
b
Department of Chemical Engineering, University of Patras, GR-26504 Greece

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

Article history: The catalytic performance of ruthenium catalysts supported on carbon nanobers for the methanation of
Received 4 August 2010 CO, CO2 and their mixture has been investigated with respect to the nature of carbon nanobers (orienta-
Received in revised form 3 September 2010 tion of graphite planes): platelet, shbone and ribbon. Experiments were conducted in the temperature
Accepted 21 September 2010
range of 200500 C using feed compositions relevant to those of reformate gas streams, both in the
Available online 29 September 2010
absence and in the presence of water. It has been found that, under conditions of solo-CO methanation,
all the investigated catalysts are able to completely and selectively convert CO at temperatures around
Keywords:
340 C, with the conversion of CO being somewhat higher for the Ru/platelet sample. For hydrogenation
CO methanation
CO2 methanation
of CO2 alone, catalytic performance is not affected by the nature of the carbon nanobers used as support.
Selective methanation of CO In combined hydrogenation of CO/CO2 mixtures, catalytic performance for all the investigated catalysts is
Ruthenium poor since they promote the undesired reverse watergas shift reaction. However, addition of 30% water
Fuel cell applications vapour in the feed inhibits the reverse watergas shift, thereby enhancing CO hydrogenation. Results of
Carbon nanobers kinetic measurements show that the turnover frequency of CO conversion becomes 23 times higher
in the presence of steam over Ru/shbone and Ru/platelet samples over the whole temperature range
examined, whereas in the case of Ru/ribbon catalyst temperatures higher than 250 C are required in
order to achieve higher turnover frequency values. Carbon dioxide hydrogenation is not affected by the
presence of steam. For all experimental conditions investigated, selectivity toward methane increases
with increasing temperature at the expense of higher hydrocarbons and is enhanced with the addition
of water vapour in the gas mixture.
2010 Elsevier B.V. All rights reserved.

1. Introduction remove CO down to 10 ppm [911]. The catalytic hydrogenation of


carbon monoxide and carbon dioxide also produces a large variety
One the major problems for the application of low temperature of products ranging form methane and methanol to higher molec-
polymer electrolyte membrane fuel cells (PEM FCs) as power source ular weight hydrocarbons and alcohols [12]. The methanation of
for electrically operated vehicles is the delivery of nearly-CO-free CO and CO2 as well as the related FischerTropsch reaction have
feed gas, which becomes problematic whenever H2 is generated been extensively studied and reviewed. COx methanation may have
from fuels. As the FC anodes can be poisoned even by traces impu- another practical application as a means of CO removal from pro-
rities of CO, this gas has to be removed to a level below 50 ppm for cess gases for gas separation purposes and is also being discussed
PtRu anode electrocatalyst [13] and below 10 ppm for Pt anode as an alternative to PROX in fuel processors for mobile fuel cell
electrocatalysts [46]. applications [13]. However, in the last years, the onboard hydro-
Theoretically, there are several methods to reduce carbon gen production has been a minor system for the application of fuel
monoxide to the levels acceptable for a fuel cell [7]. It is feasible to processing to FCV, except some particular cases. The methanation
separate hydrogen by diffusion thorough a CO ltering membrane reaction (Eq. (1)) has been widely used as a gas-purication pro-
but the membrane is expensive and it usually requires a compressor cess. It is used on a large scale for the purication of hydrogen in
owing to high pressure [8]. The most studied system for the removal ammonia and hydrogen plants where CO is a catalyst poison [14].
of nal traces of CO over the last years has been the selective oxida- The interest for the reaction has grown signicantly during the last
tion of CO in a H2 -rich atmosphere (PROX) since it has the ability to few years as a promise route to remove CO in the reformate stream
down to 50 ppm, poisoning limit for the use of polymer electrolyte
fuel cells [15,16].
Corresponding author. Tel.: +34 926295300; fax: +34 926295318.
E-mail address: vicente.jimenez@uclm.es (V. Jimnez). CO + 3H2 CH4 + H2 O, H = 206 kJ/mol (1)

0926-860X/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2010.09.026
36 V. Jimnez et al. / Applied Catalysis A: General 390 (2010) 3544

This approach has the advantage that no oxidizing and/or inert support in hydrouoric acid (70%) for 15 h under vigorous stirring,
gases are mixed with the reformate stream and that the methane ltering and washing with deionised water [40]. Upon treatment of
produced, which is inert to the PEM fuel cell, can be utilized in the catalystcarbon mixture, the metal component of the catalyst
residential PEFC systems. However, depending on the operating was generally transferred into the solution.
conditions and catalyst employed, reaction (1) may run in parallel Methanation catalysts were prepared employing the wet
with the undesired methanation of CO2 , which consumes signi- impregnation method [25] by using Ru(NO)(NO3 )3 (Alfa Products)
cant quantities of valuable hydrogen (Eq. (2)), as well as with the as metal precursor salt and the different types of CNFs obtained. The
reverse watergas shift (RWGS) reaction (Eq. (3)), which shift CO2 resulting slurry was heated at 1 C min1 up to 90 C under continu-
to CO [17]. ous stirring and maintained at that temperature until nearly all the
water evaporated. The solid residue was dried at 110 C for 24 h and
CO2 + 4H2 CH4 + 2H2 O, H = 165 kJ/mol (2)
then reduced at 400 C in H2 ow for 2 h. The metal loading of the
CO2 + H2 CO + H2 O, H = 41.1 kJ/mol (3) catalysts thus prepared was 0.5 wt.%.

Consequently, it is important to develop selective CO metha- 2.2. Support/catalyst characterization


nation catalysts characterized by high activity at sufciently low
temperatures, able to retard both the CO2 and the RWGS reactions. Surface area/porosity measurements were carried out using a
Methanation of CO over different carried metal catalysts includ- Micromeritics ASAP 2010 sorptometer apparatus with N2 at 77 K
ing Ni [1821], Ru [18,2224] and Rh [21,2426], has been widely as the sorbate. The samples were outgassed at 453 K under vac-
investigated for the production of CH4 from syngas [4,2729] and uum (6.6 109 bar) for 16 h prior to analysis. Specic surface areas
recently also from a viewpoint of residual CO removal for PEM were determined by the multi point BET method, total pore vol-
FC applications [3036]. Ruthenium catalysts dispersed on metal ume and sizes were evaluated using the standard BJH treatment
oxide carriers have been found to exhibit high activity for the solo- and micropore volume were evaluated using the t-plot equation.
methanation of CO [18,19,37] or CO2 [20,21] as well as for the The crystallinity of CNFs, the mean crystallite size and Ru species
co-methanation of CO/CO2 mixtures [19,25,33]. In a recent study was determined using XRD analyses. These analyses were carried
[25] the effects of the nature of the metallic phase on the per- out on a Philips XPert instrument using nickel ltered Cu K radia-
formance of Al2 O3 -supported Ru, Rh, Pt and Pd catalysts for the tion through a primary monochromator. The samples were scanned
methanation of CO, CO2 and their mixtures both in the absence and at a rate of 0.02 step1 over the range 5 2 80 (scan time = 2 s
in the presence of water in the feed has been studied. It has been step1 ). The metal particle size was determined after the catalyst
found that Ru and Rh are much more active hydrogenation cata- reduction in H2 ow to 400 C for 2 h. Once the sample was cooled,
lysts as compared to Pt and Pd, which promote the undesired RWGS it was passivated in a 1% (v/v) O2 /He mixture to prevent bulk oxi-
reaction. Furthermore, the nature of the support may play a crucial dation. The primary crystallite size of Ru (dRu ) was calculated by
role in the mechanism of CO/CO2 hydrogenation reactions, since means of Scherrers equation [41]:
metal-support interactions can modify the catalytic properties of
the metallic phase [26,37,38]. Supports as carbon nanobers could 0.9
dRu = (4)
be an alternative to the classical ones (alumina, silica, TiO2 , etc) B cos 
widely studied on the literature [25,16,38] due to their excellent where  is the X-ray wavelength corresponding to Cu-K radiation
characteristics as the high purity of the material, high mechani- (0.15406 nm), B is the broadening (in radians) of the ruthenium
cal strength and the mesopore nature which result in low internal (1 0 1) reection and  is the angle of diffraction corresponding to
mass-transfer resistances [39]. the peak broadening.
In the present study, the catalytic performance of Ru cata- The chemisorption measurements were carried out using a
lyst on selective COx methanation using different types of carbon dynamic pulse technique with an argon ow of 50 mL min1 and
nanobers (platelet, shbone and ribbon) has been investigated. pulses of H2 (99.9995% purity) using a Micromeritics AutoChem
The inuence of water vapour in the feed was also studied. 2950 HP apparatus according to the procedure described in ref.
[40,42,43]. In order to calculate the metal dispersion, an adsorption
2. Experimental stoichiometry of metal/H = 1 was assumed [44]. Dispersion mea-
surements with H2 pulses were carried out at 60 C to avoid the spill
2.1. Support/catalyst preparation over phenomenon [45]. Previously, the sample was pre-treated by
heating at 15 C min1 in argon ow up to 250 C and kept constant
CNFs synthesis was carried out by CVD method. The xed-bed at this temperature for 20 min. Then, the sample was reduced in
reactor, consisting on a quartz tube of 9 cm diameter and 100 cm situ. Next, the hydrogen was removed by owing argon for 30 min,
length, was located in a horizontal electric furnace (JH Hornos) with being the temperature 10 C higher than the reduction one. Finally,
an effective heating zone of 80 cm. Thermocouple type K was used the sample was cooled to the experiment temperature in an argon
for monitoring the temperature of the bed. Hydrogen and ethane gas ow. The mean crystallite size of the dispersed metal was also
ow rates were controlled by mass controllers (Brooks Instruments, estimated from hydrogen chemisorption data, assuming spherical
model 5850). Supported catalyst was taken in a quartz boat, which particles, a H:Ru stoichiometry of 1:1 and an atomic surface area of
was kept inside the heating zone during the experiment. Carbon Ru equal to 8.6 A2 [17], using the relation:
nanobers were grown at atmospheric pressure at different tem-
6
peratures: 450 C (platelet type), 600 C (shbone type) and 850 C dM = (5)
M SM
(ribbon type). In each synthesis run 5 g of the prepared catalyst
(Ni/SiO2 ) was placed in the centre of the reactor and activated by where dM is the mean crystallite diameter, M is the density of Ru
heating (10 C min1 ) in a ow dry 20% (v/v) H2 /He at the desired (12.3 g cm3 ) and SM is the exposed surface area per gram of metal.
reaction temperature. The reduced activated catalyst was thor- Temperature-programmed oxidation (TGA/TPO) was used to
oughly ushed with dry He for 1 h before introducing the H2 /C2 H4 determine the CNFs crystallinity. Analyses were performed on
ratio desired feed. The growth time was 1 h and the space veloc- 10 mg samples using a Perkin-Elmer TGA7 termogravimetric ana-
ity 25,000 h1 . Separation of CNFs and catalyst particles to recover lyzer with a ow of 50 cm3 min1 of 20% (v/v) O2 /He mixture and
the carbon material was carried out by repeated dissolution of the with a heating rate of 5 C min1 up to 1000 C.
V. Jimnez et al. / Applied Catalysis A: General 390 (2010) 3544 37

Fig. 1. Representative images of the different type of carbon nanobers (CNFs): (a) SEM picture of general CNFs, (b) TEM picture of platelet CNFs, (c) TEM picture of shbone
CNFs and (d) TEM picture of ribbon CNFs.

Support morphology was examined by scanning electron 0.5 g) were treated in 2 cm3 HCl, 3 cm3 HF and 2 cm3 H2 O2 followed
microscopy (SEM) using a JEOL JSM-6535F SEM unit with an by microwave digestion (523 K).
accelerating voltage of 10 kV. The samples were deposited on a
standard aluminium SEM holder and coated with gold (Balzers 2.3. Catalytic activity
Union, Model MED 010). Transmission electron microscopy (TEM)
analysis employed was a Philips Tecnai 20T, operated at an accel- Catalytic performance tests and kinetic measurements have
eration voltage of 200 keV. Samples were prepared by ultrasonic been carried out using an apparatus, consisting on a ow mea-
dispersion in acetone with a drop of the resultant suspension evap- suring and control system, the reactor and an on-line analysis
orated onto a holey carbon supported grid. system [25]. The ow system is equipped with a set of mass-ow
The Ru metal loading was determined by atomic absorption (AA) controllers (MKS) and a set of valves, which allows introduction
spectrophotometry, using a SPECTRA 220FS analyzer. Samples (ca. of the gas mixture to the reactor or to a by-pass loop, through

Table 1
Physicochemical properties of the CNFs and the reduced Ru catalysts.

CNFs support Ru-based CNFs catalysts

Platelet Fishbone Ribbon Platelet Fishbone Ribbon

BET surface area (m2 /g) 286.0 202.0 68.0 183.6 175.9 93.9
Micropore area (m2 /g)a 57.0 (20%) 25.0 (13%) 2.0 (3%) 11.7 (6%) 20.6 (12%) 28.7 (31%)
Mesopore area (m2 /g) 229.0 177.0 66.0 171.9 155.4 65.2
Micropore volume (cm3 /g)b 1.194 (0.62) 0.546 (0.62) 0.052 (0.64) 0.476 (0.67) 0.514 (0.66) 0.116 (0.67)
Mesopore volume (cm3 /g)c 0.409 (8.7) 0.977 (10.1) 0.246 (16.7) 0.403 (9.9) 0.378 (10.9) 0.232 (11.4)
Average pore size (nm) 6.9 9.3 14.1 6.1 8.8 9.7
npgd 7.7 8.6 12.3
Weight loss temperature range 380585 (519) 430618 (546) 545715 (640)
( C)e
NaOH added (mmol/g) 0.75 0.56 0.19 0.78 0.58 0.40
Ru dispersion (%) 67.7 63.5 58.5
dRu,TEM (nm)f 1.04 1.48 1.93
dRu, H2 (nm)g 1.41 1.50 1.63
a
In brackets: percentage of micropore area with respect to the total surface area.
b
Cumulative pore volumes obtained using HorvathKawazoe method. In brackets: mean micropore size in nm.
c
Cumulative pore volumes obtained using BJH method. In brackets: mean mesopore size in nm.
d
Number of graphene planes in the crystallites Lc /d0 0 2 where Lc is the average crystal domain size along a direction perpendicular to the basal planes in a graphitic-type
structure, and d0 0 2 is the average interlayer spacing.
e
In brackets: temperature at which the maximum of the oxidation temperature peak appears.  3  2
f
Average diameters of Ru particles determined by counting around 200 particles on the TEM images using the equation: dRu = ni di / ni di where ni is the number
of particle of diameter di .
g
Average diameter of Ru particles measured using H2 chemisorption technique using the expression dM = 6/M SM where, M is the density of Ru (12.3 g cm3 ) and SM is
the exposed surface area per gram of metal [20].
38 V. Jimnez et al. / Applied Catalysis A: General 390 (2010) 3544

350 0.125 0.5

Mesopore Volume (dV/dlogD)


Platelet CNFs

(cm
Ru/Platelet CNFs Platelet CNFs
300

N2volume adsorbed/desorbed
0.100 0.4

Micropore Volume (cm 3 /g)


3/g)
250
0.075
200 0.3

150 0.050
0.2
100
0.025
50 0.1

0.000
0
0.0
0.0 0.2 0.4 0.6 0.8 1.0 1 1 10
Pore diameter (nm)
Relative pressure (P/P0 )

0.075
N2 volume adsorbed/desorbed (cm3 /g)

350

Mesopore Volume (dV/dlogD)


Fishbone CNFs
Ru/Fishbone CNFs Fishbone CNFs
300 0.4
0.060
Micropore Volume (cm 3 /g)

250
0.045
200 0.3

150 0.030

100
0.015 0.2
50

0.000
0
0.1
0.0 0.2 0.4 0.6 0.8 1.0 1 1 10
Relative pressure (P/P0 ) Pore diameter (nm)
180 0.0075 0.25
N2 volume adsorbed/desorbed (cm3 /g)

Ribbon CNFs
160

Mesopore Volume (dV/dlogD)


Ru/Ribbon CNFs Ribbon CNFs
140 0.20
Micropore Volume (cm 3 /g)

120 0.0050

100 0.15

80
0.10
60 0.0025

40
0.05
20
0.0000
0 0.00

0.0 0.2 0.4 0.6 0.8 1.0


1 1 10
Relative pressure (P/P0 )
Pore diameter (nm)

Fig. 2. N2 adsorptiondesorption isotherms and micropore/mesopore volume distribution of the catalysts and the supports.

stainless steel tubing. When desired, water is introduced to the an electric furnace, the temperature of which is controlled using
system with the use of an HPLC pump (Marathon Scientic Sys- a second K-type thermocouple placed between the reactor and
tems), vaporized in a stainless steel evaporator maintained at the walls of the furnace. A pressure indicator is used to mea-
170 C and mixed with the gas stream coming from the mass- sure the pressure drop in the catalyst bed. The analysis system
ow controllers. The resulting gas mixture is then fed to the consists on a gas chromatograph (Shimadzu) equipped with two
reactor through stainless steel tubing maintained at 150 C by packed columns (Porapak-Q Carbosieve) and two detectors (TCD,
means of heating tapes. The reactor consists of a 40 cm long quartz FID) and operates with He as the carrier gas. The response fac-
tube (6 mm OD) with an expanded 1 cm long section in the mid- tors of the detectors were determined with the use of gas streams
dle (8 mm ID), in which the catalyst sample is placed. Reaction of known composition (Scott specialty gas mixtures). Reaction
temperature is measured in the middle of the catalyst bed by gases (He, 15% CO/He, CO2 , H2 ) are supplied from high-pressure
means of a K-type thermocouple placed within a quartz capil- gas cylinders (Messer Griesheim GMBH) and are of ultra-high
lary well, which runs through the cell. The reactor is placed in purity.
V. Jimnez et al. / Applied Catalysis A: General 390 (2010) 3544 39

a b Ru (101)
c Ru (101)

Intensity (a. u.)


Ru (101)
Ru (002) Ru (002)
Ru (002)

Ru (100) Ru (100)
Ru (100)

0 10 20 30 40 50 0 10 20 30 40 50 0 10 20 30 40 50
2 2 2

Fig. 3. XRD pattern for Ru catalysts supported on different types of CNFs: (a) platelet CNFs, (b) shbone CNFs and (c) ribbon CNFs.

The catalytic performance of the prepared samples for the selec- The graphitic character of the different supports was eval-
tive methanation of CO has been investigated in the temperature uated by XRD, where the npg values (Table 1) suggested that
range of 200500 C using a feed stream consisting of 1% CO, 15% ribbon CNFs were the most graphic carbon nanobers. Thus, Lc val-
CO2 and 50% H2 (balance He). When water was added in the feed, ues increased in the following order: platelet < shbone < ribbon,
part of the balance gas (He) was replaced by water vapour (30% and d0 0 2 values in the following one: ribbon < shbone < platelet.
H2 O). The mass of catalyst used in these experiments was typi- Higher values of Lc may reduce the adsorptive sites on CNFs sur-
cally 150 mg (particle size: 0.18 < d < 0.25 mm) and the total ow face which will be veried by N2 adsortion [46]. Carbon structural
rate was 200 cm3 min1 . Prior to each experiment the catalyst sam- order was also measured by means of TPO analyses. All CNFs exhib-
ple was reduced in situ at 300 C for 1 h under hydrogen ow ited only a single oxidation peak from the DTA curve (not shown),
(60 cm3 min1 ), purged with He and then conditioned at 170 C indicating high product purity [40]. Table 1 presents the maxi-
for 1 h with the reaction mixture. Conversions of reactants and mum weight loss temperature showing that an increase in the
selectivities toward products were then measured at that tem- CNFs synthesis temperature lead to more crystalline structures
perature using the analysis system described above. Selectivity to (ribbon > shbone > platelet).
hydrogenation product i(Si ) was calculated using the following Nitrogen adsorption/desorption isotherms and pore size dis-
expression: tribution are presented in Fig. 2. The isotherms can be clearly
assigned to the type IV IUPAC classication, characteristic of meso-
Ci,out /vi
Si = (6) porous materials, where the mean size (see Table 1) falls within the
i Ci,out /vi accepted mesoporous (250 nm) range. Also, a well dened hys-
teresis (H3 type according to IUPAC classication) was observed
where Ci,out is the outlet concentration of the product i and vi is the
indicating the existence of a substantial volume of mesopores
number of carbon atoms of product i. It should be noted that the
where irreversible capillary condensation occurred [47]. Textural
system was left at each temperature for about 1 h under isother-
parameters data are shown in Table 1. It is observed that for increas-
mal conditions in order to achieve steady state conditions. In all
ing CNFs synthesis temperature, the specic BET surface area and
cases, data points are overages of at least three measurements. All
the total pore volume of the resulting carbon material decreased,
experiments were performed at near atmospheric pressure.
in agreement with the results obtained by others authors [48].
Results obtained, along with measurements of metal dispersion,
Their corresponding pore size distribution curves are also showed
were used to determine the turnover frequencies (TOFs) of carbon
in Fig. 2. Average mesopore size increased in the following order:
monoxide and carbon dioxide, dened as moles of CO or CO2 con-
platelet < shbone < ribbon, being the principal mesopore size in
verted per surface Ru atom per second. Details of the methods and
the range 1530 nm, accompanied with a secondary mesopore
procedures employed can be found elsewhere [25].
size of 35 nm. Larger mesopores have been attributed to inter-
stices between interlaces laments whereas smaller ones have
3. Results and discussion been associated with the CNFs surface roughness. Obtained results
were indicative of the higher amount of adsorption sites present in
3.1. Support characterization platelet CNFs [46]. Thus, platelet CNFs had the largest BET surface
area and the smallest average pore diameter because of the pres-
In the representative TEM images showed in Fig. 1 are clearly dif- ence of a higher amount of 35 nm pores. Nevertheless, shbone
ferentiated the different structures of the CNFs used in this study as and ribbon CNFs had larger average pore diameters than platelet
catalyst supports. Carbon nanobers (CNFs) are an allotrope of car- ones because of much of their pores resulting mainly from aggre-
bon that has crystalline structure (Fig. 1a). Different types of CNFs gation.
can be obtained depending on their synthesis temperature. Platelet
type structures (hexagonal planes perpendicular to the ber axis)
were produced at a synthesis temperature of 450 C (Fig. 1b); sh- 3.2. Catalyst characterization
bone type structures (graphene layers terminate on the surface
with a determinate inclination angle) were predominant at a syn- The physicochemical characteristics of the synthesized Ru cat-
thesis temperatures of around 600 C (Fig. 1c), whereas, ribbon alysts are summarized in Table 1.
(hexagonal planes staked parallel to the ber axis) type structures XRD patterns for the activated catalysts are presented in Fig. 3.
(Fig. 1d) started to be produced at high temperatures (i.e. 850 C). Peaks at 2 26 correspond to the (0 0 2) graphite plane of carbon
40 V. Jimnez et al. / Applied Catalysis A: General 390 (2010) 3544

nanobers (JCPDS-ICDD Card No. 41-1487), where the implica-


tion in terms of support graphitic character has been discussed
above. Reections appearing at 38.2 , 42.3 and 44.1 correspond,
respectively, to the (1 0 0), (0 0 2) and (1 0 1) planes of metallic
ruthenium (JCPDS-ICDD Card No. 06-0663), which are consistent
with an exclusive hexagonal geometry and establish the presence
of Ru0 [49,50]. These results indicate that after activation of cat-
alysts with H2 ow previous to the methanation reaction all the
metallic phase was in the Ru0 form.
BET surface area and micropore/mesopore volume values of
Ru platelet and shbone CNFs (Table 1) decreased as compared
with the parent CNFs (before metal incorporation). This effect is
shown in Fig. 2. However, ribbon CNFs surface area and micro-
pore volume signicantly increase after metal incorporation. To
explain these results is necessary to have into account the posi-
tion of the Ru particles on the different supports. Thus, the surface
area decrease that took place with the platelet and shbone type
CNFs has been ascribed to the occupation of the adsorptive sites
present on the CNF surface by highly dispersed Ru nanoparticles
(note that the highest Ru dispersion values (Table 1) were obtained
with the platelet and shbone type CNFs). A high dispersion of
Ru nanoparticles on the CNFs is normally associated to a strong
metal-support interaction between the Ru nanoparticles and the
graphitic edges of the CNFs, as has been previously observed for a
CNF-supported Pd catalyst [51]. On the other hand, the micropore
surface area increase that was observed in ribbon CNFs could be
ascribed to the presence of additional oxygen containing groups
and defects created on the CNF surface during the Ru deposi-
tion treatment. These groups would lead to new adsorption sites
during N2 physisorption [52]. The presence of oxygen-containing
groups, which could signicantly inuence the adsorptive and sub-
sequent catalytic behaviour [53], can be conrmed by means of
standard acidbase titrations. In such titrations the acid groups
would be neutralized by NaOH. The quantities of acidic sites, which
are presented as millimoles of NaOH needed to neutralize the acid-
ity groups of pKi 7, are shown in Table 1. Results showed that
the total acidic groups determined by titrations decreased in the
sequence: platelet > shbone > ribbon. It is interesting to note that
the presence of oxygen-containing groups, after incorporation of
ruthenium, was the same in platelet and shbone CNFs, but it was
almost two times higher on the surface of ribbon CNFs.
Finally, the average diameter of ruthenium particles measured
by H2 -chemisorption and TEM are shown in Table 1. It can be
observed that the results obtained by both techniques were quite
similar. Ru particle size varied with the type of CNFs support. Par-
ticle size in platelet CNFs was the lowest if compared with that
of shbone and ribbon CNFs. As expected, the larger the Ru par-
ticles were, the lower metal dispersion was observed. In Fig. 4 is
shown the representative TEM images of different types of CNFs
with Ru.

3.3. CO and CO2 hydrogenation

The effect of the type of CNFs (platelet, shbone and ribbon)


used as support on catalytic performance for the solo-CO and CO2
hydrogenation was investigated over Ru catalysts of the same metal
content (0.5 wt.%). Results obtained are summarized in Fig. 5, where
conversions of CO (XCO ) and CO2 (XCO2 ) are plotted as functions of
reaction temperature. It is observed that conversion of CO (Fig. 5a)
increases with increasing temperature above ca. 240 C and reaches
100% at 340 C for all catalysts investigated. It should be noted that
catalytic performance is somewhat higher for Ru/platelet sample,
with XCO taking values of ca. 92% at 320 C, i.e. more than two times
higher than that measured for Ru/ribbon and Ru/shbone sam- Fig. 4. Representative TEM image illustrating dispersion of Ru particles in the CNFs:
(a) platelet CNFs, (b) shbone CNFs and (c) ribbon CNFs.
ples. This small difference may be attributed to the higher specic
surface area of platelet CNFs.
V. Jimnez et al. / Applied Catalysis A: General 390 (2010) 3544 41

a 100 Regarding the CO2 hydrogenation, it is initiated at about 300 C


and conversion reaches 50% at 500 C (Fig. 5b). As it can be seen
catalytic performance is not affected by the nature of the CNFs used
80 as support.
In addition to methane, hydrogenation of CO also results in
the production of higher hydrocarbons. Results obtained over the
(%)

60 Ru/CNFs catalysts are shown in Fig. 6 (solid symbols), where selec-


tivities toward hydrogenated products (SCx Hy ) are plotted as a
CO

function of reaction temperature. It is observed that, in the case


X

40 of Ru/platelet (Fig. 6a) and Ru/shbone (Fig. 6b) catalysts, SCH4


increases from 85% to 100% with increasing temperature from
CNFs
240 C to 400 C. The main by-products formed below 400 C are
20 Platelet C2 H6 and C2 H4 . At temperatures above 400 C, where conversion
Ribbon of CO is complete (Fig. 5a), formation of higher hydrocarbons is sup-
Fishbone pressed and the only hydrogenation product formed is CH4 (Fig. 6a
0 and b). Ruthenium catalyst supported on ribbon CNF was found
200 240 280 320 360 400 to be more selective toward CH4 (Fig. 6c) than Ru supported on
Temperature (C) platelet and shbone CNFs, with SCH4 being typically higher than
96% in the whole temperature range, while the only by-product
formed is C2 H4 . The above ndings are in good agreement with
b 100
results reported previously over noble metal catalysts supported
CNFs on Al2 O3 [25].
Platelet Qualitatively, similar results were obtained for the CO2 hydro-
80
Ribbon genation reaction. As it can be seen in Fig. 6 (open symbols)
selectivity toward methane takes more or less the same values
X CO (%)

Fishbone
(for each sample investigated) with that obtained for CO hydro-
2

60
genation, with Ru/ribbon catalyst exhibiting the highest SCH4 . It
has been proposed over noble metal catalysts supported on metal
40
oxides [17,25,26,5663] that under conditions of CO2 hydrogena-
tion appreciable amounts of CO2 are converted to CO via the reverse
WGS reaction. It is believed that carbon monoxide acts as an inter-
20 mediate, which is further hydrogenated to methane [59]. However,
this is not the case for the results of the present study since no
CO was detected under CO2 hydrogenation conditions. In pre-
0 vious investigations [17,25,64] it was found that the occurrence
200 250 300 350 400 450 500 and the onset of the RWGS reaction under conditions of solo-CO2
methanation and methanation of CO/CO2 mixtures depends sig-
Temperature (C)
nicantly on the nature of metal and the support. For example, it
Fig. 5. (a) Conversions of solo-CO (XCO ) and (b) CO2 (XCO2 ) as function of reaction was found that, for solo-CO2 methanation reaction, selectivity to
temperature obtained over Ru (0.5 wt.%) supported on the indicated CNFs. Experi- CO, at the expense of CH4 , at a given temperature decreases in the
mental conditions: mass of catalyst: 150 mg; particle diameter: 0.18 < dp < 0.25 mm; order of Pt > Pd  Ru > Rh, with Rh catalyst producing CO only at
feed composition: (a) 1% CO, 50% H2 (balance He) and (b) 15% CO2 , 50% H2 (balance
temperatures higher than 400450 C [25]. Moreover, it has been
He); total ow rate: 200 cm3 /min.
reported that production of CO via the RWGS depends strongly on
the nature of the support over Ru [17] and Rh [64] catalysts. Regard-
ing the results of the present study (Fig. 5b), it is possible that small

100
a b c
80
Selectivity (%)

60

CO/H2 CO2/H2 CO/H2 CO2/H2 CO/H2 CO2/H2


40 , CH4 , CH4 , CH4
, C2H4 , C 2H 4 , C2H 4
, C2H6 C 2H 6
20

0
240 300 360 420 480 240 300 360 420 480 240 300 360 420 480
Temperature (C)

Fig. 6. Effect of reaction temperature on the selectivity to the indicated reaction products obtained under conditions of solo-CO (solid symbols) and CO2 (open symbols) metha-
nation over Ru (0.5 wt.%) supported on (a) platelet, (b) shbone and (c) ribbon CNFs. Experimental conditions: mass of catalyst: 150 mg; particle diameter: 0.18 < dp < 0.25 mm;
feed composition: CO/H2 : 1% CO, 50% H2 (balance He) and CO2 /H2 : 15% CO2 , 50% H2 (balance He); total ow rate: 200 cm3 /min.
42 V. Jimnez et al. / Applied Catalysis A: General 390 (2010) 3544

100 temperatures as low as 275 C, as evidenced by the negative val-


CO CO2 ues of XCO . As a result, the CO content is higher at the efuent of the
----- ------
80 reactor, compared to that in the feed, and continuously increases
, platelet
, ribbon with increasing temperature.
, fishbone Results of kinetic measurements (Table 2) obtained under con-
(%)

60
ditions of combined hydrogenation of CO/CO2 mixtures indicated
CO2

that the specic reaction rate of CO conversion depends on the


40
nature of CNF. In particular turnover frequency (TOF) of CO conver-
,X

sion at 250 C is about four times higher when Ru is supported on


CO

20 ribbon CNF (0.080 s1 ) compared to platelet (0.017 s1 ) and sh-


X

bone CNFs (0.023 s1 ). However, TOF of CO2 conversion does not


0 depend signicantly on the nature of CNFs supports, exhibiting
-100 similar values at a given temperature (Table 2). Similar values of
-200 TOF of CO and CO2 conversion have been previously reported over
200 250 300 350 400 Ru/Al2 O3 catalysts under the same experimental conditions [25].
The observed increase of TOF of CO conversion when Ru is dis-
Temperature (C) persed on ribbon CNF may be due to the larger Ru crystallite size
Fig. 7. Conversions of CO and CO2 as functions of reaction temperature obtained
of this sample compared to the other catalysts examined (Table 1),
over Ru (0.5 wt.%)/CNFs catalysts for the selective methanation of CO. Solid symbols: which is well known to enhance catalytic activity for the selective
CO conversion; open symbols: CO2 conversion. Experimental conditions: mass of methanation of CO [17,33,54,55]. For example, Panagiotopoulou et
catalyst: 150 mg; particle diameter: 0.18 < dp < 0.25 mm; feed composition: 1% CO, al. [17] reported, for Ru/TiO2 catalysts, that specic activity (TOF)
15% CO2 , 50% H2 (balance He); total ow rate: 200 cm3 /min.
increases by more than one order of magnitude with increasing
Ru crystallite size from 2.1 to 4.5 nm. It should be noted however,
amounts of CO may also produced under conditions of solo-CO2 that the higher catalytic activity of Ru/ribbon catalyst can be also
methanation, which are not detectable with the analysis system attributed to interactions between the Ru crystallites and the CFN
employed, or signicantly higher temperatures are required for the support, which may be vary depending on the nature of the support.
production of measurable amounts of CO. Clearly, a nal conclusion on the role of the CNF support on the cat-
alytic performance of Ru catalysts for the title reaction requires a
3.4. Selective methanation of CO detailed investigation of the reaction mechanism, which is beyond
the scope of the present study.
The effect of the type of CNFs (platelet, shbone and ribbon) used Selectivity to CH4 at a given temperature does not depend
as support on catalytic performance for the selective methanation appreciably on the nature of the CNF support. Typical results
of CO in CO/CO2 mixtures was investigated over Ru (0.5 wt.%) cat- obtained at 250 C are presented in Table 2. In all cases, SCH4
alysts and results obtained are summarized in Fig. 7. It is observed increases from 90 to 95% with increasing temperature from 225 C
that XCO over the 0.5% Ru/ribbon catalyst increases with increas- to 275 C (not shown for clarity). It should be noted that at tem-
ing temperature above 200 C and goes through a maximum of ca. peratures higher than 275 C the estimation of selectivities toward
25% at 250 C. Further increase of temperature results in a drastic hydrogenated products is doubtful, due to the production of high
decrease of XCO , which takes negative values above 275 C. This amounts of CO (i.e. negative values of XCO ) at low reaction tem-
behaviour can be explained by considering that, under the present peratures.
conditions, CO hydrogenation (Eq. (1)) runs in parallel with the
RWGS reaction (Eq. (3)), which becomes important above ca. 275 C. 3.5. Effect of water vapour on catalytic performance
At temperatures higher than 275 C the rate of CO production via
the RWGS is higher than the rate of CO consumption via the metha- Normally, the reformate gases obtained from fuel processors
nation reaction, thereby resulting in the observed net increase of CO contain considerable amounts of steam, which may affect CO
concentration at the efuent of the reactor. Regarding conversion of methanation characteristics. In order to investigate this issue,
CO2 , it is practically zero at temperatures up to ca. 250 C, i.e., until water vapour was added to the feed (30%) and results obtained
XCO reaches its maximum value, and then progressively increases over Ru catalysts supported on CNFs are summarized in Fig. 8.
with increasing temperature (Fig. 7). This behaviour reects the fact The corresponding conversion curves obtained in the absence of
that CO interacts more strongly with the catalyst surface, compared water are also shown for comparison. It is observed that in all
to CO2 , and is in good agreement with previous studies [17,25]. cases the conversion of CO at a given temperature increases signif-
The Ru catalysts supported on platelet and shbone CNFs are icantly with the addition of water in the reaction mixture (Fig. 8).
practically inactive for the selective methanation of CO under the This is accompanied by an increase of XCO maximum from 7% (at
present experimental conditions, reaching a maximum of XCO of 7% 250 C) to 40% and 47% (at 300 C) for Ru/platelet (Fig. 8a) and
at 250 C (Fig. 7). This is attributed to the RWGS, which is operable at Ru/shbone (Fig. 8b), respectively, and from 25% (at 250 C) to 53%

Table 2
Representative results of kinetic measurements obtained over the examined catalysts for the hydrogenation of CO and CO2 mixture, in the absence and in the presence of
30% H2 O in the feed. Mass of catalyst: 150 mg; particle diameter: 0.18 < dp < 0.25 mm; feed composition: 1% CO, 15% CO2 , 50% H2 , 0% or 30% H2 O (balance He); total ow rate:
200 cm3 /min.

Catalyst (0.5% Ru) Reaction

CO/CO2 /H2 CO/CO2 /H2 /H2 O

SCH4 at 250 C (%) TOFCO at 250 C (s1 ) TOFCO2 at 330 C (s1 ) SCH4 at 250 C (%) TOFCO at 250 C (s1 ) TOFCO2 at 330 C (s1 )

Platelet 93.2 0.017 0.56 100 0.058 0.51


Fishbone 93.1 0.023 0.39 100 0.043 0.38
Ribbon 91.8 0.080 0.26 100 0.070 0.50
V. Jimnez et al. / Applied Catalysis A: General 390 (2010) 3544 43

a reported that the CO2 methanation does not occur appreciably in


100
H2O Content (%) the presence of water in the feed.
0 30 Regarding the mechanism of CO/CO2 hydrogenation reactions, it
80 0.5% Ru/platelet
, XCO has been proposed that CO methanation proceeds via dissociation
XCO of carbon monoxide to C and O atoms, followed by their hydro-
X CO, X CO (%)

,
2
60 genation into CH4 and H2 O [17,54,55,65,66]. Results of Fig. 8 show
2

that this reaction is enhanced appreciably by the presence of water.


40 This indicates that methanation of CO is favoured, compared to the
reverse WGS, under these conditions. Concerning CO2 methana-
20 tion, it is believed to involve conversion of CO2 toward CO via the
RWGS (Eq. (3)), followed by CO hydrogenation to methane [17].
0 If this is the case, then increase of water concentration in the feed
should shift Eq. (3) to the left, thereby resulting in an increase of CO
-100
conversion in agreement with results of Fig. 8. This indicates that
-200
the addition of steam in the reaction mixture retards the RWGS
200 250 300 350 400
which is responsible for the negative values of XCO (Figs. 7 and 8).
Temperature (C)
The addition of water vapour in the gas mixture also affects
selectivity toward hydrogenation products. Results show (Table 2)
b 100
that for all catalysts examined selectivity toward methane is
H 2O Content (%)
0 30 improved in the presence of water. For example, SCH4 at 250 C
80 0.5% Ru/fishbone
, X CO for Ru/ribbon catalyst increases from 91.8 to 100% with increas-
X CO ing the concentration of water in the feed from 0 to 30% (Table 2).
X CO , X CO (%)

,
2
60 It has been reported that SCH4 always increases with increasing
2

XCO in a manner which does not depend on whether this is due


40 to an increase of reaction temperature [17,25], to the effect of
metal crystallite size [17], to the effect of the oxide support [17]
or to variations of space velocity [17,25]. Furthermore, Inui et al.
20
[67] demonstrated that the formation of higher hydrocarbons has
a strong retarding effect on CO methanation, which is released
0 with increase of CO conversion. It may then be suggested that
-100 the observed increase of SCH4 with the addition of water vapour
-200 (Table 2) is related to the increase of XCO (Fig. 8).
200 250 300 350 400 Results of kinetic measurements (Table 2) showed that, for
all catalysts examined, TOF of CO2 conversion remains practically
Temperature (C)
unaffected by the presence of water, whereas TOF of CO conversion
c 100 increases with increasing water content from 0 to 30%. In partic-
H 2O Content (%)
ular, TOF of CO over Ru/shbone catalyst increases by a factor of
0 30
80 , X CO 0.5% Ru/ribbon 2 (at 250 C) with the addition of 30% water in the feed. In the
, X CO case of Ru/ribbon catalyst TOFs of CO are practically the same in
X CO , X CO (%)

60
2 the absence and in the presence of steam at temperatures lower
than 250 C, but becomes higher in the latter case above 250 C (not
2

shown for clarity). The effect is more pronounced for the Ru/platelet
40
catalyst, for which the specic activity for the CO hydrogenation
reaction becomes more than three three times higher (at 250 C)
20
with the addition of water (Table 2).
Results of the present study show that the 0.5%Ru/CNFs catalysts
0 investigated exhibit similar catalytic activity with that obtained in
-100 previous studies over the well known Ru/Al2 O3 catalyst conducted
-200 under the same experimental conditions [25]. However, it has been
200 250 300 350 400 reported that catalytic performance for the title reaction could be
Temperature (C) further improved provided that catalyst characteristics are opti-
mized and operating conditions are properly selected [17,25]. For
Fig. 8. Effect of addition of water vapour in the feed (30%) on the catalytic per- example, it has been found that catalytic activity of Ru/TiO2 and
formance of Ru (0.5 wt.%) catalysts supported on (a) platelet, (b) shbone and Ru/Al2 O3 catalysts is greatly enhanced with increase of Ru load-
(c) ribbon for co-methanation of CO/CO2 mixture. Solid symbols: CO conver-
sion; open symbols: CO2 conversion. Mass of catalyst: 150 mg; particle diameter:
ing or crystallite size [17,33]. Furthermore, results presented in
0.18 < dp < 0.25 mm; feed composition: 1% CO, 15% CO2 , 50% H2 030% H2 O (balance Figs. 58 were obtained with a relatively high space velocity (SV)
He). of ca. 48,800 h1 , compared to that usually used (5000 h1 ) in fuel
cell applications. It was found over 0.5% Ru/Al2 O3 catalysts that
(at 300 C) for Ru/ribbon (Fig. 8c) CNF. Regarding CO2 hydrogena- decreasing SV from 48,800 h1 to 12,200 h1 resulted in a shift of
tion, it is observed that this reaction remains practically unaffected the XCO curves toward lower temperatures by ca. 80 C, whereas
by increasing water content from 0 to 30% (Fig. 8) for all samples the XCO2 curve was not affected signicantly [17]. Thus, catalytic
investigated. Different results have been previously reported over performance of the investigated 0.5% Ru/CNFs catalysts could be
noble metal catalysts supported on Al2 O3 [25] and Ru/TiO2 cata- further improved by decreasing SV and/or increasing Ru loading
lysts [17], where it was found that addition of water vapour in the (i.e. crystallite size), leading to completely methanation of CO and
feed does not affect CO hydrogenation but shifts the CO2 conver- making them suitable for use in the selective methanation of CO
sion curve toward higher temperatures. Similarly, Batista et al. [58] for fuel cell applications.
44 V. Jimnez et al. / Applied Catalysis A: General 390 (2010) 3544

4. Conclusions [20] T.V. Herwijnen, H.V. Doesburg, W.A. Du Jong, J. Catal. 28 (1973) 391402.
[21] K.O. Xavier, R. Screekala, K.K.A. Rashid, K.K.M. Yusuff, B. Sen, Catal. Today 49
(1999) 1721.
Results of the present study show that catalytic performance [22] T. Utaka, T. Takeguchi, R. Kikunchi, K. Eguchi, Appl. Catal. A: Gen. 246 (2003)
and selectivity to reaction products over Ru catalysts supported 117124.
on different types of CNFs depend strongly on the experimen- [23] T. Inui, M. Funabiki, M. Suehiro, T. Sezume, J. Chem. Soc., Faraday Trans. 1 75
(1979) 787802.
tal conditions used (e.g. reaction temperature, solo-methanation [24] F. Solymosi, A. Erdhelyi, J. Mol. Catal. 8 (1980) 471474.
of CO or CO2 , co-methanation of CO/CO2 mixtures, addition of [25] P. Panagiotopoulou, D.I. Kondarides, X.E. Verykios, Appl. Catal. A: Gen. 344
water vapour etc.). Under solo-CO methanation conditions, CO con- (2008) 4554.
[26] F. Solymosi, A. Erdhelyi, T. Bnsgi, J. Catal. 68 (1981) 371382.
version reaches 100% at 340 C over all catalysts examined, with [27] K.B. Kester, E. Zagli, J.L. Falconer, Appl. Catal. 22 (1986) 311319.
Ru/platelet sample being somewhat more active than Ru/shbone [28] G.A. Somorjai, Catal. Rev. Sci. Eng. 23 (1981) 189202.
and Ru/ribbon. Under solo-CO2 methanation conditions, catalytic [29] W.J. Wang, Y.W. Chen, Appl. Catal. A: Gen. 77 (1991) 2136.
[30] S. Takenaka, K. Kawashima, H. Matsune, M. Kishida, Appl. Catal. A: Gen. 321
performance is not affected by the nature of the CNFs used as sup-
(2007) 165174.
port. In co-methanation of CO/CO2 , hydrogenation of CO is the [31] M.B.I. Choudhury, S. Ahmed, M.A. Shalabi, T. Inui, Appl. Catal. A: Gen. 314 (2006)
predominant reaction until a certain temperature is reached, above 4753.
which conversion of CO starts to decrease due to the onset of the [32] Y. Men, G. Kolb, R. Zapf, V. Hessel, H. Lwel, Catal. Today 125 (2007) 8187.
[33] R.A. Dagle, Y. Wang, G.G. Xia, J.J. Strohm, J. Holladay, D.R. Palo, Appl. Catal. A:
RWGS reaction and hydrogen starts to be unselectively consumed Gen. 326 (2007) 213218.
to methanate CO2 . All catalysts examined tend to enhance the [34] C. Galletti, S. Specchia, G. Saracco, V. Specchia, Chem. Eng. Sci. 65 (2010)
undesired RWGS reaction, exhibiting low activity for CO hydro- 590596.
[35] F. Sapountzi, M.N. Tsampas, C.G. Vayenas, Topics Catal. 44 (2007) 461468.
genation reaction. However, addition of 30% water vapour in the [36] C.G. Vayenas, M.N. Tsamplas, A. Katsaounis, Electrochemical. Acta 52 (2007)
feed enhances methanation of CO, inhibiting the RWGS reaction 22442256.
for all catalysts investigated, whereas CO2 hydrogenation is not [37] S.Y. Wang, S.H. Moon, M.A. Vannice, J. Catal. 98 (1986) 166177.
[38] J.D. Bracey, R. Burch, J. Catal. 86 (1984) 384391.
affected by the presence of steam. For all experimental conditions [39] C. Daz-Taboada, J. Batista, A. Pintar, J. Levec, Appl. Catal. B 89 (2009) 375382.
investigated selectivity toward CH4 increases with increasing tem- [40] V. Jimnez, A. Nieto-Mrquez, J.A. Daz, R. Romero, P. Snchez, J.L. Valverde, A.
perature at the expense of higher hydrocarbons and increases by Romero, Ind. Eng. Chem. Res. 48 (2009) 84078417.
[41] B.D. Cullity, Elements of X-ray Diffraction, second ed., Addison-Wesley, Read-
adding water in the reaction mixture. Among the various Ru/CNFs ing, MA, 1978.
catalysts investigated, optimal results were obtained over the [42] M.J. Ramos, A. de Lucas, V. Jimnez, P. Snchez, J.L. Valverde, Fuel Process.
Ru/ribbon catalyst, which exhibits maximum CO conversion of 53% Technol. 89 (2008) 721727.
at 300 C under realistic feed composition. [43] P. Snchez, F. Dorado, A. Fnez, V. Jimnez, M.J. Ramos, J.L. Valverde, J. Mol.
Catal. A: Chem. 273 (2007) 109113.
[44] F. Rodrguez, I. Rodrguez, C. Moreno, A. Guerrero, J.D. Lpez, J. Catal. 99 (1986)
Acknowledgement 171183.
[45] F. Dorado, R. Romero, P. Canizares, Appl. Catal. A: Gen. 236 (2002) 235243.
[46] J. Zhou, Z. Sui, X. Zhou, W. Yuan, Chin. J. Catal. 29 (2008) 11071112.
The authors gratefully acknowledge nancial support from Con- [47] K.S.W. Sing, R.T. Williams, Adsorpt. Sci. Technol. 22 (2004) 773782.
sejera de Ciencia y Tecnologa de la Junta de Comunidades de [48] T.V. Reshetenko, L.B. Avdeeva, Z.R. Ismagilov, A.L. Chuvilin, V.A. Ushakov, Appl.
Castilla-La Mancha (Projects PBI-05-038 and PCI 08-0020-1239). Catal. B: Gen. 247 (2003) 5163.
[49] M. Montiel, S. Garca-Rodrguez, P. Hernndez-Fernndez, R. Daz, S. Rojas, J.L.
Garca-Fierro, E. Fats, P. Ocn, J. Power Sources 195 (2010) 24782487.
References [50] V.I. Zaikovskii, K.S. Nagabhushana, V.V. Kriventsov, K.N. Loponov, S.V.
Cherepanova, R.I. Kvon, H. Bnnemann, D.I. Kochuby, E.R. Savinova, J. Phys.
[1] N. Fujiwara, K. Yasuda, T. Ioroi, Z. Siroma, Electrochim. Acta 47 (2002) Chem. B 110 (2006) 68816890.
40794084. [51] P. Tribolet, L. Kiwi-Minsker, Catal. Today 105 (2005) 337343.
[2] C.G. Farrell, C.L. Gardner, M. Ternan, J. Power Sources 171 (2007) 283293. [52] J.H. Zhou, Z.J. Sui, J. Zhu, P. Li, D. Chen, Y.C. Dai, W.K. Yuan, Carbon 45 (2007)
[3] X. Cheng, Z. Shi, N. Glass, L. Zhang, J. Zhang, D. Song, Z.S. Liu, H. Wang, J. Shen, 785796.
J. Power Sources 165 (2007) 739756. [53] I.I. Salame, T.J. Bandosz, J. Colloid Interface Sci. 264 (2003) 307312.
[4] Y. Borodko, G.A. Somorjai, Appl. Catal. A: Gen. 186 (1999) 355362. [54] M. Ojeda, S. Rojas, M. Boutonnet, F.J. Prez-Alonso, F.J. Garca-Garca, J.L.G.
[5] J. Divisek, H.F. Oetjen, V. Peinecke, V.M. Schmidt, U. Stimming, Electrochim. Fierro, Appl. Catal. A: Gen. 274 (2004) 3341.
Acta 43 (1998) 38113815. [55] M. Ojeda, S. Rojas, F.J. Garca-Garca, M. Lpez Granados, P. Terreros, J.L.G. Fierro,
[6] E. Passalacqua, F. Lufrano, G. Squadrito, A. Patti, L. Giorni, J. New Mater. Elec- Catal. Commun. 5 (2004) 703707.
trochim. Syst. 3 (2000) 141146. [56] M. Araki, V. Ponec, J. Catal. 44 (1976) 439448.
[7] D.L. Trimm, Appl. Catal. A: Gen. 296 (2005) 111. [57] N.W. Cant, A.T. Bell, J. Catal. 73 (1982) 257271.
[8] R.W. McCabe, P.J. Mitchell, J. Catal. 103 (1987) 419425. [58] M.S. Batista, E.I. Santiago, E.M. Assaf, E.A. Ticianelli, J. Power Sources 145 (2005)
[9] G. Kolb, V. Hessel, V. Cominos, C. Hofman, H. Lwer, G. Nikolaidis, R. Zapf, A. Zio- 5054.
gas, E.R. Delsman, M.H.J.M. de Croon, J.C. Schouten, O. de la Iglesia, J. Santamaria, [59] M.V. Twigg, Catalyst Handbook, second ed., Wolfe Publishing, London, 1989.
Catal. Today 120 (2007) 220. [60] F. Solymosi, A. Erdhelyi, M. Kocsis, J. Chem. Soc., Faraday Trans. 77 (1981)
[10] A. Luengnaruemitchai, S. Osuwan, E. Gulari, Int. J. Hydrogen Energy 29 (2004) 10031012.
429435. [61] A. Trovarelli, C. Mustazza, G. Dolcetti, J. Kaspar, M. Graziani, Appl. Catal. 65
[11] W.L. Deng, J.D. Jess, H. Saltsburg, M. Flytzani-Stephanoloulos, Appl. Catal. A: (1990) 129142.
Gen. 291 (2005) 126135. [62] K. Yaccato, R. Carhart, A. Hagemeyer, A. Lesik, P. Strasser, A.F. Volpe Jr., H. Turner,
[12] Y. Borodko, G.A. Somojai, Appl. Catal. A: Gen. 186 (1999) 355362. H. Weinberg, R.K. Grasseli, C. Brooks, Appl. Catal. A 296 (2005) 3048.
[13] Y. Men, G. Kolb, R. Zapf, V. Hessel, H. Lwe, Catal. Today 125 (2007) 8187. [63] A. Erdhelyi, M. Psztor, F. Solymosi, J. Catal. 98 (1986) 166177.
[14] A.L. Kustov, A.M. Frey, K.E. Larsen, T. Johannessen, J.K. Norskov, C.H. Christensen, [64] P. Panagiotopoulou, D.I. Kondarides, X.E. Verykios, Ind. Eng. Chem. Res. (2010),
Appl. Catal. A: Gen. 320 (2007) 98104. doi:10.1021/ie100132g.
[15] S. Takenaka, T. Shimizu, K. Otsuka, Int. J. Hydrogen Energy 29 (2004) 10651073. [65] T. Hanaoka, H. Arakawa, T. Matsuzaki, Y. Sugi, K. Kanno, Y. Abe, Catal. Today 58
[16] O. Grke, P. Pfeifer, K. Schubert, Catal. Today 110 (2005) 132139. (2000) 271280.
[17] P. Panagiotopoulou, D.I. Kondarides, X.E. Verykios, Appl. Catal. B: Environ. 88 [66] A.E. Aksoylu, A.N. Akin, Z.I. nsan, D.L. Trimm, Appl. Catal. A 145 (1996)
(2009) 470478. 185193.
[18] I. Fujita, N. Takezawa, Chem. Eng. J. 68 (1997) 6368. [67] T. Inui, M. Funabiki, Y. Takegami, Ind. Eng. Chem., Prod. Res. Dev. 19 (1980)
[19] A.E. Aksoylu, Z.I. nsan, Appl. Catal. A.: Gen. 164 (1997) 111. 385388.

View publication stats

You might also like