You are on page 1of 12

Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136

Contents lists available at SciVerse ScienceDirect

Journal of Molecular Catalysis A: Chemical


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

Pt monometallic and bimetallic catalysts prepared by acid sol–gel method for


liquid phase reforming of bioglycerol
M. El Doukkali a,∗ , A. Iriondo a , J.F. Cambra a , L. Jalowiecki-Duhamel b , A.S. Mamede d ,
F. Dumeignil b,c , P.L. Arias a
a
University of the Basque Country (UPV/EHU), Engineering School, C/Alameda Urquijo s/n, 48013 Bilbao, Spain
b
Université des Sciences et Technologies de Lille, CNRS UMR8181, Unité de Catalyse et de Chimie du Solide (UCCS), F-59655 Villeneuve d’Ascq Cedex, France
c
Institut Universitaire de France, 10 Boulevard Saint-Michel, 75005 Paris, France
d
Ecole Nationale Supérieure de Chimie de Lille, CNRS UMR8181, Unité de Catalyse et de Chimie du Solide UCCS, F-59655 Villeneuve d’Ascq Cedex, France

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

Article history: PtFe, PtCo, PtNi, Ni and Pt-based ␥-Al2 O3 catalytic systems were prepared using the acid sol–gel method
Received 3 October 2012 and characterized by appropriate techniques. Their catalytic activities in glycerol APR for H2 production
Received in revised form 1 December 2012 were investigated, and the most significant results will be fully discussed in this paper. The effect of
Accepted 9 December 2012
operating conditions, the influence of different non-noble incorporated metals, the interesting promoting
Available online 20 December 2012
effect between Pt and Ni and their dispersion effect on the catalytic performance have been the main focus
of this research. The analyses by TPR-H2 , XRD, XPS and TEM/EDX techniques confirmed high Pt and Ni
Keywords:
dispersions and the promoting effect between both elements. All this explains the best catalytic activity
Platinum–nickel
Acid sol–gel
of PtNiAl(SGA) catalyst, especially at moderate temperature/pressure reaction conditions. Highly dispersed
Promoting effect Pt/Ni and Pt spillover effects were the key for obtaining high H2 production. The Pt monometallic catalyst
Aqueous phase reforming recorded lower conversions but fairly stable activity, while the Ni-containing catalysts recorded higher
Glycerol conversions, but some initial deactivation until steady state was reached.
Hydrogen © 2012 Elsevier B.V. All rights reserved.

1. Introduction considerable attention because of its advantages: (i) low tempera-


ture, thereby reducing energy consumption because the feed needs
Fossil fuels have been the main resources used to meet energy not to be vaporized; (ii) limited decomposition reactions; (iii) the
demands all over the world. However, the use of large quantities of production of a gas stream rich in H2 and CO2 and very poor in
fossil fuels is causing the depletion of their reservoirs, global warm- CO in a simple chemical reactor. This last characteristic of the gas
ing and climate change [1], as well as damaging the environment produced renders it suitable for use in fuel cells [5].
[2]. All these drawbacks in the use of fossil fuels have generated As reported in the literature [9–14], many aqueous solutions
increasing interest in new sustainable and renewable clean fuels. of oxygenated compounds have been used as feeds in aqueous
In this sense, hydrogen is considered one of most promising future phase reforming (APR). However, the aqueous solution of glycerol
clean energy vectors. is widely used due to the possibility of revalorizing this by-product
Over the past decade, hydrogen generation from biomass- of biodiesel production through the transesterification of fats from
derived products and by-products has been considered one of biomass, as only small amounts of this glycerol can be absorbed by
the best alternatives when compared to traditional methods of its traditional markets [15].
hydrogen production, which use fossil fuels as raw materials. The During the APR process, apart from the ideal reforming reaction
conversion processes of biomass-derived materials, by contrast, (C3 H8 O3 + 3H2 O → 7H2 + 3CO2 ) [2], other reactions take place, such
present environmental benefits, as the CO2 produced is absorbed as water-gas-shift (WGS), methanation or Fischer–Tropsch ones to
during biomass growth, providing a neutral CO2 balance [3–5]. produce alkanes, partial cracking of C C/C H/C O bonds, dehydro-
One of the processes for producing hydrogen from biomass-derived genation and dehydration/hydrogenation [9–13]. In the case of C O
products [6–8], namely, the aqueous phase catalytic reforming pro- bond cleavage, this reaction leads to diol and alkane production and
cess [9–11] developed by Dumesic and co-workers, has received is favored by catalyst acidity [16]. Moreover, the dehydrogenation
and dehydration/hydrogenation of glycerol also generate diols [16].
The APR process therefore includes many reaction pathways that
∗ Corresponding author. Tel.: +34 946 017 297; fax: +34 946 014 179. are highly dependent on the catalyst used [6]. Thus, it is important
E-mail address: mohamed.eldoukkali@ehu.es (M. El Doukkali). to develop an active catalyst for C C/C H bond cleavage and WGS

1381-1169/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.molcata.2012.12.006
126 M. El Doukkali et al. / Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136

reaction in order to increase the selectivity toward H2 production, H2O/AIP solution (Molar ratio = 100)
while the parallel and consecutive reactions are either limited [5]
or the intermediate products they generate fully reformed. HNO3 (pH controlled at 3.5) Heating to 40 ºC
and stirring for 1 h
According to the literature [17,18], catalyst activity is highly
influenced by the preparation method used. Conventional prepara- Transparent alumina Sol
tion methods, such as incipient wetness impregnation, also called
Vigorously stirring for 1 h at Fe, Co, Ni or/and Pt
wet impregnation, commonly produce active phase agglomeration
40 ºC + 30 min in ultrasound precursors + C2H5OH
on the support used. This phenomenon, which is attributed to large-
sized particle formation, occurs especially when high contents of Gel precipitate
metal precursors are incorporated [17,18]. However, the sol–gel
Solvent evaporation
method presents some advantages when compared with impreg- Gellation
at 40 ºC for 24 h
nation methods, such as higher active metal dispersion, great
thermal resistance to sintering and low deactivation rate [17–23], Gel => rigid solid
thus can lead to more stable catalysts [17]. Moreover, this method Drying in an oven at
allows a better control of catalyst preparation via various parame- 100 ºC for 12 h
ters, such as: temperature, type and concentration of the alkoxide,
Dry solid
the solvent, and the pH of the preparation medium. Therefore, vary-
ing these factors allows a more flexible control of the catalysts Calcination at 550 ºC
textural properties, such as surface area, pore size distribution and for 6 h under air
metal dispersion [23].
Fresh Catalysts; PtFe, PtCo, PtNi, Ni
A variety of catalytic systems have been described in the
and Pt-based γ-Al2O3
literature for aqueous phase reforming of glycerol and other
biomass-derived products [1,2,5,6,9–13,15,24–26]. These catalytic
Fig. 1. Schematic flow of the methodology used to prepare the catalysts by the acid
systems are based on Pt, Ni, Cu, Pd or Co monometallic catalysts sol–gel method.
and PtRe and PtNi bimetallic catalysts. Pt metal phase is usually
used in both monometallic and bimetallic catalysts because of
its high activity for C C/C H bond cleavage and WGS reaction, First, aluminum isopropoxide (AIP) Al[OCH(CH3 )2 ]3
and low selectivity toward methane production via the metha- (Sigma–Aldrich 99.995%) was hydrolyzed at 40 ◦ C, using a
nation reaction [4]. According to Huber et al. [9], Pt/Al2 O3 SiO2 molar ratio H2 O/AIP of 100. This solution was then peptized
catalyst recorded higher activity in the APR of ethylene glycol by adding HNO3 to adjust the pH at 3.5. This pH was chosen
than other silica-supported Group VIII metal catalysts. A similar to be under acid conditions while the preparation took place.
conclusion was reached by Iriondo et al. [15], whereby both glyc- This plays an important role in the evolution of the hydrolysis
erol conversion and gaseous products increased when Pt and Ni reaction [23]: H3 O+ ions (acid medium) and OH− (basic medium)
were combined on ␥-Al2 O3 and ␥-Al2 O3 La2 O3 supported cata- do not have the same influence. The H3 O+ cations, attracted by
lysts. They concluded that Pt catalyst activity can be improved by oxygen, facilitate the substitution of the groups O CH(CH3 )2
the addition of Ni, Co and Fe. and, therefore, hydrolysis, while the OH− anions attracted by the
Based on this background, and consistent with our previous electronegative metal M favor the formation of M O M clusters
research work [26], the main objective here is to study new by condensation. As a result pH selection will be decisive to
Pt/transition-metal (Fe, Co and Ni) catalysts prepared via sol–gel determine the final texture of the materials prepared [23]. At the
method under acid conditions (SGA) and compare their catalytic same time, another solution was prepared using absolute ethanol
activities and stabilities with those obtained using similar Pt/Ni as solvent and the salt precursor (H2 PtCl6 ·xH2 O (Sigma–Aldrich,
catalysts prepared by sol–gel under basic conditions (SGB) and 99.99%), Ni(NO3 )2 ·6H2 O (Sigma–Aldrich, 99.99%), FeCl3 ·6H2 O
incipient wetness impregnation (IWI), which have been published (Panreac, 99.5%) and CoCl2 ·6H2 O (Quimivita)) as solute. The metal
in a previous work [26]. In fact, the selected catalytic systems con- concentration in the precursor solution was selected to obtain
sist of ␥-Al2 O3 with PtNi, PtCo and PtFe as active phases. Pt and 3 wt% of Pt and 10 wt% of Ni, Co or Fe. This salt solution was
Ni monometallic catalysts supported on ␥-Al2 O3 were also pre- added to the AIP-derived alumina sol at 40 ◦ C and the mixture was
pared for both comparative purposes and better understanding of vigorously stirred. After this, the sol formed was placed in ultra-
the properties of PtNi catalyst and their influences on catalytic sound equipment for 30 min. The alcoholic solvent was removed
activity, as commented in Section 2.3. To facilitate a comparison by drying at 40 ◦ C for 24 h. The rigid solid materials obtained by
between the results reported in this work and those shown in the SGA method were dried under air at 100 ◦ C overnight and
[26], all the parameters and the operating conditions remained then calcined at 550 ◦ C with a temperature ramp of 2 ◦ C/min for
unchanged. This research aims at the development of new cata- 6 h. The calcination procedure was carried out under an air flow
lysts for the aqueous phase reforming of glycerol, and therefore it of 2 L min−1 g−1
cat . The preparation procedure described above is
investigates their catalytic activity and stability in order to max- schematically described in Fig. 1.
imize glycerol conversion, especially to gaseous products, and to
maximize H2 yield. 2.2. Catalyst characterization techniques

In order to understand the catalytic behavior of the tested cata-


lysts, their physicochemical properties were examined at calcined
2. Experimental procedures state using different characterization techniques. The catalysts pre-
senting the best performances were also characterized at their
2.1. Catalyst preparation and reactants used reduced state to get a better understanding of their functionalities.
Specifically, calcined catalysts were characterized by induc-
As mentioned above, PtFe, PtCo, PtNi, Ni and Pt-based ␥-Al2 O3 tively coupled plasma atomic emission spectroscopy (ICP-AES),
catalysts were prepared by the SGA method following a modified N2 adsorption–desorption isotherms at 77 K, temperature pro-
preparation procedure as described in [17–20,23,26]. grammed reduction (TPR-H2 ), X-ray powder diffraction (XRD) and
M. El Doukkali et al. / Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136 127

X-ray photoelectron spectroscopy (XPS). The last two characteriza- transmission electronic microscopy, using a TEM/EDX FEI Tecnai
tion techniques were also used to characterize reduced catalysts. G2 20 equipped with a LaB6 filament, a super twin lens operating
The reduction treatment was carried out in situ within XRD and at 200 kV and EDX beam of 5 nm. Bright field images were acquired
XPS chambers using similar conditions to those used in the catalytic using a CCD camera (TVIPS GmbH). Powders of the reduced catalyst
activity tests (Section 2.3). samples were dispersed in ethanol with ultrasound, and drops were
The chemical composition of calcined catalysts was determined deposited on copper grids coated with amorphous carbon films. The
by the ICP-AES technique using a Perkin-Elmer Optima 3300DV TEM images were taken using the high resolution mode of trans-
apparatus. Around 100 mg of solid samples were first dissolved in mission microscopy. The energy dispersion X-ray was used by an
an acid solution mixture [HNO3 (3 mL), HCl (2 mL) and HF (3 mL)] EDX detector to analyze the local chemical composition of some
using a microwave with a temperature profile of 7 ◦ C/min up to catalyst particles.
180 ◦ C for 45 min. The solution obtained was diluted in water to
concentrations within the instrument’s detection range. 2.3. Activity tests
The textural properties were evaluated using the N2
adsorption–desorption technique in an Autosorb-1C instrument The catalytic activities of PtFe, PtCo and PtNi catalysts were
(Quantachrome). Prior to the analysis, the samples were outgassed measured in the bench-scale fixed-bed reactor of a Microactivity
for 12 h at 150 ◦ C. Specific areas were calculated from these modular laboratory system (PID Eng. & Tech.). In order to under-
isotherms using the BET method and taking a value of 0.162 nm2 stand the influence of the different active phases of the best Pt–Ni
for the cross-section of the physically adsorbed N2 molecule. bimetallic catalysts, Pt and Ni monometallic catalysts were also
Pore distribution and pore volume were estimated using the BJH prepared and tested under the same operating conditions.
method (desorption data). The fixed bed reactor was tubular (316-stainless steel), where
The same Autosorb-1C instrument equipped with a TCD detec- 1 g of catalyst, sieved to 0.42–0.5 mm, was diluted in inert SiC.
tor was used to analyze the reducibility of the catalysts using the The catalyst/SiC ratio was 1:3 (w/w). Prior to the catalytic tests,
TPR-H2 technique. Prior to reduction experiments, around 50 mg the Pt and PtFe catalysts were reduced at 350 ◦ C, while the PtCo,
of solid sample were thermally treated under a He stream at 300 ◦ C PtNi and Ni catalysts were reduced at 750 ◦ C. The temperature
in order to remove impurities, such as water and other contami- was increased from ambient temperature to the desired reduction
nants. TPR profiles were obtained heating the sample from 25 to temperature (TR ) using 150 mL/min of N2 over 90 min. The tem-
1000 ◦ C at a linearly programmed rate of 10 ◦ C/min, while a mix- perature was then maintained at TR for 2 h using 100 mL/min of
ture of 10 vol% H2 /Ar flowed at 50 mL/min through the sample. A H2 /N2 mixture (75 vol% of H2 ). After this in situ reduction treat-
cold trap was used to remove water from the H2 /Ar effluent before ment, the system was purged with N2 to remove the remaining H2
using TCD to measure the amount of hydrogen consumed during and the H2 O formed during the reduction treatment. The catalytic
catalyst reduction. tests were performed using an aqueous solution of 10 wt% of glyc-
X-ray powder diffraction patterns of calcined catalysts were erol/water mixture. This solution was directly fed into the reactor
recorded with a D5000 Siemens diffractometer (D8 Advance Bruker −1
with a HPLC pump maintaining a WHSV at 2.6 g glycerol g−1 cat h .
AXS) using a copper target (Cu K␣1 = 1.5418 Å ) and a secondary The temperature and pressure fixed in the activity test were 250 ◦ C
beam monochromator. Another D5000 Siemens diffractometer and 50 bar for bimetallic catalysts. The catalyst with the best per-
equipped with chamber analysis (Anton Paar XRK900) was used to formance was also tested at 230 ◦ C and 30 bar. The operating
characterize the catalysts reduced in situ. Both apparatuses worked temperature/pressure conditions were chosen to operate under
with X-ray tube at 40 kV and 40 mA, and could detect crystalline APR conditions. Aspen Plus® simulation was used to estimate the
phases with average sizes above 2 nm. The scans were taken in a temperature and pressure values for which the mixtures used are
2 = 10–80◦ range with a step size of 0.04◦ /s of acquisition time. The in the liquid phase. It is important to note that a new charge of
crystallite size values were obtained using the Scherrer equation. the same catalyst was prepared for each operating condition. The
The XPS spectra of Pt and Ni elements were recorded using high pressure used in this kind of process was maintained with N2 ,
a Kratos Axis UltraDLD electron spectrometer equipped with a which was directly fed into the gas–liquid phase separator to avoid
monochromatic Al K␣ X-ray source (1486.6 eV) and a hemispheri- increasing the space velocity. In addition, to avoid any mixing of
cal electron analyzer operating at constant pass energy (40 eV). The N2 (used to entrain the gaseous products from the Peltier) with the
apparatus operated with a power of 150 W, 10 mA and 15 kV. Bind- reactants in the reactor, a “non-return valve” was placed at the inlet
ing Energy (BE) values were referred to the C1s peak (284.6 eV) of the Peltier (see apparatus set-up in [26]).
for the calcined and reduced catalysts. The C1s, O1s, Al2p, Al2s, The stream gas produced was periodically analyzed with an
Pt4d and Ni2p core-level spectra were recorded. Superficial quan- on-line ␮GC (TCD/SM). The oxygenated hydrocarbon products and
tification was carried out by the level Al2s to avoid the influence non-reacted glycerol were condensed at 15 ◦ C (vapor–liquid sepa-
of any overlap between the peaks of Pt4f and Al2p. Therefore, rator), collected and finally analyzed by GC equipped with FID and
Pt4d was used here instead of Pt4f. The calcined samples were TCD detectors.
degasified at 10−8 bar and transferred immediately to the ion- Catalytic activity was evaluated using the following activity
pumped analysis chamber, whose residual pressure was kept below parameters [10,15,26–28].
(7–9) × 10−12 bar during data acquisition. Catalyst reduction was
carried out under the abovementioned conditions in the spectrom-
• Glycerol conversion:
eter’s preparation chamber, directly connected to the ultra-high
vacuum analysis system. CasaXPS software was used for XPS
in
NGlycerol out
(mol/min) − NGlycerol (mol/min)
spectra treatment. Different peak areas were estimated by cal-
GC (%) = × 100 (1)
culating the integral of each peak after subtracting the Shirley in
NGlycerol (mol/min)
background and fitting the experimental peak to a combination of
Lorentzian/Gaussian lines of variable proportions. The normaliza- • Conversion into gas:
tion of the estimated areas of the corresponding species allowed
calculating the relative atomic ratios.
Catomes in gas products (mol/min)
In addition to the analysis carried out by the techniques CIG (%) = × 100 (2)
mentioned above, the reduced catalysts were also analyzed by Catomes in gas products (mol/min)
128 M. El Doukkali et al. / Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136

• Flow of total gas formed: PtOx


Neighboring
−1 −1 measured volume (mL) 273 K Fe-oxides
FT (mL(STP)) min gcat. = × (3)
time (min) × Mcat. (g) Tambient
Fe-aluminate
The gaseous stream flow was normalized by catalyst mass to PtFeAl(SGA)
render it easier to compare with similar data published in the lit- PtOx CoAlxOy
Co3O4
erature. CoCl2, 6H2O CoAl2O4

CoO
• H2 selectivity: PtCoAl(SGA)

N(Hout (mol/min)

Intensity (a.u.)
2) 3 NiO-Al2O3
NiAlxOy
S(H2 ) (%) = out
× × 100 (4) NiAl2O4
NC in gas phase (mol/min) 7 PtOx-NiO

3/7 is the reforming ratio of CO2 /H2 for glycerol taking into
account the occurrence of the WGS reaction after glycerol decom- PtNiAl(SGA)
position. NiAlxOy
NiO-Al2O3
• Yield gas of H2 and alkanes: NiAl2O4

out (mol/min)
N(i) NiO
Y (i) = × 100 (5) NiAl(SGA)
in
Nglycerol (mol/min)
PtOx
where (i) are H2 or alkane products such as: CH4 , C2 H6 and C3 H8 .
PtAl(SGA)

• Yield of oxygenated hydrocarbon compounds (OHCs):


0 150 300 450 600 750 900 1050
out
Nliquid (mol/min) (exp.)
products Temperature (ºC)
Y OHCs (j) = (6)
in
Nglycerol (mol/min)
Fig. 2. TPR-H2 profiles of PtFe, PtCo, PtNi, Ni and Pt-based ␥-Al2 O3 catalysts.
where (j) are C1 (methanol), C2 (acetaldehyde, acetic acid, ethyl-
ene glycol) or C3 (propanal, 1&2-propanol, acetol, 1,2-propanediol, peaks in a low temperature range (130–320 ◦ C) were assigned to
excluding unconverted glycerol). the reduction of PtOx and, possibly, neighboring Fe oxide, which
was reduced by hydrogen spillover at low temperatures. The
3. Experimental results second and ill-defined high temperature hump, between approx-
imately 680 ◦ C and 1030 ◦ C could also be decomposed into two
3.1. Catalyst characterization results peaks, which can be attributed to the presence of trace amounts of
Fe strongly interacting with ␥-Al2 O3 originating a phase similar to
3.1.1. Chemical composition and textural properties iron–aluminate [30,31].
Chemical composition, as determined by the ICP-AES technique, For the PtCoAl(SGA) catalyst, two broad components were
and data about textural properties obtained by N2 adsorption– observed in its TPR-H2 profile. The first component was decom-
desorption isotherms for calcined catalysts are presented in Table 1. posed into four overlapping peaks between 50 ◦ C and 500 ◦ C. The
The metal amounts incorporated are slightly different from the second component was also decomposed into two other peaks
nominal values, while the catalysts have a fairly similar chemical at high temperatures (T > 500 ◦ C). The shoulder at 160 ◦ C can be
composition. attributed to the decomposition of residual cobalt chloride to
According to SBET values, in all cases the metal incorporation Co3 O4 . This shoulder depends mainly on the catalyst preparation
reduced the specific surface area of the catalysts when compared to method and calcination conditions. The following two peaks, at
the bare support. This suggests that the metal particles block espe- around 260 ◦ C, are due to the reduction of relatively large PtOx
cially small pores, as the pore diameter data confirm, resulting in an particles and Co3 O4 to CoO. The peak at approximately 330 ◦ C can
increase in the average pore diameter (Table 1). Panyad et al. [29] be attributed to the reduction of CoO to metallic Co. The com-
observed a similar effect among fresh and used sol–gel catalysts, paratively sharp peaks appearing at around 650 ◦ C and 770 ◦ C can
which was attributed to the plugging of small pores. The bimetal- be attributed to the reduction of cobalt–aluminate compounds
lic catalysts had a lower specific surface area than NiAl and PtAl CoAlx Oy and CoAl2 O4 , respectively. These two phases require high
monometallic catalysts. This observation suggests that the higher temperatures for their reduction because they strongly interact
the metal amount incorporated, the higher the SBET loss related to with ␥-Al2 O3 . This TPR-H2 diagram is fairly similar to analogous
small pore blockage. ones reported in the literature for similar systems [32,33].
For Pt and PtNi catalysts, H2 consumption can be observed
3.1.2. TPR-H2 analysis of catalysts in their reduction profiles between 200 ◦ C and 300 ◦ C. The cor-
As commented previously, TPR-H2 measurements were carried responding peak can be assigned to the reduction of Pt species
out to investigate the reducibility of the calcined PtFe, PtCo, PtNi, interacting with the oxygen of ␥-Al2 O3 . The explanation for Pt
Ni, Pt-based ␥-Al2 O3 catalysts, as well as to examine the interaction oxide detection is related to the ability of ␥-Al2 O3 to stabilize small
between metal species and alumina support in the catalysts. The particles of PtOx scattered widely over its surface, as described
TPR-H2 profiles obtained are shown in Fig. 2. elsewhere [15,26,34].
In the case of the PtFeAl(SGA) catalyst, the reducibility profile For Ni-containing catalysts, two clearly reducing compo-
revealed two components. Gaussian deconvolution was used to nents were observed in their TPR-H2 profiles, which have been
decompose each component into two peaks. The first overlapping decomposed into four peaks by Gaussian deconvolution. As the
M. El Doukkali et al. / Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136 129

Table 1
Chemical composition and N2 isotherms of calcined catalysts.

Symbol Catalyst formula Composition (wt%) N2 isotherms

Pt Ni Co Fe SBET (m2 /gcat. ) dPore (nm) VPore (cm3 /gcat. )

Al(SGA) ␥-Al2 O3 – – – – 303 2.8 0.22


PtNiAl(SGA) PtNi/␥-Al2 O3 3.2 8.0 – – 215 3.3 0.19
PtCoAl(SGA) PtCo/␥-Al2 O3 2.8 – 7.2 – 220 3.4 0.19
PtFeAl (SGA) PtFe/␥-Al2 O3 2.9 – – 6.9 225 3.5 0.20
PtAl(SGA) Pt/␥-Al2 O3 2.6 – – – 270 2.9 0.19
NiAl(SGA) Ni/␥-Al2 O3 – 8.0 – – 250 3.4 0.21

temperature increases from 300 ◦ C to 1000 ◦ C; NiO species, NiO indicate that Co species were very small (2 nm is the limit of XRD
weakly interacting with ␥-Al2 O3 , non-stoichiometric and stoichio- sensitivity) and highly distributed.
metric Ni-spinels were reduced [26,35,36]. These Ni peaks are large The Ni-containing catalysts presented reflections at 36.4◦ , 44.4◦
and broad, indicating that these species were formed with different and 64.5◦ , which were attributed to NiAl2 O4 (JCPDS 01-1299)
sizes and various degrees of interaction with the support. phases. These components are formed due to the strong Ni O Al
The oxidized Ni species of PtNiAl(SGA) are reduced at lower tem- interaction caused by the sol–gel preparation in the acidic medium.
perature when compared to the ones of NiAl(SGA) . This is explained This high interaction caused a very good dispersion of Ni species
by the H2 spillover effect [15,26]. In fact, Pt combined with Ni pro- on these catalysts. As TPR-H2 results indicated, NiO bulk was also
motes the reduction of most of the Ni at lower temperatures. It formed, especially in the NiAl(SGA) catalyst. However, reflection
is also interesting to note that in the PtNiAl(SGA) catalyst the NiO lines related to NiO oxide were not detected, indicating that if these
reduction peak overlaps those of PtOx suggesting a possible interac- species are present, they are quite small (below 2 nm) or in very low
tion between these two components (PtOx –NiO). This may generate quantities. This interpretation is in agreement with other authors
a possible promoting effect by electron transfers between Pt–Ni who have studied similar catalyst systems [3,10,11,15,27].
metals after the reduction treatment [15,26]. To gather more information about the active phases dispersion
and the average size of Pt and Ni particles on the PtNiAl(SGA) cata-
3.1.3. XRD characterization of catalysts lyst, it was decided to analyze the reduced Pt, Ni and PtNi catalysts
XRD analyses were carried out to identify the crystalline phases by in situ X-ray diffraction (i.e., at a similar state to the one just
in the prepared catalysts after their calcination. The X-ray diffrac- before the activity tests were begun). These X-ray diffractograms
tograms are presented in Fig. 3. are presented in Fig. 4. The X-ray patterns reveal that the reduc-
The patterns are broad and their intensity is very weak, imply- tion treatment slightly increased the crystallinity of the alumina
ing that the phases are not highly crystalline. All catalysts recorded support of these catalysts, but their active phases remained well
low reflections at 37.2◦ , 46.2◦ and 66.8◦ , which can be assigned to dispersed. This may indicate that the surface area and porosity
␥-Al2 O3 phase (JCPDS 29-063). Reflections at 39.6◦ , 46.3◦ and 67.8◦ , of ␥-Al2 O3 were not significantly affected by the reduction treat-
corresponding to Pt/PtOx phase (JCPDS 4-802), were also registered ment. However, new diffraction lines were registered at 44.5◦ , 51.8◦
in the Pt-containing catalysts (except in PtFeAl(SGA) ). The Pt/PtOx and 76.6◦ . These lines can be associated to metallic Ni0 (JCPDS
peaks of PtCoAl(SGA) catalyst were more intense and thinner, indi- 04-850) [3,10,11,15,27]. It should be noted that the Ni0 peaks of
cating that they correspond to bigger particles (around 46 nm). PtNiAl(SGA) are more intense than those detected in NiAl(SGA) . This
These peaks were not detected in the PtFeAl(SGA) catalyst, probably indicates that Pt promoted Ni reduction, taking into account that
due to the corresponding phase being present at very low quanti- both catalysts have similar Ni loadings (8 wt%), as indicated by the
ties or in a more amorphous structure, since Fe species reflections ICP analysis. This observation is also corroborated by previously
were not detected either. The reflection lines corresponding no Co mentioned TPR-H2 results.
species were registered in the PtCoAl(SGA) catalyst, and this could

Fig. 4. X-ray patterns of in situ reduced Pt, Ni, PtNi-based ␥-Al2 O3 catalysts and their
corresponding average size particles at calcined and in situ reduced state obtained
Fig. 3. X-ray patterns of calcined Pt, Ni, PtNi, PtFe and PtCo-based ␥-Al2 O3 catalysts. by the Scherrer equation.
130 M. El Doukkali et al. / Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136

To estimate the effect of the reduction treatment on metal After the reduction treatment, BEs for Pt0 species were detected
particle sizes, the average sizes of Pt or/and Ni species were cal- at 314.4 for PtAl(SGA) and at 315.0 eV for PtNiAl(SGA) catalysts. In
culated from the XRD data for calcined and reduced catalysts. This the latter, apart from this BE energy, an additional BE was detected
calculation involved the Scherrer equation and the most intense at around 317.4 eV and this indicates that Pt was not completely
reflections at 39.6◦ for Pt0 /PtOx, 36.4◦ for NiAlx Oy and 44.5◦ for Ni0 . reduced. A comparison of the BE of the Pt4d5/2 core level for reduced
In fact, a careful deconvolution of overlapping peaks was used to PtNi and Pt catalysts showed that BE is slightly higher for PtNi than
carry out a good estimation of the particle sizes. These results are for Pt catalysts. Bakar et al. [42,43] attributed this BE increase to the
presented in Fig. 4. formation of metallic bands or alloying of the metals involved. Some
The average Pt particle size did not change for Pt-containing Ni atoms are probably interacting with Pt modifying its electronic
catalysts, indicating that the SGA effectively yields stable Pt parti- environment via electron transfer, which can led to a promoting
cles even at high reduction temperatures (750 ◦ C for PtNiAl(SGA) ). effect between both elements. This result is also supported by TPR-
However, for Ni-containing catalysts, Ni particle sizes increased, H2 results, where the overlapping PtOx NiO peaks indicate some
indicating that after the reduction treatment the Ni particles possible Pt Ni interaction, which can occur after the reduction of
suffered some sintering. This increase is lower for PtNi cata- PtNiAl(SGA) catalyst.
lyst, which suggests that Pt combined with Ni increases the The Ni2p3/2 core level for reduced Ni-containing catalysts
resistance of PtNi catalysts to this Ni particles sintering phe- records a peak at 852.6 and 852.2 eV, which is characteristic of
nomenon. This enhancement may be related to the spillover Ni0 [44–46]. Another component was detected at 855.1–855.4 eV,
effect, since this phenomenon led to Ni reduction at lower which can be related to the non-reduced nickel ions strongly
temperatures. interacting with alumina lattice oxygen [44–46]. This indicates
that nickel species are not completely reduced. However, as com-
mented previously, the amount of reduced Ni was higher in the
3.1.4. XPS analysis of Pt, Ni and PtNi catalysts PtNiAl(SGA) catalyst (60%) as compared to its monometallic NiAl(SGA)
In order to study the physicochemical change experienced by counterpart (37%). This is mainly due to H2 spillover favored by
different elements during the reduction treatment, the calcined well-dispersed Pt on this catalyst.
(Cal.) and in situ reduced (Red.) Pt, Ni and PtNi catalysts were A comparison between the Pt/Al and Ni/Al atomic ratios
analyzed by XPS technique. The binding energy (BE), chemical (reduced catalysts with their calcined counterparts) reveals that
state and the superficial amounts of Pt or/and Ni species, as well these ratios are lower in reduced catalysts than in their calcined
as Pt/Al and Ni/Al atomic ratios, are summarized in Table 2. The counterparts. This indicates that the reduction treatment at high
Lorentzian/Gaussian deconvolutions of Pt4d5/2 , and Ni2p3/2 core- temperature (750 ◦ C) partially sintered Pt and Ni particles. This
level spectra are also illustrated in Fig. 5. result is in agreement with the results obtained from XRD analy-
The results obtained from XPS analysis showed the existence sis, as the literature [41] also indicates. The decrease in the surface
of a band at 74.4 eV (±0.4 eV) for all calcined and reduced cata- (XPS) Ni/Al ratio after reduction could also be due to the incorpora-
lysts, which is characteristic of Al3+ belonging to ␥-Al2 O3 [37,38]. tion of metal species into the porous structure of the support, as the
The deconvolution of Pt4d5/2 core-level spectra for calcined Pt and values for pore volumes derived from N2 adsorption–desorption
PtNi catalysts showed the coexistence of at least two Pt species: isotherms indicated.
one corresponds to Pt metallic at 314.3 and 314.9 eV, respectively.
Another component that can be related to Pt oxide was detected at
317.8 and 317.4 eV, respectively [39]. The detection of PtOx species 3.1.5. TEM/EDX analysis of Pt, Ni and PtNi catalysts
may be associated to the capacity of ␥-Al2 O3 to stabilize some PtOx The reduced Pt, Ni and PtNi-based ␥-Al2 O3 catalysts were exam-
particles on its surface due to the Pt O Al interaction caused by ined by transmission electron microscopy coupled with an energy
the SGA preparation method, as TPR-H2 results corroborated. The dispersive X-ray detector (TEM/EDX) (Fig. 6).
BEs of the Ni2p3/2 core level recorded for the calcined Ni and PtNi Fig. 5(a–e) illustrates the textural morphology of reduced bulk
catalysts are located at 855.6 and 855.8 eV, respectively. This BE is ␥-Al2 O3 (a), Pt/␥-Al2 O3 (b), Ni/␥-Al2 O3 (c) and PtNi/␥-Al2 O3 (d)
typically assigned to Ni2+ species formed due to the strong Ni O Al catalysts and the corresponding EDX spectra (e) at three different
interaction forming non-stoichiometric NiAlx Oy and stoichiomet- points of PtNi/␥-Al2 O3 catalysts. These points were chosen to deter-
ric NiAl2 O4 , symbolized as NiT 2+ (pseudo tetrahedral coordination) mine the local chemical composition of particles dispersed on the
[40]. For NiAl(SGA) catalyst, another BE was detected at around surface of PtNi/␥-Al2 O3 catalysts.
853.6 eV, which can also be assigned to Ni2+ in the form of NiO, The image (a) shows that the ␥-Al2 O3 support of the catalysts
symbolized as NiO 2+ (pseudo octahedral coordination) [40]. prepared by the acid sol–gel method presents a laminated tex-
A comparison of the BEs of the Pt4d5/2 core level of calcined Pt- ture with high micro-porosity, which confirms that preparation
containing catalysts showed that the BE of Pt0 was slightly higher by this procedure leads to the formation of a longitudinal fiber
for PtNiAl(SGA) than for PtAl(SGA) catalyst (see Table 2). A similar morphology. This is consistent with the information derived from
observation was reported by Pawelec et al. [41], being attributed to adsorption/desorption results. The TEM images (b–d) also show
a possible electron transfer due to higher Pt O Al interactions. For that Pt or/and Ni catalysts have well-dispersed Pt or Ni species on
calcined Ni-containing catalysts, the PtNiAl(SGA) catalyst recorded a the ␥-Al2 O3 surface. These good dispersions of Pt and Ni particles
higher proportion (100%) of NiT 2+ than the NiAl(SGA) (79%) catalyst. make it difficult to obtain a good estimate of the particle size distri-
This higher population of NiT 2+ on the catalytic surface indicates bution, but it is clear from TEM images that the particles sizes are
that the Ni species are well dispersed and, as a result, Ni surface very small. It can also be appreciated that Pt and Ni clusters have a
exposure was also higher, since the XPS atomic ratio of Ni/Al in spherical form (uniformly distributed) with an approximate aver-
PtNi catalyst (0.424) is higher than in the Ni catalyst (0.325). This age diameter of 1.6 nm and 11.5 nm for Pt/␥-Al2 O3 and Ni/␥-Al2 O3 ,
confirms that the Ni in bimetallic catalysts ends up being better respectively. The image (d) of PtNi/␥-Al2 O3 shows that the average
dispersed on the ␥-Al2 O3 surface. In addition, the Pt/Al and Ni/Al diameters of Pt (1.4 nm) and Ni (7 nm) particles are similar to Pt and
atomic ratios estimated from XPS-derived data are higher than Ni particle diameters in the monometallic catalyst counterparts.
those calculated using bulk compositions from chemical analysis The average diameters of Ni particles are consistent with those esti-
(see Table 2). This means the catalyst outer surface is enriched in mated by XRD (Table 2), while those of the Pt particles are smaller.
Pt and Ni species. This may be due to the X-ray technique’s limitation for detecting
M. El Doukkali et al. / Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136 131

Table 2
BE and population of Pt or/and Ni species in different chemical states, as well as atomic ratios for calcined (Cal.) and in situ reduced (Red.) Pt, Ni and PtNi-based ␥-Al2 O3
catalysts.

Catalyst B.E. Pt4d5/2 (eV) Pt/Al B.E. Ni2p3/2 (eV) Ni/Al

Pt0 PtOx Ni0 NiO 2+ NiT 2+

PtAl(SGA)
Cal. 314.3 (20) 317.8 (80) 0.010a – – – –
0.007b
Red. 314.4 (65) 317.4 (35) 0.010 – – – –
NiAl(SGA)
Cal. – – – – 853.6 (21) 855.8 (79) 0.325a
0.071b
Red. – – – 852.6 (37) – 855.4 (63) 0.094
PtNiAl(SGA)
Cal. 314.9 (24) 317.4(76) 0.056a – – 855.7 (100) 0.424a
0.009b 0.069b
Red. 315.0(100) – 0.015 852.2 (60) – 855.1 (40) 0.103

( ) is population of different species.


a and b are atomic ratio calculated from XPS and ICP data, respectively.

particles with diameters below 2 nm. Besides that, it is interesting 3.2. Catalytic activity results
to note that Pt/Ni agglomerations formed by Pt clusters deposited
on Ni counterparts can be observed in the PtNi/␥-Al2 O3 catalyst (see As mentioned above, PtFe, PtCo and PtNi-based ␥-Al2 O3 cata-
image d). This observation was confirmed by EDX analysis (see EDX lysts were only tested in the glycerol APR process at 250 ◦ C and
spectra (e)), although there are not many of them. 50 bar. The parameters for catalytic activity results, such as glyc-
EDX analysis was carried out to check local chemical compo- erol conversion (GC), conversion into gas (CIG) and the distribution
sitions. The analysis of point P1 (single thin particles) showed yields of gaseous products (Y(i)) and oxygenated hydrocarbon com-
Pt and Al emission lines, indicating the existence of Pt over all pounds (Y OHCs(j)), are shown in Fig. 7 (data obtained after 32 h of
the ␥-alumina matrix, as the size of the EDX beam used was ca. reaction, steady state). For a better understanding of the role played
5 nm, while the Pt particles have a diameter of approximately by each phase in the PtNi bimetallic catalyst, the ones with the best
1.4 nm. By focusing the EDX beam on the slightly higher parti- results – Pt and Ni monometallic catalysts – were then tested at
cles (point P3 ), the analysis indicated that these are formed by 250 ◦ C/50 bar and 230 ◦ C/30 bar. Their catalytic behavior as a func-
Ni. However, the EDX analysis of point (P2 ) showed the existence tion of time-on-stream is summarized in Table 3 in terms of GC,
of Ni and Pt emission lines in some agglomerations. This sug- CIG, and the distribution yields of major OHCs in liquid stream. The
gests that these agglomerations were formed by particles with very gas flow formation and H2 selectivity obtained at steady state for
close Pt/Ni atomic vicinity, which can facilitate an electronic inter- Pt, Ni and PtNi-based ␥-Al2 O3 catalysts are also presented in Fig. 8.
action or transfer between both elements, generating a possible Operating at 250 ◦ C/50 bar, the catalysts showed increasing
promoting effect. This Pt present in close vicinity to Ni parti- activity, following the order PtFeAl(SGA) < PtCoAl(SGA)  PtNiAl(SGA) .
cles confirms similar experimental findings from TPR and XPS PtFe and PtCo showed quit low activity in glycerol reforming. The
analysis. PtNi catalyst recorded the highest glycerol conversions: GC (74%)

Fig. 5. XPS spectra deconvolution of Pt4d5/2 and Ni2p3/2 core levels for calcined (Cal.) and in situ reduced (Red.) Pt, Ni and PtNi-based ␥-Al2 O3 catalysts.
132 M. El Doukkali et al. / Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136

Fig. 6. (a–d) Textural morphology of reduced bulk ␥-Al2 O3 (a), Pt/␥-Al2 O3 (b), Ni/␥-Al2 O3 (c) and PtNi/␥-Al2 O3 (d) catalysts and the corresponding EDX spectra (e) of three
different points of PtNi/␥-Al2 O3 catalysts obtained from TEM/EDX analysis.

and CIG (34%). Glycerol conversion into gas when PtNiAl(SGA) cata- yields; for example, the methane yield was as high as 0.44 (see
lyst was used was nearly nine fold higher than the conversion rates Fig. 7b). Meanwhile, several OHCs were detected in the liquid phase,
obtained using PtFe or PtCo catalysts. This indicated that the PtNi such as C1 (methanol), C2 (acetaldehyde, acetic acid, ethylene gly-
catalyst was far more active and allowed higher reforming rates of col) and C3 (propanal, 1&2-propanol, acetol, 1,2-propanediol). The
glycerol and of liquid intermediate products (OHCs) into gaseous C3 compounds accounted for 70–95% of the total liquid products
products. The major gaseous products formed on the catalysts were formed. These compounds were typically formed by dehydration
H2 , CO2 , CH4 and smaller amounts of other alkanes (C2 C3 ). Their reactions (cleavage of C O band), indicating that the acid centers
relative yields varied depending on the catalyst, and the PtNi cat- of the alumina support were also involved in the reaction, and
alysts provided the best H2 yield (0.51), but also higher alkane not only the Pt/Ni metallic centers, where C C/C H bond cracking
M. El Doukkali et al. / Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136 133

Table 3
Glycerol conversion (GC) and conversion into gas (CIG), as well as yield (Y OHCs(j)) distribution of OHCs as a function of APR time-on-stream on Pt, Ni and PtNi-based ␥-Al2 O3
catalysts at 250 ◦ C-50 bar (a)/230 ◦ C-30 bar (b) (10 wt% C3 H8 O3 /H2 O mixture, feed rate of 0.4 mL/min, 1 g of catalyst, WHSV = 2.60 h−1 ).

Catalyst Time-on-stream (h) of aqueous phase reforming reaction

Parameters 2h 8h Night 25 h 32 h
(SGA)
PtAl Conversion a/b a/b a/b a/b
CG (%) 25.5/11.4 25.0/11.1 27.6/9.9 23.4/9.3
CIG (%) 7.9/4.8 6.8/3.7 8.1/3.4 6.2/3.5

Liq. stream Y OHCs(j) × 10−2


2-Propanol 3.0/1.1 3.0/1.0 3.0/1.0 3.0/1.0
1,2-PDO 4.1/1.0 5.2/1.0 6.0/2.1 7.1/2.0
Acetic acid 0.0/3.0 0.0/3.0 0.0/3.0 0.0/3.0
Acetol 1.1/1.0 2.1/1.0 2.2/1.0 2.2/1.0
EG 0.0/0.0 1.0/0.0 2.1/0.0 2.0/0.0
Methanol 0.0/0.0 1.0/0.0 1.0/0.0 1.0/0.0

NiAl(SGA) Conversion a/b a/b a/b a/b


CG (%) 68/57 49/40 47/29 42/27
CIG (%) 22/12 10/7.5 9/5 8/4.4

Liq. stream Y OHC(j) × 10−2


2-Propanol 8.0/5.0 5.0/4.0 4.0/2.0 4.0/2.0
1,2-PDO 22.0/20.0 17.0/15.0 14.07.0 14.0/8.0
Acetic acid 0.0/0.0 0.0/0.0 0.0/0.0 0.0/0.0
Acetol 5.0/5.0 5.5/5.0 6.0/4.0 6.0/4.0
EG 5.1/4.1 4.2/3.0 3.0/1.1 2.0/1.1
Methanol 1.1/1.0 1.0/1.0 0.0/0.0 0.0/0.0

PtNiAl(SGA) Conversion a/b a/b a/b a/b


CG (%) 97/70 87/64 78/56 74/55
CIG (%) 58/36 49/29 39/25 35/22

Liq. stream Y OHCs(j) × 10−2


2-Propanol 08.1/5.3 07.1/4.2 06.1/3.0 4.4/3.0
1,2-PDO 10.3/20.0 15.1/17.0 15.0/16.0 14.8/16.0
Acetic acid 6.1/00.0 2.7/0.0 0.0/0.0 0.0/0.0
Acetol 1.5/2.0 2.2/2.0 2.7/2.0 2.6/2.0
EG 4.7/13.0 10.3/12.0 10.5/14.0 10/15.0
Methanol 3.1/3.0 3.0/3.1 2.4/2.0 2.1/2.0

uses to occur leading to gaseous products and C1 C2 compound not completely reformed. On the one hand, series reactions arose
production. from the hydrogenation of adsorbed CO/CO2 species on catalyst
An analysis of Table 3 and Fig. 8 shows that at both operating surfaces to form alkanes, which negatively affect H2 generation.
conditions, PtNi bimetallic catalyst provided higher conversion and Alkane formation can take place through different hydrogen con-
hydrogen flow rates when compared to its Pt and Ni monometal- suming reactions. The hydrogen generated in the first step could
lic counterparts. Not only were the conversions (CG, CIG) obtained be converted into alkanes, CO2 and water by hydrogenation of the
using the PtNiAl(SGA) catalyst higher than the sum of the ones dehydrated reaction intermediates [10,11].
obtained by each PtAl(SGA) and NiAl(SGA) (Table 3), but also the CH4 can be formed via methanation from CO/CO2 (Ni sites are
gaseous product flow was higher than the corresponding sum of active in these reactions), leading to a loss of H2 rate formation.
the flows generated by both monometallic catalysts (see Fig. 8). Other alkanes could possibly be formed by Fischer–Tropsch type
This observation confirmed once again the important role of the reactions [10,11]. High pressures (50 bar) favored gaseous reactions
promoting effect between the two metals on the activity of PtNi that lower total mole number, such as the reactions generating
bimetallic catalyst. The indicated effect is more pronounced at CH4 , C2 H6 and C3 H8 [10,11]. On the other hand, parallel reactions
moderate temperature/pressure conditions (230 ◦ C/30 bar). More- proceeded through C O bond cracking followed by hydrogenation
over, it can be observed that Pt monometallic catalyst presented to give diols, essentially in the form of 1,2-propanediol (recording
lower conversions, but its catalytic behavior was relatively sta- higher yields for PtNi and Ni and presenting higher glycerol con-
ble throughout the reaction time, while the Ni-containing catalyst version and lower H2 selectivity – see Table 3 and Fig. 8) and
counterparts recorded higher glycerol initial conversions but lower monoalcohols (2-propanol and methanol), or by rearrangement
stability (Table 3). However, considering only the Ni-containing pathways to form organic acids in liquid phase. Therefore, one of the
catalysts, the data presented in Table 3 show that its cat- keys for aqueous phase reforming of glycerol is to choose a selective
alytic behavior is relatively more stable at 230 ◦ C/30 bar than at and robust catalyst able to avoid, as far as possible (i) the hydro-
250 ◦ C/50 bar. genation of CO/CO2 via the methanation or/and Fischer–Tropsch
Regarding H2 selectivity, the Pt catalyst promotes H2 -rich gas, reactions, and (ii) C O bond cleavage, although it must also be
while Ni is less selective. PtNi led to an intermediate behavior, very active for C H/C C bond cleavage and promote the removal
particularly at 230 ◦ C/30 bar. This high H2 selectivity presented by of adsorbed CO throughout the WGS reaction.
PtAl(SGA) catalyst is related to its lower glycerol conversion into
gas (≈6%) as opposed to 35% for PtNiAl(SGA) and to its negligible 4. Discussion
methanation activity. At higher glycerol conversions, H2 selectivity
usually fell due to the formation of several undesired by-products The acid sol–gel preparation method used in this work gener-
via series and parallel reactions [10,11] and these intermediates are ated catalysts with fibrous and laminated morphology, and high
134 M. El Doukkali et al. / Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136

80
74 (a)
GC CIG
60
GC/CIG (%)

34 39 35
40

20

4 4
0
PtCoAl(SGA) PtFeAl(SGA) PtNiAl(SGA)

0.6
H2 CH4
0.51 (b)
C2H6 C3H8
0.5
0.44 Fig. 8. Gas flow formation (olive) and H2 selectivity (red) obtained from the APR test
of Pt, Ni and PtNi-based ␥-Al2 O3 catalysts at 250 ◦ C/50 bar and 230 ◦ C/30 bar (10 wt%
0.4
C3 H8 O3 /H2 O mixture, feed rate of 0.4 mL/min, 1 g of catalyst, WHSV = 2.60 h−1 , data
Y (i)

at steady state). (For interpretation of the references to color in this figure legend,
0.3 the reader is referred to the web version of the article.)

0.2 that Pt helped both an easier reduction and a better dispersion


0.04 0.05 of Ni active sites as compared to the ones observed in NiAl(SGA)
0.1 0.01 0.00 0.01 catalyst. Moreover, the high surface area and micro/mesoporosity
0.01
0.00 achieved by the SGA method facilitate the adsorption of intermedi-
0.000 0.000 0.001
0.0 ate molecules on highly dispersed Pt and Ni sites, converting them
PtCoAl(SGA) PtFeAl(SGA) PtNiAl(SGA) into gaseous products. In addition, the close vicinity of Pt/Ni leads
to a promoting effect, which was suggested by TPR-H2 results and
0.4 confirmed by XPS and TEM/EDX analyses, and could also have a
C1 (methanol) (c) crucial effect in the best catalytic activity shown by PtNiAl(SGA) cat-
C2 (acetaldehyde + ethylene glycol) alyst. This better performance can also be related to the high activity
C3 (propanal + acetol + 1&2-propanol +1,2-propanediol)
0.3 of Pt and Ni in C H/C C bond cleavage [10–12]. According to the
Y_OHCs (j)

0.23 literature [10–12], oxygenated molecules are decomposed on Pt


or Ni by dehydrogenation, followed by C C cleavage to give H2
0.18 0.19
0.2 and adsorbed CO, which is readily converted to CO2 by the WGS
reaction.
The poor results recorded by PtCoAl(SGA) and PtFeAl(SGA) cat-
0.1 alysts could be related, on the one hand, to lower Pt dispersion
0.06 (bigger Pt particles, around 46 nm) for PtCo catalyst and also to
0.02 the lower temperature (350 ◦ C) used to reduce the PtFe catalyst
0.000.01 0.000.01
0.0 because it was not severe enough to reduce iron active sites.
According to Davda et al. [9], these metals record higher activity
PtCoAl(SGA) PtFeAl(SGA) PtNiAl(SGA)
for methanation and Fischer–Tropsch reactions and lower activ-
Fig. 7. (a–c) Catalytic activity of PtFe, PtCo and PtNi-based ␥-Al2 O3 catalysts at ity for C C/C H bond cleavage. Thus, catalyst formulations with
250 ◦ C/50 bar; conversions of glycerol (a), distribution yields of gaseous products these metals give lower activities for glycerol APR than PtNiAl(SGA)
(b) and liquid products (c) at steady state (10 wt% C3 H8 O3 /H2 O mixture, feed rate of catalysts.
0.4 mL/min, 1 g of catalyst, WHSV = 2.60 h−1 and 30–32 h of APR).
The low glycerol conversions obtained using the Pt monometal-
lic catalyst may be due to its smaller metal loading (2.6 wt%) than
surface area and micro/mesoporosity. This method allows obtain- Ni or Pt/Ni loadings 8 wt% for NiAl(SGA) and 11.2 wt% for PtNiAl(SGA)
ing Pt and Ni particles with close vicinity and highly dispersed catalysts, approximately. Its pronounced catalytic stability could
leading to an increase in catalytic activity. be due to high Pt resistance to oxidation, which is a characteristic
The higher catalytic activity of PtNi compared to bimetallic (PtFe of noble metals (Pt active centers remained in their reduced state
and PtCo) and monometallic (Ni and Pt) catalysts can be explained because of the H2 generated throughout the APR reaction). This sta-
by its better catalytic properties compared to the other catalysts: bility could also be related to a possible oxidation/reduction cycle
higher Ni dispersion (Ni0 active phase formed from finer NiAlx Oy of Pt, since the XPS analysis showed that Pt was not completely
spinels after reduction treatment), as indicated by the results reduced just after the initial reduction treatment, and before APR
obtained from the analysis of TPR-H2 and XRD (higher dispersion) started. This part of Pt might be reduced by the H2 generated by the
and XPS (higher metallic surface exposure). A high proportion of reactions taking place and replace the oxidized Pt on the surface,
these thin particles were reduced at relatively lower temperatures maintaining the catalytic activity of the catalyst relatively stable. On
due to the H2 /Pt-spillover phenomenon. This effect avoids their sin- the other hand, the gradual decrease in catalytic activity shown by
tering, and as a result a higher number of Ni active sites on the Ni-containing catalysts along time-on-stream, as well as the quan-
catalyst were available. XRD also confirmed that Pt and Ni (see titative and qualitative changes of the OHCs yields distribution until
Fig. 4) were present on the catalyst surface as thin particles, and the overall reaction process reached its steady state might be an
M. El Doukkali et al. / Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136 135

indication of acid and Ni active centers being modified. The deacti- 5. Conclusions
vation of Ni species was also observed by Wen et al. [4], who found
that Ni atoms are highly susceptible to oxidation by water excess The following catalytic systems: PtFe, PtCo, PtNi, Ni or Pt/␥-
under APR thermodynamic conditions [26]. In addition, the overall Al2 O3 prepared by an acid sol–gel procedure were used for glycerol
carbon balance of these catalysts recorded a difference of 8–16% aqueous phase reforming to produce H2 at low temperatures and
between the carbon in the inlet and outlet streams. It is quiet prob- moderate pressures. The main conclusions of this study are:
able that part of this carbon ended as light carbonaceous molecules PtNi bimetallic systems are clearly more active than
deposited on the catalyst surface, which may have led to catalyst monometallic (Pt or Ni) or PtFe and PtCo systems for APR reactions,
deactivation. even if all of them were prepared using an acid sol–gel procedure
The values of H2 selectivity registered (Fig. 8) indicate that the and similar metal loadings were attained.
catalysts generated higher H2 yields at 230 ◦ C/30 bar. This sug- Based on reported findings [26] and on the results of this study,
gests that operating at relatively moderate temperatures/pressures it was confirmed that the properties of the catalysts studied were
is much more selective to H2 formation by the APR of glyc- directly related to the preparation method. The physicochemical
erol. This can be related to the effect of pressure, since lower properties of PtNi catalysts prepared by the acid sol–gel method
pressures implies more favorable thermodynamic conditions were greatly improved as compared to similar catalysts prepared by
for reactions increasing the total mole number in gas phase conventional impregnation [1,2,4–7,9–13,15,22,24]. The prepara-
(C3 H8 O3 + 3H2 O → 7H2 + 3CO2 ). Nevertheless, as it can be seen tion under acid conditions promoted hydrolysis reactions resulting
in Table 3 and Fig. 8 the catalysts presented higher glycerol in the formation of longitudinal fibrous materials, characterized
conversions at the highest reaction temperature and its cor- by high surface area and micro/mesoporosity, which considerably
responding total pressure (250 ◦ C/50 bar) and produced higher favored the dispersion of Pt and Ni particles. The preparation under
gas flow rates as compared to the ones obtained operating at basic conditions studied previously [26] increased the condensa-
230 ◦ C/30 bar, which is to be expected as the APR is an endothermic tion reactions and lead to cluster materials, characterized by a
process. spherical shape and mesoporosity.
As it has been showed, a comparison among Pt/Ni catalysts pre- The TPR-H2 , XRD and XPS studies of PtNi/␥-Al2 O3 catalysts sug-
pared by acid sol–gel (SGA) and those prepared by basic sol–gel gest that Pt presence improved Ni reducibility by H2 spillover, and
(SGB) and incipient wetness impregnation (IWI), presented previ- that part of the Pt and part of the Ni ended up interacting to gen-
ously in [26], revealed that catalytic properties varied depending erate a promoting effect between both metals, as suggested by
on the preparation method. The textural characteristics of the TPR-H2 and XPS and confirmed by TEM/EDX analysis. This effect
SGA-prepared ␥-alumina supported Pt/Ni catalysts are consid- was clearly manifested in catalytic activity. All these factors explain
erably different as their superficial area is approximately 50% why PtNiAl(SGA) catalytic systems had interesting and promising
higher than the ones of equivalent IWI-prepared catalysts and properties for glycerol APR. This catalyst showed better catalytic
50% lower than the ones measured for similar SGB catalysts [26]. activity than similar catalysts prepared using conventional proce-
The Pt/Ni SGA catalysts show micro/mesoporosity, while the SGB dures, being especially active for gaseous product formation from
catalysts are clearly mesoporous materials. In fact, both sol–gel glycerol reforming under liquid-phase conditions. This system is
methods lead to Ni spinel-type catalysts ensuring a very good dis- also a good catalyst for WGS reaction, as negligible amounts of
persion of Pt/Ni metallic particles, which is not achieved when CO were detected. In fact, the PtNiAl(SGA) catalyst prepared by acid
these materials are prepared by IWI method. The SGA method sol–gel presented promising results, and further work on this sys-
ensures a better compromise between platinum–nickel interac- tem could allow the development of more active and selective
tions at lower temperatures, as TPR-H2 indicated, but the SGB catalysts for APR processes.
catalysts are more resistant to the sintering effect after reduction
treatment at high temperatures, as corroborated by XRD and XPS Acknowledgements
analysis.
The results obtained using PtNiAl(SGA) catalyst prepared by acid This work was carried out in the laboratory of the “Sus-
sol–gel are much better than those recorded by a similar PtNi-based tainable Process Engineering” research group (University of the
␥-Al2 O3 catalyst prepared by incipient impregnation [15,26], but Basque Country (UPV/EHU), Spain) in collaboration with the “UCCS”
they present slightly lower catalytic activities than those obtained research unit (University of Science and Technology of Lille, France),
using the PtNiAl(SGB) catalyst reported in [26]. This can be directly supported by funds from the Spanish Ministry of Science and
related to the catalytic properties developed by each preparation Innovation (Project: ENE2011-23950) and the Basque Government.
method. However, the acid sol–gel method may still be an interest- The authors gratefully acknowledge both institutions for providing
ing alternative for preparing promoted catalysts for glycerol APR. financial support, as well as express their gratitude to L. Burylo for
Comparing the results reported in this paper to those presented in XRD analysis, A. Addad for Microscopy analysis and M. Trentesaux
several previous investigations [1,2,4–7,9–13,15,22,24], they also and A. Beaurain for XPS analysis. The European Regional Develop-
represent a promising and interesting advance in the scientific ment Fund (ERDF), CNRS, Région Nord Pas-de-Calais and the French
knowledge, especially if they are jointly considered with previ- Ministry for Higher Education and Research are acknowledged for
ous research published in [26]. As examples, operating at similar funding the XPS/LEIS/ToF-SIMS spectrometers within the Regional
operating conditions (225 ◦ C/40 bar and 240 ◦ C/40 bar), A. Iriondo Platform of Surface Analysis.
et al. [15] obtained gas flow rates of 2 and 6.5 mL(STP) min−1 g−1 cat
using an impregnated PtNi/␥-Al2 O3 catalyst (steady state (27–30 h) References
and WHSV = 2 h−1 ), which are lower than those obtained in this
work (13 and 19 mL(STP) min−1 g−1 cat ) testing the PtNiAl
(SGA) cata- [1] D.O. Özgür, B.Z. Uysal, Biomass Bioenergy 35 (2011) 822–826.
[2] Y. Guo, X. Liu, M.U. Azmat, W. Xu, J. Ren, Y. Wang, G. Lu, Int. J. Hydrogen Energy
lyst at a temperature/pressure of 230 bar/30 ◦ C and 250 ◦ C/50 bar, 37 (2012) 227–234.
respectively (steady state (30–32 h) and WHSV = 2.6 h−1 ). This is a [3] A. Iriondo, V.L. Barrio, J.F. Cambra, P.L. Arias, M.B. Güemez, R.M. Navarro, M.C.
significant improvement in the development of more active and Sanchez-Sanchez, J.L.G Fierro, Top. Catal. 49 (2008) 46–58.
[4] G. Wen, Y. Xu, H. Ma, Z. Xu, Z. Tian, Int. J. Hydrogen Energy 33 (2008) 6657–6666.
stable catalytic systems for H2 production throughout the APR pro-
[5] R.L. Manfro, A.F. Da Costa, N.F.P. Ribeiro, M.M.V.M. Souza, Fuel Process. Technol.
cess of glycerol. 92 (2011) 3330–3335.
136 M. El Doukkali et al. / Journal of Molecular Catalysis A: Chemical 368–369 (2013) 125–136

[6] N. Luo, X. Fu, F. Cao, T. Xiao, P.P. Edwards, Fuel 86 (2008) 3483–3489. [28] M. El Doukkali, A. Iriondo, P.L. Arias, J.F. Cambra, I. Gandarias, V.L. Barrio, Int. J.
[7] G.W. Huber, S. Iborra, A. Corma, Chem. Rev. 9 (2006) 4044–4098. Hydrogen Energy 37 (2012) 8298–8309.
[8] S. Adhikari, S.D. Fernando, A. Haryanto, Energy Convers. Manage. 50 (2009) [29] S. Panyad, S. Jongpatiwut, T. Sreethawong, T. Rirksomboon, S. Osuwan, Catal.
2600–2604. Today 174 (2011) 59–64.
[9] R.R. Davda, J.W. Shabaker, G.W. Huber, R.D. Cortright, J.A. Dumesic, Appl. Catal. [30] A. Sirijaruphan, J.G. Goodwin Jr., R.W. Rice, D. Wei, K.R. Butcher, G.W. Roberts,
B: Environ. 56 (2005) 171–186. J.J. Spivey, Appl. Catal. A: Gen. 281 (2005) 11–18.
[10] R.D. Cortright, R.R. Davda, J.A. Dumesic, Nature 418 (2002) 964–967. [31] V. Hacker, R. Fankhauser, G. Faleschini, H. Fuchs, K. Friedrich, J. Power Sources
[11] G.W. Huber, J.W. Shabaker, J.A. Dumesic, Science 300 (2003) 2075–2077. 86 (2000) 531–535.
[12] J.W. Shabaker, R.R. Davda, G.W. Huber, R.D. Cortright, J.A. Dumesic, J. Catal. 215 [32] D. Schanke, S. Vada, E.A. Blekkan, A.M. Hilmen, A. Hoff, A. Holman, J. Catal. 156
(2003) 344–352. (1995) 85–95.
[13] J.W. Shabaker, R.R. Davda, G.W. Huber, R.D. Cortright, J.A. Dumesic, Catal. Lett. [33] F.B. Noronha, et al., Catal. Today 28 (1996) 147–157.
88 (2003) 1–8. [34] T. Huizinga, J.V. Grondelle, R. Prins, Appl. Catal. 10 (1984) 199–213.
[14] I.O. Cruz, N.F.P. Ribeiro, D.A.G. Aranda, M.M.V.M. Souza, Catal. Commun. 9 [35] N. Ichikuni, D. Murata, S. Shimazu, T. Uematsu, Catal. Lett. 69 (2000) 33–36.
(2008) 2606–2611. [36] J.M. Rynkowski, T. Paryjczak, M. Lenik, Appl. Catal. A: Gen. 106 (1993) 73–82.
[15] A. Iriondo, J.F. Cambra, V.L. Barrio, M.B. Güemez, P.L. Arias, M.C. Sanchez- [37] M.C. Sánchez-Sánchez, R.M. Navarro, J.L.G. Fierro, Int. J. Hydrogen Energy 32
Sanchez, R.M. Navarro, J.L.G. Fierro, Appl. Catal. B: Environ. 106 (2011) 83–93. (2007) 1462–1471.
[16] A. Wawrzetz, B. Peng, A. Hrabar, J.A. Lercher, J. Catal. 269 (2010) 411–420. [38] J.C.S Araujo, D. Zanchet, R. Rinaldi, U. Schuchardt, C.E. Hori, J.L.G. Fierro, J.M.C.
[17] Z. Wang, Q. Liu, J. Yu, T. Wu, G. Wang, Appl. Catal. A: Gen. 239 (2003) 87–94. Bueno, Appl. Catal. B: Environ. 84 (2008) 552–562.
[18] I. Gandarias, P.L. Arias, J. Requies, M. El Doukkali, M.B. Güemez, J. Catal. 282 [39] D. Briggs, M.P. Seah (Eds.), Practical Surface Analysis by Auger and X-Ray Photo-
(2011) 237–247. electron Spectroscopy, vol. 1, 2nd ed., Wiley, Chichester, 1990.
[19] I.H. Cho, S.B. Park, S.J. Cho, R. Ryoo, J. Catal. 173 (1998) 295–303. [40] J. Jun, M. Dhayal, H.J. Shin, Y.H. Han, N. Getoff, Appl. Surf. Sci. 254 (2008)
[20] L. Hu, K.A. Boateng, J.M. Hill, J. Mol. Catal. A: Chem. 259 (2006) 51–60. 4557–4564.
[21] A. Vazquez, T. Lopez, R. Gomez, X. Bokhimi, J. Mol. Catal. A: Chem. 167 (2001) [41] B. Pawelec, S. Damyanova, K. Arishtirova, J.L.G. Fierro, L. Petrov, Appl. Catal. A:
91–99. Gen. 323 (2007) 188–201.
[22] A. Tanksale, C.H. Zhou, J.N. Beltramini, G.Q. Lu, J. Incl. Phenom. Macrocycl. Chem. [42] N.H.H. Abu Bakar, M.M. Bettahar, M. Abu Bakar, S. Monteverdi, J. Ismail, M.
65 (2009) 83–88. Alnot, J. Catal. 265 (2009) 63–71.
[23] J.A. Schwarz, C. Contescu, A. Contescu, Chem. Rev. 95 (1995) 477–510. [43] N.H.H. Abu Bakar, M.M. Bettahar, M. Abu Bakar, S. Monteverdi, J. Ismail, M.
[24] K. Lehnert, P. Claus, Catal. Commun. 9 (2008) 2543–2546. Alnot, J. Mol. Catal. A: Chem. 308 (2009) 87–95.
[25] G.W. Huber, J.A. Dumesic, Catal. Today 111 (2006) 119–132. [44] M. García-Diéguez, I.S. Pieta, M.C. Herrera, M.A. Larrubia, L.J. Alemany, J. Catal.
[26] M. El Doukkali, A. Iriondo, P.L. Arias, J. Requies1, I. Gandarias, L. Jalowiecki- 270 (2010) 136–145.
Duhamel, F. Dumeignil, Appl. Catal. B: Environ. 125 (2012) 516–529. [45] M. García-Diéguez, I.S. Pieta, M.C. Herrera, M.A. Larrubia, L.J. Alemany, Appl.
[27] A. Iriondo, V.L. Barrio, M. El Doukkali, J.F. Cambra, M.B. Güemez, J. Requies, P.L. Catal. A: Gen. 377 (2010) 191–199.
Arias, M.C. Sanchez-Sanchez, R.M. Navarro, J.L.G Fierro, Int. J. Hydrogen Energy [46] M. García-Diéguez, E. Finocchio, M.A. Larrubia, L.J. Alemany, Guido Busca, J.
37 (2012) 2028–2036. Catal. 274 (2010) 11–20.

You might also like