You are on page 1of 10

Applied Catalysis B: Environmental 43 (2003) 345–354

Production of hydrogen for fuel cells by steam reforming of


ethanol over supported noble metal catalysts
Dimitris K. Liguras, Dimitris I. Kondarides, Xenophon E. Verykios∗
Department of Chemical Engineering, University of Patras, GR-26504 Patras, Greece
Received 16 September 2002; received in revised form 13 October 2002; accepted 13 December 2002

Abstract
The catalytic performance of supported noble metal catalysts for the steam reforming (SR) of ethanol has been investigated
in the temperature range of 600–850 ◦ C with respect to the nature of the active metallic phase (Rh, Ru, Pt, Pd), the nature of
the support (Al2 O3 , MgO, TiO2 ) and the metal loading (0–5 wt.%). It is found that for low-loaded catalysts, Rh is significantly
more active and selective toward hydrogen formation compared to Ru, Pt and Pd, which show a similar behavior. The catalytic
performance of Rh and, particularly, Ru is significantly improved with increasing metal loading, leading to higher ethanol
conversions and hydrogen selectivities at given reaction temperatures. The catalytic activity and selectivity of high-loaded
Ru catalysts is comparable to that of Rh and, therefore, ruthenium was further investigated as a less costly alternative. It was
found that, under certain reaction conditions, the 5% Ru/Al2 O3 catalyst is able to completely convert ethanol with selectivities
toward hydrogen above 95%, the only byproduct being methane. Long-term tests conducted under severe conditions showed
that the catalyst is acceptably stable and could be a good candidate for the production of hydrogen by steam reforming of
ethanol for fuel cell applications.
© 2003 Elsevier Science B.V. All rights reserved.
Keywords: Steam reforming; Hydrogen production; Ethanol; Rhodium; Ruthenium

1. Introduction fuel processors, i.e. devices able to convert liquid or


gaseous fuels into hydrogen. Among the various feed-
The use of hydrogen for fuel cell applications stocks that have been proposed for such applications,
represents one of the most environmentally sound ethanol is very attractive because of its relatively high
methods for the production of electrical energy and is hydrogen content, availability, non-toxicity and stor-
expected to gain wide usage in the near future for both age and handling safety. More importantly, in contrast
automotive and small-to-medium scale stationary ap- to fossil fuels, such as methanol and gasoline, which
plications. Until technical and economic issues related have been proposed for the same application, ethanol
to hydrogen storage, transportation and distribution can be produced renewably from several biomass
are solved, the generation of hydrogen for fuel cells sources, including energy plants, waste materials from
is expected to be accomplished on-site with the use of agroindustries or forestry residue materials, organic
fraction of municipal solid waste, etc. [1]. The use of
∗ Corresponding author. Tel.: +30-2610-991527; biomass-derived ethanol for the production of hydro-
fax: +30-2610-991527. gen and energy has a great potential for significant
E-mail address: verykios@chemeng.upatras.gr (X.E. Verykios). contributions towards short-to-medium-term targets

0926-3373/03/$ – see front matter © 2003 Elsevier Science B.V. All rights reserved.
doi:10.1016/S0926-3373(02)00327-2
346 D.K. Liguras et al. / Applied Catalysis B: Environmental 43 (2003) 345–354

related to energy supply, environmental protection (Sorptomatic 1900) instrument. Prior to each measure-
and regional development. ment, the sample was outgased under dynamic vac-
Ethanol may be converted to a hydrogen-rich gas uum at 250 ◦ C for 5 h. The ␥-Al2 O3 and MgO supports
stream by steam reforming (SR), a reaction which has (Alfa), with specific surface areas of 83 and 21 m2 /g,
been shown to be entirely feasible thermodynamically respectively, were used as received, while TiO2 (De-
[2–4]. The challenge is, therefore, to develop cata- gussa P25) was calcined in air at 700 ◦ C prior to im-
lysts exhibiting high stability and activity for ethanol pregnation. This resulted in complete transformation
steam reforming with high yields of hydrogen and re- to its rutile form and a decrease of its specific surface
sistance to coke formation. Interest in the SR reac- area from 50 to 8 m2 /g.
tion of ethanol has grown recently, and a few studies Hydrogen adsorption isotherms were determined by
have appeared in the literature dealing with the perfor- a volumetric technique employing a modified Fisons
mance of a variety of catalysts, mainly Ni, Co and Cu Instruments (Sorptomatic 1900) apparatus. Measure-
[1,5–14] and supported noble metals, including Rh, ments were performed at 25 ◦ C, in the pressure range
Pd, and Pt [12–15]. of 0–75 Torr. Prior to each measurement, the catalyst
In our previous work [1,7] we studied the steam sample (ca. 1.0 g) was pretreated by (a) dynamic vac-
reforming of ethanol over supported Ni catalysts and uum at 250 ◦ C for 1 h; (b) reduction with 1 bar of H2
investigated the effect of the support material and at 250 ◦ C for 1 h; (c) evacuation for 30 min at 250 ◦ C;
operating conditions, such as reaction temperature, and (d) cooling down to room temperature under vac-
steam-to-ethanol ratio and space velocity, on catalytic uum. The hydrogen uptakes at monolayer coverage
activity, selectivity and stability. It was shown that were obtained for each catalyst by extrapolating the
under selected experimental conditions, the Ni/La2 O3 adsorption isotherms to zero pressure. Dispersion of
catalyst is very active and stable for the steam reform- the metallic component, expressed in terms of H/M
ing of ethanol and is characterized by high selectiv- ratio, was obtained assuming that each metal atom
ity toward hydrogen production [1,7]. In the present chemisorbs one hydrogen atom. Results obtained are
work, the same reaction is examined over noble metal summarized in Table 1.
catalysts supported on a variety of oxides. The effect
of the nature and loading of the metal and the effect 2.2. Apparatus and procedures
of the support material on catalytic performance is in-
vestigated. Catalytic performance tests have been carried out
using an apparatus which has been described in

2. Experimental
Table 1
2.1. Catalyst preparation and characterization Results of hydrogen chemisorption measurements
Catalyst Metal Hydrogen Dispersion Mean
Catalysts were prepared by impregnation of the sup- loading uptake (H/M ratio) crystallite
port (␥-Al2 O3 , MgO, TiO2 ) with an aqueous solution (wt.%) (cm3 /g) size (nm)
of the corresponding metal precursor salt (Rh(NO3 )3 , Pt/Al2 O3 1.0 0.562 0.98 1.1
Ru(NO)(NO3 )3 , (NH3 )2 Pt(NO2 )2 , (NH3 )2 Pd(NO2 )2 ). Pd/Al2 O3 1.0 0.418 0.39 2.5
The slurry was then heated slowly at 70 ◦ C under con-
Rh/Al2 O3 0.5 0.337 0.62 1.8
tinuous stirring and maintained at that temperature un- 1.0 0.491 0.45 2.4
til nearly all the water evaporated. The solid residue 2.0 1.087 0.50 2.2
was dried at 110 ◦ C for 24 h and then reduced at 300 ◦ C Ru/Al2 O3 1.0 0.158 0.14 6.8
in H2 flow for 2 h. The metal loading of the catalysts 3.0 0.739 0.22 4.4
thus prepared varied between 0 and 5 wt.%. 5.0 1.190 0.21 4.7
Specific surface areas of the carriers were measured Ru/TiO2 5.0 0.382 0.07 14.5
by the BET technique employing nitrogen physisorp-
Ru/MgO 5.0 0.120 0.02 46.4
tion at the temperature of liquid nitrogen with a Fisons
D.K. Liguras et al. / Applied Catalysis B: Environmental 43 (2003) 345–354 347

detail elsewhere [1]. It consists of a flow system, the 2 h under a hydrogen flow of 50 cm3 /min. After re-
reactor unit and the analysis system. The flow sys- duction, the reactor is heated to the desired reaction
tem is equipped with a set of mass-flow controllers temperature under He flow. When the system is equi-
(MKS) and an HPLC pump (Marathon) for feeding librated at the desired temperature, the water–ethanol
liquid reagents (ethanol–H2 O mixtures). The liquid is gas stream is introduced to the reactor and the con-
pumped into a vaporizer where it is evaporated and version of reactants and selectivities toward reaction
heated to 150 ◦ C and where it can be mixed (when de- products are determined using the analysis system
sired) with the gas stream coming from the mass-flow described above. All experiments were performed at
controllers. The resulting gas mixture is then fed to atmospheric pressure.
the reactor through stainless steel tubing maintained
at 150 ◦ C by means of heating tapes. The catalyst is
placed in a quartz microreactor and kept in place by 3. Results and discussion
means of quartz-wool. Temperatures are measured
along the reactor and the catalyst bed by means of a 3.1. Effect of reaction temperature on ethanol
K-type thermocouple placed within a quartz capillary conversion and product distribution
well running through the bed.
The analysis system consists of two gas chro- Typical results showing the influence of temperature
matographs (Shimadzu). The first one is equipped on catalytic activity and product distribution of ethanol
with two packed columns (Porapak-Q and Carbosieve) steam reforming over 1% Rh/Al2 O3 catalyst are pre-
and two detectors (TCD, FID) and operates with He sented in Fig. 1, where ethanol conversion (XEtOH )
as the carrier gas. Porapak-Q is used for the separa- and selectivities to reaction products (Si ) are plotted
tion of C2 H5 OH, H2 O, CH3 CHO, CH4 , C2 H4 and as functions of reaction temperature. It is observed
C2 H6 , while Carbosieve is used for the separation that conversion of ethanol is significant at tempera-
of CO, CO2 and CH4 . Since the presence of large tures higher than 600 ◦ C and increases continuously
amounts of water hinders analysis in the Carbosieve with increasing temperature, achieving complete con-
column, a condenser is placed before its inlet in order version at 800 ◦ C. In the entire range of temperatures
to condense and remove H2 O from the gas stream.
The second GC is equipped with a Carbosieve col-
umn, connected to the exit of the condenser, and a
TCD detector. This chromatograph uses N2 as the
carrier gas and is used solely to determine the H2
concentration in the reformate.
Determination of the response factors of the TCD
and FID detectors has been achieved with the use of
gas streams of known composition (Scott specialty gas
mixtures, self-prepared EtOH/H2 O/CH3 CHO mix-
tures, etc.). Since several substances (CO, CO2 , O2 ,
H2 ) can only be analyzed after the condenser, an inter-
nal standard is used to account for the volume change.
Reaction gases are supplied from high-pressure gas
cylinders (Air Liquid) and are of ultrahigh purity.
Analytic grade ethanol was obtained from Merck.
Catalytic activity is evaluated in terms of ethanol
conversion. Selectivities are defined as the ratios of Fig. 1. Effect of reaction temperature on the conversion of ethanol
(XEtOH ) and on the selectivities (Si ) toward products obtained over
the product moles to the consumed moles of ethanol,
the 1% Rh/Al2 O3 catalyst. Also shown the temperature drop in
accounting for stoichiometry. the catalyst bed (T) due to the SR reaction endothermicity. Ex-
In a typical experiment, 100 mg of fresh catalyst is perimental conditions: mass of catalyst: 100 mg; particle diameter:
placed in the reactor and reduced in situ at 750 ◦ C for 0.25 < dp < 0.50 mm; H2 O:EtOH = 3:1; flow rate: 340 cm3 /min.
348 D.K. Liguras et al. / Applied Catalysis B: Environmental 43 (2003) 345–354

examined, ethanol is converted with high selectivities tures, catalyst activity is too low to promote these re-
towards H2 , CO and CO2 . These are the products of actions and acetaldehyde and ethylene appear in large
both steam reforming and water gas shift reactions that concentrations in the reformate. Of course, their con-
take place simultaneously in the reactor. Significantly centration in the product mixture is also a function of
smaller quantities of byproducts, such as CH3 CHO, contact time as we will discuss below.
C2 H4 and CH4 are also observed, with their concentra- Production of methane is maximized at around
tions exhibiting a strong dependence to reaction tem- 750 ◦ C, where SH2 , SCO , and SCO2 go through weak
perature. minima, indicating that CH4 is formed by the hy-
Since the SR reaction of ethanol is highly endother- drogenation of carbon oxides. Increasing temperature
mic (H = +57 kcal/mol), a significant temperature above 750 ◦ C results in a decrease of SCH4 , due to
drop is expected along the reactor. The temperatures its reformation with H2 O and/or CO2 which becomes
shown in Fig. 1 and in subsequent figures correspond thermodynamically feasible at these temperatures.
to the temperature at the inlet of the catalyst bed. It is of interest to note that at 800 ◦ C, where com-
Fig. 1 also shows the temperature drop (T) in the plete conversion of ethanol is achieved, selectivity to-
catalyst bed, which can be related to the ethanol wards hydrogen is higher than 95% (Fig. 1) and the
conversion curve. As expected, the temperature drop only undesired byproduct is methane. Methane is un-
increases with increasing ethanol conversion. Al- desirable in this process since it competes with H2 for
though the system is not perfectly adiabatic, and the hydrogen atoms. It can, if so desired, be completely
two curves diverge at higher temperatures (probably eliminated by reformation by increasing reaction tem-
due to heat input from the hot furnace), the observed perature or contact time. From a practical point of
endotherms are in reasonably good agreement with view, however, small amounts of methane in the refor-
the ethanol conversions and can be used as a first mate may be tolerated since CH4 present in the fuel
approximation of the extend of reaction. Similar tem- cell off-gases can be burned, along with the unutilized
perature profiles were obtained over the other catalyst hydrogen, to provide the heat necessary for the en-
beds, with small variations due to different reaction dothermic reforming reaction.
rates, and, therefore, heat consumption rates. Under the experimental conditions employed, ther-
Selectivity toward ethylene, formed by dehydration modynamic equilibrium calculations indicate that the
of ethanol over the acidic sites of the support, is rela- only products are H2 , CO, CO2 and CH4 [3]. The se-
tively small and goes through a maximum of ca. 2.5% lectivity towards CH4 decreases rapidly above 500 ◦ C
at 700–750 ◦ C (Fig. 1). Since ethylene is a precursor of and becomes negligible above 800 ◦ C and this is ac-
coke formation, and may lead to catalyst deactivation, companied by a corresponding increase in selectivity
its presence is highly undesirable. Ethylene produc- towards H2 which approaches 100%. The presence of
tion, however, decreases as the temperature increases significant amounts of methane and other byproducts
above 750 ◦ C and drops to zero at 800 ◦ C. This is the in the reformate indicates that the system is far from
result of high rates of steam reforming of ethylene. equilibrium. Table 2 compares the calculated and ex-
The present catalyst exhibits high activity for this re- perimental selectivities. As the temperature increases
action, even when ethylene is present in high concen- from 735 to 800 ◦ C, the system approaches, but never
trations [14]. reaches, the equilibrium values. The same holds for
Selectivity toward acetaldehyde, which is formed the CO/CO2 equilibrium. It should be noted that none
by dehydrogenation of ethanol, follows a qualitatively of the cases examined in this study were equilibrium
similar behavior and is also completely reformed at limited.
high temperatures (Fig. 1). It can be noted here that
previous experiments [1] have shown the selectivi- 3.2. Influence of the chemical nature of the active
ties of both ethylene and acetaldehyde to rise rapidly metallic phase
at lower temperatures (below 600 ◦ C) and very low
ethanol conversions (less than 30%). Apparently, these The comparative catalytic performance of the sup-
are primary reaction products, which undergo further port material and of various noble metals such as Pd,
reactions at higher temperatures. At lower tempera- Ru, Pt and Rh for ethanol steam reforming was also
D.K. Liguras et al. / Applied Catalysis B: Environmental 43 (2003) 345–354 349

Table 2
Approach to thermodynamic equilibrium at various temperatures
800 ◦ C 760 ◦ C 735 ◦ C

Calculated Experimental Calculated Experimental Calculated Experimental

SH 2 99.61 95.74 98.95 87.61 97.99 84.79


SCO2 26.27 23.28 28.52 17.40 30.17 14.31
SCO 73.29 70.36 70.28 64.16 67.54 62.98
SCO2 /SCO 0.36 0.33 0.41 0.27 0.45 0.23
SCH4 0.44 6.09 1.20 11.32 2.29 9.38

investigated. The catalysts were prepared by disper- provides again the best conversions in the entire tem-
sion of the noble metal on the same support, ␥-Al2 O3 , perature range. Rh is also able to reform ethylene with
with the same metal loading, 1 wt.%. The results ease while Pt and Ru show some activity. Pd appears
of these experiments are summarized in Fig. 2A–C, to be inactive for ethylene reformation. None of the
where conversion of ethanol and product selectivities metals promotes the reformation of methane to a great
are plotted as functions of the reaction temperature. It extent under the conditions examined. Only Rh ex-
is observed that Rh is by far more active for the reac- hibits good activity but at temperatures above 780 ◦ C.
tion compared to the other catalysts examined; the or-
der of ethanol conversion being Rh  Pt > Pd > Ru 3.3. Influence of metal loading
(Fig. 2A). The low activity of 1% Ru may be partly
attributed to the poor dispersion of the present sample The effect of metal loading on catalytic performance
(14%) compared to that of 1% Pt (98%), 1% Pd (39%) was studied over ␥-Al2 O3 -supported Rh and Ru cat-
and 1% Rh (45%) (Table 1) since, as will be shown alysts. Results obtained from Rh/Al2 O3 samples of
below, ruthenium catalysts with higher metal loadings variable metal content are shown in Fig. 3. It is ob-
exhibit a significantly improved performance for the served that increasing Rh loading from 0.5 to 2.0 wt.%
title reaction. It is worth noting that the conversions of results in a shift of the ethanol conversion curve to-
Pt, Pd and especially of Ru, are not much higher than ward lower temperatures (Fig. 3A), while the tempera-
the conversions obtained on ␥-Al2 O3 alone, which, in ture at which complete ethanol conversion is achieved
turn, are very similar to the conversions obtained with is lowered by more than 50 ◦ C. Selectivities toward
a bed of quartz chips (i.e. homogeneous reactions) as H2 and CO2 are also substantially increased with in-
reported earlier [1]. In addition, Rh is significantly creasing Rh loading, while SCO is not significantly in-
more selective toward H2 and CO formation, and both fluenced (Fig. 3B). More importantly, the production
SH2 and SCO follow the order Rh  Pt > Ru = Pd, of undesirable reaction byproducts, i.e. acetaldehyde
as illustrated in Fig. 2B. Again, the selectivities of Ru and, especially, ethylene is significantly suppressed
and Pd towards these products are not much higher with increasing metal loading (Fig. 3C). In particu-
than those obtained on the support alone, especially lar, the maxima of both SCH3 CHO and SC2 H4 decrease
at higher temperature. The low selectivities of Pd, Ru drastically and shift toward lower temperatures with
and Pt toward hydrogen can be attributed to the rela- increasing metal loading. As a result, the concentra-
tively poor, compared to Rh, reforming capabilities of tions of acetaldehyde and ethylene in the reformate
these metals in the temperature range examined. On gas drop to practically zero over the 2% Rh/Al2 O3
the other hand, all metals show higher selectivity to- catalyst at temperatures close to 800 ◦ C.
wards CO2 indicating some ability to promote the wa- It is of interest to note that the dispersion of
ter gas shift reaction. Pt and Rh appear to be the most rhodium on the fresh Rh/Al2 O3 catalysts does not
active under these conditions. As shown in Fig. 2C, significantly change with increasing metal loading
these catalysts are able to reform acetaldehyde but from 0.5 to 2.0 wt.%, and takes values between 45
at temperatures higher than 800 ◦ C. The Rh catalyst and 62% (Table 1). This indicates that the observed
350 D.K. Liguras et al. / Applied Catalysis B: Environmental 43 (2003) 345–354

Fig. 2. Effect of reaction temperature on the conversion of ethanol Fig. 3. Effect of reaction temperature on the conversion of ethanol
(A), and on the selectivities toward reaction products (B and C) (A), and on the selectivities toward reaction products (B and
obtained over the 1% Me/␥-Al2 O3 catalysts (Me = Pd, Ru, Pt, C) obtained over Rh/␥-Al2 O3 catalysts of variable metal content
Rh) and the ␥-Al2 O3 support. Experimental conditions are same (0.5–2 wt.%). Experimental conditions are same as in Fig. 1.
as in Fig. 1.
D.K. Liguras et al. / Applied Catalysis B: Environmental 43 (2003) 345–354 351

differences in catalytic activity and selectivity (Fig. 3)


should not be attributed to differences of rhodium
mean crystallite size, which varies between 1.8 and
2.4 nm, but to the increasing number of active sites
with increasing metal loading.
Considering the fact that Rh is a rare and very
expensive metal, Ru was examined as a less costly
alternative for the SR reaction of ethanol. Results
obtained over Ru/Al2 O3 catalysts of variable metal
content are shown in Fig. 4. It is observed that in-
creasing Ru loading from 1 to 3 wt.% results in a
significant increase of activity, which is reflected in
a shift of the ethanol conversion curve toward lower
temperatures by ca. 100 ◦ C (Fig. 4A). Conversion of
ethanol is complete at 825 ◦ C over the 3% Ru catalyst,
while it reaches only 60% over the 1% Ru sample at
the same temperature. This is accompanied by a sub-
stantial increase of selectivity toward the reformation
products (Fig. 4B). For example, at 825 ◦ C, where
XEtOH = 100% over the 3% Ru catalyst, SH2 , SCO ,
and SCO2 assume values of ca. 96, 70 and 25%, re-
spectively, compared to only 60, 30 and 8% obtained
with the 1% Ru catalyst. This happens at the expense
of the byproducts (acetaldehyde, methane and ethy-
lene), which are only present in trace amounts in the
reformate at 825 ◦ C (Fig. 4C). The 3% Ru catalyst is
sufficiently active to reform these compounds at these
temperatures.
Further increasing the Ru loading to 5 wt.% results
in a small but appreciable improvement of catalytic
performance at temperatures above 750 ◦ C (Fig. 4A).
Conversion of ethanol reaches 100% at ca. 780 ◦ C,
which is about 50 ◦ C lower than the temperature re-
quired with the 3% Ru catalyst. Moreover, selectivity
toward H2 becomes higher than 95% at T > 780 ◦ C
(Fig. 4B), the only byproduct being methane (Fig. 4C).
Under these conditions, the catalytic performance of
Ru for the SR reaction of ethanol is comparable to
that of Rh (compare with Fig. 3). Since ruthenium
costs significantly less than rhodium, the former may
be considered as a good candidate for practical appli-
cations of ethanol steam reforming.

3.4. Influence of the support


Fig. 4. Effect of reaction temperature on the conversion of ethanol
As discussed above, production of ethylene, which (A), and on the selectivities toward reaction products (B and
generally is favored over acidic supports such as C) obtained over Ru/␥-Al2 O3 catalysts of variable metal content
␥-Al2 O3 , is undesirable, since it may lead to coke (1–5%). Experimental conditions are same as in Fig. 1.
352 D.K. Liguras et al. / Applied Catalysis B: Environmental 43 (2003) 345–354

formation and catalyst deactivation. This may be


avoided by neutralization of the acidic sites of Al2 O3
(e.g. by addition of potassium [8]) or by using basic
supports, such as La2 O3 [1,7] or MgO [12]. The effect
of the supporting material on the catalytic perfor-
mance has been examined over high-loaded (5 wt.%)
Ru catalysts dispersed on ␥-Al2 O3 , MgO and TiO2 .
The results obtained are summarized in Fig. 5. It is
observed that the activity of Ru is higher when dis-
persed on Al2 O3 than on MgO or TiO2 (Fig. 5A).
The Ru/Al2 O3 catalyst also exhibits higher selectiv-
ity toward reformation products (Fig. 5B) and results
in the formation of significantly lower amounts of
byproducts at any given temperature (Fig. 5C). Since
the dispersion of Ru on TiO2 and MgO is poor (7 and
2%, respectively) compared to Ru on Al2 O3 (21%)
(Table 1), the observed differences of catalytic per-
formance should not be solely attributed to the effect
of the support, but also to the number of exposed Ru
atoms.
Interestingly, selectivity toward C2 H4 over the
Al2 O3 -supported catalyst is similar to that of Ru/MgO
at temperatures up to 750 ◦ C and much lower at higher
temperatures (Fig. 5C). This may be attributed to the
high reformation activity of the 5% Ru catalysts. As
shown by Cavallaro et al. [12], the highly active Rh
catalysts act independently of the support (Al2 O3
or MgO) as far as coke formation is concerned. In
contrast, the less active Co catalyst is more selective
and stable when supported on MgO than on Al2 O3
[12]. In any case, the problem of carbon formation
requires additional study and will be discussed in a
future communication.

3.5. Effect of space time on catalytic activity and


product distribution

Although all the experiments described here were


performed with catalysts in powder form, industrial
processes typically require catalysts deposited on
structural supports, like pellets or monoliths, in order
to minimize pressure drop in the reactor. To examine
possible mass transfer limitations as one goes from
powders to pellets, we prepared a 5% Ru catalyst
supported on 1/16 in. Al2 O3 pellets (Engelhard) and Fig. 5. Effect of reaction temperature on the conversion of ethanol
investigated the effect of contact time on the catalytic (A), and on the selectivities toward reaction products (B and C)
activity and selectivity. Results obtained at 800 ◦ C obtained over 5% Ru catalysts dispersed on the indicated supports.
(catalyst bed inlet temperature) are shown in Fig. 6, Experimental conditions are same as in Fig. 1.
D.K. Liguras et al. / Applied Catalysis B: Environmental 43 (2003) 345–354 353

Fig. 7. Long-term stability test of the 5% Ru/Al2 O3 catalyst: alter-


ation of the conversion of ethylene and selectivities toward prod-
Fig. 6. Effect of space time on the conversion of ethanol and on ucts with time-on-stream. Experimental conditions: Tin = 700 ◦ C;
the selectivities toward reaction products obtained over the 5% mass of catalyst: 100 mg; particle diameter: 0.25 < dp < 0.50 mm;
Ru/Al2 O3 catalyst (1/16 in. pellets). Experimental conditions: mass H2 O:EtOH = 2:1; flow rate: 150 cm3 /min.
of catalyst: 300 mg; H2 O:EtOH = 3:1; Tin = 800 ◦ C.

where XEtOH and Si are plotted as functions of W/F. are required to completely reform 1 mole of ethanol
Mass transfer limitations are immediately apparent. to CO2 and H2 . Lower ratios increase production of
While the powdered catalyst achieves complete con- intermediates and byproducts. Additionally, higher
version and very high selectivity towards hydrogen at ratios tend to minimize carbon deposition rates and
contact times less than 0.02 g s/cm3 , the palletized cat- this is the main reason they are employed in industrial
alyst requires contact times higher than ca. 0.1 g s/cm3 steam reforming applications. Secondly, the reactor
to achieve similar performance. As contact time de- inlet temperature was limited to 700 ◦ C, while the
creases, conversion of ethanol progressively drops, outlet temperature was considerably lower due to the
accompanied by a decrease of selectivity toward the reaction endotherm. As mentioned previously, lower
reformation products (CO and H2 ) and by an increase temperatures may result in larger quantities of coke
of selectivity toward CO2 and byproducts such as formation. Finally, the system was started up and shut
CH3 CHO, and C2 H4 . At very low contact times (low down a number of times as shown in Fig. 7. This may
ethanol conversion) appreciable amounts of ethylene explain the small inconsistencies in the data. Under
are observed. Ethylene is a product of ethanol dehy- these severe experimental conditions employed, a
dration over the acidic Al2 O3 sites and the activity of small deactivation of the catalyst is observed during
the catalyst is not sufficient to reform it completely at the first 50 h on-stream, when conversion of ethanol
low contact times. drops from ∼65 to ∼50%. Further exposure of the
catalyst to the reaction mixture does not result in sig-
3.6. Long-term stability test of the 5% Ru/Al2 O3 nificant deactivation and the catalyst seams to stabilize
catalyst after 80 h on-stream. Interestingly enough, selectivity
towards hydrogen is not altered with time-on-stream
Very promising stability results were obtained and remains constant at ca. 90%. SCH4 , SCH3 CHO and
with the 5% Ru/Al2 O3 catalyst (Fig. 7) in an ex- SC2 H4 remain essentially constant and at low values
periment run under severe reforming conditions for throughout the run. It should be noted that catalyst
100 h. Firstly, the water-to-ethanol molar ratio was stability is strongly influenced by the experimen-
limited to two. Stoiciometricaly, 3 moles of water tal conditions employed and may be significantly
354 D.K. Liguras et al. / Applied Catalysis B: Environmental 43 (2003) 345–354

enhanced by increasing the water-to-ethanol ratio References


and/or reaction temperature [1].
[1] A.N. Fatsikostas, D.I. Kondarides, X.E. Verykios, Catal.
Today 75 (2002) 145.
4. Conclusions [2] E.Y. Garcia, M.A. Laborde, Int. J. Hydrogen Energy 16 (1991)
307.
[3] K. Vasudeva, N. Mitra, P. Umasankar, S.C. Dhingra, Int. J.
Supported Rh catalysts are significantly more ac- Hydrogen Energy 21 (1996) 13.
tive and selective for ethanol steam reforming for the [4] I. Fishtik, A. Alexander, R. Datta, D. Geana, Int. J. Hydrogen
production of a hydrogen-rich process gas, compared Energy 25 (2000) 31.
to Pt, Ru and Pd catalysts of similar metal content. [5] F. Haga, T. Nakajima, H. Miya, S. Mishima, Catal. Lett. 48
(1997) 223.
The catalytic performance is significantly improved [6] S. Cavallaro, S. Freni, Int. J. Hydrogen Energy 21 (1996)
with increasing metal loading. This is particularly true 465.
for the Ru catalysts where the 5% Ru/Al2 O3 sample [7] A.N. Fatsikostas, D.I. Kondarides, X.E. Verykios, Chem.
shows high activity and almost 100% selectivity to- Commun. 9 (2001) 851.
wards hydrogen at temperatures around 800 ◦ C and [8] F.J. Mariño, E.G. Cerrella, S. Duhalde, M. Jobbagy, M.A.
Laborde, Int. J. Hydrogen Energy 23 (1998) 1095.
space velocities of practical interest. Long-term tests [9] F.J. Mariño, M. Boveri, G. Baronetti, M. Laborde, Int. J.
conducted under severe conditions showed that the Hydrogen Energy 26 (2001) 665.
catalyst is acceptably stable and could be used for the [10] V.V. Galvita, G.L. Semin, V.D. Belyaev, V.A. Semikolenov,
production of hydrogen for fuel cell applications. P. Tsiakaras, V.A. Sobyanin, Appl. Catal. A 220 (2001)
123.
[11] V. Klouz, V. Fierro, P. Denton, H. Katz, J.P. Lisse, S.
Bouvot-Mauduit, C. Mirodatos, J. Power Sources 105 (2002)
Acknowledgements 26.
[12] S. Cavallaro, N. Mondello, S. Freni, J. Power Sources 102
This work was funded in part by the Commis- (2001) 198.
sion of the European Community, under contract [13] F. Auprêtre, C. Descorme, D. Duprez, Catal. Commun. 3
(2002) 263.
ERK6-CT-1999-00012. The assistance of Mr. I. Sion- [14] J.P. Breen, R. Burch, H.M. Coleman, Appl. Catal. B 39 (2002)
akides in carrying out the experiments is greatly 65.
appreciated. [15] S. Freni, J. Power Sources 94 (2001) 14.

You might also like