You are on page 1of 10

international journal of hydrogen energy 33 (2008) 4387–4396

Available at www.sciencedirect.com

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

Catalytic steam reforming of acetic acid in a fluidized


bed reactor with oxygen addition

J.A. Medrano, M. Oliva, J. Ruiz, L. Garcia*, J. Arauzo


Thermochemical Processes Group (GPT), Aragon Institute for Engineering Research (I3A), University of Zaragoza,
Marı́a de Luna 3, 50018 Zaragoza, Spain

article info abstract

Article history: Catalytic steam reforming of bio-oil is a promising process for producing hydrogen in
Received 31 October 2007 a sustainable environmentally friendly way that can improve the utilization of local
Received in revised form resources (natural sources or wastes). However, there remain drawbacks such as coke
11 April 2008 formation that produce operational problems and deactivation of the catalysts. Coprecipi-
Accepted 3 May 2008 tated Ni/Al catalysts are here used in a fluidized bed for reforming at 650  C of acetic acid as
Available online 12 August 2008 a model compound of bio-oil–aqueous fraction. Different strategies are applied in order to
study their effects on the catalytic steam reforming process: modification of the catalyst by
Keywords: increasing the calcination temperature or adding promoters such as calcium. The addition
Steam reforming of small quantities of oxygen is also tested resulting in an optimum percentage to achieve
Acetic acid a high carbon conversion process with less coke and without a hydrogen yield penalty
Pyrolysis liquids production. The results for catalytic steam reforming are compared with other ones
Oxygen addition from literature.
Fluidized bed ª 2008 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights
Nickel catalyst reserved.

1. Introduction energy carriers of the future [1], is a feasible and renewable


alternative for sustainable development.
Interest in the hydrogen economy is increasing because of its There are several methods for obtaining hydrogen from the
chemical uses and its utilization as a clean fuel in high ener- conversion of biomass classified in two main areas: biological
getic efficiency systems such as fuel cells in stationary, mobile or thermochemical. The latter includes steam reforming, one
or portable applications that can be used in vehicles. of the most suitable ways to produce a hydrogen rich gas.
Currently, hydrogen is produced through non-environ- Substantial research has been done on steam reforming
mentally friendly ways, such as steam reforming of natural from biomass or wastes, including bioethanol [2,3], trap grease
gas and petroleum derived compounds or coal gasification, [4], vegetable oils [5], glycerol [6] or bio-oil [7–9].
that deplete fossil fuels, producing corresponding CO2 emis- Steam reforming of pyrolysis liquids (bio-oil) is a promising
sions. However, hydrogen from biomass can replace fossil path for achieving hydrogen production. Lignocellulosic
fuels, reducing greenhouse gases and other contaminants, biomass is thermochemically processed (fast pyrolysis) to
contributing to a better utilization of available local natural obtain bio-oil. As bio-oil suffers from ageing during storage
resources, diversifying energy sources and reducing external [10,11] and is a complex mixture of a large number of
dependence. Waste-to-hydrogen processes are also attracting compounds, including aldehydes, alcohols, ketones, and acids,
increasing interest. Hydrogen, one of the most promising as well as more complex carbohydrate- and lignin-derived

* Corresponding author. Tel.: þ34 976 762194; fax: þ34 976 761879.
E-mail address: luciag@unizar.es (L. Garcia).
0360-3199/$ – see front matter ª 2008 International Association for Hydrogen Energy. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijhydene.2008.05.023
4388 international journal of hydrogen energy 33 (2008) 4387–4396

oligomeric materials emulsified with water, an aqueous frac- metal dispersion as the metal loading increased. Bimbela
tion is separated in order to achieve a higher stability of pyrol- et al. [22] worked with coprecipitated Ni/Al catalysts with
ysis liquids. The aqueous phase contains some organic different nickel contents in HAc steam reforming (23, 28 and
compounds such as acids, alcohols, aldehydes, ketones and 33% atomic ratio Ni/(Ni þ Al)) and reported that the highest
sugars. The ligninic fraction, insoluble in water, can be used hydrogen yield was obtained using 28% nickel content. In
to produce highly valued fine chemicals [12], such as phenolic this work this has been the nickel content selected.
resins that improve the economic viability of the process. Steam reforming catalysts are often doped with promoters.
Research into steam reforming of model compounds of Arauzo et al. [29] researched magnesium addition to a Ni/Al
bio-oil [13–22] is needed in order to develop efficient and stable catalyst that resulted in an improvement in catalyst strength
catalysts to perform this process. Generally, acetic acid (HAc), but gave a lower gas yield. Bangala et al. [28] reported that
which is soluble in water, is chosen as a model compound chromium inhibited the formation of nickel carbide, La2O3
because it has a significant presence and is one of the most led to a decrease in coke formation, MgO improved the robust-
representative constituents of the water soluble fraction of ness and TiO2 addition decreased the conversion and yield of
bio-oil [18]. gas. Galdámez et al. [18] concluded that the addition of
Some disadvantages have been found when steam reform- lanthanum in the nickel catalyst diminishes the hydrogen
ing bio-oil and other liquids derived from biomass in fixed-bed yield. Garcia et al. [7] explained that magnesium and
reactors due to problems caused by the formation of carbona- lanthanum can enhance steam adsorption that facilitates
ceous deposits limiting the operating time and requiring the gasification of surface carbon; on the other hand cobalt
a long regeneration process. A fixed bed is inadequate for pro- and chromium can modify the metal sites forming alloys
cessing the non-stable organic liquids derived from biomass with nickel and possibly reducing the crystallite size.
[23]. Therefore, it is necessary to use a fluidized bed because Another promoter is calcium. Some nickel catalysts
of one of its main advantages: carbon deposits on catalyst are employed in the partial oxidation of methane that incorpo-
better exposed to contact with steam and are consequently rates this element have demonstrated a good performance
more easily removed allowing more continuous operation. against coke formation. Morioka et al. [30] prepared Ni/Al
Another operational problem is that the bio-oil is not totally catalysts with calcium. The spc-Ni/Ca–Al catalyst (prepared
vaporized when feeding, so significant amounts of residual by solid phase crystallization) showed the highest activity
solids block the feeding line and the reactor [8]. Thus, a spray and selectivity to gas production as well as the highest
system and a cooling jacket are necessary to feed bio-oil or its sustainability against coke formation due to the stable and
fractions instead of a simple injection system. highly dispersed Ni metal particles and the basic properties
The review of Sutton et al. [24] presents various types of of the support. Lu et al. [31] reported the design of a modified
catalysts that have been employed in the gasification of Ni/Al catalyst introducing a CaAl2O4 spinel compound that
biomass. Nickel catalysts, both research and commercial, stabilized tiny nickel crystallites improving the resistance to
play a significant role. Noble metals [15,17,20] and nickel cata- coke. This work shows a first attempt to incorporate calcium
lysts [7–9,13,18–20,22] have been used in research into steam in coprecipitated Ni/Al catalysts in order to know the influ-
reforming reactions of oxygenates derived from biomass. In ence on steam reforming of acetic acid.
this paper we have selected nickel catalysts because they A common problem in the steam reforming of hydrocar-
are effective, commercially available and relatively cheap. bons reported by many authors is coke formation, which is
In order to develop the appropriate nickel catalyst for a serious operational problem and produces catalyst deactiva-
a process, some variables must be considered such as the tion. Three types of coke can be formed during steam reform-
preparation method and conditions (calcination and activa- ing: a highly reactive one constituted probably of atomic
tion), metal content and the presence of promoters. Ni/Al carbon, that is mostly gasified [32]; an amorphous type from
coprecipitated catalysts have shown a good performance in the polymerisation of the former, that is less active and may
fluidized bed in biomass steam gasification [25] and in steam be dissolved in the nickel crystallite leading to whisker forma-
reforming of model compounds of bio-oil, HAc [18] and acetol tion; and graphitic carbon. Coke can be eliminated by the use
[26], and therefore they have been selected as the catalysts in of promoters or, as some authors have reported, by oxygen
this work. The calcination temperature of Ni/Al coprecipitated addition [33]. One strategy that can be used is the incorpora-
catalysts significantly influences final properties of the cata- tion of small quantities of oxygen that can help to gasify
lyst such as, inter alia, surface area and crystalline phases, coke precursors on the catalyst surface. The incorporation of
and this variable must be considered. In previous work, the oxygen in steam reforming reactions also has a beneficial
influence of calcination temperature showed a significant effect on the energy balance decreasing the heat supply [34].
effect on catalytic biomass pyrolysis [27]. Therefore the effect Therefore, small quantities of oxygen were added to steam
of this variable on steam reforming of acetic acid is studied in reforming of acetic acid as a novelty of the research work
this paper. that is being carried out.
Metal loading is also an important parameter in catalyst This work studies the influence of the addition of small
activity. Bangala et al. [28] tested impregnated nickel–alumina quantities of oxygen to the catalytic steam reforming of HAc
catalysts with different nickel contents (5, 10, 15 and 20 wt%), as well as the influence of the calcination temperature and
and found that the conversion and gas yield increased with the addition of calcium as a promoter on the catalyst activity
metal loading up to 15%. On the other hand the coke yield and stability. The experiments were performed at 650  C in
also increased continuously up to 20% as it is related with a fluidized bed reactor using coprecipitated Ni/Al catalysts
metal loading. These results were attributed to the lower with a nickel content of 28% atomic ratio Ni/(Ni þ Al).
international journal of hydrogen energy 33 (2008) 4387–4396 4389

mixture of sand and the nickel catalyst that did not show
2. Experimental any segregation problems during the experiments. The
particle size of sand and catalyst used was 160–320 mm and
2.1. Experimental system the particle density of these solids is similar. The bed was
of 7 cm height with a nickel catalyst weight content of
The experimental system is shown in Fig. 1. This is a bench- 1.1 g. The inlet liquid (HAc aqueous solution) flow rate was
scale installation with a tubular 1 inch (2.54 cm) quartz fluid- 0.68–0.77 ml/min. The catalyst weight/HAc mass flow ratio
ized bed reactor. The reactor is externally heated by an electric (W/mHAc) was around 6 g catalyst min/g HAc, corresponding
furnace. The feeding system consists of a quartz coaxial injec- to GC1HSV values of around 6800 h1. GC1HSV was defined
tion nozzle producing a spray and is refrigerated by an as the volume of C1-equivalent species in the feed at stan-
external cooling jacket. The aqueous solution of HAc is intro- dard temperature and pressure (STP) per unit volume of
duced through the inner tube and is sprayed into the fluidized catalyst (including the void fraction) per hour. Oxygen
bed by a secondary nitrogen flow coaxial to the inner tube. addition experiments were carried out ranging from 0 to
This reactor was designed in order to process bio-oil or its 8% oxygen (percentage of oxygen necessary for the stoi-
fractions. As mentioned above, bio-oil cannot be totally vapor- chiometric combustion of HAc). In these experiments air
ized and a significant amount of residual solids can be formed was used to add the oxygen.
because of thermal decomposition blocking the feeding line, The experimental procedure involved in situ reduction of
so a spray system with external cooling is needed. the catalyst prior to the steam reforming reaction, with
The aqueous solution of HAc is delivered by a peristaltic hydrogen diluted in nitrogen (10%, v/v) at 650  C for 1 h.
pump. The product gas goes out through a recovering cooling
system (two special condensers with an ice bath) where
excess steam and non-reacted organic liquids are trapped by 2.2. Catalysts
condensation. A cotton filter is also placed after the
condensers to retain small solid particles of catalyst or carbo- Two research catalysts were prepared by coprecipitation. One
naceous residues. contains nickel and aluminium and the other also contains
The CO and CO2 concentrations of the exit gas were contin- calcium. The preparation method was similar to that
uously measured by an infrared (IR) analyzer. An Agilent P200 described by Al-Ubaid and Wolf [35]. The nickel content of
Micro GC gas chromatograph equipped with thermal conduc- these catalysts is 28% atomic ratio Ni/(Ni þ Al).
tivity (TC) detectors was used to determine concentrations of One of the catalysts (Ni/Al) was synthesised by adding
the gas products: CO2, CO, CH4, C2 (C2H4, C2H6, C2H2), N2 and H2. ammonium hydroxide (NH4OH) to a solution of Ni(NO3)2$6H2O
The experimental system was operated at atmospheric and Al(NO3)3$9H2O in distilled water until the pH attained
pressure. All the experiments were performed using a value of 7.9. The precipitation medium was maintained at
a 23 wt% HAc aqueous solution, which corresponds to an S/C 40  C and moderately stirred. The precipitate obtained was
molar ratio of 5.58. The reactor operated as a fluidized bed at filtered and washed at 40  C and dried overnight at 105  C.
u/umf ¼ 10. The precursors thus obtained were ground and sieved to
The catalytic steam reforming experiments were per- a particle size ranging from 160 to 320 mm and calcined at
formed at 650  C. The reactor bed contained a well mixed low heating rate in air atmosphere up to the final calcination

Fig. 1 – Schematic of the experimental system.


4390 international journal of hydrogen energy 33 (2008) 4387–4396

temperature for 3 h. Final calcination temperatures of 750, 850 trends were observed previously for the coprecipitated Ni/Al
and 900  C were employed. catalyst with 33% atomic ratio Ni/(Ni þ Al) [27]. The compar-
The other catalyst (Ni/Al/Ca), modified with calcium, was ison of the XRD patterns of Ni/Al and Ni/Al/Ca catalysts
synthesised with the following molar ratios: Ca/Al ¼ 0.50, (both calcined at 750  C) shows that the NiO phase is clearly
Ca/Ni ¼ 1.29. It was prepared by adding ammonium hydroxide present in the calcium modified catalyst in contrast to the
(NH4OH) to a solution of Ni(NO3)2$6H2O, Ca(NO3)2$4H2O and Ni/Al catalyst where two phases are found (NiO and NiAl2O4).
Al(NO3)3$9H2O in distilled water until the pH attained a valued Surface areas of the different modified catalysts were
of 7.9. The precipitation medium was maintained at 40  C and measured in a Micromeritics ASAP 2020 by nitrogen adsorp-
moderately stirred. The colloid was not filtered because Ca2þ tion. The values of surface area determined were 175 m2/g
could not be precipitated at the pH value and was thus dried for the catalyst calcined at 750  C, 112 m2/g for the catalyst
overnight at 70  C to evaporate the water. The precursor calcined at 850  C and 88 m2/g at 900  C of calcination temper-
obtained was ground and sieved to a particle size ranging from ature. When increasing the calcination temperature
160 to 320 mm and calcined in air atmosphere at low heating a decrease in surface area is obtained because of the catalyst
rate up to a temperature of 750  C that was maintained for 3 h. support sintering. A 50% of surface area is lost by increasing
In order to know various properties of the catalysts, the the calcination temperature from 750 to 900  C.
calcined catalysts were characterized by X-ray diffraction
(XRD) and nitrogen adsorption. Several catalyst samples 2.3. Chemicals
were also characterized by XRD and scanning electronic
microscopy (SEM) after their use in reforming. The model compound selected for the tests (HAc) was
Fig. 2 shows the results obtained for the XRD analyses (D- supplied by PANREAC (99.5% purity). Other chemicals used
Max Rigaku) of the calcined catalysts. Two different phases included commercial high purity gases (>99.999%): hydrogen,
are detected: NiO and NiAl2O4 (spinel compound) for the Ni/ nitrogen, air, helium and argon, as well as standard gas
Al catalyst calcined at different temperatures. As the calcina- mixtures of CO, CO2 and nitrogen for calibration of the CO-
tion temperature increases a higher proportion of spinel CO2 analyzer and H2, N2, CO, CO2, CH4, and C2 for the calibra-
phase is observed. Higher crystallinity is also obtained when tion of the gas chromatograph.
increasing the calcination temperature of the catalyst. Similar

3. Results and discussion

3.1. Influence of the calcination temperature


of the catalyst

The activity and stability of the catalyst can be modified by its


calcination temperature, so steam reforming experiments
were carried out to study this influence. Table 1 shows the
experimental conditions and the overall results obtained for
the experiments using the Ni/Al catalyst at different calcina-
tion temperatures (750, 850 and 900  C). The table shows the
values of some experimental variables such as the liquid
feeding rate, W/mHAc ratio and GC1HSV. Some results are also
shown such as the carbon conversion (percentage of carbon
contained in the HAc that is converted to gases: CO, CO2, CH4
and C2), yields to different gases (expressed as g gas/g HAc),
and the gas composition (% mol, N2 and H2O free).
The results indicate that the recovery for all the experi-
ments is 1.00  0.02. The gas composition (% mol, N2 and
H2O free) shows that H2 content is around 65%, CO content
around 6%, CO2 content around 27% and CH4 content up to
0.22%. Fig. 3a–d shows in more detail the influence of the calci-
nation temperature of the catalyst on gas yields and carbon
conversion. The hydrogen yield (Fig. 3a) decreases when
increasing the calcination temperature of the catalyst. This
effect is more noticeable when the catalyst is calcined at
900  C. Fig. 3b shows the same behaviour for the CO2 yield as
for H2. In contrast, an increase in the CO yield (Fig. 3c) is
observed when the calcination temperature increases. These
figures allow the study of the evolution of different gas yields
Fig. 2 – XRD patterns of Ni/Al and Ni/Al/Ca fresh catalysts. with time. A low decrease in the H2 and CO2 yields and a very
(a) Ni/Al calcined at 750 8C, (b) Ni/Al calcined at 850 8C, (c) Ni/ slight decrease in the CO yield are observed with time for the
Al calcined at 900 8C and (d) Ni/Al/Ca calcined at 750 8C. catalyst calcined at 750 and 850  C. These tendencies could be
international journal of hydrogen energy 33 (2008) 4387–4396 4391

Table 1 – Experimental results at 650 8C, with a catalyst


weight of 1.1 g and 2 h reaction time with modified
catalysts
Run # 1 2 3 4

Catalyst Ni/Al Ni/Al Ni/Al Ni/Al/Ca


Calcination temperature 750 850 900 750
( C)
Ca/Al molar ratio 0 0 0 0.50
Liquid feeding rate 0.76 0.75 0.68 0.76
(ml/min)
W/mHAc (g catalyst 6.152 6.181 6.830 6.085
min/g HAc)
GC1HSV (h1) 6772 6741 6100 6847
Gas yield (g/g liquid fed) 0.343 0.338 0.292 0.347
Liquid yield (g/g liquid fed) 0.672 0.669 0.712 0.651
Recovery 1.014 1.007 1.004 0.998
Carbon conversion (%) 98.87 98.33 87.15 100.97

Gas yields (g/g HAc)


H2 0.1303 0.1239 0.1062 0.1314
CO 0.1533 0.1655 0.1884 0.1697
CO2 1.2053 1.1784 0.9737 1.2069
CH4 0.0014 0.0014 0.0028 0.0026

Gas composition (% mol, N2 and H2O free)


H2 66.42 65.40 64.64 66.14
CO 5.58 6.24 8.20 6.10
CO2 27.92 28.27 26.93 27.60
CH4 0.09 0.09 0.22 0.16

due to catalyst deactivation probably caused by coking.


However, for the catalyst calcined at 900  C a fast decrease
in H2 and CO2 yields and a lower decrease of CO yield are
observed as the experiment evolves.
Fig. 3 – Influence of the calcination temperature on the
In Fig. 3d complete carbon conversion is observed for the
evolution of gas yields and carbon conversion with time.
catalyst calcined at 750 and 850  C but not at 900  C.
(a) H2 yield, (b) CO2 yield, (c) CO yield and (d) carbon
The main reactions involved are steam reforming (Eq. (1))
conversion; (-) 750 8C, (B) 850 8C and (6) 900 8C.
and water gas shift (Eq. (2)) that occur simultaneously:

CH3 COOH þ 2H2 O / 2CO2 þ 4H2 (1)


H2 and CO, is not capable of reducing the phases present in the
CO þ H2 O 4 CO2 þ H2 (2)
catalyst either. In addition, the surface area of the catalyst
The decrease in the H2 and CO2 yields and the increase in decreases when the calcination temperature increases, and
the CO yield when the calcination temperature increases this fact can influence reforming activity.
can be explained by the water gas shift reaction. However,
the lower carbon conversion observed with the catalyst 3.2. Influence of calcium addition as a promoter
calcined at 900  C indicates a lower reforming activity, espe-
cially with time. The experiment carried out with the catalyst These experiments were carried out in order to find out if the
calcined at 900  C also shows a higher methane yield, which addition of calcium as a modifier of the Ni/Al catalyst has
could be related to a lower reforming activity. Galdámez some significant effect on the process. The experimental
et al. [18] explained the participation of the catalyst in the conditions and the overall results obtained for the experiment
steam reforming of methane (Eq. (3)). with calcium addition as a promoter are shown in Table 1. A
comparison of the results obtained with the Ni/Al catalyst
CH4 þ H2 O 4 CO þ 3H2 (3)
(run 1) and Ni/Al/Ca catalyst (run 4), both calcined at the
XRD analyses (Fig. 2) showed that at higher calcination same temperature, indicates very similar values of H2 and
temperatures a larger proportion of crystalline spinel CO2 yields, while CO and CH4 yields are slightly higher for
(NiAl2O4) phase appears. This phase is more difficult to reduce. the calcium promoted catalyst.
The lower reforming activity of the catalyst calcined at 900  C Fig. 4 shows a comparison of the evolution of the H2 yield
could be due to the fact that the catalyst was not properly with time for the Ni/Al and Ni/Al/Ca catalysts. The same deacti-
reduced and harder reduction conditions were needed in vation tendency with time is observed for both catalysts, so the
order to obtain metallic nickel [27]. The reaction atmosphere, same resistance to coke deactivation is probably taking place.
4392 international journal of hydrogen energy 33 (2008) 4387–4396

In addition, coke could be the cause of catalyst deactivation


in steam reforming of pyrolysis oil–aqueous fraction [23].
Therefore, an addition of oxygen in small quantities can
provide less coking to the process as well as slightly helping
to compensate for the endothermicity of the steam reforming
reactions. These experiments were thus performed in order to
study the effect of oxygen addition to the steam reforming.
Table 2 shows the experimental conditions and the overall
results obtained for these experiments. The Ni/Al catalyst
calcined at 750  C was used in all cases and different concen-
trations of oxygen were fed ranging from 0 to 8% (percentage
of oxygen necessary for the stoichiometric combustion of
HAc).
As shown in Fig. 5a, the H2 yield decreases when the
percentage of oxygen addition increases. This decrease is
moderate at lower inlet oxygen concentrations (2 and 4%)
Fig. 4 – H2 yield evolution with time, influence of calcium but there is no difference when comparing these two experi-
addition. Calcination temperature of 750 8C; (-) Ni/Al ments. In contrast, with 8% oxygen an important decrease in
catalyst and (,) Ni/Al/Ca catalyst. the H2 yield occurs. The same tendency is observed for the
CO2 yield (Fig. 5b) at short reaction time. However, for times
longer than 1 h reaction CO2 yields for 2 and 4% oxygen are
higher than 0% oxygen. The opposite performance to that of
Although the results of XRD characterization (Fig. 2a and d) the H2 yield is observed for the CO yield (Fig. 5c). This is prob-
indicate different crystalline phases in the Ni/Al and Ni/Al/Ca ably because partial oxidation (Eq. (4)) is taking place at higher
catalysts, both calcined at 750  C, the performance in steam oxygen concentrations.
reforming is similar in activity and stability. It may perhaps
CH3 COOH þ O2 / 2CO þ 2H2 O (4)
be necessary to test these catalysts in other operational condi-
tions in order to know of any positive effects produced by the On the other hand, for the experiment conducted with no
addition of calcium. oxygen addition the carbon conversion decreases with time
(Fig. 5d). The addition of oxygen has a favourable effect
3.3. Influence of oxygen addition to the steam reforming because it keeps the carbon conversion of the process
process constant up to a value of 100%.
These results seem to be of interest since the O2 presence
Coke is a serious operational problem that is endemic to on the one hand maintains a constant activity over time but,
steam reforming and depends on the nature of the feed [32]. on the other hand, decreases the yield of hydrogen. This

Table 2 – Experimental results at 650 8C, with a catalyst weight of 1.1 g of Ni/Al catalyst calcined at 750 8C with oxygen
addition
Run # 1 5 6 7 8 9

Reaction time (h) 2 2 2 2 6 6


O2 addition (%) 0 2 4 8 0 4
Liquid feeding rate (ml/min) 0.76 0.76 0.77 0.75 0.75 0.74
W/mHAc (g catalyst min/g HAc) 6.152 6.096 6.026 6.179 6.205 6.246
GC1HSV (h1) 6772 6835 6914 6743 6715 6670
Gas yield (g/g liquid fed) 0.343 0.344 0.344 0.330 0.339 0.351
Liquid yield (g/g liquid fed) 0.672 0.620 0.667 0.693 0.689 0.688
Recovery 1.014 0.964 1.011 1.024 1.028 1.039
Carbon conversion (%) 98.87 99.83 100.27 100.52 98.09 102.64

Gas yields (g/g HAc)


H2 0.1303 0.1224 0.1203 0.1096 0.1255 0.1233
CO 0.1533 0.1595 0.1704 0.2467 0.1546 0.1775
CO2 1.2053 1.2134 1.2028 1.0764 1.1908 1.2246
CH4 0.0014 0.0000 0.0000 0.0037 0.0017 0.0007

Gas composition (% mol, N2 and H2O free)


H2 66.42 64.78 64.30 62.07 65.75 64.28
CO 5.58 6.03 6.50 9.97 5.78 6.65
CO2 27.92 29.19 29.20 27.70 28.36 29.03
CH4 0.09 0.00 0.00 0.26 0.11 0.05
international journal of hydrogen energy 33 (2008) 4387–4396 4393

Fig. 6 – Influence of the oxygen addition for 6 h of reaction


time. (a) H2 yield and (b) carbon conversion; (-) O% O2, (,)
4% O2. Ni/Al catalyst calcined at 750 8C.

Fig. 5 – Influence of the oxygen addition on the evolution of


gas yields and carbon conversion with time. (a) H2 yield, (b)
CO2 yield, (c) CO yield and (d) carbon conversion; (,) 0% O2,
(C) 2% O2, (6) 4% O2 and (;) 8% O2. Calcination
temperature of 750 8C.

decrease is moderate at lower inlet oxygen concentrations (2


and 4%). It follows that better operation conditions, in terms
of O2 concentration, are a compromise between keeping
high and constant levels of catalyst activity over time and
high levels of hydrogen yield. Therefore 4% of oxygen is a suit-
able value that can give a process with less coke, maintaining
the carbon conversion up to 100% with high hydrogen yield.
In order to prove if catalyst activity is maintained over the
operation time, some reforming experiments have been per-
formed for longer reaction times (6 h) in the absence and pres-
ence of O2 (0 and 4%). Results of these experiments are shown
in Table 2 (runs 8 and 9). The hydrogen yield is similar for the
two experiments, while the CO and CO2 yields are slightly
higher with 4% O2. Fig. 6a and b shows the evolution of the
H2 yield and carbon conversion with time.
The hydrogen yield (Fig. 6a) obtained with 4% O2 is kept
constant during 6 h of reaction time while a decrease at the
beginning of the experiment is observed when oxygen is not
added. The H2 yield is very similar for both runs after 2 h of
experiment, indicating almost no hydrogen penalty with Fig. 7 – SEM images of Ni/Al catalyst after 6 h reaction: (a)
oxygen addition. with no oxygen addition and (b) with 4% oxygen addition.
4394 international journal of hydrogen energy 33 (2008) 4387–4396

carbon on the surface of the catalyst when no oxygen was


added to the reaction and a clearer surface when oxygen
was added.
In order to know the crystalline phases present in the cata-
lyst after reaction, XRD analyses have been performed. Fig. 8
shows the results of the catalysts after 6 h of reaction. Higher
proportions of NiO and NiAl2O4, the oxidized phases, are
observed when 4% O2 is added, but metallic Ni is detected
with similar peak intensity. This analysis also indicates the
presence of coke in the two samples.

3.4. Literature comparison of acetic acid catalytic


steam reforming

Fig. 9a and b shows acetic acid conversion and H2 yield,


respectively, versus the inverse of space velocity, expressed
as W/mHAc, for catalytic steam reforming experiments carried
out at a reaction temperature close to 650  C (from 600 to
665  C). Some works, such as this, do not calculate acetic
acid conversion. Therefore carbon conversion has been
assimilated to acetic acid conversion.
In Fig. 9b a theoretical thermodynamic equilibrium for
hydrogen yield of acetic acid steam reforming was simulated
at the operational conditions of this work. The analysis of
these figures must be done carefully because authors use
Fig. 8 – XRD patterns for the Ni/Al catalyst after 6 h reaction:
different operating conditions (S/C molar ratio, temperature,
(a) with no oxygen addition and (b) with 4% oxygen
catalyst and fixed or fluidized bed).
addition.
In Fig. 9a the acetic acid conversion for the majority of the
studies is between 90 and 100%. This work shows an acetic
acid conversion close to 100%, which is one of the best results
On the other hand, the carbon conversion (Fig. 6b) is for the of the literature survey.
majority of time slightly higher when 4% O2 is added. This In Fig. 9b it is observed that the H2 yield obtained by many
may be due to the positive effect of oxygen on coke removal. authors is close to equilibrium and among them the results of
SEM images (Fig. 7a and b) were taken of Ni/Al catalyst after this work. The H2 yield of the study carried out by Galdámez et al.
6 h reaction in absence and presence of 4% oxygen, respec- [18] for a W/mHAc ¼ 2.0 g catalyst min/g HAc is similar to the H2 yield
tively. These images showed a major presence of filamentous obtained in this work for a W/mHAc ¼ 6.2 g catalyst min/g HAc, both

Fig. 9 – Literature comparison of acetic acid catalytic steam reforming: (a) acetic acid conversion and (b) H2 yield.
international journal of hydrogen energy 33 (2008) 4387–4396 4395

using fluidized bed reactor. The best result obtained in the work of [3] Aupretre F, Descorme C, Duprez D, Casanave D, Uzio D.
Galdámez et al. can be explained by the wall effect, because the Ethanol steam reforming over MgxNi1xAl2O3 spinel oxide-
reactor is made of stainless steel. The study of Galdámez et al. supported Rh catalysts. J Catal 2005;233:464–77.
[4] Czernik S, French RJ, Magrini-Bair KA, Chornet E. The production
also shows lower values of H2 yield that can be explained by the
of hydrogen by steam reforming of trap grease – progress in
lower W/mHAc values used. H2 yields obtained by Rioche et al. [15] catalyst performance. Energy Fuels 2004;18:1738–43.
and Basagiannis and Verykios [20] must be a consequence of the [5] Marquevich M, Coll R, Montané D. Steam reforming of
lower S/C molar ratio used, although the latter did not reach sunflower oil for hydrogen production. Ind Eng Chem Res
complete acetic acid conversion. Takanabe et al. [17], who worked 2000;39:2140–7.
with a Pt/ZrO2, also obtained an H2 yield close to equilibrium at [6] Zhang B, Tang X, Li Y, Xu Y, Shen W. Hydrogen production
similar operation conditions (S/C ¼ 5 and 602  C reaction tempera- from steam reforming of ethanol and glycerol over ceria-
supported metal catalysts. Int J Hydrogen Energy 2007;32:
ture) but at higher W/mHAc values.
2367–73.
[7] Garcia L, French R, Czernik S, Chornet E. Catalytic steam
reforming of bio-oils for the production of hydrogen: effects
4. Conclusions of catalyst composition. Appl Catal A 2000;201:225–39.
[8] Wang D, Czernik S, Chornet E. Production of hydrogen from
Coke is a serious operational problem that produces the deac- biomass by catalytic steam reforming of fast pyrolysis oils.
Energy Fuels 1998;12:19–24.
tivation of catalysts, especially those nickel based catalysts
[9] Wang D, Czernik S, Montané D, Mann M, Chornet E. Biomass
used in steam reforming. When organic feedstocks such as
to hydrogen via fast pyrolysis and catalytic steam reforming
bio-oil or its fractions are reformed, there is a greater tendency of the pyrolysis oil or its fractions. Ind Eng Chem Res 1997;36:
for the system to be fouled by coking. Therefore, different 1507–18.
strategies are needed in order to avoid these problems, and [10] Diebold JP. A review of the chemical and the physical
fluidized beds are found to perform better than fixed ones. mechanisms of the storage stability of fast pyrolysis bio-oils.
The effects of various factors have been studied in fluidized National Renewable Energy Laboratory, NREL/SR-570-27613.
[11] Qi Z, Jie C, Tiejun W, Ying X. Review of biomass pyrolysis oil
bed steam reforming with acetic acid as a model compound
properties and upgrading research. Energy Conv Manage
at 650  C. First of all, calcination temperatures of 750, 850 2007;48:87–92.
and 900  C have been studied. A lower reforming activity [12] Czernik S, Bridgwater AV. Overview of applications of
and stability has been observed when using the catalyst biomass fast pyrolysis oil. Energy Fuels 2004;18:590–8.
calcined at 900  C, probably due to the increase in spinel phase [13] Wang D, Montané D, Chornet E. Catalytic steam reforming of
that is harder to reduce. The effect of calcium as a promoter of biomass-derived oxygenates: acetic acid and
the catalyst was studied. The incorporation of a Ca/Al ratio of hydroxyacetaldehyde. Appl Catal A 1996;143:245–70.
[14] Marquevich M, Czernik S, Chornet E, Montané D. Hydrogen
0.50 under the experimental conditions did not result in an
from biomass: steam reforming of model compounds of fast-
improvement in reforming activity and stability. Finally, the pyrolysis oil. Energy Fuels 1999;13:1160–6.
influence of oxygen addition to the steam reforming process [15] Rioche C, Kulkarni S, Meunier FC, Breen JP, Burch R. Steam
indicated that an excess of oxygen (8%) can lead to a lower reforming of model compounds and fast pyrolysis bio-oil on
reforming activity decreasing the H2 and CO2 yields and supported noble metal catalysts. Appl Catal B 2005;61:130–9.
increasing the CO and CH4 yields. On the other hand, 4% [16] Hu X, Lu G. Investigation of steam reforming of acetic acid to
hydrogen over Ni–Co metal catalyst. J Mol Catal A 2007;261:
oxygen resulted in almost no penalty in hydrogen yield
43–8.
production providing a process with a probability of less
[17] Takanabe K, Aika K, Seshan K, Lefferts L. Catalyst
coke because the carbon conversion was maintained up to deactivation during steam reforming of acetic acid over Pt/
100%. ZrO2. Chem Eng J 2006;120:133–7.
[18] Galdámez JR, Garcia L, Bilbao R. Hydrogen production by
steam reforming of bio-oil using coprecipitated Ni–Al
catalysts. Acetic acid as a model compound. Energy Fuels
Acknowledgments 2005;19:1133–42.
[19] Basagiannis AC, Verykios XE. Reforming reactions of acetic
The authors express their gratitude to the Spanish Ministry of acid on nickel catalysts over a wide temperature range. Appl
Education and Science (MEC) (Research Project Ref. No. Catal A 2006;308:182–93.
CTQ2004-06279) and to the Government of Aragon (Research [20] Basagiannis AC, Verykios XE. Catalytic steam reforming of
acetic acid for hydrogen production. Int J Hydrogen Energy
Project Ref. No. PIP185/2005) for providing financial support
2007;32:3343–55.
for the work. [21] Vagia EC, Lemonidou AA. Thermodynamic analysis of
hydrogen production via steam reforming of selected
components of aqueous bio-oil fraction. Int J Hydrogen
references Energy 2007;32:212–23.
[22] Bimbela F, Oliva M, Ruiz J, Garcia L, Arauzo J. Hydrogen
production by catalytic steam reforming of acetic acid,
[1] Ni M, Leung DYC, Leung MKH, Sumathy K. An overview on a model compound of biomass pyrolysis liquids. J Anal Appl
hydrogen production from biomass. Fuel Process Technol Pyrol 2007;79:112–20.
2006;87:461–72. [23] Czernik S, French R, Feik C, Chornet E. Hydrogen by catalytic
[2] Ni M, Leung DYC, Leung MKH. A review on reforming bio- steam reforming of liquid byproducts from biomass
ethanol for hydrogen production. Int J Hydrogen Energy 2007; thermoconversion processes. Ind Eng Chem Res 2002;41:
32:3238–47. 4209–15.
4396 international journal of hydrogen energy 33 (2008) 4387–4396

[24] Sutton D, Kelleher B, Ross JRH. Review of literature on catalysts [30] Morioka H, Shimizu Y, Sukenobu M, Ito K, Tanabe E,
for biomass gasification. Fuel Process Technol 2001;73:155–73. Shishido T, et al. Partial oxidation of methane to synthesis
[25] Garcia L, Salvador ML, Arauzo J, Bilbao R. Catalytic steam gas over supported Ni catalysts prepared from Ni-Ca/Al-
gasification of pine sawdust. Effect of catalyst weight/ layered double hydroxide. Appl Catal A 2001;215:11–9.
biomass flow rate and steam/biomass ratios on gas [31] Lu Y, Liu Y, Shen S. Design of stable Ni catalysts for partial
production and composition. Energy Fuels 1999;13:851–9. oxidation of methane to synthesis gas. J Catal 1998;177:
[26] Ramos MC, Navascues AI, Garcia L, Bilbao R. Hydrogen 386–8.
production by catalytic steam reforming of acetol, a model [32] Trimm DL. Coke formation and minimisation during steam
compound of bio-oil. Ind Eng Chem Res 2007;46:2399–406. reforming reactions. Catal Today 1997;37:233–8.
[27] Garcia L, Salvador ML, Bilbao R, Arauzo J. Influence of [33] Grace JR, Li X, Lim CJ. Equilibrium modelling of catalytic
calcination and reduction conditions on the catalyst steam reforming of methane in membrane reactors with
performance in the pyrolysis process of biomass. Energy oxygen addition. Catal Today 2001;64:141–9.
Fuels 1998;12:139–43. [34] Mukainakano Y, Li B, Kado S, Miyazawa T, Okumura K,
[28] Bangala DN, Abatzoglou N, Chornet E. Steam reforming of Miyao T, et al. Surface modification of Ni catalysts with trace
naphthalene on Ni–Cr/Al2O3 catalysts doped with MgO, TiO2, Pd and Rh for oxidative steam reforming of methane. Appl
and La2O3. AIChE J 1998;44:927–36. Catal A 2007;318:252–64.
[29] Arauzo J, Radlein D, Piskorz J, Scott DS. Catalytic [35] Al-Ubaid A, Wolf EE. Steam reforming of methane on
pyrogasification of biomass. Evaluation of modified nickel reduced non-stoichiometric nickel aluminate catalysts. Appl
catalysts. Ind Eng Chem Res 1997;36:67–75. Catal 1988;40:73–85.

You might also like