Вы находитесь на странице: 1из 8

Chemical Engineering Science 61 (2006) 1195 1202

www.elsevier.com/locate/ces

Sorption-enhanced steam reforming of methane in a uidized bed reactor


with dolomite as CO2 -acceptor
K. Johnsena, , H.J. Ryub , J.R. Graceb , C.J. Limb
a Institute for Energy Technology (IFE), P.O. Box 40, NO-2027 Kjeller, Norway
b Department of Chemical and Biological Engineering, University of British Columbia, 2216 Main Mall, Vancouver, Canada V6T 1Z4

Received 22 February 2005; received in revised form 15 August 2005; accepted 18 August 2005
Available online 28 September 2005

Abstract
An experimental investigation was conducted in which carbon dioxide was captured in order to shift the steam reforming equilibrium for
the production of hydrogen. An atmospheric-pressure bubbling uidized bed reactor (BFBR) of diameter 100 mm was operated cyclically
and batchwise, alternating between reforming/carbonation conditions and higher-temperature calcination conditions to regenerate the sorbent.
Equilibrium H2 -concentration of > 98% on a dry basis was reached at 600 C and 1.013 105 Pa, with dolomite as the CO2 -acceptor. The
hydrogen concentration remained at 9899 vol% (dry basis) after four reforming/calcination cycles. The total production time decreased with
an increasing number of cycles due to loss of CO2 -uptake capacity of the dolomite, but the reaction rate seemed unaffected. Variation of the
supercial gas velocity within the bubbling bed regime showed that the overall reaction rate was sufciently fast to reach equilibrium, making
bubbling bed reactors attractive for this process.
2005 Elsevier Ltd. All rights reserved.
Keywords: Hydrogen production; Sorption-enhancement; Fluidization; Catalysis; Reaction engineering; Separations

CH4 + 2H2 O CO2 + 4H2 ,

1. Introduction
Despite efforts to decrease energy consumption and greenhouse gas emissions, fossil fuels will continue to play an important role in the coming decades as developing nations increase their standard of living and major economies are slow
to adapt to change. Hydrogen is often referred to an important
potential energy carrier, but its advantages are unlikely to be
realized unless efcient means can be found to produce it with
reduced generation of CO2 .
Steam reforming of natural gas is the predominant production
route to hydrogen for large-scale industrial applications. For
methane the reactions are:
CH4 + H2 O CO + 3H2 ,
CO + H2 O CO2 + H2 ,

0
H298
= 206.2 kJ mol1 ,

(1)

0
H298
= 41.2 kJ mol1 ,

(2)

Corresponding author: Tel.: +47 63 80 61 26; fax: +47 63 81 55 53.

E-mail address: kim.johnsen@ife.no (K. Johnsen).


0009-2509/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2005.08.022

0
H298
= 165 kJ mol1 .

(3)

Steam methane reforming (SMR) is normally carried out


at 800900 C and 1530 105 Pa, with nickel on an alumina support as the catalyst. A typical industrial reformer
contains an array of catalyst-lled tubes suspended in a huge
furnace, supplying the heat for the highly endothermic reforming reactions. These xed bed reformers suffer from a
number of limitations, making them inefcient (Adris et al.,
1996; Chen et al., 2003). One of the most serious constraints
relates to conversion of methane, which is limited by the
thermodynamic equilibrium of the reversible reactions. For
conventional xed bed reformers, reaction temperature has
to be in the region of 800900 C to achieve complete conversion of methane. At this elevated temperature the catalyst
suffers deactivation due to carbon formation, also resulting
in blockage of reformer tubes and increased pressure drops
(Trimm, 1997).
The thermodynamic equilibrium can be shifted to give more
favourable yields by removing either hydrogen or CO2 . Pure
hydrogen can be extracted from the reactor by perm-selective

1196

K. Johnsen et al. / Chemical Engineering Science 61 (2006) 1195 1202

Fig. 1. Hydrogen content at equilibrium as a function of temperature for a


pressure of 1.013 105 Pa, a H2 O : CH4 molar ratio of 3 and a CaO : CH4
molar ratio of 2.

membranes made of palladium or its alloys. A number of experimental and modelling studies have been carried out to prove
this concept (e.g. Adris et al., 1991, 1994; Chen et al., 2003;
Dogan et al., 2003).
Another way of shifting the equilibrium is by adding a CO2 acceptor to the reactor. Carbon dioxide is then converted to a
solid carbonate as soon as it is formed, shifting the reversible reforming and water-gas shift reactions beyond their conventional
thermodynamic limits. Regeneration of the sorbent releases relatively pure CO2 suitable for sequestration. Sorption-enhanced
steam reforming and the use of calcium based CO2 sorbents
been demonstrated in previous work (e.g. Balasubramanian et
al., 1999; Brun-Tsekhovoi et al., 1988; Han and Harrison, 1994;
Ortiz and Harrison, 2001; Silaban and Harrison, 1995; Silaban
et al., 1996). For example, Balasubramanian et al. (1999) added
a calcium-based CO2 acceptor to a commercial steam reforming catalyst producing > 95% H2 in a laboratory-scale xed
bed reactor. Han and Harrison (1994) used CaO to capture CO2 ,
overcoming the equilibrium limitation and achieving complete
CO conversion. In addition to reactions (1)(3) above, the noncatalytic highly exothermic carbonation reaction is included in
sorption-enhanced steam reforming, i.e.,
CaO(s) + CO2 (g) CaCO3 (s),
0
H298
= 178 kJ mol1 .

(4)

The advantages of combining steam reforming with in situ


capture of CO2 can be understood from the thermodynamics.
Fig. 1 shows the equilibrium hydrogen concentration as a function of reaction temperature at ambient pressure and at a steamto-carbon molar ratio of 3, with the predictions based on the
HSC thermodynamic software package (Outokumpu Research
Oy, Finland).
The hydrogen concentration is predicted to reach a maximum of 98% at 600 C for a CaO/CH4 ratio of 2,
whereas the equilibrium concentration of conventional steam
reforming is only 74% at that temperature. Fig. 1 shows that

sorption enhancement enables lower reaction temperatures,


which may reduce catalyst coking and sintering, while enabling use of less expensive reactor wall materials. In addition,
heat release by the exothermic carbonation reaction supplies
most of the heat required by the endothermic reforming reactions. However, energy is required to regenerate the sorbent
to its oxide form by the energy-intensive calcination reaction
(reverse of Eq. (4)), which represents a challenge in terms of
heat transfer and reactor construction. The high temperature
required for CaCO3 regeneration could also cause sintering
of solids, affecting the long-term performance of sorbent.
Previous workers (Abanades and Alvarez, 2003; Silaban and
Harrison, 1995; Silaban et al., 1996) have reported that the
absorption capacity for Ca-based sorbents decays as a function
of the number of calcinationcarbonation cycles. Silaban et al.
(1996) found that dolomite (CaCO3 MgCO3 ) was superior to
limestone as a sorbent in this respect, with better multi-cycle
performance. Its advantages were attributed to differences
between the structural properties of calcined dolomite and calcined limestone. Initial calcination produces complete decomposition of dolomite, but carbonation conditions are at such
high temperatures that only CaO forms carbonate. The excess
pore volume created by MgCO3 decomposition is believed to
be responsible for the more favourable cycling performance.
Whereas hydrogen removal by membranes can be carried
out continuously in a single reactor, continuous reforming with
CO2 removal requires either that there be parallel reactors operated alternatively and out of phase in reforming and sorbent
regeneration modes, or that sorbent be continuously transferred
between the reformer/carbonator and regenerator/calciner. Fluidized bed reactors are commonly used in processes where catalysts must be continuously regenerated, while also facilitating heat transfer, temperature uniformity and higher catalyst
effectiveness factors. However, experimental investigations for
enhancement of steam reforming by in situ removal of CO2
have mainly been conducted in small-scale xed bed reactors.
Hufton et al. (1999) used hydrotalcite as the CO2 acceptor in
a xed-bed sorption-enhanced reaction process (SERP) for hydrogen production. The sorbent had to be periodically regenerated by pressure swing adsorption (PSA). Balasubramanian
et al. (1999) demonstrated the concept of sorption enhancement with CaO as CO2 -acceptor, again based on a xed bed
reactor. Only Brun-Tsekhovoi et al. (1988) have utilized a
uidized bed reactor for sorption-enhanced reforming. While
this paper is very useful, the authors did not consider cycling, employed excessively large sorbent particles and operated for only very brief periods of time. Abanades et al.
(2004a) used a pilot-scale uidized bed reactor to investigate
the carbonation of CaO capturing CO2 from high temperature combustion ue gases. They found high capture efciencies for a typical ue gas composition of 15% CO2 in air,
comparable to the concentration of CO2 from steam methane
reforming.
A schematic illustration of a continuous sorption enhanced
SMR process based on parallel uidized bed reactors is shown
in Fig. 2. The uidized beds could operate in different ow
regimes, e.g. in bubbling or fast uidization. In this concept,

K. Johnsen et al. / Chemical Engineering Science 61 (2006) 1195 1202

1197

sorbent between cycles. Due to the lack of experimental data


on the performance of the overall process in such reactor conguration, special attention was given to the multi-cycle performance of both the sorbent and the catalyst under these circumstances. The rate of the combined reactions was also evaluated
using different supercial gas velocities within the bubbling
regime.
2. Experimental
2.1. Bubbling uidized bed reformer (BFBR)

Fig. 2. Simplied schematic of the sorption-enhanced SMR process.

catalyst and sorbent are mixed in the reformer, where sorptionenhanced steam reforming is performed. The product gas
from the reformer mainly consists of hydrogen and steam,
with minor quantities of CO, CO2 and unconverted methane.
Carbonated sorbent is transferred to the regenerator where heat
is supplied for the endothermic calcination reaction, either by
burning fuel in the regenerator or by indirect heating from an
external heat source. Heating of a dense bubbling bed could
also be accomplished by the insertion of heat transfer tubes
into the bed. Heat transfer would then occur between the uidized bed and the submerged tube surfaces. These tubes will
at the same time act as bafes reducing the gas bubble size,
enhancing interphase mass transfer. A portion of the hydrogen
produced in the reformer might be burned externally, supplying
heat to the tubes and eliminating the need for additional fossil
fuel. Indirect heating has the advantage of producing pure CO2
ready for sequestration. Direct burning of fuel inside the regenerator would be a more efcient way of supplying the heat, but
would require downstream separation of CO2 , unless hydrogen is burned with pure oxygen. To avoid separation processes
downstream, CO2 and/or H2 O can be used as the uidizing gas
in the regenerator. Calcium-based sorbents have the advantage
of being available at low cost, but cannot maintain the capture capacity upon multiple reforming/regeneration cycles. A
make-up stream of fresh sorbent must be included to maintain
capture capacity. This addition could be to the calciner, with
withdrawal in the reformer, as indicated in Fig. 2. Synthetic
sorbents, such as Li2 ZrO3 , have better multi-cycle stability,
but their cost would require them to sustain > 10, 000 cycles
to compete with natural sorbents (Abanades et al., 2004b).
Coupling of two bubbling beds would have the advantage of
low rates of attrition due to low gas and particle velocities,
and the relatively slow carbonation reaction rate will be facilitated in this ow regime. There is a lack of experimental
data on sorption-enhanced SMR in uidized beds in the open
literature.
In our work, described below, a bubbling uidized bed was
used for sequential sorption-enhanced steam reforming and regeneration of sorbent without separating the catalyst from the

A schematic of the reactor system, used previously


(Constantineau, 2004) in studies of uidized bed roasting of
zinc concentrates, is shown in Fig. 3. The major components
consist of a pre-heater, a 0.66 m high by 0.1 m ID stainless
steel uidized bed reactor with expanded freeboard, a ltration
unit and a gas cooler unit. Methane was fed to the upper part
of the pre-heater where it was mixed with steam. A removable
stainless steel gas distributor plate, with 34 drilled 1.2 mm
diameter holes on a hexagonal grid, was placed between the
pre-heater and the reactor. The preheated reactant gas passed
through a mixture of commercial Ni-based steam-reforming
catalyst (Haldor Topsoe A/S, R-67R-7H) and calcined dolomite
(Franzefoss A/S, Arctic Dolomite SHB). Dolomite was used in
preference to limestone because of initial tests indicating better ability to sustain performance in cyclical operation. Three
different zones of the reactor were each heated by electrical
furnaces, which could be controlled individually. Temperatures and pressure drops were recorded by a data acquisition
system. Teon bags were used for gas sampling from a port
at the outlet of the freeboard zone. The composition of these
samples was determined using a gas chromatograph (Shimadzu
GC-8A).
2.2. Sample preparation
The composition of the dolomite is provided in Table 1. This
dolomite was chosen because it did not contain sulphur, which
is poisonous to the reforming catalyst.
Prior to the experiments, dolomite and catalyst were sieved
to ensure particle sizes between 125300 and 150250 m,
respectively. The dolomite had to be rst calcined to obtain
the desired oxide form. This was accomplished at 850 C
in N2 without the catalyst present. A portable gas analyzer
(Horiba PG-250) was used to determine complete calcination,
corresponding to the disappearance of CO2 in the product
gas. The reactor was then cooled, and part of the calcined
dolomite removed and stored in a dessicator for later use,
while the rest was mixed with catalyst and re-injected into the
reactor.
2.3. Experimental procedure
The reactor had no feeding lines for solids, and was therefore operated batchwise, with periodic calcination at higher

1198

K. Johnsen et al. / Chemical Engineering Science 61 (2006) 1195 1202

Fig. 3. Schematic of reformer unit.

Table 1
Analysis of arctic dolomite SHB, data from Franzefoss AS
Species

CaO

MgO

SiO2

Al2 O3

Fe2 O3

Na2 O

TiO2

K2 O

Loss by ignition

Conc. [wt%]

32

20.3

0.7

0.1

0.1

0.003

0.005

0.004

46.3

temperatures to regenerate the dolomite. The experimental investigation can be divided into two parts: multi-cycle tests
and tests where the supercial gas velocity was varied. Fresh
dolomite and catalyst were used for the investigation of the effect of gas velocity in order to make the results for different
gas velocities comparable. The total initial bed mass was 3.1 kg
for all runs, with a catalyst-to-calcined dolomite mass ratio of
2.5. During the calcination stages of the multi-cycle tests, pure
N2 was fed to the reactor. No effort was made to separate the
catalyst from the dolomite between cycles. To ensure that the
catalyst was active, reduction of the catalyst was performed in
a H2 /N2 mixture at 650 C for 12 h prior to each reforming

period. This is equivalent to a continuous process where a portion of oxidized catalyst is returned to the reducing atmosphere
in the reformer, where it is reduced back to its active form. The
reforming reaction was always carried out at 600 C and ambient pressure. The experimental conditions for the BFBR unit
are summarized in Table 2.
3. Results and discussion
The reaction conditions corresponded to operation in the bubbling bed ow regime. With 0.9 kg calcined dolomite present,
the time required for complete carbonation of CaO in dolomite

K. Johnsen et al. / Chemical Engineering Science 61 (2006) 1195 1202

1199

Table 2
BFBR experimental conditions
Parameters

Values

Total mass of particles in bed (kg)


Catalyst-to-calcined dolomite mass ratio
(dimensionless)
Bulk density of mixture (kg/m3 )
Static bed height (m)
Catalyst particle size range (m)
Dolomite particle size range (m)
Reforming temperature ( C)
Supercial velocity (m/s)
Steam-to-carbon molar feed ratio
Calcination temperature ( C)
Calcination atmosphere

3.1
2.5
1300
0.3
150250
125300
600
0.032a , 0.064, 0.096
3
850
N2

a Gas velocity for multi-cycle test.

Fig. 5. Time variation of temperature in bed zones.

Fig. 4. Outlet composition (dry basis) as a function of time.

was calculated to be 170 min for a supercial gas velocity of


0.032 m/s at 600 C, based on the assumption that all carbon
fed reacted with CaO to yield CaCO3 . The total time of operation was 5 h for each run. A typical response curve is shown in
Fig. 4, with the dry gas composition plotted as a function of
time.
The hydrogen concentration is stable at 9899 vol% on a dry
basis for a period of 150180 min, often referred to as the prebreakthrough period by previous authors (e.g. Han and Harrison, 1994), before there is sudden drop in concentration to
7274%. The opposite trend is observed for the CO2 concentration, where there is a sudden increase from 0.3 to 1314%
after the same interval. This characteristic breakthrough occurs
when the amount of CO2 produced by steam reforming exceeds the sorption capacity of the CaO. Fig. 4 clearly shows
the marked enhancement of hydrogen production achieved by
in situ capture of CO2 . The shape of the curve is typical for
sorption-enhanced steam reforming. Due to the limited number of sample points, the time of the breakthrough cannot be
given precisely, but it seems to correspond to a time between
150 and 180 min, in good agreement with the calculated time
for complete carbonation (assuming 100% calcium utilization)

of 170 min. After complete carbonation, the hydrogen concentration dropped to a value corresponding to equilibrium of
steam methane reforming of 73 vol% on a dry basis. At this
point, no CaO was left to react with CO2 , so that the reaction enhancement was lost. This period is often referred to as
the post-breakthrough period, and re-calcination has to be performed to reactivate the sorbent. Combining the strongly endothermic steam reforming with the exothermic carbonation
reaction makes the overall reforming reaction almost thermally
neutral. Carbonation of CaO (Eq. (4)) is a reversible reaction,
and temperature control is very important to prevent the undesired reverse calcination reaction in the reformer. Temperature
uniformity promoted by rapid mixing of the solids makes uidized beds well suited for processes where temperature uniformity is important. Two thermocouples were placed in the dense
bed zone, one (T1) just above the distributor and the other (T2)
0.19 m above. Typical temperature traces for one run are shown
in Fig. 5.
The difference in temperature between the two positions, T1
and T2, was nearly constant at 34 C during the entire course
of reaction, conrming the excellent temperature uniformity
of the bubbling uidized bed. Another feature observed from
Fig. 5 is the temperature drop after 150 min. This corresponds to the start of the breakthrough period also observed in
Fig. 4, caused by the diminishing exothermic carbonation
reaction, while the endothermic reforming reaction continues.
3.1. Multi-cycling
In order to make the process continuous, the sorbent must be
regenerated after completion of the carbonation stage. Several
previous groups (e.g. Li et al., 2005; Ortiz and Harrison, 2001;
Silaban and Harrison, 1995) have investigated the multi-cycle
performance of CaO-based sorbents. Abanades and Alvarez
(2003) included previously published multi-cycle results when
they reported an unavoidable decay in carbonation conversion
that was dependent on the number of cycles. Most of these
investigations have been carried out using thermo-gravimetric

1200

K. Johnsen et al. / Chemical Engineering Science 61 (2006) 1195 1202

Fig. 6. H2 and CO2 concentrations (dry basis) as functions of number of


cycles.

analysis (TGA), either in pure CO2 or in simulated reforming


environments. There is little information on how sorption enhancement is affected by carbonationcalcination cycling in a
uidized bed. Our BFBR was operated batchwise, with periodic regeneration of sorbent without physically separating the
catalyst from the dolomite. The solid mixture was exposed to
hydrogen after each period of calcination to ensure that the
nickel in the catalyst was in the reduced active form, corresponding to a continuous process where the solids would be
exposed to a reducing atmosphere in the reformer.
Fig. 6 shows the concentrations of hydrogen and carbon
dioxide for different numbers of carbonationcalcination cycles. The rst cycle is not included in this gure because the
dolomite had been exposed to air, reducing its absorption capacity, hence making comparison with the other cycles difcult.
It is clear from Fig. 6 that the duration of the pre-breakthrough
period is reduced somewhat with an increase in the number of
cycles. This reduction is due to loss of CaO capacity, but the hydrogen concentration remained constant at 9899%, suggesting
that the equilibrium concentration of the combined reactions
(1)(4) was reached for each cycle. The hydrogen concentration during the post-breakthrough period was again at equilibrium for successive cycles, indicating that the catalytic activity
remained sufciently high to reach equilibrium upon cycling.
The breakthrough period was characterized by onset of the slow
carbonation reaction rate regime, where diffusion through the
solid product layer limited the rate of reaction. The slopes of
the breakthrough curves indicate that the global reaction rate
was not signicantly affected by the number of cycles. However, the limited number of cycles and the low gas velocity in
this study make it hard to conclude that the rate of reaction
would in general not be affected by multi-cycling.
Silaban and Harrison (1995) reported that the loss of capacity
was associated with a change in structural properties of the
sorbent. It was claimed that reduced porosity left the interior of
the sorbent particles inaccessible to CO2 . As a consequence, the
CO2 uptake capacity decreased with cycling, consistent with
Fig. 6.

Brun-Tsekhovoi et al. (1988) employed relatively large


dolomite particles (1.3 mm average), to facilitate their physical
separation from the catalyst (250 m) before regeneration. A
separation stage should be avoided if possible as it would add
extra components and complexity to the system, lead to additional attrition and cause extra heat losses. Khotomlyanskii
et al. (1970) studied the separation of catalyst from a heavier
heat transfer agent in a uidized bed using particles differing
signicantly in density and size. The catalyst density in the
current study was 2200 kg/m3 , whereas the sorbent density
depended on the degree of carbonation, with a possible range
from 1560 kg/m3 (fully calcined) to 2230 kg/m3 (completely
carbonated), the latter density being similar to that of the catalyst. Note that separation is more sensitive to different particle
densities than to differences in particle size (Rowe and Nienow,
1976). Since steam reforming catalysts commonly encounter
temperatures similar to those employed in calcination, it should
be possible to expose the catalyst to the calcination conditions
without separation. The present investigation shows that catalytic activity remained after 4 cycles, without any separation
of the particles. Nitrogen would not be used as the uidizing
gas for calcination in an industrial application, due to dilution
of CO2 leaving the regenerator for sequestration. In that case,
CO2 itself, or possibly steam, would be more realistic as the
medium to avoid separation processes downstream. Silaban
and Harrison (1995) reported that a CO2 atmosphere had an
adverse effect on CaO sorption capacity during multi-cycling.
It is therefore possible that the reduction in production time as
a function of cycles, observed from Fig. 6, would have been
greater if the carbonated sorbent had been calcined in atmospheres other than 100% N2 . However, Ortiz and Harrison
(2001) report no signicant difference in loss of multi-cycle
durability for different regeneration atmospheres, except when
regeneration was carried out in pure nitrogen at 950 C, using
dolomite as sorbent. It was beyond the scope of this work to
investigate the effect of different calcination atmospheres and
only nitrogen was used.

3.2. Increasing supercial gas velocity


All multi-cycle runs were conducted at a supercial gas
velocity of 0.032 m/s, a very low velocity compared to normal
commercial uidization processes. From an industrial point of
view, a higher gas throughput would be advantageous. Keeping the total mass of solid constant, the supercial gas velocity
was increased to see its effect. The highest gas velocity investigated was 0.096 m/s, again lower, though less so, than expected
commercial velocities. However, it was desirable to investigate
whether bubbles bypassing would lead to signicantly lower
conversions than reported for xed bed reactors. Fig. 7 shows
the hydrogen concentration at three different supercial gas velocities.
As the gas velocity increased, the total production time decreased, as expected. The hydrogen concentration exceeded
95% on a dry basis for all three velocities. Both fresh dolomite
and fresh catalyst were used for each run, making the results

K. Johnsen et al. / Chemical Engineering Science 61 (2006) 1195 1202

1201

overall reaction rate was sufciently fast that equilibrium was


approached, making bubbling bed reactors an appealing choice
for this process.
Acknowledgements
This work was nancially supported by the Research Council
of Norway (RCN) and the Natural Sciences and Engineering
Research Council of Canada (NSERC). K.J. would also like
to acknowledge Det Norske Veritas (DNV) in association with
The Norwegian University of Science and Technology (NTNU)
for the scholarship which made his stay at the University of
British Columbia possible. The authors also thank Dr. Pierre
Constantineau for assistance in modifying the reactor.
Fig. 7. Hydrogen concentration (dry basis) in rst cycle as a function of
supercial gas velocity.

comparable. It is evident from Fig. 7 that both mass transfer and


reaction kinetics are fast enough to reach equilibrium within
the given range of operation conditions. If there was any loss
in catalyst activity, it was too small to reduce the conversion
either before or after the capacity of the sorbent was exhausted.
Higher velocities were not tested, but high CO2 capture efciencies by CaO have been reported (Abanades et al., 2004a,b)
for ue gases at supercial velocities of 1 m/s, indicating that
further increases in gas velocity are likely to be feasible. The
experiment at U = 0.064 m/s gave a somewhat lower maximum H2 concentration than the two other runs, possibly due to
the gas analysis method. There were always nitrogen peaks in
the chromatograms, originating from the use of sampling bags
and a syringe for manual injection into the GC, making the
presence of air inevitable. The proportion of N2 in the samples
was usually 24 vol%, but for U = 0.064 m/s a higher nitrogen concentration was observed. Hydrogen is the most volatile
gas, and any leakage from a syringe or GC injection port would
introduce air at the expense of hydrogen.
4. Conclusions
Equilibrium for sorption-enhanced steam reforming at
600 C and 1.013 105 Pa was reached for gas velocities in
the range of 0.0320.096 m/s. Multiple reforming-regeneration
cycles showed that the hydrogen concentration remained at
9899 vol% after 4 cycles. The total production time was
reduced with an increasing number of cycles due to loss of
CO2 -uptake capacity of the dolomite, but the reaction rate
seemed to be unaffected for the conditions investigated. Although the experimental results presented here were obtained
from a sequence of batch reforming-batch regeneration, it is
believed that comparison can be made to coupling of two
bubbling uidized beds. The very uniform temperature within
the bed, with maximum axial differences in temperature of
only 34 C, conrms the good temperature control offered by
bubbling uidized beds. At the given reaction conditions the

References
Abanades, J.C., Alvarez, D., 2003. Conversion limits in the reaction of CO2
with lime. Energy & Fuels 17, 308315.
Abanades, J.C., Anthony, E.J., Lu, D.Y., Salvador, C., Alvarez, D., 2004a.
Capture of CO2 from combustion gases in a uidized bed of CaO. A.I.Ch.E.
Journal 50, 16141622.
Abanades, J.C., Rubin, E.S., Anthony, E.J., 2004b. Sorbent cost and
performance in CO2 capture systems. Industrial and Engineering Chemistry
Research 43, 34623466.
Adris, A.M., Elnashaie, S.S.E.H., Hughes, R., 1991. A uidized bed
membrane reactor for the steam reforming of methane. Canaidan Journal
of Chemical Engineering 69, 10611070.
Adris, A.M., Lim, C.J., Grace, J.R., 1994. The uidized bed membrane reactor
system: a pilot scale experimental study. Chemical Engineering Science
49, 58335843.
Adris, A.M., Pruden, B.B., Lim, C.J., Grace, J.R., 1996. On the reported
attempts to radically improve the performance of the steam reforming
reactor. Canadian Journal of Chemical Engineering 74, 177186.
Balasubramanian, B., Ortiz, A.L., Kaytakoglu, S., Harrison, D.P., 1999.
Hydrogen from methane in a single-step process. Chemical Engineering
Science 54, 35433552.
Brun-Tsekhovoi, A.R., Zadorin, A.N., Katsobashvili, Ya.R., Kourdyumov,
S.S., 1988. The process of catalytic steam reforming of hydrocarbons in
the presence of carbon dioxide acceptor, hydrogen energy progress, VII.
Proceedings of the World Hydrogen Energy Conference. Pergamon Press,
New York, pp. 885900.
Chen, Z., Yan, Y., Elnashaie, S.S.E.H., 2003. Novel circulating fast uidizedbed membrane reformer for efcient production of hydrogen from steam
reforming of methane. Chemical Engineering Science 58, 43354349.
Constantineau, J.P., 2004. Fluidized bed roasting of zinc sulphide concentrate:
factors affecting the particle size distribution. Ph.D. Thesis, University of
British Columbia, Vancouver, Canada.
Dogan, M., Posarac, D., Grace, J.R., Adris, A.M., Lim, C.J., 2003. Modeling
of autothermal steam methane reforming in a uidized bed membrane
reactor. International Journal of Chemical Reactor Engineering 1 A2,
114.
Han, C., Harrison, D.P., 1994. Simultaneous shift reaction and carbon dioxide
separation for the direct production of hydrogen. Chemical Engineering
Science 49, 58755883.
Hufton, J.R., Mayorga, S., Sircar, S., 1999. Sorption-enhanced reaction process
for hydrogen production. A.I.Ch.E. Journal 45, 248256.
Khotomlyanskii, L.N., Brun-Tsekhovoi, A.R., Katsobashvili, Y.R., Petrov,
V.N., Skoblo, A.I., 1970. Investigation of the separation of dissimilar
particles in uidized bed. Chemistry and Technology of Fuels and Oils 6,
1519.
Li, Y., Buchi, S., Grace, J.R., Lim, C.J., SO2 removal and CO2 capture by
limestone resulting from calcination/sulphation/carbonation cycles. 2005,
submitted for publication.

1202

K. Johnsen et al. / Chemical Engineering Science 61 (2006) 1195 1202

Ortiz, A.L., Harrison, D.P., 2001. Hydrogen production using sorptionenhanced reaction. Industrial and Engineering Chemistry Research 40,
51025109.
Rowe, P.N., Nienow, A.W., 1976. Particle mixing and segregation in gas
uidized beds: a review. Powder Technology 15, 141147.
Silaban, A., Harrison, D.P., 1995. High temperature capture of carbon dioxide:
characteristics of the reversible reaction between CaO(s) and CO2 (g).
Chemical Engineering Communication 137, 177190.

Silaban, A., Narcida, M., Harrision, D.P., 1996. Characteristics of the


reversible reaction between CO2 (g) and calcined dolomite. Chemical
Engineering Communications 146, 149162.
Trimm, D.L., 1997. Coke formation and minimization during steam reforming
reactions. Catalysis Today 37 (3), 233238.

Вам также может понравиться