You are on page 1of 9

5180 Ind. Eng. Chem. Res.

1997, 36, 5180-5188

Kinetic and Reaction Engineering Studies of Dry Reforming of


Methane over a Ni/La/Al2O3 Catalyst
Unni Olsbye,*,† Thomas Wurzel,‡ and Leslaw Mleczko‡
Sintef Applied Chemistry, P.O. Box 124 Blindern, N-0314 Oslo, Norway, and Lehrstuhl für Technische
Chemie, Ruhr-Universität Bochum, D-44780 Bochum, Germany,

Kinetic studies of CO2 reforming of methane over a highly active Ni/La/R-Al2O3 catalyst were
performed in an atmospheric microcatalytic fixed-bed reactor. The reaction temperature was
varied between 700 and 900 °C, while partial pressures of CO2 and CH4 ranged from 16 to 40
kPa. From these measurements kinetic parameters were determined; the activation energy
amounted to 90 kJ/mol. The rate of CO2 reforming was described by applying a Langmuir-
Hinshelwood rate equation. The developed kinetics was interpreted with a two-phase model of
a fluidized bed. The predictions for a bubbling-bed reactor operated with an undiluted feed
(CH4:CO2 ) 1:1) at 800 °C showed that, on an industrial scale, significantly longer contact times
(Hmf ) 7.8 m, mcat/VSTP ) 31.8 g‚s‚mL-1) are necessary for achieving thermodynamic equilibrium
(XCH4 ) 88.2%, XCO2 ) 93.6%). The performance of the reactor was strongly influenced by the
interphase gas exchange: the highest space time yields were obtained for small particles (dp )
80 µm).

1. Introduction scale have confirmed the superior performance of this


reactor type with respect to syngas yields and less
In the last decade, carbon dioxide reforming of pronounced catalyst deactivation compared to fixed-bed
methane (eq 1) has attained wide research interest. The reactors. With respect to the selection of a suitable
reaction can be applied for producing syngas with a low catalyst, supported noble metals and Ni/R-Al2O3 were
H2:CO ratio suitable for the synthesis of oxygenated proposed. For application in the fluidized-bed reactor
chemicals (Edwards, 1995). The reaction also serves for not only catalytic performance but also mechanical
the production of high-purity CO (Teuner, 1985; Kurz stability play an important role. It has previously been
and Teuner, 1990). Moreover, two molecules which shown that La/R-Al2O3 has a superior mechanical
contribute significantly to the greenhouse effect are strength compared to pure R-Al2O3, which makes it
converted into value-added products (Rostrup-Nielsen, particularly suited for fluidized-bed applications (Blom
1994). Finally, the reaction has already been studied et al., 1994). It has further been shown that the Ni/
as a storage for solar energy (Wang et al., 1996; Gadalla La/R-Al2O3 catalyst used in the present work has a
and Sommer, 1989). However, until now the title superior lifetime stability compared to Ni/R-Al2O3 when
reaction had no commercial application by itself (Ed- used in the CO2 reforming of methane in a fluidized-
wards and Maitra, 1994), but it is used downstream of bed reactor (Blom et al., 1994; Slagtern et al., 1997).
the steam-reforming process to reduce the H2:CO ratio
(Teuner, 1987). The work reported in this paper is aimed at the
elucidation of the potential of an industrial fluidized-
CH4 + CO2 ) 2CO + 2H2 bed reactor for the CO2-reforming reaction over the Ni/
La/R-Al2O3 catalyst. As a basis for the simulations,
∆RH°298 K ) +247.9 kJ/mol (1) kinetic measurements were performed in a microcata-
lytic fixed-bed reactor. The reaction engineering kinet-
Syngas yields near the thermodynamic equilibrium in ics that was developed from this study was combined
laboratory-scale fixed-bed reactors (e.g., Vernon et al., with a model of a fluidized bed. By means of reaction
1991; Richardson and Paripatyadar, 1990) were achieved engineering simulations, the influence of the hydrody-
by applying supported metal catalysts (for a review see namic conditions on the reactor performance was ana-
Wang et al., 1996). The main obstacle with respect to lyzed and main design parameters of the reactor were
the commercialization of the process is given by the determined.
severe catalyst deactivation due to carbon deposition
during reaction. This limits an application of fixed-bed
reactors for this reaction (Rostrup-Nielsen and Bak 2. Experimental Procedures and Modeling
Hansen, 1993; Seshan et al., 1994; Swaan et al., 1994; 2.1. Experimental Procedures. Catalyst Prepa-
Au et al., 1994). ration. Preparation of 0.15 wt % Ni/1.7 wt % La/Al2O3
Against this background fluidized-bed reactors have was performed by modifying θ-Al2O3 (Condea, Scc-a5/
been proposed for performing CO2 reforming of methane 90, 45-90 µm) with lanthanum nitrate (Alfa, ultrapure)
and studied (Blom et al., 1994; Mleczko et al., 1997b). by using the incipient wetness technique. The carrier
The experimental results obtained in the laboratory was heated to 1350 °C (2 °C/min) in dry air and kept at
this temperature for 12 h. The resulting carrier had a
* Author to whom correspondence is addressed. Telephone: BET surface area of 2.5 m2/g. Ni(NO3)2‚6H2O was
+47 22 06 79 01. Fax: +47 22 06 73 50. E-mail: Unni.Olsbye@ impregnated, also by using the incipient wetness tech-
chem.sintef.no.
† Sintef Applied Chemistry. nique. The catalyst was calcined at 900 °C in air for
‡ Ruhr-Universität Bochum. Telephone: +49 234 700 4102. 12 h.
Fax: +49 234 709 4115. E-mail: L.Mleczko@risc.techem.ruhr- Microcatalytic Reactor and Experimental Pro-
uni-bochum.de. cedure. Kinetic experiments were carried out at
S0888-5885(97)00246-7 CCC: $14.00 © 1997 American Chemical Society
Ind. Eng. Chem. Res., Vol. 36, No. 12, 1997 5181
Table 1. Range of Feed Gas Compositions Applied in the including the influence of backreactions (see the Results
Kinetic Experiments and Discussion section) and of the water-gas shift (wgs)
pCH4, kPa pCO2, kPa pCO, kPa pH2, kPa pHe, kPa reaction:
standard 40 40 0 0 20 CO + H2O ) CO2 + H2 (3)
dp,CH4 16-40 40 0 0 44-20
dp,CO2 40 16-40 0 0 44-20
dp,CO 40 40 0-5 0 20-15 For this reaction, the following rate expression was
dp,H2 40 40 0 0-5 20-15 used:

atmospheric pressure in a tulip-shaped fixed-bed quartz k


rwgs ) kpH2OpCO - p p (4)
reactor with inner diameter 8 mm, using the 0.15% Ni/ Keq CO2 H2
1.7% La/Al2O3 catalyst (0.05 g, 45-90 µm) diluted with
quartz chips (0.9 g, <35 mesh). The temperature in the The simulation was performed using STELLA software
catalyst bed was measured by a thermocouple inside an (Pytte and Doyle, 1984) and using Euler’s method of
axially installed quartz well with outer diameter 3 mm. integration with a 0.05 ms integration step time.
The temperature gradient in the catalyst bed was <5 Thermodynamical constants were taken from Tsang and
°C. The catalyst was reduced (5% H2/He, 970 °C, 2 h) Hampson (1986). The theoretical influence of mass- and
and aged under standard conditions (800 °C, 250 mL/ heat-transfer limitations was calculated using standard
min, CH4:CO2:He ratio ) 2:2:1) for 1000 min before the methods described in, e.g., Carberry (1976). The cal-
kinetic measurements were started. The range of culations showed that internal and external gradients
partial pressures of feed gas components applied in the were negligible.
kinetic measurements is summarized in Table 1. On- For quantifying the deactivation rate, a first-order
line gas analysis was performed using a Hewlett- expression (eq 5) was used. In order to account for the
Packard (HP) 5890 gas chromatograph (GC) and/or a effect of hydrodynamic conditions on deactivation rates,
quadrupole mass spectrometer (Fisons Sensorlab). Less deactivation constants were determined for both reac-
than 0.3% CO yield was obtained at 800 °C (125 mL/ tors (e.g., fixed and fluidized bed). Upon parameter
min, CH4:CO2:He ratio ) 2:2:1) when the catalyst was variation, deactivation constants were determined for
replaced by pure La/Al2O3 support, indicating negligible each set of parameters and the results corrected accord-
catalytic activity of this material. ingly. All data were then multiplied by a factor of f )
Determination of Catalyst Deactivation. Cata- r0/ri, where ri is the activity measured under the
lyst deactivation was taken into account by switching previous standard conditions, while r0 is the activity
the conditions to standard conditions (800 °C, 250 mL/ measured under the first standard conditions, at the
min, CH4:CO2:He ) 2:2:1) between every 1-4 conditions beginning of the experiment.
switch (i.e., every 1-4 h). The catalyst was left under
standard conditions overnight between the variation of -dr/dt ) kdeacttp,T)const. (5)
each parameter.
Laboratory-Scale Fluidized Bed. The catalytic Model of the Catalytic Fluidized Bed. Hydrody-
activity of 0.15% Ni/1.7% La/Al2O3 for CO2 reforming namics of the fluidized bed was described by applying
of methane in a fluidized-bed reactor was tested in a a modified version of the bubble assemblage model
reactor with inner diameter 23 mm. Gas was fed (BAM). The model, which was originally developed by
through a porous quartz plate at the bottom of the Kato and Wen (1969), is based on the two-phase theory
reactor. Gas was sampled at different heights above of fluidization (Davidson and Harrison, 1963). Corre-
the sinter using a vacuum source connected to an on- sponding to this theory, the bed is divided into the
line HP 5890 GC. A detailed description of the reactor emulsion phase that is in the state of incipient fluidiza-
setup is given elsewhere (Olsbye et al., 1994). The feed tion and particle free bubbles. This model assumes that
was CH4:CO2:N2 ) 2:2:1 (300 mL/min (NTP)), with 1 gas flows through the emulsion phase with minimum
atm of total pressure and 17 mL of catalyst. The fluidization velocity (umf). The remaining gas (u - umf)
catalyst was reduced (5% H2/N2, 1000 °C, 2 h) before flows as spherical bubbles which are formed directly
catalytic testing at 800 °C. above the distributor. Bubbles are assumed to grow due
2.2. Kinetic and Reaction Engineering Model- to coalescence with increasing distance from the gas
ing. Model of the Fixed-Bed Reactor and Proce- distributor. Their diameter and velocity were calculated
dure for Parameter Estimation. The reaction rate according to Werther (1992). The fluidized bed is
data obtained at 800 °C were fitted to a Langmuir- divided into a series of axial segments. The height of
Hinshelwood type rate equation by using a least-squares each segment corresponds to the local bubble diameter.
fitting procedure. The kinetic parameters correspond- Interphase gas exchange was calculated using the
ing to each variable were first estimated using an correlation of Kobayashi et al. (1967). For the gas flow
expression of the type in the bubble formation zone, plug flow was assumed.
The height of the bubble formation zone amounted to
kPi twice the initial bubble diameter (Yates, 1983). In order
r) (2) to account for the increase in the total number of moles
1 + KPi that occurs during reaction, the excess of gas (ue - umf)
due to reaction in the emulsion phase was transported
Then, the full data set (i.e., 42 data points) was fitted into the bubble phase by means of a convective stream.
to the final rate expression. The choice of a final rate A detailed description of the hydrodynamic model
expression is discussed in the Results and Discussion applied is given elsewhere (Mleczko et al., 1996).
section. Due to the differential conditions in the reactor
(i.e., <3% CH4 conversion), inlet partial pressures were 3. Results and Discussion
used for parameter estimation. The rate expression was
validated by a simulation of the full reaction scheme 3.1. Reaction Kinetics. Elucidation of the Re-
by applying a pseudohomogeneous plug flow model and action Rate Equation. The rate of CO production
5182 Ind. Eng. Chem. Res., Vol. 36, No. 12, 1997

Figure 2. Catalyst activity change observed during partial


pressure variations.

Ni/Mg(Al)O catalyst. Preliminary test results, compar-


ing the methane reforming and water-gas shift reaction
activity of the present catalyst at 800 °C after industrial
aging treatment, indicate that this is true also for the
present catalyst. On the basis of these observations,
we concluded that the CO formation rate would be a
good indication of the catalyst’s reforming activity. The
data presented in Figure 1 are corrected for deactivation
(cf. Experimental Procedures section). The observed
catalyst deactivation is exemplified in Figure 2.
The pCH4 and pCO2 data in Figure 1 were fitted to
Langmuir-Hinshelwood expressions. Both data sets
fitted well with such expressions, with the reaction order
in pCO2 approaching zero at high partial pressures (<40
Figure 1. CO production rate (mol/(s‚gcatalyst)) as a function of
reactant (i) and product (j) partial pressures over 0.15% Ni/1.7%
kPa) and the reaction order in CH4 being closer to first
La/Al2O3 in a fixed-bed reactor at 800 °C and 1 atm. (a) Variation order even at the highest partial pressure (40 kPa).
in reactant partial pressure. (b) Variation in product partial These results are in agreement with those of Rostrup-
pressure. Standard partial pressures were pCH4 ) pCO2 ) 40 kPa Nielsen and Bak Hansen (1993), who reported a zero-
and pCO ) pH2 ) 0 kPa. Symbols are experimental results. Lines order dependence in pCO2, as well as a first-order
correspond to the predictions of the kinetic model. dependence in pCH4, in CO2 reforming over Ni/Mg(Al)O
determined experimentally and estimated using the at a CO2:CH4 ratio close to 4 at 500-650 °C.
model presented further as a function of CH4, CO2, H2, Most groups who reported kinetic studies of the CO2
or CO partial pressures over 0.15% Ni/1.7% La/Al2O3 reforming reaction have proposed a Langmuir-Hin-
at 800 °C is plotted in Figure 1. The CO production shelwood type mechanism with an undefined number
rate is a sum of the reforming reaction (eq 1) and the of H atoms in the intermediate CHx species and with
reverse water-gas shift reaction (eq 3). Ideally, the dissociative adsorption of CO2 (Rostrup-Nielsen and Bak
kinetic parameters of the CO2-reforming reaction would Hansen, 1993; Turlier et al., 1993; Osaki et al., 1994).
be based on CH4 conversion data instead of CO forma- The results reported above are consistent with either a
tion rate data. However, the methane conversion was Langmuir-Hinshelwood mechanism, in which the sur-
only 0.8% to 3.0% during these experiments (in order face reaction is the rate-determining step, or an Eley-
to obtain differential conditions), and the mass balances Rideal type mechanism, in which the reaction between
were further complicated by severe deactivation during CHx and CO2,g is rate-limiting at low pCO2 while adsorp-
pCO2 variation, followed by severe reactivation during tion of CH4 is rate-limiting at high pCO2.
pH2 variation (probably due to formation and hydroge- The observed inhibiting effect of CO (Figure 1b) is in
nation of coke). Furthermore, it was observed that the line with observations reported by several groups, i.e.,
H2:CO ratio was in the range 0.9-1.0 under standard that the CO2-reforming reaction is much slower than
test conditions (i.e., CH4:CO2 ratio ) 1). This observa- the H2O-reforming reaction over various supported
tion indicates that the water-gas shift reaction ac- metal catalysts (Rostrup-Nielsen and Bak Hansen, 1993
counted for a maximum of 10% of the observed CO and references therein): CO is present in much higher
formation rate. This is in line with findings of Rostrup- concentrations in the CO2-reforming reaction compared
Nielsen and Bak Hansen (1993). They observed that to the H2O-reforming reaction. Rostrup-Nielsen and
the reverse wgs and the CO2-reforming reaction rate Bak Hansen (1993) proposed that CO adsorbs on the
constants were of the same order of magnitude over a active site, where it dissociates to activated C* and O*
Ind. Eng. Chem. Res., Vol. 36, No. 12, 1997 5183
Table 2. Best-Fit Parameters Applied for the Reactor
Simulations
k K1 K2 K3
unit mol‚s-1‚gcatalyst-1‚atm-2 atm-1 atm-1 atm-1
value 0.00445 0.52 10 27

atoms, and that the further reaction of C* is slow


compared to that of CHx*.
The negative influence of CO addition on the CO
formation rate, combined with the close to zero-order
dependence of pCO2 under standard conditions, may
suggest that CH4 and CO adsorb on the same type of
active site (i.e., Ni), while CO2 adsorbs on a different
site or reacts from the gas phase. Zhang and Verykios
(1996) studied the CO2-reforming reaction over a Ni/
La2O3 catalyst. They proposed that while CH4 is
activated by Ni, CO2 is activated by the La2O3 support Figure 3. CO yield vs contact time over 0.15% Ni/1.7% La/Al2O3
through intermediate La2O2CO3 formation. Such a in a fixed-bed reactor at 700-900 °C. Symbols are experimental
mechanism might also apply for the present Ni/La/Al2O3 results, while lines are simulated results using eqs 4 and 7.
catalyst.
Several groups have observed a negative influence of The curves corresponding to the above expression and
pH2 on the adsorption of light hydrocarbons on metal constants are given together with the experimental
surfaces (Frennet et al., 1978; Martin, 1979). Frennet results in Figure 1.
et al. (1978) proposed two possible adsorption mech- CO yield vs contact time measurements were per-
anisms: either dissociative adsorption (CnH2n+2 + ZS formed at 700-900 °C in order to calculate the activa-
) CnH2n+1* + H*; ZS ) cluster of active sites), giving a tion energy for the CO2-reforming reaction. The data
negative reaction order in pH2, or reactive chemisorption are given in Figure 3 together with simulated CO yield
(CnH2n+2 + H* + ZS ) CnH2n+1 + H2), giving a positive vs contact time data obtained using rate expressions (4)
order in pH2 at low pH2 and a negative order in pH2 at and (7). For the results obtained at 700 and 900 °C,
high pH2. The near-zero-order dependence in pH2 ob- the k value was fitted to the results, while the Ki values
served in our work may mean that adsorbed H atoms were kept unchanged. Figure 3 shows that the proposed
are easily displaced from their adsorption sites by rate expression is in fair agreement with the experi-
collision with adsorbing CH4 or that the residence time mental results. The data in Figure 3 are not corrected
of adsorbed H on one specific site is short compared to for deactivation. A significant, linear increase in CO
the residence time of adsorbed CHx (Frennet et al., yield with time on stream was observed for T ) 900 °C,
1978). This possibility is plausible, taking into account whereas for T ) 700 °C, a slight, linear decrease in CO
that the potential energy variation for H displacement yield with time on stream was observed. When correct-
on a Ni surface is about 1/10 of the heat of adsorption of ing for the observed activity changes with time on
H2 on Ni (Christmann, 1988). Rostrup-Nielsen and Bak stream, an activation energy Ea ) 90 ( 4 kJ/mol was
Hansen (1993) reported an insignificant influence of H2 calculated for the temperature range studied.
addition (CH4:CO2:H2 ) 4:16:1.6) over Ni/Mg(Al)O at Rostrup-Nielsen and Bak Hansen (1993) studied the
500-650 °C, in line with our results. activation energy of CO2 reforming over a Ni/Mg(Al)O
Based on the discussion above, the following expres- catalyst in the temperature range 500-650 °C. They
sion was used to describe the CO2-reforming reaction reported a variation in the activation energy ranging
over 0.15% Ni/1.7% La/Al2O3: from 106 to 79 kJ/mol with increasing temperature and
ascribed the differences to diffusion restrictions. Rich-
kpCH4pCO2 ardson and Paripatyadar (1990) determined an activa-
rref ) (6) tion energy of 102 kJ/mol over a Rh/Al2O3 catalyst at
(1 + K1pCH4 + K2pCO)(1 + K3pCO2) 600-700 °C. They did not report a diffusion influence.
The higher influence of diffusion restrictions in the work
In the reactor simulations, the reverse reforming reac- of Rostrup-Nielsen and Bak Hansen (1993) may be due
tion was taken into account using the following expres- to a higher Ni loading compared to the present catalyst
sion: (16 wt % vs 0.15 wt %), as well as the rather large
particle diameter used in their work (0.3-0.5 mm vs
rref )

( )
0.045-0.090 mm).
kpCH4pCO2 pCO2pH22 3.2. Model Validation and Catalyst Deactivation
1- in a Fluidized Bed. Model Validation. The CH4
(1 + K1pCH4 + K2pCO)(1 + K3pCO2) KeqpCH4 pCO2 conversion profiles measured after 1000 min on stream
in a laboratory-scale fluidized-bed reactor are presented
(7) in Figure 4. The gas velocity applied corresponds to a
The water-gas shift reaction was also included in the u/umf ratio of 8.5. The kinetic parameters applied in
model reaction scheme, assuming first-order reaction in the simulations were recalculated from the initial
each reactant and the same rate constant as the CO2- activity by applying the deactivation constant deter-
reforming reaction. It was further assumed that the mined in the fluidized bed. The steep increase of
influence of pH2 and pCO does not change at low reactant- conversions in the distributor zone and the fact that
to-product ratios. thermodynamic equilibrium is already achieved on a
Kinetic Parameters. The data were fit to the model height of 2 cm above the distributor confirm the high
by using a least-squares fitting procedure. The best fit activity of the applied Ni/La/Al2O3 catalyst. The high
was obtained with the constants presented in Table 2. catalytic activity is also confirmed by the different
5184 Ind. Eng. Chem. Res., Vol. 36, No. 12, 1997

Table 3. Influence of Time on Stream on the Bed Mass


Required To Achieve the Same Conversions Simulated
for a Catalytic Activity Corresponding to TOS ) 1000
min in the Fluidized-Bed Reactor (T ) 800 °C, u0 ) 0.1
m/s, dp ) 80 µm)
time on stream, min
0 1000 6000 40 000
activity referred to standard 1.16 1 0.521 0.19
conditions (kt/k1000 min)
solid circulation, kg‚min-1 100 000 100 16.7 2.5
bed mass, t 18 20 25 40
fixed-bed reactor than in the fluidized-bed reactor, as
reported for the title reaction (Mleczko et al., 1997b;
Olsbye et al., 1997) and for the catalytic partial oxida-
tion of methane (Mleczko et al., 1996; Santos et al.,
1994). In the fluidized beds coke-containing catalyst
Figure 4. Comparison between experimental and calculated particles are continuously transported to the non-coking
methane conversion in a laboratory-scale fluidized-bed reactor (T regions of the reactor, where coke is gasified.
) 800 °C, pCH4:pCO2:pN2 ) 40:40:20 kPa, V ) 5 mL/s). 3.3. Influence of Reaction Conditions on the
Catalytic Performance of an Industrial Fluidized-
Bed Reactor. Simulations were performed for an
atmospheric-pressure industrial-scale bubbling-bed re-
actor (DT ) 4 m). The reactor was equipped with a
perforated-plate gas distributor (800 openings‚m-2).
Particles with diameter between 80 and 500 µm were
fluidized by applying an inlet superficial gas velocity
ranging from 0.1 to 0.5 m/s. The height of the settled
bed was varied between 1.6 and 7.8 m. The methane
to carbon dioxide ratio in the undiluted feed stream and
the reaction temperature amounted to 1 and 800 °C,
respectively.
Influence of the Catalyst Deactivation on the
Design Parameters of the Industrial Unit. One of
the important advantages of the application of a fluid-
ized bed for dry reforming of methane is given by the
possibility of continuous catalyst regeneration in a
second reactor where carbon deposits are burnt off (see,
e.g., Mleczko et al., 1997a). However, continuous cata-
lyst regeneration results in higher investment and
operation costs. In order to elucidate the influence of
catalyst deactivation on the performance of an industrial-
scale unit, a sensitivity analysis was performed by
Figure 5. Catalyst deactivation observed in fixed- and fluidized- assuming different catalytic activities corresponding to
bed reactors at 800 °C (ptot ) 1 atm, CH4:CO2:He ) 40:40:20 kPa). different times on stream. The results, which are
presented in Table 3, indicate that with increasing
concentrations simulated for the bubble and the emul- residence time of the catalyst in the bed the bed height
sion phase, which indicates that the conversions in the has also to be increased to compensate the decrease of
bed are influenced by the interphase gas exchange the catalytic activity. However, due to the mass limita-
although only small bubbles (DB,max ) 0.5 cm) were tions in the industrial scale (see simulation section),
calculated. When taking into account that the concen- there is only a slight difference between the performance
tration profiles measured by means of an immersed of a fresh catalyst and a catalyst applied for 1000 min,
sampling tube represent mainly the composition of the whereas a significant increase of the bed height is
emulsion phase (see discussion in Mleczko and Wurzel, necessary for a catalyst applied for 40 000 min. In order
1997), a good agreement between the prediction of the to determine the amount of solids regenerated per unit
applied model and the experimental data is achieved, time, perfect mixing of the solids in the fluidized bed
although a slight underprediction of the conversions can was assumed. Taking the residence time distribution
be recognized. of an ideal continuous stirred tank reactor into account,
Catalyst Deactivation in a Fluidized Bed. The the solid circulation rate in order to regenerate the
experimentally observed deactivation of the catalyst complete bed mass after 1000 min on stream amounts
proceeded significantly slower in the fluidized bed than to 100 kg‚min-1, which is much lower compared to the
in the fixed bed (see Figure 5). In a fixed bed the case where the initial catalytic activity should be
present catalyst lost 60% of its initial activity during maintained (100 t‚min-1). Therefore, further simula-
the first 1000 min on stream whereas the similar 0.15% tions were performed assuming a catalytic activity
Ni/1.7% Ln/Al2O3 catalyst (Ln ) rare-earth mixture determined after 1000 min on stream in the fixed-bed
with 66% La) lost only 14% in a fluidized bed (Slagtern reactor (i.e., rate parameters given in Table 2), which
et al., 1997). The corresponding deactivation rates in corresponds to approximately 6000 min on stream in
the fluidized bed and in the fixed bed amounted to 1.8 the fluidized-bed reactor.
× 10-4 and 6.5 × 10-3 min-1, respectively. The differ- Highest Conversions. The highest conversions
ence probably results from more severe coking in the were calculated for particles with diameter dp ) 80 µm
Ind. Eng. Chem. Res., Vol. 36, No. 12, 1997 5185
being fluidized with a gas velocity u0 ) 0.1 m/s when
applying a bed height of 7.8 m. The conversion of
methane and carbon dioxide amounted to XCH4 ) 90.4%
and XCO2 ) 95%. These values correspond to the
thermodynamic equilibrium. The lowest conversion was
calculated for the highest gas velocity (0.5 m/s) by
applying a shallow bed (Hmf ) 1.6 m) containing
particles with a diameter of 200 µm. They amounted
to 34% and 52%, respectively. Since the product dis-
tribution was influenced not only by the dry reforming
but also by the water-gas shift reaction (wgs), the
conversion of carbon dioxide was higher than the one
of methane for all investigated reaction conditions. In
turn, the H2/CO ratio was always below 1.
Concentration Profiles. In order to illustrate the
effect of the bed hydrodynamics on the reaction path-
way, axial concentration profiles of reactants (CO2, CH4)
and products (H2, CO) were calculated for two different
particle diameters. In both simulations the bed height
at minimum fluidization conditions was 7.82 m and the
gas velocity amounted to 0.5 m/s.
Since the gas exchange between bubbles and the
emulsion phase is restricted, the concentration of the
feed gases in the emulsion phase was lower than that
in the bubble phase (see Figures 6a and 7a). In turn,
the concentration of the product gases was higher in
the emulsion phase than in the bubble phase (see
Figures 6b and 7b). Furthermore, the most significant
changes of the gas composition occurred in the bottom
part of the reactor, which is characterized by excellent
mass-transfer conditions. With an increasing distance
from the gas distributor, the slope of the concentration
profiles decreased; concentrations in the emulsion phase
reached those calculated for the reactor exit already at
approximately 2 m above the gas distributor. The
profiles were similar for both investigated particle
diameters (80 and 200 µm diameter). However, for the
larger particles, the exit concentration was achieved in
the emulsion phase at a lower height. In contrast, the
difference between the concentration in the bubble and
emulsion phases was more pronounced for the larger
particles.
The calculated concentration profiles indicate that the Figure 6. Concentration profiles of reactants in an industrial
performance of the industrial-scale fluidized-bed CO2 fluidized-bed reactor (DT ) 4 m, u ) 0.5 m/s, dp ) 80 µm, T ) 800
reformer is more strongly influenced by gas exchange °C, CH4:CO2 ) 1:1): (a) CH4, CO2; (b) CO, H2.
between bubbles and the emulsion phase than the
laboratory-scale reactor. This difference can be ex- The heights of the bed predicted for achieving ther-
plained by different hydrodynamic conditions in indus- modynamic equilibrium are similar to those calculated
trial and laboratory reactors; e.g., industrial reactors for catalytic partial oxidation of methane to synthesis
are mostly operated at significantly higher gas velocities gas over a Ni/Al2O3 catalyst in an atmospheric-pressure
than laboratory ones. The higher gas velocities and fluidized bed (Baerns et al., 1997; Wurzel and Mleczko,
distributors with a low open surface area, e.g., perfo- 1995). They are, however, significantly higher than the
rated plates compared to the porous plate distributors ones simulated for the pressurized fluidized-bed steam
applied in the laboratory units, result in the formation reformer (Elnashaie and Adris, 1989). This difference
of significantly larger bubbles in industrial reactors. may be explained by higher reaction rates due to better
Since the rate of gas exchange between bubbles and the mass-transfer conditions in a pressurized fluidized bed
emulsion phase is inversely proportional to the bubble caused by the formation of smaller bubbles. Although
diameter, the gas bypass through the bubbles effects this case was not simulated, it has to be mentioned that
the conversion in the industrial reactors stronger than higher pressures result in a drop of conversion due to
that in the laboratory reactors. In turn, longer contact thermodynamic reasons.
times are required to achieve the equilibrium conver- Effect of Particle Diameter and Bed Height.
sions in the industrial reactor. In the laboratory-scale Since the effective rate of the dry reforming is strongly
reactor conversions approaching the thermodynamic influenced by the interphase gas exchange, the reactor
equilibrium were achieved by applying a modified performance can be optimized by selecting proper
contact time of 4.7 g‚s‚mL-1, whereas this value in- hydrodynamic conditions in the reactor. Therefore, the
creased up to 31.8 g‚s‚mL-1 for an industrial-scale following simulations aimed at the elucidation of the
reactor operated with the smallest particles and the influence of the particle diameter on the conversion of
lowest gas velocity. methane and carbon dioxide and, in turn, on the H2-
5186 Ind. Eng. Chem. Res., Vol. 36, No. 12, 1997

Figure 8. Dependence on particle diameter (dp) and bed height


at minimum fluidization conditions in an industrial fluidized-bed
reactor (DT ) 4 m, u ) 0.3 m/s, T ) 800 °C, CH4:CO2 ) 1:1): (a)
methane conversion; (b) carbon dioxide conversion.

Figure 7. Concentration profiles in an industrial fluidized-bed


reactor (DT ) 4 m, u ) 0.5 m/s, dp ) 200 µm, T ) 800 °C, CH4:
CO2 ) 1:1): (a) reactants CH4 and CO2; (b) products CO and H2. Figure 9. Dependence of bubble diameter (DB) and gas velocity
(u) on the height above the distributor in an industrial fluidized-
bed reactor for two different particle diameters (dp ) 80 and 200
CO ratio in the product stream. Based on these results, µm, DT ) 4 m, u0 ) 0.5 m/s, T ) 800 °C, CH4:CO2 ) 1:1).
the minimum bed height and the contact time for
achieving conversions near the thermodynamic equi- diameters calculated from the correlation of Werther
librium were determined. (1992) for 80 and 200 µm particles are presented in
The calculated conversions of methane and carbon Figure 9. For small particles bubbles grew from 0.043
dioxide passed in dependence on the particle diameter m up to 0.11 m at a height of 2 m above the distributor.
through a minimum for all gas velocities and all bed However, in the upper part of the bed (h > 2 m) the
heights. For the sake of simplicity, only results obtained bubble diameter remained almost constant. In contrast,
for an inlet gas velocity u0 ) 0.3 m/s are presented (see unlimited bubble growth was calculated for the particles
Figure 8a,b). Under these conditions the minima of with 200 µm diameter; at the end of the bed the
conversions were predicted for a particle diameter diameter of a bubble amounted to almost 0.9 m. The
between 200 and 250 µm. The minima differed slightly influence of the different hydrodynamics when applying
for methane and carbon dioxide conversion. Further- different particle diameters is also given in Table 4.
more, the minimum depended on the bed height and These results illustrate that for the same modified
on the gas velocity. Conversions near the thermody- contact time but varying bed height and gas velocity
namic equilibrium were calculated only for small par- different methane and carbon dioxide conversions are
ticles (dp < 100 µm). simulated due to the different bubble dynamics. On the
Two competitive effects influencing conversion have other hand, with increasing particle diameter the bypass
to be taken into account in order to explain the of gas through the bubbles decreased and the flow rate
minimum of the conversion in the dependence on the of gas through the emulsion phase increased. This
particle diameter. With increasing particle diameter, effect was responsible for the increase of conversion for
the effective bubble diameter increases. The bubble large particles.
Ind. Eng. Chem. Res., Vol. 36, No. 12, 1997 5187
Table 4. Methane and Carbon Dioxide Conversions laboratory scale are necessary to achieve conversions
Achieved When Applying the Same Modified Contact near the thermodynamic equilibrium (XCH4 ) 88.2%,
Time but Different Hydrodynamic Conditions (T ) 800 XCO2 ) 93.6%). The performance of the reactor was
°C, CH4:CO2 ) 1:1)
strongly influenced by interphase gas exchange: the
dp ) 80 µm dp ) 200 µm highest space time yields were obtained for small
mcat., mcat., mcat., mcat., particles (dp ) 80 µm).
u, m‚s-1 t u, m‚s-1 t u, m‚s-1 t u, m‚s-1 t
0.1 20 0.5 100 0.1 20 0.5 100 Nomenclature
dp ) 80 µm dp ) 200 µm DT ) reactor diameter, m
XCH4, XCO2, XCH4, XCO2, XCH4, XCO2, XCH4, XCO2, dp ) particle diameter, µm
% % % % % % % % e ) excess
88.1 93.5 88.1 93.5 64.1 78.7 47.8 66.9 h ) height above the distributor, m
H ) bed height, m
i.d. ) inner diameter, m
Energy Demands. With respect to the large-scale k ) forward rate constant
application of a fluidized-bed reactor, the heat supply K ) equilibrium constant
due to the high endothermicity of the reforming reaction L ) length, m
is of primary importance. In order to check whether m ) mass, g
this demand may be fulfilled by immersion of a tube mf ) minimum fluidization conditions
heat exchanger into the fluidized bed, the main param- o.d. ) outer diameter, m
eters of this module were calculated. The effect of the pi ) partial pressure of component i, kPa
immersed heat exchanger on the bed hydrodynamics rref. ) reaction rate of the reforming reaction, mol/(s‚gcatalyst)
was analyzed for a reactor operated with a gas velocity u ) superficial gas velocity, m/s
of 0.5 m/s and particles having a diameter of 80 µm. Xi ) conversion of component i, %
In order to achieve conversions near the thermody-
namic equilibrium (XCH4 ) 88.2% and XCO2 ) 93.6% vs Literature Cited
XCH4,eq ) 90.4%, XCO2,eq ) 94.9%), a bed height of 7.8 m
was required. To obtain these conversions, 8.15 MW Au, C. T.; Hu, Y. H.; Wan, H. L. Pulse studies of CH4 interaction
with NiO/Al2O3 catalysts. Catal. Lett. 1994, 27, 199.
of heat have to be supplied to the bed by means of Baerns, M.; Buyevskaya, O. V.; Mleczko, L.; Wolf, D. Catalytic
immersed heat-exchanger tubes. This energy demand partial oxidation of methane to synthesis gasscatalysis and
corresponds to a specific duty of 4.26 MW/kg of syngas. reaction engineering. Stud. Surf. Sci. Catal. 1997, 107, 421.
For an overall heat-transfer coefficient of 260 W/(m2‚K) Blom, R.; Dahl, I. M.; Slagtern, A.; Sortland, B.; Spjelkavik, A.;
and a temperature difference between the heating Tangstad, E. Carbon dioxide reforming of methane over lan-
medium and the reactor of 200 K (Kunii and Levenspiel, thanum-modified catalysts in a fluidized-bed reactor. Catal.
Today 1994, 21, 535.
1991), the heat-exchange area amounts to 156 m2. This Carberry, J. J. Chemical and catalytic reaction engineering;
area is provided by immersing 90 tubes (L ) 7 m, o.d. McGraw-Hill: New York, 1976; p 199.
) 0.08 m) into the bed. These parameters are very Christmann, K. Interaction of hydrogen with solid surfaces. Surf.
similar to the configuration presented by Kunii and Sci. Rep. 1988, 9, 1.
Levenspiel (1991) for an industrial fluidized-bed reactor Davidson, J. F.; Harrison, D. Fluidised Particles; Cambridge
for the synthesis of acrylonitrile. However, due to the University Press: Cambridge, U.K., 1963.
Edwards, J. H. Potential sources of CO2 and the options for its
thermal stress, the heat-exchanger tubes made of large-scale utilisation now and in the future. Catal. Today 1995,
expensive high alloy materials are required. Since the 23, 59.
distance between the tubes (“pitch” ) 0.37 m) is more Edwards, J. H.; Maitra, A. M. The reforming of methane with
than 3 times greater than the maximum bubble size carbon dioxidescurrent status and future applications. Stud.
under these conditions, it is assumed that the immer- Surf. Sci. Catal. 1994, 81, 291.
sion of the tubes should not significantly influence bed Elnashaie, S. S. E. H.; Adris, A. M. A Fluidized bed steam reformer
for methane. Fluidization VI; United Engineering Trustees:
hydrodynamics. New York, 1989.
Frennet, A.; Lienard, G.; Crucq, A.; Degols, L. Effect of multiple
4. Conclusions sites and competition in adsorption on the kinetics of reactions
catalyzed by metals. J. Catal. 1978, 53, 150.
In this work kinetic studies of CO2 reforming of Gadalla, A. M.; Sommer, M. E. Carbon dioxide reforming of
methane over a highly active Ni/La/R-Al2O3 catalyst in methane on nickel catalysts. Chem. Eng. Sci. 1989, 44, 2825.
a microcatalytic fixed-bed reactor were performed. The Geldart, D. Types of Gas Fluidization. Powder Technol. 1973, 7,
285.
reforming of methane with carbon dioxide was described Kato, K.; Wen, C. Y. Bubble Assemblage Model for Fluidized-Bed
by applying a Langmuir-Hinshelwood type expression. Catalytic Reactors. Chem. Eng. Sci. 1969, 24, 1351.
The activation energy for this reaction was estimated Kobayashi, H.; Arai, F. Determination of Gas Cross-Flow Coef-
to 90 kJ/mol. This value is in the range reported by ficient between the Bubble and Emulsion Phases by Means of
Rostrup-Nielsen and BakHansen (1993) but lower than Measuring Residence-Time Distribution of Fluid in a Fluidized
the one published by Richardson and Paripatyadar Bed. Kagaku Kogaku 1967, 31, 239.
Kunii, D.; Levenspiel, O. Fluidization Engineering, 2nd ed.;
(1990) for a Rh/Al2O3 catalyst. The kinetics derived Heinemann Press: Oxford, U.K., 1991.
from these experiments were combined with a model of Kurz, G.; Teuner, S. Calcor process for CO production. Erdöl Kohle,
a fluidized-bed reactor based on the two-phase theory Erdgas, Petrochem. 1990, 43, 171.
of fluidization. Subsequently, this reaction engineering Martin, G. The kinetics of the catalytic hydrogenolysis of ethane
model was validated with experimental data obtained over Ni/SiO2. J. Catal. 1979, 60, 345.
in a laboratory-scale fluidized bed. Simulations of a Mleczko, L.; Wurzel, T. Experimental Studies of Catalytic Partial
Oxidation of Methane to Synthesis Gas in a Bubbling-Bed
bubbling-bed reactor operated with an undiluted feed Reactor. Chem. Eng. J. 1997, 66, 193.
(CH4:CO2 ) 1:1) at 800 °C showed that in the industrial Mleczko, L.; Ostrowski, T.; Wurzel, T. A Fluidised-Bed Membrane
scale significantly longer contact times (Hmf ) 7.8 m, Reactor for the Catalytic Partial Oxidation of Methane to
mcat/VSTP ) 31.8 g‚s‚mL-1) compared to those in the Synthesis Gas. Chem. Eng. Sci. 1996, 51, 3187.
5188 Ind. Eng. Chem. Res., Vol. 36, No. 12, 1997

Mleczko, L.; Malcus, S.; Wurzel, T. Catalytic Reformer- Teuner, S. A new process to oxo-feed. Hydrocarbon Process. 1987,
CombustorsA novel reactor concept for synthesis gas produc- 66, 52.
tion. Ind. Eng. Chem. Res. 1997a, in press. Tsang, W.; Hampson, R. F. Chemical kinetics database for
Mleczko, L.; Malcus, S.; Wurzel, T. Reaction engineering investi- combustion chemistry. Part 1. Methane and related compounds.
gations of CO2 reforming in a fluidized bed reactor. 1997b, to J. Phys. Chem. Ref. Data 1986, 15 (3), 1087.
be submitted. Turlier, P.; Pereira, E. B.; Martin, G. Proc. Int. Conf. CO2 Reform.
Olsbye, U.; Tangstad, E.; Dahl, I. M. Partial oxidation of methane 1993, 119.
to synthesis gas in a fluidized bed reactor. Stud. Surf. Sci. Catal.
1994, 81, 303. Vernon, P. D. F.; Green, M. L. H.; Cheetham, A. K.; Ashcroft, T.
Olsbye, U.; Dahl, I. M.; Slagtern, A° .; Blom, R. Catalytic partial Partial Oxidation of Methane to Synthesis Gas, and Carbon
oxidation of methane to synthesis gas in a fluidized bed reactor. Dioxide as an Oxidising Agent for Methane Conversion. Catal.
Proceedings of the ECCE-1 Conference, Florence, Italy, May Today 1991, 13, 417.
4-7, 1997; AIDIC: 1997; Vol. 1, p 367. Wang, S.; Lu, G. Q. (Max); Millar, G. J. Carbon dioxide reforming
Osaki, T.; Masuda, H.; Mori, T. Intermediate hydrocarbon species of methane to produce synthesis gas over metal-supported
for the CO2-CH4 reaction on supported Ni catalysts Catal. Lett. catalysts: state of the art. Energy Fuels 1996, 10, 896.
1994, 29, 33. Werther, J. Scale-up Modelling for Fluidized Bed Reactors. Chem.
Pytte, A.; Doyle, J. STELLA/For the Macintosh Computer, 1984. Eng. Sci. 1992, 47, 2457.
Richardson, J. T.; Paripatyadar, S. A. Carbon dioxide reforming Wurzel, T.; Mleczko, L. Reaction engineering investigations of
of methane with supported rhodium. Appl. Catal. 1990, 61, 293. catalytic partial oxidation of methane to synthesis gas in a
Rostrup-Nielsen, J. R. Aspects of CO2-reforming of methane. Stud. bubbling-reactor (German). Dechema Jahrestagungen, Wies-
Surf. Sci. Catal. 1994, 81, 25. baden, Germany, May 30-June 1, 1995; p 62.
Rostrup-Nielsen, J. R.; Bak Hansen, J. H. CO2-Reforming of Yates, J. Fundamentals of Fluidized-Bed Chemical Processes;
methane over transition metals. J. Catal. 1993, 144, 38. Butterworths: London, 1983.
Santos, A.; Menéndez, M.; Santamaria, J. Partial oxidation of
methane to carbon monoxide and hydrogen in a fluidized bed Zhang, Z. L.; Verykios, X. E. Unique performance of Ni/La2O3
reactor. Catal. Today 1994, 21, 481. catalyst in carbon dioxide reforming of methane to synthesis
Seshan, K.; ten Barge, H. W.; Hally, W.; van Keulen, A. N. J.; gas. Appl. Catal. A 1996, 138, 109.
Ross, J. R. H. Carbon dioxide reforming of methane in the
presence of nickel and platinum catalysts supported on ZrO2. Received for review March 27, 1997
Stud. Surf. Sci. Catal. 1994, 81, 285. Revised manuscript received August 4, 1997
Slagtern, A° .; Olsbye, U.; Dahl, I. M.; Blom, R.; Fjellvåg, H. Accepted August 12, 1997X
Characterisation of Ni on La modified Al2O3 during CO2
reforming of methane. Appl. Catal. A 1997, in press. IE970246L
Swaan, H. M.; Kroll, V. C. H.; Martin, G. A.; Mirodatos, C.
Deactivation of supported nickel catalysts during the reforming
of methane by carbon dioxide. Catal. Today 1994, 21, 571.
X Abstract published in Advance ACS Abstracts, October 15,
Teuner, S. Make CO from CO2. Hydrocarbon Process. 1985, 64,
106. 1997.

You might also like