Вы находитесь на странице: 1из 9

Chemical Engineering Science 56 (2001) 641}649

Heterogeneous reactor modeling for simulation of catalytic oxidation


and steam reforming of methane
A. K. Avcm , D. L. Trimm, Z. I lsen OG nsan *
Department of Chemical Engineering, Bog\ azic7 i University, 80815 Bebek, Istanbul, Turkey
School of Chemical Engineering and Industrial Chemistry, University of New South Wales, Sydney 2052, Australia

Abstract
An autothermal, dual catalyst, "xed-bed reaction system proposed for hydrogen production from methane is mathematically
investigated using di!erent catalyst bed con"gurations and feed ratios. Consecutive placement or physical mixture of the oxidation
and reforming catalysts, Pt/d}Al O and Ni/MgO}Al O , respectively, are the two con"gurations of interest. Reactor operation at
 
 
di!erent feed ratios is analyzed for both catalyst bed con"gurations on laboratory scale and industrial scale via a series of simulations
by using one-dimensional heterogeneous "xed-bed reactor model. The type of heterogeneous components implemented into the
model is decided by checking related criteria. Hydrogen production is predicted to be higher when the catalysts are in a physically
mixed state as well as at low methane-to-oxygen and high steam-to-methane ratios, which are in agreement with the experimental
results reported for a bench scale integral reactor. The optimum operating conditions for obtaining maximum hydrogen production
are also investigated.  2001 Elsevier Science Ltd. All rights reserved.
Keywords: Autothermal operation; Dual catalyst; Fixed-bed reactor; Methane oxidation; Methane steam reforming; Reactor simulation

1. Introduction
In recent years, hydrogen has become a widely used
feedstock in the chemical, petroleum re"ning and petrochemical industries. Hydrotreating and hydrocracking
processes, synthesis gas applications such as production
of ammonia and methanol, Fischer}Tropsch synthesis
and the manufacture of chemicals having speci"c end
uses such as pharmaceuticals are areas in which hydrogen is employed (Furimsky, 1998; Pena, Gomez &
Fierro, 1996). In addition, hydrogen is expected to be an
energy source in the future since it o!ers several environmental and economic advantages when compared to
other fuels (Jamal & Wyszynski, 1994; Rosen, 1991).
Steam reforming of light hydrocarbons (C }C ) run
 
on nickel catalysts is the most widely employed route for
hydrogen production due to its simple construction, operation and well-established technology (Armor, 1999;
Pena et al., 1996; Rosen, 1991). On the other hand, the
demand for considerable energy input resulting from the

* Corresponding author. Tel.: #90-212-263-15-40; fax: #90-2122872460.


E-mail address: onsan@boun.edu.tr (Z. I lsen OG nsan).

high endothermicity of the reforming reactions and the


existence of catalyst deactivation turn out to be the major
drawbacks of catalytic steam reforming (Rostrup-Nielsen, 1984; Trimm, 1999).
Catalytic partial oxidation of methane, which is mildly
exothermic (*H"!35.7 kJ/mol), gives a lower synthesis gas ratio (H :CO"2:1) (Hickman & Schmidt,

1993). Therefore, it is more suitable for use in processes
such as Fischer}Tropsch synthesis and methanol production. Direct methane cracking, carbon dioxide reforming
of methane, use of membrane reactors and of "xed-bed
reactors with reversed #ow in steam reforming, thermochemical and photocatalytic water splitting are some
of the alternative techniques for hydrogen production.
In a recently proposed reaction system using integral
microreactors, the endothermic heat and part of the
steam required for methane steam reforming are balanced by exothermic methane oxidation in the presence
of two di!erent catalysts, Pt/d}Al O for total oxidation
 
and Ni/MgO}Al O for steam reforming (Ma & Trimm,
 
1996; Ma, Trimm & Jiang, 1996). Increased hydrogen
selectivities can thus be attained in autothermal operation. Moreover, the presence of two di!erent catalysts
introduces operational #exibility, i.e. the reaction system
can be adjusted in an object-oriented manner. For

0009-2509/01/$ - see front matter  2001 Elsevier Science Ltd. All rights reserved.
PII: S 0 0 0 9 - 2 5 0 9 ( 0 0 ) 0 0 2 7 1 - 2

642

A. K. Avcn et al. / Chemical Engineering Science 56 (2001) 641}649

instance, the synthesis gas ratio may be adjusted for


a speci"c end use such as Fischer}Tropsch synthesis.
This work is a computational investigation of the
e!ects of catalyst bed con"guration and molar feed ratios
on the product distributions achieved in the dual-catalyst
autothermal operation mentioned above. For this purpose, a series of reactor simulations are performed for
bench scale and industrial scale reactors by using onedimensional heterogeneous reactor model. The operating
conditions that will lead to maximum hydrogen yields
are also evaluated.

Note that two alternative rate expressions (Eqs. (2*) and


(2)) are given for reaction (2).
Water}gas shift and methane cracking are considered
as side reactions running on the Ni/MgO}Al O cata 
lyst:
CO#H O&CO #H , *HM"!41.2 kJ/mol, (4)



CH &C#2H , *HM"74.9 kJ/mol.
(5)


The corresponding rate expressions for reactions (4) and
(5) are (Kuvshinov, Mogilnykh & Kuvshinov, 1998; Xu
& Froment, 1989a):
!r "


2. Description of the reaction system


The reaction system is an adiabatic "xed-bed reactor
in which Pt and Ni catalysts are packed in several con"gurations. During autothermal operation, oxidation,
reforming and various side reactions exist within the
reactor. Total oxidation of methane in the presence of air
is catalyzed by the Pt/d}Al O catalyst and provides the
 
exothermic heat to the system:
CH #2O PCO #2H O, *HM"!802.3 kJ/mol.




(1)
The rate expression for reaction (1) is given by Ma et al.
(1996):
k K P  (K P 
  !&
 !r "
.
(1H)
 (1#K P #(K P )
 !&
 -
A large fraction of the endothermic heat is consumed by
the steam reforming reactions:
CH #H O&CO#3H , *HM"206.2 kJ/mol,
(2)



CH #2H O&CO #4H , *HM"165.0 kJ/mol. (3)




The rate expressions for the steam reforming reactions
that run on Ni/MgO}Al O (Eq. (2*)) and Ni/MgAl O
 
 
(Eq. (2)) are given as follows (Ma, 1995; Xu & Froment,
1989a):
k K K P P 
 
!& & , (2H)
!r "
 P(1#K (P /P)#K (P /P ))
&
 !& &
& - &
!r "


k (P P  !(P  P  /K ))
 !- & !- &

,
P  (1#K P #K  P  #K  P  #K  (P  /P  ))
!- !& &
!& !&
&- &- &
&

(4H)

k (P !(P  /K ))
&
 .
(5H)
!r "  !&

(1#K P 
)L
& &
The temperature dependence of the LHHW-type rate law
parameters in these rate expressions have been taken
from the references cited.
3. Simulation of autothermal operation in 5xed-bed
reactors
Various con"gurations of the two catalysts,
Pt/d}Al O and Ni/MgO}Al O , present within the
 
 
reactor are possible. Experimental data reported show
that a change in the catalyst bed arrangement results in
di!erent product distributions (Ma & Trimm, 1996). Two
di!erent bed arrangements are studied in this work. The
con"gurations of interest are the consecutive dual bed
where Pt/d}Al O is placed in the upstream of the reac 
tor followed by Ni/MgO}Al O , and the mixed bed
 
which involves a physical mixture of the two catalysts. In
the simulation of autothermal operation in dual-bed
scheme, steam reforming and other side reactions are
assumed to be consecutive to methane oxidation while
the simultaneous occurrence of reactions (1}5) are considered in the mixed-bed scheme.
3.1. Reaction model

k (P P  !(P P  /K ))
 !& & 
!- &
,
P
(1#K P #K  P #K  P #K  (P  /P  ))
&
!- !& &
!& !&
&- &- &

A one-dimensional heterogeneous reactor model is


used for the simulation of the adiabatic autothermal dual
catalyst operation in "xed-bed reactors. The working
equations are as follows:
Model equations for the bulk yuid:

!r "


dF
H "k a (C !C ),
QH K QH
H
d=

(6)

d
h a ( !)
" Q K Q
,
d=
Fc
H H NH

(7)

(2)

k (P  P  !(P P  /K ))
 !& & !- &

.
P
(1#K P #K  P  #K  P #K  (P  /P  ))
&
!- !& &
!& !&
&- &- &

(3H)

A. K. Avcn et al. / Chemical Engineering Science 56 (2001) 641}649

dP
b
P F
M
M
2.
"!
d=
A (1!
)o P F M
M 2
A
A
Model equations for the solid:
k a (C !C )" g (l )(!r ) ,
QH K QH
H
G GH
GH Q
G
h a ( !)" g (!*H )(!r ) .
Q K Q
G
G
GQ
G
Boundary conditions:

(8)

(9)
(10)

At ="0, F "F , " , P"P .


(11)
H
HM
M
M
The particle-to-#uid mass and heat transfer coe$cients
in Eqs. (9) and (10), k and h , respectively, are evaluated
QH
Q
by using the following correlations (Wakao & Kaguei,
1982):
hD
Q N "2#1.1 Pr Re ,
(12)
j
D
k D
QH N "2#1.1 Sc Re .
(13)
D
HK
The temperature dependencies of physical properties
such as viscosity, thermal conductivity, binary and e!ective di!usivities of the species present in the reaction
mixture and their mixing rules have been reported
(Wakao & Kaguei, 1982; Rase, 1990). The ideal gas
approximation is employed in the evaluation of the density of the gas mixture.
In order to reduce the execution time in the reactor
simulations, heterogeneous terms, i.e. interfacial heat and
mass transfer resistances and the intraparticle di!usion
limitations, are not incorporated into the reactor model if
the criteria given below indicating the degree of heterogeneity are satis"ed (Fogler, 1999):
Mears' criteria for interfacial mass transfer:
(l )(r )o D n
G GH GH @ N (0.15,
2k C
QH H
(!*H )(!r )o D E
G
G
G A N  (0.15.
2h R
Q
E
Weisz}Prater criterion for intraparticle diwusion:

(14)
(15)

(l )(r )o D
G GH GH A N ;1.
(16)
4D C
C QH
These criteria are the versions modi"ed for the multiple
reaction case and give approximate indications of the
pertinent heterogeneous phenomenon (Rase, 1990).
3.2. Evaluation of the optimum operating conditions
The optimum operating conditions leading to maximum hydrogen production are determined via the

643

simultaneous treatment of the energy balance and reaction equilibrium expressions. In formulating the energy
balance, it is assumed that equilibrium values are reached
for all reactions, all the oxygen in the feed is consumed
and a steam-to-methane ratio of 3 exists for the steam
reforming reaction in order to eliminate carbon formation (Ma, 1995). Heat generated by total methane oxidation (q ) is utilized for vaporizing and heating the water

fed to the system (q ), heating the catalyst bed (q ) and


the gas mixture that exists after oxidation (q ) as well as

supplying energy to reactions (2}5) (q ) in an adiabatic

system:
q "q #q #q #q .






(17)

The explicit forms of the terms in the energy balance,


q }q , are as follows:
 
q "F  *H
,

!& \ 
 

(18)

q "F  q  ,

& - M & -

(19)

q "=



2


c
d,
N   

(20)

2
q " (F
)
c d,

H> 
NH

H

(21)


q " F  *H #F
*H .

!& \ G
G
!-\ 

G

(22)

Eq. (17) and the reaction equilibrium expressions are


solved together to obtain the optimum conditions with
maximum hydrogen yield as the system objective.
3.3. Numerical techniques
The temperature dependences of the rate-law parameters in Eq. (1*) are determined by minimizing the sum
of squares of di!erences between the reported and
calculated values. This minimization problem is handled
by using the Nelder}Mead simplex algorithm.
Eqs. (6)}(10) with the boundary conditions (11) are
simultaneously solved in MATLAB2+. The nonlinear
algebraic Eqs. (9) and (10) are solved in a separate subroutine by least-squares technique where the
Gauss}Newton method is employed together with
a mixed quadratic and cubic line search procedure. In the
solution of the di!erential Eqs. (6)}(8), combination of
a variable order sti! and a non-sti! ODE solver is found
to give the best performance in terms of yielding the
lowest CPU time. Optimum operating conditions are
evaluated by employing the same algebraic solver routine
that is used in the solution of Eqs. (9) and (10). Reactor
simulations are conducted on an IBM Net"nity 7000
M10 workstation.

640}665
800
2.5
2.9
2.2;10\
1.7;10\
2;10\(Pt/d}Al O )#3;10\(Ni/MgO}Al O ) 5;10\(Pt/d}Al O #Ni/MgO}Al O )
 
 
 
 
1.3;10\
3.4;10\

Dual-bed scheme

Bench-scale reactor (Ma & Trimm, 1996)

Table 1
Operating conditions and reactor data

Comparison of simulation outputs and their experimental counterparts for dual- and mixed-bed schemes
are given in Tables 2 and 3, respectively.
When the product distributions obtained in the experiments and simulations are compared with each other, it
is observed that the mixed-bed scheme gives higher hydrogen yields. Presence of the catalysts in the physically
mixed state enhances heat and mass transfer characteristics, i.e. reduces the degree of heterogeneity of the reaction system, which is also con"rmed by the relatively low
values of the LHS of the criteria given by Eqs. (14)}(16) in
Section 3.1.
The status of the criteria given in Section 3.1 is taken as
a measure in the implementation of the heterogeneous
components into the reactor model. In both catalyst bed
con"gurations, it is observed that the interfacial heat
transfer resistance is signi"cant during the entire operation (LHS (15)'0.15). On the other hand, interfacial
mass transfer resistance is observed to be important at
temperatures higher than ca. 1100 K. Intraparticle di!usion limitations are neglected, i.e. e!ectiveness factor
values are taken as unity due to the presence of small
catalyst particles.
Each simulation of the dual-bed scheme is composed
of two sub-simulations: one for the Pt bed where only
methane oxidation is assumed to occur, the other for the
Ni/MgO bed where reactions (2)}(5) are considered to
proceed. The bed temperature reaches to its maximum
value at the bed interface where endothermic steam
reforming reactions become signi"cant. Experimentally
determined maximum bed temperatures in dual-bed operation are less than 900 K (Table 2). Therefore, the
interfacial mass transfer is neglected and only external
heat transfer is considered as the heterogeneous component in both "rst and second sub-simulations. A single
simulation is conducted for each run in the mixed-bed
con"guration where reactions (1)}(5) are assumed to proceed in a simultaneous fashion. The experimental values
of maximum bed temperatures are less than 1010 K

Mixed-bed scheme

4.1. Bench scale simulations

Dual-bed scheme

Industrial-scale reactor (Xu & Froment, 1989b)

In order to test the reactor model, simulations are


initially conducted at di!erent feed ratios and catalyst
bed con"gurations for bench scale integral reactors and
compared with the experimental results reported by Ma
and Trimm (1996). Autothermal operation is then analyzed under the same conditions in hypothetical industrial reactors whose dimensions are determined by scaleup and by using the industrial "xed-bed reactor data
available (Xu & Froment, 1989b). The ratio of the catalyst weight to the initial methane molar #ow rate is used
as the scale-up parameter. The operating conditions and
reactor data used in the simulations are given in Table 1.

Mixed-bed scheme

4. Results and discussion

815
800
28.6
28.6
62.2
44.4
53.2(Pt/d}Al O )#79.8 (Ni/MgAl O )
133 (Pt/d}Al O #Ni/MgAl O )
 
 
 
 
1.016;10\
1.335;10\

A. K. Avcn et al. / Chemical Engineering Science 56 (2001) 641}649

(K)
M
P M (atm)
2
F   (kmol/h)
!&
=
(kg)

d (m)
P
d (m)
N

644

A. K. Avcn et al. / Chemical Engineering Science 56 (2001) 641}649

645

Table 2
Experimental results vs. bench scale simulation outputs in dual-bed scheme
Feed conditions (Ma & Trimm, 1996)

Experimental results (Ma & Trimm, 1996)

Simulation outputs

CH /O
 

H O/CH



GHSV (h\) x 
!&

y  (*)
&

y (*)
!-

(K)


x 
!&

y  (*)
&

y (*)
!-

(K)


3.53
3.53
2.98
2.98
2.51

0.88
1.75
1.75
2.34
1.75

37600
47700
50600
57500
59000

8.8
10.1
17.3
17.3
29.0

0
0
0
0
0

790
781
831
823
873

11.0
13.1
19.3
20.9
29.0

31.6
39.8
52.2
58.0
67.5

0.4
0.4
0.3
0.5
0.7

790
782
827
822
875

16.2
15.0
20.7
19.9
28.7

Table 3
Experimental results vs. bench scale simulation outputs in mixed-bed scheme
Feed conditions (Ma & Trimm, 1996)

Experimental results (Ma & Trimm, 1996)

Simulation outputs

CH /O
 

H O/CH



GHSV (h\) x 
!&

y 
&

y
!-

(K)


x 
!&

&

y
!-

(K)


2.24
1.89
1.89
1.89
1.55
1.55
1.35
1.16

1.17
1.17
1.56
2.34
1.56
2.34
2.34
2.34

37600
41100
44500
51300
49500
56000
60000
65000

47.8
74.9
76.4
82.9
105.7
107.1
119.5
120.6

6.3
11.7
12.5
9.9
15.0
19.6
30.3
35.9

839
888
889
851
931
908
953
1007

43.8
53.9
46.9
31.0
55.4
40.5
48.7
55.0

98.6
125.7
129.3
129.7
151.5
153.6
178.6
195.6

12.3
24.1
23.6
20.0
37.7
31.9
48.7
64.6

843
902
890
869
943
913
966
1009

39.5
53.3
53.8
54.9
69.1
70.1
83.6
91.7

Product yield: moles of product obtained/100 moles of methane fed.

(Table 3). This leads to the implementation of interfacial


heat transfer into the reactor model only. The heterogeneous characteristics of these reaction systems are also
con"rmed by checking the related criteria (Eqs. (14)}(15)).
For both catalyst bed con"gurations, it is observed in
the experiments and simulations that, at constant steamto-methane ratio, a decrease in the methane-to-oxygen
ratio results in higher hydrogen yields and elevated maximum bed temperatures. Since the molar #ow rate of
methane is constant in all runs, any decrease in the
methane to oxygen ratio, i.e. any increase in the molar
#ow rate of oxygen in the feed will lead to the combustion
of more methane which will increase the amount of
energy generated by total oxidation and hence the bed
temperatures. Higher bed temperatures will facilitate
endothermic steam reforming reactions and result in
higher hydrogen yields. In addition, at constant methane-to-oxygen ratio, an increase in the initial molar #ow
rate of steam lowers the maximum bed temperature. It
can be concluded that steam facilitates heat transfer
between the two beds.
In both catalyst bed con"gurations, the maximum bed
temperatures obtained from simulations are close to the
experimental ones, whereas predicted hydrogen and carbon monoxide yields seem to be positively deviating in all
runs (Tables 2 and 3). One possible explanation is the

absence of the following side reaction in the reactor


model that removes hydrogen and carbon monoxide
from the reaction medium:
CO#H &C#H O, *HM"!135.6 kJ/mol.



(23)

The change in the Gibbs free energy of the above reaction


is negative at temperatures less than ca. 950 K, indicating
its signi"cance in existing reaction conditions (Tables
2 and 3). Hence, simulated H and CO yields may

approach experimental ones if this reaction were
implemented into the model equations.
The methane conversion levels are predicted to be
lower than the experimental values at low methane-tooxygen and high steam-to-methane ratios in mixed-bed
con"guration (runs 4}8 in Table 3). It seems that excess
steam inhibits the steam reforming reaction in the simulations, which is not observed in the experiments. This
di!erence may be due to the lower performance of the
kinetic expressions in case of high amount of steam and
oxygen, which appears to be a limitation for this model in
larger scale simulations.
Pressure drop is observed to be negligible in both bed
schemes. Maximum values are obtained to be 0.063 and
0.031 atm for the bench scale dual- and mixed-bed con"gurations, respectively.

646

A. K. Avcn et al. / Chemical Engineering Science 56 (2001) 641}649

4.2. Industrial scale simulations


In order to predict characteristics of dual catalyst
autothermal operation on an industrial scale, a set of
simulations are performed in hypothetical larger scale
tubular reactors with the catalyst bed con"gurations of
interest. Since steam is fed instead of liquid water into the
reactors, the maximum bed temperatures obtained in
industrial scale simulations are higher than ones given in
Tables 2 and 3. As a result, in dual-bed operation, criteria
(14), (15) indicated the existence of interfacial mass and
heat transfer in Pt bed, both of which are considered in
the "rst subsimulation. Criterion (14) indicated the importance of interface mass transfer resistance only at the
bed interfacial, in an incremental layer of steam reforming catalyst. Therefore, interfacial heat transfer is the only
external heterogeneous component considered in secondary subsimulations. Similarly, only external heat transfer
is considered in simulating the industrial scale reactor
with mixed-bed con"guration in which the local existence of external mass transfer is neglected.
Apart from the above external "lm resistances, intraparticle di!usion limitations are taken into account
due to the presence of larger catalyst particles (LHS
(16)'1). For this purpose, e!ectiveness factors for reactions (1)}(5) are taken from the literature (De Groote
& Froment, 1996). Although the results obtained
by using constant e!ectiveness factors are approximate,
operating characteristics of the industrial scale reactor
can be identi"ed due to the similarity of the operating
conditions given.
Molar feed ratios given in Tables 2 and 3 are employed
in the industrial scale simulations. Bed temperature and
product yield trends obtained for di!erent runs are similar to the bench scale simulations. Hydrogen yields are
between 107 and 139 in the dual bed and 122 and 218 in
the mixed-bed con"gurations. The temperature and typical molar #ow rate pro"les for di!erent bed con"gurations are given in Figs. 1}4.
The bed temperatures obtained in dual-bed scheme are
greater than those calculated for mixed-bed con"guration. This is due to the complete consumption of oxygen
in the feed, which allows total oxidation to proceed until
completion and raise temperatures up to ca. 1600 K
(Fig. 1).
Dual-bed simulation results shown in Fig. 1 indicate
that the temperature in the bed rises "rst and then remains constant at a certain value, which is a typical
temperature pro"le for methane oxidation whereas
a peak is observed in temperature pro"les obtained in
mixed-bed simulations of Fig. 2. The sharp temperature
fall at the dual-bed interface is because of the high endothermicity of the reforming reactions running on
Ni/MgO catalyst, whose existence is con"rmed by
a sharp increase in hydrogen #ow rate in Fig. 4. The
increase in the methane #ow rate and the decrease in the

Fig. 1. Temperature pro"les obtained in industrial scale reactor with


dual-bed con"guration.

Fig. 2. Temperature pro"les obtained in industrial scale reactor with


mixed-bed con"guration.

hydrogen #ow rate after the maximum bed temperature


location in Figs. 3 and 4 indicate the occurrence of
reverse reactions (2), (3) and (5). Water}gas shift reaction
seems to be important in the determination of CO and
CO distribution in both the dual-bed and the mixed
bed schemes.
The shift in the locations of the maximum bed temperatures in Figs. 1 and 2 is due to the increase in the total
molar #ow rates at the reactor inlet where methane #ow
is kept constant. It is worth noting that the maximum
pressure drop is estimated to be ca. 1 atm in larger scale
reactor simulations.

A. K. Avcn et al. / Chemical Engineering Science 56 (2001) 641}649

647

Table 4
Investigation of the optimum operating conditions
CH /O
 

H O/CH



(K)


y 
&

2.09
1.62
1.52
1.43
1.33

1.81
1.47
1.33
1.27
1.15

773
873
936
973
1073

1.44
2.11
2.22
2.20
2.09

Hydrogen yield: moles of H produced/moles of CH fed.





Fig. 3. Product distribution obtained in industrial scale reactor with


mixed-bed con"guration (CH /O "2.24, H O/CH "1.17).
 



up to a certain point where the hydrogen yield reaches


a maximum value and then starts to decrease (Table 4).
Most of the methane in the feedstock is consumed by
oxidation, the bed temperatures increase, limited
amounts of methane are left for the reforming reactions,
and hence hydrogen is formed when the methaneto-oxygen ratio in the feed is low. At the other extreme,
the e$ciencies of the reforming reactions are low due to
the insu$cient temperature rise provided by oxidation.
Therefore, the presence of a maximum in H yield is

theoretically expected. Since the H O/CH molar ratio


for steam reforming is kept constant at 3 as mentioned in
Section 3.2, the H O/CH in the feed is decreased to

gether with the feed CH /O molar ratio, considering
 
that low CH /O will lead to higher H O formation.
 

Operating conditions leading to maximum hydrogen
yield constitute an optimum set. By employing these
results, approximate operating conditions for increased
hydrogen output can be determined for industrial scale
operation.

5. Conclusions

Fig. 4. Product distribution obtained in industrial scale reactor with


dual-bed con"guration (CH /O "3.53, H O/CH "0.88).
 



Several operating conditions such as di!erent molar


feed ratios are investigated thermodynamically for their
hydrogen yield values by the simultaneous solution of
Eq. (17) and reaction equilibrium expressions. Results are
given in Table 4.
It was mentioned that decrease in methane-to-oxygen
ratio in the feedstock leads to a rise in the maximum bed
temperature, and hence hydrogen production, due to the
endothermic character of the reforming reactions. In
mathematical analysis, this trend is observed to increase

Autothermal hydrogen production in the presence of


two di!erent catalyst is mathematically investigated by
a series of simulations conducted for bench scale and
hypothetical industrial scale reactors. Pt/d}Al O and
 
Ni/MgO}Al O catalyzing total methane oxidation and
 
steam reforming, respectively, are either placed consecutively (dual bed) or packed in the form of a physical
mixture (mixed bed). A one-dimensional heterogeneous
reactor model is used for the simulations. Heterogeneous
components describing heat and mass transfer resistances are incorporated into the simulations after checking the values of related criteria, and this greatly reduces
the computation time. Mixed-bed con"guration is observed to exhibit better performance when compared
with dual-bed scheme in both reactor dimensions.
Enhanced heat and mass transfer characteristics are
believed to facilitate hydrogen production in mixed-bed
scheme. Larger scale and bench scale operations possess

648

A. K. Avcn et al. / Chemical Engineering Science 56 (2001) 641}649

the same external transport characteristics except that


the interfacial mass transfer resistance is considered in
addition to the external heat transfer resistance in simulating the Pt bed in industrial scale dual-bed operation.
In both catalyst bed con"gurations, intraparticle di!usion limitations are signi"cant in industrial scale operation. Prediction of a maximum hydrogen yield indicates
the presence of an optimum set of operating conditions.

g
G
j
D
k
D
l
GH
o
@
o
D
o
A

e!ectiveness factor for reaction i


thermal conductivity of the bulk #uid, kJ/m h K
viscosity of the bulk #uid, kg/m h
stoichiometric coe$cient of j in reaction i
catalyst bed density, kg/m
density of the bulk #uid, kg/m
solid density of the catalyst, kg/m

Subscripts
Notation
a
K
c
NH
C
H
C
QH
D
N
D
C
D
HK
E

F
H
F\
H G
F>
H G
F
2
F M
2
h
Q
k
QH
k
G
K
H
K
G
n
P
H
P
2
P M
2
Pr
q 
&!r
G
Re
R
E
Sc

Q
=
x 
!&
y
H

exterior surface area per unit mass of catalyst,


m/kg cat
heat capacity of j, kJ/kg K
bulk #uid concentration of j, kmol/m
surface concentration of j, kmol/m
particle diameter, m
e!ective di!usivity inside catalyst, m/h
di!usivity of j into a mixture m, m/h
activation energy, kJ/kmol
molar #ow rate of j, kmol/h
amount of j consumed in reaction i, kmol/h
amount of j remaining after reaction i, kmol/h
total molar #ow rate, kmol/h
total molar #ow rate at reactor inlet, kmol/h
particle-to-#uid heat transfer coe$cient, kJ/m h K
particle-to-#uid mass transfer coe$cient of j, m/h
speci"c reaction rate of reaction i
adsorption equilibrium constant of j, atm\
equilibrium constant of reaction i
reaction order
partial pressure of j, atm
total pressure, atm
total pressure at reactor inlet, atm
Prandtl number
Energy required to vaporize water in the
feed, kJ/kmol
rate of reaction i, kmol/kg cat h
Reynolds number
gas constant, kJ/kmol K
Schmidt number
temperature, K
maximum bed temperature, K
inlet temperature, K
surface temperature, K
catalyst weight, kg cat
methane conversion
yield of j

Greek letters
b
parameter in the pressure drop equation

*H heat of reaction i, kJ/kmol
G

void fraction of the catalyst bed

f
i
j
o
s

bulk #uid
reaction number
component index
reactor inlet
surface conditions
total quantity

Acknowledgements
Financial support was provided by Bog\ azic7 i University through project DPT-97KI20640.

References
Armor, J. N. (1999). The multiple roles for catalysis in the production of
H . Applied Catalysis A, 176, 159}176.

De Groote, A. M., & Froment, G. F. (1996). Simulation of the catalytic
partial oxidation of methane to synthesis gas. Applied Catalysis A,
138, 245}264.
Fogler, H. S. (1999). Elements of chemical reaction engineering. (3rd ed.).
Englewood Cli!s: Prentice-Hall.
Furimsky, E. (1998). Selection of catalysts and reactors for hydroprocessing. Applied Catalysis A, 171, 177}206.
Hickman, D. A., & Schmidt, L. D. (1993). Production of syngas by
direct catalytic oxidation of methane. Science, 259, 343}346.
Jamal, Y., & Wyszynski, M. L. (1994). On-board generation of hydrogen-rich goseous fuels * a review. International Journal of Hydrogen
Energy, 19, 557}572.
Kuvshinov, G. G., Mogilnykh, Y. I., & Kuvshinov, D. G. (1998).
Kinetics of carbon formation from CH }H mixtures over a nickel
 
containing catalyst. Catalysis Today, 42, 357}360.
Ma, L. (1995). Ph.D. thesis. University of New South Wales, Sydney,
Australia.
Ma, L., & Trimm, D. L. (1996). Alternative catalyst bed con"gurations
for the autothermic conversion of methane to hydrogen. Applied
Catalysis A, 138, 265}273.
Ma, L., Trimm, D. L., & Jiang, C. (1996). The design and testing of an
autothermal reactor for the conversion of light hydrocarbons to
hydrogen I. The kinetics of the catalytic oxidation of light hydrocarbons. Applied Catalysis A, 138, 275}283.
Pena, M. A., Gomez, J. P., & Fierro, J. L. G. (1996). New catalytic
routes for syngas and hydrogen production. Applied Catalysis A,
144, 7}57.
Rase, H. F. (1990). Fixed-bed reactor design and diagnostics. London:
Butterworths.
Rosen, M. A. (1991). Thermodynamic investigation of hydrogen
production by steam-methane reforming. International Journal of
Hydrogen Energy, 16, 207}217.

A. K. Avcn et al. / Chemical Engineering Science 56 (2001) 641}649


Rostrup-Nielsen, J. R. (1984). In J. R. Anderson & M. Boudart (Eds.),
Catalysis, science and technology. Catalytic steam reforming vol. 5,
(pp. 1}117). Berlin: Springer.
Trimm, D. L. (1999). Catalysts for the control of coking during steam
reforming. Catalysis Today, 49, 3}10.
Wakao, N., & Kaguei, S. (1982). Heat and mass transfer in packed beds.
New York: Gordon and Breach Science Publishers.

649

Xu, J., & Froment, G. F. (1989a). Methane steam reforming, methanation and water}gas shift: I. Intrinsic kinetics. A.I.Ch.E. Journal, 35,
88}96.
Xu, J., & Froment, G. F. (1989b). Methane steam reforming: II. di!usional limitations and reactor simulation. A.I.Ch.E. Journal, 35,
97}103.

Вам также может понравиться