You are on page 1of 8

Full Paper

Simulation of Secondary Reformer in Industrial Ammonia Plant


By Yong-Ho Yu*

In order to develop a reactor design model for the secondary reformer in the industrial ammonia plant, the effectiveness factor
and convection heat transfer coefficient between gas and catalyst surface have been studied. The temperature and composition
of inlet gas to the catalyst bed are predicted using the kinetic equations of 32 radical reactions. The effect of oxygen content in air
on the product synthesis gas composition and the ratio of synthesis gas to nitrogen have been studied. The effectiveness factor has
been calculated with the assumption that the steam methane reforming reaction is first order in methane partial pressure. The
catalyst shape is assumed to be spherical with an equivalent volumetric diameter. The temperature and composition profiles
along the axial distance are predicted using a one-dimensional heterogeneous catalytic reaction model. The temperatures of
both gas and catalyst surface decreased with the axial distance from the top of the bed, while the reactions took place. The
temperature difference between gas and catalyst surface also decreased along the axial distance. The predicted temperature and
composition by the proposed simulation method have been verified with the data from the industrial plant.

1 Introduction In the present study, the temperature and composition of


gas in the combustion zone have been determined by the
In general, the secondary reformer in the ammonia plant 32 radical reaction mechanisms and their kinetic equations
plays an important role in further converting methane from proposed by Karim and Metwally [6]. Then, the temperature
the primary reformer and supplying nitrogen by controlling profiles of the gas and catalyst surface as well as the gas
the air flow rate, the optimum molar ratio of synthesis gas (CO composition have been predicted with a one-dimensional
+ H2) to nitrogen being 3.0. When the conventional furnace- heterogeneous catalytic reaction model, considering the heat
type steam reformer is used as primary reformer in the transfer between gas and catalyst surface and the catalyst
ammonia plant, the synthesis gas for ammonia synthesis effectiveness factor. The predicted results from the simulation
can be produced by feeding air to the secondary reformer, are compared with the data from the industrial-scale plant.
because the higher methane conversion in the primary
reformer can be achieved due to the higher temperature of
2 Experimental Section
780±850 C. Whereas, in the case of the heat-exchanger-type
reformer as a primary reformer as presented by Yu and Sosna A schematic diagram of the industrial-scale secondary
[1], oxygen-rich air should be supplied to the secondary reformer is shown in Fig. 1. The oxygen mixer, which mixes gas
reformer to give an optimum molar ratio of synthesis gas to from the primary reformer with oxygen-enriched air, is
nitrogen because of the relatively lower methane conversion mounted on the top of the secondary reformer. The gas from
in the primary reformer. On the other hand, if air was fed to the primary reformer reacts with oxygen completely in the
the secondary reformer, excess nitrogen should be separated conical combustion zone (Fig. 1).
by PSA (Pressure Swing Adsorption) system as reported by The shell of the reformer is fabricated from carbon steel,
Nirula [2]. When the oxygen-rich air is fed to the secondary while the two-layer refractory is lined to protect the carbon
reformer, the temperatures of both combustion zone and steel wall and to reduce the heat loss. The inside diameter of
catalyst bed are higher than in the case of air. Since the the reformer is 2.02 m and the thickness of lining is 390 mm.
thickness of refractory and the material selection of the The high alumina corundum plate, having an opening area of 6
oxygen mixer in the secondary reformer depend on the %, is located above the catalyst bed in order to distribute the
temperature, the predictions of temperature and gas composi- combusted gas evenly to the catalyst bed. The alumina balls of
tion are important factors for optimizing the investment costs 50 mm diameter are packed below the corundum plate with a
and determining the design criteria. depth of 300 mm. The nickel catalysts (Haldor Topsoe, RKS-
Previous studies by De Groote and Froment [3], Tsipouriari 2±7H) are packed with the volume of 11.5 m3.
et al. [4], and Witt and Schmidt [5] have included the partial The physical properties of the catalyst are summarized in
oxidation reaction of methane feed without hydrogen. Tab. 1. The average pore diameter measured with a mercury
However, there is little information on the effect of the porosimeter (Micrometrics, Autopore III ) is 0.122 lm. In
oxygen content of the oxygen-rich air on the product order to measure the temperature profiles along the axial
composition and the reaction temperature with the hydrogen distance, four thermocouples are located inside the catalyst
content of the feed gas being above 50 %. bed separated by 1 000 mm from each other, while the first
thermocouple is located at 320 mm below the top of the
±
catalyst bed. The catalyst support, covered with alumina balls
[*] Y.-H. Yu, R&D Center, Samsung Engineering Co.,LTD., 39-3 Seongbok-
Ri , Suji-Eup,Yongin, Kyunggi-Do, Korea 449-844; e-mail:yuyh@hana- of 50 mm and 25 mm, has a similar shape of cone, which is
net.net. comprised of bricks with slot holes.

Chem. Eng. Technol. 25 (2002) 3, Ó WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2002 0930-7516/02/0303-0307
0930-7516/02/0303-0307$$17.50+.50/0
17.50+.50/0 307
Full Paper

CH4 + O2 ® CO + H2O +H2 (2)

CH4 + 2O2 ® CO2 + 2H2O (3)

However, Karim and Metwally [6] have analyzed the above


thermal reforming reaction with a concept of radical reaction
mechanism. Karim and Metwally [6] analyzed the above
reactions with 32 reversible radical reactions of H, HO, O,
HO2, H2O2, CH3, CHO, and CH2O. Only Karim and Metwally
[6] have considered the hydrogen combustion mechanism,
while most previous researchers as Leroux and Mathieu [7],
De Groote and Froment [3], and Witt and Schmidt [5] have not
considered the hydrogen combustion mechanism in the partial
oxidation process in either the combustion zone or the catalyst
bed. Because the gas from the primary reformer contains
more hydrogen produced by the steam methane reforming
reaction, the hydrogen combustion kinetics should be
considered to predict the temperature and composition of
gas. Therefore, in this study, the reaction rate constants and
activation energies for 32 radical reactions proposed by Karim
and Metwally [6] are adopted to predict the composition and
temperature of gas in the combustion zone.

3.2 Catalytic Reactions

While the gas from the combustion zone flows through the
Figure 1. Schematic diagram of secondary reformer. 1 oxygen mixer; catalyst bed, it is known that the further methane steam
2 combustion chamber; 3 corundum plate; 4 alumina ball (dia. = 50 mm);
5 catalyst bed; 6 thermocouples; 7 alumina ball (dia. = 25 mm and 50 mm); reforming reaction takes place accompanied by the CO-shift
8 catalyst support reaction as follows:

CH4 + H2O = CO + 3H2 DH = 206.1 kJ/mol (4)


Table 1. Physical properties of catalyst used in this study.

Properties values
CO + H2O = CO2 + H2 DH = -41.15 kJ/mol (5)

Model Name RKS-2-7H For the above reactions, many researchers have proposed
the kinetic equations as summarized by Kim and Lee [8]. Most
Nominal size, mm 20 ” 18
kinetic equations are difficult to apply to the simulation of the
Actual size, OD, mm 20 secondary reformer, because their experimental ranges are
Actual size, ID, mm (7x)4 mainly focused on temperatures of the primary reformer
below 850 C. However, Khomenko et al. [9] proposed the
Actual size, H, mm 18
kinetic equation of methane conversion above 900 C, using
Chemical composition (typical) nickel foil to avoid the interaction of mass transfer effect. In
NiO, wt.-% 9
this case, the kinetic equation by Khomenko et al. [9] can be
MgAl2O4, wt.-% Balance
Radial crush strength, kg/cm (typical) 28 applied as the surface reaction rate for the estimation of the
Filling density, kg/L 0.9 effectiveness factor with some compensation of the nickel
content. In the case of other kinetic equations, the surface
reaction rate has to be defined with a specified thickness from
the surface as proposed by Elnashaie et al. [10].
Therefore, in the present study, the kinetic equation
3 Simulation of Secondary Reformer proposed by Khomenko et al. [9] has been adopted for the
simulation of the catalytic reaction in the industrial secondary
3.1 Combustion of Gas from Primary Reformer reformer, as the following equation1):
The overall reaction mechanism of hydrogen and methane k1 PCH
R1 ˆ 4 (6)
with oxygen in the combustion zone can be considered as 1‡K5 PH =PH
2O 2
follows:
±
H2 + 1/2 O2 ® H2O (1) 1) List of symbols at the end of the paper.

308 Ó WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2002 0930-7516/02/0303-0308 $ 17.50+.50/0 Chem. Eng. Technol. 25 (2002) 3
Full Paper

Because the reaction rate constant proposed by Khomenko reactions at steady state. The heat balance in unit volume can
et al. [9] is based on the nickel foil, it should be compensated be expressed as follows
with the nickel content for the industrial catalyst. According to 2
P
the information provided by Haldor Topsoe, the nickel oxide Qk gk Rk …1 ÿ e† ÿ 6…1 ÿ e†=Ds h…Tg ÿ Ts † ˆ 0 (12)
content and nickel surface area of the catalyst applied to the kˆ1

secondary reformer are 9 % and 1850±2500 m2/kg. Therefore, where, the film heat transfer coefficient at the catalyst surface
if the average surface area of nickel inside the catalyst is can be calculated by the following well-known equation for
2175 m2/kg, the intrinsic rate constant is expressed as following packed bed:
equation:
Nu = 0.4 Re0.64
p Pr1/3 (13)
21
k1 = 2.38 ´ 10 Ts±3 exp{±33720 / RTs}
Therefore, the heat balance for the gas phase only in axial
cm3 …STP† 10ÿ6 Nm3 kmol direction can be written as
m2 Ni atm h cm3 …STP† 22:414 Nm3
2
@Tg P
Fg Cp;g g ˆ Qk gk Rk …1 ÿ e†A (14)
2175 m2Ni900 kgcat @` kˆ1
= 2.078 ´ 1020 T±3s
kgcat m3 cat
kmol By applying Eqs. (12), (13), and (14), the profiles of the mole
exp {±33720 / RTs} 3 (7) fractions of each component as well as the temperatures of
mcat h atm
both gas and catalyst surface can be predicted.

K5 = 1.062 ´ 10±1 Ts±0.5 exp {±680 / RTs} (8)


3.4 Effectiveness Factor
The volumetric kinetic constant kv throughout the unit
conversion can be described with a correction factor, C1, to In the heterogeneous reaction, the actual reaction rate is
compensate the activity in the industrial catalyst. affected by the molecular diffusion into the micro pore inside
the catalyst. Therefore, the effectiveness factor should be
0:082 m3 atm considered in order to apply the kinetic equations to the
kv ˆ C1 k1 Ts
kmol K industrial reactor design. If the catalyst is sphere, the Thiele
modulus can be defined by the following equation with a
= C1 1.704 ´ 1019Ts±2 exp [±33620/(RTs)], m3 / m3cat h (9)
characteristic length, Ds/6, as noted by previous researchers
including Rostrup-Nielsen et al. [11] and Froment and
where, R is the gas constant as 1.987 kcal/(kmol K)
Bischoff [12].
Khomenko et al. [9] and Rostrup-Nielsen et al. [11] reported
r
that the CO-shift reaction rate is much faster and reaches to D kv
u= s (15)
the equilibrium state. Therefore, the rate equation can be 6 De …ec =s†
written as the following equation, while the intrinsic reaction
where, Ds is the equivalent diameter and defined as
constant has an increased value until the CO fraction would
not be changed. 6  volume of particle
Ds ˆ (16)
( ) surface area of particle
PCO PH
R2 ˆ k2 PCO 1 ÿ 2 2 (10) The effective diffusivity of i-component into the bulk phase
Kp2 PCO PH
2O is defined as

1 1 1
ˆ ‡ (17)
3.3 Heat and Mass Balances in Catalyst Bed De Di;k Di;mix

where, Di,mix is the molecular diffusivity of i-component into


In the secondary reformer, the gas from the combustion
the bulk phase and can be estimated from the following
zone flows through the catalyst bed, while the catalytic
equation as defined by Bird et al. [13]:
reactions of (4) and (5) take place on the nickel catalyst.
Therefore, the mass balance for i-component and heat balance …1ÿXi †
Di;mix ˆ n (18)
in axial direction based on a one dimensional heterogeneous P Xj
catalytic reaction model can be described by Dei;j
jˆ1
2
@ni P j 6ˆ i
ˆ ni;k gk Rk …1 ÿ e†A (11)
@` kˆ1
where, Dei,j is the molecular diffusivity (m2/h) of i-component
It may be assumed that the heat transfer between gas and into j-component and can be estimated from the following
catalyst surface is equivalent to the reaction heat of both equation by Sherwood et al. [14]:

Chem. Eng. Technol. 25 (2002) 3, Ó WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2002 0930-7516/02/0303-0309 $ 17.50+.50/0 309
Full Paper

1=2 Table 2. Results from the radical reaction of primary reformer gas with oxygen-
0:36  0:001858T 3=2 …1=Mi ‡1=Mj † rich air or air.
Dei;j ˆ (19)
Pri;j 2 Xi;j
Type of Primary Gas from Primay Oxygen rich Gas into catalyst
Reformer reformer air bed
On the other hand, the Knudsen diffusivity, Dk,i can be
(Plant data) (Plant data) (Estimated value)
estimated from the following equation introduced by Froment
Heat Exchanger Type
and Bischoff [12]:
 1=2 Pressure, atm 24 25 23.8
4r 2RT Temperature, C
Di;k ˆ (20) Compostion (vol.-%)
3 pMi
H2 27.47 17.01
The measured average pore radius of the catalyst is
CO 2.34 2.23
46.453 A. However, in order to estimate the effectiveness
CO2 6.08 6.46
factor, the porosity (ec) and the tortuosity (s) should be taken
CH4 12.74 6.73
into account. The measured catalyst porosity is 0.385, while
the tortuosity of 2.74 estimated by Elnashaie et al. [10] is N2 61.97 16.14

applied because of the similar catalyst, Ni/MgOAl2O3. From O2 26.34


Eqs. (17) to (20), the effective diffusivity can be estimated. H2O 51.37 11.69 50.85
Therefore, the effectiveness factor can be estimated using the Flow rate, Nm3/h 57 817 19 933 76 512
following equation as noted by Froment and Bischoff [12] with Furnace Type
the assumptions that the catalyst particles are spherical and Pressure, atm 24.8 24.2 23.8
the reaction rate of methane is first order in methane partial Temperature, C 792 535 1057
pressure. Compostion (vol.-%)
H2 38.17 18.83
1 …3u†coth…3u†ÿ1
g= (21) CO 7.38 2.98
u 3u
CO2 9.02 6.62
Lywood [15] reported that the effectiveness factor with an CH4 5.28 2.30
assumption of spherical catalyst particle can be applied to the N2 0.34 77.49 10.64
design of the industrial catalytic reformer. Moreover, because
O2 20.85
the kinetic equation suggested by Khomenko et al. [9] is first
H2O 39.82 0.73 58.48
order to methane partial pressure, it can be considered that
Flow rate, Nm3/h 98 475 24 150 179 000
this method to estimate the effectiveness factor would be
reasonable.

4 Results and Discussion

4.1 Gas from the Combustion Zone

The simulated results for the temperature and composition


of the gas from the combustion zone with the radical reactions
proposed by Karim and Metwally [6] in the industrial-scale
secondary reformer are shown in Tab. 2. It can be estimated
that 30 %, 11.8 % and 58.2 % of oxygen to the secondary
reformer are consumed for hydrogen combustion (1),
methane partial oxidation (2), and methane combustion (3),
respectively. The estimated value for the gas temperature in
the combustion zone is 1297 C. On the other hand, the
estimated temperature in the combustion zone of a conven- Figure 2. Effect of oxygen content in oxygen-rich air on the temperature of
tional ammonia plant, where air is fed to the secondary oxygen mixer inlet, combustion zone and reformer outlet and on the product
reformer, is somewhat lower than that of the process where mole fraction of H2 and CH4 in the combustion zone. The condition of gas from
the primary reformer is referred to Tab. 2; l O2 mixer inlet; combustion zone;
oxygen-rich air is fed. z reformer bottom; * H2 mole fraction; & CH4 mole fraction.
The effect of the oxygen content on temperature of both the
combustion zone and reformer bottom of the secondary
reformer are shown in Fig. 2. Though both temperatures in the content is increased to 34 %, when prereformed gas is used.
two zones increased with an increase of the oxygen content, This means that the endothermic steam-methane conversion
the temperature increase in the secondary reformer is slightly increases with an increase of inlet temperature of the catalyst,
lower than that of the combustion zone until the oxygen while the further increase in the inlet temperature of the

310 Ó WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2002 0930-7516/02/0303-0310 $ 17.50+.50/0 Chem. Eng. Technol. 25 (2002) 3
Full Paper

catalyst bed with an increase of oxygen content above 34 %


does not give any significant effect to the conversion rate of
methane.
As can be seen in Fig. 2, the methane mole fraction in the gas
from the combustion zone decreases proportionally with an
increase of oxygen content, while the rate of hydrogen
decreases because the hydrogen production from the partial
oxidation of methane. However, the hydrogen mole fraction
decreases rapidly with a further increase in oxygen content
above 34 % due to the increase of hydrogen combustion rate.
The effect of oxygen content in air on the product
synthesis gas from the secondary reformer is shown in Fig. 3.
As can be seen, hydrogen production shows a maximum at
the oxygen content of 29.5 %. The CO production increases
with an increase in the oxygen content, while the CO2
production decreases, since the exothermic CO-shift reac-
tion rate decreases with increasing reaction temperature. Figure 4. Effect of oxygen content in oxygen-rich air on the ratio of synthesis gas
The synthesis gas (CO + H2) shows a maximum at an oxygen to nitrogen. (for the condition of primary reformer gas is see to Tab. 2).
content of 29.5 %. Therefore, it can be assumed that the
oxygen content of more than 29.5 % is not useful at a given
plant operation condition, when a heat exchanger type
primary reformer is used. 4.2 Catalytic Reaction

The kinetic constant determined by Khomenko et al. [9] has


no consistency with the industrial catalytic reactor due to the
value from the experiment on the laboratory scale. If the
correction factor is not multiplied with the kinetic constant,
the simulation result represents that the methane is converted
completely since the kinetic equation represents an irrevers-
ible reaction.
The simulation results with the various correction factors C1,
from 0.378 to 0.52 are compared with the data from of industrial-
scale plant as shown in Tab. 3. In the case of plant data 1, where
oxygen-rich air is supplied, a higher outlet temperature
compared to the plant data is predicted but the methane
content is predicted correctly with the correction factor of 0.378.
However, the lower methane content is predicted with the
correction factor of 0.472, though similar temperature to the
plant data is predicted. On the other hand, the simulated
temperature and composition values with a correction factor of
Figure 3. Effect of oxygen content in oxygen-rich air on the mole fractions of H2, 0.472 show good agreement with the plant data 2. It can be
CO and CO2 in the product gas. (for the condition of gas from the primary considered that the higher reaction rate compared with the
reformer see to Tab. 2).
actual rate is estimated at the higher temperature. Therefore,
the correction factor of 0.425±0.472 would be multiplied with
For optimum production of ammonia, a value of 3 for the the reaction rate constant suggested by Khomenko et al. [9] in
ratio of synthesis gas to nitrogen should be maintained order to design the industrial-scale secondary reformer.
according to Jenning and Ward [16]. The effect of the oxygen The temperature profiles of gas and catalyst surface in axial
content in oxygen-rich air on the molar ratio of synthesis gas to direction are shown in Fig. 5. The temperatures of both gas and
nitrogen with the variation of the air flow rate is shown in catalyst surface as well as the difference between them the two
Fig. 4. The ratio of synthesis gas to nitrogen also shows a of them decrease along the axial distance from the top of the
maximum with increasing oxygen content. The oxygen bed. It can be said that the methane steam reaction rate is
content at which the ratio shows a maximum is shifted to the higher at the top of the catalyst bed, since the partial pressure
right as the air flow rate decreases. Therefore, optimum flow of methane as well as the reaction temperature are high. In Fig.
rate of air and oxygen can be determined from the trends of 5 can also be seen that the simulation result of the temperature
the production of synthesis gas and the ratio of synthesis gas to profile shows good agreement with the plant operation data
nitrogen at a given condition of gas from a primary reformer. when the correction factor C1 is 0.425. However, in the case of

Chem. Eng. Technol. 25 (2002) 3, Ó WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2002 0930-7516/02/0303-0311 $ 17.50+.50/0 311
Full Paper

Table 3. Comparison between plant data and simulation results at the reformer outlet.

Primary Heat Exchanger Type Conventional Furnace Type


reformer

Plant Simulation (deviation, %) Plant Simulation (deviation, %)


data 1 data 2

Correction ± 0.378 0.425 0.472 ± 0.425 0.472 0.52


factor, C1

Press., atm 22.5 23 23 23 23.8 23.8 23.8 23.8

Temp., C 955 972 965 960 942 951 947 942


vol.-%

CH4 0.45 0.6(33.3) 0.38(15.6) 0.23(48.9) 0.3~0.5 0.5(25) 0.42(5) 0.3(25)

H2 55.07 53.32(3.2) 53.66(2.6) 53.89(2.14) 51 52.6(3.14) 52.88(3.69) 53.9(5.7)

CO 11.95 12.52(5.1) 12.56(5.1) 12.6(5.45) 13~14 9.8(27.4) 9.86(26.9) 9.9(26.7)

CO2 10.31 9.67(6.2) 9.68(6.1) 9.68(6.1) 12.13 14.4(18.7) 14.45(19.) 14.44(19)

N2 22.22 23.87(7.42) 23.69(6.6) 23.58(6.1) 23 22.2(3.48) 22.1(3.9) 22.12(3.8)

plant 1, the predicted temperature profiles of gas and catalyst The profiles of mole fractions of each component along
surface with C1 being 0.378 and 0.472 show some deviation the axial distance in the catalyst bed are shown in Fig. 6. It
from the plant data 1, since they depend on the reaction rate. can be observed that CH4 conversion is mostly carried out in
On the other hand, in the case of plant 2, the predicted half the catalyst bed. However, the CO-shift reaction rate is
temperature profiles with a C1 value of 0.472 are in good low due to the high temperature operation even though the
agreement with the plant data 2. Therefore, it can be CO-shift reaction approaches the chemical equilibrium
considered that the correction factor for the reaction rate state, since the CO2 fraction is almost unchanged through-
constant would be 0.425±0.472 for the design of the industrial- out the catalyst bed.
scale secondary reformer.

Figure 5. Temperature profiles along the axial distance with the various Figure 6. Mole fraction profiles of each component along the axial distance for
correction factors for kinetic constant. plant 1 with a correction factor C1 of 0.425.

312 Ó WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2002 0930-7516/02/0303-0312 $ 17.50+.50/0 Chem. Eng. Technol. 25 (2002) 3
Full Paper

4.3 Effectiveness Factor reduction factor which is a power function of fractional


oxygen conversion.
The effectiveness factor is an important parameter to Throughout this study, the profiles of temperature as well as
predict the temperature profiles and composition to realize the product composition of the plant data are well predicted
the actual reaction with pore diffusion. Previous researchers with the correction factor for kinetic constant.
such as Namaguchi and Kikuchi [17], Xu and Froment, [18],
Elnashaie et al. [10], De Groote and Froment [3], and Yu and
Sosna [1] reported that the effectiveness factor for industrial 5 Conclusions
nickel catalysts would be applied.
The axial profiles of the effectiveness factor estimated from The simulation results by radical reaction in the combustion
Eqs. (15) to (21) are shown in Fig. 7. The effectiveness factor zone and catalytic reaction considering the effectiveness
for methane steam reforming increases in axial direction. The factor show good agreement with the industrial plant data.
same trend for steam reforming has been observed in a When a heat exchanger type primary reformer is used, the
previous study of Xu and Froment [18]. production of synthesis gas and the ratio of synthesis gas to
nitrogen show a maximum at 29.5 % of oxygen content in air.
The temperature gradient between gas and catalyst surface
decreases along the axial distance and both temperatures
approach the same value at the bottom of the catalyst bed. The
effectiveness factor increases along the axial distance. This
simulation model shows a high relieability for the design of an
industrial secondary reformer with the correction factor for
kinetic constant of Khomenko et al. [9].

Acknowledgement

The measurement of catalyst porosity and pore diameter


was carried out by Prof. W. M. Lee of the Taejon National
Hanbat University. The author would like to acknowledge his
contribution.
Received: August 10, 2001 [CET 1441]

Symbols used

fe;Figure 7. Effectiveness factor for methane-steam reaction along the axial


A [m2] reactor cross-sectional area
distance. C1 [±] correction factor for kinetic
constant
of steam reforming
De Groot and Froment [3] reported that the effectiveness Cp,g [kcal kg±1 K±1] heat capacity of gas mixture
factors of the CO and CO2 production by the steam-reforming De [m2 h±1] effective diffusivity of
reaction among the several reactions in the catalytic partial i-component
oxidation reactor are 0.07 and 0.06, respectively. However, in De i,j [m2h±1] diffusivity between i-component
this study, the effectiveness factor for the steam reforming and j-component
reaction in the case of oxygen-rich air has values from 0.0017 Di,k [m2 h±1] Knudsen diffusivity
to 0.0055 along the axial distance as can be seen in Fig. 7. On Di,mix [m2 h±1] molecular diffusivity of
the other hand, the effectiveness factor in the case of air only i-component into mixed phase
shows somewhat higher values ranging from 0.0035 to 0.0057. Ds [m] equivalent diameter of catalyst
The different order of effectiveness factor compared with the particle
value of De Groot and Froment [3] may be induced from the Fg [kmol h±1] mole flow rate of gas mixture
different reaction kinetic equation as well as the different h [kcal m±2 C±1 h±1] heat transfer coefficient
process by catalytic partial oxidation. De Groote and Froment k [kmol m±3 h±1 atm±1] intrinsic rate constant of reaction
[3] developed the VDR (varying degree of reduction) model kg [kcal m±1 C±1 h±1] thermal conductivity of gas
to illustrate the temperature peak in axial direction with the l [m] axial distance
consecutive reaction of partial oxidation and steam reforming. ni [kmol h±1] mole flow rate of i-component
In their model, the rates of the steam reforming reactions and Nu [±] Nusselt Numer , Ds h/kg
the water gas shift reaction were compensated with a Q [kcal m±3] reaction heat

Chem. Eng. Technol. 25 (2002) 3, Ó WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2002 0930-7516/02/0303-0313 $ 17.50+.50/0 313
Full Paper

P [atm] pressure References


Pr [±] Prandtl Number, Cp.g l/kg
r [m] average pore radius [1] Yu, Y. H.; Sosna, M. H, Korean J. Chem. Eng. 18 (2001) No.1, pp. 127±
132.
Rk [kmol m±3 h±1] reaction rate of kth reaction
[2] Nirula, S. C., Ammonia from Natural Gas by ICI ªLCAº Process, in:
R [kcal kmol±1 K±1] gas constant PEP Review No. 89±1±3, SRI Int., California 1990.
Rep [±] particle Reynolds number, [3] De Groote, A. M.; Froment, G. F., Appl. Catal. A 138 (1996) pp. 245±
264.
Ds rgV/l [4] Tsipouriari, V. A.; Zang, Z.; Verykios, X. E., J. of Catalysis 179 (1998)
T [K] temperature pp. 283±291.
Xi [±] mole fraction of component i [5] Witt, P. M.; Schmidt, L. D., J. of Catalysis 163 (1996) pp. 465±475.
[6] Karim, G. A.; Metwally, M. M., Int. J. Hydrogen Energy 5 (1980) pp. 293±
304.
[7] Leroux, P. J.; Mathieu, P. M., Chem. Eng. Prog. 57 (1954) No.11, pp. 54±
Greek symbols 59.
[8] Kim, D. H.; Lee, T. J., HWAHAK KONGHAK 29 (1991) pp. 396±406.
[9] Khomenko, A.; Apel'baum, L. O.; Shub, F. S.; Snagovskii, S.; Temkin,
fi,k [±] stoichiometric coefficient of M. I., Kinet. Katal. 12 (1971) pp. 423±430.
i-component in the kth reaction [10] Elnashaie, S. S. E. H.; Adris, A. M.; Soliman, M. A.; Al-Ubaid, A. S.,
Can. J. Chem. Eng. 70 (1992) pp. 786±793.
e [±] catalyst bed porosity [11] Rostrup-Nielsen, J.; Dybkjaer, I.; Christiansen, L. J., Steam Reforming ±
ec [±] catalyst porosity Opportunities and Limits of the Technology, in: Chemical Reactor
g [±] effectiveness factor Technology for Environmentally Safe Reactors and Products (H. I. de
Lasa et al., Eds.), Kluwer Academic Pub., The Netherlands 1993,
u [±] Thiele modulus pp. 249±281.
s [±] tortuosity [12] Froment, G. F.; Bischoff, K. B., Chemical Reactor Analysis and Design,
John Wiley & Sons, New York 1979, pp. 163±200.
[13] Bird, R. B.; Stewart, W. E.; Lightfoot, E. N., Transport Phenomena, John
Wiley & Sons, Singapore 1960, pp. 563±572.
Subscripts and Superscripts [14] Sherwood, T. K.; Pigford, R. L.; Wilke, C. R., Mass Transfer, McGraw-
Hill, Tokyo1975, pp. 6±24.
[15] Lywood, W. J., Process Design, Rating and Performance, in: Catalyst
g gas
Handbook (Twigg, M.V. 2nd ed.) Wolfe Pub, England 1989, pp. 85±139.
i ith component [16] Jennings, J. R.; Ward, S. A., Ammonia Synthesis, in: Catalyst Handbook
j jth component (Twigg, M.V. 2nd ed.) Wolfe Pub, England 1989, pp. 384±440.
[17] Namaguchi, T.; Kikuchi, K., Chem. Eng. Sci. 43 (1988) No. 8, pp. 2295±
k reaction numer 2301.
s catalyst [18] Xu, J.; Froment, G. F., AIChE J. 35 (1989) No.11, pp. 97±103.

_______________________

314 Ó WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2002 0930-7516/02/0303-0314 $ 17.50+.50/0 Chem. Eng. Technol. 25 (2002) 3

You might also like