Sei sulla pagina 1di 8

ht. J. Hydrogen Energy, Vol. 20, No. 12, pp.

949-956, 1995
Copyright @ International Associationfor Hydrogen Energy
Elsevier Science Ltd
Printed in Great Britain. All rights reserved
036&3199(94)0013&8
036&3 199/95 $9.50 + 0.00

Pergamon

HYDROGEN
PRODUCTION
FROM THE LOW-TEMPERATURE
WATER-GAS
SHIFT REACTION:
KINETICS AND SIMULATION
OF THE
INDUSTRIAL
REACTOR
N. E. AMADEO
Departamento

and M. A. LABORDE

de Ingenieria Quimica(FI)/PINMATE(FCEyN)
Universidad de Buenos Aires/CONICET,
Pabell6n de Industrias, Ciudad Universitaria 1428 Buenos Aires, Argentina

Argentina

(Receiwd j?w publication 30 December 1994)


Abstract---The

kinetics of the water-gas shift reaction (WGSR) over a copper/zinc oxide/alumina

catalyst have been

studied. The experiments were carried out at 453-503 K and atmospheric pressure. A reactive mixture of similar
composition to that employed in the industrial process was used. An integral reactor, an integral procedure and a
data treatment valid for near equilibrium conditions were employed. A number of representative models were

examined. It was found that only a Langmuir-Hinshelwood model, which considers the adsorption of four species
(CO, CO,, H, and H,O) and the surface reaction as the controlling step, adequately describesthe reaction behaviour
at the temperature and concentration ranges investigated. Values of adsorption constants and adsorption heats for
the four components involved in the WGSR are given.
An algorithm for the simulation of an adiabatic fixed-bed reactor was developed with the aim of checking the
kinetics expression. Both the industrial and simulated compositions agree. It is proved that the kinetic expression
proposed which is in harmony with a Langmuir-Hinshelwood mechanism is useful in designing industrial lowtemperature converters.

NOMENCLATURE

CP
E,
E aP
F
FCOO
k
k
Kj
&
m
;
Pj
b rco)
W
X

Average specific heat (kJ mol- K- )


Activation energy from model III (kJ molt )
Apparent activation energy from model IV (kJ
molt )
Mole flow of CO (mol s- )
Objective function
Rate coefficient (mol g- s-l at-)
Pre-exponential factor (mol g- s- at -)
Adsorption equilibrium constant for component j (at-)
Equilibrium constant
Number of parameters in model i
Number of experimental data
Pressure (at)
Partial pressure of component j
Reaction rate (mol g- s-l)
Reaction rate of CO (mol g- s- )
Temperature (K)
Catalytic mass
CO conversion

Greek letters
P
P~~PH

MPCOPH,OKJ

Vahance

AH
AH,
Y

Adsorption heat from model III (kJ mol-)


Heat of reaction (kJ mol- )
Pressure factor defined in equation (8)

Subscripts
0

e
t

Inlet
Experimental
Theoretical
INTRODUCTION

The water-gas shift reaction (WGSR)


CO + H,O = CO, + H,

AH, = -40.6 kJ mol-

is one of the steps in the overall process of producing


hydrogen and ammonia from natural gas. The role of
WGSR is to reduce carbon monoxide levels and obtain
additional hydrogen.
Demand for hydrogen, a product of the shift reaction,
will undoubtedly become even greater in the future, since
new uses for hydrogen are expected to appear, such as
in coal liquefaction and gasification, hydrotreatment of
heavier petroleum and shale oil liquids, and its use as a
direct fuel [l].
949

950

N. E. AMADEO

and M. A. LABORDE

The WGSR is a reversible and exothermic reaction,


many materials being capable of catalysing it [2]. Two
classes ofmaterials are almost exclusively used in industry
as shift catalysts, the iron-based catalysts and the copperbased catalysts.
Copper-based shift catalysts arise from a more recent
development which has gained wide industrial acceptance.
These are the so-called low-temperature shift catalysts,
operating from 453 to 523 K [3]. These catalysts have
good activity at low temperature and are therefore attractive, since equilibrium is more favourable at lower temperature. In addition to higher activity, another advantage
claimed for the low-temperature shift catalysts is higher
selectivity and fewer side reactions at high pressures.
Numerous studies on copper-based low-temperature
catalysts have been published concerning the kinetics and
mechanisms. Despite the large amount of research on
this subject, there are still controversies about the kinetics
and mechanisms of the WGSR. Two different mechanisms are propsed: a surface redox mechanism and a
Langmuir-Hinshelwood
mechanism, each having numerous supporters. The surface mechanism is based on
the adsorption and dissociation of water on the catalytic
surface [4-121. Fiolitakis and Hofmann [13] postulated
a three-path reaction model consisting of two normal
Langmuir-Hinshelwood-type
mechanisms and a redox
mechanism which regulates the oxygen activity on the
catalyst surface. The LangmuirrHinshelwoodmechanism
involves a surface formate intermediate [14-231. Salmi
and Hakkarainen [24] suggested that an alternating
oxidation-reduction
process could be significant at the
beginning of the transient reaction period on the activated
catalyst, but the main reaction path probably consists of
a Langmuir-Hinshelwood
mechanism. Finally, Campbell
and Daube [25] postulated a Langmuir-Hinshelwood
mechanism with dissociation of adsorbed water as the
determining step, but the same authors suggested that a
redox mechanism could also be probable with the same
step as the controlling rate (dissociation of water).
Kinnaird et al. [26], referring to methanol kinetics on
a copper/zinc/alumina catalyst, stated that the wide range
of experimental conditions and the variety of catalyst
preparation techniques which have been employed have
probably contributed to some of the apparent discrepancies in exprimental results, particularly since the performance of these catalysts is strongly affected by reaction
conditions. The same concept can be applied to the
WGSR.
To date, it can be concluded that catalyst texture and
structure and feed mixture composition may strongly
affect the kinetics and mechanism of the WGSR.
The aims of this work are as follows.
(1) To find a kinetics expression for the WGSR which
fits the experimental data in a range of variables similar
to those used in the industrial process when a commercial
catalyst of copper/zinc/alumina
is used.
(2) To compare the kinetic expression found with that
usually employed in the design of low-temperature industrial converters.

(3) To simulate a low-temperature industrial converter


using the kinetic expression found in this work and to
compare the computed results with those obtained from
the industrial reactor.
KINETICS:

EXPERIMENTAL

Equipment
Experiments were performed in a conventional flow
system shown in Fig. 1. Distilled water was fed to an
evaporator using a high precision liquid pump. The
evaporator was filled with a bed of catalyst in order to
retain traces of chlorine present in water.
A mixture of the water vapour produced and gases
(CO, CO,, H, and NJ was passed through a preheating
zone prior to entering the reactor. The stainless steel
reactor consisted of a cylindrical cup of 10 mm diameter
and 23 mm length, perforated on the bottom with orifices
of 0.5 mm diameter, which enabled the flow of gases. The
catalyst was retained by a plug of glass wool. A thermocouple tube passed through the catalyst bed so that
temperature could be measured at different positions
along the bed. The bed temperature was controlled within
0.5C. The cup was placed in a stainless steel pipe of
12.5 mm diameter which was heated by a tubular furnace
of three heating sections.
After leaving the reactor, the remaining steam was
condensed out. Before passing through the sampling
chamber and through a wet gas meter the gas stream
was dried by silica gel.
Feed and effluent compositions were analysed by gas
chromatography.
A Hewlett-Packard
chromatograph
model 5730, equipped with thermal conductivity detectors, was employed to analyse CO, H, and N, using
a Porapak Q column, whereas CO, was separated by a
molecular sieve column.
Chemicals
CO 99.9% from LAir Liquid and CO,, H, and N,
99.9% from La Oxigena were filtered through silica gel
and used without any further purification. Distilled water
was used.
Characteristics of the commercial catalyst used were:
composition (%p/p), CuO : 32.7, ZnO : 47, Al,O, : 11;
specific surface, 42 mz g-l; poral volume, 0.11 cm3 g-i;
and copper metallic area, 13 mz g-.
Procedure
Catalyst was loaded in the reactor and reduced in situ
for 4 h at 453 K and 12 h at 503 K using a gas mixture
of 1% H, in N,. At this stage a strict temperature control
was established in order to avoid copper oxide sintering
and/or zinc oxide reduction. Both phenomena produce
copper crystal growth and, as a consequence, catalyst
activity decay [27].
Preliminary experiments with different particle sizes
and feed rates at the highest operating temperature

HYDROGEN

PRODUCTION

-r

1. flow meter
2. sample

N2

3. water reservoir
4. pump
5. vaporiser
6. reactor
7. water trap
8. silicagel

drier

9. gas meter

H2

Fig. 1. Scheme of the equipment.

(503 K) were made in order to eliminate internal and


external diffusional resistances. These experiments indicated the absence of both resistances when operating
with a feed rate of 5 cm s-l, a catalytic mass of 0.4 g
and a particle size of 0.35.-0.42 mm. The operative
conditions used in this work are shown in Table 1.
It is well known that the feed mixture composition
affects the deactivation rate of the catalyst [28]. It was
found that the catalyst activity remained almost constant
during a long period provided that CO partial pressure
was maintained above a certain critical value [29, 303.
Experiments were performed at a constant temperature
varying the inlet partial pressure of the gases one at a
time, keeping the partial pressure of the other components
constant. Nitrogen was used in order to keep the total
gas flow constant.
RESULTS

AND DISCUSSION

In Fig. 2(aHd) the effects of partial pressures of CO,


H,O, H, and CO, on CO conversion are shown. It can
Table 1. Experimental

conditions

Water flow rate: 1.7-2.3 mg s-


Gas volumetric Row rate: 5.0-8.3 ml s-
CO molar fraction: 0.03~0.12
H, molar fraction: 0.274.52
CO, molar fraction: 0.0-0.12
Temperature: 453-503 K
Pressure: 2 x 10 Pa
Catalyst mass: 0.4 g
Particle diameter: 0.35-0.42 mm

be appreciated that CO conversion declines as inlet


partial pressures of H, and CO, are increased, whereas
it increases with increasing water partial pressure. This
set of experiments was performed at WIF,, and W/F,
constant.
In Fig. 2(a), it can be seen that CO conversion decreases
when pso is increased. It must be noted that, in this case,
the experimental data were obtained working at W/F&,
variable and W/F, constant. At this condition and if it
is assumed that the system is first order with respect to
CO, CO conversion should to remain constant when CO
partial pressure is varied. Nevertheless, this effect does
not occur. Recalling that this set of experiments was
performed at pto, = 0, the decrease of CO conversion
can be explained by the large effect of pco which, in this
case, increases proportionally
when p:o >ncreases. The
same behaviour was observed by Barreto et al. [31]
working with an Fe/Cr catalyst.
As has already been mentioned, isothermia in the bed
and absence of diffusional resistances were confirmed
experimentally. In order to find the kinetic expressions,
the plug flow hypothesis was made and an integral reactor
behaviour was assumed since the experimental conversion values were higher than 10% in a11cases.
Mass balance in the reactor is given by:
x --dX
__1
k(T) j o .f(W =x Fw

(1)

Functions f(X) and k(T) were obtained by means of the


optimization REGRE routine [32]. Equation (1) was
solved by applying Simpsons method.

N. E. AMADEO and M. A. LABORDE

952
0.6

0.5

1011(1)
0.5
0.3 -

0 Experimental values -Model


0.20*

0.05

III

D Experimental values -Model

0.10
&O

.I

0.15

0.05

0.10

0.15

0.20

x 10m5Pa

0.25

0.30

III

0.35

0.40

0.45

0.50

pi o x 10e5 Pa
2

0.5 -

0.4
(cl

0.3 -

0.2 -

80

0.2 -

L)Experimental values -Model


0.1

000

0.1 -

0.25

0.30

0.35

0.40

0.45

III
0.50

0 Experimental values -Model


0.55

0.60

0.05

p& x 10m5Pa

III

0.10

p&

0.15

x 10m5Pa

Fig. 2. (a) CO conversion vs inlet CO partial pressure (at). Experimental data: operative conditions: T = 503 K; P: atmospheric
pilo = 0.4 x lo5 Pa, p& = 0.36 x 10 Pa, pgo, = 0 0 Pa. (b) CO conversion vs inlet H,O partial pressure (at). Experimental data:
operative conditions: T = 503 K; P: atmospheric pgo = 0.12 x lo5 Pa, paH, = 0.44 x 10s Pa, p& = 0.0 Pa. (c) CO conversion vs
inlet H, partial pressure (at). Experimental data: operative conditions: T = 503 K; P: atmospheric pg, = 0.11 x lo5 Pa,
p& = 0.26 x 10s Pa, pg, = 0.0 Pa. (d) CO conversion vs inlet CO, partial pressure (at). Experimental data: operative conditions:
T = 503 K; P: atmospheric p& = 0.11 x 10 Pa, pg,, = 0.26 x 10s Pa, pi, = 0.26 x lo5 Pa.

The optimal parameters were obtained by searching


for the minimum of the following objective function:

model V
r = kpcopH20(l

FO = i (X, - X,)2.
-1

Variance of each model was calculated from:


g2 =

FO

(n-m)

(3)

The following kinetic models were fitted:


model I
r = kpn20U - PMAP,~,

- PM1 + KCO,~CO, + &P~J.

(2)

+ PCO)

model II

Models I and II, proposed by Shchibyra et a[. [4] and


Fiolitakis and Hoffman [13], respectively, represent a
redox mechanism. Models III-V are Langmuir-Hinshelwood-type models. Models III and IV postulate the
adsorption of the four species,being the critical step the
surface reaction (model III) and the CO adsorption
(model IV). Model V postulates only the adsorption of
CO and CO,, being the surface reaction the control step.
Values of the objective function and variance obtained
for each model are shown in Table 2.
From this table, it is clear that model III presents the
best fitting. The variation of CO conversion with inlet
partial pressure of the four components given by this

r = k,k,pcopn20U - B)Mk,Pco+ k2PH20+ k,pcoJ


Table 2. Objective function and variance

model III
r = kp~op&l

- B)Ml + KiPi)

model IV
r = kp,o(l

- PM1 + K~~co,~~,pco,pH,:,

+ KH2pH,

Model

Objective function

Variance

I
II

0.58
1.80
0.10
0.49
0.27

0.0050
0.0145
0.0008
0.0040
0.0020

III
IV
V

+ &,o~H,o

+ KCO~PCOJ

HYDROGEN

model is also plotted in Fig. 2. It can be seen that this


model predicts the behaviour observed experimentally.
Model III represents a LangmuirrHinshelwood
reaction mechanism and, in addition, gives information about
the adsorption constants involved in the process. Values
of the apparent activation energy and the adsorption heat
for model III were obtained from a fitting made with all
the available experimental data. The set of resultant
values is shown in Table 3. The objective function and
the variance for this fitting with 140 data were 0.12 and
0.01, respectively. These values indicate that this fitting
is as satisfactory as that obtained with data corresponding
to a sole temperature.
It can be seen that adsorption constants of CO and
H,O remain almost invariable with temperature. The
effect of temperature seems to arise from H, and CO,
adsorption constants, explaining the low variation of the
kineticcoefficient with it. Campbell [33] arrived at similar
conclusions using the same commercial catalyst.
Some authors relate the low sensibility of the kinetic
coefficient to temperature with the existence of diffusional
resistances (external or internal) in the catalyst, which is
not the case in this work, as has been mentioned
previously. In addition, the absence of diffusional resistances was verified applying theoretical criteria [34]. The
relative concentration gradients calculated, both inside
and outside the catalytic pellet, were less than lo- 5.
The agreement of the results presented here with those
of Campbell [33] is not surprising since the same catalyst
and similar operative conditions (especially feed gas
composition) were used in both works. The main difference between both experiments is that Campbell worked
in the presence of diffusional resistances while, in the
present paper, these resistances were eliminated.
In a recent paper [35], the authors, following the
surface redox mechanism, assume that both the dissociation of H,O* and the reaction between CO* and 0* may
be slow but, for large H,O/CO feed ratios, the oxidation
of CO is the controlling step (industrial converter operates at high H,O/CO ratios). In this case, these authors
postulate the following kinetic expression:

953

PRODUCTION

and they use the initial rates measured on a commercial


catalyst by van Herwijnen and de Jong [16] in order to
prove their model. They find a reasonable agreement
between both the calculated and the measured rates
except for high water contents and high temperatures
(4899501 K). At these latter conditions the calculated rate
is higher than the experimental value.
We tried to fit, with our experimental data, the model
proposed by Ovesen et ~11.[35] and we found that the
predicted conversions were always 3.5 times higher than
our experimental conversions. This result is not surprising
since these authors affirm that a possible explanation of
the lack of fitting of the redox model at high temperatures
and high H,O/CO ratios is that the observed kinetics is
sensitive to the value of the kinetics coefficient for the
CO oxidation. Therefore, the observed decrease could be
described by their model if it could calculate a CO
oxidation rate 20 times lower (as a general rule, the
reaction rate increases two times when temperature
increases 10 K).
SIMUL.ATION
The kinetics expression proposed in this work, replacing the values of adsorption constants and kinetics
coefficient, is:

(-rco) =
0.92e~454~3~7p,opH20(1
- /I)
+ 0.0047e2737~9~Tpco2
(1 + 2.2e101.5/TpC0+ 0.4e158.~TpHzo
454.3%JOPH,0(1 +0.92e~,yje1596.1!TpH,)2

1855.5
>

Table 3. Adsorption

co

co*

AH

-0.91 +
0.02

-- 24.72 k 0.4

KO

2.21 *
0.11

0.0047 f
0.0002

H,O

energy of

H*
-- 14.4 k
0.50

0.40 i
0.02

0.052 +
0.002

4.08 + 0.06

k,:

0.92 + 0.08

P)

(5)

- 1.42 f
0.05

Ea, :

kOPH,Ofl

02,

heats and apparent activation


model 111

(4)

where ( -rco) is given in moles per gram per minute and


T in kelvins.
The most popular kinetics expression used in the design
of the industrial reactor is that proposed by Moe [2, 36,
371:
r = 1.85 x 10m5 exp 12.88 -T

Pco,

P)

which is also given in moles per gram per minute. This


is an empirical expression and no mechanism can be
inferred from it.
Prior to carrying out the simulation of the industrial
reactor a comparison of the kinetics with that obtained
in this work [equation (4)] was made. As the reactor
operates adiabatically, the comparison was made considering an adiabatic path, which means that the reaction
rate in equations (4) and (5) is only a function of the
conversion since:
T = T + yto -?-AHr (X - X0).
1,

(6)

954

N. E. AMADEO
1,/r

(l,mol;g/

and M. A. LABORDE

s)

6200
P

5200

!"
/

4200
3200

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

conversion

--Et

model

III

Moe (2)

Fig. 3. Verification of the kinetic model used in this work. (Reaction rate)- vs conversion, adiabatic path, for Moe [2] and our

kinetics expressions. To = 453 K.

In Fig. 3, l/r vs x for the two kinetics expressions


[equations (4) and (5)] and for T = 453 K is shown. It
can be seen that the agreement is very good, especially
at low conversion values. The differences appear at high
conversion values, i.e. at the lower values of the reaction
rate, therefore they will have no influence on the industrial
reactor simulation, as can be seen later.
In industry this reaction is carried out in a packed bed
adiabatic reactor, at temperatures between 453 and
523 K and pressures of 15-20 at. The hypothesis of plug
flow is assumed.
The integral fixed bed reactor can be represented as a
series of N subreactors with a length AZ = L/N and a
space time of A( W/pco) = (Wr/Pc,)/N [37].
The material balance for CO in each subreactor is
given by:

r = Yl.85

x lo-

exp 12.88 -T

1855.5
!

PCOPH,OU- /3).

(10)
The solution of the state differential equation (7) taking
into account equations (6) and (9) [or (lo)], has been
found by the fourth-order Runge-Kutta method.
The computed results, using the reaction rates given
by equations (9) and (lo), were compared with those
available from industrial data, as shown in Table 4. The
operating conditions used in the simulation are typical
of the industrial low-temperature converter employed in

Table 4. Verification of the kinetic parameters usedin this work*


A( W/F;,)

AX

= ( - rcok

(7)

As the reactor operates adiabatically equation (6) is valid.


The kinetics expressions given in equations (4) and (5)
were obtained at atmospheric pressure; however, the
effect of this variable can be considered through the factor
proposed by Rase [36]:
Y = 0.86 + 0.14

for P < 24.S at.

(8)

This factor also includes the diffusional effects into the


catalytic slab, therefore the reaction rate expressions used
in the simulation, considering equations (4), (5) and (8),
are given by:

(- rco),
= w - rco)p
=1

(9)

Exit gas composition (vol.%)

H2
co
co2
CH,
Exit temperature
(K)

Industrial

Simulation
using Moes
[2] expression

Simulation
using our
expression

77.42
0.43
20.36
1.79
505

17.40
0.42
20.50

17.23
0.48
19.79
1.74
502

1.80

505

*Operating conditions (from Gonzilez Velasco et al. [37]):


P, 17 x lo5 Pa; To, 480 K; steam feed flow, 0.208 mol s- ;
dry-gas feedRow,0.339mols~. Compositionofdrygas(vol.%):
H,: 76.8; CO: 3.2; CO?: 18.2; CH,: 1.8.

HYDROGEN

PRODUCTION

955

~ll/&/TF

F
0

0.2

Model III

0.4

r/

MOE (2)
0.6

0.8

1.0

4.50.
0

0.2

MOE (:2)
0.4

Model III
0.6

0.8

1.0

Dimensionless position
Fig. 4. (a, b) Conversion and temperature profiles predicted using MOoe[2] and our kinetics expressions. T = 453 K, P = 22 x 10
Pa, $o = 2.5 x 10 Pa, pi.o = 8.14 x lo5 Pa, pi, = 6 x 10 Pa. Pco, = 1.65 x IO5 Pa, M = 2.043 x IO g, F& = 0.216 mol s-l.

manufacture of ammonia. The good agreement between


industrial and simulated compositions indicates the adequacy of the kinetics expression obtained in this work
and the algorithm used in the simulation. The agreement
between both simulations should also be noted.
This agreement is not limited to the prediction of the
outlet values of conversion and temperature. The conversion and temperature profiles obtained applying equations (9) or (lo), for different industrial operating conditions, are shown in Fig. 4(a) and (b). It can be seen that
the simulation using the reaction rate deduced from this
work predicts conversion and temperature profiles very
similar to those obtained using the empirical reaction
rate proposed by Moe [2].

CONCLUSIONS
In the present paper attention was centred on reproducing the industrial procedure, i.e. kinetic data were
obtained using a commercial catalyst (CuO/ZnO/Al,O,)
and feed compositions similar to those of an industrial
reactor.
WGSR kinetic was determined in an integral reactor,
using an integral procedure and a data treatment valid
for near equilibrium conditions. A Langmuir-Hinshelwood kinetic expression which considers adsorption of
the four components and the surface reaction as the
controlling step was found to fit accurately the experimental data.
Values of adsorption constants and adsorption heats
for the four components involved in the WGSR are given.
In addition, the kinetic was compared successfully, in the
complete range of conversions and considering an
adiabatic path, with the empirical kinetics proposed by
Moe and largely used in the design of this kind of reactor.
Finally, our kinetic expression obtained at atmospheric
pressure and modified with the pressure factor suggested
by Rase [36] was used in the simulation of an industrial
reactor. The computed results agree very well with the
data produced from an industrial low-temperature converter employed in the ammonia process.

The conversion and temperature profiles predicted by


simulation using our kinetic expression are very similar
to those predicted using the empirical expression proposed by Moe [Z].
In summary, a kinetic expression has been proposed
in harmony with a Langmuir- Hinshelwood-type mechanism at least at the operating conditions and with the
commercial catalyst used in this work. In addition, it has
been proved that this expression can be used to simulate
and, in consequence, to design an industrial low-temperature converter.
It was not the aim of this work to elucidate the WGSR
mechanism; but, if it is accepted that the chemical
kinetics is a field in which mechanistic knowledge and
hypotheses are translated into useful equations that
describe these processes in terms ofconstants independent
of the composition of reactants and products and of the
reaction conditions [23, 381, it must be concluded that
the kinetics expression proposed in this work satisfies
these requirements.
Acknowlrdgrments---The
authors thank CONICET
(Consejo
National de Investigaciones Cientificas y Tecnicas), the Universidad de Buenos Aires and the Antorchas Foundation for
financial support. They also thank Dr Graciela Cerrella for the
experimental work.

REFERENCES
1. D. Newsome, Cutc~l.Rec. Sci. Engng 21, 275 (1980).
2. J. M. Moe, Chem. Enyng Progr. 58, 33 (1962).
3. H. Bohlbro, Acta Chem. Scund. 16, 431 (1962).
4. G. Shchibrya, N. Morozov and M. Temkin. Kinrt. Cutul..
USSR 6, 1010 (1965).
5. E. Cherednik, N. Morozov and M. Temkin, Kinet. Caful.,
USSR 10, 94 (1969).
6. T. Semenova. B. Lyudkovskaya. M. Markina, Ya. Volynkina, G. Cherkasov, V. Sharkina, N. Khitrova and G. Shpiro,
Kinet. Katd., USSR 18, 1014 (1977).
7. E. G. M. Kuijpers, R. B. Tjepkema and W. J. J. van der
Wal, Appl. Cutal. 25, 139 (1986).
8. G. C. Chinchen, M. S. Spencer, K. C. Waught and D. A.
Whan. J. Chem. Sot.. Furaduv Trans. 1 83. 2193 (1987).
9. J. Nakamura, J. M. Campbelland C. T. Campbell: J. Chem.
Sot. Faraduy Trans. 86, 2725 (1990).

956

N. E. AMADEO

and M. A. LABORDE

10. E. Colbourn, R. A. Hadden, H. D. Vandervell. K. C. Waugh


and G. Webb, JI. C~ltuI. 130, 514 (1991).
11. K. H. Ernst, C. T. Campbell and G. Moretti. J. Cata/. 134,
66 (1992).
12. S. 1. Fujita, S. I., M. Usui and N. Takezawa, J. Cutul. 134,
220 (1992).
13. E. Fiolitakis and H. Hofmann, J. Cutal. 80, 328 (1983).
14. T. Van Herwijnen, On the kinetics and mechanism of the
CO-shift conversion on a copper/zinc oxide catalyst, Ph.D.
Thesis, Delft University, The Netherlands (I 973).
15. Y. Amenomiya, J. C&l. 57, 64 (1979).
16. T. Van Herwiinen and W. de Jonu. J. Cutal. 63. 83 (1980).
17. T. Van Herwiinen, R. Guzalski and W. de Jong, J.Cut;l.
63, 94 (1980).
18. D. C. Grenoble, M. M. Estadt and D. F. Ollis, J. Catal. 67,
90 (1981).
19. C. Rofer - De Porter, Chem. Rev. 81, 447 (1981).
20. Y. Amenomiya and G. Pleizier, J. Catal. 76, 345 (1982).
21. J. Edwards and G. Scharader, J. Phys. Chem. S&5620(1984).
22. D. G. Bybell, P. P. Deutsch, R. G. Herman, P. B. Himelfarb,
J. G. Nunan, C. W. Young, C. E. Bogdan, G. W. Simmons
and K. Klier, Syposium on Fundamental Chemistry of Promoters and Poisons in Heterogeneous Catalysis, p. 116, New
York, 13-18 April (1986).
23. K. Klier, C. W. Young and J. G. Nunan, Ind. Engng Chem.
Fundam. 25, 36 (1986).
24. T. Salmi and R. Hakkarainen, Appl. Cutal. 49, 285 (1989).
25. C. T. Campbell and K. A. Daube, J. Catal. 104, 109 (1987).
26. S. Kinnaird, G. Webb and G. Chinchen, J. Chem. Sot.,
Faraday Trans. 1 83, 3399 (1987).

27. C. Quincoces. N. Amadeo. M. Gonzalez, XII Simposio


Iheroamericano de Catcilisis, Segovia. Espaiia. Vol. I, p. 529
( 1992).
28. G. Barreto, M. Gonzalez. M. Laborde, N. Moreno and
J. Viiias, 7mas Jornadas Argentinas de Catilisis. Mar del
Plata, Argentina (1991).
29. N. E. Amadeo, E. G. Cerrella. M. A. Laborde and F.
Pennella, XIII Sin~po.sio Iberoumericuno de Cutiilisis,
Segovia, Espana, Vol. II, p. 779 (1992).
30. N. E. Amadeo, E. G. Cerrella, M. A. Laborde and F.
Pennella, Latin Am. Appl. Rex (in press).
31. G. Barreto, M. Gonzalez, M. Laborde, N. Moreno and J.
Viiias, XII Simposio Iheroamericano de Cat&e.
Rio de
Janeiro (1990).
32. 0. Quiroga and J. Gottifredi, Acances en /a Determinwidn
de Pwimetros Cinkticos en Sistemus de Reacciones Complejas. Fundacion para la education, la ciencia y la cultura,
Buenos Aires, Argentina (1982).
33. J. S. Campbell, Ind. Engng Chem. Proc. Res. Dev. 9,588 (1970).
34. G. Froment and K. BischolT, Chemical Reactor, Analysis and
Desian. Wilev. New York (1979).
35. C. VT Ovesen; P. Stoltze, J. K. Nbrskov and C. T. Campbell,
J. Catal. 134, 445 (1992).
36. H. F. Rase, Chemical Reuctors Design Jar Process Plants,
Vol. II. Wiley, New York (1977).
37. J. R. Gonzalez Velasco, M. A. Gutitrrez Ortiz, J. A. Gonzalez
Marcos, N. E. Amadeo, M. A. Laborde and M. Paz, Chem.
Engng Sci. 47, 1495 (1992).
38. M. Boudart, Kinetics of Chemicul Processes. Prentice Hall,
New York (1968).

Potrebbero piacerti anche