You are on page 1of 14

Applied Catalysis A: General, 97 (1993) 145-158 145

Elsevier Science Publishers B.V., Amsterdam

APCAT A2480

Kinetic mechanism for the reaction between methanol


and water over a Cu-ZnO-Al,O, catalyst

C.J. Jiang, D.L. Trimm and M.S. Wainwright


School of Chemical Engineering and Industrial Chemistry, University of New South Wales,
P.O. Box 1, Kensington, NSW2033 (Australia)

and
N.W. Cant
School of Chemistry, Macquarie University, NSW 2109 (Australia)
(Received ‘29October 1992, revised manuscript received 12 January 1993)

Abstract

The steam reforming of methanol over a Cu/ZnO/Al,O, catalyst has been investigated. The reaction
yields carbon dioxide and hydrogen in the ratio of one to three, with small amounts of dimethyl ether
and carbon monoxide being produced at high conversion. Comparison of the rates of methanol dehy-
drogenation and of steam reforming over the same catalyst indicate that steam reforming proceeds via
dehydrogenation to methyl formate. Methyl formate then hydrolyses to formic acid which decomposes
to carbon dioxide and hydrogen. Detailed studies of the kinetics of the reactions show that methanol
dehydrogenation controls the rate of steam reforming. Langmuir-Hinshelwood modelling indicates that
hydrogen extraction from adsorbed methoxy groups is rate determining to the overall processes.

Keywords: copper/zinc oxide/alumina; formate; kinetics; mechanism; methanol synthesis; steam


reforming

INTRODUCTION

Hydrogen production by steam reforming of methanol is best carried out


over a copper based catalyst [ 11.
CH30H+H20-rC02+3H2 (1)
There is general agreement that metallic copper is the active ingredient and
that zinc oxide is an effective promoter. The fact that copper is a good catalyst
both for methanol synthesis and for the water-gas shift equilibration has led
to speculation that a decomposition-shift mechanism exists [ 2-51. It is sup-

Correspondence to: Prof. M.S. Wainwright, School of Chemical Engineering and Industrial
Chemistry, University of New South Wales, P.O. Box 1, Kensington, NSW 2033, Australia. Tel.
(+61-2)6975000, fax. (+61-2)6622904.

0926-860X/93/$06.00 0 1993 Elsevier Science Publishers B.V. All rights reserved.


146 C.J. Jiang et al. / Appl. Catal. A 97 (1993) 145-158

posed that carbon monoxide is first generated and then reacts on the surface
with water to form carbon dioxide.

CH,0H-CO+2H2 (2)
CO+H,O+CO,+H, (3)

Thermodynamic calculations by Pour et al. [ 51 showed that theoretically it


was possible for methanol almost completely to decompose to carbon monox-
ide and hydrogen under typical conditions for reaction (1). However their ex-
perimental work indicated that the concentration of carbon monoxide in the
products was usually well below equilibrium. The only explanation offered was
that part of the carbon monoxide produced by reaction (2) was consumed to
produce some higher alcohols, acetic acid and formaldehyde. This explanation
does not seem reasonable as the production of the two condensed products
needs much higher pressures and temperatures [ 6,7]. Furthermore attempts
to interpret the results in terms of the decomposition-shift mechanism using
Langmuir-Hinshelwood models gave a rather poor fit.
In our recent investigation of the kinetics of reaction (1) over a Cu/ZnO/
A1203 catalyst [ 81, no carbon monoxide could be detected in the products be-
low 523 K and significant amounts were formed only at temperatures over 573
K. Furthermore, the introduction of carbon monoxide along with methanol
and water had little influence on the rate of hydrogen production or the Hz/
CO, ratio in the products. Perhaps most significantly it was found that, al-
though the catalyst gave high conversions for the water-gas shift reaction (2 )
when this was carried out alone, the reaction did not occur when methanol was
present. Thus decomposition-shift appears untenable as the sole mechanism
of hydrogen production from methanol over this catalyst.
Further tests showed that methyl formate was the major product when
methanol alone was reacted over the catalyst. Previous studies [ 9,101 have also
shown that a wide variety of copper catalysts are active and selective for the
dehydrogenation of methanol to methyl formate at atmospheric pressure over
the temperature range 453 to 513 K. It was therefore speculated that, following
the suggestion of Takahashi et al. [ 111, methanol steam reforming proceeded
by the following set of reactions

2CH,OH-+CH,OCHO+2H, (4)
CH30CHO+H~O-+CH30H+HCOOH (5)
HCOOH-CO, +Ha (6)

In the present work the rates of the three individual reactions (4)) (5 ) and
(6) have been measured in comparison to that of the overall methanol steam
reforming reaction (1). The origin of deactivation which occurs during reac-
C.J. Jiang et al. /Appl. Catal. A 97 (1993) 145-158 147

tion (4) but not with reaction ( 1) [ 81 has also been investigated and a kinetic
model consistent with the above reaction scheme has been developed.

EXPERIMENTAL

The catalytic measurements were carried out using the single pass flow sys-
tem described previously [B] with 0.05 to 0.5 g samples of a commercial Cu/
ZnO/A1203 methanol synthesis catalyst (BASF S3-85 ) crushed and sieved to
a 250 to 500 pm particle fraction. The unreduced catalyst contained 31.7%
CuO, 49.5% ZnO and 18.8% Al,O,. The total BET surface area was 83 m2/g
and the copper metal area 15.5 m2/g as measured by nitrous oxide chemisorp-
tion and dissociation after hydrogen reduction [ 121. Smaller masses of catalyst
were diluted to similar bed volumes using a-Al,O, of the same particle size.
For steam reforming, methanol decomposition and formic acid decomposition
the liquid reactants (or mixtures of them with water) were pumped into a
heated vaporiser and carried over the reduced catalyst in a stream of nitrogen
at a total pressure of one atmosphere. The organic fraction leaving the reactor
was analyzed by periodic sampling into one gas chromatograph. Products not
condensable at 195 K (i.e. CO, C02, H2 and N,) were then determined using a
second chromatograph with a thermal conductivity detector and argon as the
carrier gas. Further details of the analytical methods and other procedures are
reported elsewhere [ 81. The hydrolysis of methyl formate was studied simi-
larly except that the formate was introduced separately from water using a high
pressure syringe pump (ISCO Model LC-2600) to avoid problems with inter-
mittent vaporisation.

RESULTS AND DISCUSSION

Methanol steam reforming

Blank tests showed that the stainless steel reactor, and the cu-Al,O, used as
a catalyst diluent in some experiments, were inactive for reactions between
water and methanol over the temperature range of interest (400 to 600 K ) .
Over the same range the Cu/ZnO/A120, catalyst was both active and selective
for reaction (1). Dimethyl ether and carbon monoxide were the only detectable
by-products but each amounted to less than 1% for methanol conversions to
over 80%. Fig. 1 shows the result of one ten hour test at 493 K. The conversion
was unchanged at 62 2 2% throughout and the yields of hydrogen and carbon
dioxide were in the expected ratio of 3 : 1 within experimental error.
Experiments carried out under differential conditions over a wide range of
temperature and pressure and described elsewhere [ 81, have established that
the steam reforming of methanol (rate expressed as rsn) is well represented
by the power law rate expression
148 C.J. Jiang et al. f Appl. Catal. A 97 (1993) 145-158

C
80_‘=’ - l l -. . l :. -*

(1 _.. - - 1.
60- .”

40-
_. _.__ b
- -.
ZO-

O-O 12
Time, hours

Fig. 1. Time-on-stream data for stream reforming of methanol over 0.4 g of Cu/ZnO/A1,03 at 493
K with methanol and water partial pressures each equal to 33 kPa and total input flow-rate of 140
ml(STP)/min: (a), methanol conversion; (h), rate of carbon dioxide production; (c), rate of
hydrogen production.

~~‘SR
= 5 3.1012 e - 105 kJ mol -‘/RT $W20H p”.03
.
Hz0
(7)

with the rate in mmol(CH3OH) kg-’ s-’ and pressures in kPa. An additional
term P -o.2o, needs to be included if hydrogen is added at pressures in excess
of 7 bPaHi8]. The rate expression [ eqn. ( 7 ) ] is purely empirical and is reliable
for reactor design within the range of state variables used in the experiments.

Methanol dehydrogenation

Methanol dehydrogenation, reaction (4), is equilibrium limited with a max-


imum conversion of 2540% under the present conditions of interest. As re-
ported by Tonner et al. [ 91 for copper catalysts in general, the selectivity falls
off steeply for temperatures above 500 K and for conversions exceeding 50%
of the equilibrium value. With the present Cu/ZnO/A1203 catalyst, the chief
by-products were dimethyl ether, carbon dioxide and a trace of carbon mon-
oxide. It is likely that dimethyl ether is produced by dehydration of methanol
on the Al303 component of the catalyst since it is a minor by-product for copper
catalysts formulated with other supports [ 91. The carbon dioxide presumably
arises by hydrolysis of methyl formate using the water produced during di-
methyl ether production.
Reaction rates were measured under differential conditions (conversions
less than 30% of equilibrium). The selectivity to methyl formate then exceeded
95% and the rate of the methanol dehydrogenation reaction could be calculated
using
rate = FMX/ W
C.J. Jiang et al. / Appl. Catal, A 97 (1993) 145-158 149

where FM is the input flow-rate of methanol in molar terms, W is the catalyst


weight and X the conversion of methanol to methyl formate calculated accord-
ing to

x= 2Y&F
2( Y&F + YDME)
+ YMeOH
+ yco2
whereYM~F, YDME,YNI~OHandYCO* are the mole fractions of methyl formate,
dimethyl ether, methanol and carbon dioxide of carbon containing compounds
in the stream leaving the reactor as determined by gas chromatographic anal-
ysis. Fig. 2 shows rates measured as a function of time on stream for dehydro-
genation at 453 K, 478 K and 493 K. Unlike steam reforming, this reaction
exhibits continuous deactivation. Previous infrared measurements [ 131 sug-
gest that the deactivation may be due to polymerisation of formaldehyde which
itself has been observed as a minor product with unsupported copper catalysts
[lo]. Formaldehyde was detectable in trace quantities in the liquid products
trapped at 195 K in the present experiments.
Deactivation processes can be expressed [ 14,151 by the following general
equation:

where kd is the rate constant for deactivation, the subscripts A, D or P, refer


to reactants, products or poisons respectively (applied singly or in combina-
tion depending on the source or sources of deactivation), m is the order in the
components and d the order in the activity, a.
Based on a study of deactivation during the hydrolysis of acrylonitrile over

240
: \
s200- ‘\
5 \
5 160- 0‘1
‘.

=b-
”f 120- --a-...n
--n---_-o
5 -q
E IjO- -O--=--o___
-*---_-o__
g 40- -0

II --o--o- -o--o_____ _o-___-o--o


O*
0 30 60 90 120 150
Time , minutes
Fig. 2. Rate of methanol dehydrogenation vs. time-on-stream for reaction over 0.2 g of Cu/ZnO /
A&O, with methanol partial pressure of 23.8 kPa. ( 0 ) 453 K; (a ) 478 K; ( 0 ) 493 K.
150 C.J. Jiang et al. / Appl. Catal. A 97 (1993) 145-158

Raney copper, Onuoha and Wainwright [ 161 proposed a second order mech-
anism associated with polymer fouling. In that reaction system C~‘,D,Pj is
constant and eqn. (8) reduces to

(9)

If activities are expressed relative to the initial reaction rate, then integra-
tion of eqn. (9) gives the following relationship between the reaction rate over
the fouled catalyst and time.

(10)

where, FDH is the reaction rate at particular time t in mmol g-’ s-‘; rLH is the
reaction rate at time zero (i.e. the rate over the fresh surface) and kd is the
deactivation rate constant in s-l.
Fig. 3 shows plots of l/r nn against time using the data from Fig. 2. Straight
lines are obtained. Thus the course of reaction can be represented by second
order deactivation kinetics at all three temperatures. Values for deactivation
constants ( kd), and initial reaction rates over a fresh catalyst surface (rkH),
obtained from the slopes and intercepts of the plots of Fig. 3 are listed in Table
1. The apparent activation energy for deactivation, as estimated from the tem-
perature dependence of kd, is 43 kJ/mol. Such a low value is consistent with a
fouling process [ 161.
The apparent activation energy for methanol dehydrogenation itself, ob-
tained from the initial rate values is 103 kJ/mol, which is close to that observed
for the steam reforming reaction. Experiments carried out with two different

60 90
Time, minutes
Fig. 3. Test of second order deactivation model, eqn. (lo), against time-on-stream data of Fig. 2
for methanol dehydrogenation over Cu/ZnO/Al,O,: ( 0 ) 493 K; q ) 478 K; (+ ) 453 K. The solid
lines represent the model fit.
C.J. Jiang et al. / Appl. Catal. A 97 (1993) 145-158 151

TABLE 1

Deactivation constants ( kd) and initial dehydrogenation rates ($4 in comparison to steam re-
forming rates ( rsR)

Temperature 453 K 478 K 493 K

kd W-‘1 5.4.10-5 10.4*10-5 13.8.10-5


r& (mmol (CH,OH) kg-’ s-l)0 22 103 210
rSR (mmol(CH,OH) kg-’ s-l)= 12 52 116

’ For a methanol partial pressure of 20 kPa and a water partial pressure of 20 kPa.

methanol pressures (24 and 42 kPa) gave identical time-on-stream plots for
methanol dehydrogenation. Thus the kinetic order in methanol is zero order
both for dehydrogenation and for the deactivation process.
The final row of Table 1 gives values for the rate of steam reforming when
measured at the same temperature and pressure as dehydrogenation or as cal-
culated by eqn. (7). It is apparent that the rate of steam reforming is about
one-half the initial dehydrogenation rate. This is as expected if steam reform-
ing follows the dehydrogenation/hydrolysis route represented by eqns. (4) to
(6) with step (4) rate determining. Reaction (5) returns one of the two meth-
anol molecules consumed in reaction (4) thus halving the overall rate of meth-
anol consumption.
In order to determine the reason why the same catalyst showed markedly
different stabilities for methanol dehydrogenation and methanol steam re-
forming, attention was focused on the influence of water on the catalyst sta-
bility. An experiment was carried out in which steam reforming was com-
menced with the water-to-methanol ratio initially set at two. The ratio of water
to methanol was then reduced in steps to zero, while the partial pressure of
methanol was kept constant (28 kPa) by make-up with nitrogen gas. The rates
of methanol consumption by steam reforming and dehydrogenation were cal-
culated from carbon dioxide and methyl formate analyses respectively made
at each step. Methanol conversions were less than 10% throughout. The results
are shown in Fig. 4.
For partial pressures of water above 7 kPa (water:methanol= 1:4) no methyl
formate production was detected and the steam reforming rate was indepen-
dent of water pressure as expected from eqn. (7). At lower pressures the carbon
dioxide production (steam reforming) rate dropped off with the dehydroge-
nation becoming dominant below 5 kPa water pressure. The falloff in steam
reforming is approximately linear as expected if the rate of reaction with water
is largely complete. Deactivation was less than that indicated in Fig. 2 when-
ever water was present, and completely absent when it was in excess.
It seems that water reduces deposition of the foulant that causes the catalyst
deactivation during dehydrogenation. In order to confirm this, a further ex-
C.J. Jiang et al. /Appl. Catal. A 97(1993) 145-158

Water pressure, kPa

Fig. 4. Influence of water partial pressure on the rate of methanol disappearance by steam reform-
ing (m) and dehydrogenation (0 ) over 0.1 g of Cu/ZnO/A1203 at 473 K with methanol pressure
of 28 kPa.

Ol
0 4 8 2
Time, hours

Fig. 5. Rate of dehydrogenation of methanol (53 kPa) over Cu/ZnO/A&O, at 473 K as a function
of time: (a) fresh catalyst; (b) following regeneration in steam as per text.

periment was carried out in which reaction was initiated with methanol alone
at 473 K (Fig. 5). After six hours the catalyst was treated with 60 kPa of water
vapour in nitrogen for one hour followed by nitrogen alone for 30 min. The
initial activity on recommencement of reaction with methanol, and the sub-
sequent deactivation profile, was similar to the original values. Thus water can
regenerate a catalyst deactivated during methanol dehydrogenation as well as
maintain stable activity during methanol steam reforming.
C.J. Jiang et al. / Appl. Catal. A 97 (1993) 145-158 153

Hydrolysis of methyl formate

The proposed dehydrogenation/hydrolysis route for methanol steam re-


forming, [eqns. (4), (5) and (S)], is tenable only if the methyl formate hy-
drolysis step, (5)) is fast. Trials with the Cu/ZnO/A120s catalyst showed that
this was indeed the case. At conversions up to 75% the observed products of
methyl formate hydrolysis were carbon dioxide, hydrogen and methanol in the
ratio 1:1:1. This is as expected if eqn. (5 ) describes hydrolysis and the follow-
ing step, formic acid decomposition, is complete. Above 75% conversion the
methanol product was found to be involved in steam reforming and the yields
of carbon dioxide and hydrogen increased.
The kinetics of hydrolysis were then investigated under differential condi-
tions (conversion < 15% ) where steam reforming did not affect the rate. Formic
acid decomposition (reaction 6) was found to be fast (see below) and mea-
surements of the production of carbon dioxide provided a measure of the rate
of hydrolysis. The results showed that methyl formate slightly retards the re-
action, (Fig. 6), that water pressure had almost no effect on rate and that
inclusion of hydrogen at pressures up to 30 kPa had no effect. Multiple linear
regression analyses on all rate data for methyl formate hydrolysis showed that
the following power rate law fitted the rate data well.
rHDR= 3.g2.10'3 e-87.4kJmol-1/RTp~~~~~H0 (11)

where runR is the hydrolysis rate in mmol (CH,OCHO ) kg- ’ s- ’ and pres-
sures are in kPa.
This empirical equation provides predictions of rate that are within + 1% of
experimentally determined values for the range of state variables used in the
experiments.

: 300
$w - -cmc_,

6
o’OO- w
m
6
% so-
EE -lx_
ai xF\
$20-
5 10 20 30 40 50
Pressure (CH3 OCHO), kPa

Fig. 6. Effect of methyl formate partial pressure on the rate of hydrolysis of methyl formate over
0.05 g of Cu/ZnO/Al,O, with water partial pressure of 23 kPa: ( x ) 443 K; (e) 453 K, (0) 463
K, (0) 473 K.
154 C.J. Jiang et al. / Appl. Catal. A 97 (1993) 145-158

Formic acid decomposition

The dehydrogenation/hydrolysis route also requires that formic acid decom-


position, reaction (6), is fast. Most metals are active for this reaction [ 171,
with copper and zinc particularly so [ 17,181. Tests with the Cu/ZnO/AlzOS
catalyst showed that the decomposition of the acid was fast at temperatures
used for methanol steam reforming. At 453 K, for example, the rate of decom-
position was 41.6 mmol kg-is-l, which is about twice as high as the rate of
methyl formate hydrolysis at the same temperature. Although the kinetics of
reaction were not measured in detail, the apparent activation energy, based on
measurements of the rate of production of carbon dioxide at 433,453 and 473
K using a 88% formic acid:12% water in nitrogen feed was calculated to be 70
kJ mol-‘. No carbon monoxide could be detected amongst the products.
Table 2 summarises kinetic parameters obtained for the four reactions over
the Cu/ZnO/Al,O, catalyst. The final row gives reaction rates at 473 K for
reactant partial pressures of approximately 25 kPa. The values are consistent
with the view that steam reforming of methanol proceeds by the dehydroge-
nation/hydrolysis with methanol dehydrogenation, eqn. (4)) rate determining.
The activation energy for this reaction is very similar to that for the overall
reaction and the rate of steam reforming is one half that for dehydrogenation
as required by the overall stoichiometry of the three steps. The later steps, (5)
and (6), are certainly faster in isolation than reaction (4), as required for
reactions subsequent to a rate determining one. However the measured rates
of (5) and (6) seem insufficiently rapid to explain the complete absence of

TABLE 2

Comparison of the kinetic parameters for methanol steam reforming, methanol dehydrogenation,
methyl formate hydrolysis and formic acid decomposition over Cu/ZnO/A120s

Kinetic parameters Reaction”

SR DH HDR FAD

Order in methanol 0.26 0.00 -


Order in water 0.03 0.00 -
Order in methyl formate -0.23 -
lz, (mmol kg-’ s-i kPa_” ) * 5.3.10’2 -1l.l.Io’x 3.9*1013 nm
E. (kJ mol-‘) 105 103 67 71
Rate (mm01 (carbon) kg-‘s-l) c 52 102 380d 220

a SR: steam reforming of methanol, DH: dehydrogenation of methanol, based on the rate when
initially on stream, HDR: hydrolysis of methyl formate, FAD: decomposition of formic acid.
* x Is the sum of kinetic orders.
’ For reactant partial pressures of approximately 25 kPa and a temperature of 473 K.
d Twice the value calculated by eqn. (11) since one mole of methyl formate contains two moles of
carbon.
C.J. Jianget al. /Appl. Catal. A 97(1993) 145-158 155

methyl formate and formic acid as products during steam reforming. The ex-
planation for the absence of these products probably lies in the relative strengths
of adsorption. If this were in the order formic acid > methyl formate > meth-
anol the products of reactions (5 ) and (6) would displace their respective reac-
tants and both would be driven to completion. This is not unreasonable.
The alternative decomposition-shift mechanism, reactions (2) and (3)) can
be ruled out over the present catalyst. The catalyst does have considerable
activity for the water-gas shift reaction, (3), when tested with carbon mon-
oxide/water mixtures alone [ 81. However the conversion of carbon monoxide
was very low when it was added to a methanol/water feed passed over the
catalyst under methanol steam reforming conditions and the overall produc-
tion of carbon dioxide and hydrogen was only slightly dependent on the carbon
monoxide pressure. The absence of water-gas shift activity when methanol is
present is again likely to be a consequence of different adsorption strengths.
Infrared measurements [ 191 show that methanol readily displaces carbon
monoxide from copper under the conditions used here.
Similarly another decomposition-shift mechanism based on the decompo-
sition of formaldehyde according to reaction (12) can be ruled out.
HCHO--+CO+H, (12)
An extensive investigation in this laboratory [lo] of the dehydrogenation
of methanol over copper-based catalysts has shown that the decomposition
reaction (12) does not occur under the conditions of the present investigation.
For example, when a 30% formaldehyde in helium mixture was passed over
two copper-based catalysts at 453 K the major product was methyl formate
which was formed by the dimerization of formaldehyde
2HCHO-CHBOCH0 (13)
At high conversions of formaldehyde equimolar amounts of carbon monox-
ide and methanol were formed as major products by decomposition of methyl
formate
CH,OCHO+CH,OH+CO (14)
These results indicate that formaldehyde decomposition [reaction ( 12 ) ] does
not occur under the conditions of methanol steam reforming.
Because of the low heats of adsorption of carbon dioxide and hydrogen it is
unlikely that the trace quantities of carbon monoxide observed during steam
reforming result from the reverse water-gas shift reaction. It is more probable
that they arise by decarbonylation of methyl formate (reaction 14 ) occurring
parallel with hydrolysis. This reaction is well known in the context of methanol
dehydrogenation [ 9,20,21].
The dehydrogenation/hydrolysis mechanism suggested by this study has been
modelled using the following reaction sequence where * represents an adsorp-
156 C.J. Jiang et al. / Appl. Catal. A 97 (1993) 145-158

tion site on the copper surface:

CH30H 9t CH,O*+H* (15)

CH,O* $ CH,O*+H* (16)


ki

Hz0 3 HzO* (17)

Hz 3
t 2H* (18)

2CHz 0* 3 CH, OCHO* (19)

CH,OCHO* +H,O* k,5HCOOH*+CH,OH (26)


k6
HCOOH* --+ H, +CO, (21)
The experimental results require that reaction steps 18 to 21 are all very
fast, and that the dehydrogenation of methanol is relatively slow. Either re-
action (15) or (16) should thus be the rate determining step. Attempts to
correlate the kinetic data for steam reforming using a standard Langmuir-
Hinshelwood-Hougen-Watson formulation [ 221 showed that only the as-
sumption of reaction (16) as the rate determining step was acceptable. This is
in accord with the observation of a large kinetic isotope effect for methanol
dehydrogenation when C2H30H or C2H,02H is used in place of CH,OH [ 231.
All other assumptions were rejected either due to poor correlation or to nega-
tive parameter estimates. The rate equation then becomes, when K2 is small:
kK K-1’2PMP$‘2
rSR= (,+KlK$P;Pfil,2 +Kr1/2P$/2)2 (22)

where the subscripts M and H refer to methanol and hydrogen respectively.


The kinetic parameters estimated on the basis of this model are presented in
Table 3. The values for K3 rely mainly on experiments with added hydrogen
[8] while those for Kl are based on all experiments. The activation energy
obtained from the values of k is 110+5 kJ/mol in good agreement with the
experimental value of 105 kJ/mol [eqn. (7) 1. Fig. 7 shows that eqn. (22) pro-
vides a good fit to the rate of steam reforming of methanol under the conditions
studied.
Approximate values for the heats of adsorption of methanol and of hydrogen
on the catalyst can be determined by applying the van ‘t Hoff equation to the
adsorption coefficients obtained for the proposed model. Heats of adsorption
of 25 ? 5 and 33 2 5 kJ/mol for methanol and the hydrogen respectively can be
calculated. The value of the heat of adsorption of hydrogen is in reasonable
C.J. Jiang et al. / Appl. Catal. A 97 (1993) 145-158 157

TABLE 3

Langmuir-Hinshelwood parameter estimates for methanol steam reforming over Cu/ZnO/AlxOs


according to eqn. (22 )

Temperature k & K3 Correlation


(K) (mmol (CH,OH) kg-’ a-‘) (kPa_' ) (kPa- ’ ) coefficient

443 32 6.36.10-3 9.5.10-3 0.990


463 69 4.59*10-3 4.2.10-3 0.982
513 174 2.79.10-3 2.6~10-~ 0.992
523 226 2.13*10-3 2.1.10-3 0.998

r SR (mmol/kg/sec) (measured)

Fig. 7. Predicted rats of methanol steam reforming [ eqn. (22) ] versus experimental rate.

agreement with that reported by Salmi and Hakkaranian [24] (32.8 kJ/mol
for a Cu/ZnO catalyst) and by Eley and Rossington [ 22,251 (33.5 kJ/mol for
copper films ). It is because of this weak adsorption that hydrogen inhibits the
steam reforming reaction only to an kinetic order of approximately - 0.2 [ 81.

CONCLUSIONS

1. A Cu/ZnO/Alz03 catalyst exhibits high and stable activity for the steam
reforming of methanol, producing carbon dioxide and hydrogen in high
selectivity.
2. In the absence of water the catalyst dehydrogenates methanol to methyl
formate at a rate initially faster than that of steam reforming. Dehydrogena-
tion is accompanied by deactivation, probably attributable to fouling by form-
aldehyde based polymers.
3. Hydrolysis of methyl formate and formic acid decomposition are faster
than steam reforming or dehydrogenation of methanol over Cu/ZnO/AlzO,
158 C.J. Jiang et at. / Appl. Catal. A 9?(1993) 145-158

4. Steam reforming of methanol over Cu/ZnO/A1203 can he explained in


terms of: methanol dehydrogenation, methyl formate hydrolysis and formic
acid decomposition. Rate data for steam reforming can be interpreted accord-
ing to this scheme using a Langmuir-Hinshelwood-Hougen-Watson model in
which the rate determining step is the abstraction of hydrogen from an ad-
sorbed methoxy group.

ACKNOWLEDGEMENT

The financial support by the Australian Research Council is gratefully


acknowledged.

REFERENCES

1 H. Kobayashi, N. Takezawa and C. Minochi, Chem. Lett., (1976) 1347.


2 J.C. Amphlett, M.J. Evans, R.F. Mann and R.D. Weir, Can. J. Chem. Eng., 63 (1985) 605.
3 E. Santacesaria and S. Carra, Appl. Catal., 5 (1983) 345.
4 Y. Okamoto, Y. Konishi, K. Fukino, T. Imanaka and S. Teranishi, Proceedings 8th Inter-
national Congress on Catalysis, Berlin, July 1984, VerIag-Chemie, Weinheim, 1984, Vol. V,
p. 159.
5 V. Pour, J. Barton and A. Benda, Coll. Czech. Chem. Commun., 40 (1975) 2923.
6 F. Mashner and F.W. Moeller, in B.E. Leach (Editor), Methanol Synthesis, Applied Indus-
trial Catalysis, Vol. 2, Academic Press, 1983, p. 234.
7 V.N. Ipatieff and V. Haensel, J. Org. Chem., 7 (1942) 189.
8 C.J. Jiang, D.L. Trimm, M.S. Wainwright and N.W. Cant, Appl. Catal., 93 (1993) 245.
9 S.P. Tonner, D.L. Trimm, M.S. Wainwright and N.W. Cant, Ind. Eng. Chem. Prod. Res.
Dev., 23 (1984) 3.
10 S.P. Tonner, PhD Thesis, University of New South Wales, (1984).
11 K. Takahashi, N. Takezawa and H. Kobayashi, Appl. CataI., 2 (1982) 383.
12 J.W. Evans, M.S. Wainwright, A. Bridgewater and DJ. Young, Appl. Catal., 7 (1983) 75.
13 D.M. Monti, N.W. Cant, D.L. Trimm and M.S. Wainwright, J. Catal., 100 (1986) 17.
14 0. Levenspiel, J. Catal., 25 (1972) 265.
15 A. Wheeler, Adv. Catal., 2 (1955) 105.
16 N.I. Onuoha and M.S. Wainwright, Chem. Eng. Commun., 29 (1984) 13.
17 G.C. Bond, Catalysis by Metals, Academic Press, London, 1962, p. 412.
18 A. Ueno, T. Onishi and K. Tamaru, Trans. Faraday Sot., 66 (1970) 756.
19 K.J. Soerensen and N.W. Cant, unpublished results.
20 B.W. Higdon, C.C. Hobbs and M.U. Onore, US Patent 3 812 210, (1974).
21 T. Ushikubo, H. Hattori and K. Tanabe, Chem. Lett., (1984) 649.
22 J.M. Thomas and W.J. Thomas, Introduction to the Principles of Heterogeneous Catalysis,
Academic Press, London, 1967.
23 N. W. Cant, S.P. Tonner, D.L. Trimm and M.S. Wainwright, J. Catal., 91 (1985) 197.
24 T. SaImi and R. Hakkarainen, Appl. Catal., 49 (1989) 285.
25 D.D. Eley and D.R. Rossington, in W.E. Garner (Editor), Chemisorption, Butterworths,
London, 1957, p. 137.

You might also like