Вы находитесь на странице: 1из 10

Chemical Engineering and Processing 44 (2005) 393–402

The dehydrogenation of methanol to methyl formate


Part I: Kinetic studies using copper-based catalysts
X. Huang, N.W. Cant, M.S. Wainwright∗ , L. Ma
School of Chemical Engineering and Industrial Chemistry, The University of New South Wales (UNSW),
Sydney 2052, Australia

Received 5 September 2003; accepted 26 May 2004


Available online 20 August 2004

Abstract

Kinetics of the dehydrogenation of methanol to methyl formate (MF) have been determined for a commercial copper-chromite catalyst and
for a skeletal copper catalyst that has undergone deactivation to a steady-state activity. The activation energy over the copper-chromite catalyst
was found to be around 78 kJ/mol, considerably lower than the approximately 120 kJ/mol observed for the skeletal copper catalyst. The reaction
order with respect to methanol was found to be approximately 0.5 for both catalysts whilst hydrogen and methyl formate both inhibited the
reaction significantly. This inhibition is consistent with a Langmuir–Hinshelwood model in which methanol is adsorbed dissociatively and the
rate-determining step is the loss of hydrogen from the resultant methoxy species. The model parameter values imply that a significant fraction
of the copper sites on the skeletal catalyst are occupied by formaldehyde while the coverage is low on copper-chromite. These differences can
be used to explain the presence of deactivation and the considerably higher activation energy for dehydrogenation observed for the skeletal
copper catalyst.
© 2004 Elsevier B.V. All rights reserved.

Keywords: Kinetics; Methanol dehydrogenation; Methyl formate; Skeletal copper catalyst; Copper-chromite catalyst

1. Introduction copper system improves the structure of skeletal copper and


thereby promotes the stability of the catalysts for dehydro-
Dehydrogenation of methanol to methyl formate (MF): genation of methanol [8]. The improvement of performance
was attributed to the presence of Cr2 O3 on the surface of cop-
2CH3 OH → CH3 OCHO + 2H2 (1) per, which minimised polymerization of adsorbed species on
is important as a first step in the synthesis of acetic acid and active copper sites [8]. However, detailed support for this
formamide [1]. Copper-based catalysts appear to be uniquely interpretation was lacking.
effective. Extensive research has been conducted to improve This study addresses the kinetics of the gas phase dehydro-
catalyst activity and selectivity [2–4]. genation of methanol to methyl formate over a significantly
Amongst copper-based catalysts, pure skeletal copper cat- deactivated (stable) skeletal copper catalyst and a commer-
alysts exhibit very high initial activity for the reaction [6]. cial copper-chromite catalyst, as a way to elucidate the role
However, rapid deactivation is observed with two-thirds of of chromia in improving the stability of copper catalysts.
the initial activity lost due to fouling attributed to polymeriza-
tion of formaldehyde on the surface of the copper [7]. Recent
studies have shown that the addition of Cr2 O3 to the skeletal 2. Experimental

The gas phase dehydrogenation reaction (1) was con-


∗ Corresponding author. Tel.: +61 2 93852700; fax: +61 2 93858008. ducted in a conventional stainless steel micro-reactor
E-mail address: m.wainwright@unsw.edu.au (M.S. Wainwright). system operated at atmospheric pressure. Methanol was in-

0255-2701/$ – see front matter © 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.cep.2004.05.012
394 X. Huang et al. / Chemical Engineering and Processing 44 (2005) 393–402

troduced to the reactor by passing helium through saturators The range of conditions explored were: 448–468 K, GHSV
(inert material) connected in series. Methyl formate was in- 150,000–200,000 h−1 , mol% methanol from 2 to 10%, added
troduced when required in the same way via separate satura- methyl formate from 0 to 5% and added hydrogen from 0 to
tors. The gas flow were controlled by mass flow controllers 10%.
(MFCs). The products were analyzed using two on-line gas Rates were calculated using the differential reactor ap-
chromatographs (GCs) equipped with thermal conductivity proximation:
detectors and a Porapak Q column in GC-1 (for analysis of X
methanol and methyl formate) and a CTR-1 column in GC-2 −r = Fin (2a)
W
(for CO, CO2 and H2 ). An ice-bath between GC-1 and GC-2
was used to condense liquid products from the gas stream where −r is the reaction rate of methanol dehydrogenation,
to improve the performance of the CTR-1 column. Prior to mol/h g-cat; Fin flow rate of methanol entering the reactor,
the kinetic measurements a blank test was performed with mol/h; X the fractional conversion of methanol and W the
␣-Al2 O3 (the diluent used in the normal runs) over the tem- mass of catalyst, g.
perature range used for the catalytic measurements. There In experiments without added products, or with hydrogen
was no detectable reaction due to the reactor walls. alone added, the conversion was calculated from the compo-
In each catalytic experiment, approximately 0.03 g of sition of the outlet stream in mol% (Y), i.e.
the skeletal copper, prepared as described in [5], or 0.06 g 2YCH3 OCHO + YCO2
of Harshaw-type 0203 copper-chromite catalyst particles X= (2b)
YCH3 OH + 2YCH3 OCHO + YCO2
(211–325 ␮m) were diluted with six times the volume of ␣-
Al2 O3 of the same particle size in order to ensure bed isother- With methyl formate added, the rate was calculated di-
micity and charged to the constant temperature zone of the rectly from the difference between methanol entering and
stainless steel reactor. The ratio of bed length to diameter was leaving the reactor, Fin − Fout , with the outlet molar flow
greater than 7. Prior to use, the catalysts were slowly reduced obtained from the corresponding values for YCH3 OH with ad-
in stages in hydrogen starting from room temperature with a justment for small differences in volumetric flow rate.
hold for 2 h at 403 K and a final period of 4 h at 513 K.
In order to efficiently measure the kinetics of methanol
dehydrogenation over skeletal copper, it was necessary to 3. Results
stabilize the catalyst. Tonner et al. [6] showed that skeletal
copper was highly active for the reaction but rapidly deacti- 3.1. Thermodynamic analysis
vated due to the fouling. In this study, we have allowed the
copper surface to be fouled by the polymers until a stable Under the low conversion conditions used here, the selec-
steady state operating condition was obtained. The pretreat- tivity of the conversion of methanol to methyl formate and
ment conditions used to stabilize the skeletal copper cata- hydrogen was always >90% with CO2 the only significant
lyst were: reaction temperature: 468 K, time on stream: 2 h, by-product. The latter is probably formed by the methanol
methanol vapor flow rate: 10 cm3 /min. After the process of steam reforming reaction as a result of the trace amounts of
quick deactivation, the activity was significantly lower than water in the feed.
that obtained on a fresh catalyst but the stable skeletal cop- The main reaction (1) is equilibrium limited. Table 1
per catalyst enabled kinetic measurements to be made with summarises the equilibrium conversions, Xe , for the
minimal further change in activity. extremes of the conditions used here. The equilib-
Kinetic measurements were made under steady-state con- rium constant was calculated using standard methods
ditions with a total pressure of around 115 kPa and conver- for ideal gases. The value for ∆Hfo (CH3 OCHO)
sions below 15%. The total flow rate through the reactor was was taken from the measurements of Hall and Baldt
maintained constant using He as diluent gas and the flow [9], and that for methanol from Chao et al. [10]
rates were high enough to avoid mass transfer limitations. and for hydrogen from the tabulations of Barin [11].

Table 1
Equilibrium conversions calculated for methanol dehydrogenation
T (K) Ka Initial partial pressure (kPa) Conversions (%) Xe /Xmax
CH3 OH CH3 OCHO H2 Xe Xmax b
448 0.015 10.1 0.0 0.0 45.0 3.8 0.08
448 0.015 10.1 2.2 0.0 37.8 0.8 0.02
448 0.015 10.1 0.0 11.0 14.3 0.1 0.01
468 0.027 10.1 0.0 0.0 50.9 11.2 0.22
468 0.027 10.1 4.2 0.0 39.9 1.2 0.03
468 0.027 10.1 0.0 11.0 20.3 2.8 0.14
a The equilibrium constants are for reaction (1) as written (i.e. with 2 mol of CH3 OH).
b Maximum conversions obtained when using the copper-chromite catalyst.
X. Huang et al. / Chemical Engineering and Processing 44 (2005) 393–402 395

Fig. 1. log–log plots of the rates of methanol consumption vs. methanol partial pressure for skeletal copper (a) and copper-chromite (b) catalysts.

Fig. 2. Arrhenius plots of the rate constants for methanol dehydrogenation using power law equation (Eq. (3b)) over skeletal copper () and copper-chromite
(䊉).
396 X. Huang et al. / Chemical Engineering and Processing 44 (2005) 393–402

Table 2
Power law kinetic parameters for methanol dehydrogenation
Catalyst T (K) k × 100 (mol/h g-cat kPa−␣ ) A (mol/h g-cat kPa−␣ ) Ea (kJ/mol) α
453 1.27
Skeletal copper 458 1.96 4.2 × 1011 117 ± 10 0.50 ± 0.02
463 2.71
468 3.45
448 1.13
Copper-chromite 1.3 × 107 78 ± 1.5 0.46 ± 0.01
458 1.81
468 2.76

The equilibrium conversion was then obtained by 468 K. Thus, the conversion reached did not exceed 22% of
solution of the cubic expression for the equilibrium cons- the equilibrium value, with most values much less, and thus
tant. the differential reactor approximation used to obtain rates is
The values for Xmax are the maximum conversions reached justified. With products in the feed equilibrium conversions
in the present work for the conditions shown. As may be seen, fell substantially (to 14% when 11 kPa of H2 were added at
Xe is in the range 45–51% when no products are added com- 448 K). However, as described later, the rate declined more
pared with a maximum experimental conversion of 11.2% at steeply than this so the reaction was operated further from

Fig. 3. Plots of calculated rate using the power law model (Eq. (3)) vs. observed rate for methanol dehydrogenation over skeletal copper (a) and copper-chromite
(b) catalysts.
X. Huang et al. / Chemical Engineering and Processing 44 (2005) 393–402 397

Table 3
Summary of Langmuir–Hinshelwood models in terms of reaction steps
Model CH3 OH adsorption Rate-determining step Route to methyl formate
1 Molecular CH3 OH* + * → CH3 O* + H* 2CH2 O* → CH3 OCHO* + *
2 Molecular CH3 OH + * → CH3 OH* 2CH2 O* → CH3 OCHO* + *
3 Molecular CH3 OCHO* → CH3 OCHO + * 2CH2 O* → CH3 OCHO* + *
4 Molecular H* + H* → H2 + 2* 2CH2 O* → CH3 OCHO* + *
5 Molecular CH3 OH* + * → CH3 O* + H* CH2 O + CH2 O* → CH3 OCHO*
6 Molecular CH3 OH* + * → CH3 O* + H* CH3 O* + CHO* → CH3 OCHO* + *
7 Molecular 2CH2 O* → CH3 OCHO* + * 2CH2 O* → CH3 OCHO* + *
8 Molecular CH3 O* + * → CH2 O* + H* CH3 O* + CHO* → CH3 OCHO* + *
9 Dissociative CH3 O* + * → CH2 O* + H* 2CH2 O* → CH3 OCHO* + *
10 Dissociative CH3 O* + * → CH2 O* + H* CH2 O + CH2 O* → CH3 OCHO*
11 Molecular CH3 O* + * → CH2 O* + H* CH2 O + CH2 O* → CH3 OCHO*
12 Molecular CH3 O* + * → CH2 O* + H* CH2 O* + CH3 OH* → CH3 OCHO + 2H*
13 Molecular CH3 O* + * → CH2 O* + H* 2CH2 O* → CH3 OCHO* + *
14 Molecular CH3 O* + * → CH2 O* + H* CH3 OCH2 O* + * → CH3 OCHO* + H*
15 Dissociative CH3 O* + * → CH2 O* + H* CH3 OCH2 O* + * → CH3 OCHO* + H*
16 Dissociative CH3 O* + * → CH2 O* + H* CH3 OCH2 O* + CH2 O* → CH3 OCHO* + CH3 O*
* Indicates bound to surface.

Fig. 4. Effects of methyl formate partial pressure on the rate of dehydrogenation of methanol (ca. 8.0 kPa) over skeletal copper (a) and copper-chromite (b)
catalysts.
398 X. Huang et al. / Chemical Engineering and Processing 44 (2005) 393–402

Fig. 5. Effects of hydrogen partial pressure on the rate of dehydrogenation of methanol (ca. 10 kPa) over skeletal copper (a) and copper-chromite (b) catalysts.

equilibrium when products were added than when they were where k is the apparent rate constant in the power law model,
not. mol/h g-cat kPa−α ; A the pre-exponential factor, mol/h g-cat
kPa−α ; Ea the activation energy, kJ/mol; PCH3 OH the par-
tial pressure of methanol, kPa; and α the reaction order of
3.2. Kinetic studies methanol.
The influence of methanol over both catalysts was ex-
An empirical power law expression (Eq. (3)) was used to plored by changing the partial pressure. The plots used to
assess kinetic parameters. determine reaction order with respect to methanol are shown
α in Fig. 1. The slopes are all close to 0.5. Similar behavior
−r = kPCH3 OH
(3a)
has been reported by Ai [12], who varied the initial concen-
−r = Ae−Ea /RT PCH
α
(3b) tration of methanol from 1.6 to 16.5 vol.%, while fixing the
3 OH other reaction conditions.
Table 4
Langmuir–Hinshelwood model kinetic parameters for methanol dehydrogenation over the skeletal copper catalyst
T (K) k (mol/h g-cat) K1 × 102 (kPa−1 ) K3 K4 (kPa) K5 (kPa) R2 Ea (kJ/mol)
458 0.52 9.8 3.66 0.53 0.50 0.97 127 ± 9.4
463 0.78 0.9 0.09 12.53 9.41 0.95
468 1.06 0.2 0.47 8.03 113.7 0.96
R2 is the regression coefficient for the fit to the linearised rate expression.
X. Huang et al. / Chemical Engineering and Processing 44 (2005) 393–402 399

Table 5
Formaldehyde coverage calculated using kinetic model parameters for the skeletal copper catalyst
T (K) Feed methanol partial Feed methyl formate Feed hydrogen partial Coverage by
pressure (kPa) partial pressure (kPa) pressure (kPa) formaldehyde (%)
463 9.30 0.00 0.00 16.0
463 9.30 0.00 5.00 7.91
463 8.37 4.00 0.00 39.0
468 9.30 0.00 0.00 16.1
468 9.30 0.00 5.00 9.91
468 8.37 4.00 0.00 34.5

Arrhenius plots for the reaction rate constants are shown in 3.4. Mechanistic considerations
Fig. 2. The slopes, and hence the activation energies are quite
different for the two catalysts. The values for the constants in The inhibition data in Figs. 4 and 5 were tested jointly with
the power law Eqs. (3a) and (3b) are listed in Table 2. A likely those for the dependence on methanol in Fig. 1 for conformity
reason for the difference in activation energy is discussed to 16 Langmuir–Hinshelwood-type models listed in Table 3.
later. They differed according to the mode of methanol adsorption
Fig. 3(a–b) is a plot of the rates calculated using the power (molecular or dissociative), the rate-determining step (con-
law equation versus the observed rates. They show that the version of methanol to a methoxy species, dehydrogenation
power law model fits the experimental data very well for both of methoxy to formaldehyde, methyl formate formation or
catalysts. desorption) and the chemistry of methyl formate formation
(dimerisation of adsorbed CH2 O, CH3 O + CHO or CH3 OH
+ CH2 O).
3.3. Effect of products
Initial testing was carried out with linearised versions of
each rate expression using the Excel function Linest. Most
The effect of the principal products of the dehydrogena-
models could be excluded immediately because they gave
tion of methanol on the reaction rate was also investigated.
one or more negative adsorption coefficients, sometimes in
As noted previously, the presence of additional methyl for-
combination with a negative rate constant, both of which are
mate or hydrogen reduces the equilibrium conversion but the
physically unreasonable. In the case of the skeletal catalyst,
observed effect on the rate was greater still since, as shown
the only model giving realistic solutions was as follows:
in Table 1, the maximum conversions obtained experimen-
K1
tally were further from equilibrium when either product was CH3 OH(g) + 2∗ ←→ CH3 O∗ + H∗ (4a)
included in the feed.
k
The effect of added methyl formate on rate is illustrated CH3 O∗ +∗ −→ CH2 O∗ + H∗ (4b)
in Fig. 4. With both catalysts, the apparent order is −0.5
K3
within the limits of the accuracy of the measurements. The 2CH2 O∗ ←→ CH3 OCHO∗ +∗ (4c)
negative order implies that methyl formate is adsorbed more K4
strongly on copper than is methanol. This agrees with the CH3 OCHO∗ ←→ CH3 OCHO(g)+∗ (4d)
conclusions of previous studies of the dehydrogenation of K5
higher alcohols to aldehydes or ketones over copper where H∗ + H∗ ←→ H2 (g) + 2∗ (4e)
the adsorption coefficients for the carbonyl compounds are This scheme assumes that methanol is adsorbed disso-
several times those for the alcohols [13]. ciatively with the subsequent conversion of the resultant
The retarding effect of hydrogen on the rate of methanol methoxy species to adsorbed formaldehyde as the rate-
dehydrogenation is shown in Fig. 5. The inhibition is some- determining step. The rate expression is:
what greater than for methyl formate with apparent kinetic
−1/2
orders of −0.8 for skeletal copper and −0.7 for copper- 1/2
kK1 K5 PCH3 OH PH2
chromite. Such inhibition has been previously noted for a −r = −1/2
(5)
+ K4−1 PCH3 OCHO
1/2
Cu/ZnO catalyst although the authors did not rule out a ther- (1 + K1 K5 PCH3 OH PH2
−1/2 −1/2 1/2 −1/2 −1/2 2
modynamic explanation [14]. +K3 K4 PCH3 OCHO + K5 PH2 )

Table 6
Langmuir–Hinshelwood model kinetic parameters for methanol dehydrogenation over the commercial copper-chromite catalyst
T (K) k (mol/h g-cat) K1 × 102 (kPa−1 ) K4 (kPa) K5 (kPa) R2 Ea (kJ/mol)
448 0.28 0.03 1.55 643 0.99 89 ± 9.5
458 0.43 0.48 0.60 30 0.99
468 0.78 0.22 1.25 104 0.99
R2 is the regression coefficient for the fit to the linearised rate expression.
400 X. Huang et al. / Chemical Engineering and Processing 44 (2005) 393–402

Fig. 6. Plots of calculated rate using Langmuir–Hinshelwood model (Eq. (5)) vs. observed rate for methanol dehydrogenation over skeletal copper (a) and
copper-chromite (b) catalysts.

Table 4 shows the calculated values for the five parame- fouling of the surface by polymerized formaldehyde. Further-
ters in the scheme with the fit between predicted and observed more, the presence of methyl formate in the feed significantly
rates given in Fig. 6(a). The average deviation overall is ca. increases the coverage by formaldehyde. On the other hand,
14%. This is rather good bearing in mind that the concen- the presence of hydrogen in the feed halves the coverage by
trations of methyl formate and hydrogen vary by almost two formaldehyde. The results indicate that the deactivation of the
orders of magnitude between the lowest produced by reaction skeletal copper catalyst appears to be associated with surface
and the maximum included in the feed. coverage by formaldehyde and/or polymers derived from it.
A slightly better fit (average deviation 12%) could be ob- These results are consistent with the experimental findings
tained by a further iterative non-linear refinement using the of Evans, who showed that the deactivation of skeletal cop-
Excel tool Solver. However, a statistics package operated in per catalyst could be reversed by passing hydrogen instead of
conjunction with Excel [15] did indicate that some of the methyl formate over the deactivated catalyst during a study
parameters had quite large uncertainties. Thus, the values in of methyl formate hydrogenolysis [16].
Table 4 should be regarded as indicative rather than accurate By contrast, the data for the copper-chromite catalyst gave
Langmuir–Hinshelwood coefficients. The parameter values physically reasonable parameters according to the above rate
in Table 4 were used to calculate the coverage of formalde- expression, Eq. (5), only if the second last term in the denomi-
hyde on the surface. Table 5 shows that at 463 and 468 K, the nator, which arises algebraically from coverage of the surface
coverage of formaldehyde is more than 15%, indicating that by formaldehyde, was very small (i.e. K3 was very large). The
the severe deactivation observed for skeletal copper is due to improvement to the stability for methanol dehydrogenation
X. Huang et al. / Chemical Engineering and Processing 44 (2005) 393–402 401

over the copper-chromite catalyst, which was studied recently hyde to form methyl formate [24]:
[8], can be explained by the presence of chromium oxide re-
stricting formation of a polymer that can bind to the surface of CH3 O∗ + CH2 O∗ → CH3 OCH2 O∗ +∗ (6a)
active copper. Making the assumption that this term (CH2 O) CH3 OCH2 O∗ + HCHO∗ → CH3 OCHO∗ + CH3 O∗ (6b)
could be neglected gave the parameters shown in Table 6 and
the good fit between calculated and observed rates shown in Substitution of this route in place of step (4c) in the earlier
Fig. 6(b). scheme led to a rate expression that did not fit the present data
The data for the copper-chromite catalyst could be equally for either catalyst with all parameters positive.
well fitted to a second model in which methanol initially ad- The most interesting aspect of the modeling is the infer-
sorbed as molecules with the conversion to a methoxy species ence of significant amounts of formaldehyde on the surface of
being rate determining and with the additional assumption the skeletal catalyst but with negligible amounts for copper-
that coverage of sites by both methoxy and formaldehyde chromite. The former is consistent with the observation of
species was very low. This model could be excluded on the Tonner et al. [6] that trace amounts of gaseous formalde-
basis that it was not compatible with the existence of a large hyde are seen as a product with a skeletal catalyst when it is
deuterium kinetic isotope effect which shows that C H bond substantially deactivated. The presence of formaldehyde can
breaking is rate-determining over this catalyst [17]. It also provide an explanation for the existence of deactivation with
required high hydrogen coverage to duplicate the observed the skeletal catalyst and none for the chromite one. If cover-
inhibition. That is not reasonable given the weakness of hy- age by formaldehyde is higher on the extended copper planes
drogen adsorption. It may be noted here that in the scheme on the skeletal catalyst then the probability of polymerisation
described by Eqs. 4(a) and 4(e), inhibition by hydrogen arises to foulant will be enhanced and deactivation will be faster.
primarily from its mass action effect on step (4a) which re- Infrared studies using Cu/SiO2 are consistent with the con-
duces the concentration of the surface methoxy species. version of formaldehyde to polyoxymethylene and this is di-
From Fig. 6(a–b), it can be seen that good correlations rectly associated with deactivation [25]. On the other hand,
were obtained for both catalysts. Tables 4 and 6 showed if chromia were to scavenge gaseous formaldehyde, then the
good fit to the Arrhenius plots of rate constants using the probability and extent of fouling and deactivation would be
Langmuir–Hinshelwood model Eq. (5) and the activation en- reduced and the catalyst should be more stable as observed.
ergies of the reaction were close to the results which were The higher activation energy observed with the skele-
found using the power law equation for skeletal copper and tal catalyst can be rationalised in a similar way. If
copper-chromite catalysts respectively. formation of foulant is to some extent a reversible
polymerization–depolymerisation process, then increasing
temperature could lead to less coverage of the surface by
4. Discussion foulant and an increased number of active sites. In this case,
one would expect the apparent activation energy to be higher
The reaction scheme above is consistent with other knowl- for the skeletal catalyst, due to an increase in the number of
edge. Methanol is readily adsorbed as methoxy groups on active sites with temperature, than for the chromite system
copper surfaces [18,19] and they have also been observed by where this effect is significantly reduced or absent.
infrared spectroscopy on supported copper catalysts [20–22].
Reversal of the adsorption facilitates exchange between D2
and the OH group of CH3 OH and this occurs at a rate that 5. Conclusions
is much faster than that of methanol dehydrogenation over
both the catalysts studied here. By comparison, deuterium Kinetic studies have shown that the dehydrogenation of
exchange into the methyl group is considerably slower than methanol over copper-based catalysts is near half order
that of dehydrogenation and the existence of a large deu- with respect to the partial pressure of methanol. The ac-
terium kinetic isotope effect when the CH3 group in normal tivation energy is ca. 120 kJ/mol for skeletal copper and
methanol is substituted by CD3 indicates that breaking of a ca. 78 kJ/mol for a commercial copper-chromite catalyst.
C H bond is rate determining during dehydrogenation [17]. Methyl formate and hydrogen, the reaction products, sig-
The formation of methyl formate by the formaldehyde nificantly inhibit dehydrogenation. The kinetic data can
dimerisation (the Tischenko reaction, 4c) is also consistent be fitted to a Langmuir–Hinshelwood model in which the
with suggestions from isotope studies at low pressure [23] and rate-determining step is the dissociation of a C H bond
with catalysts similar to those used here [17]. In the latter case, in a methoxy species formed by dissociative adsorption of
the interpretation was based on detailed arguments concern- methanol with methyl formate then formed by a Tischenko-
ing the deuterium distribution in the methyl formate formed type reaction between two formaldehyde molecules. The
by reaction of CH3 OH/CD3 OD mixtures and may not be to- implied coverage by formaldehyde is significant for the
tally definitive. Other authors have favoured a hemi-acetal skeletal copper but negligible with copper-chromite cata-
route involving the combination of adsorbed methoxy and lyst. The difference can explain the initial deactivation seen
formaldehyde followed by reacting with adsorbed formalde- with the skeletal catalyst and can also account for the higher
402 X. Huang et al. / Chemical Engineering and Processing 44 (2005) 393–402

activation energy for methanol dehydrogenation over this cat- [10] J. Chao, K.R. Hall, K.N. Marsh, R.C. Wilholt, Thermodynamic prop-
alyst. erties of key organic oxygen compounds in the carbon range C1 to
C4. Part 2. Ideal gas properties, J. Phys. Chem. Ref. Data 15 (1986)
1369.
[11] I. Barin, Thermochemical Data of Pure Substances, Part I, VCH
Acknowledgements
Press, p. 640.
[12] M. Ai, Dehydrogenation of methanol to methyl formate over copper-
Financial support of an Australian Postgraduate Award based catalysts, Appl. Catal. 11 (1984) 259.
(APA) for Xinwei Huang is gratefully acknowledged. We [13] G. Ertl, H. Knozinger, J. Weitkamp, Handbook of Heterogeneous
gratefully acknowledge the on-going support of the Aus- Catalysis, VCH, Weinheim, Germany, 1997.
[14] M.J. Chung, D.J. Moon, K.Y. Park, S.K. Ihm, Mechanism of methyl
tralian Research Council.
formate formation on Cu/ZnO catalyst, J. Catal. 136 (1992) 609.
[15] E.J. Billo, Excel for Chemists, Wiley, New York, 2001.
[16] J.W. Evans, Studies of the copper catalysed hydrogenolysis of
References alkyl esters, Ph.D. Thesis, The University of New South Wales,
1983.
[1] T. Abe, Y. Nishide, N. Muro, H. Higuchi, Process for producing [17] N.W. Cant, S.P. Tonner, D.L. Trimm, M.S. Wainwright, Isotopic
carboxylic acid esters and formamide, J 91-345190, Mitsubishi Gas labeling studies of the mechanism of dehydrogenation of methanol
Chemical Co. Inc., Japan. to methyl formate over copper-based catalysts, J. Catal. 91 (1985)
[2] H. Yamashita, T. Kaminade, M. Yoshikawa, T. Funabiki, S. Yoshida, 197.
Amorphous Cu67Ti33 powder alloy as a new catalyst for dehydro- [18] I.E. Wachs, R.J. Madix, The selective oxidation of CH3 OH to H2 CO
genation of methanol to methyl formate C1 , Mol. Chem. 1 (1986) on a Cu(1 1 0) catalyst, J. Catal. 53 (1978) 208.
491. [19] M. Bowker, R.J. Madix, XPS, UPS and thermal desorption studies of
[3] A. Guerrero-Ruiz, I. Rodriguez-Ramos, G. Fierro, Dehydrogenation alcohol adsorption on copper(1 1 0): methanol, Surf. Sci. 95 (1980)
of methanol to methyl formate over supported copper catalysts, Appl. 190.
Catal. 72 (1991) 119. [20] D.B. Clarke, D.K. Lee, M.J. Sandoval, A.T. Bell, Infrared studies
[4] T.P. Minyukova, N.V. Shtertser, L.P. Davydova, I.I. Simentsova, of the mechanism of methanol decomposition on Cu/SiO2 , J. Catal.
A.V. Khasin, T.M. Yurieva, CO-free methyl formate from methanol: 150 (1994) 81.
the control of the selectivity of the process on Cu-based catalysts, [21] G.J. Millar, C.H. Rochester, K.C. Waugh, Infrared study of the ad-
Khimiya v Interesakh Ustoichivogo Razvitiya 11 (2003) 189. sorption of methanol on oxidized and reduced copper/silica catalysts,
[5] L. Ma, M.S. Wainwright, Development of skeletal copper-chromia J. Chem. Soc., Faraday Trans. 87 (1991) 2795.
catalysts. I. Structure and activity promotion of chromia on skeletal [22] I.A. Fisher, A.T. Bell, A mechanistic study of methanol decompo-
copper catalysts for methanol synthesis, Appl. Catal. A 187 (1999) sition over Cu/SiO2 , ZrO2 /SiO2 , and Cu/ZrO2 /SiO2 , J. Catal. 184
89. (1999) 357.
[6] S.P. Tonner, D.L. Trimm, M.S. Wainwright, N.W. Cant, Dehydro- [23] E. Miyazaki, I. Yasumori, Kinetics of the catalytic decomposition
genation of methanol to methyl formate over copper catalysts, Ind. of methanol, formaldehyde, and methyl formate over a copper-wire
Eng. Chem. Prod. Res. Dev. 23 (1984) 384. surface, Bull. Chem. Soc. Jpn. 40 (1967) 2012.
[7] S.P. Tonner, The copper-catalysed dehydrogenation of methanol, [24] K. Takahashi, K. Takezawa, H. Kobayashi, Mechanism of formation
Ph.D. Thesis, University of New South Wales, 1984. of methyl formate from formaldehyde over copper catalysts, Chem.
[8] L. Ma, M.S. Wainwright, in: D. Morrell (Ed.), Catalysis of Organic Lett. (1983) 1061.
Reactions-Chemical Industries, Marcel Dekker, New York, 2002, p. [25] D.M. Monti, N.W. Cant, D.L. Trimm, M.S. Wainwright, Hy-
225. drogenolysis of methyl formate over copper on silica. II. Study
[9] H.K. Hall, J.H. Baldt, Thermochemistry of strained-ring bridgehead of the mechanism using labeled compounds, J. Catal. 100 (1986)
nitriles and esters, J. Am. Chem. Soc. 93 (1971) 140. 17.

Вам также может понравиться