Sunteți pe pagina 1din 12

Applied Catalysis A: General 238 (2003) 211–222

Formaldehyde synthesis from methanol over silver catalysts


Min Qian∗ , M.A. Liauw, G. Emig
Institute of Chemical Technology I, University Erlangen-Nuremberg, Egerlandstr. 3, D-91058 Erlangen, Germany
Received 5 February 2002; received in revised form 11 March 2002; accepted 10 June 2002
Dedicated to Professor J. Weitkamp on the occasion of his 60th birthday

Abstract
The formaldehyde synthesis from methanol was investigated over a polycrystalline silver catalyst at temperatures up to
993 K. Water was added to the feed (water ballast process) like in the commercial BASF process. The conversion of methanol
and the selectivity to formaldehyde appeared to increase with respect to the methanol ballast process with no added water. A
long-time experiment was carried out lasting over more than 300 days time-on-stream. While the methanol conversion does
not change significantly, there is a pronounced change in hydrogen and CO2 selectivity. A most noteworthy observation is
that over months of operation, the width of the temperature region where formic acid is formed increases in a linear manner.
Finally, interrupting the oxygen supply for a few hours led to a temporary deterioration of the product selectivity after oxygen
re-admittance. All observations may well be interpreted in the framework of the commonly discussed silver–oxygen chemistry
with its three different oxygen species (O␣ , O␤ and O␥ ).
© 2002 Elsevier Science B.V. All rights reserved.
Keywords: Methanol oxidation; Oxidative dehydrogenation; Formaldehyde; Silver catalyst; Catalyst deactivation; Oxygen species

1. Introduction route is performed over an electrolytic silver catalyst


at air-lean conditions, which has been commercial-
With an ever increasing production of (2.5–2.7) × ized at the beginning of the 20th century. The second
107 t per year worldwide in 2000 for the production of route, industrialized since the 1950s, is by the oxi-
urea–phenolic, melamine and acetal resins, formalde- dation in excess air over a ferric molybdate catalyst.
hyde is one of the world’s most important chemicals Other routes have not yet been commercialized, e.g.
[1]. It may be synthesized by either dehydrogenation the sodium catalyzed dehydrogenation of methanol
(Eq. (1)) or partial oxidation (Eq. (2)) of methanol: to anhydrous formaldehyde [7] and the partial oxida-
CH3 OH  CH2 O + H2 , H = +84 kJ/mol (1) tion of methane over silica catalysts [8]. With about
55% of the industrial production capacity in western
CH3 OH + 21 O2 → CH2 O + H2 O, Europe being based on the silver-catalyzed route in
H = −159 kJ/mol (2) 2000 [1], it plays an important role. Its advantages
are the relatively low investment cost, the high yield
Two important routes are prominent in the industrial and the stable production.
formaldehyde production [2–6]. The first synthesis In commercial production, the strongly exothermic
∗ Corresponding author. Tel.: +49-9131-85-27-426; reaction is carried out under adiabatic conditions [2].
fax: +49-9131-85-27-421. A selection of parallel and consecutive reactions is
E-mail address: min.qian@rzmail.uni-erlangen.de (M. Qian). given in Table 1. They may inevitably lead to many

0926-860X/02/$ – see front matter © 2002 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 6 - 8 6 0 X ( 0 2 ) 0 0 3 4 0 - X
212 M. Qian et al. / Applied Catalysis A: General 238 (2003) 211–222

Table 1 Despite a century of industrial production, still


Some possible secondary reactions in the system with the compo- no comprehensive insight into the silver-catalyzed
nents formic acid, hydrogen, carbon dioxide and carbon monoxide,
formaldehyde synthesis is available. It has been
coke, etc. [2,4]
pointed out for the methanol oxidation that the high
Number Reaction HR,973 K (kJ/mol) yields and selectivities may seem to render any at-
(3) CH2 O → CO + H2 +12 tempt to study the reaction with the aim to improve it
(4) CH3 OH + 23 O2 → CO2 + 2H2 O −676 superfluous, but that, considering the potential market
(5) CH2 O + O2 → CO2 + H2 O −519 demand of formaldehyde, even small improvements
(6) CH2 O + 21 O2 → CH2 O2 −314
will be economically interesting [16]. This is certainly
(7) CH3 OH → C + H2 O + H2 −31
(8) CO + H2 C + H2 O −136 one of the reasons that quite some research work is
(9) CO + H2 O CO2 + H2 −35 devoted to this system.
One approach is the surface science approach where
the interaction of reactants and well-defined single
byproducts, e.g. carbon dioxide, carbon monoxide, crystal surfaces is studied under high vacuum condi-
formic acid, hydrogen, coke, etc. The combination tions. In order to move to industrially more relevant
of the reactions yields a rather complex reaction conditions, some studies have been carried out with
network. polycrystalline electrolytic silver at atmospheric pres-
Various versions of the conventional silver-catalyzed sure and temperatures up to 930 K. From the body of
process are adopted commercially, described in many literature, a mechanistic model begins to emerge that
different patents [9–14]. Two main variations are used may explain a number of experimental observations.
industrially. The first process (hereafter referred to A brief outline will be given in the next paragraphs.
as methanol ballast process) in which only air and It is generally accepted that understanding well the
pure methanol are fed is used in numerous companies interaction of oxygen with silver is a prerequisite to ex-
(e.g. ICI and Degussa). Features are an incomplete plain the mechanism of total reaction routes [17–30].
conversion of methanol and its distillative recov- Three different oxygen species O␣ , O␤ and O␥ are as-
ery. BASF uses another important process (hereafter sumed to form during the reaction [31,32]. Molecular
referred to as water ballast process) feeding extra oxygen dissociates on the silver and forms the weakly
water with the reactant mixture, thereby achieving a bound atomic surface oxygen species O␣ which bene-
complete conversion of methanol [2,9,15]. However, fits especially the formation of HCOOH and the com-
water addition is limited by the requirement of the plete oxidation of products to H2 O and CO2 . Deng
final products strength and of the additional water et al. even supposed that molecularly adsorbed oxygen
needed for tail-gas scrubbing. Normally, the commer- acts as a precursor and transforms to atomic oxygen
cial production is carried out with a H2 O/CH3 OH in this process [33]. With similar atomic dimensions
molar ratio of 40/60 or 0.67 [2]. Due to its large (φ Ag : 60 Å, φ O : 62 Å), oxygen may dissolve in the
heat capacity, water vapor may remove a great deal silver lattice. This O␤ cannot participate directly in
of reaction heat, thereby both preventing detrimen- the reaction. The strongly bound oxygen species O␥
tal overheating as well as sintering of the catalyst. is formed from O␤ , when the latter segregates from
Moreover, because addition of water vapor helps to the bulk to the surface. O␥ exists predominantly in
burn away the coke on the catalyst surface, it is re- the uppermost layer of silver catalyst and tends to
ported for the industrial production that the life time only catalyze the dehydrogenation of methanol. The
of Ag catalyst in the water ballast process is signifi- oxygen–silver chemistry also seems to influence the
cantly longer than that in the methanol ballast process adsorption of methanol which has been found to hardly
[5]. occur on an oxygen-free silver surface [34]. Under
In the present study, both methanol and water bal- reaction conditions, the silver is constantly exposed to
last processes were investigated comparatively in the oxygen, part of which will penetrate into the crystal
same laboratory-scale setup. The influence of water lattice of silver and form the oxygen species of O␣ , O␤
addition on catalyst life time, formaldehyde selectiv- and O␥ species that are depleted by the corresponding
ity and generation of byproducts was studied. reactions, which shows a dynamic equilibrium [35].
M. Qian et al. / Applied Catalysis A: General 238 (2003) 211–222 213

The population is a strong function of parameters Table 2


and pre-treatment. The three oxygen species have, Standard settings and operation range in the laboratory-scale setup
e.g. stability regions at different temperatures. Another Standard Operation
factor is the dynamic restructuring of the silver: SEM setting range
images reveal that the morphology of silver catalyst Temperature (K) 823 473–993
changes strongly during the reaction [32,36–39]. Mil- Residence time (s; at 823 K) 0.085 0.06–0.45
lar et al. observed that “pinholes” appear firstly pre- GHSV (h−1 ; at 298 K) 1.5 × 104
dominantly in the vicinity of initial surface defects WHSV (h−1 ; at 298 K) 440
CH3 OH fraction in the feed (vol.%) 17.5
and spread gradually from the region of defects to the O2 fraction in the feed (vol.%) 7
entire silver surface [40]. This reaction-induced mor- CH3 OH/O2 molar ratio 2.5
phological restructuring reinforces the formation of H2 O/CH3 OH molar ratioa 0.67 0–2.0
grain boundary defects over the catalyst, which en- a In the methanol ballast process, no water was fed and the
ables more oxygen to penetrate into the silver lattice other conditions were the same as those in the water ballast
and, in turn, to intensify the reaction [30,31,41–43]. process. The space velocity was kept constant by varying the N2
Recent studies have focused on using silver sup- flow accordingly.
ported catalysts [44], on elucidating the mechanism
with sophisticated vacuum experiments, such as tem- mesh) in a tubular fixed-bed reactor. With the help of
poral analysis of products (TAPs) [45] or on optimiz- two HPLC pumps, methanol and water were pumped
ing the thermal management [46]. The present exper- separately into the two different vaporizers (kept at 363
imental work addresses three industrially relevant is- and 393 K, respectively) which were set up in series.
sues that have been little studied in the past: Assisted by N2 as a carrier gas, evaporation yielded
partial pressures with a standard deviation of ±1.5%.
• the investigation of the water ballast process with
Oxygen was added behind the second vaporizer and
H2 O/CH3 OH of 0–2.0;
the mixture flowed through the catalyst bed from the
• (detrimental) formic acid formation as a function of
top of the reactor. Table 2 gives the range and stan-
control parameters and time;
dard settings of all relevant operating parameters. In
• long-time behavior (10 months) of conversion and
this paper, yields and selectivities are expressed with
selectivities.
respect to the excess component CH3 OH and not with
Two byproducts that were closely monitored were respect to oxygen.
carbon dioxide and formic acid. Carbon dioxide, The reactor was made of a 50 cm long Al2 O3 tube
one of the major byproducts, was monitored with and had an inner diameter of 10 mm. The blank ac-
time-on-stream to gauge the catalyst activity during a tivity proved negligible. The reactor was filled with
long-time operation. Formic acid is a strongly corro- quartz chips (Ø 1 mm) before the catalyst bed to im-
sive byproduct and an undesired contaminant of the prove the mixing and pre-heating of the feed. The 3 cm
final product. It is, therefore, also strictly controlled long Ag catalyst bed was strongly diluted with quartz
in commercial operation (normally <0.01 wt.% in chips (Ag/quartz: 0.1 g/3 g) in order to weaken the lo-
the product [2]). Before the onset of catalyst deacti- cal heating effects.
vation, the formation of formic acid is restricted to Heating was provided by an electric heat coil con-
only a limited temperature range. After deactivation, trolled by a Eurotherm temperature controller, which
normally 4–9 months, formic acid is formed at all monitored the temperature with the help of a thermo-
temperatures, which forces a shutdown and catalyst couple, located directly in the middle of catalyst bed
replacement in industrial production. through a capillary Al2 O3 tube. At the given param-
eter settings, no isothermal operation was possible
(see Section 4). The temperatures given in this paper
2. Experimental are those located on the centerline, at an axial posi-
tion of approximately 1 cm into the catalyst bed (T1 ).
Catalytic partial oxidation of methanol was carried In order to prevent the product formaldehyde from
out over an electrolytic silver powder (Bayer AG, 20 decomposing at the high temperature, the reaction
214 M. Qian et al. / Applied Catalysis A: General 238 (2003) 211–222

was carried out with a very short residence time of 3. Results


<0.086 s. Moreover, the product mixture was rapidly
quenched to about 413 K subsequently with a cooling 3.1. Influence of operating parameters
jacket behind the catalyst bed. The quench down-
stream caused a strong axial temperature gradient: the 3.1.1. Influence of reaction temperature
temperature on the centerline 2 cm downstream of T1 , Fig. 1 depicts conversions and yields as a function of
just before the quench (T2 ), was about 42 K higher. temperature. A light-off temperature was observed at
This strong gradient is the reason that, compared with about 570 K. CO2 displayed a maximal yield at the rel-
the results of Panzer [31], different runs may exhibit atively low temperature of 575 K and then dropped off
relative shifts on the temperature axis even if their with temperature. The yield of formaldehyde increased
qualitative behavior is identical. gradually with temperature and reached a maximum
Analysis of the products was made by online gas at about 923 K, which corresponds well with the com-
chromatography (HP 5890 series II) with two TCD de- mercial operation temperature. The abrupt decrease
tectors. Qualitative and quantitative separation of both of the formaldehyde yield above 923 K was accom-
high and low boiling point compounds was possible panied by a yield increase of CO and H2 , suggesting
in this setup with a mol-sieve (5A, HP) and a polar a gas phase decomposition of formaldehyde to CO
capillary column (Poraplot-Q, HP). The separation of and H2 at the high temperature. Formic acid appeared
formaldehyde and water is relatively difficult, but was only in a limited temperature region (approximately
attained in this study within a reasonable accuracy 570–850 K) and could not be observed in the high
(see error bars in Fig. 3). The carbon balance is about temperature region before the deactivation of catalyst
100 ± 5% with an acceptable margin of error. For the (see Section 3.3).
temperature variation, a heating rate of 0.16 K/min was
selected. Combined with a sampling period of 45 min 3.1.2. Influence of residence time
and a sampling time of seconds, it may be safely as- Methanol conversion and the selectivities to
sumed that the values obtained at this extremely slow formaldehyde and hydrogen were determined at dif-
heating rate are representative of the steady state val- ferent residence times (0.06–0.45 s; Fig. 2). The
ues with a temperature resolution of 7.5 K. higher the residence time was, the more methanol

Fig. 1. Variation of the operating temperature in the water ballast process with 225 days time-on-stream.
M. Qian et al. / Applied Catalysis A: General 238 (2003) 211–222 215

Fig. 2. Variation of the residence time (823 K).

was converted (open triangles). However, the longer 3.1.3. Influence of molar ratio of H2 O/CH3 OH in
residence time was not beneficial for the formalde- the feed
hyde formation: its selectivity decreased apparently The influence of water vapor in the reaction gas on
under the longer residence time, which may be partly the formaldehyde selectivity was estimated. Water va-
due to the fast decomposition of formaldehyde in the por content was varied in the region of H2 O/CH3 OH
gas phase to H2 and CO at high operation tempera- molar ratio of 0–2.0. The space velocity was kept
tures. The H2 selectivity did increase with residence constant by varying the N2 flow accordingly. This
time, albeit not to the extent that the formaldehyde led to a constant CH3 OH/O2 molar ratio. Each re-
selectivity decreased. sult was an average over a 15 h lasting stationary

Fig. 3. Selectivity to HCHO and CO2 vs. the molar ratio of H2 O/CH3 OH (823 K; CH3 OH, 8.5%; O2 , 3.5%).
216 M. Qian et al. / Applied Catalysis A: General 238 (2003) 211–222

test. Fig. 3 indicates that the conversion of methanol


increased with the H2 O/CH3 OH molar ratio, how-
ever, the selectivity to formaldehyde passed through
a maximum around a H2 O/CH3 OH molar ratio of
about 0.75, which corresponds basically well with
the above-mentioned molar ratio of 0.67 in industrial
formaldehyde manufacture (indicated by the vertical
dashed line). Because of the experimental error in the
formaldehyde detection, the experiment was repro-
duced at different feed concentrations, supporting the
conclusions reported above. It is also shown in Fig. 3
that the selectivity to CO2 decreased with the molar
ratio of H2 O/CH3 OH. The more water vapor was fed
in the reaction gas, the less CO2 was detected.

3.2. Comparison of water and methanol ballast


Fig. 4. Comparison between the water and methanol ballast pro-
processes
cess.

In order to estimate the effect of water vapor on


the reaction, during the long-time water ballast study ments carried out subsequently and could be neg-
some experiments were carried out where methanol lected.
ballast conditions were ensured. The two modes of In comparison with the methanol ballast process,
operation were, therefore, studied at identical space the water ballast process displayed a higher conversion
velocity in the same setup. The gradual decrease of of methanol and selectivity to formaldehyde. Fig. 4
catalyst activity (see Section 3.3) was slow com- shows the methanol conversion rose by 10% and the
pared to the time elapsed between the two experi- formaldehyde selectivity increased by 4% at 823 K, if

Fig. 5. Comparison of HCOOH selectivity between the water and methanol ballast process (lines indicate a moving average over a 10 K
temperature interval).
M. Qian et al. / Applied Catalysis A: General 238 (2003) 211–222 217

Fig. 6. The life time of silver catalyst in the water ballast process (823 K).

adding water vapor to the reaction gas. The selectivity 3.3. Aging behavior
to CO2 , the product of total oxidation, was obviously
suppressed in the water ballast process as compared In order to gauge the long-time behavior, conver-
to the methanol ballast process. sions and selectivities were measured for over 300
In addition, the selectivity to formic acid was also days time-on-stream. After a short initialization, both
clearly smaller in the water ballast process. Fig. 5 methanol conversion and formaldehyde selectivity
shows the result of two experiments carried out one were basically constant (steady state performance).
immediately after the other. The selectivity to formic The latter may have displayed a slight decrease
acid in the water ballast process was almost 30% lower (Fig. 6). A much more pronounced change was ob-
at 620 K than that in the methanol ballast process. served for the selectivities to CO2 and H2 (Fig. 7).

Fig. 7. Selectivity to CO2 and H2 vs. time-on-stream in the water ballast process (823 K).
218 M. Qian et al. / Applied Catalysis A: General 238 (2003) 211–222

Fig. 8. Selectivity to HCOOH vs. time-on-stream in the water ballast process.

S(CO2 ) increased gradually over the entire experi- two trends made the region of HCOOH formation
ment. S(H2 ) decreased from 14 to 8% during the first broader with time.
100 days. It then dropped to about 5–6%, where it To study the nature of this deactivation, the opera-
remained basically constant for the remainder of the tion was interrupted once during the time while the up-
experiments. per limit still changed (180 days time-on-stream) and
Fig. 8 shows the selectivity to formic acid as a once when it was constant (280 days time-on-stream).
function of temperature for various times-on-stream. The interruption consisted of 4 weeks during which
No formic acid was detected during the first 100 the catalyst bed was kept under flowing N2 at 523 K.
days. After this period, formic acid appeared, but only In both cases, after the experimental conditions had
in a limited temperature region of 600–690 K with
S(HCOOH) < 0.05. However, this region grew with
time, such that after 8 months (240 days), formic acid
could be observed over a wide range of temperatures
(580–870 K). In order to visualize the slow extension
of formic acid formation into the high temperature
region, we define an upper (Tupper ) and a lower
(Tlower ) temperature limit such that these two bound
the region where formic acid could be detected in the
outlet with S(HCOOH) > 0.01. The two tempera-
tures are plotted in Fig. 9 versus time-on-stream. The
upper temperature increased linearly to 870 K during
the first 8 months with a correlation coefficient of
R 2 > 0.97. After this time, the shift into the higher
temperature region stopped and the upper tempera-
ture stayed constant with time. The lower temperature
limit decreased with time-on-stream, but in a less Fig. 9. Upper and lower detection temperature of HCOOH vs.
pronounced manner and without a discontinuity. The time-on-stream.
M. Qian et al. / Applied Catalysis A: General 238 (2003) 211–222 219

been restored, no marked discontinuities were ob- oxygen (and hence the normal feed composition) was
served (Fig. 9). switched back on. The exit gas composition was av-
eraged over 5 h following the oxygen-free treatment.
3.4. Oxygen-free treatment Fig. 10a shows that, at three different temperatures
(623, 773 and 823 K, respectively), the conversion of
For more insight into the interaction between oxy- methanol and the yield of CH2 O decreased strongly af-
gen and silver during the partial oxidation of methanol, ter oxygen-free treatment as compared to steady state.
the catalyst was temporarily exposed to an oxygen-free In addition, the yields of carbon dioxide and formic
atmosphere. Starting from steady state, the oxygen acid had increased while the hydrogen yield was lower.
feed was interrupted for 5 h. Methanol, water vapor The decrease of the ratio of S(CH2 O)/S(CO2 ) after
and nitrogen kept flowing over the catalyst. After this, oxygen-free treatment is depicted in Fig. 10b.

Fig. 10. Influence of oxygen-free treatment on: (a) the conversion and the yield; (b) the ratio of S(CH2 O)/S(CO2 ) as compared to steady
state in the water ballast process.
220 M. Qian et al. / Applied Catalysis A: General 238 (2003) 211–222

4. Discussion of Millar et al. demonstrated that subsurface OH(a)


species were formed very rapidly in a well-defected
In order to classify the results, it should firstly be Ag sample, whereas extremely slowly over the smooth
noted that laboratory-scale operation was neither adi- surface of a fresh Ag sample [50]. It is assumed that
abatic nor isothermal. Like in [25], a pronounced tem- the initial activation of silver catalysts is due to the
perature gradient was measured in the catalyst bed. diffusion of OH(a) species and the consequent forma-
When discussing the results, it should be kept in mind tion of defected sites.
that the temperatures given are those measured at T1 A notable trend is that S(CO2 ) increased with time
and used for the temperature control. Rather than the whereas S(H2 ) decreased (Fig. 7). This may point to
absolute values of T1 , the trend should be considered. an increase of the number of weakly bound atomic
The excellent isothermicity of catalytic wall reactors O␣ on the silver surface with time. The degree of
as reported by Redlingshöefer et al. [47] has also been total oxidation is commonly attributed to the surface
used for the methanol oxidation to formaldehyde [31]. adsorbed atomic O␣ . While this is particularly true
However, more development is required for the sys- in the low temperature region, after extended times it
tem prior to publication. may also hold in the high temperature region, where
A second caveat is that even if we will discuss the re- O␥ plays the prominent role in the dehydrogenation
ported findings in the framework of the state-of-the-art of methanol.
of silver–oxygen chemistry, it should be obvious that HCHO is normally unstable at high temperatures,
due to the absence of spectroscopic evidence, this pa- being easily decomposed to H2 and CO or, in the pres-
per cannot contribute much to the body of evidence. ence of oxygen, even totally oxidized to CO2 and H2 O.
However, the phenomenological approach taken here It is also assumed that HCHO could be further oxi-
may indeed give suggestions for new experiments. As dized to HCOOH in the presence of the weakly bound
a matter of fact, transferring promising results with surface atomic O␣ [31]:
fiber-optic in situ spectroscopy [48] to this reaction
HCHO + O␣ → HCOOH
system is currently underway. Raman spectroscopy
seems to be an appropriate method [49]. As mentioned above, the influence of O␣ became
The comparison of water and methanol ballast pro- more and more notable with time, HCOOH was con-
cesses indicates that an appropriate water vapor ad- sequently formed in a wider temperature region and
dition plays a great role in the reaction. Bao et al. even at high temperatures after several months of op-
[35] found the oxygen-induced restructuring process eration (Fig. 8). Because the O␣ species plays an im-
can be greatly enhanced by adding a small amount portant role mainly in the lower temperature region,
of water vapor to O2 , which makes the formation of it is reasonable that the amount of HCOOH increased
oxygen species more favorable. An evident increase particularly at low temperatures around 600 K with
of CH3 OH conversion shown in Fig. 4 supports this time-on-stream. The reason that no formic acid was
hypothesis. As well as reinforcing the conversion of detected above 870 K is that this is the decomposition
methanol, water vapor can suppress effectively the to- temperature and does not reflect a discontinuity in the
tal oxidation and reduce consequently the generation deactivation process.
of CO2 shown in Figs. 3 and 4. Various tentative ex- The experimental results after oxygen-free treat-
planations can be given: water vapor may block spe- ments confirmed this hypothesis well on another as-
cific surface sites at which total oxidation takes place; pect. The dynamic equilibrium of the O-development
it may help burn away the coke on the catalyst surface is disturbed during treatment with N2 gas only. Since
to expose more active sites. oxygen species could not evolve well during nitro-
The study shows an initialization phase to ignite gen treatment, the dominant oxygen species in the
the silver catalyst, which was also observed by Nagy subsequent oxidative reaction would initially be the
et al. [32]. This is necessary not only in the methanol weakly bound atomic O␣ , which shows high oxidative
ballast process but also in the water ballast process, power and helps promoting the formation of carbon
which was also observed in the industrial plants from dioxide and formic acid. With the help of in situ Ra-
some hours to several days. The in situ Raman studies man spectroscopy, Pettinger et al. discerned a gradual
M. Qian et al. / Applied Catalysis A: General 238 (2003) 211–222 221

re-development of O␥ formation over the Ag catalyst support by the Bayerische Forschungsstiftung in the
when it was exposed again to O2 after a 30 min treat- framework of FORKAT and by the DFG (Li 669/2-1)
ment with pure N2 [51]. As a result, treatment un- is gratefully acknowledged. The authors thank their
der oxygen-free conditions followed by re-admission industry partners H.-G. Pirkl (Bayer), A. Nagy (for-
of oxygen should temporarily lead to an increase of merly Bayer), and G. Stochniol and J. Sauer (Degussa
O␣ -catalyzed formation of carbon dioxide and formic AG) for their support.
acid as well as a decrease of O␥ -catalyzed hydrogen
formation as compared to the steady state. This is in-
deed found (Fig. 10a). Consequently, the oxygen-free References
treatment should lead to an Ag catalyst on which the
[1] E. Burridge, Eur. Chem. News 75 (1983) 16.
steady state distribution of oxygen species does not [2] G. Reuss, W. Disteldorf, A.O. Gamer, A. Hilt, Formaldehyde,
fully develop immediately. Hence, a temporary de- in: Ullmann’s Encyclopedia of Industrial Chemistry, 6th
crease of methanol conversion and CH2 O yield would Edition, Vol. A11, VCH, Weinheim, 2001, p. 619.
be expected after oxygen-free treatment. This is also [3] B.A.J. Fischer, Formaldehyde, in: Encyclopaedia of Chemical
supported well by Fig. 10a. A decrease of the ratio of Technology, 4th Edition, Vol. 11, Wiley, New York, 1992,
p. 929.
S(CH2 O)/S(CO2 ) appeared after the treatment under [4] M.V. Twigg, Catalyst Handbook, 2nd Edition, Manson
oxygen-free condition, especially with almost 50% at Publishing, London, 1996, pp. 490.
low temperature where O␣ is normally developed well [5] K. Weissermel, H.-J. Arpe, Industrielle Organische Chemie,
and plays a dominant role (shown in Fig. 10b). 5th Edition, Wiley/VCH, New York/Weinheim, 1998, pp. 40–
41.
[6] A.R. Chauvel, P.R. Courty, R. Maux, C. Petitpas, Hydrocarbon
5. Conclusions Process. 9 (1973) 180.
[7] S. Ruf, A. May, G. Emig, Appl. Catal. A 213 (2001) 203.
[8] F. Arena, F. Frusteri, A. Parmaliana, Appl. Catal. A 197
The silver-catalyzed synthesis of formaldehyde (2000) 239.
from methanol was studied at near industrial condi- [9] G. Halbritter, W. Mühlthaler, H. Soerber, H. Diem, C. Dudeck,
tion. The reaction proves sensitive to some operating G. Lehmann, Verfahren zur herstellung von Formaldehyd, DE
parameters, e.g. reaction temperature, residence time Patent 2,442,231 (1974).
and molar ratio of H2 O/CH3 OH. [10] K. Yoshikawa, T. Matsuzawa, Process for producing aqueous
formaldehyde solution, US Patent 4,594,457 (1985).
Methanol can be converted more completely and [11] A. Aicher, G. Lehmann, N. Petri, W. Pitteroff, V. Reuss,
the selectivity to formaldehyde increased if water is V. Schreiber, R. Sebastian, Preparation of formaldehyde, US
added to the reaction gas. In addition, an appropriate Patent 4,584,412 (1986).
amount of water vapor decreases the total oxidation [12] W.P. McMillan, C.C. Hobbs Jr., H.R. Gerberich, M.L. Junker,
in the reaction system and suppresses the generation Process for converting methanol to formaldehyde, US Patent
4,450,301 (1984).
of formic acid. Besides removing a great deal of heat [13] B. Knuth, R. Diercks, Process for the preparation of
to avoid the sintering of the catalyst, water vapor may formaldehyde, US Patent 5,990,358 (1999).
also block total oxidation sites, or burn off coke, to [14] I.E. Wachs, C.-B. Wang, Formaldehyde production, US Patent
improve the silver activity significantly. 6,147,263 (1999).
The amount of carbon dioxide, the main byprod- [15] H. Sperber, Chem. Ing. Tech. 41 (1969) 962.
[16] G. Centi, F. Cavani, F. Trifirò, Selective oxidation by
uct, grew with time-on-stream, revealing a gradual en- heterogeneous catalysis, Kluwer Academic Publishers/Plenum
hancement of oxidation in the reaction system. An- Press, Dordrecht/New York, 2001, pp. 441.
other major byproduct, formic acid, was first formed [17] L. Lefferts, J.G. van Ommen, J.R.H. Ross, Appl. Catal. A 23
in a narrow temperature region. However, the region (1986) 385.
grew wider for more than 10 months. [18] C.T. Au, S. Singh-Boparai, M.W. Roberts, R.W. Joyner, J.
Chem. Soc., Faraday Trans. 79 (1983) 1779.
[19] G. Rovida, F. Pratesi, M. Maglietta, E. Ferroni, Surf. Sci. 43
(1974) 230.
Acknowledgements
[20] X. Bao, J.V. Barth, G. Lehmpfuhl, R. Schuster, Y. Uchida,
R. Schlögl, G. Ertl, Surf. Sci. 283 (1993) 14.
Ernst Panzer is thanked for invaluable groundwork [21] A. Nagy, G. Mestl, D. Herein, G. Weinberg, E. Kitzelmann,
and advice regarding this reaction system. Financial R. Schlögl, J. Catal. 182 (1999) 417.
222 M. Qian et al. / Applied Catalysis A: General 238 (2003) 211–222

[22] J.-H. Wang, W.-L. Dai, J.-F. Deng, X.-M. Wei, Y.-M. Cao, [36] L. Lefferts, J.G. van Ommen, J.R.H. Ross, Appl. Catal. A 34
R.-S. Zhai, Appl. Surf. Sci. 126 (1998) 148. (1987) 329.
[23] J.-H. Wang, X.-H. Xu, J.-F. Deng, Y. Liao, B.-F. Hong, Appl. [37] P.J.R. Uwins, G.J. Millar, M.L. Nelson, Microsc. Res. Tech.
Surf. Sci. 120 (1997) 99. 36 (1997) 382.
[24] B. Pettinger, X. Bao, I. Wilcock, M. Muhler, R. Schlögl, G. [38] X. Bao, G. Lehmpfuhl, G. Weinberg, R. Schlögl, G. Ertl, J.
Ertl, Angew Chem. 106 (1994) 113. Chem. Soc., Faraday. Trans. 88 (1992) 865.
[25] A. Nagy, G. Mestl, Appl. Catal. 188 (1999) 337. [39] G.J. Millar, M.L. Nelson, P.J.R. Uwins, J. Chem. Soc.,
[26] C.-B. Wang, G. Deo, I.E. Wachs, J. Phys. Chem. B 103 Faraday Trans. 94 (1998) 2015.
(1999) 5645. [40] G.J. Millar, M.L. Nelson, P.J.R. Uwins, J. Catal. 169 (1997)
[27] G.J. Millar, J.B. Metson, G.A. Bowmaker, R.P. Cooney, Appl. 143.
Catal. 147 (1994) 385. [41] C. Rehren, G. Isaac, R. Schlögl, G. Ertl, Catal. Lett. 11 (1991)
[28] X.-D. Wang, W.T. Tysoe, R.G. Greenler, K. Truszkowska, 253.
Surf. Sci. 258 (1991) 335. [42] I.E. Wachs, R.J. Madix, Surf. Sci. 76 (1978) 531.
[29] M. Voß, D. Borgmann, G. Wedler, Surf. Sci. 465 (2000) [43] G.J. Millar, M.L. Nelson, P.J.R. Uwins, Catal. Lett. 43 (1997)
211. 97.
[30] L. Lefferts, J.G. van Ommen, J.R.H. Ross, Appl. Catal. A 31 [44] W.-L. Dai, J.-L. Li, Y. Cao, Q. Liu, J.-F. Deng, Catal. Lett.
(1987) 291. 64 (2000) 37.
[31] E. Panzer, Reaktionstechnische Untersuchungen zur kataly- [45] A.C. van Veen, H.W. Zanthoff, O. Hinrichsen, M. Muhler, J.
tischen Oxidation von Methanol zu Formaldehyd an Silber, Vac. Sci. Technol. A 19 (2001) 651.
Doctoral thesis, Universität Erlangen-Nuremberg, Germany, [46] G. Groppi, E. Tronconi, Catal. Today 69 (2001) 63.
2000. [47] H. Redlingshöefer, O. Kröcher, W. Böck, K. Huthmacher, G.
[32] A. Nagy, G. Mestl, T. Rühle, G. Weinberg, R. Schlögl, J. Emig, Ind. Eng. Chem. Res. 41 (2002) 1445.
Catal. 179 (1998) 548. [48] R. Philipps, S. Walter, M. Liauw, Chem. Eng. J., in press.
[33] J.-F. Deng, X.-H. Xu, J.-H. Wang, Y.-Y. Liao, B.-F. Hong, [49] G. Mestl, J. Mol. Catal. A Chem. 158 (2000) 45.
Catal. Lett. 32 (1995) 159. [50] G.J. Millar, J.B. Metson, G.A. Bowmaker, R.P. Cooney, J.
[34] J.-F. Deng, X. Bao, J. Catal. 99 (1986) 391. Chem. Soc., Faraday Trans. 91 (22) (1995) 4149.
[35] X. Bao, M. Muhler, B. Pettinger, R. Schlögl, G. Ertl, Catal. [51] B. Pettinger, X. Bao, I.C. Wilcock, M. Muhler, G. Ertl, Phys.
Lett. 22 (1993) 215. Rev. Lett. 72 (10) (1994) 1561.

S-ar putea să vă placă și