Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Methanol oxidation 167

Ayse Hilal Ulukardesler1 Research Article


Süheyda Atalay1
Ferhan S. Atalay1 Determination of Optimum Conditions and
1
Department of Chemical the Kinetics of Methanol Oxidation
Engineering, Ege University,
Bornova-Izmir, Turkey.
In this study, the catalytic oxidation of methanol to formaldehyde was investi-
gated in a laboratory-scale fixed-bed catalytic reactor, under a large number of
different conditions. Iron-molybdate catalysts supported by silica or alumina
with a molybdenium/iron (Mo/Fe) ratio of 1.5, 3 and 5 were studied for the gas
phase reaction. In order to obtain the optimum conditions, six different tempera-
tures in the range of 250–375 °C and three different space times of 50.63, 33.75
and 20.25 g/(mol/h) were investigated. After determining the optimum condi-
tions for this reaction, experiments aimed at understanding the reaction kinetics,
were carried out. These experiments were performed on the catalyst favoring the
formation of formaldehyde, which has a (Mo/Fe) ratio of 5 on a silica support.
Seven reaction models derived by the mechanisms cited in the literature were
tested to elucidate the kinetics of the reaction and the surface reaction controlling
model was found to be the most suitable reaction mechanism.

Keywords: Fixed-bed catalytic reactor, Formaldehyde, Iron-molybdate catalyst,


Kinetics of methanol oxidation, Methanol oxidation
Received: July 7, 2009; revised: September 28, 2009; accepted: October 19, 2009
DOI: 10.1002/ceat.200900349

1 Introduction cess, respectively [7]. Most of the newly built formaldehyde


plants (more than 70 %) are based on the metal oxide catalyst
Methanol, one of the world’s most important chemical inter- (Fe-Mo) due to almost complete conversion of methanol
mediates, is the starting material for the synthesis of various (exceeding 99 %) as well as very high formaldehyde selectivity
products such as formaldehyde [1]. Formaldehyde is a reactive (95 %) and, most importantly, due to the imposition of strin-
molecule and the first in the series of aliphatic aldehydes [2] gent environmental regulations [3]. The most widely used cat-
and is a base chemical of major industrial importance. In spite alyst consists of mixtures of iron-molybdate with molybdenum
of fluctuations in the world economy, the growth of formalde- trioxide often modified by different additives. Its catalytic
hyde production has been remarkably steady and is expected performance is highly satisfactory and practically unbeatable
to continue [3]. Formaldehyde consumption has grown by (98–100 % methanol conversion and 92–94 % selectivity to
2–3 % per year on an average over the past two decades, due formaldehyde) [8].
primarily to increasing demand in the construction sector for Over the last decade several researchers have studied the
engineered wood products manufactured using formaldehyde- kinetics of the partial oxidation of methanol to formaldehyde.
based resins [4]. The annual worldwide production capacity of Jiru et al. [9] suggested a redox mechanism similar to that ob-
formaldehyde now exceeds 25–27 million tons (calculated as served for the Mars-van Krevelen type reaction mechanism for
37 % solution) [5]. the description of the oxidation of aromatic hydrocarbons over
The main production process for formaldehyde is air oxida- V2O5. They found that the formaldehyde concentration affects
tion of methanol with high conversion; by an exothermic reac- the reaction rate, while the water concentration does not [4].
tion at atmospheric pressure and at a temperature of 300– Santacesaria et al. studied the kinetics of the oxidation of
400 °C [6]. All industrial production of formaldehyde uses methanol to formaldehyde over an iron-molybdate catalyst
methanol and air as raw materials. Two technologies are avail- [10]. The ratio Mo/Fe of the employed catalyst was 2.5 and it
able, which are referred as the silver process and the oxide pro- contained small amounts of cobalt molybdate (1.8 % by weight
of CoO). The results were described with a kinetic model de-
rived on the basis of a redox mechanism partially hindered by
– the absorption of water [10]. McCormick et al. studied the oxi-
Correspondence: Dr. A. H. Ulukardesler (ahulukardesler@gmail.com), dation of methane, methanol, formaldehyde and carbon mon-
Department of Chemical Engineering, Ege University, 35100, Bornova- oxide over precipitated silica catalysts examined over a range
Izmir, Turkey. of reactant and oxygen partial pressures [8]. The conversion

Chem. Eng. Technol. 2010, 33, No. 1, 167–176 © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.cet-journal.com
168 A. H. Ulukardesler et al.

and selectivity relationships were used to assess the reaction 2.2 Experimental Setup
network and differential reactor experiments were employed
to determine the global reaction kinetics. The authors applied All parts of the setup were made of 316 stainless steel. A pre-
the Power Law model to their results. They found that the heater with an inside diameter of 6 cm and a length of 30 cm
reaction orders were 0.4 and 0.43 in oxygen and methanol, was heated using a Ni-Cr electrical wire that surrounded the
respectively [8]. pipe. Methanol was fed to the preheater using a variable speed
Yao et al. concluded that methanol was highly oxidized in peristaltic pump. The reactor had a length of 308 mm and a
the temperature range of 400–600 °C over SiO2 [11]. They diameter of 27 mm. A catalyst bed, having a special arrange-
constructed the methanol oxidation model on two assump- ment of holes, was located at the center of the reactor. An
tions, i.e., the first-order reaction of methanol and the first- iron-constantan thermocouple connected to a PID tempera-
order reaction of formaldehyde. These assumptions were ture controller was inserted into the reactor to control and
consistent with the experimental results when the oxygen con- measure the working temperature. A condensation system that
centration in the gas phase was higher than 1.5 % [11]. In the included a condenser and drum was used to collect and ana-
study carried out by Forzatti and Buzzi-Ferraris, methanol oxi- lyze the reaction mixture. The vapor products exiting the reac-
dation over a silica-supported Fe2O3-MoO3 catalyst at 370 °C tor were condensed by circulation of an alcohol-water refriger-
was determined from the 13 deactivation runs that were per- ant system through the shell-side of the exchanger. The
formed under integral reactor conditions [12]. The authors condensed products were collected in the drum [14].
derived two reliable reaction-deactivation models on the basis The analyses of liquid products collected in the drum were
of previous chemical investigations on fresh and deactivated performed by using a HP 5860 series 2 type Gas Chromato-
catalysts and/or information on methanol oxidation kinetics graph. A GA 94 A Gas Analyzer (Geotechnical Instruments
over Fe2O3-MoO3 based catalysts [12]. Limited, UK), was used for the measurement of the mole per-
Although significant research exists on the subject, there are centage of carbon dioxide in the noncondensable exit gas. An
still areas to be clarified from the points of both catalysts used IR-470 Shimadzu Analyzer was used for the infra red (IR)
and operating conditions affecting the reaction. In order to analysis of all fresh and used catalysts. The surface areas and
make a further contribution to the existing findings in the lit- the static adsorption and desorption isotherm plots for the
erature, the ratio of Mo to Fe, support type, and the effects of fresh and used catalysts were measured using an Omnisorp
temperature and space time are examined in detail. For this 100CX Volumetric Adsorption Analyzer.
purpose, the catalytic oxidation of methanol is studied over six
different iron-molybdate catalysts supported by silica or alu-
mina with three different (Mo/Fe) ratios of 1.5, 3 and 5. Six 2.3 Experiments
different temperatures in the range of 250–375 °C, and three
different space times (W/FA0, i.e., catalyst amount/feed flow Iron-molybdate catalysts having Mo/Fe ratios of 1.5, 3 and 5,
rate) of 50.63, 33.75 and 20.25 g/(mol/h) are tested on all pre- on two different supports, i.e., silica and alumina, were pre-
pared catalysts1). In this manner, the optimum conditions for pared. Experiments were performed at three different metha-
the production of formaldehyde are simultaneously deter- nol flow rates. The amount of catalyst was kept constant at
mined for a wide range of different conditions. Furthermore, 5 g. By changing methanol flow rates, three different space
attempts are made to clarify the kinetics of the gas phase oxi- times, W/FA0, of 50.63, 33.75 and 20.25 g/(mol/h) methanol
dation of methanol to formaldehyde and to introduce a new were examined. All experiments were performed under atmo-
approach involving a metal oxide catalyst prepared in the labo- spheric pressure and the temperature was changed by 25 °C
ratory. increments in the range of 250–375 °C. The operating condi-
tions for the experiments are listed in Tab. 1.
No intermediates were observed during the analysis of the
2 Experimental Section experiments, and only formaldehyde was detected as the reac-
tion product. Therefore, the following reactions are assumed
2.1 Catalyst Preparation to be possible main reactions:

A patent that was registered by Cairati and di Fiore was fol- CH3OH + 1/2O2 → CHOH + H2O X1 (1)
lowed for the preparation of the catalysts [13]. This invention
concerned the catalysts used for the oxidation of methanol to
formaldehyde, and the recommended atomic ratios of molyb- CH3OH + 3/2O2 → CO2 + 2H2 X2 (2)
denum to iron were 1.53:1 and 2.2:1. The ratios applied in this
study were chosen by taking these ratios given into account where X1 is the conversion of CH3OH to CHOH (partial
and in order to see the real effect of the ratio, the Mo/Fe ratio oxidation), and X2 is the conversion of CH3OH to CO2 (total
ranges were increased to 1.5, 3 and 5. oxidation). The results are evaluated by using the following
definitions of the selectivity to formaldehyde, Eq. (3), and the
total conversion of methanol, Eq. (4):

Moles of CHOH produced


– X1 ˆ (3)
1) List of symbols at the end of the paper.
Moles of CH3 OH in feed

www.cet-journal.com © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eng. Technol. 2010, 33, No. 1, 167–176
Methanol oxidation 169

Table 1. Operating conditions for the determination of optimum conditions. 2.3.3 Kinetic Experiments

Catalyst Amount 5g The kinetic experiments were carried out after


Catalyst Support the determination of the safe region of reaction
conditions for negligible mass transfer effects.
Silica Alumina
The kinetic experiments were performed in the
Temperatures (°C) Mo/Fe W/FA0 (gcat h/mol) W/FA0 (gcat h/mol) temperature range of 250–375 °C. For each
temperature (250, 275, 300, 325, 350, 375 °C),
250, 275, 300, 325, 350, 375 1.5 50.63 50.63
five space time values of 0.51, 0.68, 1.69, 2.53
250, 275, 300, 325, 350, 375 1.5 33.75 33.75 and 3.38 g/(mol/h) values were tested.
250, 275, 300, 325, 350, 375 1.5 20.25 20.25
250, 275, 300, 325, 350, 375 3 50.63 50.63 3 Results and Discussion
250, 275, 300, 325, 350, 375 3 33.75 33.75
3.1 Catalyst Characterization
250, 275, 300, 325, 350, 375 3 20.25 20.25
250, 275, 300, 325, 350, 375 5 50.63 50.63 In order to retain the activity of the catalyst,
after each set of experiments, the catalyst was
250, 275, 300, 325, 350, 375 5 33.75 33.75
regenerated using dry air at 300 °C and then
250, 275, 300, 325, 350, 375 5 20.25 20.25 reused. After the completion of each set of

Moles of CO2 produced Table 2. Operating conditions for external diffusion experiments.
X2 ˆ (4)
Moles of CH3 OH in feed
Catalyst Support Silica

Therefore, the total conversion is given by Eq. (5): Catalyst Amount 1g

Moles of CH3 OH reacted Reactor Temperature 375 °C


XT = = X1 + X2 (5)
Moles of CH3 OH in feed Methanol Flow Rate 0.0034 mol/min
X
In addition, the selectivity to CHOH is given by 1 Nitrogen Flow Rate (Carrier Gas) 0.0024 mol/min
XT
Experiment Air (mol/min) Total Flow rate (mol/min)

2.3.1 External Mass Transfer Experiments I 0.0082 0.0116


II 0.0108 0.0142
It is known that external diffusion has greater effects at high
temperatures and low flow rates. The external diffusion effects III 0.0133 0.0167
were investigated by changing the air flow rate, while keeping IV 0.0158 0.0192
the temperature, catalyst amount and methanol flow rate con-
V 0.0182 0.0216
stant. The experiments were performed at the highest tempera-
ture, which is 375 °C and with a space time of 4.9 g/(mol/h). VI 0.0206 0.0240
During the experiments, the total flow rate of the reactant
VII 0.0230 0.0264
stream was increased by increasing the flow rate of air and
following this, formaldehyde and total conversions were calcu- VIII 0.0254 0.0288
lated. The operating conditions for these experiments are IX 0.0277 0.0311
detailed in Tab. 2.
X 0.0299 0.0333

2.3.2 Internal Mass Transfer Experiments experiments, the first experiment of the set was repeated. The
results showed that the catalysts did not lose activity during
Internal diffusion is the mass transfer of the reactants from the experimental studies. IR spectroscopy and adsorption-de-
outside of the surface into the pores of the catalyst. In this part sorption isotherms of fresh and used catalysts were used to
of the study, the effect of internal diffusion was examined by examine the possible change in the structure of catalysts. The
keeping the temperature and space time (4.9 g/(mol/h)) con- example IR spectroscopy results for the fresh and used catalysts
stant, and changing the catalyst diameter. Three different cata- with silica support and Mo/Fe = 5 are given in Fig. 1. The
lyst diameters were used at the lowest and the highest tempera- main peaks are observed at 1620 cm–1 (H2O bending band),
tures of 250 and 375 °C, respectively. The total conversions and 3400 cm–1 (–OH stretching band), and at 1400 cm–1 and
reaction rates were calculated at the end of the experiments. 2400 cm–1 (CH stretching band). It can be seen from the spec-
The operating conditions for these experiments are listed in tra in Fig. 1, that there is no structural difference between the
Tab. 3. fresh and the used catalysts.

Chem. Eng. Technol. 2010, 33, No. 1, 167–176 © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.cet-journal.com
170 A. H. Ulukardesler et al.

Table 3. Operating conditions for internal diffusion experiments. Table 4. Results for fresh and used catalysts with Mo/Fe = 5 sup-
ported by alumina and silica.
Catalyst Support Silica
Surface Area (m2/g)
Catalyst Amount 1g
Catalyst Support BET T-Plot Langmuir D-R-A
Methanol Flow Rate 0.0034 mol/min
Fresh Alumina 100.17 134.27 139.96 108.27
Air Flow Rate 0.0277 mol/min
Used Alumina 121.79 184.49 173.76 148.81
Nitrogen Flow Rate (Carrier Gas) 0.0024 mol/min
Fresh Silica 354.89 549.30 500.80 439.89
Experiment Temperature (°C) Catalyst Diameter (mm)
Used Silica 419.92 581.12 585.91 567.35
I 250 3.034
II 250 2.746
Although there was no structural analysis undertaken in this
III 250 2.240 study, the increase in the surface area and the change in the
IV 375 3.034 pore distribution might be the result of the migration of
molybdenum particles and their adhesion to the outside sur-
V 375 2.746
face of the catalyst. Andersson et al. [7] have also studied this
VI 375 2.240 subject and they found that during the operation of the cata-
lyst there is migration of Mo species from the upper part of
the reactor towards the outlet. They concluded that the deacti-
The surface areas of the fresh and used catalysts with vation is primarily due to formation of volatile species formed
Mo/Fe = 5 supported by alumina and silica, were measured by the MoO3 surface reacting with methanol, resulting in a de-
using an Omnisorp 100CX Volumetric Adsorption Analyzer. crease of the MoO3/Fe2(MoO4)3 mole ratio in the catalyst [7].
The samples were degassed at 300 °C for 3 h. The results for
the fresh and used catalysts are listed in Tab. 4. The sample
result of the analysis for the catalyst supported by silica and 3.2 Testing of the Catalysts
the static adsorption and desorption isotherm plot of these
catalysts are shown in Fig. 2. In addition, the pore size distri- It if known that for methanol oxidation, both the selectivity to
butions of these catalysts are given in Fig. 3. Similar findings formaldehyde and total conversion depend highly on the reac-
were also observed from the catalysts with alumina support. It tion temperature, space time, Mo/Fe ratio of the catalyst and
can be seen from the results given in Tab. 4 that the surface the type of support [16].
areas increased after experiments for the catalysts supported
by both silica and alumina. In addition, there is a substantial
change in pore size distribution of the fresh and used catalyst 3.2.1 Effect of Temperature
supported by both silica and alumina. This can be explained
due to the change in the pore size. The effect of temperature on both selectivity and total conver-
sion was investigated with all prepared catalysts. The results

100.0 100.0

80.0 80.0

60.0 60.0

40.0 40.0

20.0 20.0

0.0 0.0
3000 2000 1500 1000 500 400 3000 2000 1500 1000 500 400

(a) Fresh catalyst (b) Used catalyst


Figure 1. Infrared spectra of silica supported fresh and used catalysts with Mo/Fe = 5.

www.cet-journal.com © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eng. Technol. 2010, 33, No. 1, 167–176
Methanol oxidation 171

500 500

400 400

Volume adsorbed, (cc/g. STP)


Volume adsorbed, (cc/g. STP)

300 300

200 200

100 100

0.0 0.0
0.0 .2 .4 .6 .8 1 0.0 .2 .4 .6 .8 1
P/P0 P/P0

(a) Fresh catalyst (b) Used catalyst


Figure 2. Static adsorption/desorption isotherm plot of silica supported fresh and used catalysts with Mo/Fe = 5.

40 400

30 300
DVP/DRP
DVP/DRP

20 200

10 100

0.0 0.0
100 1000 10 100 1000
RP (A) RP (A)

(a) Fresh catalyst (b) Used catalyst


Figure 3. Pore size distribution of the fresh and used catalysts supported by silica with Mo/Fe = 5.

obtained for a catalyst with Mo/Fe = 1.5 and for all W/FA0 val- 300 °C, this effect may be suppressed by controlling the equi-
ues are presented in Fig. 4. It can be seen from Fig. 4 that the librium. Similar results were obtained for the other catalysts
total conversion increased as the temperature was increased, [14].
whereas selectivity decreased on increasing the temperature.
The increase in total conversion and decrease in selectivity
with increasing temperature may be due to the fact that in- 3.2.2 Effect of Spacetime
creasing temperature favors the total oxidation reaction. Total
conversion increased up to 300 °C and beyond this tempera- It can also be seen from Fig. 4 that for all the temperatures
ture it remained almost constant at ca. 0.6. Between the tem- studied, the total conversion and selectivity both increase on
perature ranges of 250–275 °C, the selectivity was ca. 1, after increasing space time. In addition, at constant temperature
which it decreased slightly. and Mo/Fe ratio, the selectivity increased on increasing space
It is known that temperature has an important effect on the time. The same trend was also observed for the catalysts pre-
reaction rate. Between 250 °C and 300 °C this effect was very pared on alumina. Although the results obtained for the space
obvious due to the possible control by the reaction. Above time values of 20.25 and 33.75 g/(mol/h) are very close at the

Chem. Eng. Technol. 2010, 33, No. 1, 167–176 © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.cet-journal.com
172 A. H. Ulukardesler et al.

1 much more active than that with Mo/Fe = 1.5


[15]. The same trends were obtained in the current
study.
0.9

0.8 3.2.4 Effect of Support Type

0.7 The effect of the support type is shown in Fig. 6


for a space time of W/FA0 = 20.25 g/(mol/h) and
at 300 °C. Both supports exhibited the same trend
0.6
for all Mo/Fe ratios. It can be determined from
Fig. 6, that the silica support gives better results
0.5 Sel. W/FA0=50.63 for conversion and selectivity than the alumina
Sel. W/FA0=33.75
Sel. W/FA0=20.25 support, for all experiments.
0.4 Conv. W/FA0=50.63
Conv. W/FA0=33.75
Conv. W/FA0=20.25
0.3 3.3 Kinetic Study
250 275 300 325 350 375
3.3.1 External Mass Transfer Experiments
Temperature, (°C)

Figure 4. Change in total conversion and selectivity with temperature and W/FA0
It can be seen from Fig. 7, that when the flow rate
for silica supported catalyst with Mo/Fe = 1.5. is increased, the total conversion increases and
then remains constant at ca. 36 %. Since the flow
rate corresponding to constant conversion was
point of total conversion, the results for the space time value 0.0277 mol/min air, all kinetic experiments were performed at
of 50.63 g/(mol/h) are higher than those for the experiments flow rates higher than this value. Therefore, according to the
using the lower space time values. The same finding is also results obtained at the highest temperature, it can be seen that
determined by the selectivity results. The change in total con- external diffusion effects can be neglected at temperatures be-
version with changing space time has a similar trend for all low 375 °C, with an air flow rate higher than 0.0277 mol/min.
temperatures but the selectivity does not maintain the same In addition, the effect of external diffusion was investigated
trend. At increasing temperature values, the increase in the theoretically. The criterion proposed by Hougen to test exter-
interval for the selectivity as the space time increases, can be nal mass transfer resistance was taken into consideration [16].
explained by the fact that the overall
reaction is determined by the first reac-
1
tion.

0.9
3.2.3 Effect of Mo/Fe Ratio
0.8
The effect of Mo/Fe ratio for a space
time of 33.75 g/(mol/h) with alumina 0.7
support and the temperatures studied
can be seen in Fig. 5. As the Mo/Fe ra- 0.6
tio increase, both total conversion and
selectivity also increase. For Mo/Fe = 5, 0.5
the conversion is between 70–80 %,
whereas for the other two ratios of 0.4 Sel. M o/Fe=1.5
Mo/Fe, this range is between 50–60 %. Sel. M o/Fe=3
It can be said that Mo/Fe = 5 favors the 0.3 Sel. M o/Fe=5
first reaction although the temperature Conv. M o/Fe=1.5
for the reaction is further increased 0.2 Conv. M o/Fe=3
above 300 °C. The experiments per- Conv. M o/Fe=5
formed with alumina support demon- 0.1
strated the same trend.
250 275 300 325 350 375
Soares et al. [15] studied methanol
oxidation using a Mo/Fe catalyst. They
Temperature, (°C)
found that the conversion of methanol
increased with increasing temperature Figure 5. Change in total conversion and selectivity with Mo/Fe ratios for alumina supported
and the catalyst with Mo/Fe = 3 was catalyst at W/FA0 = 33.75 g/(mol/h).

www.cet-journal.com © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eng. Technol. 2010, 33, No. 1, 167–176
Methanol oxidation 173

Table 5. Results for the theoretical calculation of the external dif-


1 fusion experiments.
0.9
T (°C) Air Flow rate (mol/min) Cb Cs
0.8 Cb
0.7
375 0.0082 2.21 · 10–8
0.6
0.5 375 0.0299 1.52 · 10–7
0.4
0.3 the results are given in Fig. 8. It can be seen from Fig. 8 that
0.2 internal diffusion effects may not be neglected.
0.1 In order to attempt to determine the internal diffusion
0 effect, theoretical calculations were performed by following the
1.5 3 5 Mo/Fe Ratios procedure described elsewhere [17]. The effectiveness factors
Si-Conversion Al-Conversion were calculated for all the temperatures studied and are listed
Si-Selectivity Al-Selectivity
in Tab. 6.
Figure 6. Change in total conversion and selectivity with respect
to Mo/Fe ratios at W/FA0 = 20.25 g/(mol/h) and 300 °C.
Table 6. Results for the theoretical calculation of the internal dif-
fusion experiments.
0.40 T (°C) g (–)

250 0.980
0.35
275 0.976
Total Conversion, %

0.30 300 0.971


325 0.965
0.25 350 0.961
375 0.956
0.20

0.15 3.3.3 Kinetic Experiments


X total
0.10 The kinetic experiments were performed at six temperatures
0.01 0.02 0.03 and five values of W/FA0. The reaction rates were calculated
from the experimental results and the calculated rates are listed
F (methanol+air) mole/min
in Tab. 7. The experiments carried out for this purpose were
Figure 7. Results for the external diffusion experiments. evaluated and the conversions, as explained before, were calcu-

Therefore, the condition necessary to neglect the effect of 0.5


external diffusion is given by Eq. (6): 250 C
375 C
Cb Cs 0.4
< 0:1 (6)
Cb
Total Conversion, XT

0.3
where Cb is the molar concentration of methanol in the
gas phase [mol/cm3], and Cs is the molar concentration of
methanol at the catalyst surface [mol/cm3]. 0.2
The results of the theoretical calculation of the external
diffusion effect are detailed in Tab. 5. It can be seen from
0.1
Tab. 5 that external diffusion resistances can be neglected.

0
3.3.2 Internal Mass Transfer Experiments 1.0 1.1 1.2 1.3 1.4 1.5 1.6
Catalyst Radius, (mm)
The results of the experiments conducted for internal mass
transport effects, as explained above, were evaluated, and Figure 8. Results for the internal diffusion experiments.

Chem. Eng. Technol. 2010, 33, No. 1, 167–176 © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.cet-journal.com
174 A. H. Ulukardesler et al.

Table 7. Results for the kinetic experiments. 3.3.4 Kinetic Model Study

T (°C) W/FA0 (gcat/mol) XT rA (XT/(W/FA0) The data collected in the experiments were used to determine
250 0.51 0.0278 0.055 the kinetics of the reaction. The studied kinetic models were
derived by using tables from the literature assuming one of the
250 0.68 0.0358 0.053
consecutive steps as a rate limiting step, as mentioned by Fog-
250 1.69 0.0692 0.041 ler [18]. Following the derivation of these models, the expres-
sions were linearized in order to find the kinetic parameters.
250 2.53 0.0603 0.034
In addition, the simplest case, i.e., the Power Law Model, was
250 3.38 0.0601 0.028 tested for the reaction, as represented in three different forms
275 0.51 0.0289 0.057 in Eq. (7).

275 0.68 0.0378 0.056 1⁄2O + CH3OH > CHOH + H2O (7a)
2
275 1.69 0.0827 0.049
A+B>R+S (7b)
275 2.53 0.1038 0.041
275 3.38 0.1148 0.034 Oxygen + methanol > formaldehyde + water (7c)
300 0.51 0.0309 0.061
The model equations are listed in Tab. 8. In order to check
300 0.68 0.0432 0.064 the suitability of the tested models, the following restrictions
300 1.69 0.0962 0.057 should be satisfied simultaneously:
– The reaction rate and adsorption constants cannot be nega-
300 2.53 0.1215 0.048
tive;
300 3.38 0.1350 0.040 – The change of reaction rate constants with temperature
should obey the Arrhenius equation;
325 0.51 0.0354 0.070
– The adsorption coefficients should decrease with increasing
325 0.68 0.0486 0.072 temperature, and
325 1.69 0.1148 0.068 – The results of the model equations should fit the experimen-
tal results well.
325 2.53 0.1443 0.057 By performing the required numerical regression analysis
325 3.38 0.1654 0.049 for the seven model equations, the reaction rates and their
kinetic parameters were calculated. A numerical regression
350 0.51 0.0425 0.084
method was applied depending on the minimization of the
350 0.68 0.0614 0.091 square of the errors explained by R(rexp – rmodel)2.
Thus, the percentage errors between the experimental rate
350 1.69 0.1384 0.082
and rate obtained from the model equations were determined.
350 2.53 0.1747 0.069 By taking the arithmetic mean value of this percentage error,
350 3.38 0.1992 0.059 Tab. 9 has been prepared to compare the model equations.
As it can be seen from Tab. 9, models 4 and 6 gave better re-
375 0.51 0.0501 0.099 sults compared to the other kinetic model equations. Although
375 0.68 0.0689 0.102 the error percentage of model 4 is also low, the kinetic parame-
ters of this model do not satisfy some of the restrictions men-
375 1.69 0.1553 0.092
tioned above. In model 6, surface reaction control, with and
375 2.53 0.1899 0.075 without the dissociation of oxygen and without adsorption of
375 3.38 0.2160 0.064
methanol, was seen to be a better fit for the kinetic data than
the other models. It gave an average error of 1.52 %, and the
Arrhenius factor and activation energy were determined as
5146.13 and 2.71 kcal/mol, respectively. The kinetic constants
lated. If the conversion values are included, it can be seen that obtained for model 6 are listed in Tab. 10.
the reactor exhibits considerably different behavior. Therefore, A broad change of activation energy can be seen with the
the reaction rates were calculated by considering the behavior different type of catalysts and are discussed elsewhere by Kim
of the reactor. The partial pressures of the species were also et al. [1] and Waterhouse et al. [4]. The activation energy
calculated. In addition, the values of the equilibrium constants found in this study is at the lower range found in the literature.
at the different temperatures were calculated theoretically. The This finding emphasizes the fact that the catalyst used is very
arithmetic mean of the inlet and outlet partial pressures of active, and that it results in a lowering of the energy barrier
oxygen, methanol and nitrogen were used as the partial pres- that must be overcome.
sures of the species involved. Soares et al. studied the kinetics of methanol oxidation over
an iron-molydate catalyst. They found that the kinetic mecha-

www.cet-journal.com © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eng. Technol. 2010, 33, No. 1, 167–176
Methanol oxidation 175

Table 8. Kinetic model equations [10].

Kinetic Model Definition

Power Law r = kPAmPBn

Adsorption of A constant without dissociation (Model 1) kA PA


rA ˆ h KA PR PS
i
…1 ‡ ‡ KB PB ‡ KR PR ‡ KS PS †
KPB

Adsorption of A constant without dissociation (Model 2) kA PA


rA ˆ  r 
KA PR PS
…1 ‡ ‡ KB PB ‡ KR PR ‡ KS PS †2
KPB

Adsorption of B constant (Model 3) kB PB


rA ˆ h K P P
i
…1 ‡ B R S ‡ KA PA ‡ KR PR ‡ KS PS †
KPA

Surface reaction constant without dissociation of A (Model 4) PA PB 1


ˆ …1 ‡ KA PA ‡ KB PB ‡ KR PR ‡ KS PS †2
rA ksr KA KB

Surface reaction constant with dissociation of A (Model 5) P P 1


… A B †1=3 ˆ …1 ‡ KA PA ‡ KB PB ‡ KR PR ‡ KS PS †
rA …ksr KA KB †1=3

Surface reaction contstant, with or without dissociation of A, B PA PB 1


ˆ …1 ‡ KA PA ‡ KB PB ‡ KR PR ‡ KS PS †2
but no adsorption (Model 6) rA ksr KA

Table 9. Comparison of the model equations. nism of the reaction depends on the reaction conditions over-
all, and mainly the temperature [19]. The results in Fig. 9
Model Name Mean Error Percentage (%)
summarize the experimental reaction rates and the calculated
Power Law 8.91 rates using the equations of model 6.
Model 1 19.15
Model 2 42.31 0,10
250 275
Model 3 38.45
300 325
Model 4 3.75 350 375
0,08
Model 5 9.93
Model 6 1.52 0,06
rAmodel

Table 10. Kinetic constants calculated for model 6. 0,04

T (°C) KA KB KR KS ksr

250 2620 1800 1.20 1.30 400 0,02

275 2070 1770 1.15 1.20 415.14 ∑R2=0.9962


300 1870 1750 1.09 1.12 450.29 0,00
0,00 0,02 0,04 0,06 0,08 0,10
325 1650 1724 1 1.05 512.23
rAexp.
350 1621 1637 0.99 1.02 595.78
Figure 9. Deviations for experimental and model rates for
375 1600 1526 0.93 0.985 639.12
model 6.

Chem. Eng. Technol. 2010, 33, No. 1, 167–176 © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.cet-journal.com
176 A. H. Ulukardesler et al.

Acknowledgement [4] G. I. N. Waterhouse, G. A. Bowmaker, J. B. Metson, Appl.


Catal., A 2004, 266, 257.
This study was supported by the Ege University Research Fund [5] M. Qian, M. A. Liauw, G. Emig, Appl. Catal., A 2003, 238,
through the project, “2000/MUH/039-Catalytic Oxidation of 211.
Methanol to Formaldehyde”. [6] L. E. Wade, R. B. Gengelbach, J. L. Trumbley, W. L. Hall-
bauer, in Encyclopedia of Chemical Technology (Eds: H. Mark,
D. F. Othmer, C. G. Overberger, G. T. Seaborg), John Wiley
Symbols used & Sons, New York 1981, pp. 398-415.
[7] A. Andersson, M. Hernelind, O. Augustsson, Catal. Today
W [g] total catalyst mass 2006, 112, 40.
FA0 [mol/h] feed flow rate [8] R. L. McCormick, M. B. Al-Sahali, G. O. Alptekin, Appl. Cat-
X [–] conversion al., A 2002, 226, 129.
Cb [mol/cm3] molar concentration of [9] P. Jiru, B. Wichterlova, J. Tichy, in Proc. of the 3rd Int. Con-
methanol in the gas phase gress on Catalysis, Royal Dutch Chemical Society, Amsterdam
Cs [mol/cm3] molar concentration of 1965, p. 1.
methanol at the catalyst surface [10] E. Santacesaria, M. Morbidelli, S. Carra, Chem. Eng. Sci.
rA [mol/gcat h] rate of reaction 1981, 36 (5), 909.
K [–] equilibrium constants [11] S. Yao et al., Appl. Catal., A 2000, 198, 43.
[12] P. Forzatti, G. Buzzi-Ferraris, Ind. Eng. Chem. Process Des.
Dev. 1982, 21 (1), 67.
Greek symbol [13] L. Cairati, L. di Fiore, US Patent 4 181 629, 1980.
[14] A. H. Yilmaz, Ph.D. Thesis, Ege University, Izmir, Turkey
g [–] effectiveness factor 2004.
[15] A. P. V. Soares et al., Appl. Catal., A 2001, 206, 221.
[16] J. M. Smith, Chemical Engineering Kinetics, McGraw-Hill,
References New York 1981.
[17] G. F. Froment, K. B. Bischoff, Chemical Reactor Analysis and
[1] T. H. Kim et al., Catal. Lett. 2004, 98, 161. Design, John Wiley & Sons, New York 1990.
[2] A. P. V. Soares, M. Farinha Portela, A. Kiennemann, Catal. [18] H. S. Fogler, Elements of Chemical Reaction Engineering,
Rev. 2005, 47 (1), 125. Prentice-Hall, Englewood Cliffs, NJ 1999.
[3] S. A. R. K. Deshmukh, M. van Sint Annaland, J. A. M. Kui- [19] A. P. V. Soares, M. Farinha Portela, A. Kiennemann, Stud.
pers, Appl. Catal., A 2005, 289, 240. Surf. Sci. Catal. 2001, 133, 489.

www.cet-journal.com © 2010 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eng. Technol. 2010, 33, No. 1, 167–176

You might also like