You are on page 1of 12

Applied Catalysis A: General 255 (2003) 289300

Influence of the reaction conditions on the activity properties


of vanadiumaluminium oxynitride propane
ammoxidation catalyst
Mihaela Florea, Ricardo Prada Silvy, Paul Grange
Universit Catholique de Louvain, Catalyse et Chimie des Matriaux Diviss, Croix du Sud 2/17, 1348 Louvain-la-Neuve, Belgium
Received 25 June 2003; received in revised form 10 July 2003; accepted 13 July 2003

Abstract
The influence of different reaction conditions on the catalytic activity of vanadiumaluminium oxynitride VAlON in
propane ammoxidation is investigated. This new catalytic material exhibits amorphous character, high specific surface area,
basic-redox properties and remarkable acrylonitrile production per hour and amount of catalyst. Optimal performance is
achieved when the reaction temperature is 500 C and the molar ratio of propane, oxygen, and ammonia in the gas feed is
1.25:3:1. Under these optimal reaction conditions, the catalyst showed a propane conversion level of 60%, an acrylonitrile
selectivity of 56%, and an acrylonitrile productivity of 812 l/kg h. The formation of a propylene intermediate compound as
well as hydrogen cyanide was detected only for those reaction conditions where the NH3 :C3 H8 ratio was lower than 1. This
suggests differences in the reaction mechanism with respect to conventional VMe-oxide propane ammoxidation catalysts.
In propane ammoxidation over VAlON catalysts, ammonia seems to play a double role: (i) as a reagent necessary for the
thermal nitridation and generation of nitrogen species, (ii) participating in N-insertion step for the acrylonitrile formation from
propane. The catalytic performances depend on the generation of nitrogen species as a function of the reaction conditions,
which determine the level of propane conversion and selectivity to the reaction products.
2003 Elsevier B.V. All rights reserved.
Keywords: Oxynitride catalysts; Nitrogen species; Contact time; Propane ammoxidation

1. Introduction
The high abundance and lower cost of light
paraffins, in comparison with olefins and acetylene
compounds, has motivated the development of new
catalytic processes to use this feedstock for different
applications in the chemical industry. One example
of this is acrylonitrile (ACN) production, which is
widely used as a chemical intermediate in the produc-

Corresponding author. Fax: +32-10-473649.


E-mail address: grange@cata.ucl.ac.be (P. Grange).

tion of synthetic fibers, rubbers, nitriles, resins and


other important products.
Actually, acrylonitrile is mainly produced by propylene ammoxidation [1]. Around 80 plants located in 22
countries produce more than 5 million metric tons of
acrylonitrile per year. An alternative to this industrial
application is to replace the current propylene ammoxidation technology by propane, which was proposed
by the Standard Oil Company (SOHIO) in 1970.
However, the propane ammoxidation route shows
some difficulties. The lower reactivity of alkanes
requires very active and stable catalysts and severe
operating conditions (higher reaction temperature and

0926-860X/$ see front matter 2003 Elsevier B.V. All rights reserved.
doi:10.1016/S0926-860X(03)00591-X

290

M. Florea et al. / Applied Catalysis A: General 255 (2003) 289300

contact times). Under these conditions, acrylonitrile


yield can be affected by catalytic and thermal decomposition leading to the formation of undesirable
carbon and nitrogen oxygenated compounds.
Propane ammoxidation reaction requires catalysts
with specific surface properties for alkane activation
molecules and the subsequent nitrogen insertion reaction. The development of improved catalytic systems for this reaction is thus necessary. According to
Sokolovskii et al [2], the hydrocarbon activation is the
controlling step in the ammoxidation reaction mechanism. According to these authors, this reaction should
occur via heterolytic CH bond dissociation by interaction of the hydrocarbon with acidbase neighbouring sites of the catalyst leading to carbanion species
(C3 H7 ) formation.
This surface intermediate could be subsequently
activated on a basic site to form an allyl-type
(C3 H5 ) intermediate which would react with activated ammonia-adsorbed species (NHx ) to form
acrylonitrile. The reactivity of the catalyst depends on
the activated surface concentration species. Therefore,
optimal reaction conditions (temperature, contact
time) and balance among the reactants composition
in the gas feed are absolutely necessary.
Different catalytic systems have been developed to perform the propane ammoxidation reaction: (i) V-antimonate promoted with W, Al, Fe,
etc., which present rutile-like structures [36]; (ii)
BiV-molybdate with scheelite structure [79]; (iii)
vanadyl pyrophosphate catalysts [1012]; and (iv)
Ga-antimonate pure or modified with W, Ni and P
[13,14]. Most of these formulations contains vanadium as key element. Among these catalyst formulations, VSbWAl oxides system [15] and VMo
mixed oxides modified with Nb and Te [16,17],

have shown the best catalytic performance (acrylonitrile yields between 35 and 55%). However, these
multi-component catalyst systems are complex and
the literature does not provide clear information about
the direct contribution of each metallic element on
the catalytic activity.
Using vanadyl pyrophosphate-based catalysts, Centi
et al [12] observed that the acrylonitrile selectivity
considerably depends on the surface concentration of
the ammonia species. Other catalytic results obtained
using V-antimonate catalysts [18,19] evidenced that
the acrylonitrile selectivity depends on the gas feed
composition and the nature of the catalyst system.
In some patents, a gas feed composition with higher
concentration of oxygen and ammonia compared to
propane was used, while a recent published patent
[20] claims a gas feed composition rich in propane
(propane:oxygen:ammonia ratio of 5:2:1) that delivers
optimal catalytic activity.
Table 1 summarizes a series of vanadium-based
propane ammoxidation catalysts, which clearly show
differences in gas feed composition for obtaining optimal performance.
We have previously described the behavior of
vanadiumaluminium oxynitride catalysts in propane
ammoxidation reactions [22]. This novel nitrogencontaining catalytic system showed high selectivity
and productivity in acrylonitrile. As the nature and
concentration of surface species depend on the reaction conditions, in this work, we investigated the
effect of the reaction temperature (450550 C) and
the gas feed composition on the catalytic activity
for the propane ammoxidation reaction. This kind of
experiment may help to understand the activity behavior of this new material under different operating
conditions and optimize its catalytic functionality.

Table 1
Reaction conditions corresponding to different conventional propane ammoxidation catalysts
Catalyst system

Treaction
( C)

C3 H 8
(mol)

O2 (mol)

NH3 (mol)

Selectivity
ACN (%)

Conversion
C3 H8 (%)

Reference

VMoNbTeOx
VSbOx
VSbAlOx
VSbAlOx
VSbWAlOx
VSbAlOx
VSbPWAlOx

520
480
500
480
480
500
470

1.0
1.0
1.0
1.0
1.0
1.0
5.0

2.8
2.0
2.0
2.0
2.0
5.0
1.0

1.2
1.0
2.0
1.0
1.0
5.0
2.0

56
37
51
45
48
50
17

52
25
44
30
77
45
20

[28]
[4]
[5]
[15]
[15]
[18]
[20]

M. Florea et al. / Applied Catalysis A: General 255 (2003) 289300

2. Experimental
2.1. Catalyst preparation
Vanadiumaluminium oxide precursor, with V:Al
ratio of 0.25, was prepared by coprecipitation of solutions containing 102 M ammonium meta-vanadate
and 102 M aluminium nitrate with ammonia (25%) at
pH 5 and 60 C. The obtained precipitate was washed
with ethanol and isopropanol and dried in a vacuum
oven at 60 C under a pressure of 30 mbar. Thermal
nitridation of the vanadiumaluminium oxide was realized in situ at 500 C under different reaction mixtures. The heating rate was 10 C min1 .
2.2. Physico-chemical characterization
The measurements of the BET surface area using
single-point method were performed on a Micromeritics Flow Sorb II 2300.
XRD spectra were recorded using a Siemens
D-5000 powder diffractometer using Ni-filtered Cu
K radiation ( = 1.5418 ).
The total nitrogen content (NT ), including bulk and
surface nitrogen, was determined using the method
described by Guyader et al. [23]. Briefly, this method
consists in the measurement of ammonia produced by
the reaction between melted potassium hydroxide and
nitrogen (nitride N3 or NHx surface species), following the equation
xN3 + 3xKOH xNH3 +

3x
3x
+
2K2 O 2O2

A mixture of 20 mg sample and 5 g KOH was heated


at 400 C. The released ammonia was carried away by
a nitrogen flow (50 ml/min) and the output was connected to a trap containing distilled water and titrated
with a sulfuric acid solution of 102 N.
2.3. Activity measurements
The catalytic tests were performed in a fixed-bed
quartz micro reactor at atmospheric pressure, on-line
coupled with a gas chromatograph (Hewlett Packard
5890 Serie II), equipped with FID and TCD detectors. The organic compounds were separated on a capillary column WCOT fused silica CP 8CB, and O2 ,

291

CO2 , CO, N2 were analyzed on two columns, MS 5A,


80100 mesh (2 m 1/8 in.) and Haysep Q 100120
mesh (2 m 1/8 in.). In order to have isothermal conditions, the catalysts (particle size: 100 m) were diluted with quartz beads having the same particle size.
An on-line mass spectrometer was used to check the
NOx formation.
The composition of the reaction mixture was variable, with molar ratios C3 H8 :O2 :NH3 of 1:1:1 to
1.5:4:1.5. Different reaction conditions were used to
evaluate the effect of temperature, feed composition
and space velocity on the performance of the catalyst.
To avoid the condensation or the polymerization of
acrylonitrile, the connection between the reactor and
the chromatograph was heated to 150 C. However, the
mass balance observed evidenced that these problems
cannot be fully resolved.
Propane conversion is defined as the % ratio between the mole of propane consumed per mole of
propane in the feed; the ACN selectivity as mole of
ACN in the product per mole of propane consumed;
and the ACN yield as mole of ACN in the product per
mole of propane in the feed.

3. Results
3.1. Catalyst preparation and characterization
The oxide precursor dried at 60 C is totally amorphous to XRD and presents a high surface area
of 250 m2 /g. For the preparation of the vanadium
aluminium oxynitride, the precursor was heated under reaction mixture, namely ammonia, propane and
oxygen, at 500 C, for 24 h corresponding to the reaction time. No calcination was performed before
nitridation, because when the vanadiumaluminium
precursor is heated under air, it crystallizes at 550 C
in AlVO4 and the surface area drops to 0.1 m2 /g. No
nitrogen can be incorporated at low temperature once
the crystallization has occurred [24].
The oxynitrides obtained after reaction are amorphous and exhibit a specific area between 130
160 m2 /g as function of the activation conditions, as
is presented in Table 2. Analysis of total nitrogen
content on the catalysts after the activation process
showed values from 1.8 to 4.9 wt.% nitrogen, as
function of the ratio of the reactants used (Table 2).

Table 2
Surface area and nitrogen content of VAlON catalysts as function
of the molar reactants ratio
C3 H8 :O2 :NH3
ratio

Surface area
(m2 /g)

Total nitrogen
content (wt.%)

1:1:1
1:2:1
1:3:1
1:3.5:1
1:4:1
1.25:3:0.75
1.25:3:1
1.25:3:1.5

158
156
154
140
130
146
150
140

1.8
2.2
3.2
2.8
2.6
2.2
4.5
4.9

The amorphous character, the high surface area and


the presence of bulk nitrogen species on the used
catalysts represent important novel physico-chemical
characteristics with respect to the conventional catalysts used in propane ammoxidation and reported in
the literature.
3.2. Catalytic activity
The catalytic test performed in absence of the catalyst at 500 C, indicated a conversion of 5% and the
formation of carbon oxides as reaction product.
Analysis of the reaction products in the presence of
the catalyst always indicated the following products:
acrylonitrile, acetonitrile (AcCN) and carbon oxides
(COx = CO + CO2 ). Traces of butanedinitrile have
also been identified but not quantified. Small amounts
of nitrogen as reaction product were observed due
to the oxidation of ammonia as parallel reaction. No
propylene and hydrogen cyanide formation was detected for the sample tested under optimal reaction
condition. A carbon mass balance of 8090% was obtained due to the polymerization and/or condensation
of acrylonitrile inside of the connection-line between
the reactor and the gas chromatograph.
Total oxygen conversion was reached even when
the oxygen to propane molar ratio was higher than the
stoichiometric value.
3.2.1. Activity evolution as a function of time on
stream
Fig. 1 shows the conversion of propane and the selectivities to the reaction products as a function of time
on stream in the following conditions: molar ratios of
C3 H8 :O2 :NH3 = 1.25:3:1 and 500 C. The conversion

conversion and yield (%)

M. Florea et al. / Applied Catalysis A: General 255 (2003) 289300

100
80
60
40
20
0

selectivities (%)

292

10

20

30

40

50
60
time (h)

10

20

30

40

50

100
80
60
40
20
0
60
time(h)

Fig. 1. Catalytic behavior over VAlON catalyst (V:Al = 0.25)


as function of time on stream: GHSV = 16.8 l/g h,
C3 H8 :O2 :NH3 = 1.25:3:1, 500 C (( ) propane conversion; ( )
yield ACN; () selectivity to ACN; () selectivity to AcCN; ()
selectivity to COx ; and () selectivity to N2 ).

of propane is almost constant with time, while the selectivity to acrylonitrile increases with time on stream,
as presented in Fig. 1. Acrylonitrile yield is also reported. Steady state conditions are reached after 12 h
of reaction. Under steady state conditions, propane
conversion is 60% and ACN selectivity is 56%. In parallel, the selectivity to COx decreased with time on
stream, meaning that the surface composition of the
catalyst changes during the initial reaction step and
the total oxidation is limited in time. Propylene and
hydrogen cyanide formation was not observed during
the 24 h of reaction.
No deactivation of the catalyst occurred during
60 h of time on stream, which shows the stability
of the VAlON catalyst in the propane ammoxidation
reaction.
To confirm this stability of the VAlON catalyst and
to prove the reproducibility of the experiment, a reaction of 60 h was performed, and the conversion of

M. Florea et al. / Applied Catalysis A: General 255 (2003) 289300


Table 3
Total nitrogen content, ACN selectivity and propane conversion as
function of time on stream (same reaction conditions as in Fig. 1)
Time (h)

(wt.%)a

N content
ACN selectivity (%)
C3 H8 conversion (%)
a

0.5

10

12

24

1.8
12.3
56.9

3.1
24.7
58.6

3.8
35.1
60.8

4.2
46.0
60.8

4.5
54.5
60.8

4.5
56.6
60.8

Total nitrogen content.

propane and the selectivity to the reaction products


were constant with time on stream (Fig. 1).
Several reactions were performed under the same
experimental conditions and the reaction was stopped
after different times on stream. The catalyst was then
withdrawn from the reactor and the nitrogen content
was analyzed.
The results concerning the evolution of total nitrogen content (Grekov analysis) on the VAlON during propane ammoxidation are presented in Table 3.
Data presented in this table indicate an increase of
the nitrogen content of the catalyst until the catalyst
reached the steady-state conditions (12 h). After this
time on stream, no change in the nitrogen content was
observed.

conversion and selectivities(%)

3.2.2. Effect of the reaction temperature


For the formation of unsaturated nitriles, such as
acrylonitrile from propane, the hydrocarbon activation
is an important step, which proceeds at rather high
temperatures compared with the formation of acry-

293

lonitrile from propylene. In order to identify the optimal range of temperature for propane ammoxidation
over vanadiumaluminium oxynitrides, catalytic tests
were carried out in the range of 400550 C. Fig. 2
shows the conversion of propane and product selectivities as a function of the reaction temperature over
the vanadiumaluminium-based catalyst. The conversion of propane increased continuously with reaction
temperature, while the selectivity to acrylonitrile increased with the temperature up to 500 C and then
decreased. Under these conditions, the maximum in
selectivity to acrylonitrile corresponding to 55% was
obtained at 500 C.
The selectivity to COx exhibited an opposite variation, namely, it progressively decreased up to 500 C,
while above this temperature most of the propane was
oxidized. After 500 C, traces of heavier hydrocarbons
such as butanedinitrile were also detected.
3.2.3. Effect of space velocity
The influence of the space velocity on the propane
ammoxidation was investigated by varying the amount
of the catalysts and keeping the total flow rate constant. Fig. 3 shows the changes in the conversion of
propane and selectivities to the reaction products as a
function of space velocity in steady-state conditions.
The selectivity to acrylonitrile increases with increasing space velocity, up to a value of 56%, while the
propane conversion diminishes with increasing space
velocity. These dependencies demonstrate that longer
residence times of propane in the gas phase lead to a
total oxidation of alkane.

100
80
60
40
20
0
400

450

500

550
temperature ( C)

Fig. 2. Catalytic behavior over VAlON catalyst (V:Al = 0.25) as function of the reaction temperature: GHSV = 16.8 l/g h,
C3 H8 :O2 :NH3 = 1.25:3:1 (( ) propane conversion; () ACN selectivity; () AcCN selectivity; and () COx selectivity).

294

M. Florea et al. / Applied Catalysis A: General 255 (2003) 289300

conversion and selectivities(%)

80

60

40

20

0
5

10

15

20

25

30

35

40

GHSV (l/g.h)

Fig. 3. Catalytic behavior over VAlON catalyst (V:Al = 0.25) as function of space velocity: C3 H8 :O2 :NH3 = 1.25:3:1, 500 C (( ) propane
conversion; () ACN selectivity; () AcCN selectivity; and () COx selectivity).

3.2.4. Effect of feed composition


The influence of the gas feed composition on
propane ammoxidation was investigated under reaction conditions in which the catalyst showed the
highest ACN selectivity value (T = 500 C, GHSV =
16.800 ml/g h).
The catalytic tests were performed by changing the
contact time of one reactant and keeping the others
constant. The total flow and the space velocity were
kept constant by changing correspondingly the flow
of helium used as diluent gas.
For a better understanding, Table 4ac present the
gas composition and the corresponding contact time
of oxygen, propane and ammonia. Fig. 4ac show
the conversion of propane and the selectivities to the
reaction products as function of the contact time of
oxygen, propane and ammonia used in the catalytic

reactions. As is shown in Fig. 4a, the conversion of


propane depends on the oxygen contact time. When
the oxygen is in excess of the amount required by
the stoichiometry of the reaction, the conversion of
propane increases considerably, more than 80% for a
molar ratio corresponding to C3 H8 :O2 :NH3 = 1:4:1,
but the selectivity to acrylonitrile is 19%. Under
lean-oxidation conditions, the conversion of propane
was 27% (C3 H8 :O2 :NH3 = 1:1:1). According to
the thermodynamics of this reaction, conversion of
propane is favored by increased amounts of oxygen.
But the selectivity to acrylonitrile increased continuously with decreasing oxygen content. A maximum of
45% can be reached for a contact time of 3.11 g h/mol
oxygen, after which a rapid decrease of the selectivity to acrylonitrile occurs. At lower contact time of
oxygen, the conversion of propane is highest, but the

Table 4
The gas composition and the corresponding contact time of oxygen
(a) Ratio among the reactants and corresponding contact time of oxygen
Molar ratio C3 H8 :O2 :NH3
1:1:1
1:2:1
1:2.8:1
W:F (g h/mol O2 )
9.33
4.67
3.39

1:3:1
3.11

1:3.25:1
2.87

(b) Ratio among the reactants and corresponding contact time of propane
Molar ratio C3 H8 :O2 :NH3
0.75:3:1
1:3:1
1.1:3:1
12.5
9.33
8.54
W:F (g h/mol C3 H8 )

1.25:3:1
7.69

1.5:3:1
6.25

(c) Ratio among the reactants and corresponding contact time of ammonia
Molar ratio C3 H8 :O2 :NH3
1.25:3:0.75
1.25:3:1
1.25:3:1.25
W:F (g h/mol NH3 )
12.5
10.2
7.69

1:3.5:1
2.66

1:4:1
2.33

M. Florea et al. / Applied Catalysis A: General 255 (2003) 289300

295

Fig. 4. (a) Catalytic behavior over VAlON catalyst (V:Al = 0.25) as function of oxygen contact time: GHSV = 16800 ml/g h,
C3 H8 :O2 :NH3 = 1:X:1 (( ) propane conversion; ( ) ACN yield; () ACN selectivity; () AcCN selectivity; () COx selectivity, and
() N2 selectivity). (b) Catalytic behavior over VAlON catalyst (V:Al = 0.25) as function of propane contact time: GHSV = 16800 ml/g h
C3 H8 :O2 :NH3 = X:3:1 (( ) propane conversion; ( ) ACN yield; () ACN selectivity; () AcCN selectivity; () COx selectivity; ()
N2 selectivity; and () propylene selectivity). (c) Catalytic behavior over VAlON catalyst (V:Al = 0.25) as function of ammonia contact
time: GHSV = 16800 ml/g h; C3 H8 :O2 :NH3 = 1.25:3:X (X = 0.75; 1; and 1.5), 500 C (( ) propane conversion; ( ) ACN yield; ()
ACN selectivity; () AcCN selectivity; () COx selectivity; () N2 selectivity; and () propylene selectivity).

296

M. Florea et al. / Applied Catalysis A: General 255 (2003) 289300

oxidation to COx becomes critical and the selectivity to acrylonitrile decreases to 19%. It is thus necessary to limit the total oxidation reaction and, from
this point of view, it is preferable to use a higher contact time of oxygen, since it can be a limiting reactant.
Even at higher contact time of oxygen, the formation of propylene was not observed. In order to have
higher selectivity to acrylonitrile but also to limit the
production of COx , the ammoxidation reaction on the
VAlON catalyst must be realized at moderate propane
conversion and the contact time of oxygen must be
limited at 3.11 g h/mol oxygen. This mixture was then
considered as a standard point in modification of the
propane content.
Fig. 4b shows the conversion of propane and the
selectivities to the reaction products as function of the
contact time of propane.
These variations allowed the optimization of the
gas composition of the mixtures subjected to the reaction. The contact time of propane has an important
role in the catalytic activity of the VAlON catalysts.
The minimum propane conversion but maximum
acrylonitrile selectivity is obtained for a propane contact time of 7.69 g h/mol propane corresponding to
the reactants ratio of 1.25:3:1. A further increase in
propane conversion leads to a decrease of the selectivity to acrylonitrile, probably as a consequence of the
consecutive reactions. Low contact time of propane

leads to the detection of propylene as reaction product


with a very low selectivity of 3%.
The maximum yield in acrylonitrile was obtained over the VAlON catalyst for W:F ratio of
7.69 g h/mol propane which correspond to a reactants
molar ratio of C3 H8 :O2 :NH3 = 1.25:3:1.
The last step of the optimization of the composition
was the content in ammonia, which is varied while
keeping the contact time of propane and oxygen constant. Ammonia is the reactant necessary for the acrylonitrile formation, but it is also necessary to maintain
the superficial nitrogen species of the oxynitrides catalysts. Table 4c presents the gas compositions and the
corresponding contact times used in this experiment.
The catalytic results are presented in Fig. 4c. The
conversion of propane presents a minimum as a function of the contact time of ammonia. High contact time
of ammonia led to lower selectivities in acrylonitrile
and, in these conditions, to the formation of propylene, probably due to a deficit of ammonia in the gas
feed and an excess of oxygen and propane. A maximum in selectivity to acrylonitrile was obtained for
10.2 g h/mol ammonia which correspond to a molar
ratio of C3 H8 :O2 :NH3 = 1.25:3:1. The formation of
acetonitrile and nitrogen does not depend on the ammonia contact time.
A series of transient experiments were conducted
varying the O2 :NH3 ratio in order to investigate the

O2/NH3 = 3/075
O2/NH3 = 3/1

O2/NH3 = 3/1

conversion and selectiv(%)

100
80
60
40
20
0
0

10

15

20

25

30

35

40

time (h)

Fig. 5. Catalytic behavior over VAlON catalyst (V:Al = 0.25) function of time on stream and function of O2 :NH3 ratio:
GHSV = 16800 ml/g h; C3 H8 :O2 :NH3 = 1.25:3:X (X = 0.75 and 1), 500 C (( ) propane conversion; () selectivity to ACN; ()
selectivity to AcCN; () selectivity to COx ; and () propylene selectivity).

M. Florea et al. / Applied Catalysis A: General 255 (2003) 289300

effect of the oxidation/reduction reaction environment


on the catalytic activity of the VAlON system. The
results are presented in Fig. 5. After 24 h on stream,
the O2 :NH3 ratio was varied from 3:1 to 3:0.75 for
5 h. Subsequently, the reaction conditions used were
the same as initially. Under these conditions, decreasing the O2 :NH3 molar ratio had the following effects:
(i) propane conversion increased, (ii) ACN selectivity
decreased, (iii) COx selectivity increased, (iv) AcCN
selectivity underwent a slight increase, and (v) propylene formation was detected. By changing the O2 :NH3
ratio and returning to the initial conditions, the catalyst
recovered the same catalytic activity and selectivity,
leading to the same selectivity to acrylonitrile, namely
56%. At the same time, propylene disappeared from
the reaction products.

4. Discussion

time on stream, which demonstrates the stability of


the VAlON catalyst in the propane ammoxidation
reaction (Fig. 1). In fact, it is suggested that during the
initial period of the reaction, the substitution of oxygen by nitrogen is realized and after that an equilibrium between oxygen and nitrogen surface species is
reached. These results are in good concordance with
the nitrogen content measured after different reaction
times, which increased with the time on stream in parallel with the acrylonitrile selectivity (see Table 3).
If we assume that in situ activation under catalytic
conditions can lead to the same results as nitridation
with pure ammonia, then one can expect two possible effects induced by this substitution [25]: (i) the
cationic network is not modified, and three oxygen
atoms are replaced by two nitrogen atoms; (ii) the
cationic network is modified and one oxygen is replaced by one nitrogen and the reduction of vanadium
to a lower oxidation state occurs. The substitution of
oxygen by nitrogen enables the modification of the
superficial acidbase properties [21]. Based on the hypothesis of Sokolovskii et al. [2] that the acrylonitrile
selectivity increases with the basicity, one can then
explain the evolution of the acrylonitrile selectivity as
a function of the nitrogen content of the catalyst measured at different times on stream, shown in Fig. 6.
The conversion of propane does not depend on the
nitrogen content, while the selectivity to acrylonitrile
increases with the increasing of total nitrogen content

100

100

80

80

60

60

40

40

20

20

conversion (%)

ACN selectivity (%)

The tests performed on the VAlON indicated that


the catalytic activity depends on the reaction temperature, space velocity and gas feed composition. Moreover, an activation process is observed as a function
of time on stream (Fig. 1) and in consequence an increase of the selectivity. The steady state conditions
are reached after 12 h of reaction when propane conversion is 60% and ACN selectivity is 56%, and no
deactivation of the catalyst occurred during 60 h of

297

0
0

5
nitrogen (%)

Fig. 6. Evolution of propane conversion and selectivity of ACN as function of nitrogen content measured at different time on stream (( )
propane conversion and () selectivity to CAN).

298

M. Florea et al. / Applied Catalysis A: General 255 (2003) 289300

on the catalyst. After 12 h of reaction, the total nitrogen content in the vanadiumaluminium oxynitrides
remained constant as well as the production of acrylonitrile.
Also, as function of the reaction mixtures after 24 h
of time on stream, the nitrogen content in VAlON
catalysts is different, as presented in Table 2 as well as
the catalytic activity and is depending on the NH3 :O2
ratio.
The activity results indicate that the catalytic properties of the VAlON system depends on two factors:
(i) the NH3 :O2 ratio and in consequence, and (ii) the
amount and nature of nitrogen species. Forward, we
shall attempt to explain the catalytic behavior of the
VAlON system from these two different fundamental points of view.
In our case, ammonia seems to play a double role:
(i) as a reagent necessary for the thermal nitridation
and generation of nitrogen species, and (ii) participate in N-insertion step for the acrylonitrile formation
from propane. To provide conditions to ensure these
functions, a certain amount of ammonia is necessary,
which in our case corresponded to 10.2 g h/mol ammonia (Fig. 4c).
The nature of the surface oxygen species is very
important, since it may favor an electrophilic or nucleophilic oxidation, namely a non- or selective oxidation
[27]. The oxygen content from the reaction mixture
could control (i) the optimal degree of nitridation, and
(ii) the redox balance of vanadium species. Therefore,
the nitrogen content of the used catalysts depends on
the contact time of oxygen used in the reaction mixture
(Table 2). Higher oxygen contents favor the oxidation
of ammonia and under these conditions the VNHx
species are transformed into VO electrophilic ones,
as already suggested [24]. However, a contact time of
oxygen up to 3.11 g h/mol oxygen led to a large coverage of the catalyst surface with NHx species, and
for this value, the nitrogen content on the corresponding catalyst is about 3.2%. Under these conditions, the
selectivity to acrylonitrile achieved 46.5%.
Propane can also be a limiting reagent in the
propane ammoxidation over VAlON catalysts and it
seems that the propane:oxygen and propane:ammonia
ratios have a significant influence on the catalyst selectivity. When very low contact times of propane are
employed propylene is formed. Under these conditions, the conversion of propane reached 75% and the

main reaction products were carbon oxides. The same


phenomenon was observed over the FeSbO system
[29], lower selectivity in acrylonitrile and the presence of propylene at shorter propane contact times.
This suggests that when long gas residence times
are involved, the propylene produced is converted to
either acrylonitrile or carbon oxides by the efficient
propylene activation sites on the catalyst. However, at
shorter contact times, not all the propylene produced
on the surface is transformed and it can escape into
the product stream before conversion to acrylonitrile
or carbon oxides, evidencing itself as a major reaction
intermediate. Probably a deficit of ammonia in the
gas feed and an excess of oxygen and propane can
also explain this phenomena.
However, the activity of the VAlON catalyst is
maximal when the molar ratio among C3 :O2 :NH3 reactants is 1.25:3:1. This optimal molar ratio corresponds to a space time (W:F) of 7.69 g h/mol C3 ,
3.11 g h/mol of O2 and 10.2 g h/mol NH3 .
Thus, a proper feed composition provides a correct
equilibrium between the active nitrogen and oxygen
species. Transient experiments confirmed this behavior (Fig. 5). These data clearly demonstrate that the
catalytic behavior of the surface does not only depend on the presence of a particular surface species,
but also depends on the nature of the changes in the
surface reactivity induced by the adsorption of the reactants. It is also necessary to pay a special attention
to the surface restructuring phenomena during the catalytic reaction, as a central aspect of understanding
the surface reactivity at oxynitride surfaces. On the
vanadiumaluminium oxynitrides, this phenomenon is
possible as the nitridation process is reversible. This
behavior was observed on AlPON catalysts [30] and
should be confirmed in the case of VAlON.
The positive effect of pre-nitridation clearly indicates the role of these nitrogen species in ammoxidation, which results from the evolution of the propane
conversion and selectivities level of the reaction products as a function of nitrogen content measured after
different reaction conditions (Fig. 7). The selectivity
results depicted in Fig. 7 are in agreement with those
presented in Fig. 6, meaning that the selectivity to
ACN depends on the nitrogen content. Concerning
the propane conversion, two tendencies were observed (i) when the nitrogen content is measured at
different times on stream, but the same molar ratio

299

100

100

80

80

60

60

40

40

20

20

0
1

conversion (%)

selectivities (%)

M. Florea et al. / Applied Catalysis A: General 255 (2003) 289300

0
5
nitrogen content (%)

Fig. 7. Evolution of propane conversion and selectivity of ACN as function of nitrogen content measured after different C3 H8 :O2 :NH3
molar ratios (( ) propane conversion and () selectivity to ACN).

of C3 :O2 :NH3 = 1.25:3:1, the propane conversion


does not depend on the nitrogen content (Fig. 6) and
(ii) when nitrogen content is measured after different
reaction conditions, with different C3 :O2 :NH3 molar
ratios, the conversion varies with the nitrogen content
(Fig. 7). This suggests that the propane conversion depends more on the molar ratio between the reactants
than on the nitrogen content, which is responsible for
the evolution of the acrylonitrile selectivity.
The positive effect of the nitrogen species from the
VAlON catalysts also results from the very large difference in the contact time value of propane (W:F),
corresponding to high selectivities. For the VAlON
catalyst [22], a selectivity as high as 55% is reached
for a contact time of 8 g h/mol of C3 H8 , which is
much higher than that corresponding to the conventional ammoxidation systems operating at values often
above 100 g h/mol of C3 H8 . Andersson and coworkers [15], by using VSbWAl oxides, reported a
selectivity of 48% for 77% propane conversion at
53 g h/mol of propane. We estimated the ACN productivity (ACN l/kg of catalyst per hour) from the W:F
value and the ACN yield. For VSbWAlOx catalysts,
the ACN productivity is 164 l/kg h, while for VAlON
the obtained value is 812 l/kg h.
The fact that propylene was not observed as a reaction product in the optimal conditions, might suggest
that the mechanism of acrylonitrile formation on the
nitridated vanadiumaluminium oxides is different as

compared with the mechanism reported for the promoted catalysts.


Until now, little information has been presented in
the literature concerning the formation of ACN in one
stage from propane ammoxidation. In 1979, Osipova
and Sokolovskii [26] did not exclude the possibility of
ACN formation from propane without propylene desorption from the surface. Propylene can be formed as
intermediate, which, without desorption from surface,
quickly transforms into allyl and subsequent nitrile
formation. Centi et al [6] suggested that the formation
of acrylonitrile results from two parallel pathways, one
directly from propane and the second through the intermediate formation of propylene. The latter is the
prevalent path.

5. Conclusions
These results underline the influence of the contact time of each reactant on the catalytic behavior of
vanadiumaluminium oxynitrides. Therefore, the positive effect of pre-nitridation clearly indicates the role
of nitrogen-adsorbed species in ammoxidation reaction. Optimal performance is achieved when the reaction temperature is 500 C and the molar ratio of
propane, oxygen, ammonia and inert in the gas feed is
1.25:3:1. Under these optimal reaction conditions, the
catalyst showed a propane conversion level of 60%,

300

M. Florea et al. / Applied Catalysis A: General 255 (2003) 289300

an acrylonitrile selectivity of 56% and an acrylonitrile


productivity of 812 l/kg h.
Taking into account the fact that the difference
between the conventional oxide catalysts used in ammoxidation of propane and the oxynitrides catalysts
studied in this paper is the presence of different nitrogen species (N3 , NHx ), we can admit the positive
influence of these species on the reaction mechanism.
This represents an important parameter which could
determine the level of the propane conversion and the
selectivity to the reaction products. However, propylene was not observed as a reaction product in the
optimal reaction conditions. It is a key aspect which
underlines that the mechanism of the acrylonitrile
formation on the vanadiumaluminum oxynitrides
should be different from conventional catalyst in
propane ammoxidation.
Acknowledgements
The authors would like to thank the Rgion
Wallonne, the Communaut Franaise de Belgique,
as well as the FNRS (Belgium) for financial support.
References
[1] J.F. Brazdil, Catal. Sci. Technol. 1 (1991) 3.
[2] V.D. Sokolovskii, A.A. Davydov, O.YU. Ovsitser, Catal. Rev.
Sci. Eng. 37 (3) (1995) 425.
[3] G. Centi, S. Perathoner, F. Trifiro, Appl. Catal. A 157 (1997)
143.
[4] R. Nilsson, T. Lindblad, A. Andersson, J. Catal. 148 (1994)
501.
[5] O. Ratajczak, H.W. Zanthoff, S. Geisler, Stud. Surf. Sci.
Catal. 130 (2000) 1685.
[6] G. Centi, R.K. Grasselli, E. Patane, F. Trifiro, Stud. Surf. Sci.
Catal. 55 (1990) 515.

[7] Y.C. Kim, W. Ueda, Y. Moro-oka, Stud. Surf. Sci. Catal. 55


(1990) 491.
[8] J.S. Kim, S.I. Woo, Appl. Catal. A 110 (1994) 207.
[9] US Patents 4,760,154 (1988); 4,835,125 (1989); 4,837,191
(1989); 4,843,055 (1989).
[10] G. Centi, D. Pesheva, F. Trifiro, Appl. Catal. A 33 (1987)
343.
[11] G. Centi, S. Perathoner, J. Catal. 142 (1993) 84.
[12] G. Centi, T. Tosarelli, F. Trifiro, J. Catal. 142 (1993) 70.
[13] A.A. Davydov, A.A. Budneva, V.D. Sokolovskii, Kinet. Katal.
(1979) 179.
[14] G.A. Zenkovets, G.N. Kryukova, S.V. Tsybulya, V.F.
Anufrienko, T.V. Larina, E.B. Burgina, Kinet. Catal. 43 (2002)
384.
[15] J. Nilsson, A.R. Landa-Cnovas, S. Hansen, A. Andersson,
J. Catal. 186 (1999) 442.
[16] K. Asakura, K. Naakatani, T. Kubota, Y. Iwasawa, J. Catal.
194 (2000) 309.
[17] T. Ushikubo, K. Oshima, A. Kayou, M. Vaarkamp, J. Catal.
169 (1997) 394.
[18] G. Centi, R.K. Grasselli, F. Trifiro, Catal. Today 13 (1992)
661.
[19] R. Catani, G. Centi, F. Trifiro, Ind. Eng. Chem. Res. 31
(1992) 107.
[20] US Patents: 4,746,641 (1988); 4,784,979 (1988); 4,788,317
(1988); 4,871,706 (1989); 4,879,264 (1989).
[21] H. Wiame, C. Cellier, P. Grange, J. Catal. 190 (2000) 406.
[22] M. Florea, R. Prada Silvy, P. Grange, Catal. Lett. 87 (2003)
53.
[23] J. Guyader, F.F. Grekov, R. Marchand, J. Lany, Rev. Chim.
Miner. 15 (1978) 431.
[24] H. Wiame, L. Bois, P. LHaridon, P. Grange, J. Eur. Ceram.
Soc. 17 (1997) 2017.
[25] R. Marchand, Y. Laurent, J. Guyader, P. LHaridon, P. Verdier,
J. Eur. Ceram. Soc. 8 (1991) 197.
[26] Z.G. Osipova, V.D. Sokolovskii, Kinet. Katal. 20 (1979)
420.
[27] E. van Steen, G. Kuwert, A. Naido, M. Williams, Stud. Surf.
Sci. Catal. 110 (1997) 423.
[28] M. Vaarkamp, T. Ushikubo, Appl. Catal. A 174 (1998) 99.
[29] M. Bowker, C.R. Bicknell, P. Kerwin, Appl. Catal. A 136
(1996) 205.
[30] M.A. Centeno, M. Debois, P. Grange, J. Phys. Chem. 102 (35)
(1998) 6835.

You might also like