Вы находитесь на странице: 1из 14

Applied Catalysis A: General, 84 (1992) 155-168 155

Elsevier Science Publishers B.V., Amsterdam

APCAT 2259

Deamination of set-butylamine over acidic zeolites

M. Lequitte, F. Figueras and C. Moreau


Laboratoire de Chimie Organique Physique et Cine’tique Chimique Applique& CNRS URA 418,
Ecole Nationale Supe’rieure de Chimie, 8, Rue de l%cole Normale, 34053 Montpellier, (France)

and
S. Hub
Centre de Recherches Rh&ze Alpes, ATOCHEM, 2, rue H. Moissan, BP 20,693lO Pierre Benite
(France)
(Received 9 December 1991, revised manuscript received 20 February 1992)

Abstract

The conversion of set-butylamine (BA) has been investigated over MFI, MOR, FAU, BEA and MAZ
zeolites. Faujasites were used in the H and Ce forms. The products of the reaction are l- and 2-butenes,
and di-set-butylamines (di-set-BA). Butene formation follows a zero order relative to the pressure of
butylamine, while dibutylamine formation follows a first order. The catalytic activity for deamination
to butenes is related to the total number of acid sites. An effect of the zeolite structure is observed: at
similar aluminium content, MFI and MOR appear more active than FAU and BEA for butene forma-
tion. The formation of dibutylamines is observed on large pore zeolites only, and results from two dif-
ferent pathways, di-set-butylamine appearing as a secondary product in the presence of non-dealumi-
nated H-form of Y zeolites, and as a primary product in the presence of Ce3+ exchanged and dealuminated
Y zeolites. This change is interpreted in terms of a classical consecutive reaction scheme in the former
case, through addition of set-butylamine to butene intermediates catalyzed by Brensted acids, and in
terms of a parallel reaction scheme in the latter case, through condensation of two set-butylamine mol-
ecules catalyzed by Lewis acids.

Keywords: acid sites, butylamine deamination, cerium, large-pore zeolites.

INTRODUCTION

Direct reaction of ammonia and alkenes is an attractive process for the man-
ufacture of amines. This reaction has been recently investigated in the pres-
ence of RE-Y zeolites, and alumina-, ferro- or boro-zeolites of the pentasil
group [l-6]. Deeba and co-workers [ 7-141 reported good yields using offretite

Correspondence to: Prof. F. Figueras, Laboratoire de Chimie Organique Physique et Cinetique


Appliqukes, CNRS URA 418, Ecole Nationale Supkrieure de Chimie, 8 Rue de 1’Ecole Normale,
34053 Montpellier Cedex 1, France. Tel. ( +33)67144300, fax. (+33)67144353.

9926~86W92/$05.OO 0 1992 Elsevier Science Publishers B.V. All rights reserved.


156 M. Lequitte et al./Appl. Catal. A 84 (1992) 155-168

and dealuminated mordenites. They pointed out the problem of catalyst life
due to cokefaction of wide pore zeolites [ 121 and the limitations introduced by
unfavorable thermodynamics of the equilibrium. Amination is then difficult to
study because of these two factors, and presumably also because of possible
diffusional limitations at the high temperatures required for alkene activation.
For academic purposes it appears interesting to investigate the reverse re-
action of deamination, which has no thermodynamic limitations, in the tem-
perature range used for amination, and which is slow enough to permit a ki-
netic study in the absence of diffusional limitations.
The adsorption of butylamines has been investigated by many authors [ 15
211, using temperature-programmed desorption or infrared spectroscopy. A
strong adsorption of the amine occurs at temperatures < 473 K, with an initial
heat of adsorption of about 20 kcal/mol (1 cal= 4.1868 J) [ 161, and butylam-
ine forms a l/l adsorbed complex with each aluminium atom in H-ZSM-5
[ 211. Above 573 K the adsorption of butylamines is reversible, and the sub-
strate is partly decomposed into butenes and ammonia, part of the butenes
being transformed into hydrocarbons of higher molecular weight [ 19,211. The
rate of decomposition of butylamines has been reported to be limited by the
rate of desorption of ammonia [ 181.
Around 623 K, the conversion can be kept low, and the alkene should be
displaced by the amine, thus ensuring a good stability of the catalyst with time-
on-stream. Moreover, the dialkylamine is also formed in this reaction [ 171 and
the mechanism of this secondary reaction, which represents the main loss of
selectivity in amination can then be investigated more safely. The reaction of
alkylation of isobutene with arylamines has been reported to be useful as a
synthetic tool at high pressures [ 221.
The present work is devoted to the study of deamination of set-butylamine
on a series of zeolites in order to check the respective influence of the structure
and chemical composition of the catalyst.

EXPERIMENTAL

Catalysts preparation

Zeolite /I? (BEA) was synthesized using tetraethylammonium hydroxide


(TEAOH) as organic template following the procedure described by Nicolle
[ 231. The crystals appeared as spheroids with an average size of 0.6 pm. Hex-
agonal crystals of offretite (OFF) and omega (MAZ) were synthesized accord’
ing to patented procedures [ 24,251. The average size of the crystals was 0.7 x 1.5
,um for OFF and 1 x 3.5 pm for MAZ. These samples were calcined at 773 K
(zeolite p> or 873 K (OFF and MAZ) for 8 h under dry air (120 ml min-I),
then converted to the ammonium form by ion exchange in a 1 M NH,NO,
solution at 373 K.
M. Lequitte et al./Appl. Catal. A 84 (1992) 155-168 157

MFI samples of different compositions were obtained from Conteka. Mor-


denites (MOR) with silicon to aluminium ratios of 6.9, 11 and 49 were ob-
tained from Norton and Zeocat. The ammonium forms of all samples were
activated according to the procedure described for offretites. Mordenites con-
sisted of polycrystalline aggregates, with an average particle size of about 2 pm.
The Y zeolites were Linde LZ-Y62 (Si/Al= 2.5), and three dealuminated
samples ZF510,ZF515 and ZF520 obtained from Zeocat.
An amorphous silica-alumina with 13% Al2O3 was obtained from Ketjen,
and an acid leached montmorillonite clay from Siid Chemie.

Catalyst churacterizations

The chemical composition of the catalyst was determined by atomic absorp-


tion analysis after dissolution of the sample (SCA-CNRS, Solaize, France).
The crystallinity of zeolites was checked by X-ray powder diffraction on a CGR
Theta 60 instrument using Cu Kcwl filtered radiation. Diffraction patterns
showed a good crystallinity of the samples.
The size and morphology of the crystals were determined by electron scan-
ning microscopy. The micropore volume and BET surface area were deduced
from the isotherms of adsorption of nitrogen at 77 K.
Acid properties were determined by stepwise thermal desorption of ammo-

TABLE 1

Chemical compositions and textural properties of the catalysts used here

Sample Suppliers Si/Al Vol. micro Vol. meso Acidity > 523 K
reference (ml g-‘) (ml g-‘) (me&)

HY2.5 Linde LZ62 2.5 0.36 0.04 1.5


HYlO ZF510 10
HY15 ZF515 15 0.46
HY20 ZF520 20 0.36
MOR6.9 Zeolon 1OOH 6.9 0.22 0.014 0.88
MORll ZM510 11 0.21 0.04 0.82
MOR49 ZM760 49 0.22 0.07 0.29

ZSM5( 14) CBV3010 14 0.45


ZSM5(25) CBV5010 25
ZSM5(70) CBV1501 70

BEA ZIP1 15 0.85


BEA ZIP4 23 0.48
BEA ZIP6 26 0.30 0.04 0.32
OFF 3.5 0.51
MAZ 3.2 0.08 0.02 1.04
158 M. Lequitte et al./Appl. Catal. A 84 (1992) 155-168

nia. The catalyst was first activated in flowing nitrogen in situ, saturated with
ammonia at 373 K, then swept by a flow of dry nitrogen while the temperature
was raised by steps of 50 K. The amount of ammonia evolved by the solid was
monitored by conductometry. The chemical compositions, the textural and
acid properties of these catalysts are reported in Table 1.

Catalytic tests

Tests were performed in a conventional flow reactor operated at atmospheric


pressure. 100 mg of catalyst were first reactivated in situ in flowing nitrogen
at 773 K overnight, then the temperature was decreased to the desired value.
In a typical experiment, the carrier gas for the reaction, nitrogen, was saturated
with the vapour pressure of see-butylamine (Aldrich, purity > 99% ), at 273 K
(partial pressure of substrate 6.65 kPa), and the conversion was determined
at several temperatures between 553 and 673 K. The composition of the ef-
fluent was determined by an on-line gas chromatograph equipped with a cap-
illary column (30 m x 0.75 mm I.D. ), bonded with an apolar phase. All con-
necting lines were heated to avoid condensation and adsorption of the reactants
and products. The organics produced by the reaction were identified by com-
parison with authentic samples, and by GC-MS analysis.

RESULTS

Preliminary results

Preliminary experiments were carried out to find the conditions in which


diffusional limitations are minimized. Using a weight of HY = 100 mg, the rate
and selectivity are independent of the flow of reactant above 1 1 h-l, then
external limitations can be discarded. Typical values of the rate of butene for-
mation are 10m6mol s-l g-’ over most zeolites, using a concentration of sub-
strate of 10m6mol ml-‘, with average crystal sizes of about 1 pm. The criterion
of Weisz and Swegler [ 261: dN/dt* 1/CO*R2D,.f< 0.1, predicts that the kinetics
is controlled by the chemical process if the diffusion coefficient Deff is higher
than 10e7, which is probably the case for large pore zeolites. For dialkylamine
formation the rates are usually higher, and diffusional limitations of the rate
are expected for the more active zeolites.

Kinetic study of deamination

The reaction yields ammonia and l- and 2-butenes, with different amounts
of di-set-butylamine which vary with the type of catalyst used. On HY 2.5 the
selectivity to di-set-butylamine is about 6%, but reaches 87% on HY 15. There-
M. Lequitte et al. jApp1. Catal. A 84 (1992) 155-168 159

fore, two different kinetic studies were carried out on these catalysts to clarify
the mechanisms of the two reactions.
The rates of reaction as a function of the partial pressure of set-butylamine
are reported in Figs. 1 and 2 for HY 2.5 and HY 15, respectively. The rate of
the monomolecular reaction of deamination to butenes is zero order above 6.65
kPa at 613 K, as expected for the conversion of a strongly adsorbed substrate.
The kinetics follow a Langmuir Hinshelwood law, with a rate constant
I~=7*10-~ mol h-l g-l, and an adsorption coefficient 2=4.25 Pa-l. The ac-
tivation energy determined in the zero order range is 40 kcal/mol, and the
activation energy at low partial pressure, then in the first order range is 20
kcal/mol. The average heat of adsorption of set-butylamine on HY 2.5 can
then be calculated to be 20 kcal/mol, which appears reasonably consistent with
the calorimetric determination of the heat of adsorption of ammonia by Au-
roux [ 271 for zeolites (25-30 kcal/mol).
The rate of the reaction of see-butylamine to di-set-butylamine over HY 15

I
0 5 10 15 20 25
BA partial pressure (kPa)

Fig. 1. Influence of the partial pressure of set-butylamine on the rate of reaction at 613 K over HY
(LZ62).

s
._ 40
.
E
+?a .

z mm
0 0 L- I
0 10 20 30 40 50
BA partial pressure (Pa)

Fig. 2. Influence of the partial pressure of set-butylamine on the rate of reaction at 613 K over HY
dealuminated to a Si/Al ratio of 15.

Fig. 3. Variations of the selectivity to dibutylamines as a function of conversion, at 613 K, over


HY (LZ62) and MOR (Zeolon lOOH).
160 M. Lequitte et al./Appl. Catal. A 84 (1992) 155-168

is a first order fast reaction, with a low activation energy, about 11 kcal/mol.
Di-set-butylamine can be formed either by a parallel mechanism of reaction
of two amines leading to elimination of ammonia, or consecutive addition of
see-butylamine to the butenes produced by the reaction.
The selectivity of the reaction was determined at different conversions on
HY 2.5, H-MOR and CeY at the same temperature of 613 K. The results re-
ported in Figs. 3 and 4 show that on the two protonic zeolites di-see-butylamine
(di-set-BA) is a secondary product of the reaction.
However, on CeY, di-see-butylamine appears as a primary product. The se-
lectivity to di-set-butylamine obtained, at low conversion, on dealuminated Y
zeolites suggests that this is a primary product in that case too.

Fig. 4. Variations of the selectivity to dibutylamines as a function of conversion, at 613 K, over


HCeY (72% exchange).

TABLE 2

Conversion of n-butylamine (n-BA) over HY 15 at 613 K, pressure of amine= 6650 Pa

Time-on-stream Conversion Selectivities to


(min) (%)
butenes (% ) di-n,sec-BA (%) di-n-BA (%)

5 16.4 2.2 6.4 0.6


60 4 0.65 0.8 0.8

TABLE 3

Comparison of the reactivities of butylamines over zeolites at 613 K

Sample Conversion of Selectivity to butenes in the conversion of

n-butyl (%) set-butyl (%) tert-butyl (%) n-butyl (%) see-butyl (%) tert-butyl (%)

HY2.5 22.4 52.8 94 98


HY15 16.4 28 99.7 13 14 98
HCeY 35 44 11 98
M. Lequitte et al./Appl. Catal. A 84 (1992) 155-168 161

Some additional experiments on the conversion of n-butylamine, and tert-


butylamine were then performed using HY 15 in order to define exactly the
reaction scheme. The results are reported in Tables 2 and 3. The reactivity of
amines on a HY 15 catalyst parallels that expected for alcohols, since the con-
version increases in the sequence n- < set- < tert-butylamines. As for alcohols,
n and set-butylamines can give dibutylamines, which are analogues of ethers.
The distribution of the products formed by conversion of n-butylamine shows
that dibutylamines can be formed by two differents mechanisms: the first giv-
ing the formation of di-n,sec-butylamine, in which the alkyl chain is ramified,
and the other leading to di-n-butylamine, in which the ramification of the alkyl
chain is not changed. The catalyst deactivates with time-on-stream. On the
deactivated catalyst, equal selectivities are obtained for the two types of reac-
tion, and the bimolecular process of elimination of ammonia is then favored.

Influence of the zeolite structure on activity

All samples were compared in conditions in which the reaction is zero order,
at low conversion. The respective influences of structure and composition were

TABLE 4

Catalytic properties of a series of zeolites of different composition and structure for the conversion
of set-butylamine at 613 K

Zeolite %/Al Conversion Selectivity Rates of reaction (mmol/h g) E,(BA)


BA C4 (kcal/mol)
(%) (%) r(G) r(DBA) r(BA)

HY2.5 2.5 22.4 94 6.8 0.6 7.1 44


HYlO 10 24.5 25 2.7 2.7 8 11
HY15 15 28.1 14 1.6 3.3 8.2 9
HY20 20 17.4 17 1.2 1.7 5.1 12
MOR6.9 6.9 25.5 97 8.8 0.04 8.9 39
MORll 11 13.7 93 5.3 0.05 5.3 46
MOR49 49 0.3 100 0.07 0 0.07
ZSM514 14 14.6 96 4.9 0.05 4.9 45
ZSM525 25 11.9 99 3.8 0.07 3.9 48
ZSM570 70 0.3 79 0.08 0.06 0.09
BEA 15 14.5 41 2.6 1.2 4.9 13
BEA 23 14.8 38 2.4 1.2 4.8 22
BEA 26 12.5 39 2 1.0 3.9 21
OFF 3.5 5.3 94 1.5 0.1 1.7 45
MAZ 3.2 3.5 73 1.4 0 1.4 52
Ketjen 2.5 4.6 59 1.1 0.07 1.3 35
KlO 6.2 67 1.5 0.2 2 34
162 M. kquitte et al./Appl. Catal. A 84 (1992) 155-168

0 095 1 1,5 2
Acidity (meq.g-1)

Fig. 5. Rate of formation of butenes at 613 K, partial pressure of see-butylamine=6650 Pa, as a


function of the number of milliequivalents of ammonia released above 523 K. (0) Faujasites, ( n )
BEA, (0 ) ZSM-5, (+) MOR, (0 ) OFF.

separated by comparing zeolites of different Si/Al ratios, obtained either by


synthesis, as for MFI and BEA or by dealumination as for Y and MOR zeolites.
The activities determined at 613 K are reported in Table 4 as a function of the
Si/Al ratio for several zeolites. For mordenites and faujasites, the activity de-
creases when the zeolite is dealuminated. Mazzite and offretite show a low
activity compared to their aluminium content, which suggests that diffusional
limitations are important in this case.
As reported in Fig. 5, activity is well correlated with the number of acid sites
which retain ammonia above 523 K, but two different lines can be drawn, one
related to pentasils and the other to FAU and BEA zeolites.

Influence of the zeolite structure on selectivity

The results obtained at about l&30% conversion are reported in Table 4,


and show that the selectivity to di-see-butylamine changes with the structure
of the zeolite and its chemical composition. The selectivity to butenes is high
on HY 2.5, mordenites and ZSM-5, but decreases for amorphous silica-alu-
mina, dealuminated faujasites or zeolite /3. It is interesting to note however
that while HY is selective for butenes, CeY shows a high selectivity for di-sec-
butylamine, as illustrated in Table 5.
M. Lequitte et al./Appl. Catal. A 84 (1992) 155-168 163

TABLE 5

Catalytic properties of HY and CeY faujasites

Sample Exchange Conversion Selectivity Reaction rates (mol s-’ g-l) E,(BA)
(W) BA C4 (kcal/mol)
(%) (mol-%) r(G) r(DBA) r(BA)

HY2.5 100 22 94 6.6 0.16 7.1 44


CeY 72 35 11 1.8 4.8 11.5 7

Catalyst deactivation

Some deactivation is noticed as a function of time. The amounts of carbon


and nitrogen retained by the solid after reaction were determined by chemical
analysis, and are reported in Figs. 6 and 7, respectively, as a function of the
catalytic activity and number of aluminium atoms in the solid. The amount of
nitrogen is correlated to the number of acid sites, except for MAZ which prob-
ably reflects a lower accessibility to the internal surface. The amount of carbon
is proportional to the reaction rate. This suggests that deactivation is due to
one of the products of the reaction.
It can be remarked that for zeolite p for instance, the C/N atomic ratio of
the residue, determined by chemical analysis is 5.6, to compare with 4 for sec-
butylamine remaining adsorbed, and 8 for di-see-butylamine. It appears then
that deactivation could be attributed, not to the formation of coke or oligomers,
but to the formation of di- or trialkylamines, more basic than the substrate,
which remain adsorbed.

0
0 I 2 3 4 5

Rsclctlon rate (mmol.g-1. am-l) Al content (mmol.g-1)

Fig. 6. Amount of carbon retained by the catalyst after reaction, as a function of its steady-state
activity.

Fig. 7. Amount of nitrogen retained by the catalyst after reaction as a function of ita aluminium
content.
164 M. Lquitte et al./Appl. Catal. A 84 (1992) 155-168

DISCUSSION

The formation of l- and 2-butenes can easily be accounted for by the well
known Hofmann degradation of amines represented in Scheme 1, as already
assumed in the literature for this kind of reaction performed over zeolites [ 201.
In the liquid phase this reaction is catalyzed by protonic acids, and the good
relationship observed between the rate of formation of butenes and the number
of acid sites is consistent with such an acid mechanism. As reported in Table
6, the butenes produced on HY, MFI and BEA are at thermodynamic equilib-
rium. By contrast MOR shows a high selectivity to cis-butene-2. These differ-
ences of selectivity are accounted for by consecutive reactions which are ap-
parently unfavored in the one directional channels or mordenite, as reported
earlier by Fajula and Dirringer from their results on butene-deuterium ex-
change [ 291.
The activity per protonic site follows the pattern: HMFI, HMOR > HBEA,
HFAU. It can be pointed out that diffusional limitations would lead to an
underestimation of the activity of small or medium pore zeolites, such as MFI
or MOR then the particular position of these zeolites merits a discussion.
For large Si/Al ratios, about 15, a critical examination of previous experi-

CHs-CC-CH -CH3 - P-BUTENES


P I
H NH2
c
3 I
Zeol.0 . H*

CH3-CH2--CH -C-H - I-BUTENE


I <I
NH2 H
c
f
H+ _O.Zeol

Scheme 1. Formation of butenes through Hofmann /T-elimination reaction mechanism.

TABLE 6

Distribution of the butenes (in molar % ) produced in the reaction at 613 K

Sample FAU ZSM5 BEA MAZ OFF MOR

But-l 25 30 34 4 37 8
&-But-2 38 39 29 70 23 69
trans-But-2 37 31 37 26 40 23
M. Lequitte et al./Appl. Catal. A 84 (1992) 155-168 165

ments, using high-resolution solid state NMR concluded that aluminium at-
oms are isolated in the lattice of faujasite [ 301, which has then the maximum
acidity permitted in this zeolite. The observation of higher activity for butene
formation on ZSM-5 (14) compared to HY15 could then be attributed to the
influence of the structure on the acid strength.
According to Rabo and Gajda [ 311, the rigidity of the lattice can affect the
acid strength since it is well known that acidity is controlled by the Si-O-Al
bond angle [ 321. Recent quantum chemical calculations demonstrate that de-
protonation of the zeolite lattice leads to large local changes in geometry that
change acidity [ 331. These calculations account for the modifications of the
infrared spectrum of NH,H and HY zeolites related to the relaxation of the
lattice when the proton is lost and then fur the higher acidity, at the same
aluminium content, of pentasil zeolites which possess a rigid lattice.
Some effect on the catalytic properties is then expected for those reactions
catalyzed by protons, and this effect should be larger for those reactions in-
volving strong bases which can completely displace the proton from the lattice,
and weaker for hydrocarbons, for the absorption of which Kazansky and Sen-
chenya [34] proposed the formation of a covalent bond. The present results
support this hypothesis, but also demonstrate that this influence of the struc-
ture is small and represents at most a factor of 3 on the activity at 613 K for
the deamination of set-butylamine.
The formation of di-set-butylamine is more complex, and can occur by two
different pathways:
(i) A consecutive alkylation of l- and/or 2-butenes by the non-reacted amine,
in agreement with the formation of di-set-butylamine as a secondary product
(Scheme 2). In addition, it should be noted that this consecutive reaction
scheme must lead to an isomerisation of the alkyl chain. This isomerisation is
indeed clearly observed in the conversion of n-butylamine with the high initial
selectivity to di-n,sec-butylamine.

CH3--HP-CH -CH,
I

CH3 -CH = CH -CH3 CH3 -CHy--CH =CH2

\CH3-CH2-i; -CH, )

I
CH3-CH2-CH -CH3

Scheme 2. Formation of dibutylamine through addition of set-butylamine to l- or 2-butene.


166 M. Lequitte et al./Appl. Catal. A 84 (1992) 155-168

CH3-CH H

-0 Al -

Scheme 3. Formation of di-set-butylamine by self condensation of set-butylamine over Lewis acid


sites.

(ii) A parallel reaction of bimolecular condensation of see-butylamine with


elimination of ammonia, not detected by flame-ionization detection (FID).
This is the main path on CeY and dealuminated Y, on which di-set-butylamine
appears as a primary product.
The kinetics of the reaction does not permit to choose between these two
hypotheses. It was reported above that deamination to butenes was a slow zero
order reaction. The observation of a first order relative to set-butylamine agrees
with a reaction of gaseous butylamine with either a butylammonium ion or an
adsorbed butene, the concentration of which would be constant at the surface.
From the expected parallelism between the reactions of amines and alcohols,
in which dialkylamine would be the analogue of ether, a mechanism similar to
that already proposed in the literature for conversion of alcohols into ethers,
can be considered [ 351. Indeed, it is well known that alumina, which is a typical
Lewis acid, catalyzes the formation of ethers from alcohols [ 361. Alumina was
also reported by Hogan and Pasek [37] to catalyze diethylamine dispropor-
tionation to ethyl- and triethylamine. In the present case, the exchange of
protons by Ce3+, or steam dealumination which also corresponds to an in-
crease of Lewis acidity [38,39] promote the formation of dibutylamines. It is
also well known that BEA contains after activation an appreciable amount of
extra framework aluminium, and Lewis acid centres [ 40,411. Therefore the
selectivity to dialkylamine can be rationalized assuming a bimolecular reaction
of the type shown in Scheme 3.This reaction is fast, and can therefore be lim-
ited by diffusion on small pore zeolites: in that case, the first order bimolecular
reaction would be much more sensitive to this effect than the zero-order mon-
omolecular reaction leading to butenes. Thus small-pore zeolites appear to be
more selective for alkene formation, even when they contain extra-framework
aluminium such as steam-dealuminated mordenites.

CONCLUSION

Deamination of set-butylamine is, in a first approximation, controlled by


the acidity of the zeolite, with a small effect of the structure, at equal alumin-
ium content. The influence of the conditions of activation of the zeolite is large
M. Lequitte et al./Appl. Catal. A 84 (1992) 155-168 167

on the selectivity because the formation of dibutylamine is favored by Lewis


acidity. Therefore the selectivity of faujasites can be controlled by
dealumination.
As a result of the large differences in the activation energies for the forma-
tion of butenes and dibutylamines from butylamines, the reaction of deami-
nation appears to be a sensitive probe for the characterization of Lewis acidity
of solid catalysts, even in the presence of Brnrnstedspecies.

ACKNOWLEDGMENTS

We thank Mr. P. Espiau for the synthesis and characterization of the sam-
ples of zeolites BEA, the Service Central d’Analyses du CNRS (Solaize) for
the chemical analysis, and Dr. A. Germain for fruitful discussions.

REFERENCES

J. Peterson and H.S. FaIes, U.S. Patent 4 307 250 (April 8,198l).
J. Peterson and H.S. FaIes, U.S. Patent 4 375 002 (Feb. 22,1983).
J. Peterson andH.S. FaIes, European Patent 39 918 (Feb. 29,1984), assigned to Air Products.
M. Deeba, W.J. Ambs, Eur. Pat. 77016 (June 19,1985), assigned to Air Products.
V. Taglieber, W. Holderich, R. Kummer, W.D. Mross and G. Saladin, DBP 3 326 576 (Jan.
31,1985).
6 V. Taglieber, W. Holderich, R. Kummer, W.D. Mross and G. Saladin, DBP 3 327 000 (Feb.
7,1985), assigned to BASF AG.
7 M. Deeba, M.E. Ford and T.A. Johnson, in J.W. Ward (Editor), Catalysis 1987, Amsterdam,
1988, p. 221.
8 M. Deeba, M.E. Ford and T.A. Johnson, in D.W. Blackburn (Editor), Catalysis of Organic
Reactions, M. Dekker, New York, 1990, p. 241.
9 M. Deeba, Eur. Pat. 305 564 (March 8,1987) assigned to Air Products.
10 W.J. Ambs, M. Deeba and J.F. White, U.S. Pat. 403 889 (July 30,1982).
11 W.J. Ambs, M. Deeba and J.F. White, Eur. Pat. 101 921 (April 16, 1986), assigned to Air
Products.
12 M. Deeba and M.E. Ford, Zeolites, 10 (1990) 794.
13 M. Deeba, M.E. Ford and T.A. Johnson, J. Chem. Sot., Chem. Commun., (1987) 562.
14 M. Deeba and M.E. Ford, J. Org. Chem., 53 (1988) 4594.
15 P.A. Jacobs, B.K.G. Theng and J.B. Uytterhoeve, J. Catal., 26 (1972) 191.
16 J. Koubek, J. Volf and J. Pasek, J. Catal., 38 (1975) 385.
17 M. Takahashi, Y. Iwasawa and S. Ogawasara, J. Catal., 45 (1976) 15.
18 L.V. MaIysheva, E.A. Paukshitis and N.S. Kotsarenko, React. Kinet. Catal. Lett., 24 (1984)
97.
19 J.H. Zhu and Q.H. Xu, Chin. Sci. Bull., 34 (1989) 384.
20 C.G. Pope, Zeolites, 10 (1990) 28.
21 D.J. Parillo, A.T. Adamo, G.T. Kokotailo and R.J. Gorte, Appl. Catal., 67 (1990) 107.
22 D.D. Dixon and W.F. Burgoyne, Appl. Catal., 62 (1990) 161.
23 M.A. Nicolle, Ph.D. Thesis, University of Montpellier, France, 1991.
24 F. Fajula, F. Figueras and L. Moudafi, Eur. Pat. 118 382 (1987), assigned to CNRS.
168 M. Lequitte et al./Appl. Catal. A 84 (1992) 155-168

25 F. Fajula, F. F&eras, L. Moudafi, M. Vera Pacheco, S. Nicolas, P. Dufresne and C. Guegen,


French Pat. 8 507 772 (1985), assigned to ELF and IFP.
26 P.B. Weisz and E.W. Swegler, J. Phys. Chem., 59 (1955) 823.
27 A. Auroux in B. Imelik and J.C. Vedrine (Editors), Les Techniques Physiques d’Etude des
Catalyseures, Technip, Paris 1988, p. 823.
28 N. Petronovic and U.B. Mioc, J. Serb. Chem. Sot., 54 (1989) 211.
29 F. Fajula and M.C. Dirringer, Zeolites, 3 (1983) 163.
30 J. Dwyer, in P.J. Grobet, W.J. Mortier, E.F. Vansant and G. Schutz-Ekloff (Editors), In-
novation in Zeolite Materials Science (Studies in Surface Science and Catalysis, Vol. 37),
Elsevier, Amsterdam, 1988, p. 333.
31 J.A. Rabo and G.J. Gajda, in D. Barthomeuf, E. Derouane and W. Holderich (Editors),
Physicochemical Properties of Zeolitic Systems, NATO AS1 series, Plenum, New York, Vol.
221,1990, p. 273.
32 A.G. Pelmenshchikov, E.A. Paukshitis, V.G. Stepanov, V.I. Pavlov, E.N. Yerchenko, K.G.
Ione, G.M. Zhidomirov and S. Beran, J. Phys. Chem., 93 (1989) 6725.
33 R.A. van Santen, A.J.M. de Man, W.P.J.H. Jacobs, E.H. Teunissen and G.J. Kramer, Catal.
lett., 9 (1991) 273.
34 V.B. Kazansky and I.N. Senchenya, J. Catal., 119 (1989) 108.
35 W.H. Saunders, Jr. and A.F. Cockerill, Mechanisms of Elimination Reactions, Wiley, New
York, 1973, p. 256.
36 J.H. de Boer, R.B. Fahim, B.G. Linsen, W.J. Visseren and W.F.N.M. de Vleesschauwer, J.
Catal., 7 (1967) 163.
37 P. Hogan and J. Pasek, Collect. Czech. Chem. Commun., 39 (1974) 3696.
38 Y. Ben Taarit, M.V. Mathieu and C. Naccache, Molecular Sieve Zeolites II, Adv. Chem. Ser.
102, Amer. Chem. Sot., Washington, 1971, p. 362.
39 J.W. Ward, in J.A. Rabo (Editor), Zeolite Chemistry and Catalysis, ACS Monograph 171,
Washington, DC, 1976, p. 118.
40 A. Corma, V. Form%, F. Melo and J. Perez Pariente, in M. Occelli (Editor), Fluid Catalytic
Cracking, ACS Symposium Series 375, Amer. Chem. Sot., Washington, 1988, p. 49.
41 S.G. Hegde, R. Kumar, R.N. Bhat and P. Ratnasamy, Zeolites, 9 (1989) 231.

Вам также может понравиться