You are on page 1of 11

Applied Catalysis A: General 248 (2003) 291301

Beckmann rearrangement over phosphotungstic


acid/SiMCM-41 cyclohexanone oxime
to -caprolactam
R. Maheswari
a
, K. Shanthi
a
, T. Sivakumar
a
, Sankarasubbier Narayanan
b,
a
Department of Chemistry, Anna University, Chennai 600025, India
b
Inorganic and Physical Chemistry Division, Indian Institute of Chemical Technology,
Hyderabad 500007, India
Received 14 September 2002; received in revised form 6 January 2003; accepted 20 February 2003
Abstract
Vapor-phase Beckmann rearrangement of cyclohexanone oxime to -caprolactam is reported for the rst time on phospho-
tungstic acid (PWA)-supported SiMCM-41 catalyst. The catalyst was prepared in the laboratory following standard procedures
and was characterised by XRD and FT-IR. By N
2
adsorption, BET surface area, pore size and pore diameter were measured.
The conversion of cyclohexanone oxime over the catalyst was studied in the temperature region of 250400

C using mainly
acetonitrile as a solvent at different contact times. Increase in the reaction temperature to 325

C enhances oxime conversion


and the selectivity of -caprolactam. An impressive catalytic performance of cyclohexanone oxime conversion >99% and
-caprolactamselectivity of 75%was observed on 30 wt.%PWA/SiMCM-41 at 325

Cusing a feed containing10 wt.%oxime


in acetonitrile at WHSV = 3.24 h
1
. The catalytic efciency for this rearrangement reaction was studied, varying the PWA
content from 10 to 50 wt.% on SiMCM-41 and also using benzene and ethanol as alternate solvents in place of acetonitrile.
The rearrangement reaction and -caprolactam selectivity with respect to catalyst physical characteristics, PWA content and
type of solvent are discussed.
2003 Elsevier Science B.V. All rights reserved.
Keywords: Beckmann rearrangement; Cyclohexanone oxime; -Caprolactam; SiMCM-41; Phosphotungstic acid
1. Introduction
-Caprolactam is an important precursor for nylon-6
and plastics. The conventional method of prepara-
tion of -caprolactam from cyclohexanone oxime via
Beckmann rearrangement results in the formation of
a large amount of by-products, especially ammonium
sulfate and approximately 45 tonnes of (NH
4
)
2
SO
4

Corresponding author. Tel.: +91-40-27-19-32-00;


fax: +91-40-2716-09-21.
E-mail address: snarayanan@iict.ap.nic.in (S. Narayanan).
per tonne of -caprolactam is inevitably obtained
[1,2]. Further problems encountered include handling
of a large amount of oleum and corrosion of the ap-
paratus. To overcome the above-mentioned problems,
solid acid catalysts were investigated for vapor-phase
Beckmann rearrangement of cyclohexanone oxime.
There are a number of reports on the use of hetero-
geneous catalysts such as alumina, heteropolyacids,
boronphosphate, phosphoric acid, silicaalumina and
boric acid on alumina and silica-supported tantalum
oxide catalysts [3] for the rearrangement of a variety
of ketoximes to amides [2,413]. There are sev-
0926-860X/$ see front matter 2003 Elsevier Science B.V. All rights reserved.
doi:10.1016/S0926-860X(03)00182-0
292 R. Maheswari et al. / Applied Catalysis A: General 248 (2003) 291301
eral studies on the rearrangement of cyclohexanone
oxime to -caprolactam over zeolites Y [4,14], ZSM-5
[15], Beta [16] and mordenite [17]. The principal
by-product was cyanopentene together with traces
of cyclohexanone and cyclohexanol. Aucejo et al.
[14] reported that the by-product cyanopentene was
formed mainly on the Na
+
ions of the zeolite. The
Beckmann rearrangement is believed to take place
even at 100

C, but the desorption of the product


takes place only at temperatures higher than 300

C
[18]. Above 360

C, there is a decrease in selectivity


due to the decomposition of -caprolactam on the cat-
alyst surface [19]. Medium polar solvents are more
favorable than high polar or non-polar solvents [20].
Chung et al. [21] proposed that the high polarity is
preferred for promoting the migration of OH
2
+
group
in the rearrangement step. However, the efciency of
the solid acid catalysts is low towards -caprolactam
formation, because of the rapid deactivation during
the reaction [2224]. Sato et al. [15,25] have reported
that the Beckmann rearrangement occurs on the ex-
ternal surface of ZSM-5 type zeolites. Rseler et al.
[26] and Dahlhoff and co-workers [27,28] claim that
only extremely weak acidic sites on zeolitic cata-
lysts are needed for the Beckmann rearrangement.
The silanol nests of SiMCM-41 are the most suit-
able for the reaction; the vicinal silanol groups are
more favorable than the terminal silanols. Yashima
et al. [29] concluded by studying the vapor-phase
Beckmann rearrangement of cyclohexanone oxime
on zeolites with different pore windows that the
selective rearrangement reaction occurred on the ex-
ternal surface of zeolite crystals. Takahashi et al.
[22] studied the kinetics of Beckmann rearrangement
on HZSM-5 and the effect of acid strength on the
reaction. They concluded that, if the reaction had
proceeded on the outer surface of the zeolite, the
rate constant should be directly proportional to the
outer surface area. However, the rate constant was
not proportional to the outer surface area, so the re-
sult indicated that the acid sites on the outer surface
area were not necessarily effective for the reaction
of cyclohexanone oxime. Other reports [19,30] on
AlMCM-41 and MCM-22 also reveal the importance
of weak acidic sites, larger outer surface and pore
structure for the above transformation. Silanol groups
on SiMCM-41 are not sufciently acidic [31,32] to
catalyse the rearrangement and the neutral silanol
groups cannot be the active centres for Beckmann
rearrangement [33]. However, there are reports claim-
ing that the medium and weak Bronsted acid sites
or even neutral silanol groups are most favourable
for Beckmann rearrangement [7,14,3436]. Even
though silanol groups provide considerable conver-
sion and selectivity towards -caprolactam, they are
also responsible for the formation of by-products.
Chaudhari et al. [19] reported that, while acidic sites
appear to catalyse the rearrangement, SiOH groups
catalyse the formation of hydrolysis product. It has
also been reported that siliceous materials are not
good enough due to high hydrophobicity and that
the main by-product is cyanopentene [37]. There
is a considerable inuence of moderate acidity and
silanol groups on the Beckmann rearrangement;
most of the reports support the assertion that the
silanol groups on the external surface are responsible
[3840].
The production of -caprolactam via cyclohexanone
oxime intermediate involves three different steps:
(i) the synthesis of cyclohexanone;
(ii) amoximation of cyclohexanone to its oxime;
(iii) the Beckmann rearrangement of the oxime to
-caprolactam.
The rst step involves the conversion of benzene
to cyclohexene/cyclohexane and to cyclohexanol/
cyclohexanone. Cyclohexanone can also be obtained
from phenol. Narayanan and Krishna [4144] have
reported a highly selective synthesis of cyclohex-
anone from phenol over Pd/hydrotalcite. The present
work concerns the Beckmann transformation of the
cyclohexanone oxime to -caprolactam. There is a
considerable scope for understanding the Beckmann
rearrangement of oxime from the point of view of the
mechanism, the dependence on physical factors and
acidity of solid acids. Here, the focus of attention is
on the preparation of SiMCM-41 and phosphotungstic
acid (PWA)-loaded (1050 wt.%) SiMCM-41 and
evaluation of them as catalysts for the rst time in the
Beckmann rearrangement reaction of cyclohexanone
oxime to -caprolactam. The effects of various reaction
parameters, such as temperature, time on stream, use
of different solvents, contact time and concentration
of oxime in the feed are investigated. The inuences
of physical properties of the PWA/SiMCM-41 are cor-
R. Maheswari et al. / Applied Catalysis A: General 248 (2003) 291301 293
Table 1
Physical characteristics of SiMCM-41 and PWA/SiMCM-41
Sample PWA (wt.%) d
1 0 0
spacing
()
S
BET
(m
2
/g)
Pore volume
(ml/g)
Pore
diameter ()
SiMCM-41 41.63 1120 0.80 31
SiMCM-41/PWA (10) 10 37.02 963 0.63 28
SiMCM-41/PWA (30) 30 34.92 789 0.47 28
SiMCM-41/PWA (50) 50 30.56 606 0.23 26
related with oxime transformation and -caprolactam
selectivity.
2. Experimental
2.1. Catalyst preparation
SiMCM-41 material was synthesised by an isother-
mal method [45]. Sodium meta silicate (E-Merck) was
used as silica source. Sodium meta silicate (10.6 g)
was dissolved in 60 g of water and the mixture was
thoroughly stirred until a clear solution was obtained.
Cetyltrimethylammonium bromide (3.36 g; CTAB,
E-Merck, India) was dissolved in 20 g of ethanol.
To this, sodium meta silicate solution was added
dropwise. The resultant mixture was stirred for 1 h;
then the pH of the resulting gel was adjusted to 11
using 4N sulfuric acid, followed by 3 h stirring. The
resulting homogenous solution was transferred to an
autoclave and heated to 140

C in static conditions
for 12 h.
The molar composition of the gel was
SiO
2
: 9.0 EtOH : 0.20 CTAB : 160 H
2
O
The resulting precipitate was recovered by ltra-
tion, washed with deionised water, dried in air at
ambient temperature and nally calcined at 550

C
for 1 h in a nitrogen ow and then in air for
12 h.
Catalyst samples with different loadings of PWA
(E-Merck, AR Grade, India), were prepared by stir-
ring 1.0 g of SiMCM-41 with 10 ml of an aqueous
solution containing 0.1, 0.3 or 0.5 g PWA at room
temperature for 24 h. The resultant mixture was l-
tered, dried at 80

C and calcined at 200

C for
2 h.
2.2. Characterisation
The mesoporous materials were characterised by
XRD (Rigaku, D-Max/IIIVC model) using nickel l-
tered Cu K radiation, = 1.5406 ; surface area
measurements were done by a NOVA 1200, Quan-
tachrome (USA) instrument. FT-IR spectra were taken
on the instrument Bio-RAD FTS 175c FT-IR using
catalyst samples supported on KBr wafers. Surface
area, pore volume and pore size description details of
the catalysts are given in Table 1.
2.3. Reaction procedure
Cyclohexanone oxime (Aldrich), acetonitrile, ben-
zene and ethanol (Analytical Grade, E-Merck, India)
were used as solvents for dissolving oxime without
further purication. The vapor-phase Beckmann rear-
rangement of cyclohexanone oxime to -caprolactam
was carried out using a xed bed vertical down
ow reactor (length, 45 cm; diameter, 19 mm) with
a catalyst charge of 1.5 g (1020 mesh). Prior to the
start of the experiment, the catalyst was activated
at 330

C for 3 h in owing air. The temperature of


activation is kept below 350

C, because above this


temperature the heteropoly acid would decompose.
The catalyst was brought to the reaction tempera-
ture (250400

C) in a ow of dry nitrogen. The


reaction feed containing a solution of 10 wt.% cyclo-
hexanone oxime in acetonitrile was introduced into
the reactor from the top using a calibrated motorised
syringe, along with nitrogen carrier gas (20 ml/min).
The products along with solvent in liquid state was
collected in a trap at room temperature and are anal-
ysed by gas chromatography (HP 6890 series) with
HP-1 (100% dimethylpolysiloxane, 30 m 0.25 mm)
capillary column and FID detector. The analysis was
294 R. Maheswari et al. / Applied Catalysis A: General 248 (2003) 291301
conrmed by GC-MS (QP 2000A, Shimadzu SE-52
column).
3. Results and discussion
3.1. Catalyst characterization
Table 1 gives the physical characteristics of
SiMCM-41 and PWA-supported catalysts. SiMCM-41
has a large surface area of 1120 m
2
/g with pore vol-
ume 0.80 ml/g and diameter 31 . Impregnation of
this SiMCM-41 with 10, 30 and 50 wt.% of PWA de-
creases surface area, pore volume and pore diameter.
Addition of 10 wt.% PWA brings down the surface
area of SiMCM-41 by approximately 15% from 1120
to 963 m
2
/g. The surface area of SiMCM-41 decreases
to nearly half with the addition of 30 and 50 wt.%
PWA. The decrease of surface area may be mainly
attributed to the corresponding decrease of pore vol-
ume as well as pore diameter. A signicant decrease
in the pore volume suggests that PWA occupies vol-
ume inside the pore as well as around the pore mouth.
A reasonably large surface area of PWA-loaded
SiMCM-41 suggests that the mesoporous structure
is rather maintained and this is supported by XRD
(Fig. 1).
XRD patterns of SiMCM-41 and PWA/SiMCM-41
are shown in Fig. 1. SiMCM-41 has a typical XRD
pattern exhibiting an intense peak at d = 41.63 and
a much less intensive peak at d = 17.42 , which is
in accordance with the earlier observations reported
in [45,46]. Addition of 10 wt.% of PWA did not af-
fect the XRD pattern very much. However, addition
of 30 wt.% PWA alters the intensity of the peaks at
d = 41.63 as well as at d = 17.42 , and new peaks
start to appear. In the case of 50 wt.% PWA-loaded
SiMCM-41, the intensity of the peaks around d =
17.42 increases along with the decrease of the peak
at d = 41.63 . A comparison of the XRD patterns
of pristine SiMCM-41 and PWA-loaded SiMCM-41
shows that, with the PWA loading, the mesoporous
structure is rather intact. This is corroborated with the
reasonable surface area of PWA/SiMCM-41, which
is one of the characteristics of mesoporous structure.
On higher loading, PWA spreads on the surface espe-
cially in and around the pore mouth of mesoporous
SiMCM-41, contributing to the crystallinity observed
Fig. 1. XRD pattern of (a) SiMCM-41; (b) 10 wt.% PWA/SiMCM-
41; (c) 30 wt.% PWA/SiMCM-41; (d) 50 wt.% PWA/SiMCM-41.
in the XRD pattern as well as to the decreases in the
pore volume and surface area.
FT-IR spectra of bulk PWA as well as SiMCM-41-
supported PWA are shown in Fig. 2. The IR frequen-
cies of bulk PWA are in good agreement with the
reported values (cm
1
): 1081 (PO), 985 (W=O) and
897, 803 (WOW) [47]. The IR spectra of supported
SiMCM-41 are similar to that of bulk PWA, even
though the peaks are less intensive. It appears that
the PWA keggin structure is retained in the supported
sample. One sees that the bands for outer groups
(W=O, corner- and edge-bridging WOW) show a
small shift of 412 cm
1
compared to the spectrum of
bulk PWA. This may be due to a chemical interaction
between the PWA anion and the SiMCM-41 inner
surface.
3.2. Catalysis
3.2.1. Effect of temperature
The transformation of cyclohexanone oxime was
studied in the temperature region 250400

C on
R. Maheswari et al. / Applied Catalysis A: General 248 (2003) 291301 295
Fig. 2. FT-IR spectra of (a) PWA; (b) 10 wt.% PWA/SiMCM-41; (c) 30 wt.% PWA/SiMCM-41; (d) 50 wt.% PWA/SiMCM-41.
SiMCM-41 as well as on PWA-supported SiMCM-41
catalysts. Fig. 3 shows the variation of oxime con-
version as well as the selectivity of prime product
-caprolactam and others. Cyclohexanone oxime
conversion increases from around 70% at 250

C
and reaches a steady state of nearly 100% around
300

C. With increase of temperature from 250

C,
-caprolactam selectivity generally decreases and the
decrease is signicant beyond 325

C. Of the cat-
alysts studied, 30 wt.% PWA shows a conversion
of 100% and -caprolactam selectivity of 75%
at 300

C. The pristine SiMCM-41 and 10 wt.%


PWA/SiMCM-41 show -caprolactam selectivity of
60% at the same temperature. Fifty percent by weight
of PWA/SiMCM-41 shows the least selectivity to
-caprolactam at same temperature, even though the
oxime conversion is more or less similar to that of
other catalysts. The other products such as cyanopen-
tene and cyclohexanone follow a selectivity pattern
opposite to that of -caprolactam with respect to tem-
perature. Of the side products formed, cyanopentene
is always more than cyclohexanone at higher tem-
peratures. From conversion and selectivity patterns,
it may be said that the 30 wt.% PWA/SiMCM-41 is
selective for -caprolactam at 325

C.
3.2.2. Effect of space velocity
The inuence of feed rate of the reactant oxime on
the surface of 30 wt.% PWA/SiMCM-41 was studied
by using 10 wt.% cyclohexanone oxime in acetoni-
trile at 325

C (Fig. 4). While the catalyst weight was


kept at 1.5 g, the contact time was varied by chang-
296 R. Maheswari et al. / Applied Catalysis A: General 248 (2003) 291301
Fig. 3. Effect of reaction temperature on cyclohexanone oxime transformation: (a) SiMCM-41; (b) 10 wt.% PWA/SiMCM-41; (c)
30 wt.% PWA/SiMCM-41; (d) 50 wt.% PWA/SiMCM-41. Catalyst weight = 1.5 g; feed = 10 wt.% cyclohexanone oxime in acetonitrile;
WHSV = 3.24 h
1
; N
2
carrier = 20 ml/h; time on stream = 2 h.
ing the feed rate, for example, 3, 5, 10 and 15 ml/h
(i.e. 1.59.6 WHSVh
1
). With increase of feed rate,
the oxime conversion decreases from nearly 100 to
70%, as expected, and the -caprolactam selectiv-
ity increases from 25 to 80%. The other products,
cyanopentene and cyclohexanone, follow the oppo-
site trend to that of -caprolactam. Similar studies on
the effect of contact time on conversion and selectiv-
ity were carried out on 30 wt.% PWA/SiMCM-41 at
325

C, by changing the concentration of oxime in the


feed and keeping the feed rate at 5 ml/h (Table 2). With
increase in oxime concentration from 5 to 20 wt.% in
the feed, the oxime conversion decreases from nearly
100 to 90%. The -caprolactam selectivity under the
same condition decreases from 84 to 62%, while the
other products, cyanopentene and cyclohexanone in-
crease from 11 to 23% and 5 to 16%, respectively.
The contact time studies also suggest an inverse
relationship in the selectivity of -caprolactam and
other products, as seen in the case of the temperature
effect.
3.2.3. Effect of solvent and time on stream
Several polar as well as non-polar solvents have
been used for dissolving the cyclohexanone oxime. It
has been reported that these solvents, depending on the
polarity, contribute to 1,2-H shift or migration of pro-
ton during this transformation [21]. The mechanism
R. Maheswari et al. / Applied Catalysis A: General 248 (2003) 291301 297
Fig. 4. Effect of WHSV (h
1
) on cyclohexanone oxime transformation over 30 wt.% PWA/SiMCM-41. Catalyst weight = 1.5 g;
feed = 10 wt.% cyclohexanone oxime in acetonitrile; temperature = 325

C; N
2
carrier = ml/h; time on stream = 2 h.
of proton shift is visualised to take place as follows:
We have chosen benzene (non-polar), ethanol
(medium polar) and acetonitrile (highly polar) as sol-
vents for cyclohexanone oxime to study the polarity
of solvents on conversion and the selectivity of the
products for this reaction. Each reaction has been
carried out for a period of 7 h, collecting the product
samples at every hour and analysing them. The sol-
vent effect and time on stream on cyclohexanone
oxime transformation are shown in Fig. 5. Ten per-
cent by weight of cyclohexanone oxime in the se-
lected solvent has been chosen and the feed rate has
been maintained at WHSV 3.24 h
1
and reaction
298 R. Maheswari et al. / Applied Catalysis A: General 248 (2003) 291301
Table 2
Effect of oxime concentration in the feed on cyclohexanone oxime transformation over 30 wt.% PWA/SiMCM-41
Oxime concentration (wt.%) WHSV (h
1
) Oxime conversion (%) Product selectivity (%) -Caprolactam
yield (%)
A B C
5 2.92 100 84 11 5 84
10 3.24 99 74 18 9 73
20 3.88 90 62 23 16 56
(A) -caprolactam; (B) cyanopentene +cyanopentane; (C) cyclohexanone +cyclohexenone. Catalyst weight = 1.5 g; solvent = acetonitrile;
feed = 5 ml/h; N
2
carrier = 20 ml/h; temperature = 325

C.
Fig. 5. Effect of solvent and time on stream on cyclohexanone oxime transformation using (a) benzene; (b) ethanol; (c) acetonitrile as
solvent over 30 wt.% PWA/SiMCM-41. Catalyst weight = 1.5 g; feed = 10 wt.% cyclohexanone oxime in solvent; WHSV = 3.24 h
1
;
temperature = 325

C; N
2
carrier = 20 ml/h.
temperature at 325

C in all the cases. Irrespective of


the solvent used, the conversion is nearly 100% during
the experimental period up to 7 h. The solvents seem
to affect only the selectivity patterns. For example,
when benzene and acetonitrile are used as solvents,
the initial -caprolactam selectivity is around 70% and
it decreases with time on stream; the decrease however
is sharper beyond 4 h in the case of benzene than in
the case of acetonitrile. The selectivity is maintained
around 55% up to 7 h. When ethanol is used as a
solvent, the -caprolactam selectivity is only between
20 and 40% during the reaction period studied. Gen-
erally, cyanopentene increases with time on stream
and increases signicantly in the acetonitrile solvent
from 20 to 45%. Correspondingly, the cyclohexanone
selectivity is low in the case of acetonitrile solvent.
The solvent and time on stream studies bring out
the interdependence of -caprolactam and cyanopen-
tene selectivity and their independence with respect
to conversion. This is due to the competitive nature
of formation of caprolactam and cyanopentene. The
reaction mechanism of formation of the by-products
R. Maheswari et al. / Applied Catalysis A: General 248 (2003) 291301 299
involving silanol group is given below:
3.3. Effect of PWA loading
Cyclohexanone oxime transformation was studied
on SiMCM-41 as well as on 10, 30 and 50 wt.%
PWA-loaded SiMCM-41. The detailed studies on the
effect of temperature on such catalysts have been dis-
cussed already in the previous section (Fig. 3). Here,
we compare the oxime conversion and -caprolactam
selectivity on these catalysts at a chosen tempera-
ture of 325

C for the reactions carried out using


10 wt.% cyclohexanone oxime in acetonitrile as feed,
WHSV = 3.24 h
1
(Fig. 6). Pristine SiMCM-41 and
PWA-supported SiMCM-41 (irrespective of the PWA
loading) show nearly the same oxime conversion of
>99%. The -caprolactam selectivity of SiMCM-41
and 10 wt.% PWA/SiMCM-41 is the same, around
300 R. Maheswari et al. / Applied Catalysis A: General 248 (2003) 291301
Fig. 6. Effect of PWA loading of SiMCM-41on cyclohexanone oxime transformation. Catalyst weight = 1.5 g; feed = 10 wt.% cyclohexanone
oxime in acetonitrile; WHSV = 3.24 h
1
; temperature = 325

C; N
2
carrier = 20 ml/h.
50% whereas 30 wt.% PWA catalyst shows a higher
selectivity of 75%. Increasing the PWA loading to
50 wt.% brings down the -caprolactam selectivity to
20%, even though the conversion is maintained at
>99%. The fact that there is a good conversion com-
parable to that of the PWA-loaded SiMCM-41 shows
that the SiMCM-41 itself is active for this transforma-
tion reaction. Many reports claim that the SiMCM-41
is nearly neutral or weakly acidic. It appears that
the inherent acidity in SiMCM-41 is sufcient for
the oxime conversion. The observation that 10 wt.%
PWA shows the same conversion and -caprolactam
selectivity as that of pristine SiMCM-41 indicates
that the added PWA is not available for the reaction.
This is conrmed by the identity of the XRD pat-
terns of SiMCM-41 and 10 wt.% PWA/SiMCM-41
being the same (Fig. 1). On 30 wt.% PWA loading of
SiMCM-41, XRD patterns show crystalline PWA on
the surface. These crystalline PWA, which may be well
dispersed, contribute to the -caprolactam selectivity.
However, further loading of SiMCM-41 with PWA
(50 wt.%) brings down the -caprolactam selectivity,
though the oxime conversion is the same as in the
case of other catalysts. In 50 wt.% PWA/SiMCM-41,
the PWA may be large and bulky agglomerates and
these may not contribute to the selectivity. These
observations lead to the conclusion that the cyclohex-
anone oxime transformation can take place even on
pristine SiMCM-41. Addition of PWA up to 30 wt.%
contributes to the selectivity of -caprolactam, be-
cause of the contribution to acidity by well-dispersed
PWA. However, bulk PWA that may be too strong in
acidity is not helpful for -caprolactam selectivity.
4. Conclusions
The cyclohexanone oxime transformation reactions
over SiMCM-41 as well as PWA-loaded SiMCM-41
are studied at different reaction conditions of temper-
ature, contact time and time on stream. SiMCM-41
itself is an active catalyst for this reaction. Ad-
dition of PWA improves mainly the selectivity of
-caprolactam. Between 300 and 325

C, a con-
version of >99% with -caprolactam selectivity of
75% is achieved on 30 wt.% PWA/SiMCM-41 using
10 wt.% cyclohexanone oxime in acetonitrile feed.
Well-dispersed PWA helps -caprolactam selectivity,
even though oxime conversion is insensitive to PWA
addition. Conversion being the same, the selectivity
of -caprolactam and of other products, cyanopentene
and cyclohexanone, depend on reaction variables;
the selectivity of -caprolactam is inversely related
to that of cyanopentene because of the competi-
tive nature of formation of these products from the
oxime.
R. Maheswari et al. / Applied Catalysis A: General 248 (2003) 291301 301
Acknowledgements
SN thanks the CSIR, New Delhi for the Emeritus
Scientist Position. The authors wish to thank Depart-
ment of Science and Technology (DST), New Delhi,
for providing nancial support. We thank the Direc-
tor of IICT for extending the facilities to carry out the
characterization work.
References
[1] Kirk-Othmer. Encyclopedia of Chemical Technology, vol. 4,
fourth ed., Wiley, New York, 1995, p. 827.
[2] O. Immel, H.H. Schwarz, G. Starke, W. Swodenk, Chem.
Ing. Technol. 56 (1984) 612.
[3] T. Ushikubo, K. Wada, J. Catal. 148 (1994) 138.
[4] P.B. Venuto, P.S. Landis, J. Catal. 6 (1966) 245.
[5] D.C. England, Process for Manufacturing Caprolactam. US
Patent 2,634,269 (1953).
[6] I.S. Fisher, Improvements in or Relating to the Production of
Amides and Lactams, UK Patent 881,276 (1961).
[7] A. Costa, P.M. Deya, J.V. Sinisterra, J.M. Marinas, Can. J.
Chem. 58 (1980) 1266.
[8] P. Vitarelli, S. Cavallaro, R. Maggiore, G. Cimino, C. Caristi,
S. Galvagno, Gazz. Chim. Ital. 112 (1982) 493.
[9] T. Curtin, J.B. McMonagle, B.K. Hodnett, Appl. Catal. A 93
(1992) 91.
[10] T. Curtin, J.B. McMonagle, B.K. Hodnett, Appl. Catal. A 93
(1992) 75.
[11] T. Curtin, J.B. McMonagle, M. Ruwet, B.K. Hodnett, J. Catal.
142 (1993) 172.
[12] S. Sato, K. Hirose, M. Kitamura, H. Tojima, N. Ishii,
European Patent 236,092 (1987).
[13] S. Sato, H. Sakurai, K. Urabe, Y. Izumi, Chem. Lett. 3 (1985)
277.
[14] A. Aucejo, M.C. Burguet, A. Corma, V. Fornes, Appl. Catal.
22 (1986) 187.
[15] H. Sato, K. Hirose, M. Kitamura, Y. Nakamura, Stud. Surf.
Sci. Catal. 49 (1989) 1213.
[16] L.-X. Dai, R. Hayasaka, Y. Iwaki, K.A. Koyano, T. Tatsumi,
Chem. Commun. 9 (1996) 1071.
[17] H. Sato, K. Hirose, N. Ishii, Y. Umada, European Patent
234,088 (1986).
[18] H. Sato, K. Hirose, Y. Nakamura, Chem. Lett. (1993) 1987.
[19] K. Chaudhari, R. Bal, A.J. Chandwadkar, S. Sivasanker, J.
Mol. Catal. A 177 (2002) 247.
[20] T. Komatsu, T. Maeda, T. Yashima, Microporous Mesoporous
Mater. 3536 (2000) 173.
[21] Y.M. Chung, H.K. Rhee, J. Mol. Catal. A: Chem. 159 (2000)
389.
[22] T. Takahashi, M. Nishi, Y. Tagawa, T. Kai, Microporous
Mater. 3 (1995) 467.
[23] J.S. Reddy, R. Rvaisankar, S. Sivasanker, P. Ratnasamy, Catal.
Lett. 17 (1993) 139.
[24] P. OSullivan, L. Forni, B.K. Hodnett, Ind. Eng. Chem. Res.
40 (2001) 1471.
[25] H. Sato, N. Ishii, K. Hirose, S. Nakamura, Stud. Surf. Sci.
Catal. 28 (1986) 755.
[26] J. Rseler, G. Heitmann, W.F. Hlderich, Appl. Catal. A 144
(1996) 319.
[27] G.P. Heitmann, G. Dahlhoff, W.F. Hlderich, J. Catal. 186
(1999) 12.
[28] G. Dahlhoff, G.P. Heitmann, J.P.M. Niederer, W.F. Hlderich,
J. Catal. 194 (2000) 122.
[29] T. Yashima, K. Miura, T. Komatsu, Stud. Surf. Sci. Catal. 84
(1994) 1897.
[30] G. Dahlhoff, U. Barsnick, W.F. Hlderich, Appl. Catal. A
210 (2001) 83.
[31] M. Busio, J. Janchen, J.H.C. Van Hoff, Microporous Mater.
5 (1995) 211.
[32] A. Corma, V. Fornes, M.T. Navarro, J. Perez-Pariarte, J. Catal.
148 (1994) 569.
[33] M. Kitamura, H. Ichihashi, Stud. Surf. Sci. Catal. 90 (1994)
67.
[34] N. Katada, T. Tsubouchi, M. Niwa, Y. Murakami, Appl. Catal.
A 124 (1995) 1.
[35] E.R. Herrero, O.A. Anunziata, L.B. Pierella, O.A. Oria, Latin
Am. Appl. Res. 24 (1994) 195.
[36] T. Curtin, J.B. McMonagle, B.K. Hodnett, Catal. Lett. 17
(1993) 145.
[37] A. Corma, H. Garcia, J. Primo, Zeolites 11 (1991) 593.
[38] H. Sato, K. Hirose, Chem. Lett. (1993) 1765.
[39] G.P. Heitmann, G. Dalhlhoff, W.F. Holderich, J. Catal. 186
(1999) 12.
[40] T. Yashima, K. Miura, T. Komatsu, Stud. Surf. Sci. Catal. 84
(1994) 1897.
[41] S. Narayanan, K. Krishna, Chem. Commun. (1997) 1991.
[42] S. Narayanan, K. Krishna, Appl. Catal. A 174 (1998)
221.
[43] S. Narayanan, K. Krishna, Catal. Today 49 (1999) 57.
[44] S. Narayanan, K. Krishna, Appl. Catal. A 198 (2000) 13.
[45] J.S. Beck, J.C. Vartuli, W.J. Roth, M.E. Leonowicz, C.T.
Kresge, K.D. Schmitt, C.T.-W. Chu, D.H. Olson, E.W.
Sheppard, S.B. McCullen, J.B. Higgins, J.L. Schlenker, J.
Am. Chem. Soc. 114 (1992) 10834.
[46] C.-Y. Chen, H.-X. Li, M.E. Davis, Microporous Mater. 2
(1993) 17.
[47] C. Rocchiccioli-Deltcheff, M. Fournier, R. Frank, R.
Thouvenot, Inorg. Chem. 22 (1983) 207.

You might also like