Você está na página 1de 11

Applied Catalysis A: General 268 (2004) 139149

Characterization of Co,Al-MCM-41 and its activity in


the t-butylation of phenol using isobutanol
M. Karthik
a
, A.K. Tripathi
b
, N.M. Gupta
b
, A. Vinu
c
, M. Hartmann
c
,
M. Palanichamy
a
, V. Murugesan
a,
a
Department of Chemistry, Faculty of Science and Humanities, Anna University, Chennai 600025, India
b
Applied Chemistry Division, BARC, Mumbai 400085, India
c
Department of Chemistry, Chemical Technology, Kaiserslautern University of Technology, P.O. Box 3049, D-67653 Kaiserslautern, Germany
Received in revised form 3 March 2004; accepted 18 March 2004
Available online 26 April 2004
Abstract
Co,Al-MCM-41 catalysts with various n
Si
/(n
Co
+n
Al
) ratios were synthesized and extensively characterized by low-angle XRD, TGA/DTG,
BET, AAS, DRIFT, UV-Vis DRS and ESR. UV-Vis DRS and ESR studies reveal that cobalt in Co,Al-MCM-41 is highly symmetrical and
occurs in tetrahedral coordination. Some of the cobalt atoms are transformed into the cobalt oxide form when n
Si
/(n
Co
+n
Al
) is increased to
20. t-Butylation of phenol with isobutanol was studied in the vapor phase as a model reaction at temperatures between 200 and 500

C. The
products obtained were O-tert-butyl phenol (OTBP), 2-tert-butyl phenol (2TBP) and 4-tert butyl phenol (4TBP). O-Butenyl phenol (OBP) and
2-butenyl phenol (2BP) were also observed along with normal alkylated products. The phenol conversion drastically increased with temperature
over all the catalysts. The activity of the catalysts followed the order of Co,Al-MCM-41 (20) > Co,Al-MCM-41 (50) > Co,Al-MCM-41
(80) > Al-MCM-41 (23). The inuences of various parameters such as temperature, reactant feed ratio and feed rate, time on stream on
conversion and products selectivity were studied and the salient results are discussed.
2004 Elsevier B.V. All rights reserved.
Keywords: Mesoporous materials; Co,Al-MCM-41; t-Butylation; Phenol; Isobutanol
1. Introduction
The M41S family of mesoporous materials has attracted
substantial research attention since its report by Mobil Cor-
poration in 1991 [1,2]. The past decade has witnessed a
dramatic increase in the design, synthesis, characterization
and property evaluation of mesoporous molecular sieves for
catalysis, adsorption and separation of bulky molecules and
environmental pollution control. MCM-41, the most impor-
tant member of the M41S family, possesses a regular hexag-
onal array of uniform pores with diameters in the range
210 nm. These hexagonal MCM-41 molecular sieves have
attracted much research attention owing to their high surface
area, thermal and hydrothermal stability and ordered meso-

Corresponding author. Tel.: +91-44-22301168;


fax: +91-44-22200660.
E-mail addresses: karthik annauni@yahoo.co.in (M. Karthik),
v murugu@hotmail.com (V. Murugesan).
porous nature. They nd potential applications as catalysts,
adsorbents and hosts for different kinds of molecules [35].
Pure siliceous mesoporous molecular sieves are of limited
use as catalysts because of the presence of the neutral frame-
work. In order to obtain materials for catalytic applications,
one must modify the nature of the amorphous walls by in-
corporation of heteroatoms. The isomorphous substitution
of Al
3+
for some of Si
4+
leads to negatively charges in the
framework that are balanced by protons. The resulting acid
sites can be formally represented as SiO(H)Al. The bron-
sted acid sites in aluminosilicates are due to OH bridged
between Si and Al present in the framework [6,7]. MCM-41
containing metal ions such as Al, V, Fe, Mn, B, Ga, Ti, Zr,
Cs, Cu and Zn have been synthesized and their catalytic
properties have been reported [810]. The incorporation of
cobalt into the framework imparts both acidic and redox
properties, which make such materials potentially interest-
ing for catalytic applications [11]. It is interesting to note
that substitution of Co
2+
into the framework of mesoporous
0926-860X/$ see front matter 2004 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2004.03.028
140 M. Karthik et al. / Applied Catalysis A: General 268 (2004) 139149
MCM-41 leads to moderate acidity. Alkylation of phenol is
an industrially important reaction because many alkylated
phenols have commercial applications. t-Butyl derivatives
of phenol are the precursors for a number of commercially
important antioxidants in the synthesis of various agrochem-
icals, fragrants, thermoresistant polymers and protecting
agents for plastics. The products like O-butyl and O-butenyl
phenols are used in the synthesis of avoring and fragrant
compounds, pharmaceutical intermediates, chromans, ne
chemicals and petrochemicals [12]. In general, this reaction
is carried out over the conventional acid catalysts like AlCl
3
or sulfated zirconia in liquid phase [13,14], these catalysts
are environmentally hazardous, require tedious work-up
and cannot be used at higher temperatures. Zeolite catalysts
with isobutanol/t-butanol in vapor phase have also been
attempted for this reaction, but it produces a mixture of
ortho, para and meta butylated products or predominantly
para products and they undergo deactivation signicantly
with time on stream [15]. However, t-butylation of phe-
nol with isobutanol has not been studied over mesoporous
Al-MCM-41 catalysts. In the present investigation, attempt
has been made to synthesize mesoporous Co,Al-MCM-41
(n
Si
/(n
Co
+ n
Al
) various ratios) by hydrothermal process.
t-Butylation of phenol with isobutanol over these catalysts
was studied extensively in the vapor phase at 200, 250,
300, 350, 450 and 500

C, the results are discussed in this


paper.
2. Experimental
2.1. Preparation of the catalysts
Co,Al-MCM-41 samples with various n
Si
/(n
Co
+ n
Al
)
ratios (the n
Co
/n
Al
was xed to 1) were synthesized from
gels with the following gel composition: 10 SiO
2
:5.4
C
14
TMABr:0.030.125 Al
2
O
3
:0.060.25 CoO:4.25 Na
2
O:
1.3 H
2
SO
4
:480 H
2
O. The typical synthesis procedure for
Co,Al-MCM-41 using tetradecyltrimethylammonium bro-
mide was performed as follows: 32 g of tetradecyltrimethy-
lammonium bromide and 115 g of water were mixed and
stirred for 30 min. Thereafter, 37.4 g of sodium silicate so-
lution was added drop wise to the surfactant solution under
vigorous stirring. After stirring for another 30 min, the ap-
propriate amounts of cobalt acetate tetrahydrate and sodium
aluminate for the desired n
Si
/(n
Co
+ n
Al
) ratio were dis-
solved in 5 ml of water, added to the synthesis mixture and
the resulting gel was stirred for another 15 min. Then 2.4 g
H
2
SO
4
in 10 g of water was added to the above mixture
to reduce the pH to ca. 11.5. The stirring was continued
for another 30 min, the resulting gel was transferred into
a polypropylene bottle and kept in an oven at 100

C for
24 h. After cooling to room temperature, the resultant solid
was recovered by ltration, washed with distilled water and
dried in an oven at 100

C for 6 h. Finally, the materials


were calcined in a mufe furnace at 540

C for 10 h.
2.2. Characterization
The powder X-ray diffraction patterns of the Co,Al-
MCM-41 materials were collected on a Siemens D5005
diffractometer using Cu K ( = 0.154 nm) radiation. The
diffractograms were recorded in the 2 range from 0.8 to
10

with a 2 step size of 0.01

and a step time of 10 s.


High-resolution thermogravimetric analysis (SETARAM
setsys 16MS) was carried out under nitrogen atmosphere
in the temperature range 20600

C with a heating rate of


5

C/min. Nitrogen adsorption and desorption isotherms


were collected at 196

C on a Quantachrome Autosorb 1
sorption analyzer. All the samples were outgassed for 3 h at
250

C under vacuum (P < 10


5
hPa) in the degas port of
the adsorption analyzer. The specic surface area was calcu-
lated using the BET model. The pore size distributions were
obtained from the desorption branch of the isotherm using
the corrected form of the Kelvin equation by means of the
BarrettJoynerHalenda method as proposed by Kruk et al.
[16]. UV-Vis diffuse reectance spectra were recorded with
a Perkin-Elmer Lambda 18 spectrometer equipped with a
PrayingMantis diffuse reectance attachment. BaSO
4
was
used as reference. Elementary analysis was done by us-
ing an Analyst AA 300 spectrometer. X-band EPR spectra
were recorded at 267

C using an ESP 380E spectrom-


eter (Bruker) equipped with an ER 041 XK-D microwave
bridge (Bruker) operated with the resonance frequency of
9.7 GHz and 100 KHz eld modulation. DRIFT spectra
of the samples were recorded on a Nicolet (Avatar 360)
FT-IR spectrophotometer equipped with a high temperature
vacuum chamber. Approximately 30 mg of the sample was
taken in the sample holder and dehydrated at 400

C for 6 h
under vacuum (10
5
mbar). The sample was then cooled
to room temperature and the spectra were recorded. Then
pyridine was adsorbed at room temperature. The physically
adsorbed pyridine was removed by heating the sample at
150

C under vacuum (10


5
mbar) for 30 min, after the
sample was cooled to room temperature, the spectrum was
recorded. The same sample was used to record the infrared
spectrum in the range 17001400 cm
1
(pyridine adsorp-
tion region). The number of Bronsted and Lewis acid sites
was calculated by measuring the integrated absorbance of
bands representing pyridinium ion formation and coordi-
natively bonded pyridine [17]. The method developed for
porous aluminosilicates by Emeis [18] was adopted for the
determination of the acidity of the catalysts.
2.3. Catalytic studies
The reactor system was a xed-bed, vertical, downward-
ow type made up of a quartz tube of 40 cm length and
2 cm i.d. The reactor was heated to the requisite temperature
with the help of a tubular furnace controlled by a digital
temperature controller cum indicator. The temperature was
measured by a chromalalumel thermocouple placed inside
the furnace. About 0.5 g of the catalyst was placed in the
M. Karthik et al. / Applied Catalysis A: General 268 (2004) 139149 141
middle of the reactor and supported on either side with a thin
layer of quartz wool and ceramic beads. The reactants were
fed into the reactor using a syringe infusion pump that could
be operated at different ow rates. The reaction was carried
out at atmospheric pressure. The liquid products collected
in the rst 15 min were discarded and the analysis was made
only with the products collected after this time. After each
catalytic run, the catalyst was regenerated by passing mois-
ture and carbon dioxide free air through the reactor for 6 h
at 500

C. The liquid products were analyzed using a Shi-


madzu gas chromatograph (Model: GC-17A) using a DB-5
capillary column (30 m) equipped with a ame ionization de-
tector (FID) and nitrogen as a carrier gas. The identication
of products was also made using a GC-MS Perkin-Elmer
Auto System XL gas chromatograph (Perkin-Elmer elite
series PE-5 capillary column (30 m 0.25 mm 1 m)
equipped with a turbo mass spectrometer (EI, 70 eV) with
helium as carrier gas at a ow rate of 1 ml/min.
3. Results and discussion
3.1. Characterization
The well-dened XRD pattern of Co,Al-MCM-41 in
Fig. 1 is typical of MCM-41 as described by Kresge et al.
[2]. All the four XRD reections (1 0 0), (1 1 0), (2 0 0) and
(2 1 0) are well resolved and can be indexed to a hexagonal
lattice. The (1 1 0) and (2 0 0) reections for Co,Al-MCM-41
(20) are somewhat less well resolved. Upon calcination
of the as-synthesized materials, the XRD patterns of
Co,Al-MCM-41 become better resolved and the intensity
of the XRD patterns increased signicantly as a result of
removal of the intercalated organic molecules (template).
The unit cell parameters of Co,Al-MCM-41 are given in
Table 1. It has been reported that the unit cell parameter
increases with increasing metal content [19,20]. However,
we observe that unit cell parameters of all Co,Al-MCM-41
samples, though similar to that of Si-MCM-41, are more
than that of Al-MCM-41 (23). The intensity of (1 0 0) peak
decreases with increasing Co and Al content. This may be
due to the formation of non-framework metal oxide species
inside the mesopores upon calcination. Further, the broad-
ening of XRD peaks is increased with increasing Co and
Al content, suggesting a decrease in the structural integrity.
Table 1
Synthesis and physicochemical parameters of pure siliceous MCM-41 and Co,Al-MCM-41
Catalyst n
Si
/n
Al
n
Si
/n
Co
a
o
(nm) A
BET
(m
2
/g) d
p
, BJH (nm) Pore volume (cm
3
/g)
Gel Product Gel Product
Si-MCM-41 3.96 1273 2.40 0.85
Al-MCM-41 (23) 23 22 3.63 1210 2.43 0.80
Co,Al-MCM-41 (80) 160 142 160 145 3.90 1147 2.49 0.77
Co,Al-MCM-41 (50) 100 104 100 93 3.95 1099 2.49 0.74
Co,Al-MCM-41 (20) 40 39 40 38 3.91 1015 2.44 0.66
Fig. 1. XRD powder patterns of calcined MCM-41 and Co,Al-MCM-41
with different n
Si
/(n
Co
+ n
Al
) ratios: (a) MCM-41; (b) Co,Al-MCM-41
(80); (c) Co,Al-MCM-41 (50); (d) Co,Al-MCM-41 (20).
Fig. 2. TGA and DTG analysis curves for as-synthesized purely siliceous
MCM-41 and Co,Al-MCM-41 with different n
Si
/(n
Co
+ n
Al
) ratios ()
MCM-41; (. . . ) Co,Al-MCM-41 (80); ( ) Co,Al-MCM-41 (50); ( )
Co,Al-MCM-41 (20).
Fig. 2 shows TGA and DTG curves obtained for the
as-synthesized pure siliceous sample and Co,Al-MCM-41
samples with different n
Si
/(n
Co
+ n
Al
) ratios. There are
four weight losses that appear in the TGA curve. The rst
weight loss between 20 and 120

C is attributed to the re-


lease of physisorbed water molecules, it is very low for
siliceous MCM-41 (2.7%) compared to the corresponding
metal substituted analogue (45%). This indicates that the
hydrophilicity increases with increase of metal incorpora-
tion in the framework. The second and third weight losses,
142 M. Karthik et al. / Applied Catalysis A: General 268 (2004) 139149
which are centered around 210 and 265

C, respectively, are
attributed to the decomposition of loosely bound organic
template within the framework. The fourth weight loss is
centered around 390

C, it increases as the n
Si
/(n
Co
+ n
Al
)
ratio decreases and develops at the expense of the second
and third weight loss centered at 210 and 265

C, respec-
tively. This could be attributed to the strongly bonded tem-
plate cation with the metallosilicate framework. As more
aluminum and cobalt atoms are incorporated into the frame-
work, more of occluded template molecule is protonated so
as to balance the resulting negative charge. This indicates
that most of the aluminum and cobalt atoms are incorporated
inside the framework of the materials.
The n
Si
/n
Al
and n
Si
/n
Co
molar ratios of all the samples
under investigation are summarized in Table 1. It can be seen
from the table that incorporation of Si, Co and Al into the
solid is in close agreement with the input gel composition.
In the case of Co,Al-MCM-41 (80), the Co and Al content in
the solid phase is higher than that of the input gel, suggesting
preferential incorporation of both metal atoms compared to
silicon. Similar observations were previously reported in the
case of zeolites [21].
The nitrogen adsorption isotherms of purely siliceous
MCM-41 and Co,Al-MCM-41 samples are shown in Fig. 3.
All the samples exhibit isotherms of type 1V of IUPAC
classication, featuring a narrow step due to the capil-
lary condensation of N
2
in the primary mesopores. The
samples containing Co and Al possess an often observed
type-H4 hysteresis loop at p/p
o
between 0.5 and 1. The
assignment of hysteresis is made according to de Boers
classication. Such a hysteresis loop has already been ob-
served in aluminum containing MCM-41 materials and is
attributed to the capillary condensation of nitrogen within
interparticles and/or some impurity phases generated dur-
ing synthesis [22]. The incorporation of Co and Al into
the walls of MCM-41 has a signicant effect on the spe-
cic surface area and specic pore volume of the materials
Fig. 3. Nitrogen adsorption isotherm (adsorption: closed symbols; des-
orption: open symbols) of Co,Al-MCM-41 with different n
Si
/(n
Co
+n
Al
)
ratios: () Co,Al-MCM-41 (80); () Co,Al-MCM-41 (50); ()
Co,Al-MCM-41 (20).
(Table 1). With increasing metal content, the pore volume
is reduced from 0.85 to 0.66 cm
3
/g and the specic surface
area declines from 1270 to 1015 m
2
/g. This is attributed to
a slight reduction in the structural integrity of the samples
with increasing metal content. A similar behavior has also
been observed in monometal-substituted MCM-41 materi-
als [23,24]. Moreover, it is interesting to note that the pore
diameter of Co,Al-MCM-41 samples is larger than the pure
silica MCM-41. As the unit cell parameter is almost identi-
cal for both Si-MCM-41 and Co,Al-MCM-41 samples, the
increase in the pore diameter of Co,Al-MCM-41 samples
indicates the reduction in wall thickness. However, the pore
diameter of Co,Al-MCM-41 (20) is smaller than those of
Co,Al-MCM-41 (50) and (80). This could be due to the
formation of more metal oxide clusters in the mesopores of
Co,Al-MCM-41 (20) than in (50) and (80).
The coordination geometry of cobalt incorporated in
Al-MCM-41 is indicated by the UV-Vis absorption spec-
tra. Fig. 4 shows the absorption spectra of calcined
Co,Al-MCM-41 samples. It exhibits a strong absorption
around the 15 50020 000 cm
1
region. This absorption
consists of three components, with maxima at 19 685,
17 241 and 15 384 cm
1
. These bands can be assigned to
4
A
2
(F)
4
T
1
(P) transition of the tetrahedrally coordi-
nated divalent cobalt species [25]. A broad band between
20 800 and 22 000 cm
1
centered at 21 505 cm
1
is also
observed for all the samples. The intensity of this band
increases monotonically with increasing Co content. This
band could be assigned to the strong bonding of oxygen
ligands to Co
2+
ions. A similar band has also been reported
in the zeolites with low Si/Al ratio [26]. However, the de-
tailed geometry of the Co
2+
ion is not clearly known [27].
Moreover, a broad band in the UV region between 34 400
and 45 500 cm
1
centered at 40 816 cm
1
is also observed
for all samples. This has been assigned to a low-energy
charge transfer between the oxygen ligands and central
Co
2+
ion in tetrahedral symmetry [28] or to self-absorption
of the molecular sieves [26,27]. All these results reveal that
Fig. 4. UV-Vis spectra of calcined Co,Al-MCM-41 with various
n
Si
/(n
Co
+n
Al
) ratios: (a) Co,Al-MCM-41 (20); (b) Co,Al-MCM-41 (50);
(c) Co,Al-MCM-41 (80).
M. Karthik et al. / Applied Catalysis A: General 268 (2004) 139149 143
Fig. 5. ESR spectra of calcined Co,Al-MCM-41 with various
n
Si
/(n
Co
+n
Al
) ratios: (a) Co,Al-MCM-41 (20); (b) Co,Al-MCM-41 (50);
(c) Co,Al-MCM-41 (80).
the majority of the cobalt ions in the Co,Al-MCM-41 sam-
ples occupy framework positions in the surface layer of the
Co,Al-MCM-41 channel walls.
ESR spectroscopy is a sensitive technique to study
the coordination of high spin Co
2+
ions in molecular
sieves. The X-band ESR spectrum at 267

C of calcined
Co,Al-MCM-41samples synthesized with different n
Si
/n
Co
ratios in the synthesis gel is shown in Fig. 5. The spectrum
shows two major components, at g = 5.4 and 2.05. With
increasing n
Si
/n
Co
ratio from 160 to 40, the correspond-
ing spectra show ESR signal increases in intensity. The
ESR signal at g = 5.4 and 2.05 are attributed to Co
2+
in
an elongated tetrahedral environment [29]. Similar spectra
were also reported in cobalt-substituted aluminophosphate
molecular sieves [30,31]. It is important to note that the in-
tensity of ESR signal in the calcined material is higher than
in the as-synthesized material. The increase in the intensity
of ESR signal in the calcined materials indicates that there
is no oxidation of Co
2+
to Co
3+
during calcination [31].
Fig. 6. DRIFT spectra of pyridine adsorption region of mesoporous materials: (a) Co,Al-MCM-41 (20); (b) Al-MCM-41 (23); (c) Co,Al-MCM-41 (50);
(d) Co,Al-MCM-41 (80); (L: Lewis acid; B: Bronsted acid; L +B: Lewis and Bronsted acid sites).
Table 2
Bronsted and Lewis acidity values for Al-MCM-41 and Co,Al-MCM-41
Catalyst Bronsted (B) acid
site (mmol/g)
Lewis (L) acid
site (mmol/g)
B/L acid
site ratio
Al-MCM-41 (23) 0.114 0.121 0.942
Co,Al-MCM-41 (20) 0.190 0.284 0.669
Co,Al-MCM-41 (50) 0.117 0.107 1.093
Co,Al-MCM-41 (80) 0.057 0.026 2.192
The in situ DRIFT spectra of pyridine adsorbed on meso-
porous Al-MCM-41 and Co,Al-MCM-41 molecular sieves
are shown in Fig. 6. The acidity of the catalysts was cal-
culated using extinction coefcients of the bands of Bron-
sted and Lewis acid sites adsorbed pyridine [18] and the
acidity values are given in Table 2. The DRIFT spectra of
chemisorbed pyridine have shown that Co,Al-MCM-41 and
Al-MCM-41 contain both Bronsted and Lewis acid sites.
The existence of Bronsted acid sites in the samples is clearly
shown by the bands at 1545 and 1636 cm
1
due to ring vibra-
tions of pyridine bound to Bronsted acid sites [32,33]. The
bands at 1451, 1493 and 1617 cm
1
are assigned to pyridine
associated with Lewis acid sites [34]. The band at 1493 cm
1
is attributed to pyridine chemisorbed on both Bronsted and
Lewis acid sites [32,35]. The band at 1400 cm
1
is attributed
to pyridine hydrogen bonded to defective SiOH groups
[17,36]. The peak at 1230 cm
1
is assigned to the asymmet-
ric TOT vibration of the framework [37,38]. Comparison
of spectra reveals that Co,Al-MCM-41 (20) has a greater
amount of acid sites (Bronsted and Lewis) than Al-MCM-41
(23) as the corresponding bands in the former are more in-
tense than those in the latter. The acid sites are found to be
same in both Co,Al-MCM-41 (50) and Al-MCM-41 (23),
whereas Co,Al-MCM-41 (80) has a smaller amount of acid
sites than either Co,Al-MCM-41 (50) or Co,Al-MCM-41
(20). The intensity of the bands due to Lewis acid sites in
Co,Al-MCM-41 (20) is more than in either Co,Al-MCM-41
144 M. Karthik et al. / Applied Catalysis A: General 268 (2004) 139149
(50) or Co,Al-MCM-41 (80). This is ascribed to the forma-
tion of more cobalt oxide in Co,Al-MCM-41 (20) upon cal-
cination. It is obvious that, with increasing cobalt content
in Al-MCM-41, the number of surface acid sites increases.
This study reveals that the acidity of the catalysts decreases
in the order Co,Al-MCM-41 (20) > Al-MCM-41 (23)
Co,Al-MCM-41 (50) > Co,Al-MCM-41 (80). Hence, it is
concluded that the isomorphic substitution of Co
2+
into the
framework of mesoporous Al-MCM-41 creates more acid
sites (Bronsted and Lewis) and enhances the catalytic activ-
ity of Co,Al-MCM-41.
3.2. Catalytic activity
Alkylation of phenol with isobutanol was studied over
Co,Al-MCM-41 (20, 50 and 80) and Al-MCM-41 (23) at
200, 250, 300, 350, 400, 450 and 500

C by co-feeding
phenol and isobutanol in the feed ratio of 1:3 and feed
rate of 3 ml/h. The plot of conversion versus temperature
is shown in Fig. 7. There is a nearly linear increase of
phenol conversion with temperature over all the catalysts.
It can be seen from the gure that the butylation activity
in Co,Al-MCM-41 (20) is more than in other catalysts.
Al-MCM-41 (23) possesses less density of acid sites than
Co,Al-MCM-41 (20), as isomorphic substitution of Al
3+
by Co
2+
would provide two Bronsted acid sites, and hence
the former shows less activity than the latter. As the reac-
tion requires formation of isobutyl cation in order to have
electrophilic attack on either the chemisorbed phenol or the
free phenol, the density of Bronsted acid sites, which are
the active sites in this reaction, is important. As the density
of acid sites in Co,Al-MCM-41 decreases in the following
order: Co,Al-MCM-41 (20) > Co,Al-MCM-41 (50) >
Co,Al-MCM-41 (80), the conversion also follows the same
trend as that shown in Fig. 7. The density of acid sites are
Fig. 7. Effect of temperature on conversion of phenol. Phenol:
isobutanol1:3; feed rate3 ml/h; catalyst weight0.5 g. Conversion
of phenol() Co,Al-MCM-41 (20); () Co,Al-MCM-41 (50); ()
Co,Al-MCM-41 (80); () Al-MCM-41 (23).
in good agreement with DRIFT (pyridine adsorption) mea-
surements. The activity of Co,Al-MCM-41 (80) is expected
to be less than Al-MCM-41 (23), but the results indicate the
reverse trend. Hence, in addition to acidity, the associated
hydrophilic and hydrophobic properties of the catalysts are
also important factors to account for the difference in the
activity of the catalysts. Though Al-MCM-41 (23) pos-
sesses more density of acid sites, it is less hydrophobic than
Co,Al-MCM-41 (80). Hence, the adsorption of hydropho-
bic isobutanol on the active sites is to be less and for this
reason the former gives less conversion than the latter. This
trend is similar to the previously reported hydrophilic and
hydrophobic properties of molecular sieves in acetalyzation
[35] and isopropylation of m-cresol [37]. In a similar way,
Co,Al-MCM-41 (20) can be expected to give less conversion
than even Al-MCM-41 (23), but it gives more conversion.
There might be yet another route to yield isobutyl carbo-
nium ion in addition to the above Bronsted acid-assisted
route. The formation of non-framework cobalt oxide in the
channels of MCM-41 may provide an alternate route to yield
tert-butyl cations as shown in reaction (Scheme 1). Such
dissociative adsorption of phenol on the Lewis acid sites,
so that the released proton can rest on the adjacent basic
site, has already been proposed by many workers [39,40].
Isobutyl alcohol should be chemisorbed on the Bronsted
site in order to yield isobutyl cation. This cation without or
with rearrangement is to react with the adjacent phenoxide
to yield O-or C-alkylated product, as shown in the reac-
tion (Scheme 1(a)(c)). Based on the conversion, the order
of activity of Co,Al-MCM-41 catalysts is Co,Al-MCM-41
(20) > Co,Al-MCM-41 (50) > Co,Al-MCM-41 (80),
which is also the order of the amount of non-framework
cobalt oxide in the pores of MCM-41. These particles are
also expected to be of nanodimensions close to 1 nm, so that
the reactants and products can very well diffuse in and out of
the pores of the catalysts without any diffusional constraints.
In general, alkylation of phenol is a reaction sensitive to
the acidbase properties of the catalysts employed. It has
been observed that O-alkylation of phenol is favored by
strong acid sites [41], while C-alkylation is favored by weak
acidic (or) strong basic sites [42]. Tleimaat-Manzaliji et al.
[43] and Marczewski et al. [44] have claimed that weak
acids sites favor C-alkylation, while Velu and Swamy [45]
reported that C-alkylation occurred due to the presence of
higher acidity. The formation of O-alkylated products de-
pends on the intrinsic properties of the alcohol and acidbase
properties of the catalysts [46]. Based on the above obser-
vations, we have also tried to explain the formation of prod-
ucts as a function of acidity and activity of the catalysts for
the t-butylation of phenol.
Table 3 shows the selectivity of the products obtained over
the catalysts in the temperature range 350500

C. The prod-
ucts were identied by GC and GC-MS. The main products
of the catalytic butylation of phenol were O-tert butyl phe-
nol (OTBP), 2-tert-butyl phenol (2TBP), 4-tert-butyl phe-
nol (4TBP), O-butenyl phenol (OBP) and 2-butenyl phenol
M. Karthik et al. / Applied Catalysis A: General 268 (2004) 139149 145
Scheme 1.
Table 3
Effect of temperature on t-butylation of phenol over Al-MCM-41 and Co,Al-MCM-41
Catalyst
a
Temperature (

C) Conversion (wt.%) Selectivity of products (wt.%)


OTBP 2TBP 4TBP OBP 2BP Others
Al-MCM-41 (23) 350 13.54 31.21 50.31 10.03 7.21
400 19.43 26.60 53.43 10.47 9.40
450 24.32 25.43 47.21 13.04 14.23
500 37.52 19.57 44.72 10.10 25.01
Co,Al-MCM-41 (20) 350 32.71 43.24 14.26 12.10 30.20
400 44.34 37.65 13.40 13.43 20.24 9.10 6.06
450 50.24 30.72 14.01 12.12 18.31 10.03 14.08
500 48.98 17.21 22.09 11.01 16.03 13.12 20.21
Co,Al-MCM-41 (50) 350 21.22 56.81 14.06 5.62 22.98
400 34.14 42.02 18.79 8.74 20.18 7.68 2.24
450 42.08 31.41 20.12 9.55 22.10 8.01 8.21
500 46.19 20.67 24.22 10.52 18.24 10.21 16.14
Co,Al-MCM-41 (80) 350 19.94 80.24 6.92 12.14
400 26.82 69.85 11.58 18.26
450 34.28 52.26 12.01 6.94 19.24 2.10 7.12
500 43.34 39.52 13.44 10.22 20.52 4.84 10.75
Phenol:isobutanol feed ratio: 1:3; feed rate: 3 ml/h.
a
Catalyst weight: 0.5 g.
146 M. Karthik et al. / Applied Catalysis A: General 268 (2004) 139149
Scheme 2.
(2BP). Small amounts of other products were also detected,
which were unidentied polymerized products. The forma-
tion of products with t-butyl group evidently proves rapid
isomerization of isobutyl cation into t-butyl cation on the
catalyst surface. The formation of OBP and 2BP has also
been reported by earlier workers [47]. The proposed route
as shown in Scheme 2, which is a little different from that
given by previous workers. But isobutyraldehyde was not
observed in present study. It is visualized that, as and when
isobutyraldehyde is formed, it may react immediately with
phenol to yield the products. Selectivity to ring alkylation
over Al-MCM-41 (23) is found to be more than that to
O-alkylation. Generally, the electrophilic substitution may
take place in the ortho and para positions of the phenyl ring
[46,48]. As there is more probability for ortho position than
para position for electrophilic reactions between phenol in
the vapor phase and tert-butyl cation adsorbed on the cat-
alyst surface, more selectivity to 2TBP is observed. More-
over, the presence of phenolic group kinetically favors ortho
alkylation [49]. It is clearly seen from Table 3 that OTBP is
one of the main products of the reaction between phenol and
isobutanol over Co,Al-MCM-41 catalysts. Co,Al-MCM-41
(80) gives about 80% selectivity to OTBP. This suggests that
Scheme 1(a) is the most probable one for the high selectiv-
ity of OTBP. When phenol is dissociatively adsorbed on the
Lewis acid site, the plane of the aromatic ring is to be in-
clined towards the catalyst surface. Hence, if the ring is ro-
tated about the cobalt-phenol oxygen bond axis and brought
M. Karthik et al. / Applied Catalysis A: General 268 (2004) 139149 147
close to the adjacently adsorbed tert-butyl cation, there could
be much steric hindrance for para position. Hence, in the
Lewis acid-adsorbed state of phenol, the phenoxide is to re-
act nucleophilically with tert-butyl cation to yield OTBP.
The selectivity of OTBP over Co,Al-MCM-41 (80) is higher
than that over Co,Al-MCM-41 (50) and Co,Al-MCM-41
(20) due to lower density of acid sites. However, a decrease
in the selectivity of OTBP is observed over all the catalysts
with an increase in temperature, as it is thermally unstable.
It can be seen from Table 3 that the selectivity to 2TBP
increases with increase in temperature over all the Co,Al-
MCM-41 catalysts. But the selectivity over Co,Al-MCM-41
(80) is much less than over other catalysts, this may be due
to the much lower density of Bronsted acid sites available
on Co,Al-MCM-41 (80). Similarly, Co,Al-MCM-41 (80)
with a lower density of Bronsted acid sites gives less yield
of 4TBP, whereas Co,Al-MCM-41 (50) and Co,Al-MCM-
41 (20) give more selectivity due to a greater density of acid
sites. The selectivity to OBP increases with decrease in the
n
Si
/(n
Co
+n
Al
) ratio of the catalysts. But this product is not
observed over Al-MCM-41 (23). Hence, the dehydrogena-
tion of isobutanol to isobutyraldehyde and its subsequent
reaction with free phenol to give OBP and/or 2BP depends
on the amount of non-framework cobalt oxide inside the
pores of the catalysts (Scheme 2). Formation of this prod-
uct only over Co,Al-MCM-41 and not over Al-MCM-41
evidently proves the reaction is Lewis-acid dependent; the
former catalysts have been proved to have non-framework
cobalt oxide. It has also been reported in the literature that
the modes of adsorption of phenol on catalytic surfaces also
play an important role in the formation of O-butenyl phenol
[50].
Table 4
Effect of feed rate on t-butylation of phenol over Al-MCM-41 and Co,Al-MCM-41
Catalyst
a
Feed rate (ml/h) Conversion (wt.%) Selectivity of products (wt.%)
OTBP 2TBP 4TBP OBP 2BP Others
Al-MCM-41 (23) 1 9.04 37.21 59.63 3.16
2 15.21 31.91 53.74 9.82 4.21
3 19.43 26.60 53.43 10.47 9.40
4 12.86 20.58 47.98 19.62 11.32
Co,Al-MCM-41 (20) 1 34.42 43.03 10.08 10.14 24.19 12.17
2 40.18 40.08 11.21 12.10 22.84 10.72 2.64
3 44.30 37.65 13.43 13.41 20.24 9.10 6.06
4 32.81 30.82 12.10 21.16 18.89 10.21 6.12
Co,Al-MCM-41 (50) 1 28.74 47.16 12.21 6.18 22.28 11.87
2 30.12 44.12 15.81 7.91 21.88 10.21
3 34.14 42.02 18.79 8.74 20.18 7.68 2.24
4 23.16 35.72 17.81 17.10 20.02 7.90 1.04
Co,Al-MCM-41 (80) 1 18.42 71.21 7.51 19.24 2.04
2 22.18 70.77 7.83 18.42 2.98
3 26.82 69.85 11.58 18.26
4 15.46 69.62 12.07 18.21
Phenol:isobutanol feed ratio: 1:3; temperature: 400

C.
a
Catalyst weight: 0.5 g.
Fig. 8. Effect of feed ratio and feed rate on conversion of phenol over
Co,Al-MCM-41 (20). Catalyst weight0.5 g; temperature400

C. Con-
version of phenol in the feed ratio (phenol:isobutanol)() 1:2; ()
1:3; () 1:4.
The effects of feed ratio and feed rate on conversion were
studied at 400

C over Co,Al-MCM-41 (20). The feed ratios


were kept at 1:2, 1:3 and 1:4. The variation in conversion
with feed rate for each feed ratio is illustrated in Fig. 8. In-
crease in conversion with increase in feed rate up to 3 ml/h
for feed ratios 1:2 and 1:3 may be attributed to the force ex-
erted on the reactants to move closer to the channel surface,
which could facilitate their chemisorption. Above this feed
rate (3 ml/h) the conversion decreases, which may be due to
the fast diffusion of the reactants. The feed ratio 1:4 gives a
gradual decrease in conversion with increase in feed rate.
Table 4 shows the effect of feed rate on phenol con-
version and product selectivity over Al-MCM-41 and
148 M. Karthik et al. / Applied Catalysis A: General 268 (2004) 139149
Table 5
Effect of time on stream on t-butylation of phenol over Al-MCM-41 and Co,Al-MCM-41
Catalyst
a
Time on stream (h) Conversion (wt.%) Selectivity of products (wt.%)
OTBP 2TBP 4TBP OBP 2BP Others
Al-MCM-41 (23) 1 19.43 26.60 53.43 10.47 9.40
2 15.82 24.26 47.21 13.13 15.39
3 10.14 20.14 44.80 15.81 22.51
4 8.26 15.82 40.28 18.20 25.64
5 7.12 13.21 37.12 19.70 29.96
Co,Al-MCM-41 (20) 1 44.34 37.65 13.40 13.43 20.24 9.10 6.06
2 39.45 33.12 12.22 16.08 18.22 10.12 10.21
3 30.47 29.34 11.36 16.89 17.81 13.24 11.36
4 25.89 22.83 9.98 24.17 15.88 15.16 11.98
5 20.84 19.16 9.02 29.54 14.02 15.82 12.44
Co,Al-MCM-41 (50) 1 34.14 42.02 18.77 8.74 20.18 7.68 2.24
2 34.10 40.16 17.24 13.14 18.36 7.92 3.18
3 31.68 38.44 15.98 15.09 17.12 8.14 5.23
4 28.36 35.28 13.22 17.83 16.82 8.97 7.88
5 23.65 31.17 12.81 22.70 15.72 9.39 8.21
Co,Al-MCM-41 (80) 1 26.82 69.85 11.58 18.26
2 26.76 69.81 11.02 18.21
3 26.05 69.68 10.96 18.20
4 25.86 68.94 10.54 18.00 2.52
5 25.34 68.51 10.06 17.98 3.00
Phenol:isobutanol feed ratio: 1:3; feed rate: 3 ml/h; temperature: 400

C.
a
Catalyst weight: 0.5 g.
Co,Al-MCM-41. The reaction was carried out at different
feed rates at 400

C with feed ratio 1:3. An appreciable


increase in phenol conversion is observed up to the feed
rate of 3 ml/h, above which the conversion decreases due
to the fast diffusion of the reactants through the catalyst
[51,52]. On the other hand, the low conversion at low feed
rate could be attributed to coke formation due to more con-
tact time of the catalyst [51]. It is observed that increase in
the feed rate decreases the selectivity of OTBP, while the
4TBP selectivity increases. A similar increase in selectivity
of 4TBP has been reported in the tert-butylation of phenol
over H-AlMCM-41 [51]. The selectivity values of OBP and
2BP are observed to remain the same with increase in the
feed rate.
The time on stream study was conducted to assess its
effect on conversion and selectivity of the products. The
temperature was maintained at 400

C and the feed ra-


tio was set at 1:3 with a feed rate of 3 ml/h. Al-MCM-
41 (23), Co,Al-MCM-41 (20), Co,Al-MCM-41 (50) and
Co,Al-MCM-41 (80) were examined; the results are pre-
sented in Table 5. There is a decrease in conversion over
Al-MCM-41 (23), Co,Al-MCM-41 (20) and Co,Al-MCM-
41 (50) with increase in the stream. This is attributed to
the formation of coke and subsequent blocking of active
sites. Co,Al-MCM-41 (20) and Al-MCM-41 (23) are more
rapidly deactivated than Co,Al-MCM-41 (50) due to more
acid sites [53,54]. Furthermore, the product selectivity also
changes as a consequence of deactivation. With increase in
stream, the selectivity of 4TBP and 2BP gradually increases,
which is accompanied by decreases in the selectivity of
OTBP, 2TBP and OBP. The increase in the selectivity of
4TBP with increase in stream has already been reported in
[55,56]. The increase in the selectivity of 2BP with increase
in stream is attributed to enhancement in the transport of
isobutanol to cobalt oxide particles of specic dimension.
It is assumed that only such particles encourage dehydro-
genation of isobutanol to isobutyraldehyde, which results
in high selectivity of 2BP. The deactivation rate of the cat-
alysts is in the order Co,Al-MCM-41 (20) > Al-MCM-41
(23) > Co,Al-MCM-41 (50) > Co,Al-MCM-41 (80). The
catalytic activity of Co,Al-MCM-41 (80) is not affected
even at the end of 5 h, due to a lower density of acid sites.
4. Conclusion
Cobalt containing mesoporous aluminoslicate molec-
ular sieves that contain tetrahedrally coordinated cobalt
atoms has been hydrothermally synthesized. A series of
Co,Al-MCM-41 with n
Si
/(n
Co
+ n
Al
) ratios from 20 to 80
was synthesized successfully and characterized by physic-
ochemical methods such as XRD, nitrogen adsorption,
UV-Vis DRS, ESR, AAS, DRIFT (pyridine adsorption) and
TGA/DTG. UV-Vis DRS and ESR studies reveal that cobalt
occurs in Co,Al-MCM-41 is highly symmetrical tetrahedral
coordination and that some of the cobalt atoms are trans-
formed into the oxide form when the n
Si
/(n
Co
+n
Al
) ratio is
increased to 20. The adsorbed pyridine measured by FT-IR
M. Karthik et al. / Applied Catalysis A: General 268 (2004) 139149 149
indicates the presence of both Bronsted and Lewis acid sties.
Alkylation of phenol with isobutanol over Al-MCM-41 (23)
and Co,Al-MCM-41 catalysts shows an increase in conver-
sion with increase in temperature. Isomerization of isobutyl
cation to tert-butyl cation is very evident in this study.
Co,Al-MCM-41 (20) is more active than other catalysts.
The activity of Al-MCM-41 and Co,Al-MCM-41 catalysts
follows the order: Co,Al-MCM-41 (20) > Co,Al-MCM-41
(50) > Co,Al-MCM-41 (80) > Al-MCM-41 (23). The
selectivity of O-tert-butyl phenol is found to be more than
that of other products. Formation of non-framework cobalt
oxide, which is evident in the spectral studies, is expected
to give more selectivity to O-butenyl phenol. The interest-
ing observation with Co,Al-MCM-41 (80) is its uniform
activity with increase of time.
Acknowledgements
The authors gratefully acknowledge the nancial support
from BRNS-DAE, Mumbai (Sanction No. 98/37/23/BRNS/
Cell/720), for this research work. The authors M.K. and
A.V. are thankful to BRNS-DAE, Mumbai, for the research
fellowship.
References
[1] J.S. Beck, US Patent 5,057,296 (1991).
[2] C.T. Kresge, M.E. Leonowicz, W.J. Roth, J.C. Vartuli, J.S. Beck,
Nature 359 (1992) 710.
[3] X.S. Zhao, G.Q. Lu, G.J. Millar, Ind. Eng. Chem. Res. 35 (1996)
2075.
[4] L. Mercier, T.J. Pinnavaia, Adv. Mater. 9 (1997) 500.
[5] A. Corma, Chem. Rev. 97 (1997) 2373.
[6] J.M. Thomas, Faraday Discuss. 105 (1996) 1.
[7] H. Knozinger, S. Huber, J. Chem. Soc., Faraday Trans. 94 (1998)
2047.
[8] P.T. Tanev, M. Chibwe, T.J. Pinnavaia, Nature 368 (1994) 321.
[9] A. Tuel, Micropor. Mesopor. Mater. 27 (1999) 151.
[10] S. Lim, G.L. Haller, Appl. Catal. A: Gen. 188 (1999) 277.
[11] D.B. Akolekar, Catal. Lett. 28 (1994) 249.
[12] B. Elvers, S. Hawkins, G. Schulz (Eds.), Ullmanns Encylopedia of
Industrial Chemistry, 5th ed., A.19, (1991) 325341.
[13] V.A. Koshchii, B.Ya. Kozlikovskii, A.A.Zh. Matyusha, Org. Khim.
24 (7) (1989) 1508.
[14] A.R. Rajadhyaksha, D.D. Chaudhari, Ind. Eng. Chem. Res. 26 (7)
(1987) 1276.
[15] H.C. Kuizhang, H. Zhang, S. Xiang, S. Liu, D. Xu, H. Li, Appl.
Catal. A: Gen. 166 (1998) 89.
[16] M. Kruk, M. Jaroniec, A. Sayari, Langmuir 13 (1997) 6267.
[17] E.P. Parry, J. Catal. 2 (1963) 371.
[18] C.A. Emeis, J. Catal. 141 (1993) 347.
[19] C.F. Cheng, J. Klinowski, J. Chem. Soc., Faraday Trans. 92 (1996)
289.
[20] N. Ulagappan, C.N.R. Rao, J. Chem. Soc. Chem. Commun. (1996)
1047.
[21] E.G. Derouane, S. Detremmerie, Z. Gabelica, N. Blom, Appl. Catal.
1 (1981) 201.
[22] Q. Huo, R. Leon, P.M. Petroff, G.D. Stucky, Science 268 (1995)
1324.
[23] M. Kruk, M. Jaronice, A. Sayari, Langmuir 15 (1999) 5683.
[24] Y.W. Chen, Y.H. Lu, Ind. Eng. Chem. Res. 38 (1999) 1893.
[25] J. poner, J. ejka, J. D`deek, B. Wichterlov, Micropor. Mesopor.
Mater. 37 (2000) 117.
[26] J. D`deek, D. Kaucky, B. Wichterlova, Micropor. Mesopor. Mater.
35/36 (2000) 483.
[27] D. Kaucky, J. D`deek, B. Wichterlova, Micropor. Mesopor. Mater.
31 (1999) 75.
[28] R.A. Schoonheydt, in: F. Delannay (Ed.), Diffuse Reectance Spec-
troscopy Characterization of Catalysis, Marcel Dekker, New York,
1984, p. 220.
[29] S. Thomson, V. Luca, R. Howe, Phys. Chem. Chem. Phys. 1 (1999)
615.
[30] B.M. Weckhuysen, A.A. Verberckmoes, M.G. Uytterhoeven, F.E.
Mabbs, D. Collison, E.D. Boer, R.A. Schoonheydt, J. Phys. Chem.
B 3742 (2000) 104.
[31] V. Kurshev, L. Kevan, D. Parillo, G.T. Kokotailo, R.J. Gorte, J. Phys.
Chem. 98 (1994) 10160.
[32] A. Corma, V. Fornes, M.T. Navarro, J. Perez-Pariente, J. Catal. 148
(1994) 569.
[33] B. Chakraborty, B. Viswanathan, Catal. Today 49 (1999) 253.
[34] M.L. Occelli, S. Biz, A. Aurous, G.J. Ray, Micropor. Mesopor. Mater.
26 (1998) 193.
[35] M.J. Climent, A. Corma, S. Iborra, S. Miquel, J. Primo, F. Rey, J.
Catal. 76 (1999) 783.
[36] A. Corma, Chem. Rev. 95 (1995) 559.
[37] V. Umamaheswari, M. Palanichamy, V. Murugesan, J. Catal. 210
(2002) 367.
[38] S. Biz, M.L. Occelli, Catal. Rev. Sci. Eng. 40 (3) (1998) 329.
[39] M. Bowker, R.W. Pelts, K.C. Waugh, J. Catal. 99 (1986) 33.
[40] T. Yashima, H. Suzuki, N.J. Hara, J. Catal. 33 (1974) 486.
[41] V.V. Rao, K.V.R. Chary, V. Durga Kumari, S. Narayanan, Appl.
Catal. 61 (1990) 89.
[42] V. Durga Kumari, G. Sreekanth, S. Narayanan, Res. Chem. Intermed.
14 (1990) 223.
[43] R. Tleimaat-Manzaliji, D. Bianchi, G.M. Pajonk, Appl. Catal. A:
Gen. 101 (1993) 339.
[44] M. Marczewski, J.P. Bodibo, G. Perot, M. Guisnet, J. Mol. Catal.
50 (1989) 211.
[45] S. Velu, C.S. Swamy, Appl. Catal. A: Gen. 145 (1996) 141.
[46] E. Dumitriu, V. Hulea, J. Catal. 218 (2003) 249.
[47] A.H. Padmasri, V. Durga Kumari, P. Kanta Rao, Recent
Trends in Catalysis, Narosa Publishing House, New Delhi, 1999,
pp. 189196.
[48] R.T. Parton, J.M. Jacobs, D.R. Hoybrechts, P.A. Jacobs, Stud. Surf.
Sci. Catal. 46 (1998) 163.
[49] A.J. Kolka, J.P. Napolitano, G.G. Elike, J. Org. Chem. 21 (1956) 712.
[50] A.H. Padmasri, V. Durga Kumari, P. Kanta Rao, Stud. Surf. Sci.
Catal. 113 (1998) 563.
[51] A. Sakthivel, S.K. Badamali, P. Selvam, Micropor. Mesopor. Mater.
39 (2000) 457.
[52] J.W. Yoo, C.W. Lee, H.C. Jeong, Y.K. Park, S.E. Park, Catal. Today
60 (2000) 255.
[53] V. Umamaheswari, M. Palanichamy, B. Arabindoo, V. Murugesan,
Indian J. Chem. 39A (2000) 1241.
[54] Z. Liu, P. Moreau, F. Fajula, Appl. Catal. A 159 (1997) 305.
[55] S.K. Badamali, A. Sakthivel, P. Selvam, Catal. Lett. 65 (2000)
153.
[56] A. Sakthivel, N. Saritha, P. Selvam, Catal. Lett. 72 (2001) 225.

Você também pode gostar