You are on page 1of 6

Materials Research Bulletin 70 (2015) 914919

Contents lists available at ScienceDirect

Materials Research Bulletin


journal homepage: www.elsevier.com/locate/matresbu

Characterizations and synergistic catalytic activity of bimetallic


Al-Fe-TUD-1
Vinju Vasudevan Srinivasana , Anand Ramanathanb,* , Rajamanickam Maheswaria,**,
Gaffar Imrana , Rajamanickam Rajalakshmia , Anbazhagan Nilamadanthaia
a
b

Department of Chemistry, Anna University, Chennai 600025, India


Center for Environmentally Benecial Catalysis, The University of Kansas, Lawrence, KS 66047, USA

A R T I C L E I N F O

A B S T R A C T

Article history:
Received 26 April 2015
Received in revised form 18 June 2015
Accepted 24 June 2015
Available online 29 June 2015

Amorphous 3D mesoporous silicate TUD-1 was incorporated with monometallic Al3+, Fe3+ and bimetallic
Al3+Fe3+ by direct one pot synthesis. The incorporation of these metal ions created both Lewis and
Brnsted acid sites. A synergy between Al and Fe sites in TUD-1 was noticed and is found responsible for
observed higher activity in the synthesis of 1,4-dihydropyridine and b-amino carbonyl compounds
through Hantzsch and Mannich reactions respectively.
2015 Elsevier Ltd. All rights reserved.

Keywords:
Amorphous materials
Solgel chemistry
Nuclear magnetic resonance (NMR)
X-ray diffraction
Transmission electron microscopy (TEM)
Infrared spectroscopy
Catalytic properties

1. Introduction
Solid acids such as zeolites are well known catalysts for
petroleum rening due to their well-dened number and nature of
acid sites [1,2]. However, for acid-catalyzed organic transformations involving bulkier substrates, incorporation of heteroatoms
such as Al was successfully carried out in MCM-type (pore size
23 nm) mesoporous molecular sieves [3,4]. In general, the
incorporation of Al in these mesoporous silicates generates both
Lewis and Brnsted acidity that can be tuned based on the loading
of Al. Further, for the synthesis of larger pore silicates (>5 nm, e.g.
SBA-type materials), the incorporation of Al was achieved by
adjusting the pH of the synthesis gel [5] under acidic conditions.
On the other hand, amorphous disordered mesoporous material
with relatively narrow pore size distribution, TUD-1, is prepared
under basic conditions. It was shown that with the nonsurfactantaggregated route for the mesopore formation of
TUD-1, a higher loading of Al (Si/Al = 4) can be achieved and its
Brnsted site is exploited as an ionic carrier [6]. We have shown
that Al-TUD-1 can be utilized as a FriedelCrafts alkylation catalyst

* Corresponding author. Fax: +1 785 864 6051.


** Corresponding author.
E-mail addresses: anand@ku.edu (A. Ramanathan),
maheswarianand@annauniv.edu (R. Maheswari).
http://dx.doi.org/10.1016/j.materresbull.2015.06.039
0025-5408/ 2015 Elsevier Ltd. All rights reserved.

both in liquid and gas phase [7]. Independently, Al-TUD-1 was


also shown to be an effective catalyst for degradation of high
density polyethylene (HDPE) [8] and for transformation of
bio-mass based substrates such as conversion of furfuryl alcohol
into ethyl levulinate and conversion of saccharides into furanic
aldehydes [9,10].
The acidity of Al-TUD-1 can further be tuned by incorporating
another metal ion such as Zr4+ and a synergy between Brnsted
and Lewis acid sites were demonstrated in AlZr-TUD-1 for Prins
cyclization of citronellal [11]. Similarly, the incorporation of Fe to
Al-MCM-41 leads to a higher catalytic activity in the t-butylation of
phenol compared to Al-MCM-41 [12]. The activity increase was
attributed to strengthening of the Brnsted acid sites (generated
by Al3+) by incorporation of Fe in tetrahedral co-ordination
in close proximity to the Brnsted acid site. Though bimetallic
Al-Fe-SBA-15 was prepared by direct synthesis, leaching of Fe
species that are not close to framework Al ions was observed
[13]. Moreover, similar catalyst (FeAl-SBA-15) is shown to be
active in benzene hydroxylation [14].
Encouraged by these results, we intent to synthesis and explore
incorporation of both Al and Fe into amorphous mesoporous
TUD-1 material and its activity was compared with monometallic
Al-TUD-1 and Fe-TUD-1. The synthesis of 1,4-dihydropyridines and
b-amino carbonyl compounds through Hantzsch and Mannich
reactions respectively studied by us earlier [15,16] was chosen as
probe reactions.

V.V. Srinivasan et al. / Materials Research Bulletin 70 (2015) 914919

915

2. Experimental

2.3. Catalytic activity

2.1. Synthesis of bimetallic Al-Fe-TUD-1

For Hantzsch reaction [15], benzaldehyde (1 mmol), ethyl


acetoacetate (2 mmol), ammonium acetate (1.2 mmol), ethanol
(4 mL) and 100 mg of catalyst (preheated at 200  C for 3 h) were
taken in a 25 mL two neck round bottom ask tted with reux
condenser. The ask was then immersed in an oil bath and the
reaction was carried out under reux conditions (typically 80  C for
3 h). After completion of reaction (monitored by TLC using hexane:
ethyl acetate 7:3), the reaction mixture was cooled down to room
temperature and poured into crushed ice with stirring. The crude
product was ltered and washed with distilled water followed by
treatment with brine solution and mixed with ethyl acetate to
separate the compound and dried over anhydrous Na2SO4. The
crude mixture was dissolved in hot ethanol to separate the solid
catalyst and the crude product was further puried by recrystallization from ethanol. The isolated pure compound was conrmed
by 1H NMR, 13C NMR and FT-IR (not shown) with the reported
literature.
For Mannich reaction [16], a mixture of aniline (1 mmol),
benzaldehyde (1 mmol), acetophenone (1 mmol) and catalyst
(0.1 g) in ethanol (5 mL) were added to a 15 mL glass vial and the
reaction mixture was stirred at room temperature for 10 h.
The progress of the reaction was monitored by thin-layer
chromatography (TLC). After reaction, the catalyst was ltered
and the ltrate was washed with aqueous NaCl followed by
distilled water and dried over sodium sulphate (Na2SO4).
After evaporation of the solvent, the crude yellow product
was further puried by recrystallization from ethanol. The
isolated pure compound was characterized and conrmed by
1
H NMR, 13C NMR and FT-IR (not shown) with the reported
literature.

Monometallic Al-TUD-1 and Fe-TUD-1 with (Si/M = 40, M = Al or


Fe) were synthesized according to previously reported literature
[7,17]. Al-Fe-TUD-1 materials with Si/(Al + Fe) molar ratio of
40 with Al/Fe ratio of 1 were synthesized by using triethanolamine
(TEA) as complexing agent and tetraethyl ammonium hydroxide
(TEAOH) as base in a one pot surfactant-free procedure. In a typical
synthesis, aluminum isopropoxide (0.18 g, 98%, Aldrich) and iron
(III) nitrate nanohydrate (0.36 g, 98%, Aldrich) were dissolved in a
1:1 mixture of isopropanol (2 mL) and absolute ethanol (2 mL) and
were added to tetraethyl orthosilicate (15.0 g, 98%, Aldrich). After
stirring for a few minutes, a mixture of TEA (10.5 g) and distilled
water (6 mL) was added, followed by addition of TEAOH (10.5 g,
35 wt% in H2O, Aldrich) under vigorous stirring. The clear gel
obtained after these steps was then aged at room temperature for
24 h and dried at 98  C for 24 h. Hydrothermal treatment was
carried out in a Teon-lined autoclave at 180  C for 8 h and nally
the organic moieties were removed by calcination in the presence
of air at 600  C with a temperature ramp of 1  C min 1.
2.2. Characterization
The powder X-ray diffraction patterns were recorded on a
Rigaku instrument with Cu-Ka (l = 1.54 ) in the 2u range of
1065 . Nitrogen adsorption and desorption isotherms were
measured at 77 K using a Quantachrome porosimeter (QuantosorbSI). Chemical compositions were analyzed by ICP-OES using a
PerkinElmer OES Optima 5300 DV spectrometer. Fourier-transform
infrared spectroscopy (FT-IR) was carried out on a Nicolet 5700
FT-IR instrument measured at 4 cm 1 resolutions. Samples were
mixed and ground with KBr followed by pressing into pellets for IR
measurement in the range of 4000400 cm 1. Diffuse reectance
UVvis spectra were recorded by a Shimadzu UV-2401 spectrophotometer with BaSO4 as reference. Temperature programmed
desorption (TPD) of ammonia measurements were performed
using the Micromeritics TPR/TPD 2900 equipped with a thermal
conductivity detector (TCD). For pyridine FTIR 100 mg of the
catalyst was dried at 200  C for 2 h to remove the adsorbed species.
Then the sample was wetted with pyridine and equilibrated for
12 h. The sample was then heated at 110  C in the atmospheric
pressure to remove the physisorbed pyridine and FTIR analysis was
carried out in the absorption mode.

3. Results and discussion


The low angle powder XRD patterns of Al-TUD-1, Fe-TUD-1 and
Al-Fe-TUD-1 are presented in Fig. 1a. A growing peak starting about
2 (2u) toward the low angle 0.5 (2u) was noticed in all these
samples indicating mesostructured and amorphous nature of
these materials [1618]. A broad peak ranging from 2u = 15 to 30
in the high angle XRD patterns of these samples (Fig. 1b) conrmed
the amorphous nature of the silica. Further no crystalline
oxide (Al2O3 or Fe2O3) species were evidenced in the high
angle XRD suggesting homogeneous dispersion of Al and Fe
species in these samples.

Fig. 1. Powder XRD patterns of Al-TUD-1, Al-Fe-TUD-1 and Fe-TUD-1 in (a) low angle and (b) high angle.

916

V.V. Srinivasan et al. / Materials Research Bulletin 70 (2015) 914919

The N2 physisorption isotherms of Al-TUD-1, Fe-TUD-1 and


Al-Fe-TUD-1 (Fig. 2a) exhibit a sharp inection at a relative
pressure (P/P0) approximately from 0.45 to 0.85 indicating
capillary condensation within mesopores, a characteristic of
typical mesoporous materials [1618]. With Al-TUD-1, the
structure presented a hysteresis close to H1 type and with
increasing Fe content and in Fe-TUD-1, H2 type hysteresis was
noticed. The specic surface area (BET), total pore volume and
average pore diameter calculated from N2 adsorption isotherms
using the BJH model are summarized in Table 1. A relatively broad
pore size distribution from 3 nm to 20 nm was noticed for all these

samples (Fig. 2b). Nevertheless, the average pore diameter was


estimated to be 4.56.5 nm and the total pore volume increased
with Al content and was observed in the range of 0.720.91 cm3/g.
FTIR spectra of KBr diluted Al-TUD-1, Fe-TUD-1 and Al-Fe-TUD1 samples in the framework region are shown in Fig. 3a. The strong
bands observed at around 1090 cm 1 and 806 cm 1 are characteristic of the Si O Si asymmetric and symmetric stretching
vibrations respectively. The absorption band at 1640 cm 1
corresponds to O H bending vibration arising from absorbed
water in silica network. In addition, in all these samples a weak
band at 966 cm 1 due to the presence of Si O M (M = Al or Fe)

Fig. 2. (a) N2 sorption and (b) pore size distributions of Al-TUD-1, Al-Fe-TUD-1 and Fe-TUD-1.

Table 1
Properties of TUD-1 catalysts.
Catalyst (Si/M)a

Al (wt%)b

Fe (wt%)b

Si/Mc ratio (M = Al + Fe)

SBETd (m2/g)

Vtpe (cc/g)

dP,BJHf (nm)

Total acidity (mmol NH3/g)

Al-TUD-1 (40)
Fe-TUD-1 (40)
Al-Fe-TUD-1 (40)

1.3

0.6 (0.6)g

2.5
1.4 (1.1)g

33
35
35

558
634
521

0.91
0.72
0.85

6.5
4.4
6.5

0.30
0.15
0.21

a
b
c
d
e
f
g

Molar ratio and wt% of Al and Fe in the synthesis gel.


The wt% of Al and Fe in calcined samples.
Actual molar ratio in sample determined by ICP-OES.
SBET = BET specic surface area.
Vtp = total pore volume at 0.99 P/P0.
dP,BJH = average pore diameter calculated from N2 adsorption isotherms using the BJH model.
The wt% of Al and Fe after fourth cycle of Hantzsch reaction by ICP-OES.

Fig. 3. (a) FTIR spectra and (b) diffuse reectance UVvis spectra of Al-TUD-1, Al-Fe-TUD-1 and Fe-TUD-1.

V.V. Srinivasan et al. / Materials Research Bulletin 70 (2015) 914919

917

Fig. 4. (a) NH3-TPD and (b) pyridine adsorbed FTIR spectra of Al-TUD-1, Al-Fe-TUD-1 and Fe-TUD-1.

bending vibration, suggesting framework incorporation of Al or Fe


ions in TUD-1 silicate network [19].
The diffuse reectance spectra of Al-TUD-1, Fe-TUD-1 and AlFe-TUD-1 samples are depicted in Fig. 3b. No signicant absorption
band was noticed for Al-TUD-1 but a sharp and intense peak
centered around 245250 nm was observed for Fe-TUD-1 and
Al-Fe-TUD-1 samples due to the presence of Fe3+ ions in
tetrahedral coordination [20,21]. However, a broad absorption
band ranges between 450 nm and 600 nm in Fe-TUD-1 indicates
the formation of extra-framework small oligomeric Fe2O3
clusters on the surface of the material [20,21].
The acidity of Al-TUD-1, Fe-TUD-1 and Al-Fe-TUD-1 samples
was characterized from ammonia-TPD studies and the results are
shown in Fig. 4a. All the samples displayed a broad desorption of
ammonia between 100  C and 350  C due to the presence of weak
and medium type acid strength. From the amount of ammonia
desorbed, the total acidity of samples followed the order
Al-TUD-1 > Al-Fe-TUD-1 > Fe-TUD-1 (see Table 1). Further,
Al-Fe-TUD-1 showed additional ammonia desorption between
380  C and 500  C evidencing formation of strong acid sites
probably from the interaction of Al and Fe species [12,22].
The FTIR spectra of pyridine adsorbed on Al-TUD-1,
Fe-TUD-1 and Al-Fe-TUD-1 samples are shown in Fig. 4b. A
predominant intense band at 1448 cm 1 ascribed to strong
Lewis acid sites which is found to increase in the order of

Fig. 6.

27

Al-MAS-NMR of (a) A-TUD-1 and (b) Al-Fe-TUD-1.

Al-TUD-1 > Al-Fe-TUD-1 > Fe-TUD-1 similar to the observation in


ammonia-TPD studies. The Brnsted acid sites usually observed at
1545 cm 1 are clearly absent in Fe-TUD-1 whereas it is found
weak in Al-Fe-TUD-1. On the other hand, increases in this band
(1545 cm 1) intensity along with 1490 cm 1 band (due to both
Lewis and Brnsted acid sites) were observed for Al-TUD-1.
These observation clearly suggests that bimetallic Al-Fe-TUD-1
sample presents both the Lewis and Brnsted acid sites
in moderate concentration due to stabilization and high
dispersion of isolated Fe3+ sites which are shown to be active
for various acid catalyzed reactions [12,22].

Fig. 5. TEM images of Al-Fe-TUD-1 at a magnication of (a) 100 nm and (b) 10 nm.

918

V.V. Srinivasan et al. / Materials Research Bulletin 70 (2015) 914919

R
O

O
O

Aldehyde
O
O

O O
Ethyl acetoacetate

Al-Fe-TUD-1

Ethanol,
Refluxed at 80 C

NH4
O

O
O

N
H

O
Ammonium acetate
Scheme 1. Hantzsch reaction over Al-Fe-TUD-1 catalyst.

R2

R3
O HN

R1
H2N

Al-Fe-TUD-1

R3

R1

R2

Scheme 2. Mannich reaction over Al-Fe-TUD-1 catalyst.

The three-dimensional sponge-like mesoporous structure of


Al-Fe-TUD-1 is further evidenced in the TEM image (see Fig. 5a).
Moreover, no diffraction fringes of Fe2O3 or Al2O3 were observed
even at higher magnication (see Fig. 5b), suggesting again
homogeneous distribution of Al3+ and Fe3+ species in Al-Fe-TUD-1.
These results are in correlation with the observations from XRD,
UVvis and FTIR studies.
27
Al-MAS-NMR measurement of Al-TUD-1 and Al-Fe-TUD-1
(Fig. 6) displayed a major resonance peak at 53 ppm suggesting
that majority of aluminum is framework incorporated in tetrahedral coordination [7,23]. Further, a sharp resonance at 0.5 ppm in
Al-TUD-1 indicates the presence of a small amount of extra
framework Al species in the octahedral coordination. Al-Fe-TUD1 also exhibited a peak at 12 ppm that can be assigned to six
coordinated octahedral AlO6 or distorted tetrahedral Al species.
The spinning sideband at 106 ppm arises due to the presence of
Fe3+ in the Al-Fe-TUD-1 [24].
All the three catalysts are subjected to two different acid
catalyzed reaction (see Scheme 1) namely Hantzsch synthesis of
1,4-dihydropyridines (DHP) [15] and Mannich reaction (synthesis
of b-amino carbonyl compounds, Scheme 2) [16] and the results
are presented in Table 2. Both these reactions did not proceed with
siliceous TUD-1 due to the absence of acid sites (typical acidity

Fig. 7. Reusability of Al-Fe-TUD-1 for Hantzsch and Mannich reaction.

0.03 mmol NH3/g). Similarly, no desired product was obtained


with homogeneous AlCl3 or FeCl3 (yield <10%) for Mannich
reaction, whereas considerable yield of DHP (24% and 38%
respectively) was obtained (Table 2, entries 6 and 7).
In sharp contrast, desired products with signicantly higher
yields were obtained over both the heterogeneous Al-TUD-1 and
Fe-TUD-1 catalyst. It is interesting to note that despite higher total
acidity of Al-TUD-1 signicant enhanced activity was observed
over Fe-TUD-1 catalyst (Table 2, entries 1 and 2). This could
possibly be attributed to strong interaction of bases such as
ammonium acetate and aniline with strong Brnsted acid sites of
Al-TUD-1 thus an observed drop in the nal yield of 1,4-DHP and
b-Amino carbonyl compound respectively. Moreover, these
catalysts performed excellent compared to their homogeneous
salts AlCl3 and FeCl3 (Table 2, entries 6 and 7). On the other hand,
Al-Fe-TUD-1 having intermediate Lewis and Brnsted acid sites
compared to both Al-TUD-1 and Fe-TUD-1 gave signicantly higher
yields of 1,4-DHP and b-Amino carbonyl compound (Table 2, entry
3). Zr-SBA-16 and Zr-TUD-1 having predominant Lewis acid sites
[15,25] gave slightly higher yields in Hantzsch and Mannich
reaction (Table 2, entries 4 and 5) compared to Fe-TUD-1(40),
however, showed lower activity compared to bimetallic Al-FeTUD-1(40). Further, commercial catalyst such Hb, montmorillonite
K-10 and bentonite were also compared for this reaction and the
results are given in Table 2 (entries 810). Signicant lower yield in
the strong Brnsted acid catalyst such as Hb clearly indicates that

Table 2
Activity of TUD-1 catalysts for Hantzsch and Mannich reaction.
Entry

Catalyst

Time (h)a

Hantzsch reaction yield (%)b,c

Time (h)a

Mannich reaction yield (%)b,d

1
2
3
4
5
6
7
8
9
10

Al-TUD-1 (40)
Fe-TUD-1 (40)
Al-Fe-TUD-1 (40)
Zr-TUD-1 (40)
Zr-SBA-16 (50)
AlCl3
FeCl3
Hb
Montmorillonite K-10
Bentonite

3
3
3
3
3
3
3
6
2
8

52
71
86
73
69
24
38
56
60
48

10
10
10
10
8
10
10
10
6
12

58
65
74
70
72
Trace
<10
40
48
61

a
b
c
d

Completion of the time monitored by TLC.


Isolated yield.
Reaction conditions: catalyst (100 mg), benzaldehyde (1 mmol), ethyl acetoacetae (2 mmol) and ammonium acetate (1 mmol) at T = 80  C.
Reaction conditions: catalyst (100 mg), benzaldehyde (1 mmol), aniline (1 mmol) and acetophenone (1 mmol) at room temperature.

V.V. Srinivasan et al. / Materials Research Bulletin 70 (2015) 914919

the Brnsted acid sites lead to formation of intermediate and other


products both in Hantzsch and Mannich reaction. Hence, we
conclude that the moderate Lewis and Brnsted acid sites arising
from incorporation of both Al and Fe sites in TUD-1 promotes the
formation of 1,4-DHP and b-Amino carbonyl compound compared
to its monometallic counterparts as observed by other authors in a
different type of reactions [12,22].
Further, to investigate the heterogeneity of the Al-Fe-TUD1 catalyst, after reaction, the catalyst was washed with water and
acetone and activated at 200  C for 2 h, then the reactivated
catalyst was employed in the reaction. Though the catalyst was
reusable up to four cycles slight lowering in the yields was
observed (Fig. 7) which might be attributed to slow leaching of Fe3+
ions (about 27% of Fe in the solid was found leached, Table 2).
4. Conclusions
We have successfully synthesized bimetallic Al and Fe
incorporated TUD-1 material and is characterized by presence of
both Lewis and Brnsted acid sites. Diffuse reectance UVvis
spectrum conrms the presence of isolated Fe3+ in the framework
and further supported by FTIR studies. Al-Fe-TUD-1 is shown to be
an efcient catalyst for the synthesis of 1,4-dihydropyridine and
b-amino carbonyl compounds through Hantzsch and Mannich
reaction due to the synergy between Al and Fe sites in TUD-1.
Further, the catalyst can be easily recovered and reused without
signicant leaching of active metal ions.
References
[1] J. Weitkamp, Zeolites and catalysis, Solid State Ion. 131 (2000) 175188, doi:
http://dx.doi.org/10.1016/S0167-2738(00)00632-9.
[2] N. Rahimi, R. Karimzadeh, Catalytic cracking of hydrocarbons over modied
ZSM-5 zeolites to produce light olens: a review, Appl. Catal. Gen. 398 (2011)
117, doi:http://dx.doi.org/10.1016/j.apcata.2011.03.009.
[3] H. Kosslick, G. Lischke, B. Parlitz, W. Storek, R. Fricke, Acidity and active sites of
Al-MCM-41, Appl. Catal. Gen. 184 (1999) 4960.
[4] R. Mokaya, W. Jones, Grafting of Al onto purely siliceous mesoporous
molecular sieves, Phys. Chem. Chem. Phys. 1 (1999) 207213, doi:http://dx.doi.
org/10.1039/A807919F.
[5] G. Muthu Kumaran, S. Garg, K. Soni, M. Kumar, J.K. Gupta, L.D. Sharma, et al.,
Synthesis and characterization of acidic properties of Al-SBA-15 materials
with varying Si/Al ratios, Microporous Mesoporous Mater. 114 (2008) 103109,
doi:http://dx.doi.org/10.1016/j.micromeso.2007.12.021.
[6] C. Simons, U. Hanefeld, I.W. Arends, R.A. Sheldon, T. Maschmeyer, Noncovalent
anchoring of asymmetric hydrogenation catalysts on a new mesoporous
aluminosilicate: application and solvent effects, Chem. Eur. J. 10 (2004) 5829
5835.
[7] R. Anand, R. Maheswari, U. Hanefeld, Catalytic properties of the novel
mesoporous aluminosilicate AlTUD-1, J. Catal. 242 (2006) 8291, doi:http://dx.
doi.org/10.1016/j.jcat.2006.05.022.
[8] I. Neves, G. Botelho, A. Machado, P. Rebelo, S. Rama, M. Pereira, et al.,
Feedstock recycling of polyethylene over AlTUD-1 mesoporous catalyst,
Polym. Degrad. Stab. 92 (2007) 15131519.

919

[9] P. Neves, S. Lima, M. Pillinger, S.M. Rocha, J. Rocha, A.A. Valente, Conversion of
furfuryl alcohol to ethyl levulinate using porous aluminosilicate acid catalysts,
Catal. Today 218219 (2013) 7684, doi:http://dx.doi.org/10.1016/j.
cattod.2013.04.035.
[10] S. Lima, M.M. Antunes, A. Fernandes, M. Pillinger, M.F. Ribeiro, A.A. Valente,
Acid-catalysed conversion of saccharides into furanic aldehydes in the
presence of three-dimensional mesoporous Al-TUD-1, Molecules 15 (2010)
38633877.
[11] S. Telalovi
c, A. Ramanathan, J.F. Ng, R. Maheswari, C. Kwakernaak, F. Soulimani,
et al., On the synergistic catalytic properties of bimetallic mesoporous
materials containing aluminum and zirconium: the Prins cyclisation of
citronellal, Chem. Eur. J. 17 (2011) 20772088, doi:http://dx.doi.org/10.1002/
chem.201002909.
[12] R. Savidha, A. Pandurangan, M. Palanichamy, V. Murugesan, A comparative
study on the catalytic activity of Zn and Fe containing Al-MCM-41 molecular
sieves on t-butylation of phenol, J. Mol. Catal. A Chem. 211 (2004) 165177, doi:
http://dx.doi.org/10.1016/j.molcata.2003.10.022.
[13] Y. Li, Z. Feng, H. Xin, F. Fan, J. Zhang, P.C. Magusin, et al., Effect of aluminum on
the nature of the iron species in Fe-SBA-15, J. Phys. Chem. B 110 (2006) 26114
26121.
[14] R.A. van Santen, J. Emiel, Iron-functionalized Al-SBA-15 for benzene
hydroxylation, Chem. Commun. (6) (2008) 774776, doi:http://dx.doi.org/
10.1039/B717079C.
[15] R. Maheswari, V.V. Srinivasan, A. Ramanathan, M.P. Pachamuthu, R.
Rajalakshmi, G. Imran, Preparation and characterization of mesostructured
Zr-SBA-16: efcient Lewis acidic catalyst for Hantzsch reaction, J. Porous
Mater. 22 (2015) 705711, doi:http://dx.doi.org/10.1007/s10934-015-9943-7.
[16] M.P. Pachamuthu, K. Shanthi, R. Luque, A. Ramanathan, SnTUD-1: a solid acid
catalyst for three component coupling reactions at room temperature, Green
Chem. 15 (2013) 21582166, doi:http://dx.doi.org/10.1039/C3GC40792F.
[17] K. Kandasamy, M.P. Pachamuthu, M. Muthusamy, S. Ganesabaskaran, A.
Ramanathan, Synthesis of novel pyrazolylbiscoumarin derivatives using
FeTUD-1 as a mesoporous solid acid catalyst, RSC Adv 3 (2013) 25367
25373, doi:http://dx.doi.org/10.1039/C3RA43913E.
[18] M.P. Pachamuthu, V.V. Srinivasan, R. Maheswari, K. Shanthi, A. Ramanathan,
The impact of the copper source on the synthesis of meso-structured CuTUD1: a promising catalyst for phenol hydroxylation, Catal. Sci. Technol. 3 (2013)
33353342.
[19] Z. Luan, J.A. Fournier, In situ FTIR spectroscopic investigation of active sites and
adsorbate interactions in mesoporous aluminosilicate SBA-15 molecular
sieves, Microporous Mesoporous Mater. 79 (2005) 235240, doi:http://dx.doi.
org/10.1016/j.micromeso.2004.11.012.
[20] M.P. Pachamuthu, S. Karthikeyan, G. Sekaran, R. Maheswari, A. Ramanathan,
Fenton-type oxidative degradation of N,N-diethyl-p-phenylenediamine by a
mesoporous wormhole structured FeTUD-1 catalyst, CLEAN Soil Air Water 43
(2015) 375381, doi:http://dx.doi.org/10.1002/clen.201300747.
[21] A.M. Balu, A. Pineda, K. Yoshida, J.M. Campelo, P.L. Gai, R. Luque, et al., Fe/Al
synergy in Fe2O3 nanoparticles supported on porous aluminosilicate
materials: excelling activities in oxidation reactions, Chem. Commun. 46
(2010) 78257827.
[22] A.C. Akah, G. Nkeng, A.A. Garforth, The role of Al and strong acidity in the
selective catalytic oxidation of NH3 over Fe-ZSM-5, Appl. Catal. B Environ. 74
(2007) 3439, doi:http://dx.doi.org/10.1016/j.apcatb.2007.01.009.
[23] S. Telalovi
c, S.K. Karmee, A. Ramanathan, U. Hanefeld, Al-TUD-1: introducing
tetrahedral aluminium, J. Mol. Catal. A Chem. 368369 (2013) 8894, doi:
http://dx.doi.org/10.1016/j.molcata.2012.11.018.
[24] P.J. Dirken, Solid-state MAS NMR study of pentameric aluminosilicate groups
with 180 intertetrahedral Al O Si angles in zunyite and harkerite, Am.
Mineral. 80 (1995) 3945.
[25] M.P. Pachamuthu, V.V. Srinivasan, R. Maheswari, K. Shanthi, A. Ramanathan,
Lewis acidic ZrTUD-1 as catalyst for tert-butylation of phenol, Appl. Catal. Gen.
462463 (2013) 143149, doi:http://dx.doi.org/10.1016/j.apcata.2013.05.008.

You might also like