Você está na página 1de 8

Journal of Molecular Catalysis A: Chemical 390 (2014) 198205

Contents lists available at ScienceDirect

Journal of Molecular Catalysis A: Chemical


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

Amino-functionalized Zr(IV) metalorganic framework as


bifunctional acidbase catalyst for Knoevenagel condensation
Yang Yang, Hong-Fei Yao, Fu-Gui Xi, En-Qing Gao
Shanghai Key Laboratory of Green Chemistry and Chemical Processes, Department of Chemistry, East China Normal University, Shanghai 200062, China

a r t i c l e i n f o a b s t r a c t

Article history: The amino-functionalized metalorganic framework of Zr(IV) with 2-aminoterephthalate, UiO-66-NH2 ,
Received 6 December 2013 was studied as a solid catalyst for Knoevenagel condensation. The material can efciently catalyze the
Received in revised form 1 April 2014 condensation reaction of benzaldehyde with ethyl cyanoacetate or malononitrile in highly polar solvents
Accepted 3 April 2014
such as DMF, DMSO and ethanol. The catalytic system has also been tested for various aromatic alde-
Available online 13 April 2014
hydes, the conversion easily reaching more than 90% under mild conditions. It was demonstrated that the
catalytic process is heterogeneous and shows size effects, characteristic of a porous catalyst. The catalyst
Keywords:
can be recycled without losing its framework integrity and catalytic activity. The catalytic activity has
Metalorganic frameworks
Bifunctional acidbase catalysts
been compared with dimethyl 2-aminoterephthalate and the isostructural amino-free MOF (UiO-66).
Knoevenagel condensation The superior performance of UiO-66-NH2 has been attributed to the site-isolated acidbase bifunctional
Heterogeneous catalysis character. It has been proposed that the Zr site in close proximity to the amino group activates aldehydes
to promote the formation of aldimine intermediates from the aldehydes and the amino group.
2014 Elsevier B.V. All rights reserved.

1. Introduction may be inherent in the frameworks or generated by post-synthetic


methods [11,1722].
In recent years, metalorganic frameworks (MOFs) have The three-dimensional (3D) Zr(IV) MOFs of formula
attracted considerable attention due to their special characteristics [Zr6 O4 (OH)4 (L)6 ], such as UiO-66 [l = 1,4-benzenedicarboxylate
such as hybrid compositions, diverse networks, tunable porosity (BDC)] and UiO-66-NH2 [l = 2-amine-1,4-benzenedicarboxylate
and tailorable surfaces [1,2]. Owing to these characteristics, MOFs (BDC-NH2 )], are based on the octahedral [Zr6 (3 -O)4 (3 -OH)4 (2 -
are promising materials for various technological applications such COO)12 ] cluster and featured by alternating octahedral and
as gas storage/capture [35], separation [6,7], and heterogeneous tetrahedral cages sharing triangular windows [23] (Fig. 1). The
catalysis [811]. Heterogeneous catalysis is superior to homoge- MOFs contain potential Lewis-acid sites (Zr(IV) centers) upon
neous catalysis for easier separation, reusability, minimized waste activation and/or base sites (NH2 groups), and remarkably, they
and cleaner products. Heterogeneous catalysts also offer the possi- show exceptionally high thermal and chemical stability. However,
bility of combining isolated acidic and basic sites for cooperative or for only a few reactions have these MOFs been studied as hetero-
tandem catalysis [1215]. Furthermore, porous solid catalysts may geneous catalysts, including the cross-aldol condensation between
provide conned space to inuence reactivity and selectivity. The benzaldehyde and heptanal [24], the cyclization of citronellal to
application of MOFs as porous heterogeneous catalysts alternative isopulegol [25,26], the cycloaddition of CO2 to styrene oxide [27],
or complementary to microporous zeolites is especially interesting, the acetalization of benzaldehyde with methanol [28], and photo-
since the pore size and chemical functionality of MOFs can be mod- catalytic reactions [2934]. Electronic effects of linker substitution
ulated within a wider range [2,11,16]. In principle, the active sites on Lewis acid catalysis have been studied [28]. It was found that
of MOF catalysts can be metal centers with unsaturated (or labile) the Lewis acidity and the catalytic activity increase in the order of
coordination environments, catalytic functional groups attached UiO-66-NH2 < UiO-66 < UiO-66-NO2 . It was also demonstrated that
to any components of the frameworks, or other catalytic species the introduction of amino groups into linkers leads to an increase
(molecules, metals, etc.) encapsulated in the pores, and the sites in basicity [25,28].
In this paper, we report the catalytic study of UiO-66-NH2 for
Knoevenagel condensation. The Knoevenagel condensation of a
carbonyl group with the methylene group activated by two elec-
Corresponding author. Tel.: +86 21 62233404; fax: +86 21 62233404. tron withdrawing groups (Scheme 1) is an important C C bond
E-mail address: eqgao@chem.ecnu.edu.cn (E.-Q. Gao). coupling reaction and has been widely used in the synthesis of ne

http://dx.doi.org/10.1016/j.molcata.2014.04.002
1381-1169/ 2014 Elsevier B.V. All rights reserved.
Y. Yang et al. / Journal of Molecular Catalysis A: Chemical 390 (2014) 198205 199

2.2. Characterization

The X-ray powder diffraction patterns of the samples were


measured using a Rigaku Ultima IV X-ray Diffractometer with
Cu K radiation ( = 1.54 A) at a scanning rate of 10 C/min, with
accelerating voltage and current of 35 kV and 25 mA, respec-
tively. SEM study was carried out with a S-4800 HITACHI scanning
electron microscope. Elemental analyses were determined on an
Elementar Vario ELIII analyzer. FT-IR spectra were recorded in
the range 5004000 cm1 using KBr pellets on a Nicolet NEXUS
670 spectrophotometer. The leached metal amount was detected
by inductively coupled plasma atomic emission spectroscopy
(ICP-AES) using a IRIS Intrepid II XSP spectrometer. Gas chromatog-
Fig. 1. Structure of UiO-66. (a) A [Zr6 (3 -O)4 (3 -OH)4 (2 -COO)12 ] cluster. (b) The
raphy (GC) was conducted using a Linghua GC 9890E instrument
porous systems in which octahedral and tetrahedral cages share triangular windows. equipped with an FID detector and an SE-54 capillary column
(30 m 0.25 mm 0.25 m). The temperature program for GC
analysis was set as follows: the temperature was held at 40 C for
chemicals and pharmaceuticals. The condensation can be catalyzed 1 min, then raised to 260 C at 30 C/min and held for 5 min. Inlet
by bases or Lewis acids, either homogeneous (such as amines and detector temperatures were 280 C. The analysis was carried
[35,36], ZnCl2 and Mg(ClO4 )2 [37,38]) or heterogeneous (such as out directly after sampling to avoid any additional conversion.
hydrotalcite [39], silica-supported basic catalysts [40,41]). Consid-
ering the demand for environmentally friendly heterogeneous 2.3. Catalytic test
catalysts and the intriguing features of MOFs, the exploration of
MOF catalysts for the Knoevenagel condensation has become a In a typical catalytic experiment, 0.144 g of the catalyst (corre-
topic of increasing interest in recent years [4250]. The MOFs sponding to 0.45 mmol amino groups) was placed in a ask and
used are diverse in composition and structure, and most of them heated at 150 C for 4 h. After three cycles of vacuum pumping
contain the amino group as Lewis base sites, such as the BDC-NH2 - and nitrogen injection, the solution of ethyl cyanoacetate or other
base metalcarboxylate frameworks IRMOF-3, MIL-101(Al)-NH2 , active methylene compounds (10 mmol) in an appropriate solvent
MIL-53-NH2 , and UMCM-1-NH2 [45,46]. MIL-101(Cr) with post- (5 mL) was added with stirring and heated to a given temperature.
synthetically introduced amino groups has also been studied for After temperature equilibrium, benzaldehyde (5 mmol) was added
Knoevenagel condensation [51]. In this work, we demonstrate that to initiate the reaction. The reaction mixture was stirred under a
UiO-66-NH2 is an efcient, size-selective, stable, and recyclable static nitrogen atmosphere, and small aliquots of the supernatant
heterogeneous catalyst and that the superior catalytic performance were withdrawn at different time intervals to monitor the reaction
to UiO-66 is attributable to the bifunctional acidbase character. conversion by GC.

2. Experimental 3. Results and discussion

2.1. Synthesis 3.1. Characterization of the UiO-66-NH2 catalyst

All chemicals (A. R. grade, purity 98 wt% or higher) were UiO-66-NH2 was synthesized according to literature proce-
obtained from commercial sources and used without further dures. The XRD pattern of the as-synthesized sample (Fig. 2a) is
purication. UiO-66-NH2 and UiO-66 synthesized by follow- in good agreement with the literature data and with the simu-
ing a solvothermal procedure reported in the literature [52,53], lated pattern of pristine UiO-66 [24], evidencing good synthetic
using acetic acid as modulator. ZrCl4 (0.7 mmol, 0.164 g) and reproducibility and phase purity. The SEM morphology (Fig. 2b)
H2 BDC-NH2 (0.7 mmol, 0.127 g) were dissolved in the mixture shows that the particles are of nanometer size with a distribu-
of N,N -dimethylformamide (DMF, 8 ml), acetic acid (1.2 ml) and tion in the range 4090 nm. After thermal activation at 150 C, the
deionized water (0.05 ml) at room temperature. After stirring for reections at large angles become less intensive, but the main char-
about 5 minutes, the mixture was sealed in a 23 ml Teon liner, acteristic peaks remain, indicating that the overall phase integrity
heated in an oven at 120 C for 48 h, and then cooled to room is retained although the activation leads to some degree of disor-
temperature. A yellow solid powder was obtained by ltration, der in the framework. Fig. 2a also shows the XRD patterns of the
washed three times with DMF. The product was soaked and stirred UiO-66-NH2 sample after one and three cycles of catalytic reac-
in reuxing methanol for 24 h, ltered and dried at 70 C in an oven. tions. It is interesting to note that the reections at large angles
The formula of the material thus obtained was estimated to be are recovered after the catalytic reaction, indicating that the dis-
[Zr6 O4 (OH)4 (BDC-NH2 )6 ]14H2 O according to elemental analysis. order caused by thermal treatment is repaired. It can be deduced
Found: C 29.2, H 3.8, N 4.4%; calculated: C 28.7, H 3.1, N 4.2%. that the original crystalline phase of the material is retained with-
out signicant degradation in structural integrity after the catalytic
reactions.

3.2. Catalytic properties

3.2.1. Inuence of catalytic dose, solvent, and temperature


The condensation reaction of benzaldehyde with ethyl cyanoac-
etate, which affords -cyanocinnamate, was performed as test
reaction to study the effects of different reaction conditions, includ-
Scheme 1. Knoevenagel condensation. ing the catalytic dose, the solvent, and the reaction temperature.
200 Y. Yang et al. / Journal of Molecular Catalysis A: Chemical 390 (2014) 198205

effects are similar to those observed for some basic organocatalysts


[35,36] and for the amine-functionalized MOF catalysts IRMOF-
3 [45,46] and NH2 -MIL-101(Al) [48], although sometimes ethanol
was found to be better than DMF and DMSO due to its amphiprotic
character. On the other hand, the effects are in clear contrast with
those observed for some amino-tagged silicas, for which toluene is
better than polar solvents [40,41]. For some other catalysts, the
solvent effects are not necessarily related to the polarity or the
amphiprotic nature. For example, the catalytic activity of ZIF-8, -
9 and -10 (ZIF = zeolite imidazolate framework) for Knoevenagel
condensation in toluene was found to be intermediate between
that in THF (higher) and that in DCM (lower) [49,50]. For solid-
supported base catalysts, it has been reported that the higher the
hydrophilicity (or the surface polarity) of the support, the lower the
effect of the solvent [41,46], but the opposite trend has also been
reported [40]. Therefore, there is no general trend for the effect of
solvents on the Knoevenagel reaction. Nevertheless, the results for
UiO-66-NH2 , IRMOF-3 and NH2 -MIL-101(Al) consistently indicate
that amino-tagged MOFs perform better in highly polar solvents.
The temperature dependence of the catalytic performance of
UiO-66-NH2 was checked by performing the Knoevenagel reaction
in DMF at 40, 60 and 80 C (Table 1, entries 57). As expected, the
reaction rate increases quickly with the temperature, the conver-
sion within 30 min changing from 27% at 40 C to 91% at 80 C. The
conversion at 40 C can reach a high level (>90%) within 2 h.

3.2.2. The inuence of different substrates


Having established that UiO-66-NH2 is a good catalyst for
the Knoevenagel condensation between benzaldehyde and ethyl
cyanoacetate, we extend the study to various substrates, includ-
ing different aromatic aldehydes and methylene compounds. The
results are listed in Table 2.
Diethyl malonate and malononitrile have been compared with
ethyl cyanoacetate (entries 13). Under the same catalytic condi-
tions, diethyl malonate led to a trace conversion of benzaldehyde
after 2 h, whereas malononitrile caused a conversion of 98% within
40 min. The different reactivity of the methylene compounds is
consistent with their different acidity, which increases in the
Fig. 2. PXRD patterns (a) and SEM picture (b) of UiO-66-NH2 .
order of diethyl malonate (pKa = 16.4 [54]) < ethyl cyanoacetate
(pKa = 13.1 [55]) < malononitrile (pKa = 11.1 [56]). Actually, no Ar-
The results obtained for the reactions performed in ethanol at NH2 -functionalized materials (including MOFs) have been found
80 C in the presence of different amount of catalysis are given to be able to efciently catalyze the Knoevenagel reaction between
in Table 1 (entries 13). While the blank reaction in the absence benzaldehyde and diethyl malonate, because the pKa is too high
of any catalyst gave a benzaldehyde conversion of only 27% after for the amine group or the in situ formed imine group to induce
2 h, introducing UiO-66-NH2 and increasing the amount led to deprotonation.
signicant increase in the conversion. With 0.144 g UiO-66-NH2 Various aromatic aldehydes have been tested for catalytic Kno-
(corresponding to 9 mol% amino groups), the conversion reached evenagel condensation with malononitrile. It was found that the
94% after 2 h. For comparison, the blank reaction needs 17 h to reach activity of UiO-66-NH2 for 2-nitrobenzaldehyde is much higher
a conversion of >90% (complete conversion needs more than 21 h). than that for benzaldehyde, as the reaction for the former alde-
These results clearly conrm that UiO-66-NH2 is active in catalyz- hyde can reach completion within only 5 min (entry 4 in Table 2).
ing the Knoevenagel reaction. No by-product was observed during This reects the great accelerating effect of the strong electron-
the reaction, so the selectivity for -cyanocinnamate is 100%. withdrawing nitro group, as would be expected for a reaction
Strong effects of solvent polarity have been observed for cat- involving nucleophilic attack at the carbonyl group. The catalyzed
alytic Knoevenagel reactions. To establish the solvent effects for reactions of the three methyl-substituted benzaldehydes proceed
UiO-66-NH2 , we performed the reaction in different solvents readily to give 97% conversions after 40 min (entries 57). The
(Table 1). It proved that the catalyst has the highest activity (>90% conversions are similar to that for benzaldehyde, but the initial
conversion within 2 h) in DMSO and DMF (entries 5, 8), which have reaction rates of these substrates are different. As shown in Fig. 3,
the highest polarity (dielectric constants = 48.9 for DMSO and 36.7 compared with the non-substituted benzaldehyde, the initial reac-
for DMF). Ethanol (protic, = 24.3) is also a good medium for the tion rate of 2-methylbenzaldehyde is similar (or slightly faster),
reaction (entry 4). By contrast, nonprotic and less polar solvents, the reaction of the 3-isomer is obviously slower, and notably, the
including DCM ( = 9.1), THF ( = 7.5), EtOAc ( = 6.02) and toluene 4-isomer was more reactive, though methyl is an electron-donating
( = 2.4), are not good media (entries 912). The effects can be group. The position effect of methyl substitution on the reactivity
generally explained by the assumption that polar solvents help to of benzaldehyde is similar to that observed for the Knoeve-
stabilize the charged transition-state complex of the reaction [36]. nagel condensation catalyzed by ZIF-8 [49] and ZIF-9 [50], where
Besides, the strong hydrogen-bonding acceptor ability of DMSO and the reactivity varies in the order of 4-methylbenzaldehyde 2-
DMF can facilitate proton transfer [45]. On the one hand, the solvent methylbenzaldehyde > benzaldehyde > 3-methylbenzaldehyde.
Y. Yang et al. / Journal of Molecular Catalysis A: Chemical 390 (2014) 198205 201

Table 1
UiO-66-NH2 catalyzed Knoevenagel condensation between benzaldehyde and ethyl cyanoacetate: inuence of catalytic dose, solvent, and temperature.

.

Entry Solvent T ( C) Time (h) Conv. (%)

1 EtOH 80 2 27
2 80 2 67
3 80 2 94
4 40 0.5/2 25/79
5 DMF 40 0.5/2 27/92
6 60 0.5 58
7 80 0.5 91
8 DMSO 40 0.5/2 31/95
9 DCM 40 2 5
10 EtOAc 40 2 3
11 THF 40 2 2
12 Toluene 40 2 <1

Conditions: reactants: benzaldehyde (5 mmol), ethyl cyanoacetate (10 mmol); solvent: 5 ml; Cat: UiO-66-NH2 , 0.144 g (corresponding to 0.45 mmol amino groups), except
for entry 1 (no catalyst) and entry 2 (0.072 g, corresponding to 0.23 mmol amino groups).

The catalyst system is also efcient for the Knoevenagel UiO-66-NH2 , the reaction with ethyl cyanoacetate shows a much
reactions of larger aromatic aldehydes (1-naphthaldehyde and higher conversion (56%), indicating a catalytic process. By contrast,
9-anthraldehyde) with malononitrile, though the conversion is the use of UiO-66-NH2 for the reaction with tert-butyl cyanoac-
somewhat decreased when the aldehyde substrate becomes more etate leads to a conversion (15%) comparable to (or even slightly
bulky (comparing entries 3, 8 and 9 in Table 2). The decreased lower than) that for the reaction without any catalyst, indicating
conversion may be a (weak) indicator of size effects. The effects that UiO-66-NH2 is not active for the two bulky substrates. Thus, it
are more appreciable for the reactions of different aldehydes with is very likely that the catalytic reaction over UiO-66-NH2 proceeds
ethyl cyanoacetate (molecular dimensions 3.4 7.3 A 2 ), which inside the pores. The pore system of perfect UiO-66 crystals con-
are bulkier than malononitrile (2.8 4.3 A 2 ). As shown by entries sists of interconnected tetrahedral and octahedral cages with free
1 and 10, the conversion of 9-anthraldehyde (56%, 5.9 9.2 A 2 ) diameters of about 8 and 11 A, respectively. Notably, some recent
reacting with ethyl cyanoacetate is much lower than that of benz- studies have demonstrated that real UiO-66 and analogs contains
aldehyde (92%, 6.0 4.3 2 ). The results could reect that the an amount of defects due to missing linkers between Zr6 clusters
probability of two bulky substrates forming transition-state com- [26,57,58]. The linker vacancies lead to expanded pores so that
plexes is signicantly reduced due to the limited space in the the real pore dimensions can be larger. The failure in catalyzing
catalyst. the reaction between 9-anthraldehyde and tert-butyl cyanoacetate
To further conrm the size effects, 9-anthraldehyde was could be because the pores are still too small for the two bulky sub-
reacted with a still bulkier methylene substrate, tert-butyl strates diffuse into the pores of the catalyst to access the active sites
cyanoacetate (4.3 8.2 A 2 ). The independent control reactions of and to form the transition state required for the reaction.
9-anthraldehyde with ethyl and tert-butyl cyanoacetates in the
absence of any catalyst give similar conversions (around 20% after 3.2.3. Heterogeneity and reusability
2 h, entries 10 and 11), indicating that the intrinsic reactivity of With a solid catalyst for liquid-phase reactions, an issue of great
the two methylene substrates is comparable. In the presence of concern is whether the catalytic process is heterogeneous or homo-
geneous. In the latter case, the species leached into the liquid phase
is partially or fully responsible for the observed catalytic activity.
This is practically undesirable. The heterogeneous nature of our cat-
alyst system for Knoevenagel reactions is implied in some results
of the above catalytic experiments. For example, the dependence of
benzaldehyde conversion on the amount of solid catalyst (Table 1)
could be an implication. The solid shows no appreciable solvabil-
ity. It means that if there were any dissolved species, it would be
quite easy for the liquid phase to be saturated by the species. Once
saturated, the amount of the species in a given volume of liquid
would be independent of the dose of the solid. Then, if the dissolved
species were responsible for the catalytic activity, the reaction con-
version would be independent of the amount of the solid. Therefore,
the dependence of conversion on solid dose may be taken as a
(weak) indication of heterogeneous catalysis. A stronger evidence
is the size selectivity for substrates (Table 2), which is difcult to
achieve with homogeneous catalysts but characteristic of porous
heterogeneous catalysts. In order to give a more direct conrma-
tion, a control reaction was performed between benzaldehyde and
Fig. 3. Time dependence of reaction conversion for the condensation of differ-
ethyl cyanoacetate in DMF. After the reaction proceeded at 40 C
ent monosubstituted bezaldehydes (5 mmol) with malononitrile (10 mmol) in DMF
(5 ml, 40 C) in the presence of 0.144 g catalyst (corresponding to 0.45 mmol amino for 2 h, the catalyst was removed by hot ltration. The ltrate was
groups). further heated at the same temperature and the composition was
202 Y. Yang et al. / Journal of Molecular Catalysis A: Chemical 390 (2014) 198205

Table 2
UiO-66-NH2 catalyzed Knoevenagel condensation of different substrates.a

Entry Aldehyde Methylene Time (min) Conv. (%)

O
N
1 120 92
O

2 120 <2

3 40 98

4 5 100

5 40 98

6 40 97

7 40 98

8 40 92

9 40 87

10 120 56 (21)b

11 120 15 (19)b

a
Conditions: aldehyde (5 mmol), methylene compound (10 mmol), DMF (5 ml), Catalyst (0.144 g, corresponding to 0.45 mmol amino groups), 40 C.
b
The conversion for the control test in the absence of any catalyst is given in parentheses for comparison.

monitored at given time intervals. The results are compared in Fig. 4 cycles for the reaction between benzaldehyde and malononitrile.
with those obtained for the reaction under the same conditions but After each cycle, the solid was ltered out, soaked and stirred in
without the ltering procedure. Obviously, no appreciable reaction methanol for 24 h, ltered out again and washed with methanol.
took place after removal of the catalyst, indicating that no active The recovered solid was heated at 150 C for 4 h and then reused
species were leached into the liquid phase and providing a direct in the next cycle. The results of the three cycles are very simi-
evidence for the heterogeneity of the catalyst system. This is also lar. As shown in Fig. 5a, the conversion of benzaldehyde after 1 h
conrmed by ICP-AES analysis, which suggests no detectable Zr in remains at the high level of 97%, indicating that the catalyst could
the liquid phase. be reused without signicant degradation in catalytic performance.
Another issue of great concern for solid catalysts is the recycla- The recyclability is consistent with the good chemical resistance
bility. To check this for UiO-66-NH2 , the catalyst was used for three of the structure against the reaction conditions, which has been
Y. Yang et al. / Journal of Molecular Catalysis A: Chemical 390 (2014) 198205 203

Fig. 4. Filtration test of the catalyst. Conditions: benzaldehyde (8 mmol), ethyl


cyanoacetate (7 mmol), Cat. 0.10 g (corresponding to 0.31 mmol amino groups),
solvent: DMF (5 ml), 40 C. The conversion is based on ethyl cyanoacetate.

Fig. 6. (a) Knoevenagel condensation with different catalysts. Conditions: benzal-


dehyde (5 mmol), ethyl cyanoacetate (10 mmol), solvent: EtOH (5 ml), 80 C, Cat.
0.45 mmol Zr or NH2 , 2 h. (b) The reactions of benzaldehyde (1 mmol) with DMBDC-
NH2 (1.2 mmol) in EtOH (6 ml, 80 C) in the absence/presence of UiO-66 (0.014 g).

demonstrated by XRD measurements (see Fig. 2a). The intactness


of the catalyst is also conrmed by the fact the FT-IR spectra of the
catalyst shows no appreciable difference after catalytic reactions
(Fig. 5b).

3.2.4. Mechanism considerations


To determine the role of the amine group in the catalyst, the
amine-free MOF (UiO-66) prepared and activated by similar proce-
dures was tested as catalyst for the reaction between benzaldehyde
and ethyl cyanoacetate. As can be seen from Fig. 6a, under the
same catalytic conditions, the amino-free MOF led to much lower
conversion than UiO-66-NH2 . Since the two MOFs are isostruc-
tural, the superior performance of UiO-66-NH2 indicates that the
basic amine group is important in promoting the Knoevenagel
reaction, as is normally expected for NH2 -tagged catalysts. It has
been demonstrated that the introduction of amino groups into
the linkers of UiO-66 leads to an appreciable increase in basicity
[28]. The organic analog of UiO-66-NH2 , dimethyl 2-amine-1,4-
benzenedicarboxylate (DMBDC-NH2 ), was also tested under the
same conditions. The homogeneous reaction showed a low conver-
Fig. 5. (a) Catalyst recycling studies. Conditions: benzaldehyde (5 mmol), malonon- sion of 29%, which is only slightly higher than the reaction without
itrile (10 mmol), solvent: DMF (5 ml), Cat. 0.144 g, 40 C, 1 h. (b) IR spectra of the fresh any catalysts and much lower than the heterogeneous reaction
and used catalysts.
using UiO-66-NH2 (see Fig. 6a), though the amount of amino groups
204 Y. Yang et al. / Journal of Molecular Catalysis A: Chemical 390 (2014) 198205

was set to be equal in two catalytic reactions. This could indicate


that the catalytic activity of amine is enhanced in the MOF struc-
ture or that the reaction is promoted by other sites besides amine.
The Knoevenagel reactions with aniline as homogenous catalyst
have also been reported to give much lower conversion than those
with IRMOF-3 as heterogeneous catalyst [45,46]. The high activity
of IRMOF-3 compared to aniline was attributed to the increased
basicity of the aromatic amino group when incorporated within
the MOF structure. It was proposed that the increase of basicity
could be due to the intramolecular hydrogen bonding interaction
between the amino group and an carboxylate oxygen (coordinated
to Zn(II)) [45]. However, this explanation fails for DMBDC-NH2 ,
which could have similar hydrogen interactions but does not show
signicant catalytic activity. A DFT study [59] suggested that the
basicity of the aromatic amine (Ar-NH2 ) and the aldimine derivative
(Ar-N = CHC6 H5 ) indeed increases in going from aniline to DMBDC-
NH2 and that the increase is more signicant in going further to
the amine embedded in IRMOF-3. The increase has been related to
the stabilization of the protonated aminium and iminium species,
which are hydrogen bonded to a carboxylate oxygen to generate
a 6-membered planar ring. The theoretical study also indicated
that the aldimine intermediates are the most basic species in the
reaction systems and should be the active species that deproto-
nate ethyl cyanoacetate. However, the DFT energetic study on the
catalytic cycles of Knoevenagel condensation indicated that the
increased catalytic activity of IRMOF-3 seems to be unrelated to
its increased basicity but could be explained by its adsorption abil-
ity for the water by-product which would otherwise poisons the
catalytic amino sites.
It should be noted that the theoretical study was based on the
hypothesis that the amine groups are the only catalytic sites in
the MOF, not considering the possibility of the existence of other
active sites. A recent reinvestigation of IRMOF-3 for Knoevenagel
condensation suggested that the MOF behaves as an unexpected Scheme 2. (a) Mechanism for amine-catalyzed Knoevenagel condensation. (b)
Aldimine formation promoted by the Lewis acid site in close proximity to the amine
bifunctional acidbase catalyst [15]. The unintentional acid sites
group.
come from a defective origin, either ZnOH species formed upon
partial hydrolysis of the framework or ZnO particles entrapped
inside the cavities during the synthesis. in Fig. 6a, ZrO2 and DMBDC-NH2 show very low activity (the con-
As already mentioned, UiO-66 is much less active than UiO- versions are higher than the blank test by only 8% and negligibly
66-NH2 , but it is noteworthy that the NH2 -free MOF still shows 2%, respectively), while the mixture show enhanced activity (15%
signicant catalytic activity (50% after 2 h, See Fig. 6a) when com- higher than the blank test), indicating the presence of some coop-
pared with the blank non-catalytic reaction. This suggests that the eration, although weak. The much higher activity of UiO-66-NH2
Knoevenagel reaction could also be promoted by Zr(IV) centers than UiO-66 may indicate much stronger cooperation in the porous
(Lewis acid sites), so UiO-66-NH2 could be a bifunctional acidbase framework. Probably, the framework arranges Zr and amino sites
catalyst for Knoevenagel condensation. The bifunctional charac- in close proximity on the inner surface so that the two sites can
ter has recently been proposed to explain the high activity of the cooperate more efciently.
material for the cross-aldol condensation between heptanal and Knoevenagel condensation, as a modication of aldol conden-
benzaldehyde [24]. The acid sites in UiO-66 and UiO-66-NH2 could sation, may proceed via different mechanisms that depend on the
be the open Zr(IV) sites generated by dehydration of the Zr6 O4 (OH)4 nature of the catalyst used. A Lewis acid site usually interacts
cluster to give Zr6 O6 . It has been shown that the dehydration begins with the carbonyl oxygen of benzaldehyde, thus the C O bond is
at about 100 C for UiO-66 and even below the temperature for further polarized to facilitate the nucleophilic attack at the car-
UiO-66-NH2 [24]. Therefore, the activation procedure (heating at bon atom. Strong bases would cause direct deprotonation of the
150 C) that we applied before catalytic reactions could induce par- methylene group, generating the carbanion that attacks the car-
tial dehydration. Another origin of the Lewis acid site could be the bonyl carbon atom of benzaldehyde. For amino-based catalysts, it
presence of defects. Since the UiO-66 MOFs are much stable than is generally accepted that the reaction proceeds via aldimine inter-
MOF-5 and IRMOF-3, one would not expected that the Zr MOFs con- mediates [45,59]. Two nucleophilic addition-elimination processes
tains a large density of defects of similar origin to those in the Zn are involved, as illustrated in Scheme 2a. The rst is the formation
MOFs. However, recent studies have demonstrated the presence of an aldimine intermediate (and water) from aldehyde and the
of linker vacancies (defects due to the missing of organic linkers amine group at the catalyst surface. The aldimine intermediate is
between Zr6 clusters) even in well-crystallized UiO-66 materials more basic than the original amine and also more reactive than the
[26,57,58]. The defects could generate open metal sites upon heat- original aldehyde [36]. Thus, the methylene compound is activated
ing. Whatever the origin of the acid sites is, the presence of both (deprotonated) by aldimine rather than by amine, and meanwhile
acid and base sites could explain the superior performance of UiO- the aldimine group is also activated by the proton transfer. There-
66-NH2 . fore, the second addition-elimination process is facilitated, which
Comparisons among ZrO2 , DMBDC-NH2 and their mixture may occurs between methylene and aldimine (rather than aldehyde)
give some information about the acidbase cooperation. As shown groups to give the product and to regenerate the amine catalyst.
Y. Yang et al. / Journal of Molecular Catalysis A: Chemical 390 (2014) 198205 205

Spectroscopic evidences for the formation of intermediate aldimine [8] M. Yoon, R. Srirambalaji, K. Kim, Chem. Rev. 112 (2012) 11961231.
have been recently reported for amino-tagged silicas and IRMOF-3 [9] P. Valvekens, F. Vermoortele, D. De Vos, Catal. Sci. Technol. 3 (2013) 14351445.
[10] A. Dhakshinamoorthy, M. Opanasenko, J. Cejka, H. Garca, Catal. Sci. Technol. 3
[45,60]. (2013) 25092540.
According to this mechanism, the formation of the aldimine [11] A. Corma, H. Garca, F.X. Llabrs i Xamena, Chem. Rev. 110 (2010) 46064655.
intermediate is important. We have studied the formation by reac- [12] B. Voit, Angew. Chem. Int. Ed. 45 (2006) 42384240.
[13] N.R. Shiju, A.H. Alberts, S. Khalid, D.R. Brown, G. Rothenberg, Angew. Chem. Int.
ting benzaldehyde with DMBDC-NH2 under the conditions used for Ed. 50 (2011) 96159619.
the Knoevenagel reaction. As can be seen from Fig. 6b, the forma- [14] R. Srirambalaji, S. Hong, R. Natarajan, M. Yoon, R. Hota, Y. Kim, Y.H. Ko, K. Kim,
tion of aldimine in the absence of any catalyst is much slower than Chem. Commun. 48 (2012) 1165011652.
[15] F.X. Llabrs i Xamena, F.G. Cirujano, A. Corma, Microporous Mesoporous Mater.
in the presence of UiO-66. The results clearly suggest the catalytic
157 (2012) 112117.
role of the Zr MOF for aldimine formation. The promotion is most [16] A. Dhakshinamoorthy, M. Alvaro, A. Corma, H. Garca, Dalton Trans. 40 (2011)
likely due to the activation of carbonyl groups by Zr sites. 63446360.
[17] S.M. Cohen, Chem. Rev. 112 (2012) 9701000.
Based on the above considerations, the cooperative acidbase
J. Ferrando-Soria, I. Luz, P. Serra-Crespo, E. Skupien, V.P. Santos,
[18] J. Juan-Alcaniz,
bifunctional catalysis of UiO-66-NH2 can be speculated as follows E. Pardo, F.X. Llabrs i Xamena, F. Kapteijn, J. Gascon, J. Catal. 307 (2013) 295.
(Scheme 2b). The Zr site serves to activate benzaldehyde so that [19] B.Y. Huang, S.Y. Gao, T.F. Liu, J. L, X. Lin, H.F. Li, R. Cao, ChemPlusChem 77
the amino group in close proximity the Zr site can readily attack (2012) 106112.
[20] Y.B. Huang, T. Ma, P. Huang, D.S. Wu, Z.J. Lin, R. Cao, ChemCatChem 5 (2013)
the carbonyl group to generate the aldimine intermediate, which 18771883.
serves not only as a base (the N atom) to extract proton from methy- [21] Y.B. Huang, Z.J. Lin, R. Cao, Chem. Eur. J. 17 (2011) 1270612712.
lene but also as an acid (the C atom) to couple with methylene. [22] Y.B. Huang, S.J. Liu, Z.J. Lin, W.J. Li, X.F. Li, R. Cao, J. Catal. 292 (2012) 111.
[23] J.H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga, K.P. Lillerud,
This synergic process is a little different from that proposed for J. Am. Chem. Soc. 130 (2008) 1385013851.
the cross-aldol condensation between heptanal and benzaldehyde [24] F. Vermoortele, R. Ameloot, A. Vimont, C. Serre, D. De Vos, Chem. Commun. 47
[24], where the reaction was supposed to occur between benz- (2011) 15211523.
[25] F. Vermoortele, M. Vandichel, B. Van de Voorde, R. Ameloot, M. Waroquier, V.
aldehyde activated by Zr(IV) and the heptanal methylene group Van Speybroeck, D.E. De Vos, Angew. Chem. Int. Ed. 51 (2012) 48874890.
activated by amine. [26] F. Vermoortele, B. Bueken, G. Le Bars, B. Van de Voorde, M. Vandichel,
K. Houthoofd, A. Vimont, M. Daturi, M. Waroquier, V. Van Speybroeck, C.
Kirschhock, D.E. De Vos, J. Am. Chem. Soc. 135 (2013) 1146511468.
4. Conclusions [27] J. Kim, S.N. Kim, H.G. Jang, G. Seo, W.S. Ahn, Appl. Catal. A 453 (2013) 175180.
[28] M.N. Timofeeva, V.N. Panchenko, J.W. Jun, Z. Hasan, M.M. Matrosova, S.H. Jhung,
In this work, we reported the use of UiO-66-NH2 as solid Appl. Catal. A 471 (2014) 9197.
[29] D. Sun, Y. Fu, W. Liu, L. Ye, D. Wang, L. Yang, X. Fu, Z. Li, Chem. Eur. J. 19 (2013)
catalyst for Knoevenagel condensation. It has been demonstrated 1427914285.
that the catalyst shows better performance in highly polar solvents [30] L. Shen, W. Wu, R. Liang, R. Lin, L. Wu, Nanoscale 5 (2013) 93749382.
such as DMF, DMSO and ethanol. It can efciently catalyze the [31] L. Shen, S. Liang, W. Wu, R. Liang, L. Wu, Dalton Trans. 42 (2013) 1364913657.
[32] J. Long, S. Wang, Z. Ding, S. Wang, Y. Zhou, L. Huang, X. Wang, Chem. Commun.
condensation reactions of aromatic aldehydes with cyanoacetate 48 (2012) 1165611658.
and malononitrile, the conversion easily reaching more than 90% [33] C. Wang, Z. Xie, K.E. deKraff, W. Lin, J. Am. Chem. Soc. 133 (2011) 1344513454.
under mild conditions within a short time (dependent on the [34] C.G. Silva, I. Luz, F.X. Llabrs i Xamena, A. Corma, H. Garca, Chem. Eur. J. 16
(2010) 1113311138.
substrates). The catalytic process is heterogeneous and shows [35] M.J. Climent, A. Corma, I. Domnguez, S. Iborra, M.J. Sabater, G. Sastre, J. Catal.
size effects, characteristic of a porous catalyst with active sites 246 (2007) 136146.
inside the pores. The catalyst can be recycled without losing its [36] I. Rodriguez, G. Sastre, A. Corma, S. Iborra, J. Catal. 183 (1999) 1423.
[37] S.K. Mohamed, A.A. Abdelhamid, A.M. Maharramov, A.N. Khalilov, A.V. Gur-
framework integrity and catalytic activity. Compared with the
banov, M.A. Allahverdiev, J. Chem. Pharm. 4 (2012) 17871793.
homogeneous amino precursor (2-aminoterephthalate) and the [38] G. Bartoli, M. Bosco, A. Carlone, R. Dalpozzo, P. Galzerano, P. Melchiorre, L.
isostructural amino-free MOF (UiO-66), the superior performance Sambri, Tetrahedron Lett. 49 (2008) 25552557.
[39] B.M. Choudary, M. Lakshmi Kantam, V. Neeraja, K. Koteswara Rao, F. Figueras,
of UiO-66-NH2 has been attributed to the acidbase bifunctional
L. Delmotte, Green Chem. 3 (2001) 257260.
character. It was proposed that the Zr site in close proximity to [40] D.J. Macquarrie, D.B. Jackson, Chem. Commun. (1997) 17811782.
the amino group facilitates the formation of aldimine interme- [41] A. Corma, S. Iborra, I. Rodrguez, F. Snchez, J. Catal. 211 (2002) 208215.
diates, which are the active species reacting with the methylene [42] S. Hasegawa, S. Horike, R. Matsuda, S. Furukawa, K. Mochizuki, Y. Kinoshita, S.
Kitagawa, J. Am. Chem. Soc. 129 (2007) 26072614.
compounds. This study further demonstrates that MOFs can offer [43] Y.K. Hwang, D.-Y. Hong, J.-S. Chang, S.H. Jhung, Y.-K. Seo, J. Kim, A.
extraordinary opportunity of achieving site-isolated acidbase Vimont, M. Daturi, C. Serre, G. Frey, Angew. Chem. Int. Ed. 47 (2008)
catalysts, which are desirable for more efcient and more facile 41444148.
[44] S. Neogi, M.K. Sharma, P.K. Bharadwaj, J. Mol. Catal. A: Chem. 299 (2009)
organic synthesis. Working along this line, we are exploring such 14.
catalysts for multi-component reactions. [45] J. Gascon, U. Aktay, M.D. Hernandezalonso, G.P.M. van Klink, F. Kapteijn, J. Catal.
261 (2009) 7587.
[46] A.R. Burgoyne, R. Meijboom, Chem. Lett. 143 (2013) 563571.
Acknowledgements [47] M. Opanasenko, A. Dhakshinamoorthy, M. Shamzhy, P. Nachtigall, M. Horcek,

H. Garca, J. Cejka, Catal. Sci. Technol. 3 (2013) 500507.
This work is supported by the National Science Foundation of [48] P. Serra-Crespo, E.V. Ramos-Fernandez, J. Gascon, F. Kapteijn, Chem. Mater. 23
(2011) 25652572.
China (NSFC nos. 21173083 and 91022017) and the Research Fund [49] U.P.N. Tran, K.K.A. Le, N.T.S. Phan, ACS Catal. 1 (2011) 120127.
for the Doctoral Program of Higher Education of China. [50] L.T.L. Nguyen, K.K.A. Le, H.X. Truong, N.T.S. Phan, Catal. Sci. Technol. 2 (2012)
521.
[51] S.-N. Kim, S.-T. Yang, J. Kim, J.-E. Park, W.-S. Ahn, Cryst. Eng. Comm. 14 (2012)
References 4142.
[52] A. Schaate, P. Roy, A. Godt, J. Lippke, F. Waltz, M. Wiebcke, P. Behrens, Chem.
[1] C. Wang, D. Liu, W. Lin, J. Am. Chem. Soc. 135 (2013) 1322213234. Eur. J. 17 (2011) 6643.
[2] H. Furukawa, K.E. Cordova, M. OKeeffe, O.M. Yaghi, Science 341 (2013) [53] S.J. Garibay, S.M. Cohen, Chem. Commun. 46 (2010) 77007702.
12304441230456. [54] B.R. Gelin, M. Karplus, J. Am. Chem. Soc. 26 (1975) 69967006.
[3] M.P. Suh, H.J. Park, T.K. Prasad, D.W. Lim, Chem. Rev. 112 (2012) 782835. [55] F.G. Bordwell, H.E. Fried, J. Org. Chem. 46 (1981) 43284331.
[4] K. Sumida, D.L. Rogow, J.A. Mason, T.M. McDonald, E.D. Bloch, Z.R. Herm, T.H. [56] W.N. Olmstead, F.G. Bordwell, J. Org. Chem. 45 (1980) 32993305.
Bae, J.R. Long, Chem. Rev. 112 (2012) 724781. [57] H. Wu, Y.S. Chua, V. Krungleviciute, M. Tyagi, P. Chen, T. Yildirim, W. Zhou, J.
[5] S. Xiang, Y. He, Z. Zhang, H. Wu, W. Zhou, R. Krishna, B. Chen, Nat. Commun. 3 Am. Chem. Soc. 135 (2013) 10525.
(2012), 956(19). [58] L. Valenzano, B. Civalleri, S. Chavan, S. Bordiga, M.H. Nilsen, S. Jakobsen, K.P.
[6] J.-R. Li, Y.-G. Ma, M.C. McCarthy, J. Sculley, J.-M. Yu, H.-K. Jeong, P.B. Balbuena, Lillerud, C. Lamberti, Chem. Mater. 23 (2011) 17001718.
H.-C. Zhou, Coord. Chem. Rev. 255 (2011) 17911823. [59] R. Cortese, D. Duca, Phys. Chem. Chem. Phys. 13 (2011) 1599516004.
[7] Z. Zhang, Y. Zhao, Q. Gong, Z. Li, J. Li, Chem. Commun. 49 (2013) 653661. [60] R. Wirz, D. Ferri, A. Baiker, Langmuir 22 (2006) 36983706.

Você também pode gostar