Sei sulla pagina 1di 7

Highly dispersed palladium nanoparticles supported on amino

functionalized metal-organic frameworks as an efcient and reusable


catalyst for Suzuki cross-coupling reaction
Reihaneh Kardanpour, Shahram Tangestaninejad
*
, Valiollah Mirkhani
*
,
Majid Moghadam
*
, Iraj Mohammadpoor-Baltork, Ahmad R. Khosropour,
Farnaz Zadehahmadi
Department of Chemistry, Catalysis Division, University of Isfahan, Isfahan 81746-73441, Iran
a r t i c l e i n f o
Article history:
Received 3 December 2013
Received in revised form
9 March 2014
Accepted 12 March 2014
Keywords:
Palladium nanoparticle
Metal-organic framework
SuzukieMiyaura cross-coupling reaction
a b s t r a c t
Palladium nanoparticles (Pd NPs) supported on amino functionalized UiO-66-NH
2
(UiO University of
Oslo) metal-organic framework was prepared using a direct anionic exchange method followed by
chemical reduction with sodium acetate in methanol. This nano palladium containing catalyst was
characterized by X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), N
2
adsorption,
inductively coupled plasma atomic emission spectroscopy (ICP-AES), eld emission scanning electron
microscopy (FE-SEM) and transmission electron microscopy (TEM). The Pd/UiO-66-NH
2
catalyst showed
excellent activity in SuzukieMiyaura cross-coupling reaction and reused several times without any
appreciable loss of activity even after ve consecutive times.
2014 Elsevier B.V. All rights reserved.
Introduction
Palladium-catalyzed SuzukieMiyaura cross-coupling reaction of
aryl halides with arylboronic acids have been accredited as one of
the most powerful and convenient synthetic tools available to
construct unsymmetrical biaryl compound [1]. Homogenous
palladium catalysts provide high turnover number and high activ-
ity but they often suffer from difculty of recycling and separation
from the products [2]. So it is favorable to develop heterogeneous
catalysts for Suzuki reactions to solve such problems [3]. However,
palladium nanoparticles as an important kind of heterogeneous
catalysts usually undergo aggregation and lose their catalytic ac-
tivity in the absence of a suitable stabilizer or protecting agent [4]. It
is a well-known fact that immobilization of Pd nanoparticles on
solid support makes the catalyst effective and recyclable and
minimizes the leaching of the particles.
Metal-organic frameworks (MOFs), a new family of organice
inorganic hybrid materials, consist of coordination centers boun-
ded by organic linkers. Due to their tunable cavities, various to-
pologies, porosity and extraordinary surface areas, MOFs have
emerged as promising functional materials for gas sorption or
storage [5], drug release [6], luminescence [7], optoelectronics [8,9],
chemical sensing [10] and catalysis [11e14].
In recent decades, a number of solid materials, such as polymers
[15], carbon structures [16], zeolites [17] and mesoporous silica [18]
have been successfully employed as catalyst supports. MOFs can be
used as supports for Pd NPs owing to their unique structural fea-
tures such as well-dened pore structure, narrow micropore dis-
tribution and ultimately large surface areas [19]. In fact, the pores of
MOFs can prevent migration and aggregation of the metal nano-
particles. Several reviews describe the use of metal-organic
frameworks (MOFs) as porous matrices to embed metal nano-
particles (MNPs) [20,21]. A recent review by Moon et al. summa-
rized different methods for preparation of metal NPs within MOF
materials [22]. Although there are several methods for loading
palladium nanoparticles onto porous MOFs using surface grafting
methods and solution inltration [23e25], because of the easy
agglomeration and leaching into solution, developing of an effec-
tive and facile method to incorporate Pd NPs on MOFs remains a big
challenge [14,19,26,27]. Recently amino functionalized metal-
organic frameworks have attracted tremendous attention for this
purpose. It seems that the uncoordinated amine groups of these
MOFs can stabilize Pd NPs and prevent their agglomeration. Huang
et al. reported the preparation of Pd NPs supported on amine-
functionalized MIL-53(Al)-NH
2
and MIL-53(Al) and their
* Corresponding authors. Tel.: 98 311 7932705; fax: 98 311 6689732.
E-mail addresses: stanges@sci.ui.ac.ir (S. Tangestaninejad), mirkhani@sci.ui.ac.ir
(V. Mirkhani), moghadamm@sci.ui.ac.ir, majidmoghadamz@yahoo.com
(M. Moghadam).
Contents lists available at ScienceDirect
Journal of Organometallic Chemistry
j ournal homepage: www. el sevi er. com/ l ocat e/ j organchem
http://dx.doi.org/10.1016/j.jorganchem.2014.03.012
0022-328X/ 2014 Elsevier B.V. All rights reserved.
Journal of Organometallic Chemistry 761 (2014) 127e133
application in Suzuki and Heck reactions [28,29]. Yamashita et al.
immobilize Pd NPs within the pores of MIL-125 MOF and its amine-
functionalized NH
2
-MIL-125 by photo-assisted and ion exchange
deposition methods and used them for hydrogen production at
ambient temperature [30]. They showed that the amine groups in
NH
2
-MIL-125 increased the H
2
-generating activity and acted as a
stabilizer of Pd NPs. Xu et al. successfully immobilized bimetallic
AuePd nanoparticles (NPs) in the MIL-101 MOF and ethylenedi-
amine (ED)-grafted MIL-101 (ED-MIL-101) using a simple liquid
impregnation method as a rst highly active MOF immobilized
metal catalysts for conversion of formic acid to hydrogen [31].
Lillerud et al. synthesized the rst example of a zirconium(IV)
dicarboxylate MOF named UiO-66 by a conventional solvothermal
method [32]. Recently Cohen et al. synthesized and characterized
amino functionalized UiO-66-NH
2
and introduced specialized func-
tionalities through a postsynthetic modication (PSM) approach
[33]. UiO-66-NH
2,
which has uncoordinated amine groups, is based
on a 3-D structure of zirconium-oxo clusters [33]. It has been estab-
lished that the amino group in UiO-66-NH
2
are chemically available,
as evidenced by the H/D exchange experiment, so this MOF is very
interesting in the eld of catalysis [34]. One of the important key
points for using of MOFs is their thermal and chemical stability. The
UiO-66-NH
2
contains a highly stable secondary building units (SBU),
comparable to the parent UiO-66 material, which leads to an
Fig. 1. XRD patterns of UiO-66-NH
2
samples: (a) activated UiO-66-NH
2
; (b) fresh Pd/
UiO-66-NH
2
and (c) Pd/UiO-66-NH
2
after ve catalytic cycles.
Fig. 2. FE-SEM image of: (a) activated UiO-66-NH
2
and (b) Pd/UiO-66-NH
2
. SEM-EDX spectrum of: (c) activated UiO-66-NH
2
and (d) Pd/UiO-66-NH
2
.
R. Kardanpour et al. / Journal of Organometallic Chemistry 761 (2014) 127e133 128
inimitable thermal stability (decomposition temperature is above
500

C) and a strong resistance to many chemicals such as water,
benzene, ethanol, and so on [35]. In this work, we have successfully
employed amino functionalized UiO-66-NH
2
for immobilization of
Pd NPs as a highly active catalyst in the SuzukieMiyaura cross-
coupling reaction. The amine groups of UiO-66-NH
2
can stabilize
palladium nanoparticles to avoid agglomeration [36].
Experimental
Catalyst preparation
Synthesis of UiO-66-NH
2
Preparation of UiO-66-NH
2
was performed based on the pro-
cedure reported by Cohen et al. [33]. In a typical reaction, ZrCl
4
(0.35 mmol) and 2-amino terephthalic acid (0.35 mmol) were
dissolved in dimethyl formamide (4 mL) and the mixture was
placed in a Teon-lined stainless steel autoclave. The mixture was
heated at 120

C for 24 h, and the resulting yellow powders were
then isolated by centrifugation and heated at 100

C for 1e2 h.
Activation was achieved by washing samples of UiO-66-NH
2
(w60 mg) with methanol (3 15 mL), followed by dispersing the
material in MeOH for 3 days. Finally, heating the material at 150

C
for 5 h activated the pores of UiO-66-NH
2
.
Synthesis of UiO-66-NH
2
containing Pd nanoparticles, Pd/UiO-66-
NH
2
A mixture of PdCl
2
(240 mg, 1.36 mmol) and NaCl (88 mg,
1.52 mmol) in methanol (8 mL) was stirred at roomtemperature for
Fig. 3. TEM images of Pd/UiO-66-NH
2
.
Fig. 4. The histogram of Pd nanoparticles distribution.
Fig. 5. BET adsorptionedesorption isotherms of: (a) UiO-66-NH
2
and (b) Pd/UiO-66-
NH
2
.
R. Kardanpour et al. / Journal of Organometallic Chemistry 761 (2014) 127e133 129
24 h and a clear orange solution was obtained. The ltrates were
diluted with methanol (40 mL) and freshly prepared UiO-66-NH
2
(1.5 g) was added to this solution. The mixture was then stirred at
60

C for 24 h. At the end of the reaction, the mixture was cooled to
room temperature and the resulting sample was reacted with so-
dium acetate (0.76 g, 9.28 mmol) for reducing the Pd at room
temperature for 1 h. The solid was ltered, washed with methanol
to remove the unreacted starting materials and then dried in vac-
uum to afford Pd/UiO-66-NH
2
catalyst as a yellow gray solid.
Physico-chemical measurements
The surface area, pore volume, and pore size distribution of the
support and the catalyst were determined by adsorptionedesorp-
tion of N
2
gas at 77 K with ASAP 2000 micromeritics instrument X-
ray diffraction (XRD) analysis was carried out a D8 Advanced Bruker
anode X-ray Diffractometer with Cu Ka (l 1.5406

A) radiation. The
X-ray photo-electron spectroscopy (XPS) measurements were
performed using a Gammadata-scienta ESCA200 hemispherical
analyzer equipped with an Al (Ka 1486.6 eV) X-ray source. The
ICP analyses were carried out by a PerkineElmer optima 7300 DV
spectrometer. GC analyses were performed using a Shimadzu GC-
16A gas chromatograph equipped with a ame ionization detec-
tor and a 2 m column packed with silicon DC-200 or Carbowax
20m. The transmission electron microscopy (TEM) was carried out
on a Philips CM10 instrument operating at 100 kV. The crystal size
and microscopic morphological features were obtained on a Hitachi
S-4700 eld emission-scanning electron microscope (FE-SEM).
General procedure for SuzukieMiyaura cross-coupling reaction
catalyzed by Pd/UiO-66-NH
2
Typically, a mixture of aryl halide (1 mmol), phenylboronic acid
(1.1 mmol), K
2
CO
3
(1.5 mmol) and the Pd/UiO-66-NH
2
catalyst
(10 mg, 0.25 mol%) in a 2:1 solution of DMF/H
2
O (3 mL) was stirred
at 60

C under air atmosphere. The reaction progress was moni-
tored by GC. After completion of the reaction, the catalyst was
separated by centrifugation and the products were extracted with
ethyl acetate (3 10 mL) and. The organic phase was washed with
water (2 10 mL), dried over anhydrous Na
2
SO
4
and evaporated.
The residue was recrystallized from ethyl acetate and ether (1:3) to
afford the pure product.
Catalyst recovery and reuse
In the recyclability experiment, the reaction of 4-iodoanisole
and phenylboronic acid was chosen as model reaction. At the end
of each reaction, the catalyst was recovered by centrifugation,
washed with water and DMF, respectively, and dried at 60

C under
vacuum and used with fresh 4-iodoanisole and phenylboronic acid.
Fig. 6. The XPS spectrum of: (a) Pd/UiO-66-NH
2
showing Pd 3d
5/2
and Pd 3d
3/2
binding
energies, (b) the elemental survey scan of Pd/UiO-66-NH
2
.
Table 1
Optimization of reaction conditions in the SuzukieMiyaura cross-coupling of 4-iodoanisole (Ia) with phenylboronic acid (PBA) catalyzed by Pd/UiO-66-NH
2
.
Entry Base Ia:PBA:base [mmol] Solvent
a
Catalyst amount (mol%) T (

C) Time (min) Yield (%)


b
1 K
2
CO
3
1:1.1:1.5 Toluene 0.25 60 40 55
2 K
2
CO
3
1:1.1:1.5 EtOH 0.25 60 80 68
3 K
2
CO
3
1:1.1:1.5 EtOH/H
2
O (1:1) 0.25 60 60 71
4 K
2
CO
3
1:1.1:1.5 DMF 0.25 60 90 68
5 K
2
CO
3
1:1.1:1.5 H
2
O 0.25 60 90 53
6 K
2
CO
3
1:1.1:1.5 DMF/H
2
O (1:1) 0.25 60 50 80
7 K
2
CO
3
1:1.1:1.5 DMF/H
2
O (2:1) 0.25 60 10 95
8 NEt
3
1:1.1:1.5 DMF/H
2
O (2:1) 0.25 60 60 40
9 Na
2
CO
3
1:1.1:1.5 DMF/H
2
O (2:1) 0.25 60 35 50
10 K
3
PO
4
1:1.1:1.5 DMF/H
2
O (2:1) 0.25 60 50 45
11 K
2
CO
3
1:1.1:1.5 DMF/H
2
O (2:1) 0.37 60 10 95
12 K
2
CO
3
1:1.1:1.5 DMF/H
2
O (2:1) 0.16 60 10 69
13 K
2
CO
3
1:1:1 DMF/H
2
O (2:1) 0.25 60 30 80
14 K
2
CO
3
1:1.1:1.3 DMF/H
2
O (2:1) 0.25 60 30 85
15 K
2
CO
3
1:1.1:1.5 DMF/H
2
O (2:1) 0.25 25 10 35
16 K
2
CO
3
1:1.1:1.5 DMF/H
2
O (2:1) 0.25 40 10 62
17 K
2
CO
3
1:1.1:1.5 DMF/H
2
O (2:1) 0.25 80 10 95
a
Reaction was performed using 3 mL of solvent.
b
GC yield.
R. Kardanpour et al. / Journal of Organometallic Chemistry 761 (2014) 127e133 130
For determination of Pd leaching during the reaction, the ltrate
was washed with water (5 mL) and DMF (5 mL), and the liquid
phase was analyzed by inductively coupled plasma (ICP).
Results and discussion
Preparation and characterization of Pd/UiO-66-NH
2
The UiO-66-NH
2
was synthesized under solvothermal condi-
tions by combining equimolar amounts of ZrCl
4
with 2-amino-1,4-
benzenedicarboxylate (NH
2
-BDC) as previously reported [33].
During the preparation of this paper, we found a new published
paper about the synthesis of Pd/UiO-66-NH
2
nanocomposite and its
application for reduction of aqueous Cr(VI) under visible light
illumination [37]. They immobilized Pd NPs in UiO-66(NH
2
), by
hydrothermal method in a Teon-lined stainless-steel autoclave at
180

C in the presence of poly vinylpyrrolidone (PVP), PdCl
2
and
NaI.
In this work, amino functionalized MOF was used as support and
Pd NPs were immobilized on it by a direct ion-exchange method
followed by reduction [38]. Recently, Yaghi et al. showed that the
amine groups in the UiO-66-NH
2
exist as a mixture of amino and
ammonium chloride [39]. In fact hydrolysis of ZrCl
4
produces HCl
and NH
3

Cl BDC salt moiety is produced during MOF formation. The


chloride anions in the UiO-66-NH
2
were exchanged with [PdCl
4
]
2
precursors and the supported Pd NPs were obtained by reduction
with sodium acetate in methanol at roomtemperature [38]. The Pd
content of the supported catalyst, measured by ICP, revealed that
the amount of Pd in the Pd/UiO-66-NH
2
is about 2.65%.
The prepared catalyst was characterized by XRD analysis. As can
be seen, after loading of Pd, there is almost no apparent loss of
crystallinity according to the XRD pattern which indicates that the
basic lattice structure of UiO-66-NH
2
was well maintained (Fig. 1).
The sharp peaks indicate the excellent crystallinity of the frame-
work. The XRD pattern did not exhibit the characteristic peak of Pd
(111) at 2q 40.1

. It is maybe due to the low Pd loading and small


diameter of palladium nanoparticles [40]. FE-SEM image clearly
showed that the small cubic inter-grown crystals of UiO-66-NH
2
have diameters in the range of 80e110 nm. There is no signicant
change in the morphologies of the crystals upon immobilization of
Pd nanoparticles (Fig. 2a and b). The energy dispersive X-ray (EDX)
results, obtained from SEM analysis for the UiO-66-NH
2
and Pd/
UiO-66-NH
2
(Fig. 2c and d) clearly show the presence of Pd nano-
particles in the Pd/UiO-66-NH
2
catalyst. The TEM images showed
that Pd nanoparticles exist as dark-gray areas, with a mean diam-
eter of 5.28 0.5 nm, which are easily recognized from the sur-
rounding light gray regions of UiO-66-NH
2
(Fig. 3). As can be seen,
the Pd NPs were homogeneously dispersed on the surface of the
UiO-66-NH
2
support and no obvious aggregation was observed.
Also, the histogram of the Pd nanoparticles size distribution, esti-
mated by TEM image, is shown in Fig. 4. On the basis of these re-
sults, the Pd nanoparticles are too large to enter into the pores of
UiO-66-NH
2
and attached on the surface of the support. The
presence of amino group on the linker has been proven to be
benecial for the stabilization of Pd species by electrostatic in-
teractions [41]. The N
2
adsorption isothermmeasurement indicated
that UiO-66-NH
2
exhibited a BrunauereEmmetteTeller (BET) sur-
face area about 730 m
2
g
1
(Fig. 5a). Upon immobilization of Pd
nanoparticles, the specic surface area of Pd/UiO-66-NH
2
decreased to 441 m
2
g
1
(Fig. 5b) which can be attributed to the
blocking of UiO-66 cavities by surface located palladium NPs.
In the X-ray photoelectron spectroscopy (XPS) of Pd/UiO-66-
NH
2
, the 3d
5/2
and 3d
3/2
peaks of the Pd
0
appear at 336.05 and
341.52 eV, respectively. Due to the interaction between the Zr and
Pd, these peaks appear at higher energy (Fig. 6a) [42]. It is note-
worthy to note that the two additional peaks in Fig. 6a, are ascribed
to the Zr 3p
1/2
and Zr 3p
3/2
, respectively [37]. No obvious peak of
Pd
2
is observed, indicating that palladium is in the reduced form.
The peaks corresponding to oxygen, carbon, nitrogen, chlorine,
zirconiumand palladiumare also clearly observed in XPS elemental
survey of the catalyst (Fig. 6b).
The heterogeneous Pd/UiO-66-NH
2
catalyst was applied in the
SuzukieMiyaura coupling of phenylboronic acid and 4-iodoanisole
as model reaction and the reaction parameters such as solvent, base
types, temperature and molar ratios of substrates were optimized.
The results are summarized in Table 1. Initially, the model reaction
was carried out in different single and mixed solvents. Among
them, DMF/H
2
O (2:1) gave the best results (entries 1e7). Then
different organic and inorganic bases such as NEt
3
, K
3
PO
4
, Na
2
CO
3
and K
2
CO
3
were used and it was found that K
2
CO
3
is the most
effective base (entries 7e10). The amount of catalyst was also
optimized in which in the presence of 0.25 mol% (10 mg) of Pd/UiO-
66-NH
2
the highest yield was observed (entry 7, 11 and 12). The
ratio of starting materials was also investigated and the highest
yield was obtained with phenyl iodide (1 mmol), phenylboronic
acid (1.1 mmol) and K
2
CO
3
(1.5 mmol) (entries 7, 13 and 14). The
reaction temperature was also optimized and 60

C was chosen as
reaction temperature (entries 7 and 15e17).
Pd
0
Ar-X
Oxidative-addition
[Ar-Pd
II
-X]
X-B(OH)
3
-
Transmetallation
Ar-Pd
0
-Ar'
Ar-Ar'
Reductive-elimination
X= I, Br, Cl, ...
Ar'-B(OH)
2
+ H
2
O
Ar'-B(OH)
3
-
Base
1
2
3
4
5
Fig. 7. Catalytic cycle for the SuzukieMiyaura cross-coupling reaction.
Table 2
SuzukieMiyaura coupling reaction of different aryl halides with phenylboronic acid
catalyzed by Pd/UiO-66-NH
2
.
a
Entry R X Time (min) Yield (%)
b
TOF (h
1
)
1 H I 10 92 2190.5
2 4-MeO I 10 95 2261.9
3 4-Me I 10 93 2214.3
4 4-Ac I 13 90 1666.7
5 H
c
Cl 41 80 467.8
6 4-CHO
c
Cl 60 85 340
7 H Br 30 90 697.7
8 4-Ac Br 32 93 699.2
9 4-CHO Br 35 91 623.3
a
Reaction conditions: Aryl halide (1 mmol), phenylboronic acid (1.1 mmol),
K
2
CO
3
(1.5 mmol), Pd/UiO-66-NH
2
(0.01 g, 0.25 mol% Pd), H
2
O/DMF (1 mL/2 mL) at
60

C.
b
GC yield.
c
The reaction was performed at 80

C.
R. Kardanpour et al. / Journal of Organometallic Chemistry 761 (2014) 127e133 131
The scope and generality of this method was investigated using
different aryl aldehydes. In this manner, different substituted
phenyl chloride, bromide and iodides were reacted with phenyl-
boronic acid in the presence of 0.25 mol% (10 mg) of Pd/UiO-66-
NH
2
using K
2
CO
3
as base and DMF/H
2
O (2:1) as solvent (Table 2).
The experimental results showed that reactions proceeded
extraordinarily well and electronic properties of the substituents
on the aromatic rings had no important effect on the reaction.
However, as expected, aryl iodides were more reactive than aryl
bromides and aryl chlorides in this catalytic system. It is important
to note that aryl chlorides are cheaper and more readily available
but less reactive than aryl iodides and bromides. Due to lower
reactivity, the coupling reactions with aryl chlorides have been
generally investigated at higher temperatures (Table 2, entries 5, 6)
[43e48].
The catalytic process is as following: The rst step is the
oxidative-addition of palladium 1 to the halide 2 to form the
organo-palladium species 3. Arylboronic acid reacts with water in
the presence of base to produce the arylboronate complex. Reaction
of organo-palladium species 3 with the arylboronate complex via a
transmetalation reaction forms the organopalladium species 4.
Reductive-elimination of the desired product 5 restores the original
palladium catalyst 1 (Fig. 7).
The results of this work were compared with some results re-
ported in the SuzukieMiyaura cross-coupling reaction catalyzed by
MOF supported Pd nanoparticles. As can be seen in Table 3, this
catalytic system is superior to the others.
The recycling and reusability of a catalyst is very important for
industrial uses. Thus, the reusability of Pd/UiO-66-NH
2
was studied
in the SuzukieMiyaura cross-coupling of 4-iodoanisole with phe-
nylboronic acid under the optimal reaction conditions (Table 4). In
this manner, after each catalytic cycle, the catalyst was ltered,
washed with water and DMF, respectively, and dried at 60

C. The
recovered catalyst was used with fresh starting materials. The re-
sults showed that the catalyst could be reused for ve consecutive
cycles. The nature of the recovered catalyst was monitored by XRD
analysis. The XRD patterns indicated that the basic lattice structure
of UiO-66-NH
2
was not altered after ve cycles (Fig. 1c). The
amount of Pd leaching was determined by ICP analysis, which
showed that only small amounts of Pd is leached in two rst runs
(Table 4). These results showed that UiO-66-NH
2
is a good host for
distribution of Pd NPs.
Conclusions
In conclusion, a heterogeneous catalyst containing highly
dispersed Pd nanoparticles using functionalized UiO-66-NH
2
as
host compound for the SuzukieMiyaura cross-coupling reactions is
reported. The NH
2
groups on the organic linkers lead to high
dispersion of Pd nanoparticles on the support. The catalyst is of
high catalytic activity and reusability in the CeC coupling reactions.
Acknowledgment
We are thankful to the University of Isfahan for nancial support
of this work.
References
[1] R. Martin, S.L. Buchwald, Acc. Chem. Res. 41 (2008) 1461e1473.
[2] V. Calo, A. Nacci, A. Monopoli, P. Cotugno, Angew. Chem. 121 (2009) 6217.
Angew. Chem. Int. Ed. 48 (2009) 6101e6103.
[3] C. Evangelisti, N. Panziera, P. Pertici, G. Vitulli, P. Salvadori, C. Battocchio,
G. Polzonetti, J. Catal. 262 (2009) 287e293.
[4] C. Evangelisti, N. Panziera, A. DAlessio, L. Bertinetti, M. Botavina, G. Vitulli,
J. Catal. 272 (2010) 246e252.
[5] U. Mueller, M. Schubert, F. Teich, H. Puetter, K. Schierle-Arndt, J. Pastr,
J. Mater. Chem. 16 (2006) 626e632.
[6] P. Horcajada, C. Serre, M. Vallet-Reg, M. Sebban, F. Taulelle, G. Frey, Angew.
Chem. Int. Ed. 45 (2006) 5974e5978.
[7] M.D. Allendorf, C.A. Bauer, R.K. Bhakta, R.J.T. Louk, Chem. Soc. Rev. 38 (2009)
1330e1352.
[8] A. Schaate, M. Schulte, M. Wiebcke, A. Godt, P. Behrens, Inorg. Chim. Acta 362
(2009) 3600e3606.
[9] L. Jiang, Z.X. Li, Y. Wang, G.D. Feng, W.X. Zhao, K.Z. Shao, C.Y. Sun, L.J. Li,
Z.M. Su, Inorg. Chem. Commun. 14 (2011) 1077e1081.
[10] G. Lu, J.T. Hupp, J. Am. Chem. Soc. 132 (2010) 7832e7833.
[11] M.H. Alkordi, Y. Liu, R.W. Larsen, J.F. Eubank, M. Eddaoudi, J. Am. Chem. Soc.
130 (2008) 12639e12641.
[12] D. Jiang, T. Mallat, F. Krumeich, A. Baker, J. Catal. 257 (2008) 390e395.
[13] S. Horike, M. Dinc, K. Tamaki, J.R. Long, J. Am. Chem. Soc. 130 (2008) 5854e
5855.
[14] J.Y. Lee, O.K. Farha, J. Roberts, K.A. Scheidt, S.T. Nguyen, J.T. Hupp, Chem. Soc.
Rev. 38 (2009) 1450e1459.
[15] J.C. Garcia-Martinez, R. Lezutekong, R.M. Crooks, J. Am. Chem. Soc. 127 (2005)
5097e5103.
[16] G.M. Scheuermann, L. Rumi, P. Steurer, W. Bannwarth, R. Mulhaupt, J. Am.
Chem. Soc. 131 (2009) 8262e8270.
[17] M. Dams, L. Drijkoningen, B. Pauwels, G. van Tendeloo, D.E. der Vos,
P.A. Jacobs, J. Catal. 209 (2002) 225e236.
[18] S.M. Quarrie, B. Nohair, J.H. Horton, S. Kaliaguine, C.M. Crudden, J. Phys. Chem.
C 114 (2010) 57e64.
[19] M. Meilikhov, K. Yusenko, D. Esken, S. Turner, G.V. Tendeloo, R.A. Fischer, Eur.
J. Inorg. Chem. (2010) 3701e3714.
[20] A. Dhakshinamoorthy, H. Garcia, Chem. Soc. Rev. 41 (2012) 5262e5284.
[21] H.-L. Jiang, Q. Xu, Chem. Commun. 47 (2011) 3351e3370.
[22] H.R. Moon, D.-W. Lim, M.P. Suh, Chem. Soc. Rev. 42 (2013) 1807e1824.
[23] S. Opelt, S. Trk, E. Dietzsch, A. Henschel, S. Kaskel, E. Klemm, Catal. Commun.
9 (2008) 1286e1290.
[24] W. Wang, Y. Li, R. Zhang, D. He, H. Liu, S. Liao, Catal. Commun. 12 (2011) 875e
879.
[25] D. Esken, X. Zhang, O.I. Lebedev, F. Schroder, R.A. Fischer, J. Mater. Chem. 19
(2009) 1314e1319.
[26] L. Ma, C. Abney, W. Lin, Chem. Soc. Rev. 38 (2009) 1248e1256.
[27] A. Corma, H. Garca, F.X.L.I. Xamena, Chem. Rev. 110 (2010) 4606e4655.
[28] Y. Huang, Z. Zheng, T. Liu, J. Lu, Z. Lin, H. Li, R. Cao, Catal. Commun. 14 (2011)
27e31.
[29] Y. Huang, Sh Gao, T. Liu, J. Lu, Z. Lin, H. Li, R. Cao, Chem. Plus. Chem. 77 (2012)
106e112.
[30] M. Martis, K. Mori, K. Fujiwara, W.-S. Ahn, H. Yamashita, J. Phys. Chem. C 117
(2013) 22805e22810.
[31] X. Gu, Z.H. Lu, H.L. Jiang, T. Akita, Q. Xu, J. Am. Chem. Soc. 133 (2011) 11822e
11825.
[32] J.H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga,
K.P. Lillerud, J. Am. Chem. Soc. 130 (2008) 13850e13851.
[33] S.J. Garibay, S.M. Cohen, Chem. Commun. 46 (2010) 7700e7702.
[34] M. Kandiah, M.H. Nilsen, S. Usseglio, S. Jakobsen, U. Olsbye, M. Tilset, C. Larabi,
E.A. Quadrelli, F. Bonino, K.P. Lillerud, Chem. Mater. 22 (2010) 6632e6640.
Table 3
Comparison of the results obtained in the SuzukieMiyaura coupling reaction of
phenyl iodide with phenylboronic acid catalyzed by Pd/UiO-66-NH
2
with those
obtained by the recently reported catalysts.
Catalyst T

C TOF (h
1
) Ref.
Pd/MIL-53(Al)-NH
2
40 396 [28]
Pd/MIL-53(Al) 40 98 [28]
IRMOF-3-PI-Pd 80 2037 [49]
Pd/UiO-66-NH
2
60 2190.5 Present work
Table 4
Recycling of the Pd/UiO-66-NH
2
catalyst in the SuzukieMiyaura cross-coupling of 4-
iodoanisole with phenylboronic acid in 10 min.
a
Run Yield (%)
b
Pd leached [%]
c
TOF (h
1
)
1 95 3 2261.9
2 90 1 2142.9
3 88 0 2095.2
4 88 0 2095.2
5 88 0 2095.2
a
Reaction conditions: 4-iodoanisole (1 mmol), phenylboronic acid (1.1 mmol),
K
2
CO
3
(1.5 mmol), Pd/UiO-66-NH
2
(0.01 g, 0.25 mol% Pd), H
2
O/DMF (1 mL/2 mL) at
60

C.
b
GC yield.
c
Determined by ICP.
R. Kardanpour et al. / Journal of Organometallic Chemistry 761 (2014) 127e133 132
[35] S.S. Kaye, A. Dailly, O.M. Yaghi, J.R. Long, J. Am. Chem. Soc. 129 (2007) 14176e
14177.
[36] Y.K. Hwang, M.I. Hong, J.S. Chang, S.H. Chung, Y.K. Seo, J. Kim, A. Vimont,
M. Daturi, C. Serre, G. Frey, Angew. Chem. Int. Ed. 47 (2008) 4144e4148.
[37] L. Shen, W. Wu, R. Liang, R. Lin, L. Wu, Nanoscale 5 (2013) 9374e9382.
[38] A. Landarani Isfahani, I. Mohammadpoor-Baltork, V. Mirkhani,
A.R. Khosropour, M. Moghadam, S. Tangestaninejad, R. Kia, Adv. Synth. Catal.
355 (2013) 957e972.
[39] W. Morris, C.J. Doonan, O.M. Yaghi, Inorg. Chem. 50 (2011) 6853e6855.
[40] M.S. El-Shall, V. Abdelsayed, A.E.R.S. Khder, H.M.A. Hassan, H.M. El-Kaderi,
T.E. Reich, J. Mater. Chem. 19 (2009) 7625e7631.
[41] W. Kleist, M. Maciejewski, A. Baiker, Thermochim. Acta 499 (2010) 71.
[42] B. Yuan, Y. Pan, Y. Li, B. Yin, H. Jiang, Angew. Chem. 122 (2010) 4148e4152.
Angew. Chem. Int. Ed. 49 (2010) 4054e4058.
[43] C.L. Deng, S.M. Guo, Y.X. Xie, J.H. Li, Eur. J. Org. Chem. (2007) 1457e1462.
[44] I.D. Kostas, B.R. Steele, A. Terzis, S.V. Amosova, A.V. Martynov, N.A. Makhaeva,
Eur. J. Inorg. Chem. (2006) 2642e2646.
[45] Q. Luo, S. Eibauer, O. Reiser, J. Mol. Catal. A Chem. 268 (2007) 65e69.
[46] I.D. Kostas, F.J. Andreadaki, D. Kovala-Demertzi, C. Prentjas, M.A. Demertzis,
Tetrahedron Lett. 46 (2005) 1967e1970.
[47] T. Mino, Y. Shirae, M. Sakamoto, T. Fujita, J. Org. Chem. 70 (2005) 2191e2194.
[48] A. Corma, H. Garcia, A. Leyva, J. Catal. 240 (2006) 87e99.
[49] D. Saha, R. Sen, T. Maity, S. Koner, Langmuir 29 (2013) 3140e3151.
R. Kardanpour et al. / Journal of Organometallic Chemistry 761 (2014) 127e133 133

Potrebbero piacerti anche