You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/281091420

Oxidative cross-dehydrogenative coupling of amines and alpha-carbonyl


aldehydes over heterogeneous Cu-MOF-74 catalyst: a ligand- and base-free
approach

Research · August 2015


DOI: 10.13140/RG.2.1.3053.2329

CITATIONS READS

0 374

1 author:

Dang Huynh Giao


Can Tho University
22 PUBLICATIONS   223 CITATIONS   

SEE PROFILE

All content following this page was uploaded by Dang Huynh Giao on 19 August 2015.

The user has requested enhancement of the downloaded file.


Accepted Manuscript

Title: Oxidative cross-dehydrogenative coupling of amines


and ␣-carbonyl aldehydes over heterogeneous Cu-MOF-74
catalyst: a ligand- and base-free approach

Author: Thanh Truong Giao H. Dang Nam V. Tran Ngoc T.


Truong Dung T. Le Nam T.S. Phan

PII: S1381-1169(15)30033-9
DOI: http://dx.doi.org/doi:10.1016/j.molcata.2015.07.022
Reference: MOLCAA 9573

To appear in: Journal of Molecular Catalysis A: Chemical

Received date: 11-6-2015


Revised date: 16-7-2015
Accepted date: 18-7-2015

Please cite this article as: Thanh Truong, Giao H.Dang, Nam V.Tran, Ngoc
T.Truong, Dung T.Le, Nam T.S.Phan, Oxidative cross-dehydrogenative coupling of
amines and rmalpha-carbonyl aldehydes over heterogeneous Cu-MOF-74 catalyst:
a ligand- and base-free approach, Journal of Molecular Catalysis A: Chemical
http://dx.doi.org/10.1016/j.molcata.2015.07.022

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Oxidative cross-dehydrogenative coupling of amines and-carbonyl aldehydes over

heterogeneous Cu-MOF-74 catalyst: a ligand- and base-free approach

Thanh Truong*, Giao H. Dang, Nam V. Tran, Ngoc T. Truong, Dung T. Le, Nam T. S. Phan*

Department of Chemical Engineering, HCMC University of Technology, VNU-HCM,

268 Ly Thuong Kiet, District 10, Ho Chi Minh City, Viet Nam
*
Email: tvthanh@hcmut.edu.vn, ptsnam@hcmut.edu.vn

Ph: (+84 8) 38647256 ext. 5681 Fx: (+84 8) 38637504

Research Highlights

 An efficient heterogeneous catalyst due to its exceptional large pore aperture.

 Ligand- and base-free conditions with broad reaction scope.

 The catalyst can be facilely recovered and reused.

 Graphic abstract

Abstract

A crystalline porous metal-organic framework Cu-MOF-74 was synthesized and its properties

were examined by a host of techniques, including X-ray powder diffraction (XRD), scanning
electron microscopy (SEM), transmission electron microscopy (TEM), thermogravimetric

analysis (TGA), Fourier transform infrared (FT-IR), atomic absorption spectrophotometry

(AAS), and nitrogen physisorption measurements. The Cu-MOF-74 exhibited high activity in

oxidative cross-dehydrogenative coupling of amines and -carbonyl aldehydes to form -

ketoamides using air oxidant. High yields were achieved in the presence of a catalytic amount of

the Cu-MOF without any added ligand and base. The activity of Cu-MOF-74 was showed to be

higher than that of other frequently reported Cu-MOFs and comparable to common copper salts.

Large pore aperture was proposed to strongly contribute to Cu-MOF-74 activity. Furthermore,

leaching test indicated no contribution from homogeneous leached species and the catalyst could

be recovered and reused several times without a significant degradation in catalytic activity.

Keywords: Metal-organic framework;-ketoamides; oxidative couplings; heterogeneous

catalyst; Cu-MOF-74.

1. Introduction

-Ketoamides have emerged as valuable structural motifs in a variety of pharmaceutical

candidates and agrochemicals [1-2]. They were frequently used as versatile and important

intermediates in numerous functional group transformations [1-4]. Conventionally, these

structures were achieved by several synthetic strategies, including the condensation of -keto

acids with amines [5], the oxidative coupling of aliphatic aldehydes with isocyanides [6], the

oxidation of -hydroxyamides and -aminoamides [7], the double carbonylative amination of

aryl halides [8, 9], and the reaction of carbamoylstannane and carbamoysilane with acid

chlorides [10]. Recently, aerobic oxidative cross-dehydrogenative coupling transformations have

exhibited several advantages in the construction of -ketoamides. Jiao and Zhang previously
reported an approach to -ketoamides by the CuBr2-catalyzed oxidative amidation-

diketonization reaction of terminal alkynes using oxygen as the oxidant and as a reactant via

dioxygen activation [11]. Jiao and co-workers subsequently changed the synthetic strategy to the

CuBr-catalyzed aerobic oxidative coupling of aryl acetaldehydes with anilines [1]. Ji and Du also

reported the CuI-catalyzed direct oxidative transformation to form -ketoamides from aryl

methyl ketones, amines, and molecular oxygen via a one-pot tandem process [4]. To overcome

the existing drawbacks, Jiao and co-workers performed the CuBr-catalyzed aerobic oxidative

cross-dehydrogenative coupling of amines with -carbonyl aldehydes in the presence of 2,2’-

bipyridine ligand and pyridine base [12]. To achieve greener approaches to -ketoamides, in

terms of the ease of handling, simple workup, recyclability and reusability, heterogeneous

catalysts should be investigated to replace conventional copper salts.

Metal-organic frameworks (MOFs), a new class of crystalline porous materials, have recently

exhibited potential applications in a variety of areas, including energy technologies, gas storage,

gas separation, sensors, optics, biomedicine, and catalysis [13-17]. The combination of organic

linkers and metal-connecting points in the framework leads to several interesting properties for

MOFs, such as high surface areas, high porosity, well-defined structures, structural diversity, and

the ability to tune the pore size as well as the surface hydrophobicity/hydrophilicity [13, 14, 18-

25]. Although studies on applications of MOFs in catalysis are relatively lagging behind other

fields, this topic should be undoubtedly a promising area that will attract further research in the

near future [26-28]. Technically, the catalytic active sites could be from the organic linkers and

the metals in MOF structure or from the synergistic effects of both components [27, 29-35].

During the last few years, many MOFs have been explored as heterogeneous catalysts or catalyst
supports in organic synthesis [36-54]. One of the challenges in MOFs catalysis is to make

crystals whose pore apertures are of a size suitable for the inclusion of large organic molecules.

Recently, the discovery of MOF-74 with paddle wheel structure, open metal sites, high

coordinate metal ion, and exceptionally large pore apertures offered a promising type of MOFs

for catalysis [19b, 53b, c]. By inspiring the frequent use of Cu-MOFs in catalysis [52-61], herein,

we report an efficient route for -ketoamides synthesis via oxidative cross-dehydrogenative

coupling of amines and -carbonyl aldehydes over Cu-MOF-74 catalyst using atmospheric

oxygen oxidant. Unlike homogeneous reported protocols, excellent yields were still obtained in

the absence of ligand and base. To the best of our knowledge, the aerobic oxidative cross-

coupling approaching to -ketoamides under heterogeneous catalysis condition was not

previously mentioned in the literature.

2. Experimental

2.1. Materials and instrumentation

All reagents and starting materials were obtained commercially from Sigma-Aldrich and Merck

and were used as received without any further purification unless otherwise noted. Nitrogen

physisorption measurements were conducted using a Micromeritics 2020 volumetric adsorption

analyzer system. Samples were pretreated by heating under vacuum at 150 oC for 3 h. A Netzsch

Thermoanalyzer STA 409 was used for thermogravimetric analysis (TGA) with a heating rate of

10 oC/min under nitrogen atmosphere. X-ray powder diffraction (XRD) patterns were recorded

using a Cu Kα radiation source on a D8 Advance Bruker powder diffractometer. Scanning

electron microscopy studies were conducted on a S4800 Scanning Electron Microscope (SEM).

Transmission electron microscopy studies were performed using a JEOL JEM 1400
Transmission Electron Microscope (TEM) at 100 kV. The Cu-MOF-74 sample was dispersed on

holey carbon grids for TEM observation. Elemental analysis with atomic absorption

spectrophotometry (AAS) was performed on an AA-6800 Shimadzu. Fourier transform infrared

(FT-IR) spectra were obtained on a Nicolet 6700 instrument, with samples being dispersed on

potassium bromide pallets.

Gas chromatographic (GC) analyses were performed using a Shimadzu GC 2010-Plus equipped

with a flame ionization detector (FID) and an SPB-5 column (length = 30 m, inner diameter =

0.25 mm, and film thickness = 0.25 μm). The temperature program for GC analysis held samples

at 100 oC for 1 min; heated them from 100 to 280 oC at 40 oC/min; held them at 280 oC for 2

min. Inlet and detector temperatures were set constant at 280 oC. Diphenyl ether was used as an

internal standard to calculate reaction conversions. GC-MS analyses were performed using a

Hewlett Packard GC-MS 5972 with a RTX-5MS column (length = 30 m, inner diameter = 0.25

mm, and film thickness = 0.5 μm). The temperature program for GC-MS analysis heated samples

from 60 to 280 oC at 10 oC/min and held them at 280 oC for 10 min. Inlet temperature was set

constant at 280 oC. MS spectra were compared with the spectra gathered in the NIST library. The
1
H NMR and 13C NMR were recorded on Bruker AV 500 spectrometers using residual solvent

peak as a reference.

2.2. Synthesis of the metal-organic framework Cu-MOF-74

In a typical preparation [62], a solid mixture of 2,5-dihydroxyterephthalic acid (H2dhtp, 0.186 g,

0.97mmol), and Cu(NO3)2.3H2O (0.500g, 2.07 mmol) was dissolved in a mixture of N,N’-

dimethylformamide (DMF, 20 mL), and water (1 mL). The resulting solution was distributed to
three 10 mL vials. The vials were then heated at 85 oC in an isothermal oven for 18 h. After

cooling the vials to room temperature, the solid product was removed by decanting with mother

liquor and washed in DMF (3 x 20 mL) for 3 days. Solvent exchange was carried out with

methanol (3 x 20 mL) at room temperature for 3 days. The material was then evacuated under

vacuum at 150 oC for 5 h, yielding 0.260 g of Cu-MOF-74 in the form of reddish black crystals

(60 % based on H2dhtp).

2.3. Catalytic studies

In a typical experiment, a mixture of phenylglyoxal monohydrate (0.11 mL, 1.0 mmol),

pyrrolidine (0.246 mL, 3.0 mmol), and diphenyl ether (0.07 mL) as an internal standard in

toluene (4 mL) was added into a 25 mL flask containing the pre-determining amount of Cu-

MOF-74 catalyst. The catalyst amount was calculated with respect to the copper/phenylglyoxal

molar ratio. The reaction mixture was stirred at 80 oC for 120 min. Reaction conversion was

monitored by withdrawing aliquots from the reaction mixture at different time intervals,

quenching with aqueous KOH (1 mL). The organic components were then extracted into ethyl

acetate (2 mL), dried over anhydrous Na2SO4, analyzed by GC with reference to diphenyl ether.

The product identity was further confirmed by GC-MS and 1H NMR. To investigate the

recyclability of Cu-MOF-74, the catalyst was separated from the reaction mixture by simple

filtration, washed with copious amounts of DMF and methanol, dried 150oC under vacuum in 2

h, and reused if necessary. For the leaching test, a catalytic reaction was stopped after 20 min,

analyzed by GC, and filtered to remove the solid catalyst. The reaction solution was then stirred

for a further 100 min. Reaction progress, if any, was monitored by GC as previously described.
3. Results and discussion

Scheme 1. The oxidative coupling of pyrrolidine and phenylglyoxal using Cu-MOF-74 catalyst.

The characterization results of XRD, SEM, TEM, TGA, FT-IR, AAS, and nitrogen physisorption

measurements were in good agreement with previous reports (Fig. S14 – Fig. S20) [63]. Briefly,

powder X-ray diffraction pattern showed the typical reflections of MOF-74 phase. The basically

type-1 adsorption/desorption isotherm indicated the permanent micro-porosity with Brunauer-

Emmett-Teller specific surface area of 1064 m2/g, a pore volume of 0.48 cm3/g and an average

pore diameter of about 14 Ao. Scanning electron microscopy analysis revealed homogeneity with

respect to needle-shaped crystals. Thermal gravimetric analysis (TGA) of activated Cu-MOF-74

shows high thermal stability (>360 °C) and the measured mass percent of residue CuO is

consistent with the EA data. ICP-MS provided 27.7 % copper content which is close to the

calculated value of 28.2 %. Finally, FT-IR spectra of Cu-MOF-74 indicated the presence of

bonded carboxylate organic linkers.

In optimization studies, the Cu-MOF-74 was assessed for its catalytic activity in the aerobic

oxidative cross-dehydrogenative coupling of pyrrolidine and phenylglyoxal (Scheme 1). Initial

studies addressed the effect of different reagent molar ratios of phenylglyoxal: pyrrolidine on

reaction yields (Fig. 1). The reactions were carried out in toluene at 90 oC under air, in the

presence of 5 mol% Cu-MOF-74 catalyst. It was found that the coupling reaction using 1

equivalent of pyrrolidine proceed with difficulty, affording only 23% GC yield after 120 min.
Increasing the amount of pyrrolidine led to a significant enhancement in the reaction rate, with

47%, 83%, and 94% yields for the reaction using 1.5, 2, and 3 equivalents of pyrrolidine,

respectively. However, using more than 3 equivalents of pyrrolidine resulted in a moderate drop

in the reaction yield. It is worth mentioning that in the first example of the homogeneous version,

10 mol % of 2,2’-bipyridine as ligand and 2 equivalents of pyridine as base were required to

obtained good yields [12]. The use of heterogeneous Cu-MOF-74 catalyst allowed reactions to

proceed efficiently without added ligand and base. This can be rationalized by the organometallic

cooperative effects of linkers and metal clusters [39, 63].

1:1 1:1.5 1:2 1:3 1:4

100

80
Yield (%)

60

40

20

0
0 20 40 60 80 100 120

Time (min)

Fig. 1. Effect of reagent molar ratios on the reaction yields.

Effect of temperature on the reaction yield was then kinetically investigated (Fig. 2). In

particular, the reaction could not occur at room temperature, with no trace amount of desired

product after 120 min. As expected, increasing the temperature led to a dramatic enhancement in

the reaction rate. The coupling reaction carried out at 60 oC and 70 oC afforded 80% and 89 %

yield, respectively. Similar results were achieved when reactions were conducted at 80 oC and 90
o
C with about 95 % yield in 2 hours. In the first example under homogeneous copper catalysis,

Jiao and co-workers carried out the reaction at 90 oC in 12 hours [12]. It is well-known that low

reactivity was frequently observed with heterogeneous catalysts. However, Cu-MOF-74

exhibited comparable catalytic activity with reported homogeneous systems even the absence of

ligand and base.

60 °C 70 °C 80 °C 90 °C
100

80
Yield (%)

60

40

20

0
0 20 40 60 80 100 120

Time (min)

Fig. 2. Effect of temperatures on the reaction yields.

Table 1. Reaction condition development. a

Entry Amount of Solvent Oxidant GC yield (%)


catalyst (%)
1 1 toluene Air 58
2 3 toluene Air 70
3 7 toluene Air 96
4 0 toluene Air 5
5 5 p-xylene Air 95
6 5 mesitylene Air 94
7 5 DMF Air 92
8 5 NMP Air 82
9 5 PhCl Air 76
10 5 CH3CN Air 73
11 5 Dioxane Air 36
12 5 toluene O2 94
13 5 toluene TBHP 95
14 5 toluene None 9
a
Volume of solvent 4 mL, 1.0 mmol scale.

Further reaction optimization was conducted with respect to catalyst loading, solvents, and type

of oxidants (Table 1). Specifically, remarkable drop was obtained when < 5 % catalyst was

employed (entries 1, 2). Increasing catalyst loading to 7 % was not necessary while only 5 %

yield of desired product was formed in the absence of catalyst (entries 3, 4). Furthermore, the

impact of solvents on reaction efficiency is not significant. In details, common aromatic non-

polar solvent such as toluene, p-xylene, or mesitylene afforded about 95 % yields (entries 5, 6).

A slightly decrease in reaction yields were observed when the transformation was carried out in

amide solvents (entries 7, 8). Reasonable yields were still achieved in chlorobenzene and

acetonitrile (entries 9, 10) thought dioxane was showed to be not sufficiently active (entry 11).

Replacing reaction atmosphere from air to molecular oxygen did not affect to reaction yields

(entry 12). Reactions using peroxide oxidants such as tert-butyl hydroperoxide (TBHP) also

offered excellent yield (entry 13). The necessity of oxidant in reaction progress was confirmed

by the low yield obtained from reactions conducted in inert atmosphere (entry 14). In

comparison, similar reported reactions under homogeneous copper catalysis employed 10 – 20 %

catalysts and toluene solvent and atmospheric oxygen were showed to be optimal [4, 11, 12].
Table 2. Reactivity of other related catalysts a

Reported pore aperture


Entry Type Catalyst GC yield (%)
(Å)
1 MOFs Cu-MOF-74 14 (ref. 19) 95
2 Cu3(BTC)2 (MOF-199) 6.0-6.5 (ref. 64) 80
3 Cu(BDC) 5.6 (ref. 65) 77
4 Cu2(BDC)2(DABCO) 4.7 (ref. 66) 82
5 Cu2(BDC)2(BPY) 5.2 - 5.6 (ref. 67) 76
6 Cu2(BPDC)2(BPY) 5.7 (ref. 68) 77
7 Other CuFe2O4 61
8 heterogeneous SCR-Cu-Zeolite 53
9 catalysts Co-MOF-74 10-11 (ref. 19) 62
10 Ni-MOF-74 55
11 Common salts CuI 66
12 CuCl 90
13 Cu(NO3)2 94
14 Cu(OBz)2 96
a
Volume of solvent 4 mL, 1.0 mmol scale.

To highlight the advantages of Cu-MOF-74 catalyst for the oxidative cross-dehydrogenative

coupling reaction, the catalytic activity of the Cu-MOF-74 was compared with that of other Cu-

MOFs containing open metal sites (Table 2). These Cu-MOFs were synthesized by solvothermal

method and characterized according to literature procedures [50, 69-71]. Cu3(BTC)2 (BTC =

benzene-1,3,5-tricarboxylic acid) and Cu(BDC) (BDC = benzene-1,4-dicarboxylic acid), which

possess aperture size of about 6 Å and were commonly used as catalyst for previously reported

organic transformations, afforded 80 % and 77 % yield, respectively (entries 2, 3). Though


introduction of additional ligands such as 2, 2’-bipyridine (BPY) or 1,4-

diazabicyclo[2.2.2]octane (DABCO) resulted in a slightly decrease aperture pore, they was

demonstrated to often enhance the catalytic activities [39b, 44, 45]. However,

Cu2(BDC)2(DABCO), Cu2(BDC)2(BPY) exhibited similar activity as compared to Cu(BDC)

with about 79 % yield (entries 4, 5). Notably, Cu 2(BPDC)2(BPY) (BPDC = biphenyl

dicarboxylic acid) possessing larger pore size and similar aperture pore in comparison with

Cu2(BDC)2(BPY) afforded only 77 % yield (entry 6). Thus, high reaction efficiency was

achieved when using Cu-MOF-74 probably due to its exceptionally large aperture size.

Additionally, other common heterogeneous catalysts including copper magnetic nanoparticles

(CuFe2O4) and Cu-zeolite-SCR showed poor activity (entries 7, 8). Remarkable drops were

observed for Co-MOF-74 and Ni-MOF-74 confirming the vital role of copper active sites

(entries 9, 10). Interestingly, the catalytic activity of Cu-MOF-74 is comparable with that of

homogeneous copper salts (entries 11-14).

5 mol% Leaching test


TEMPO (1 equiv.)
100

80

60
Yield (%)

40

20

0
0 20 40 60 80 100 120

Time (min)
Fig. 3. Leaching test and reaction with added TEMPO

In order to confirm the reaction heterogeneity, a control experiment was carried out using a

simple catalyst filtration during the course of the reaction (Fig. 3). The coupling reaction was

carried out under optimized condition: 3 equivalents of pyrrolidine, 5 mol% Cu-MOF-74, in

toluene at 80 oC for 120 min using air oxidant. After 20 min reaction time with 67% yield, the

Cu-MOF-74 catalyst was separated from the reaction mixture by hot filtration. The liquid

reaction mixture was then transferred to a new reactor vessel, and magnetically stirred for an

additional 100 min at 80 oC with aliquots being sampled at different time intervals. Experimental

results showed that almost no further yield of desired product was detected after the solid Cu-

MOF catalyst was removed from the reaction mixture. Furthermore, ICP-MS of reaction filtrate

provided that < 10 ppm of copper was detected. It was therefore proposed that the the aerobic

oxidative cross-dehydrogenative coupling of pyrrolidine and phenylglyoxal could only proceed

in the presence of the solid Cu-MOF-74 catalyst, and the contribution of leached active copper or

other species, if any, was negligible. In addition, mechanistic experiment with added (2,2,6,6-

tetramethylpiperidin-1-yl)oxy (TEMPO, 1 equiv.) indicated that radical pathway is unlikely.

100

80
Yield (%)

60

40

20

0
1 2 3 4 5 6 7 8 9

Run
Fig. 4. Catalyst recycling studies.

The ability to recover and reuse offers heterogeneous catalytic systems with high potential on

achieving more environmentally benign processes. The Cu-MOF-74 was therefore investigated

for recoverability and reusability. The coupling reaction was carried out under optimal

conditions. After the coupling reaction was complete, the catalyst was separated from the

reaction mixture by simple filtration, washed with copious amounts of DMF and methanol to

remove any physisorbed reagents, dried 150oC under vacuum in 2 h, and reused in further

reactions under identical conditions of previous run. It was found that the Cu-MOF-74 catalyst

could be recovered and reused several times without a significant degradation in catalytic

activity. Indeed, up to 90% yield was still achieved in the 9th run (Fig. 4). The XRD result of the

reused Cu-MOF-74 catalyst revealed that the crystallinity of the Cu-MOF was maintained during

the course of the reaction (Fig. 5). Furthermore, the FT-IR spectra of the reused Cu-MOF-74

exhibited a similar absorption as compared to that of the fresh catalyst (Fig. 6).

1600

1400

1200

1000
Relative intensity

800

600
b)

400

200
a)

0
5 10 15 20 25 30
2-Theta scale
Fig. 5. X-ray powder diffractograms of the fresh (a) and reused (b) Cu-MOF-74 catalyst.

80

60
Transmittance [%]

40 a)

20

b)
0
4000 3400 2800 2200 1600 1000 400

-1
Wavenumber [cm ]

Fig. 6. FT-IR spectra of the fresh (a) and reused (b) Cu-MOF-74 catalyst.

The generality of optimized conditions was tested by extending the reaction scope with various

N-H amines (Table 3). Corresponding to 95 % GC yield, a 92 % isolated yield was obtained for

reaction of of phenylglyoxal with pyrrolidine (entry 1). Amination using piperidine and its

derivatives gave products excellent yields (entries 2, 3). Morpholine is also active and its cross

coupling reaction provided 88 % yield (entry 4). Optimized condition under Cu-MOF-74 is not

limited to cyclic N-H amines. Particularly, reaction of phenylglyoxal with dibutylamine afforded

product in moderate yield (entry 5). Moreover, the optimal condition is also applicable for

aromatic amines and 78 % of product was formed in the reaction of N-methylaniline (entry 6).

Gratifyingly, the synthesis of -ketoamides from primary amine is also possible and aminated

product using benzylamine was achieved in 50 % yield (entry 7).


Table 3. Reaction scope with respect to different amines a

Entry Amines Products Isolated yields (%)


O
N
1 NH 92
O
O
2 NH N 85
O

4 81

O O
3 O NH N 88
O
NH O
5 b N 47
O
O
NH
b N
6 78
O

7 50

a
Volume of solvent 4 mL, 1.0 mmol scale.b Reactions under dried oxygen.

4. Conclusions
In summary, the metal-organic framework Cu-MOF-74 was synthesized by a solvothermal

method. Its properties were fully characterized by several techniques including XRD, SEM,

TEM, FT-IR, TGA, AAS, and nitrogen physisorption measurements. The Cu-MOF-74 was

showed to be an efficient heterogeneous catalyst for oxidative cross-dehydrogenative coupling of

amines and -carbonyl aldehydes to form -ketoamides. Optimized conditions involved the use

of 5 % catalyst in toluene solvent at 80 oC under air atmosphere. The method provide a more

simple, practical, and benign protocol in approaching -ketoamides without added ligand and

base as compare to the reported homogeneous routes. The experimental results indicated that

catalytic activity of Cu-MOF-74 is higher than that of other related Cu-based MOFs and

heterogeneous catalysts and is comparable to common copper salts. Leaching test confirmed the

heterogeneity and contribution from leached species is unlikely. Furthermore, the Cu-MOF-74

catalyst could be recovered and reused several times without a significant degradation in

catalytic activity. Thorough investigation about the correlation between aperture size and activity

for MOF-74, in general, is undertaken to highlight to its potential in catalysis applications due to

the exceptional large pore aperture.

Acknowledgements

The Vietnam National University – Ho Chi Minh City (VNU-HCM) is acknowledged for

financial support through project No. B2015-20-03.

References

1. C. Zhang, Z. Xu, L. Zhang, N. Jiao, Angew. Chem., Int. Ed. 50 (2011) 11088-11092.

2. A. Y. Shaw, C.R. Denning, C. Hulme, Tetrahedron Lett. 53 (2012) 4151-4153.

3. R. Deshidi, M. Kumar, S. Devari, B.A. Shah, Chem. Commun. 50 (2014) 9533-9535.

4. F.-T. Du, J.-X. Ji, Chem. Sci. 3 (2012) 460-465.


5. I. Ojima, N. Yoda, M. Yatabe, T. Tanaka, T. Kogure, Tetrahedron 40 (1984) 1255-1268.

6. J. M. Grassot, G. Masson, J. Zhu, Angew. Chem., Int. Ed. 46 (2007) 947-950.

7. A. Papanikos, J. Rademann, M.Medal, J. Am. Chem. Soc. 123 (2001) 2176-2181.

8. J. Liu, R. Zhang, S. Wang, W. Sun, C. Xia, Org. Lett. 11 (2009) 1321-1324.

9. E. R. Murphy, J.R. Martinelli, N. Zaborenko, S.L. Buchwald, K.F. Jensen, Angew.

Chem., Int. Ed. 46 (2007) 1734-1737.

10. R. Hua, H.-A. Takeda, Y. Abe, M. Tanaka, J. Org. Chem. 69 (2004) 974-976.

11. C. Zhang, N. Jiao, J. Am. Chem. Soc. 132 (2010) 28-29.

12. C. Zhang, X. Zong, L. Zhang, N. Jiao, Org. Lett. 14 (2012) 3280-3283.

13. H. K. Chae, D.Y. Siberio-Perez, J. Kim, Y. Go, M. Eddaoudi, A.J. Matzger, M. O'Keeffe,

O.M. Yaghi, Nature 427 (2004) 523-527.

14. D. J. Tranchemontagne, M.O.K Z. Ni, O.M. Yaghi, Angew. Chem. Int. Ed. 47 (2008)

5136-5147.

15. D. T. Genna, A.G. Wong-Foy, A.J. Matzger, M.S. Sanford, J. Am. Chem. Soc. 135

(2013) 10586-10589.

16. J. A. Mason, M. Veenstra, J. R. Long, Chem. Sci. 5 (2014) 32-51.

17. M.-L. Ma, C. Ji, S.-Q. Zang, Dalton Trans. 42 (2013) 10579-10586.

18. S. S. Kaye, A. Dailly, O.M. Yaghi, J.R. Long, J. Am. Chem. Soc. 129 (2007) 14176-

14177.

19. (a) H. Furukawa, N. Ko, Y.B. Go, N. Aratani, S.B. Choi, E. Choi, A.O. Yazaydin, R.Q.

Snurr, M. O'Keeffe, J. Kim, O.M. Yaghi, Science 239 (2010) 424-428. (b) H. Deng, S.

Grunder, K. E. Cordova, C. Valente, H. Furukawa, M. Hmadeh, F. Gándara, A. C.


Whalley, Z. Liu, S. Asahina, H. Kazumori, M. O’Keeffe, O. Terasaki, J. F. Stoddart, O.

M. Yaghi, Science 336 (2012) 1018-1023.

20. P. Horcajada, T. Chalati, C. Serre, B. Gillet, C. Sebrie, T. Baati, J.F. Eubank, D.

Heurtaux, P. Clayette, C. Kreuz, J.S. Chang, Y.K. Hwang, V. Marsaud, P.N. Bories, L.

Cynober, S. Gil, G. Fe´rey, P. Couvreur, R. Gref, Nature Mater. 9 (2010) 172-178.

21. R. J. Kuppler, D.J. Timmons, Q.-R. Fang, J.-R. Li, T.A. Makal, M.D. Young, D. Yuan,

D. Zhao, W. Zhuang, H.-C. Zhou, Coord. Chem. Rev. 253 (2009) 3042-3066.

22. H. Li, M. Eddaoudi, M. O'Keeffe, O.M. Yaghi, Nature 402 (1999) 276-279.

23. J. L. C. Rowsell, O.M. Yaghi, Micropor. Mesopor. Mater. 73 (2004) 3-14.

24. Z.-Q. Li, L.-G. Qiu, T. Xu, Y. Wu, W. Wang, Z.-Y. Wu, X. Jiang, Mater. Lett. 63 (2009)

78-80.

25. B. Chen, S. Xiang, G. Qian, Acc. Chem. Res. 43 (2010) 1115-1124.

26. Z.-Y Gu, J. Park, A. Raiff, Z. Wei, H.-C. Zhou, ChemCatChem 6 (2014) 67-75.

27. P. Valvekens, F. Vermoortelea, D.D. Vos, Catal. Sci. Technol. 3 (2013) 1435-1445.

28. J. Liu, L. Chen, H. Cui, J. Zhang, L. Zhang, C.-Y. Su, Chem. Soc. Rev. 43 (2014) 6011-

6061.

29. A. Corma, H. García, F.X. Llabrés i Xamena, Chem. Rev. 110 (2010) 4606-4655.

30. A. Dhakshinamoorthy, H. Garcia, Chem. Soc. Rev. 41 (2012) 5262-5284.

31. M. Yoon, R. Srirambalaji, K. Kim, Chem. Rev. 112 (2012) 1196-1231.

32. F. Zadehahmadi, S. Tangestaninejad, M. Moghadam, V. Mirkhani, I. Mohammadpoor-

Baltork, A.R. Khosropour, R. Kardanpour, Appl. Catal. A. Gen. 477 (2014) 34-41.

33. T. Otto, N. N. Jarenwattananon, S. Glöggler, J.W. Brown, A. Melkonian, Y.N. Ertas, L.-

S. Bouchard, Appl. Catal. A. Gen. 488 (2014) 248-255.


34. Q. Luo, X.-D. Song, M. Ji, S.-E. Park, C. Hao, Y.-q. Li, Appl. Catal. A. Gen. 478 (2014)

81-90.

35. K. Leus, Y.-Y. Liu, M. Meledina, S. Turner, G.V. Tendeloo, P.V.D. Voort, J. Catal. 316

(2014) 201-209.

36. P. Wu, C. He, J. Wang, X. Peng, X. Li, Y. An, C. Duan, J. Am. Chem. Soc. 134 (2012)

14991-14999.

37. A. S. Roy, J. Mondal, B. Banerjee, P. Mondal, A. Bhaumik, S.M. Islam, Applied

Catalysis A: General 469 (2014) 320-327.

38. N. T. S. Phan, T.T. Nguyen, P. Ho, K.D. Nguyen, ChemCatChem 5 (2013) 1822-1831.

39. (a) T. Truong, V.T. Nguyen, H.T.X. Le, N.T.S. Phan, RSC Adv. 4 (2014) 52307-52315.

(b) H. T. N. Le, T. V. Tran, N. T. S. Phan, T. Truong, Catal. Sci. Technol. 5 (2015) 851-

859.

40. Y. Luan, N. Zheng, Y. Qi, J. Tang, G. Wang, Catal. Sci. Technol. 4 (2014) 925-929.

41. G. Yu, J. Sun, F. Muhammad, P. Wang, G. Zhu, RSC Adv. 4 (2014) 38804-38811

42. Y.-R. Lee, Y.-M. Chung, W.-S. Ahn, RSC Adv. 4 (2014) 23064-23067

43. A. S. Roy, J. Mondal, B. Banerjee, P. Mondal, A. Bhaumik, S.M. Islam, Appl. Catal. A.

Gen. 469 320-327.

44. G. H. Dang, D.T. Nguyen, D.T. Le, T. Truong, N.T.S. Phan, J. Mol. Catal. A: gen. 395

(2014) 300-306.

45. G. H. Dang, T.T. Dang, D.T. Le, T. Truong, N.T.S. Phan, J. Catal. 319 (2014) 258-264.

46. P. Valvekens, M. Vandichel, M. Waroquier, V.V. Speybroeck, D.D. Vos, J. Catal. 317

(2014) 1-10.
47. M. Savonnet, S. Aguado, U. Ravon, D. Bazer-Bachi, V. Lecocq, N. Bats, C. Pinel, D.

Farrusseng, Green Chem. 11 (2009) 1729-1732.

48. E. Pérez-Mayoral, Z. Musilová, B. Gil, B. Marszalek, M. Položij, P. Nachtigall, J. Čejka,

Dalton Trans. 41 (2012) 4036-4044.

49. M. Opanasenko, M. Shamzhy, J. Čejka, ChemCatChem 5 (2013) 1024-1031.

50. N. T. S. Phan, P. H. L. Vu, T.T. Nguyen, J. Catal. 306 (2013) 38-46.

51. O. V. Zalomaeva, A.M. Chibiryaev, K.A. Kovalenko, O.A. Kholdeeva, B.S.

Balzhinimaev, V.P. Fedin, J. Catal. 298 (2013) 179-185.

52. H. T. N. Le, T.V. Tran, N.T.S. Phan, T. Truong, Catal. Sci. Technol. 5 (2015) 851-859.

53. (a) I. Luz, A. Corma, F.X.L.i. Xamena, Catal. Sci. Technol. 4 (2014) 1829-1836. (b) G.

R. Calleja, R. Sanz, G. Orcajo, D. Briones, P. Leo, F. Martínez, Catalysis Today 227

(2014) 130–137. (c) D. Ruano, M. D. Garcia, A. Alfayate, M. S. -Sanchez,

ChemCatChem, 7 (2015) 674-681.

54. T. Maity, D. Saha, S. Koner, ChemCatChem 6 (2014) 2373-2383.

55. A. Dhakshinamoorthy, M. Alvaro, H. Garcia, Advanced Synthesis & Catalysis 351

(2009) 2271-2276.

56. A. Dhakshinamoorthy, M. Alvaro, H. Garcia, Advanced Synthesis & Catalysis 352

(2010) 711-717.

57. A. Dhakshinamoorthy, M. Alvaro, H. Garcia, Advanced Synthesis & Catalysis 352

(2010) 3022-3030.

58. L. Alaerts, E. Séguin, H. Poelman, F. Thibault-Starzyk, P.A. Jacobs, D.E. De Vos,

Chemistry – A European Journal 12 (2006) 7353-7363.

59. I. Luz, F.X. Llabrés i Xamena, A. Corma, Journal of Catalysis 285 (2012) 285-291.
60. I. Luz, F.X. Llabrés i Xamena, A. Corma, Journal of Catalysis 276 (2010) 134-140.

61. (a) G. H. Dang, D.T. Le, T. Truong, N.T.S. Phan, J. Mol. Catal. A, 400 (2015) 162-169.

(b) G. H. Dang, Y.T.H. Vu, Q.A. Dong, D.T. Le, T. Truong, N.T.S. Phan, Appl. Catal. A.

Gen. 491 (2015) 189-195.

62. R. Sanz, F. Martínez, G. Orcajo, L. Wojtas, D. Briones, Dalton Trans. 42 (2013) 2392-

2398.

63. Z. Zhang, H. Yoshikawa, K. Awaga. J. Am. Chem. Soc. 136 (2014) 16112-16115.

64. (a) D. Peralta, G. Chaplais, A. Simon-Masseron, K. Barthelet, C. Chizallet, A.-A.

Quoineaud and G. D. Pirngruber, J. Am. Chem. Soc. 134 (2012) 8115–8126. (b) B. Zornoza, B.

Seoane, J. M. Zamaro, C. Tellez, J. Coronas, ChemPhysChem. 12 (2011) 2781-2787.

65. J. Liu, B. Lukose, O. Shekhah, H. K. Arslan, P. Weidler, H. Gliemann, S.

Bräse, S. Grosjean, A. Godt, X. Feng, K. Müllen, I.-B. Magdau, T. Heine, C. Wöll,

Scientific Reports 2, 921 (2012) doi:10.1038/srep00921.

66. M. Maes, S. Schouteden, K. Hirai, S. Furukawa, S. Kitagawa, D. E. De Vos, Langmuir

27 (2011) 9083–9087.

67. (a) M. Tu, S. Wannapaiboon, R. A. Fischer, Dalton Trans. 42 (2013) 16029 - 16035. (b)

T. A. Makal, A. A. Yakovenko, H.-C. Zhou, |J. Phys. Chem. Lett. 2 (2011) 1682–1689.

68. A. Pichon, C. M. Fierro, M. Nieuwenhuyzen, S. L. James CrystEngComm, 9 (2007)

449–451.

69. (a) L. T. L. Nguyen, T.T. Nguyen, K.D. Nguyen, N.T.S. Phan, App. Catal. A. Gen. 425-

426 (2012) 44-52. (b) N. T. T. Tran, Q. H. Tran, T. Truong. J. Catal 320 (2014) 9-15.

70. H. Furukawa, J. Kim, N.W. Ockwig, M. O’Keeffe, O.M. Yaghi, J. Am. Chem. Soc. 130

(2008) 11650-11661.
71. K. Tan, N. Nijem, P. Canepa, Q. Gong, J. Li, T. Thonhauser, Y.J. Chabal, Chem. Mater.

24 (2012) 3153-3167.

View publication stats

You might also like