Pd@Cu (II) - MOF-Catalyzed Aerobic Oxidation of Benzylic Alcohols in Air With High Conversion and Selectivity

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Article

pubs.acs.org/IC

Pd@Cu(II)-MOF-Catalyzed Aerobic Oxidation of Benzylic Alcohols in


Air with High Conversion and Selectivity
Gong-Jun Chen,* Jing-Si Wang, Fa-Zheng Jin, Ming-Yang Liu, Chao-Wei Zhao, Yan-An Li,
and Yu-Bin Dong*
College of Chemistry, Chemical Engineering and Materials Science, Collaborative Innovation Center of Functionalized Probes for
Chemical Imaging in Universities of Shandong, Key Laboratory of Molecular and Nano Probes, Ministry of Education, Shandong
Normal University, Jinan 250014, P. R. China
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via INDIAN INST OF TECH INDORE on November 7, 2019 at 18:42:12 (UTC).

ABSTRACT: A new 3D porous Cu(II)-MOF (1) was


synthesized based on a ditopic pyridyl substituted diketonate
ligand and Cu(OAc)2 in solution, and it features a 3D NbO
motif which is determined by the X-ray crystallography.
Furthermore, the Pd NPs-loaded hybrid material Pd@Cu(II)-
MOF (2) was prepared based on 1 via solution impregnation,
and its structure was confirmed by HRTEM, SEM, XRPD, gas
adsorption−desorption, and ICP measurement. 2 exhibits
excellent catalytic activity (conversion, 93% to >99%) and
selectivity (>99% to benzaldehydes) for various benzyl alcohol
substrates (benzyl alcohol and its derivatives with electron-
withdrawing and electron-donating groups) oxidation reac-
tions in air. In addition, 2 is a typical heterogeneous catalyst,
which was confirmed by hot solution leaching experiment, and it can be recycled at least six times without significant loss of its
catalytic activity and selectivity.

■ INTRODUCTION
Selective catalytic oxidation of benzyl alcohols to the
metal nodes and organic linkers. The structurally ordered and
molecularly tunable features make them an ideal scaffold to
corresponding benzaldehydes is one of the most important integrate various functional moieties. Their incorporation
issues in modern synthetic chemistry and chemical industry. between distinct individual doped species will endow such
Hitherto, a great deal of research effort has been devoted to composite materials with great potential applications, especially
finding novel and efficient catalysts for this type of reaction.1−3 in the catalytic field. High surface areas, controllable pore sizes,
Over the past decades, nanoparticles (NPs) have been and tunable pore environments of MOFs would facilitate the
intensively pursued in a wide range of important catalytic MOF supports to entrap various NPs.44−50 In addition, the
processes owing to their large surface area-to-volume ratios. For crystalline porous structure together with heteroatom donors
example, Cu,4−6 Ru,7 Pd,8−11 Au,12−16 and Au−Pd alloy
on MOFs would effectively limit the aggregation and migration
NPs17−22 were demonstrated to be the efficient species for
alcohol oxidation. Among these NPs, Pd NPs have attracted of small active catalytic NPs in the solid state, consequently,
more attention owing to their excellent catalytic perform- making the NPs involved NPs@MOFs catalysts to be highly
ance.23,24 Although Pd NPs show good catalytic behaviors for active and reusable.51,52
many useful reactions under reaction conditions, Pd NPs are In this contribution, we report the synthesis and structure of
prone to aggregation and form Pd black because of their high a new porous Cu(II)-MOF (1) and its Pd NPs-embedded
surface energy.25 Therefore, the immobilization of Pd NPs in composite Pd@Cu(II)-MOF (2) via solution impregnation
porous supports is regarded as one of the most practical ways to (Scheme 1). Notably, 2 can be a highly active catalyst to
address this problem. So far, various inorganic materials, such as promote the aerobic oxidation of various benzylic alcohols in
CeO226−28 SiO2,29−34 ZrO2,35 and porous carbon,36−40 and air with high conversion rates (93% to >99%) and chemo-
organic materials, such as polystyrene,41 amphiphilic resin,42
selectivity (aldehydes, >99%).53 Moreover, the catalyst can be
and copolymers,43 have been utilized as supports to incorporate
Pd NPs. However, the inorganic−organic hybrid materials as a easily recovered and reused due to its heterogeneous catalytic
heterogeneous supporting matrix for metal NPs have not yet nature.
attracted much attention.
Metal−organic frameworks (MOFs) are a new class of Received: December 25, 2015
inorganic−organic hybrid materials composed of inorganic Published: March 9, 2016

© 2016 American Chemical Society 3058 DOI: 10.1021/acs.inorgchem.5b02973


Inorg. Chem. 2016, 55, 3058−3064
Inorganic Chemistry Article

Scheme 1. Synthesis of Cu(II)-MOF (1) and Pd@Cu(II)- an operating voltage of 200 kV. Scanning electron microscopy (SEM)
MOF (2)a analysis was performed on a Gemini Zeiss Supra 55. Gas
chromatography (GC) analysis was performed on an Agilent 7890B
GC.
X-ray Crystallography. Diffraction data for the complex were
collected at 293(2) K, with a Bruker Smart 1000 CCD diffractometer
using Mo−Kα radiation (λ = 0.71073 Å) with the ω-2θ scan
technique. An empirical absorption correction was applied to raw
intensities.55 The structure was solved by direct methods (SHELX-97)
and refined with full-matrix least-squares technique on F2 using the
SHELX-97.56 The hydrogen atoms were added theoretically, and
riding on the concerned atoms and refined with fixed thermal factors.
The details of crystallographic data and structure refinement
parameters are summarized in Table S1. The selected bonds lengths
and angles for 1 are summarized in Table S2. CCDC 1443366
contains the supplementary crystallographic data for this paper. These
data can be obtained free of charge from The Cambridge
Crystallographic Data Centre via www.ccdc.cam.ac.uk/data_requeat/
cif.
Synthesis of Cu(II)-MOF (1). A solution of Cu(OAc)2 (2.4 mg,
0.01 mmol) in MeOH (1 mL) was layered onto a solution of HL (4.8
mg, 0.02 mmol) in CH2Cl2 (2 mL). The solutions were left for about 3
days at room temperature, and compound 1 was obtained as bright
green crystals (Scheme 1). Yield, 75%. IR (KBr pellet, cm−1):
3068(w), 1607 (s), 1569 (s), 1536 (s), 1458 (m), 1299 (m), 1250
a
The photographs of 1 and 2 are inserted. (m), 1193 (s), 1145 (s), 1073 (m), 787 (s), 701 (s). Elemental


analysis (%) calcd for C96H54Cu3F18N6O18 (desolvated): C, 54.59, H,
2.58, N, 3.98; Found: C, 54.48, H 2.71, N 3.96.
EXPERIMENTAL SECTION Synthesis of Pd@Cu(II)-MOF (2). 1 (100 mg, 0.02 mmol) was
Materials and Instrumentation. The reagents and solvents added to a CH3CN (10 mL) solution of palladium nitrate (90 mg, 0.2
employed were commercially available and used without further mmol). The mixture was stirred for 1 h at room temperature. The
purification. Ligand (HL = 4,4,4-trifluoro-1-(4-(pyridin-4-yl)phenyl)- resulting solid was isolated by centrifugation and washed with CH2Cl2.
butane-1, 3-dione) was synthesized according to our reported The obtained green-yellow crystalline solids (Figure S1) were mixed
procedure.54 Elemental analyses for C, H, and N were obtained on with NaBH4 (50 mg, 1.3 mmol) in water (4 mL), and the mixture was
a PerkinElmer analyzer model 240. The powder diffractometer (XRD) stirred for an additional 15 h to afford 2 as a dark brown solid (Scheme
patterns were collected by a D8 ADVANCEX-ray with Cu Kα 1). The obtained crystalline solids were washed with water and EtOH
radiation (λ = 1.5405 Å). Electron spin resonance (ESR) spectra were and dried in air. ICP measurement indicated that the encapsulated
obtained from a Bruker A300-10/12/S-LC. Infrared (IR) samples were amount of Pd NPs in 2 is up to 5.1% (mass fraction).
prepared as KBr pellets, and spectra were obtained in the 400−4000 The General Catalytic Reaction Procedure. The catalytic
cm−1 range using a PerkinElmer 1600 FTIR spectrometer. The total activity of 2 for the aerobic oxidation of benzyl alcohol to
surface areas of the catalysts were measured by the BET (Brunauer− benzaldehyde was tested in air. A mixture of benzyl alcohols (25 μL,
Emmett−Teller) method using carbon dioxide adsorption at 195 K. 0.21 mmol), 2 (20 mg, 5% Pd), and xylene (3.0 mL) was stirred at 130
This was done by the Micromeritics ASAP 2000 sorption/desorption °C for 25 h in the air (monitored by gas chromatography (GC)) to
analyzer. ICP-LC was performed on an IRIS Interpid (II) XSP and afford the corresponding aldehydes. The conversion and selectivity
NU AttoM. HRTEM (high-resolution transmission electron micros- were determined by GC. 2 was recovered by centrifugation. After it
copy) analysis was performed on a JEOL 2100 electron microscope at was washed with ethanol (3.0 mL × 3) and dichloromethane (3.0 mL

Figure 1. (a) Simulated and measured XRPD patterns of 1. (b) Coordination sphere of Cu(II) in 1. (c) Single cubic unit of 1. (d) 3-fold
interpenetrating NbO-net of 1 with 1D channel down the crystallographic c axis.

3059 DOI: 10.1021/acs.inorgchem.5b02973


Inorg. Chem. 2016, 55, 3058−3064
Inorganic Chemistry Article

Figure 2. Left: SEM and elemental maps of Cu and Pd images of 2. Simulated and measured XRPD patterns of 2. Right: HRTEM image of 2.

Figure 3. Left: CO2 adsorption isotherms for Cu(II)-MOF (1) and Pd@Cu(II)-MOF (2) at 195 K. Right: The pore widths of 1 and 2 are centered
at 1.3 and 0.9 nm, respectively.

× 3) and dried at 90 °C in vacuum, 2 was used in the next run under molecules can be removed at 110 °C based on the TGA trace
the same reaction conditions. (Figures S2 and S3). The desolvated sample of 1 is stable up to
Leaching Test. The solid catalyst Pd@Cu(II)-MOF (2) was ca. 200 °C (demonstrated by XRPD, Figure S4). The high
separated from the hot solution right after reaction for 5 h. The
reaction was continued with the filtrate in the absence of 2 for an
thermal stable feature of 1 would make it suitable to be a
additional 15 h. No further increase in either the conversion of benzyl catalyst support for uploading active NPs to promote organic
alcohol or the selectivity of aldehyde was detected, which confirms that reactions at relative high temperature.
the catalytically active sites for this oxidation reaction were located on Palladium-embedded Pd@Cu(II)-MOF (2) was prepared by
2. the impregnation with the Pd(NO3)2 in CH3CN for 5 h,


yielding Pd(II)@Cu(II)-MOF (Scheme 1). The Pd NPs-loaded
RESULTS AND DISCUSSION Pd@Cu(II)-MOF (2) was obtained by the reduction of Pd(II)
@Cu(II)-MOF with NaBH4 in aqueous solution. The uploaded
Synthesis and Structural Analysis of Cu(II)-MOF (1)
amount of Pd, as determined by inductively coupled plasma
and Pd@Cu(II)-MOF (2). The combination of HL with
Cu(OAc)2 in a CH2Cl2/MeOH mixed solvent system to afford (ICP) measurement, is up to 5.1 wt %. Presumably, the
1 ([Cu3(L)6]) as bright green crystals in 75% yield. Simulated generated Pd NPs might be stabilized by the multitudinous F
and measured XRPD patterns match each other very well, atoms in the framework. The energy-dispersive X-ray spectrum
indicating that 1 was obtained in pure phase (Figure 1a). (EDS) measurement (Figure 2) shows that the uploaded
Compound 1 crystallizes in the hexagonal space group R3̅. Each palladium homogeneously distributes in 2. In addition, the
Cu(II) center in 1 adopts a 4 + 2 octahedral coordination structure of Cu(II)-MOF (1) remained intact after the Pd
sphere (Figure 1b), in which two coplanar chelating β-diketone uploading and no changes in the XRPD patterns were detected,
units form the square (Cu−O distances of 2.107(3)−2.124(3) which supports that Cu(II)-MOF (1) was chemically stable
Å), and the axial positions are occupied by two pyridyl N- during the palladium reducing process (Figure 2). In addition,
donors (Cu−N distance of 2.016(3) Å). As shown in Figure 1c, the ESR spectra show that no valence state change of the
octahedral Cu(II) centers are linked together by L to a 4- Cu(II) in Pd@Cu(II)-MOF (2) was observed after the
connected 3D NbO network. The shortest opposite Cu(II)··· treatment by NaBH4 in aqueous solution under the reaction
Cu(II) distance in a single cubic unit is 26.6139(4) Å. Notably, conditions (Supporting Information, Figure S5).
1 is 3-fold interpenetrating, and the three sets of NbO To get a further insight into the structure of 2, high-
frameworks are displaced by ca. (1/3, 1/3, 1/3) relative to each resolution transmission electron microscopy (HRTEM) was
other so that the large hexagonal channels appear along the used to investigate the dispersion and size distribution of the Pd
crystallographic c axis (Figure 1d). The opposite Cu···Cu NPs in 2. HRTEM analysis revealed that the Pd NPs were
distance in the channel is 22.9330(3) Å. All the fluorine atoms crystalline and highly dispersed with an average particle size of
on the framework face toward the center of the channels, and ca. 2 nm (Figure 2). The lattice fringes had an interplanar
the opposite F···F distance is 14.4423(3) Å. In addition, solvent spacing of 0.24 nm, corresponding to 1/3 (4 2 2) fringes of
molecules are located in 1, and the encapsulated solvent face-centered cubic (fcc) Pd.57
3060 DOI: 10.1021/acs.inorgchem.5b02973
Inorg. Chem. 2016, 55, 3058−3064
Inorganic Chemistry Article

BET Surface Area Measurement of 1 and 2. In order to xylene is the best solvent for this oxidation reaction in
investigate the BET surface area of Cu(II)-MOF (1) and Pd@ atmospheric air. As shown in Table 1 (entries 5 and 6), the
Cu(II)-MOF (2), gas adsorption−desorption experiment was reaction conversion and selectivity (to benzaldehyde) in
performed on 1 and 2. A N2 adsorption isotherm was first toluene (110 °C, 25 h, Figure S12) are 88% and 91%,
measured, but it exhibits negligible uptake by the framework respectively. Meanwhile, the corresponding conversion and
even at low temperature. Such a phenomenon has been selectivity in CH3CN (70 °C, 25 h, Figure S13) are 15% and
observed for some MOFs.58 However, CO2 adsorption of 1 and 95%, respectively. Besides benzaldehyde, benzoic acid and an
2 at 195 K revealed a typical type IV mode, showing pore unidentified species were also detected in entries 5 and 6 by
condensation with pronounced adsorption−desorption hyste- GC.
resis (Figure 3). It indicates the existence of mesopores in 1; The conversion and the selectivity of benzyl alcohol to
meanwhile, 2 shows no adsorption−desorption hysteresis, benzaldehyde in xylene at different reaction times are shown in
indicating that 2 is microporous. The surface areas of 1 and Figure 4. The initial conversion of benzaldehyde is continu-
2 are 1129.24 and 373.24 m2/g, respectively. The different
adsorption capacities and behaviors of 1 and 2 are clearly
caused by the Pd NPs doping. The mesopore size distribution
curve, calculated from Barrett−Joyner−Halenda analysis, shows
a narrow pore diameter distribution at ca. 1.3 nm and ca. 0.9
nm for 1 and 2, respectively (Figure 3), which is well consistent
with the single-crystal analysis of 1. Thus, most of the Pd NPs
might not be located inside the MOF cavity, as they are larger
than the pore size. Therefore, these Pd NPs are presumably
sandwiched in the crystallite matrix of Cu(II)-MOF (1) and
partly stabilized by F atoms on the framework.
Oxidation of Benzyl Alcohol in Air. Reaction Con- Figure 4. Reaction time examination (black line) and leaching test
ditions. The aerobic oxidation reactions (with 2, 5% Pd) of the (red line) for benzyl alcohol reaction catalyzed by 2. Reaction
benzylic alcohols with different substituted groups in xylene conditions: air, Pd@Cu(II)-MOF (2) (5% Pd), benzyl alcohol (0.21
were carried out in air (monitored by GC). We found that the mmol), xylene (3 mL). The solid catalyst was filtrated from the
reaction solution after 5 h, whereas the filtrate was transferred to a new
temperature, solvent system, and reaction time are the key
vial and reaction was carried out under the same conditions for an
parameters for this Pd@Cu(II)-MOF-catalyzed oxidation additional 15 h.
reaction. The optimized reaction conditions are obtained
based on the benzyl alcohol oxidation in atmospheric air
(Table 1). The reaction temperature was first investigated for
ously increased, and the maximum yield (95%) appeared at 25
Table 1. Oxidation of Benzyl Alcohol in Different Solvent h (Figures S14−S19). No more obvious changes in the
Systems at Different Temperaturesa conversions can be detected after 25 h; the TON and TOF are
19 and 0.76 h−1. In order to gain insight into the heterogeneous
nature of 2, the hot leaching test was carried out. As indicated
in Figure 4, no further reaction took place without 2 after
ignition of the oxidation reaction at 5 h (Figures S20 and S21).
entry T (°C) solvent t (h) conv. (%) select (%) This finding demonstrated that no leaching of the catalytically
1 25 xylene 25 0 0
active sites occurs and that 2 exhibits a typical heterogeneous
2 70 xylene 25 15 >99 catalyst nature (Figure 4). In addition, the catalytic activities of
3 110 xylene 25 88 >99 Cu(II)-MOF (1) and treating it with NaBH4 without Pd NPs
4 130 xylene 25 95 >99 as the catalysts for this oxidation reaction were also examined
5 110 toluene 25 88 91 under the same reaction conditions; however, the benzyl
6 70 CH3CN 25 15 95 alcohol conversions are only up to 16% and 15%, respectively
a (Figures S22 and S23).
Reaction conditions: air, benzyl alcohol (0.21 mmol), Pd@Cu(II)-
MOF (2) (5% Pd), solvent (3 mL).
Recyclability Studies. Reusability of catalysts is very
important for industrial applications. The recyclability of 2
was tested. The Pd@Cu(II)-MOF (2) was recycled seven times
the impact on the efficiency of oxidation reactions in air. As for benzyl alcohol oxidation (Figure 5). After each run, the
shown in Table 1, upon increase of the reaction temperature solid catalyst was easily collected by centrifugation, washed with
from 25 to 130 °C in xylene (entries 1−4), the conversion of ethanol/dichloromethane, dried at 90 °C, and reused in the
benzyl alcohol dramatically enhanced. Notably, >99% selectiv- next run under the same conditions. Interestingly, the reaction
ity toward benzaldehyde was maintained during the temper- conversion increased from 95% at the first run to >99% at the
ature increasing process. It should be noted that an excellent 4−6 runs (Figures S24−S29). Although we distance ourselves
yield (95%) was achieved when the reaction was carried out at from any type of explanation and say forthright that we do not
130 °C (Table 1, entry 4). However, the chemoselectivity to know why the conversion enhanced after three catalytic cycles,
benzaldehyde is still >99% at this temperature, which was it is possible because of some smaller solid catalyst particles
confirmed by GC analysis (Figures S6−S11). It, therefore, formed and dispersed more uniformly in the MOF matrix
appears that 130 °C is the optimum temperature for this during the reactions (Figure S30). The size and dispersion of
benzaldehyde oxidation reaction. In addition, we also the Pd NPs in 2 did not show any detectable aggregation after
investigated the effect of the solvent system and found that six runs, and the atomic lattice fringes (0.24 nm) corresponding
3061 DOI: 10.1021/acs.inorgchem.5b02973
Inorg. Chem. 2016, 55, 3058−3064
Inorganic Chemistry Article

Figure 5. Recycling catalytic test. Left: conversion and selectivity obtained at 25 h in repeated runs of the benzyl alcohol oxidation catalyzed by 2
(air, 130 °C, benzyl alcohol (0.21 mmol), 2 (5% Pd), xylene (3.0 mL)). The average conversion and chemoselectivity based on six catalytic cycles are
>97% and >99%, respectively. Right: corresponding XRPD patterns (a−g corresponding to 1−7 runs).

Table 2. Summary of the Reported Benzyl Alcohol Oxidation Reactions Using Pd NPs as the Heterogeneous Catalysts
catalyst cond. run conv. (%) select (%) ref.
Pd/NaX zeolite toluene/100 °C/O2 (3 mL/min) 1 66 97 32
Pd/hybrid SiO2 CO2/O2 (18 MPa)/O2 (8 vol %)/80 °C 1 99.5 89.8 33
Pd/A20E10 toluene/water/80 °C/air (5 bar) 1 85.6 97 61
Pd/Al2O3 solvent free/120 °C/O2 (50 mL/min) 1 80.1 94.3 62
Pd/GC solvent free/110 °C/O2 (20 mL/min) 1 72.5 98.3 36
Pd/SBA-16 water/50 °C/air or O2 11 >99 >99 63
Pd-pol water (K2CO3)/100 °C/air (1 atom) 6 93−99 >99 64
CM-CeO2-Pd-1.0 solvent free/O2 1 82.1 62.9 26
Pd/CB 353 K/O2 1 99 95 37
Pd/AC water/60 °C/O2 (1.5 atm) 1 50 >99 38
ARP-Pd water/100 °C/O2 (1 atm) 1 97 40
Pd/PEG solvent free/80 °C/O2/scCO2 1 83.1 99.8 65
Pd@U-E15 water/K2CO3/90 °C/O2 (1 atm) 1 90 90 66
Pd cluster colloid water/32 °C/air bubbling/(pH 3.5) 10 86−93 100 41
PdHAP-0 trifluorotoluene/90 °C/O2 (1 atm) 1 >99 99 67
water/110 °C/O2 (1 atm) 1 >99 90
Pd@hmC water/K2CO3/80 °C/O2 1 48 37 68
MNP−Pd toluene/K2CO3/85−90 °C/air 4 86−100 90−93 69
Pd@Cu(II)-MOF xylene/130 °C/air (atmospheric pressure) 6 95 to >99 >99 this work

to 1/3(422) of face-centered cubic (fcc) Pd were also clearly exposed metallic palladium faces; therein, the decarbonylation
observed (Figure S27). often occurred and the undesired products preferentially
The conversion, however, decreased to 85% (Figure S31) at formed on hollow sites of Pd particle faces.31,59 However, in
the seventh run, indicating that 2 began to be deactivated. our experiments, almost no decarbonylating products were
HRTEM analysis revealed that some of the Pd NPs in 2 began detected during reaction, so we proposed that the sites on face-
to aggregate (Figure S32), which could be the reason for the centered cubic Pd are the primary sites for benzyl alcohol
low catalysis efficiency of 2 after six runs. Such an observation oxidation.60,67 The distance of the H(Cα-H)−H(hydroxy) on
was further confirmed by the elemental maps of Cu and Pd in 2 benzyl alcohol is about 0.21 nm, which is close to the Pd−Pd
after six runs (Figure S33). In addition, ICP measurement bonding distance of 0.24 nm found in 2. This might effectively
indicated that the amount of Pd NPs in 2 has no significant enhance the interactions between Pd NPs and reaction
changes after six runs; the amount of Pd species, however, substrates, consequently, leading to a high oxidation catalytic
dropped to 3.85% after the seventh cycle, which could be an efficiency.
additional reason for the catalyst inactivation (Table S3). Compared to various reported Pd NPs-loaded heterogeneous
The XRD patterns of 2 and that after being reused for six catalysts (Table 2), 2 herein exhibits good catalytic activity and
cycles indicated that the structural integrity of 2 was well selectivity for benzyl alcohol oxidation in the air. Therefore, it
preserved. Therefore, Cu(II)-MOF can be an ideal support to can be a valuable complement for the MOFs to heterogeneous
upload Pd NPs, furthermore, effectively preventing their aromatic carbinols oxidation composite catalysts.
sintering and aggregation in the solid state even at a higher Oxidation of Other Substituted Benzylic Alcohols. The
reaction temperature (Figure 5). scope of the oxidation catalytic system was explored by
It has been reported that the oxidative dehydrogenation of performing the oxidation reactions of various substituted
alcohols to the corresponding aldehydes happened on all benzylic alcohols. Table 3 summarizes the results of these
3062 DOI: 10.1021/acs.inorgchem.5b02973
Inorg. Chem. 2016, 55, 3058−3064
Inorganic Chemistry Article

Table 3. Oxidation Reactions of Benzyl Alcohols with Notes


Different Substituted Groups Catalyzed by 2 in Aira The authors declare no competing financial interest.

■ ACKNOWLEDGMENTS
We are grateful for financial support from NSFC (Grant Nos.
21475078, 21271120, and 21301109), the 973 Program (Grant
Nos. 2012CB821705 and 2013CB933800), the Taishan
Scholar’s Construction Project, the Reward Fund for Out-
standing Young and Middle Aged Scientists of Shandong
Province (no. BS2012CL032), and the Jinan Science and
Technology Bureau (OUT_06623).

■ REFERENCES
(1) Backvall, J. E. Modern Oxidation Reactions, 1st ed.; Wiley-VCH:
Weinheim, Germany, 2004.
(2) Fernandez, M. I.; Tojo, G. Oxidation of Alcohols to Aldehydes
and Ketones. In A Guide to Current Common Practice; Springer: New
York, 2006.
(3) Long, R.; Huang, H.; Li, Y. P.; Song, L.; Xiong, Y. J. Adv. Mater.
a
Reaction conditions: air, 130 °C, benzylic alcohols (0.21 mmol), 2
2015, 27, 7025−7042.
(5% Pd), xylene (3.0 mL). Reaction conversion and selectivity are
(4) Hu, Z.; Kerton, F. M. Appl. Catal., A 2012, 413−414, 332−339.
determined based on GC analysis (Figures S34−S39).
(5) Ryland, B. L.; McCann, S. D.; Brunold, T. C.; Stahl, S. S. J. Am.
Chem. Soc. 2014, 136, 12166−12173.
(6) Li, L. C.; Matsuda, R. I.; Sato, T.; Kanoo, H. P.; Jeon, H. J.; Foo,
oxidation reactions. We found that all benzylic alcohols with M. L.; Wakamiya, A.; Murata, Y.; Kitagawa, S.; Tanaka, I. J. Am. Chem.
Soc. 2014, 136, 7543−7546.
both electron-withdrawing (−F, −NO2) and electron-donating (7) Sheldon, R. A.; Arends, I. W. C. E.; ten Brink, G.-J.; Dijksman, A.
groups (−OMe, −Me) at either para- or meta-position gave Acc. Chem. Res. 2002, 35, 774−781.
excellent conversions (93−99%). In addition, the reaction (8) Tonucci, L.; Nicastro, M.; d'Alessandro, N.; Bressan, M.;
selectivity toward benzaldehydes is basically 100% in all cases. D’Ambrosio, P.; Morvillo, A. Green Chem. 2009, 11, 816−820.
Therefore, Pd@Cu(II)-MOF is indeed a good heterogeneous (9) Edwards, J. K.; Pritchard, J.; Lu, L.; Piccinini, M.; Shaw, G.;
catalyst for aromatic carbinols oxidation. Carley, A. F.; Morgan, D. J.; Kiely, C. J.; Hutchings, G. J. Angew. Chem.,

■ CONCLUSIONS
A new porous Cu(II)-MOF (1) with a NbO network was
Int. Ed. 2014, 53, 2381−2384.
(10) Klotter, F.; Studer, A. Angew. Chem., Int. Ed. 2014, 53, 2473−
2476.
(11) Guo, Z.; Liu, B.; Zhang, Q. H.; Deng, W. P.; Wang, Y.; Yang, Y.
synthesized. Pd NPs-loaded Pd@Cu(II)-MOF (2) was H. Chem. Soc. Rev. 2014, 43, 3480−3524.
prepared based on 1 by impregnation and reduction reaction (12) Miyamura, H.; Matsubara, R.; Miyazaki, Y.; Kobayashi, S. Angew.
in solution. The Cu(II)-MOF with fluorine atoms can Chem., Int. Ed. 2007, 46, 4151−4154.
effectively stabilize the palladium nanoparticles in the crystals. (13) Shang, C.; Liu, Z. P. J. Am. Chem. Soc. 2011, 133, 9938−9947.
More importantly, the obtained Pd NPs-embedded Pd@ (14) Asao, N.; Hatakeyama, N.; Menggenbateer, T.; Minato, E. I.;
Cu(II)-MOF can be a highly active heterogeneous catalyst Hara, M.; Kim, Y.; Yamamoto, Y.; Chen, M.; Zhang, W.; Inoue, A.; Ito,
for the oxidation of aromatic carbinols to the corresponding E. Chem. Commun. 2012, 48, 4540−4542.
benzaldehydes in the air with high conversion rates and almost (15) Tsukamoto, D.; Shiraishi, Y.; Sugano, Y.; Ichikawa, S.; Tanaka,
100% selectivity. The catalyst can be recovered and reused for S.; Hirai, T. J. Am. Chem. Soc. 2012, 134, 6309−6315.
(16) Liu, H. L.; Liu, Y. L.; Li, Y. W.; Tang, Z. Y.; Jiang, H. F. J. Phys.
at least six runs without a significant loss in its activity and
Chem. C 2010, 114, 13362−13369.
selectivity. Further work on exploring new catalytic reactions (17) Yang, X.; Huang, C.; Fu, Z. Y.; Song, H. Y.; Liao, S. J.; Su, Y. L.;
with such NPs@MOF catalytic systems is currently underway.


Du, L.; Li, X. J. Appl. Catal., B 2013, 140−141, 419−425.
(18) Cui, W. J.; Xiao, Q.; Sarina, S.; Ao, W. L.; Xie, M. X.; Zhu, H. Y.;
ASSOCIATED CONTENT Bao, Z. Catal. Today 2014, 235, 152−156.
*
S Supporting Information (19) Cao, E. H.; Sankar, M.; Nowicka, E.; He, Q.; Morad, M.;
The Supporting Information is available free of charge on the Miedziak, P. J.; Taylor, S. H.; Knight, D. W.; Bethell, D.; Kiely, C. J.;
ACS Publications website at DOI: 10.1021/acs.inorg- Gavriilidis, A.; Hutchings, G. J. Catal. Today 2013, 203, 146−152.
chem.5b02973. (20) Wang, H. W.; Wang, C. L.; Yan, H.; Yi, H.; Lu, J. L. J. Catal.
2015, 324, 59−68.
Single-crystal data, TGA traces, XRPD patterns, photo- (21) Wang, J. C.; Kondrat, S. A.; Wang, Y. Y.; Brett, G. L.; Giles, C.;
graphs for Pd(II)@Cu(II)-MOF, ICP measurement, Bartley, J. K.; Lu, L.; Liu, Q.; Kiely, C. J.; Hutchings, G. J. ACS Catal.
HRTEM images, EDS measurement, and GC analysis 2015, 5, 3575−3587.
(PDF) (22) Sun, D. H.; Zhang, G. L.; Jiang, X. D.; Huang, J. L.; Jing, X. L.;
Crystallographic data for 1 (CIF) Zheng, Y. M.; He, J.; Li, Q. B. J. Mater. Chem. A 2014, 2, 1767−1773.


(23) Dhakshinamoorthy, A.; Alvaro, M.; Garcia, H. ACS Catal. 2011,
1, 48−53.
AUTHOR INFORMATION (24) Ding, S. Y.; Gao, J.; Wang, Q.; Zhang, Y.; Song, W. G.; Su, C. Y.;
Corresponding Authors Wang, W. J. Am. Chem. Soc. 2011, 133, 19816−19822.
*E-mail: yubindong@sdnu.edu.cn. (25) Iwasawa, T.; Tokunaga, I. M.; Obora, Y.; Tsuji, Y. J. Am. Chem.
*E-mail: gongjchen@126.com. Soc. 2004, 126, 6554−6555.

3063 DOI: 10.1021/acs.inorgchem.5b02973


Inorg. Chem. 2016, 55, 3058−3064
Inorganic Chemistry Article

(26) Chen, C.; Fang, X. L.; Wu, B. H.; Huang, L. J.; Zheng, N. F. (60) Williams, R. M.; Medlin, J. W. Langmuir 2014, 30, 4642−4653.
ChemCatChem 2012, 4, 1578−1586. (61) Benyounes, A.; Louisia, S.; Axet, R.; Mahfoud, Z.; Kacimi, M.;
(27) Wang, H. H.; Kong, W. P.; Zhu, W. J.; Wang, L. X.; Yang, S.; Serp, P. Catal. Today 2015, 249, 137−144.
Liu, F. J. Catal. Commun. 2014, 50, 87−94. (62) Wang, X. M.; Wu, G. J.; Guan, N. J.; Li, L. D. Appl. Catal., B
(28) Karimi, B.; Khorasani, M.; Vali, H.; Vargas, C.; Luque, R. ACS 2012, 115−116, 7−15.
Catal. 2015, 5, 4189−4200. (63) Ma, Z. C.; Yang, H. Q.; Qin, Y.; Hao, Y. J.; Li, G. J. Mol. Catal.
(29) Lee, K.-B.; Lee, S.-M.; Cheon, J. Adv. Mater. 2001, 13, 517−520. A: Chem. 2010, 331, 78−82.
(30) Kang, H.; Jun, Y. W.; Park, J. I.; Lee, K. B.; Cheon, J. Chem. (64) Michel; Dell'Anna, M. M.; Mali, M.; Mastrorilli, P.; Cotugno, P.;
Mater. 2000, 12, 3530−3532. Monopoli, A. J. Mol. Catal. A: Chem. 2014, 386, 114−119.
(31) Qiao, Z. A.; Zhang, P. F.; Chai, S. H.; Chi, M. F.; Veith, G. M.; (65) Hou, Z. S.; Theyssen, N.; Brinkmann, A.; Leitner, W. Angew.
Gallego, N. C.; Kidder, M.; Dai, S. J. Am. Chem. Soc. 2014, 136, Chem. 2005, 117, 1370−1373.
11260−11263. (66) Karimi, B.; Khorasani, M.; Vali, H.; Vargas, C.; Luque, R. ACS
(32) Li, F.; Zhang, Q. H.; Wang, Y. Appl. Catal., A 2008, 334, 217− Catal. 2015, 5, 4189−4200.
226. (67) Mori, K.; Hara, T.; Mizugaki, T.; Ebitani, K.; Kaneda, K. J. Am.
(33) Hou, Z.; Theyssen, N.; Brinkmann, A.; Klementiev, K. V.; Chem. Soc. 2004, 126, 10657−10666.
Grunert, W.; Buhl, M.; Schmidt, W.; Spliethoff, B.; Tesche, B.; (68) Harada, T.; Ikeda, S.; Hashimoto, F.; Sakata, T.; Ikeue, K.;
Weidenthaler, C.; Leitner, W. J. Catal. 2008, 258, 315−323. Torimoto, T.; Matsumura, M. Langmuir 2010, 26, 17720−17725.
(34) Parlett, C. M. A.; Bruce, D. W.; Hondow, N. S.; Lee, A. F.; (69) Wang, D.; Deraedt, C.; Salmon, L.; Labrugère, C.; Etienne, L.;
Wilson, K. ACS Catal. 2011, 1, 636−640. Ruiz, J.; Astruc, D. Chem.Eur. J. 2015, 21, 6501−6510.
(35) Harada, T.; Ikeda, S.; Hashimoto, F.; Sakata, T.; Ikeue, K.;
Torimoto, T.; Matsumura, M. Langmuir 2010, 26, 17720−17725.
(36) Wu, G. J.; Wang, X. M.; Guan, N. J.; Li, L. D. Appl. Catal., B
2013, 136−137, 177−185.
(37) Dimitratos, N.; Villa, A.; Wang, D.; Porta, F.; Su, D. S.; Prati, L.
J. Catal. 2006, 244, 113−121.
(38) Villa, A.; Janjic, N.; Spontoni, P.; Wang, D.; Su, D. S.; Prati, L.
Appl. Catal., A 2009, 364, 221−228.
(39) Prestianni, A.; Ferrante, F.; Sulman, E. M.; Duca, D. J. Phys.
Chem. C 2014, 118, 21006−21013.
(40) Uozumi, Y.; Nakao, R. Angew. Chem., Int. Ed. 2003, 42, 194−
197.
(41) Dun, R. R.; Wang, X. G.; Tan, M. W.; Huang, Z.; Huang, X. M.;
Ding, W. Z.; Lu, X. G. ACS Catal. 2013, 3, 3063−3066.
(42) Tanaka, A.; Hashimoto, K.; Kominami, H. J. Am. Chem. Soc.
2012, 134, 14526−14533.
(43) Sarina, S.; Zhu, H.; Jaatinen, E.; Xiao, Q.; Liu, H.; Jia, J.; Chen,
C.; Zhao, J. J. Am. Chem. Soc. 2013, 135, 5793−5801.
(44) Dybtsev, D. N.; Nuzhdin, A. L.; Chun, H.; Bryliakov, K. P.;
Talsi, E. P.; Fedin, V. P.; Kim, K. Angew. Chem., Int. Ed. 2006, 45,
916−920.
(45) Farrusseng, D.; Aguado, S.; Pinel, C. Angew. Chem. 2009, 121,
7638−7649; Angew. Chem., Int. Ed. 2009, 48, 7502−7513.
(46) Ma, L.; Abney, C.; Lin, W. B. Chem. Soc. Rev. 2009, 38, 1248−
1256.
(47) Yoon, M.; Srirambalaji, R.; Kim, K. Chem. Rev. 2012, 112,
1196−1231.
(48) Canivet, J.; Aguado, S.; Schuurman, Y.; Farrusseng, D. J. Am.
Chem. Soc. 2013, 135, 4195−4198.
(49) Zhang, Y. G.; Ying, J. Y. ACS Catal. 2015, 5, 2681−2691.
(50) Chughtai, A. H.; Ahmad, N.; Younus, H. A.; Laypkov, A.;
Verpoort, F. Chem. Soc. Rev. 2015, 44, 6804−6849.
(51) Guo, Z.; Xiao, C.; Maligal-Ganesh, R. V.; Zhou, L.; Goh, T. W.;
Li, X.; Tesfagaber, D.; Thiel, A.; Huang, W. ACS Catal. 2014, 4, 1340−
1348.
(52) Li, X.; Guo, Z.; Xiao, C.; Goh, T. W.; Tesfagaber, D.; Huang, W.
ACS Catal. 2014, 4, 3490−3497.
(53) Aijaz, A.; Xu, Q. J. Phys. Chem. Lett. 2014, 5, 1400−1411.
(54) Hou, G. G.; Liu, Y.; Liu, Q. K.; Ma, J. P.; Dong, Y. B. Chem.
Commun. 2011, 47, 10731−10733.
(55) Sheldrick, G. M. SHELXS 97: Program for the Solution of Crystal
Structures; University of Göttingen: Göttingen, Germany, 1997.
(56) Sheldrick, G. M. SHELXL 97: Program for the Refinement of
Crystal Structures; University of Göttingen: Göttingen, Germany, 1997.
(57) Huang, X. Q.; Tang, S. H.; Yang, J.; Tan, Y. M.; Zheng, N. F. J.
Am. Chem. Soc. 2011, 133, 15946−15949.
(58) Cheng, J. Y.; Wang, P.; Ma, J. P.; Liu, Q. K.; Dong, Y. B. Chem.
Commun. 2014, 50, 13672−13675.
(59) Rana, S.; Maddila, S.; Jonnalagadda, S. B. Catal. Sci. Technol.
2015, 5, 3235−3241.

3064 DOI: 10.1021/acs.inorgchem.5b02973


Inorg. Chem. 2016, 55, 3058−3064

You might also like