Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles

1 downloads 0 Views 3MB Size Report
Feb 3, 2014 - She is now performing graduate research with Professor Hyunjoon Song at ...... J. Choi, Y. Jun, S. Yeon, H. C. Kim, J. Shin, and J. Cheon, J. Am.
Review Journal of Nanoscience and Nanotechnology

Copyright © 2014 American Scientific Publishers All rights reserved Printed in the United States of America

Vol. 14, 1872–1883, 2014 www.aspbs.com/jnn

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles Aram Kim1 , Ji Chan Park2 , Mijong Kim3 , Eunjung Heo1 , Hyunjoon Song3 , and Kang Hyun Park1 ∗ 1

Department of Chemistry and Chemistry Institute for Functional Materials, Pusan National University, Busan 609-735, Korea 2 Clean Fuel Department, Korea Institute of Energy Research, Daejeon 305-343, Korea 3 Department of Chemistry, Korea Advanced Institute of Science and Technology, Daejeon 305-701, Korea This paper reviews recent developments in the field of hybrid Pd nanoparticles and their catalytic activity in the Suzuki coupling reaction, which is used extensively in the fabrication of both simple and complex biaryl compounds. We developed three types of Pd-silica hybrid nanoparticles. Pd/SiO2 nanobeads containing tiny Pd clusters, Pd@nickel phyllosilicate yolk-shell nanoparticles, Pd@porous SiO2 yolk-shell nanoparticles were synthesized, and they displayed highly efficient catalytic activity and excellent reusability. The hybrid nanoparticles also catalyzed the Suzuki coupling reaction with various substrates, including bromobenzene and chlorobenzene. This review also Deliveredbriefly by Publishing Technology to: Korea Advanced Institute ofand Science Technology discusses the synthesis procedure, structural characterization, catalytic&activity of hybrid(KAIST) Pd nanoparticles. IP: 143.248.37.33 On: Mon, 03 Feb 2014 07:05:42

Copyright: American Scientific Publishers Keywords: Suzuki Coupling Reaction, Palladium, Silica, Hybrid, Nanocatalyst, Core–Shell, Yolk-Shell.

CONTENTS 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. General Experimental Details . . . . . . . . . . . . . . . . . . . . . . . . . . 2.1. General Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2. Synthesis of Hybrid Pd/SiO2 Nanobeads . . . . . . . . . . . . . . 2.3. Synthesis of Pd Nanoparticles . . . . . . . . . . . . . . . . . . . . . . 2.4. Preparation of Hybrid Pd@Nickel Phyllosilicate Yolk-Shell Nanoparticles . . . . . . . . . . . . . . . . . . . . . . . . . . 2.5. Synthesis of Hybrid Pd@pSiO2 Yolk-Shell Nanoparticles . 3. Hybrid Pd Nanoparticles Catalyzed Suzuki Coupling Reaction . 3.1. Suzuki Coupling Reaction Using Hybrid Pd/SiO2 Nanobeads . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Suzuki Coupling Reaction Using Hybrid Pd@Nickel Phyllosilicate Yolk-Shell Nanoparticles . . . . . . . . . . . . . . . 3.3. Suzuki Coupling Reaction Using Hybrid Pd@pSiO2 Yolk-Shell Nanoparticles . . . . . . . . . . . . . . . . . 4. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References and Notes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1872 1874 1874 1875 1875 1875 1875 1875 1875 1877 1878 1881 1881 1881

1. INTRODUCTION Since its discovery in 1979,1 the Suzuki coupling reaction has evolved into a broadly useful carbon–carbon bond ∗

Author to whom correspondence should be addressed.

1872

J. Nanosci. Nanotechnol. 2014, Vol. 14, No. 2

transformation for synthesizing biaryls. The reaction is important for various applications in the chemical, pharmaceutical, and biochemical industries.2–12 The main problems associated with the Suzuki coupling reaction are: (i) activation of aryl chlorides,13 14 which generally show low conversion and yields under normal conditions, and (ii) achievement of easy separation and reusability of the catalyst.13 15–21 Using Pd nanoparticles as a catalyst is not only industrially important, but also scientifically interesting since they provide details of the sensitive correlation between the catalytic activity and the features of particles size and shape as well as the nature of the surrounding media.22 23 The well-known application of Pd nanoparticles as a catalyst in organic synthesis is in reactions involving transformations, including the hydrogenation of alkenes and alkynes,24–26 the oxidation of alcohols,27 and carbon–carbon coupling reactions.28 29 Especially carbon–carbon coupling reactions which include the Heck,30 31 Sonogashira,31 32 Suzuki,17 31 33–40 Stlle,41 Negishi,42 Hiyama,43 CorriuKumada,44 Tsuji-Trost,45 and Ullmann reactions46 are used as excellent methods for the synthesis of natural products, pharmaceutical products, fine chemicals and the 1533-4880/2014/14/1872/012

doi:10.1166/jnn.2014.9100

Kim et al.

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles

manufacturing of long chain organic molecules for organoelectronics applications. Pd nanoparticles are widely used as catalysts in organic reactions; in particular, they are the most active heterogeneous catalyst in the Suzuki coupling reaction.16 17 33–40 Compared to homogeneous catalysts, heterogeneous catalysts offer many advantages, including easy recovery of the catalyst after the reaction and catalyst recyclability.33 47–57 The synthesis of well-defined nanoparticles is of special interest in nanoscience, particularly for their catalytic applications due to the interesting properties associated

with their nanoscale dimension. As metal and metal oxide nanoparticles are very stable both physically and chemically, they have been frequently used as catalysts.58 59 In addition, their distinct characteristics as nanoparticles with large surface areas make them applicable to a wide range of fields. In recent years, various studies have been conducted into hybrid nanostructure, with the aim of designing and synthesizing multicomponent structures that combine the physical and chemical properties of the individual components.60–70 Many efforts have been made to

Aram Kim received her B.S. (2010) and M.S. (2013) in inorganic chemistry at Pusan National University. She is currently working towards a Ph.D. in chemistry under the supervision of K. H. Park at Pusan National University. Her research interests are in development of transition metal nanoparticles and organometallic complexes, and their application in organic reactions.

Ji Chan Park received his M.S. and Ph.D. from Korea Advanced Institute of Science and Technology (KAIST) in 2007 and 2010. Since 2010, he has been working as a senior researcher at Korea Institute of Energy Research (KIER). His main research interests Delivered by Publishing Technology to: Korea Advanced Institute of Science & Technology (KAIST) are nanocatalysts reactions such as hybrid nanostructures and IP:heterogeneous 143.248.37.33 On: Mon,and 03 catalytic Feb 2014 07:05:42 Fischer-Tropsch synthesis. Copyright: American Scientific Publishers

Mijong Kim earned a B.S. degree in chemistry from Sungkyunkwan University in 2009. She is now performing graduate research with Professor Hyunjoon Song at Korea Advanced Institute of Science and Technology (KAIST) as a Ph.D. candidate. Her current research focuses on the fabrication of metal/metal oxide nanostructures, and their catalytic reactions for organic transformation.

Eunjung Heo received her B.S. and M.S. in chemistry under the supervision of K. H. Park at Pusan National University. Currently, she is a research engineer in polyolefin R&D group at Hanwha Chemical.

J. Nanosci. Nanotechnol. 14, 1872–1883, 2014

1873

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles

Kim et al.

Hyunjoon Song received his M.S. and Ph.D. from Korea Advanced Institute of Science and Technology (KAIST) in 1996 and 2000. He worked as a postdoctoral fellow at KAIST and University of California, Berkeley until 2004. He was an Assistant Professor in 2005, and promoted to an Associate Professor in 2008 in the department of chemistry at KAIST. His research interests are morphology control of metal nanocrystals and their application for surface plasmon monitoring and energy catalysts.

Kang Hyun Park received his Ph.D. in 2005 under the supervision of Y. K. Chung and worked on the synthesis of transition metal nanoparticles and their application in organic reactions. He did his postdoctoral work with Professor Seung UK Son at Sungkyunkwan University in 2006–2007. In 2008, he was appointed as an Assistant Professor and Associated Professor at Pusan National University. His research interests include the development of new transition metal-nanoparticles catalyzed reactions.

synthesize these multicomponent nanostructures in order demonstrate highly efficient catalytic activity and excellent to achieve increased functionality.71–81 This increase in reusability. In addition, the relationship between the catafunction, combined with their enhanced chemical and lyst structure and catalytic activity is discussed. physical properties, makes these hybrid nanostructures Delivered by Publishing Technology to: Korea Advanced Institute of Science & Technology (KAIST) IP: such 143.248.37.33 applicable in many research fields as magnetics,On: plas-Mon, 03 Feb 2014 07:05:42 2. GENERAL EXPERIMENTAL DETAILS Copyright: metalAmerican nanopar- Scientific Publishers monics, and semiconducting.82 In addition, 2.1. General Remarks ticles immobilized on metal oxide supports can be applied Reagents were purchased from Aldrich Chemical Co. and to an organic reaction as catalysts. They have the advanStrem Chemical Co. and used as received. Reaction prodtages of higher catalytic activity and better recyclability ucts were analyzed by 1 H NMR. 1 H NMR with specthan are offered by single metal nanoparticles, because of tra were obtained on a Varian Mercury Plus spectrometer the electron transfer across the interface as well as their (300 MHz). Chemical shift values were recorded as parts heterogeneous properties.31 52 83–87 per million relative to tetramethylsilane as an internal stanFor example, core–shell nanostructures, which generally dard, unless otherwise indicated, and coupling constants consist of an encapsulated Pd core particle and a porous were in Hertz. Reaction products were assigned by comsilica shell, exhibit high catalytic activity; their stability is parison with the literature value of known compounds. also superior to that of the corresponding unencapsulated The nanoparticles were characterized by high-resolution core–shell nanostructures. These core–shell structures have transmission electron microscopy (HRTEM; Philips F20 high thermal stability and they completely prevent particle Tecnai operated at 200 kV, KAIST) and high-angle annuaggregation/agglomeration, which makes them promising lar dark field transmission electron microscopy (HAADFrecoverable and efficient catalysts for various liquid-phase TEM; Hitachi HD-2300A operated at 200 kV; National catalytic reactions.88–95 The use of Pd nanoparticles on Nanofabrication Center, NNFC, at KAIST). Samples were widely available active carbon or inorganic supporting prepared by placing a few drops of the corresponding colmaterials has also been found to be an efficient heterogeloidal solutions on carbon-coated copper grids (Ted Pelneous catalyst for the Suzuki reaction.96 lar, Inc.). X-ray powder diffraction (XRD) patterns were In this review, we highlight recent advances made in the recorded on a Rigaku D/MAX-RB (12 kW) diffractomedevelopment of hybrid Pd nanoparticles and discuss their ter. The Pd loading amount was measured by inductively application to the Suzuki coupling reaction of aryl halides coupled plasma-atomic emission spectrometry (ICP-AES; and arylboronic acid. This review also introduces three POLY SCAN 60 E). Nitrogen sorption isotherms were types of hybrid Pd nanoparticles: (i) Pd/SiO2 nanobeads measured at 77 K using a BELSORP mini-II (BEL Japan containing tiny Pd clusters;97 (ii) Pd@nickel phyllosiliInc.). Before measurements, the samples were degassed in cate yolk-shell nanoparticles;98 and (iii) Pd@porous SiO2 a vacuum at 423 K for at least 6 h. yolk-shell nanoparticles;99 here, they are shown to clearly 1874

J. Nanosci. Nanotechnol. 14, 1872–1883, 2014

Kim et al.

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles

2.2. Synthesis of Hybrid Pd/SiO2 Nanobeads In a 125 mL polypropylene (PP) bottle containing 50 mL of cyclohexane and 1.0 mL of NH4 OH solution (28% aqueous solution), Na2 PdCl4 (0.2 g, 0.68 mol) was added and completely mixed by sonication for 20 min. After sufficient stirring for 20 min and the subsequent addition of 8.0 mL of Igepal CO-630 to the mixture solution, a pale yellow sediment was formed at the bottom of the mixture solution. Then, 1.0 mL of TMOS and 1.0 mL of C18 TMS were simultaneously added and stirred at 200 rpm for 1 h. After 1 h, the products were precipitated by adding 15 mL of methanol and centrifuging at 10,000 rpm for 30 min. The precipitates were washed by carrying out a repetitive dispersion/precipitation cycle in ethanol; then, the precipitates were dried with acetone. Finally, the Pd(II) complexes/SiO2 were obtained as a yellowish powder. The yellowish Pd(II) complexes/SiO2 powders were placed in a ceramic boat in a glass tube oven, and they were slowly heated at a ramping rate of 4 K/min up to 773 K under a hydrogen flow of 200 cc/min. In order to carry out complete reduction of the Pd(II) complexes to Pd(0) clusters and to remove organic moieties from the sample, the samples were allowed to stand at 773 K for 4 h under the continuous hydrogen flow. After the reduction, the resulting dark brown Pd/SiO2 powder was cooled to room temperature.

2.4.3. Preparation of Pd@SiO2 –Niphy Nanoparticles (3) The synthetic procedure is identical to that of 2, except the added amount of Ni(OAc)2 · 4H2 O (50 mL, 0.16 M, 1.2 equiv with respect to the silica precursor concentration). 2.4.4. Preparation of Pd@Niphy Yolk-Shell Nanoparticles (4) 3 was dispersed in ethanol (60 mL). HF (0.20 mL) was added into the dispersion (30 mL), and the mixture was stirred for 1 h at room temperature. The product was centrifuged and thoroughly washed with ethanol.

2.5. Synthesis of Hybrid Pd@pSiO2 Yolk-Shell Nanoparticles Cyclohexane (25 mL) was mixed with igepal CO-630 (8.0 mL) and aqueous ammonia solution (0.80 mL), and the mixture was stirred for 20 min. The Pd particle dispersion in cyclohexane (25 mL, 6 mM with respect to the precursor concentration) was added into the mixture and allowed to stir for 30 s. TMOS (1.2 mL) and C18 TMS (1.2 mL) were simultaneously added into the reaction mixture, and the mixture was stirred at room temperature for 1 h. After the reaction, the product was precipitated by 30 mL of methanol, and thoroughly washed with ethanol. The aqueous dispersion of the Pd@SiO2 2.3. Synthesisby ofPublishing Pd Nanoparticles core–shell nanoparticles (20 mL, 7.5 mM with(KAIST) respect to Delivered Technology to: Korea Advanced Institute of Science & Technology concentration) 143.248.37.33 On: Mon,the 03precursor Feb 2014 07:05:42 was put in a plastic bottle and A mixture of Pd(acac)2 (91 mg,IP: 0.30 mmol), trioctylphosCopyright: Publishers placed in an oven at 383 K for 12 h. After the reaction, the phine (1.0 mL, 2.3 mmol), and oleylamine (10 American mL) was Scientific product was precipitated by centrifugation at 10,000 rpm heated to 230  C for 20 min under a nitrogen atmosphere for 20 min, thoroughly washed with water and ethanol, and was allowed to stir for an additional 40 min. The reacand dried in air to yield dark brown powders. The resulttion mixture was cooled down to room temperature, and ing Pd@pSiO2 yolk-shell nanoparticles were placed in a the Pd nanoparticles were collected by centrifugation in ceramic boat in a glass tube oven, slowly heated in a rampethanol. The black precipitate was redispersed in cycloing rate of 4 K/min from room temperature to 773 K, and hexane (50 mL). aged at the same temperature for 4 h under a continuous hydrogen flow of 200 cm3 /min. After the thermal treat2.4. Preparation of Hybrid Pd@Nickel Phyllosilicate ment, the resulting dark brown powders were cooled down Yolk-Shell Nanoparticles to room temperature. 2.4.1. Preparation of Pd@SiO Core–Shell 2

Nanoparticles (1) Cyclohexane (25 mL) was mixed with igepal CO-630 (8.0 mL) and an aqueous ammonia solution (0.80 mL, 28% in water), and the mixture was stirred for 20 min. The Pd nanoparticle dispersion in cyclohexane (25 mL) and TMOS (1.0 mL) were subsequently added, and the resulting mixture was stirred for 1 h at room temperature. The product was precipitated by the addition of methanol (45 mL) and thoroughly washed with ethanol. 2.4.2. Preparation of Pd@SiO2 –Niphy Nanoparticles (2) 1 was dispersed in an aqueous solution of Ni(OAc)2 ·4H2 O (50 mL, 0.080 M, 0.59 equiv with respect to the silica precursor concentration). The mixture was heated under reflux for 1 h. The product was precipitated by centrifugation and thoroughly washed with ethanol. J. Nanosci. Nanotechnol. 14, 1872–1883, 2014

3. HYBRID Pd NANOPARTICLES CATALYZED SUZUKI COUPLING REACTION 3.1. Suzuki Coupling Reaction Using Hybrid Pd/SiO2 Nanobeads Pd/SiO2 nanobeads consisting of tiny Pd nanoparticles in a porous silica sphere were synthesized in two simple steps. In the first step, the encapsulation of Pd(II) complexes in a silica bead was carried out using microemulsions,100 101 and in the second, the complete reduction of the Pd(II) complexes to Pd(0) clusters was carried out under a continuous hydrogen flow (Scheme 1). The Pd(II) complexes encapsulated inside the silica bead were obtained by adding a Pd precursor to a water-in-oil microemulsion, which consisted of Igepal CO-630 as a 1875

Kim et al.

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles

Table I. Pd/SiO2 nanobeads catalyzed Suzuki coupling reaction of bromobenzene with phenylboronic acid.

B(OH)2

Br +

Scheme 1. The proposed procedures for the formation of encapsulated Pd/SiO2 nanobead. Reprinted with permission from [97], M. Kim, et al., Catal. Lett. 142, 588 (2012). © 2012, Springer.

Entry

Pd (mol%)

Temp. (K)

Time (h)

Solvent

Conv. (%)a

001 001 05 001 0005

473 473 473 473 473

3 5 1.5 3 3

DMF DMF DMF:H2 O (10:0.5) DMF:H2 O (10:0.5) DMF:H2 O (10:0.5)

22 38 83 >99 65

1 2 3 4 5

nonionic surfactant, aqueous ammonia solution, and cyclohexane. The pale yellow sediment that was observed after complete sonication was encapsulated in silica spheres by adding tetramethyl orthosilicate (TMOS) as a silNote: a Determined by 1 H NMR spectra. K2 CO3 (2.0 equiv.) was used as a base. ica source; C18 TMS could be added as a porogen to introduce irregular pores via heat treatment. The transand a mixture of DMF and H2 O as a solvent at a reacmission electron microscopy (TEM) and high resolution tion temperature of 200  C, the expected coupling product transmission electron microscopy (HRTEM) images show obtained at a 99% conversion is achieved using only a the obtained spherical Pd(II) complexes/SiO2 nanobeads 0.01 mol% Pd/SiO2 catalyst (Table I, entry 4). The opti(Figs. 1(a) and (b)). In order to reduce the Pd(II) commal conditions were successfully obtained by adding a plexes to a Pd cluster, the Pd(II) complexes/SiO2 were small amount of water as a solvent. Also when the catalyst calcined at 773 K under a hydrogen atmosphere. Conseamount dropped in half, a satisfactory conversion of 65% quently, tiny Pd nanoclusters with an average diameter of was achieved (Table I, entry 5). 17 ± 03 nm can be confirmed in the dark spots in the To evaluate the scope of the substrate, we performed TEM image (Fig. 1(c)). The distance between the adjacent a number of coupling reactions using several aryl brolattice fringes, as calculated from the HRTEM image, is mides (Table II). In the optimized conditions, bromobenapproximately 0.224 nm, and the Pd nanocluster in the silzenes substituted by either an electron-withdrawing such ica shell has a single-crystalline nature (Fig. 1(d)). The Pd as NO , COMe, COOEt, CF , and CN or electron-donating nanobeads was determined to be loading in the Pd/SiO 2 Delivered by Publishing Technology to: Korea Advanced 2Institute of Science3 & Technology (KAIST) group such2014 as OMe produced the corresponding biaryls 8.27 wt% by inductively coupled plasma atomic emission IP: 143.248.37.33 On: Mon, 03 Feb 07:05:42 spectrometry (ICP-AES). Copyright: American Scientific Publishers The hybrid Pd/SiO2 nanobeads were employed as a catTable II. Pd/SiO2 nanobeads catalyzed Suzuki coupling reaction of various aryl bromide with phenylboronic acid.a alyst in the Suzuki coupling reaction of bromobenzene and phenylboronic acid. When bromobenzene was treated with Entry Aryl bromide Product Conv. (%)b phenylboronic acid under 2.0 equiv. of K2 CO3 as a base 1

82

O2N

2

Br

O2N

82

O

O Br

3

O

66

O

Br EtO

EtO

4

59 F3C

Br

F3C

NC

Br

NC

5

100

6

53 MeO

Br

MeO

7

100 Br

Figure 1. TEM and HRTEM images of (a), (b) Pd(II) complexes/SiO2 beads and (c), (d) Pd/SiO2 nanobeads. Inset images of (c), (d) present the size distribution of Pd nanoparticles and single Pd/SiO2 nanobead. The bars represent 5 nm (inset of (d)). Reprinted with permission from [97], M. Kim, et al., Catal. Lett. 142, 588 (2012). © 2012, Springer.

1876

O

O

Notes: a Reaction conditions: Aryl halide, arylboronic acid, Pd/SiO2 nanobeads (0.01 mol% with respect to aryl halide concentration) K2 CO3 (2.0 equiv) as a base, DMF:H2 O (10:0.5) at 473 K; b Determined by 1 H-NMR spectra.

J. Nanosci. Nanotechnol. 14, 1872–1883, 2014

Kim et al.

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles

at good conversion. Reactions of various aryl bromides including electron-withdrawing and electron-donating substrates proceeded readily at very low catalyst loading, and various functional groups were tolerated in the reaction. High reactivity with excellent yield was also observed in the coupling of activated o-bromobenzaldehyde with phenylboronic acid (Table II, entry 7). Furthermore, the reaction with electron-rich 4-methoxyphenylboronic acid exhibited a high reactivity to afford the corresponding 4-methoxybiphenyl in a 100% yield.

1.2 equiv., the average length and thickness of the branches were estimated to be 26 ± 3 and 13 ± 03 nm, respectively (Figs. 2(e) and (f)). In the core–shell nanoparticle, as the porosity of the shell increased, the surface of the core metal inside the silica shell became significantly more accessible by a reagent, and the reaction rate depended on the diffusion rate through the pores. In order to dissolve the silica and form pores surrounding the Pd core, the residual silica was completely etched with HF for 1 h at the room temperature. The Pd core and Niphy were resistant to HF; consequently, Niphy branches formed spherical hol3.2. Suzuki Coupling Reaction Using Hybrid low shells surrounding the Pd core (Figs. 2(g) and (h)). Pd@Nickel Phyllosilicate Yolk-Shell Nanoparticles The average size of the Pd cores is estimated to be 41 ± Well-defined Pd@SiO2 bifunctional nanostructures were 03 nm, which is almost identical to their original size, manufactured via the formation of nickel phyllosilicate and the average thickness of the Niphy shells is estimated (Niphy), Ni3 Si2 O5 (OH)4 , which generates a high pore dento be 26 ± 03 nm (Figs. 2(g) and (h)).We synthesized sity in the silica shell.102–105 The synthesis procedure is four types of hybrid Pd nanoparticles: Pd@SiO2 core–shell shown in Scheme 2. Tiny Pd nanoparticles, which had an (1), Pd@SiO2 -Niphy (2, 3), and Pd@Niphy yolk-shell (4). average diameter of 42 ± 04 nm, were synthesized via the All the four types of nanoparticles were analyzed by N2 thermal decomposition of Pd-oleylamine complexes. The adsorption at 77 K and BET (Brunauer–Emmett–Teller) TEM image shows uniform spherical and monodisperse measurements (Fig. 3). The obtained results showed the Pd nanoparticles (Fig. 2(a)). The uniform Pd nanopartifollowing order of the pore volume and surface area: 1 < cles were coated with a silica shell when TMOS was used 2 < 3 < 4. The pore volumes of catalysts 1–4 were 0.502, as a silica source; the resulting silica shell of Pd@SiO2 0.585, 0.744, and 0.996 cm3 g−1 , and the surface areas core–shell nanoparticles had a thickness of approximately were 170, 493, 523, and 551 m2 g−1 , respectively, as esti79 ± 03 nm (Fig. 2(b)). mated by BET measurements. In order to by prepare the Pd@SiO Delivered Publishing Technology to: Korea Advanced Institute of Science & Technology (KAIST) 2 -Niphy nanoparticles, In order to compare the catalytic activity of hybrid were dispersed in the Pd@SiO2 core–shell nanoparticles IP: 143.248.37.33 On: Mon, 03 Feb 2014 07:05:42 nanoparticles 1–4, the Suzuki coupling reactions were carCopyright: Publishers an aqueous solution of nickel acetate solution American and then Scientific ried out in the presence of 2 mol% Pd with respect to heated under reflux. Consequently, needle-like branches the substrate using 2 equiv. of Cs2 CO3 as a base, with were generally formed on the surface of the silica shell as a mixture of ethanol and water at room temperature for a result of basification of the nickel(II) solution under a 15 h (Table III). The reaction of 1-bromo-4-ethylbenzene hydrothermal condition. Interestingly, two different experand phenylboronic acid carried out by employing cataiments were carried out by using different amounts of the lyst 1 resulted in a low conversion of 42%; on the other added nickel(II) solution with respect to the silica precurhand, when the same reaction was carried out with catsor concentration. It was found that as the added amount alyst 4, a conversion of greater than 99% of the desired of nickel salts was increased from 0.59 equiv. to 1.2 equiv., product was achieved (Table III, entries 1 and 4). In the longer needle-like branches were formed that penetrated control experiment that employed a SiO2 -Niphy hollow the silica shells more deeply, with the maximum penetraparticle without Pd, no reaction occurred because of the tion depth of 11 ± 1 nm (Figs. 2(e) and (f)). In the case absence of the Pd core (Table III, entry 9). In the cases of 0.59 equiv., the average length and thickness of the of both 1 mol% and 2 mol% Pd loading, the catalytic branches were estimated to be 11 ± 2 and 13 ± 02 nm, activity followed the sequence: 4 > 3 > 2 > 1 (Table III, respectively, with an average penetration depth of approxentries 1–8). These results are reasonable, considering that imately 36 ± 04 nm (Figs. 2(c) and (d)). In the case of

Scheme 2. The proposed procedures for the formation of four distinct catalyst structures: Pd@SiO2 core–shell (1), Pd@SiO2 –Niphy (2, 3), and Pd@Niphy yolk-shell (4) nanoparticles. Reprinted with permission from [98], M. Kim, et al., Langmuir 28, 6441 (2012). © 2012, American Chemical Society.

J. Nanosci. Nanotechnol. 14, 1872–1883, 2014

1877

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles

Kim et al.

demonstrated complete conversion, was reused more than five times without any loss of conversion, and it displayed remarkable aggregation-resistant behaviors because of the chemical inertness of the Niphy shells under the basic reaction conditions used in this study. The shells also maintained their original structure, with the Pd cores surrounded by branched hollows. 3.3. Suzuki Coupling Reaction Using Hybrid Pd@pSiO2 Yolk-Shell Nanoparticles The fabrication of Pd@pSiO2 yolk-shell nanoparticles, which bore tiny palladium cores and highly porous silica hollow shells, was achieved by direct silica coating, partial etching, and thermal treatment of the silica layers. To accomplish the extremely high catalytic activity through the large pores generated on the silica shells, all surfactants around the catalyst surface were removed at a high temperature (Scheme 3). The first step in the synthesis was accomplished by thermal decomposition of the Pd-surfactant complexes.106 A mixture of palladium(II) acetylacetonate and trioctylphosphine (TOP) was slowly heated to 503 K and then aged. The TEM images clearly show that the resulting Pd nanoparticles are uniformly dispersed and spherical, with an average diameter of 34 ± 02 nm (Fig. 5(a)). In the HRTEM image, the atomic lattice fringes, whose distance betweenofadjacent images was(KAIST) 0.225 nm, Delivered by Publishing Technology to: Korea Advanced Institute Sciencefringe & Technology correspond to the (111) planes in the face-centered cubic IP: 143.248.37.33 On: Mon, 03 Feb 2014 07:05:42 (fcc) Pd (Fig. 5(b)). In the second step, more than 90% Copyright: American Scientific Publishers of the Pd nanoparticles were successfully coated with a silica layer, using Igepal CO-630 as a nonionic surfactant, TMOS as a silica source, and C18 TMS as a porogen, following a previously reported water-in-oil microemulsion method.107 Igepal CO-630 forms a microemulsion. Ammonia catalyzes the decomposition of the silica precursor on the surface of the Pd core during the reaction. The TEM image of the Pd@SiO2 core–shell nanoparticles indicates that the average thickness of the silica shells is 97 ± 15 nm (Figs. 5(c) and (d)). Consequently, through the partial dissolution of the silica under a benign basic soluFigure 2. TEM images of (a) Pd, (b) Pd@SiO2 core–shell (1), (c)–(f) tion and hydrothermal etching, dark brown powders were Pd@SiO2 -Niphy (2, 3), and (g), (h) Pd@Niphy yolk-shell nanopartiobtained as Pd@SiO2 yolk-shell nanoparticles. As shown cles (4). Reprinted with permission from [98], M. Kim, et al., Langmuir in the TEM images, the Pd@SiO2 yolk-shell nanoparticles 28, 6441 (2012). © 2012, American Chemical Society. had a silica shell that is less dense than that of the core– shell nanoparticles, and which consists of multiple large the pore volumes of the silica shells and the porosity of pores (Figs. 6(a) and (b)). the shells influence the diffusion rate, which leads to an Next, a thermal treatment at 773 K under a hydrogen efficient catalytic reaction. flow generated Pd@pSiO2 yolk-shell nanoparticles, with a To test the recyclability, catalyst 1 and 4 were used total pore volume of 0.57 cm3 g−1 and a surface area of as catalysts under the condition of 2 mol% Pd catalysts 145 m2 g−1 , through the removal of all organic residues at room temperature for 24 h (Fig. 4). After the fifth such as C18 TMS and surfactants. The tiny Pd nanopartirecycling, the conversion yield was significantly dropped cles had an average core size of 35 ± 03 nm, which is to 45%, and the silica layers of the recovered Pd@SiO2 identical to that of the original Pd nanoparticles, thus indicore–shell nanoparticle (1) were partially dissolved and cating the high thermal stability of the tiny Pd (Fig. 6(c)). the particles were found to agglomerate. On the other The Pd loading in Pd/SiO2 yolk-shell nanoparticles was hand, the Pd@Niphy yolk-shell nanoparticle (4), which determined to be 1.5 wt% by ICP-AES. 1878

J. Nanosci. Nanotechnol. 14, 1872–1883, 2014

Kim et al.

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles

Figure 3. (a) N2 adsorption–desorption isotherms and (b) total pore volume and surface area analyses of 1–4. Reprinted with permission from [98], M. Kim, et al., Langmuir 28, 6441 (2012). © 2012, American Chemical Society.

The Pd@pSiO2 yolk-shell nanocatalysts were employed core–shell nanoparticles, and Pd@pSiO2 yolk-shell as a catalyst in the Suzuki coupling reaction. Experiments nanoparticles, prior to thermal treatment under optimal were carried out at different temperatures and for differconditions (Table IV, entries 8–10). The order of the ent reaction times (Table IV, entries 1–3). The coupling of conversion yield is: Pd@pSiO2 yolk-shell nanocatalysts bromobenzene and phenylboronic acid resulted in 100% (100%) > Pd@pSiO2 yolk-shell nanoparticles before therconversion after 6 h at room temperature in the presmal treatment (72%) > Pd@SiO2 core–shell nanoparticles ence of 0.1 mol% Pd catalysts (Table IV, entry 1). The (40%) > free-standing Pd nanoparticles (35%). The factors reaction temperature is one of the important factors in that lead to a high conversion are: (i) completely exposed maximizing the reaction rate. At 473 K, the conversion surfaces of the Pd cores, and (ii) rapid diffusion of the approached 100% for a reaction time of 3 h in the presreactants through the large pores in the silica layers genence of 0.005 mol% Pd catalysts. The best results were erated by the thermal treatment. obtained, with 100% conversion, when a mixture of DMF To evaluate the scope of this reaction condition, we Delivered Technology Korea Advanced Institute of Science & Technology (KAIST) and H2 O was by usedPublishing as the solvent and Cs2 COto: performed a number of coupling reactions using several 3 as the base IP:of143.248.37.33 On:KMon, 03 Feb 2014 07:05:42 in the presence of 0.003 mol% the catalyst at 473 Copyright: American Scientific Publishers (Table IV, entry 5). Because the yolk-shell nanoparticles displayed a notable advantage in high-temperature stability, which is consistent with previous reports,108–113 it is possible to conduct the reaction at a high temperature. In addition, in order to investigate the factors that lead to such catalytic activity of yolk-shell nanoparticles, we carried out a control experiment using different catalysts, including free-standing Pd nanoparticles, Pd@SiO2 Table III. Pd@Niphy yolk-shell nanoparticles catalyzed Suzuki coupling reactions of 1-bromo-4-ethylbenzene with phenylboronic acid. B(OH)2

Br +

Entry 1 2 3 4 5 6 7 8 9

EtOH:H2O (10:0.4)

Catalysts

Pd (mol%)

Temp.

Time (h)

Conv. (%)a

Pd@SiO2 core–shell, 1 Pd@SiO2 -Niphy, 2 Pd@SiO2 -Niphy, 3 Pd@Niphy yolk-shell, 4 Pd@SiO2 core–shell, 1 Pd@SiO2 -Niphy, 2 Pd@SiO2 -Niphy, 3 Pd@Niphy yolk-shell, 4 SiO2 -Niphy

2 2 2 2 1 1 1 1 –

r.t. r.t. r.t. r.t. r.t. r.t. r.t. r.t. r.t.

15 15 15 15 36 36 36 36 24

42 81 89 100 (92) 71 76 87 >99 (98) 0

Notes: a All conversion yields were estimated by 1 H-NMR spectra, except the isolated yields in parentheses; Cs2 CO3 (2.0 equiv.) was used as a base.

J. Nanosci. Nanotechnol. 14, 1872–1883, 2014

Figure 4. (a) Conversion yields during five recycling runs with 1 and 4. (b), (c) TEM images of the recovered catalysts 1 and 4, respectively, after the fifth cycle. Reprinted with permission from [98], M. Kim, et al., Langmuir 28, 6441 (2012). © 2012, American Chemical Society.

1879

Kim et al.

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles

Scheme 3. The proposed procedures for the formation of Pd@pSiO2 yolk-shell nanoparticle. Reprinted with permission from [99], J. C. Park, et al., J. Phys. Chem. C 115, 15772 (2011), © 2011, American Chemical Society.

aryl halides and arylboronic acids that were catalyzed by Pd@pSiO2 yolk-shell nanocatalysts (Table V). In general, the order of reactivity of aryl halides in the Pd-catalyzed Suzuki coupling reaction is: I > Br  Cl. Using the Pd@pSiO2 yolk-shell nanoparticles as a catalyst produced a 100% conversion in the coupling of chlorobenzene with phenylboronic acid within 3 h (Table V, entry 1). Both bromobenzenes and chlorobenzenes with either electron-withdrawing substituents such as CHO and CF, or electron-donating substituents such as OCH3 , were successfully coupled with arylboronic acids (Table V, entries 2–7). Finally, it was found that the coupling of 4-bromobenzene with various substituted arylboronic acids such as electron-rich 4-methoxy and 4-methylphenylboronic acid was efficiently catalyzed yolk-shell in high yields by Pd@pSiO 2 of Delivered by Publishing Technology to: Korea Advanced Institute Sciencenanocatalysts & Technology (KAIST) (Table V, entries 8 and 9). IP: 143.248.37.33 On: Mon, 03 Feb 2014 07:05:42 Copyright: American Scientific Publishers In Pd-catalyzed cross-coupling reaction, it is important to investigate whether Pd catalyst system operates Figure 5. TEM and HRTEM images of (a), (b) Pd nanoparticles, (c), (d) Pd@SiO2 core–shell nanoparticles. Reprinted with permission in a homogeneous or heterogeneous manner. In our catfrom [99], J. C. Park, et al., J. Phys. Chem. C 115, 15772 (2011). © 2011, alytic system, it is confirmed that the Pd@pSiO2 yolkAmerican Chemical Society. shell catalysts appear to catalyze the coupling reactions in a heterogeneous manner. Poly(4-vinylpyridine), which trap homogeneous Pd species most strongly through chelation Table IV. Pd@pSiO2 yolk-shell nanoparticles catalyzed Suzuki coupling reactions of bromobenzene with phenylboronic acid. Br

Entry 1 2 3b 4 5 6 7 8c 9d 10e 11 Figure 6. TEM and HRTEM images of (a, b) Pd@pSiO2 yolk-shell nanoparticles and (c), (d) Pd@pSiO2 yolk-shell nanocatalysts after thermal treatment. Reprinted with permission from [99], J. C. Park, et al., J. Phys. Chem. C 115, 15772 (2011). © 2011, American Chemical Society.

1880

+

B(OH)2

Pd (mol %)

Temp. (K)

Time (h)

Base

Conv. (%)a

0.1 0.005 0.003 0.003 0.003 0.003 0.003 0.003 0.003 0.003 0.002

298 473 523 473 473 473 473 473 473 473 473

6 3 1 1 1 0.50 0.17 1 1 1 1

K2 CO3 K2 CO3 K2 CO3 Na2 CO3 Cs2 CO3 Cs2 CO3 Cs2 CO3 Cs2 CO3 Cs2 CO3 Cs2 CO3 Cs2 CO3

100 100 17 71 100 45 (30) 39 (23) 35 72 40 63

Notes: a Determined by 1 H-NMR spectra; b NMP/H2 O (20:1) as a solvent mixture; c Free-standing Pd nanoparticles; d Pd@pSiO2 yolk-shell nanoparticles before thermal treatment; e Pd@SiO2 core–shell nanoparticles; The values inside the parentheses are isolated yields of the product.

J. Nanosci. Nanotechnol. 14, 1872–1883, 2014

Kim et al.

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles

Table V. Pd@pSiO2 yolk-shell nanoparticles catalyzed Suzuki coupling reaction of various aryl bromide with arylboronic acid.a

Entry

Aryl bromide

Arylboronic acid

Cl

B(OH)2

1 2 MeO

Br

Br O

Time (h)

Yield (%)b

3

100

3

64

3

64

3

100

3

61

3

68

Acknowledgment: This research was financially supported by the Ministry of Education (MOE) and National Research Foundation of Korea (NRF) through the Human Resource Training Project for Regional Innovation (No. 2012H1B8A2026225) and National Research Foundation grant funded by the Korea Government (MSIP) (2012-005624).

3

100

References and Notes

B(OH)2

3 B(OH)2

4 Br B(OH)2

O

5 CF3

Br

B(OH)2

Cl

B(OH)2

6 O

7 Cl

porous silica hollow shells, were synthesized by direct silica coating and partial etching of the silica layers. In comparison to existing homogeneous and heterogeneous Pd catalysts, these three types of Pd-silica hybrid nanostructure possess outstanding thermal stability, excellent catalytic performance, and reusability without appreciable loss in catalytic activity, which comes from their distinctive hybrid structure. In addition, these represent good activity in the coupling of both bromobenzenes and chlorobenzenes with arylboronic acid, which include either electron-withdrawing or electron-donating substituents.

1. N. Miyaura, K. Yamada, and A. Suzuki, Tetrahedron Lett. 36, 343 (1979). 2. K. C. Colaou, P. G. Bulger, and D. Sarlah, Angew. Chem. Int. Ed. B(OH)2 O 44, 4442 (2005). 8 3 100 3. M. Miura, Angew. Chem. Int. Ed. 43, 2201 (2004). Br MeO B(OH)2 4. S. P. Stanforth, Tetrahedron 54, 263 (1998). Delivered by Publishing Technology to: Korea Advanced of Commun. Science38,&4759 Technology (KAIST) 5. A.Institute Suzuki, Chem. (2005). 9 3 87 6. N. Miyaura and07:05:42 A. Suzuki, Chem. Rev. 95, 2457 (1995). IP: 143.248.37.33 On: Mon, 03 Feb 2014 Br B(OH)2 7. A. Suzuki, J. Organomet. Chem. 576, 147 (1999). Copyright: American Scientific Publishers 8. W. Zhang, Q. Wang, F. Qin, H. Zhou, Z. Lu, and R. Chen, a Notes: Reaction conditions: Aryl halide, arylboronic acid, Pd@pSiO2 yolk-shell J. Nanosci. Nanotechnol. 11, 7794 (2011). nanocatalysts (0.003 mol% with respect to aryl halide concentration), Cs2 CO3 9. A. S. Kumar, K. K. R. Datta, T. S. Rao, K. V. Raghavan, b (2.0 equiv), DMF/H2 O (20:1) at 473 K; Isolated yield. M. Eswaramoorthy, and B. V. Reddy, J. Nanosci. Nanotechnol. 12, 2000 (2012). 10. A. B. Patil and B. M. Bhanage, J. Nanosci. Nanotechnol. 13, 5061 in the solution phase, was successfully used as poisons for (2013). Pd-catalyzed coupling reactions. The coupling reaction in 11. S. M. Reddy, K. K. R. Datta, C. Sreelakshmi, M. Eswaramoorthy, the presence of a poisoning reagent showed no obvious and B. V. Reddy, Nanosci. Nanotechnol. Lett. 4, 420 (2012). change in catalytic activity. Furthermore, no coupled prod12. X. Cheng, J. Li, Q. Wang, L. Gao, D. Zhang, and A. Zhang, uct was detected after the catalyst removal by microfiltraNanosci. Nanotechnol. Lett. 5, 384 (2013). 13. M. Choi, D.-H. Lee, K. Na, B.-W. Yu, and R. Ryoo, Angew. Chem. tion, demonstrating the absence of dissolved Pd species in Int. Ed. 48, 3673 (2009). the reaction solution. 14. M.-J. Jin and D.-H. Lee, Angew. Chem. Int. Ed. 49, 1119 (2010). 15. Y. Wan, H. Wang, Q. Zhao, M. Klingstedt, O. Terasaki, and D. Zhao, J. Am. Chem. Soc. 131, 4541 (2009). 4. CONCLUSION 16. G. M. Scheuermann, L. Rumi, P. Steurer, W. Bannwarth, and We have synthesized various hybrid Pd nanoparticles such R. M˝ulhaupt, J. Am. Chem. Soc. 131, 8262 (2009). as Pd/SiO2 nanobeads, Pd@Niphy yolk-shell nanoparti17. Y.-H. Chen, H.-H. Hung, and M. H. Huang, J. Am. Chem. Soc. 131, 9114 (2009). cles, Pd@pSiO2 yolk-shell nanoparticles. And they dis18. B. Baruwati, D. Guin, and S. V. Manorama, Org. Lett. 9, 5377 played highly efficient catalytic activity and excellent (2007). reusability in the Suzuki coupling reaction which is used 19. H. Sakurai, T. Tsukuda, and T. Hirao, J. Org. Chem. 67, 2721 extensively in the fabrication of both simple and complex (2002). biaryl compounds. Pd/SiO2 nanobeads containing tiny Pd 20. H. Qiu, S. M. Sarkar, D. H. Lee, and M. J. Jin, Green Chem. 10, 37 (2008). clusters in porous silica beads were obtained via a sol–gel 21. W. Mai and L. Gao, Synlett 2006, 2553 (2006). process and hydrogen reduction methods. Pd@Niphy yolk22. Y. Xiong, J. Chen, B. Wiley, Y. Xia, S. Aloni, and Y. Yin, J. Am. shell nanoparticles were obtained via Niphy formation for Chem. Soc. 127, 7332 (2005). pore generation of the silica shells, and exhibit high con23. V. P. Ananikov, N. V. Orlov, I. P. Beletskaya, V. N. Khrustalev, version rate with respect to its porosity. Pd@pSiO2 yolkM. Y. Antipin, and T. V. Timofeeva, J. Am. Chem. Soc. 129, 7252 shell nanoparticles, consist of tiny Pd core and highly (2007).

J. Nanosci. Nanotechnol. 14, 1872–1883, 2014

1881

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles

Kim et al.

60. H. Gu, Z. Yang, J. Gao, C. K. Chang, and B. Xu, J. Am. Chem. 24. R. Schlögl and S. B. A. Hamid, Angew. Chem. Int. Ed. 43, 1628 Soc. 127, 34 (2005). (2004). 61. H. R. Choi, H. Woo, S. Jang, J. Y. Cheon, C. Kim, J. Park, K. H. 25. A. Roucoux, J. Schulz, and H. Patin, Chem. Rev. 102, 3757 (2002). Park, and S. H. Joo, Chem. Cat. Chem. 4, 1587 (2012). 26. S. U. Son, Y. Jang, J. Park, H. B. Na, H. M. Park, H. J. Yun, J. Lee, 62. J. C. Park, A. Y. Kim, J. Y. Kim, S. Park, K. H. Park, and H. Song, and T. Hyeon, J. Am. Chem. Soc. 126, 5026 (2004). Chem. Commun. 48, 8484 (2012). 27. T. Mallat and A. Baiker, Chem. Rev. 104, 3037 (2004). 63. Y. Li, Y. J. Kim, A. Y. Kim, K. Lee, M. H. Jung, N. H. Hur, K. H. 28. S. Jansat, M. Gómez, K. Philippot, G. Muller, E. Guiu, C. Claver, Park, and W. S. Seo, Chem. Mater. 23, 5398 (2011). S. Castillón, and B. Chaudret, J. Am. Chem. Soc. 126, 1592 (2004). 64. Y. Choi, H. S. Bae, E. Seo, S. Jang, K. H. Park, and B.-S. Kim, 29. M. Moreno-Mañas and R. Pleixats, Acc. Chem. Res. 36, 638 (2003). J. Mater. Chem. 21, 15431 (2011). 30. C. C. Cassol, A. P. Umpierre, G. Machado, S. I. Wolke, and 65. A. Kim, B. Sharma, B.-S. Kim, and K. H. Park, J. Nanosci. J. Dupont, J. Am. Chem. Soc. 127, 3298 (2005). Nanotechnol. 11, 6162 (2011). 31. M. Lamblin, L. N. Hardy, J. C. Hierso, E. Fouquet, and F. X. 66. J. C. Park, H. J. Lee, H. S. Jung, M. Kim, H. J. Kim, K. H. Park, Felpin, Adv. Synth. Catal. 352, 33 (2010). and H. Song, Chem. Cat. Chem. 3, 755 (2011). 32. C. Xue, K. Palaniappan, G. Arumugam, S. A. Hackney, J. Liu, and 67. G. M. Kumar, J. Kawakita, and R. Jayavel, J. Nanosci. NanotechH. Liu, Catal. Lett. 116, 94 (2007). nol. 11, 3867 (2011). 33. N. T. S. Phan, M. Van Der Sluys, and C. W. Jones, Adv. Synth. 68. K. S. Yang, J. S. Yun, J. C. Kim, J. Min, T. J. Park, J. K. Ahn, and Catal. 348, 609 (2006). D. H. Kim, J. Nanosci. Nanotechnol. 12, 377 (2012). 34. K. Sawai, R. Tatumi, T. Nakahodo, and H. Fujihara, Angew. Chem. 69. M. Estevez, J. R. Rodriguez, S. Vargas, J. A. Guerra, W. Brostow, 120, 7023 (2008). and H. E. H. Lobland, J. Nanosci. Nanotechnol. 13, 4446 (2013). 35. K. Sawai, R. Tatumi, T. Nakahodo, and H. Fujihara, Angew. Chem. 70. M. Na, H. Park, M. Ahn, H. Lee, and I. Chung, J. Nanosci. Int. Ed. 47, 6917 (2008). Nanotechnol. 10, 6992 (2010). 36. P. J. Ellis, I. J. S. Fairlamb, S. F. J. Hackett, K. Wilson, and A. F. Lee, Angew. Chem. 122, 1864 (2010). 71. J. Choi, Y. Jun, S. Yeon, H. C. Kim, J. Shin, and J. Cheon, J. Am. 37. A. F. Lee, P. J. Ellis, I. J. S. Fairlamb, and K. Wilson, Dalton Trans. Chem. Soc. 128, 15982 (2006). 39, 10473 (2010). 72. P. V. Kamat, J. Phys. Chem. C 111, 2834 (2007). 38. N. Hoshiya, M. Shimoda, H. Yioshikawa, Y. Yamashita, S. Shuto, 73. S. J. Guo, S. J. Dong, and E. W. Wang, ACS Nano 4, 547 (2010). and M. Arisawa, J. Am. Chem. Soc. 132, 7270 (2010). 74. J. Yang, C. Tian, L. Wang, and H. Fu, J. Mater. Chem. 21, 3384 39. A. V. Gaikwad, A. Holyuigre, M. B. Thathagar, Dr. J. E. ten Elshof, (2011). 75. K. Vinokurov, J. E. Macdonald, and U. Banin, Chem. Mater. and G. Rothenberg, Chem. Eur. J. 13, 6908 (2007). 24, 1822 (2012). 40. S. MacQuarrie, J. H. Horton, J. Barnes, K. McEleney, H.-P. Loock, 76. Y. Deng, Y. Cai, Z. Sun, J. Liu, C. Liu, J. Wei, and D. Zhao, J. Am. and C. M. Crudden, Angew. Chem. 120, 3324 (2008). Chem. Soc. 132, 8466 (2010). 41. V. Calo, A. Nacci, A. Monopoli, and F. Montingelli, J. Org. Chem. Delivered by Publishing Technology to: Korea Advanced ofR.Science & Technology 77. X.Institute Liu, Y. Shen, Yang, S. Zou, X. Ji, L. Shi, and(KAIST) G. D. Stucky, 70, 6040 (2005). IP: 143.248.37.33 Mon, 03 Nano Feb Lett. 2014 07:05:42 12, 5733 (2012). 42. J. Liu, Y. Deng, H. Wang, H. Zhang, G. Yu, B. Wu, and On: A. Lei, Copyright: American Scientific Publishers 78. H. Sharghi, R. Khalifeh, and M. M. Doroodmand, Adv. Synth. Org. Lett. 10, 2661 (2008). Catal. 351, 207 (2009). 43. B. C. Ranu, R. Dey, and K. Chattopadhyay, Tetrahedron Lett. 79. X. Zhang, W. Jiang, Y. Zhou, S. Xuan, C. Peng, L. Zong, and 49, 3430 (2008). X. Gong, Nanotechnology 22, 375701 (2011). 44. A. Castonguay, D. Zargarian, and A. L. Beauchamp, 80. Y. Zhu, J. Shen, K. Zhou, C. Chen, X. Yang, and C. Li, J. Phys. Organometallics 27, 5723 (2008). Chem. C 115, 1614 (2010). 45. Z. Y. Jiang, C. H. Zhang, F. L. Gu, K. F. Yang, G. Q. Lai, L. W. 81. P. Worathanakul, J. Jiang, P. Biswas, and P. Kongkachuichay, Xu, and C. G. Xia, Synlett 2010, 1251 (2010). J. Nanosci. Nanotechnol. 8, 6253 (2008). 46. J. Y. Kim, J. C. Park, A. Kim, A. Y. Kim, H. J. Lee, H. Song, and 82. D. Steiner, T. Mokari, U. Banin, and O. Millo, Phys. Rev. Lett. K. H. Park, Eur. J. Inorg. Chem. 28, 4219 (2009). 95, 056805 (2005). 47. L. Yin and J. Liebscher, Chem. Rev. 107, 133 (2007). 83. M. Haruta, N. Yamada, T. Kobayashi, and S. Iijima, J. Catal. 48. G. Marck, A. Villiger, and R. Buchecker, Tetrahedron Lett. 35, 3277 115, 301 (1989). (1994). 84. M. S. Chen and D. W. Goodman, Science 306, 252 (2004). 49. D. J. Cole-Hamilton, Science 299, 1702 (2003). 85. B. M. Choudary, S. Madhi, N. S. Chowdari, M. L. Kantam, and 50. P. D. Stevens, G. Li, J. Fan, M. Yen, and Y. Gao, Chem. Commun. B. Sreedhar, J. Am. Chem. Soc. 124, 14127 (2002). 35, 4435 (2005). 86. L. Strimbu, J. Liu, and A. E. Kaifer, Langmuir 19, 483 (2003). 51. C. M. Crudden, M. Sateesh, and R. Lewis, J. Am. Chem. Soc. 127, 10045 (2005). 87. C. R. LeBlond, A. T. Andrews, Y. Sun, and J. R. Sowa, Org. Lett. 52. S. W. Kim, M. Kim, W. Y. Lee, and T. Hyeon, J. Am. Chem. Soc. 3, 1555 (2001). 124, 7642 (2002). 88. M. Xiao, C. Zhao, H. Chen, B. Yang, and J. Wang, Adv. Funct. 53. B. Yuan, Y. Pan, Y. Li, B. Yin, and H. Jiang, Angew. Chem. Int. Mater. 22, 4526 (2012). Ed. 49, 4054 (2010). 89. H. Song, Y. T. Yu, and P. Norby, J. Nanosci. Nanotechnol. 9, 5891 54. B. J. Gallon, R. W. Kojima, R. B. Kaner, and P. L. Diaconescu, (2009). Angew. Chem. Int. Ed. 46, 7251 (2007). 90. Y. S. Wang, S. J. Lin, Q. H. Guo, B. K. Chen, and D. Zheng, 55. N. Li, Z. Wang, K. Zhao, Z. Shi, S. Xu, and Z. Gu, J. Nanosci. J. Nanosci. Nanotechnol. 10, 2208 (2010). 91. M. X. Jing, C. Han, Z. Wang, and X. Q. Shen, J. Nanosci. Nanotechnol. 10, 6748 (2010). Nanotechnol. 13, 4949 (2013). 56. N. V. Long, C. M. Thi, Y. Yong, M. Nogami, and M. Ohtaki, 92. M. Ghahari, R. Aghababazadeh, T. Ebadzadeh, A. Mirhabibi, J. Nanosci. Nanotechnol. 13, 4799 (2013). R. Brydson, P. Fabbri, and F. Najafi, J. Nanosci. Nanotechnol. 57. S. C. Kim, W. G. Shim, and J. Y. Ryu, J. Nanosci. Nanotechnol. 11, 5311 (2011). 10, 3521 (2010). 93. S. Vaidya, P. Thaplyal, K. V. Ramanujachary, S. E. Lofland, and 58. B. Bridier and J. Perez-Ramirez, J. Am. Chem. Soc. 132, 4321 A. K. Ganguli, J. Nanosci. Nanotechnol. 11, 3405 (2011). (2010). 94. Z. Ma, H. Han, and J. Xue, J. Nanosci. Nanotechnol. 9, 3188 59. B. L. Liu, Y. P. Fu, and M. L. Wang, J. Nanosci. Nanotechnol. (2009). 9, 1491 (2009).

1882

J. Nanosci. Nanotechnol. 14, 1872–1883, 2014

Kim et al. 95. Q. He, J. Liu, and R. Hu, Nanosci. Nanotechnol. Lett. 5, 995 (2013). 96. C. W. Yang, K. Chanda, P. H. Lin, Y. N. Wang, C. W. Liao, and M. H. Huang, J. Am. Chem. Soc. 133, 19993 (2011). 97. M. Kim, E. Heo, A. Kim, J. C. Park, H. Song, and K. H. Park, Catal. Lett. 142, 588 (2012). 98. M. Kim, J. C. Park, A. Kim, K. H. Park, and H. Song, Langmuir 28, 6441 (2012). 99. J. C. Park, E. Heo, A. Kim, M. Kim, K. H. Park, and H. Song, J. Phys. Chem. C 115, 15772 (2011). 100. K. Osseo-Asare and F. J. Arriagada, Colloid Surf. 50, 321 (1990). 101. D. K. Yi, S. T. Selvan, S. S. Lee, G. C. Papaefthymiou, D. Kundaliya, and J. Y. Ying, J. Am. Chem. Soc. 127, 4990 (2005). 102. A. Mcdonald, B. Scott, and G. Villemure, Microporous Mesoporous Mater. 120, 263 (2009). 103. P. Burattin, M. Che, and C. Louis, J. Phys. Chem. B 102, 2722 (1998).

Suzuki Coupling Reaction Using Hybrid Pd Nanoparticles 104. Z. Guo, F. Du, G. Li, and Z. Cui, Chem. Commun. 25, 2911 (2008). 105. J. C. Park, H. J. Lee, J. U. Bang, K. H. Park, and H. Song, Chem. Commun. 47, 7345 (2009). 106. S. U. Son, Y. Jang, K. Y. Yoon, E. Kang, and T. Hyeon, Nano Lett. 4, 1147 (2004). 107. Y. Han, J. Jiang, S. S. Lee, and J. Y. Ying, Langmuir 24, 5842 (2008). 108. J. C. Park and H. Song, Nano Res. 4, 33 (2011). 109. J. Lee, J. C. Park, and H. Song, Adv. Mater. 20, 1523 (2008). 110. J. C. Park, J. U. Bang, C. H. Ko, and H. Song, J. Mater. Chem. 20, 1239 (2010). 111. J. C. Park, H. J. Lee, J. Y. Kim, K. H. Park, and H. Song, J. Phys. Chem. C 114, 6381 (2010). 112. P. M. Arnal, M. Comotti, and F. Schüth, Angew. Chem. Int. Ed. 45, 8224 (2006). 113. X. Huang, C. Guo, J. Zuo, N. Zheng, and G. D. Stucky, Small 5, 361 (2009).

Received: 24 July 2013. Revised/Accepted: 18 August 2013.

Delivered by Publishing Technology to: Korea Advanced Institute of Science & Technology (KAIST) IP: 143.248.37.33 On: Mon, 03 Feb 2014 07:05:42 Copyright: American Scientific Publishers

J. Nanosci. Nanotechnol. 14, 1872–1883, 2014

1883