Facile synthesis of palladium nanoparticles supported on silica: An

29 downloads 0 Views 849KB Size Report
Facile synthesis of palladium nanoparticles supported on silica: An efficient ..... [27] S.N. Jadhav, A.S. Kumbhar, C.V. Rode, R.S. Salunkhe, Green Chem.
Catalysis Communications 86 (2016) 32–35

Contents lists available at ScienceDirect

Catalysis Communications journal homepage: www.elsevier.com/locate/catcom

Short communication

Facile synthesis of palladium nanoparticles supported on silica: An efficient phosphine-free heterogeneous catalyst for Suzuki coupling in aqueous media Debojeet Sahu a, Ana R. Silva b, Pankaj Das a,⁎ a b

Department of Chemistry, Dibrugarh University, Dibrugarh, Assam, India Department of Chemistry, CICECO-Aveiro Institute of Materials, University of Aveiro, Aveiro, Portugal

a r t i c l e

i n f o

Article history: Received 26 April 2016 Received in revised form 14 July 2016 Accepted 2 August 2016 Available online 3 August 2016 Keywords: Palladium nanoparticles N-ligand Heterogeneous catalyst Suzuki-coupling Aqueous media Aryl chloride

a b s t r a c t A convenient one-step method for synthesizing highly dispersed palladium nanoparticles supported on silica, without taking assistance from any external reductant or stabilizer, has been developed. The supported nanoparticles were characterized by N2-adsorption desorption, XRD, HRTEM, SEM-EDX, XPS, ICP analyses and applied as catalyst for Suzuki-Miyaura reactions of aryl halides. The reactions with aryl bromides were performed in neat water at room temperature; while the reactions with aryl chlorides were conducted in aqueous-ethanol at 90 ° C. The catalyst could be reused at least three times without compromising with its activity, however from the fourth cycle a progressive decrease in yield was noticed. No aggregation of NPs was observed by the TEM analysis of the six-time used catalyst. © 2016 Elsevier B.V. All rights reserved.

1. Introduction Over the past few decades, remarkable developments have been accomplished in the palladium (Pd) catalyzed Suzuki-Miyaura cross-coupling reaction, which is one of the most popular methods for the construction of carbon–carbon bonds [1–4]. Usually, in most cases homogeneous ligand-based molecular complexes are used as catalysts. Although such homogeneous systems often display amazing activities particularly while dealing with substrates like aryl chlorides, however, their tedious syntheses and high costs pose a serious limitation to their industrial-scale applications [2,5]. Recent examples suggest that Pd nanoparticles (NPs), either in their colloidal form [6,7] or supported on solids [8–11], are attractive alternatives to these molecular catalysts. Compared to colloidal NPs, supported NPs are more advantageous owing to their ease in separations and better recycling efficiencies. As a result, during the past few years, a number of materials such as silica [12–14], MCM-41(Mobil Composition of Matter No. 41) [15], SBA-15 (Santa Barbara amorphous type silica No. 15) [16], carbon [17,18], Metal-Organic Framework (MOF) [19], etc. have been employed as solid supports for PdNPs to design efficient heterogeneous catalysts for Suzuki-Miyaura reaction. Typically, supported PdNPs are synthesized using a PdII precursor, an organically modified support, a reducing

⁎ Corresponding author. E-mail address: [email protected] (P. Das).

http://dx.doi.org/10.1016/j.catcom.2016.08.005 1566-7367/© 2016 Elsevier B.V. All rights reserved.

agent and a stabilizing or dispersing agent via chemical reduction techniques [20,21]. Recently, we have demonstrated that PdNPs could also be synthesized without the assistance of an external reducing and stabilizing agent using a commercially available phosphine ligand bound to silica [22]. However, the high cost of the ligand (approx. $430/g) along with its inherent toxicity and air sensitivity forms the bottleneck in the large-scale use of the afore-mentioned catalyst. Thus, it is of great importance to develop new cost-effective methods for synthesizing PdNPs under phosphine-free conditions. In line with this, herein, we report synthesis of a new Pdnanocatalyst using PdCl2 and a supported nitrogen-based ligand without using an external reductant or stabilizer. The effectiveness of the catalyst was tested for the Suzuki-Miyaura cross-coupling reactions of aryl halides in aqueous media. It is worth noting that significant efforts have been devoted nowadays to explore aqueous condition in Suzuki reactions, however most of the successful systems are still homogeneous in nature [23,24]. To the best of our knowledge, very few heterogeneous Pd catalysts are known to date that can promote the Suzuki reactions in aqueous media [16–19,25–27]. Moreover, majority of those systems showed promising activity mostly with aryl bromides and failed when aryl chlorides were used as substrates. It may be noted that from the cost and accessibility point of view, aryl chlorides are the most desirable substrates in cross-coupling reactions. Hence, the development of new heterogeneous Pd catalysts for efficient activations of aryl halides including less reactive aryl chlorides in aqueous media is one of the most attractive areas of research in Suzuki reaction.

D. Sahu et al. / Catalysis Communications 86 (2016) 32–35

2. Experimental 2.1. Synthesis of the supported Pd nanocatalyst 2.15 g of NMe2@SiO2 was added to 40 mL of acetonitrile solution of 1.5 mmol of PdCl2 under continuous stirring at 70 °C. After 6 h, the reaction mixture was filtered to get a black solid residue which was washed several times through soxhlet extraction using acetonitrile for 12 h in each time. No coloration of the washed extract confirmed that the solid catalyst was free from physically adsorbed PdCl2. The solid was then dried at 80 °C in an oven in air for 12 h. The material so obtained was designated as PdNP-NMe2@SiO2. 2.2. Suzuki-Miyaura cross-coupling reaction For Suzuki-Miyaura reaction, appropriate amount of the catalyst, PdNP-NMe2@SiO2, was added to a mixture of aryl halide (0.5 mmol), arylboronic acid (0.65 mmol), K2CO3 (1.5 mmol) in 6 mL solvent. The reaction was then stirred under desired temperature for the required time. The initial progress of the reaction was monitored by TLC using aluminum coated TLC plates (Merck) under UV light and the product formation was determined using GC–MS. After completion, the catalyst was collected by filtration and washed with isopropanol-water. The filtrate was diluted with water and extracted with ether and then dried over Na2SO4. After evaporation of the solvent under reduced pressure, the residue was chromatographed (silica gel, ethyl acetate–hexane, 1:9) to obtain the desired product. 3. Results and discussion 3.1 Synthesis and characterization of the catalyst

The palladium nanoparticles (PdNPs) were synthesized by a simple one-step reaction between PdCl2 and an amine-functionalized silica gel (NMe2@SiO2) (Scheme 1). No separate reducing or stabilizing agent was used. The formation of highly dispersed NPs was clearly visible from TEM picture (Fig. 1). The particles are usually spherical/elliptical in nature (ESI: Fig. S1a) and have the sizes in the range of 2–7 nm with majority of the particles fall between 4 and 5 nm (Fig. 1a, inset). The crystalline nature of the NPs was evident from the selected area electron diffraction (SAED) pattern which shows a fringe width of about 0.2 nm consistent with the face centre cubic (fcc) arrangement of the nanocrystals (ESI: Fig. S1b) [28,29]. The SEM micrograph (ESI: Fig. S2a) of the catalyst material shows silica particles are of different dimensions and the EDX spectra shows the presence of Pd (ESI: Fig. S2a). The Pd content from the ICP-AES analysis was found to be

33

0.39 mmol g− 1 of silica gel. The N2-adsorption-desorption isotherm (ESI: Fig. S3) of the catalyst shows a type IV isotherm with a specific surface area (ABET) of 328 m2/g. Compared to the amine-supported silica material (438 m2/g) this value is much lower indicating immobilization of Pd onto silica (ESI: Table S1). The powder XRD patterns (ESI: Fig. S4) of the Pd-based material shows a broad peak for silica at 2θ = 22.40, along with four other low intensity peaks at (111), (200), (220) and (311) planes corresponding to Pd(0) nanocrystals (JCPDS card No. 001–1201). The XPS survey spectrum of the Pd-nanocatalyst, PdNPNMe2@SiO2, confirms the presence of Pd, C, N, O and Si (ESI: Fig. S5a). The N 1s high resolution spectrum (Fig. S5b) can be deconvoluted into two peaks centered at 402.3 and 400.3 eV respectively. The higher energy peak is very close to the value (401.9 eV) usually obtained for an iminium ion [29], while the lower binding energy peak is typical of an amine group [30,31]. Although the chemical pathway for formation of this iminium ion is not very clear; there are literature evidences where tertiary amines in the presence of transition metals like Pd are converted to iminium ion [32,33], which possibly isomerized to enamine and/or hydrolyzed to aldehyde. In fact, the weak binding energy peaks in the XPS spectrum for C1S at 287.6 eV and O1S at 534.1 (ESI Fig. S6) might also suggest the presence of a carbonyl group. The binding energies of the Pd 3d5/2 and 3d3/2 electrons in the high resolution spectrum (Fig. 2) appear at 335.7 eV and 340.8 eV respectively and these values are consistent with the formation of Pd(0) NPs on silica matrix [34]. 3.1. Catalytic studies The Suzuki-Miyaura activity of PdNP-NMe2@SiO2 was first tested with aryl bromides employing p-bromoanisole and phenylboronic acid as model substrates using water as solvent and K2CO3 as base with 0.5 mol% Pd. We were delighted to see that the coupling reaction was completed smoothly in water at room temperature and, almost quantitative conversion was obtained (ESI: Table S2, entry 1). In a controlled experiment, when the same reaction was performed with homogeneous PdCl2 as catalyst without using the supported ligand, only 44% yield was obtained (Table S2, entry 2). Interestingly, under similar experimental condition, when a water-soluble Pd salt such as, Pd(NO3)2, was used as catalyst only slight improvement in product yield was observed (entry 11). Hence, it can be concluded that the cooperative influence between Pd and supported amine is responsible for the improved activity of PdNP-NMe2@SiO2 over homogeneous Pd species. Besides K2CO3, our system can also tolerate other bases like Na2CO3, Cs2CO3 and NaOH. Optimization of catalyst quantity revealed that the use of 0.1 mol% metal loading resulted in completion of the reaction within 4 h (Table S2, entry 10). Motivated by our initial optimization study, we extended the scope of the catalyst for a range of sterically

Pd

Scheme 1. Synthesis of silica supported Pd nanocatalyst, PdNP-NMe2@SiO2.

34

D. Sahu et al. / Catalysis Communications 86 (2016) 32–35

(b)

40

35

35

30

Distribution (%)

Distribution (%)

(a)

30 25 20 15 10 5 0 2

3

4

5

Size (nm)

6

25 20 15 10 5 0

7

2

3

4

5

Size (nm)

6

7

Fig. 1. TEM spectra of PdNP-NMe2@SiO2 along with particle size distribution histogram; (a) Before catalysis and (b) After catalysis.

and electronically different aryl bromides (Table 1). Good-to-excellent yields of cross-coupling products were obtained in most of the cases. Indeed, the efficient activities were also observed in case of hindered substrates like, 2-bromobenzaldehyde (entry 5) or 1-bromo-2,5dimethoxybenzene (entry 6). Interestingly, biaryls containing heterocyclic groups were also accessible by our catalytic system under slightly warm conditions (entries 8, 9 and 11). It may also be noted that changing the boronic acid from phenyl to tolyl (entry 13)/chloro (entry 12) had very little influence in our catalytic system. However the heterocyclic boronic acid viz. 2-thienylboronic acid gave relatively low yield (entry 14). With high yielding results with aryl bromides, the potentiality of our catalyst was tested for aryl chlorides. The initial reaction in water was conducted between p-chloronitrobenzene and phenylboronic acid by following the same experimental conditions used for aryl bromide (ESI: Table S3, entry 1). Unfortunately, the catalytic system was found to be almost inactive at room temperature. Increasing the temperature to 90 °C and catalyst loading to 1.0 mol%, no significant improvement in yield was observed (Table S3, entry 2). Interestingly, changing the solvent system had a dramatic effect on the catalytic system. For instance, when water was replaced with EtOH as solvent, yield of 4nitrobiphenyl was increased from 12 to 80% (Table S3, entries 2 and 5). Similarly when DMF was used as solvent almost quantitative product formation was achieved. DMF and H2O-EtOH (1:1) was found to be the best solvent systems in our study (Table S3, entries 6 and 8). Pd 3d

Pd 3d5/2 335.7

Pd 3d3/2 341.1

Although the activity in H2O-EtOH was slightly lower compared to DMF (88 and 96% respectively), considering economic and environmental benefits, we chose to use H2O-EtOH system for further studies (Table 2). It is noteworthy that both electron-poor (Table 2; entries 1, 2 and 5) and electron-rich (entries 6 and 7) aryl chlorides underwent smooth coupling with phenylboronic acid and produced the corresponding biaryls in good-to-excellent yields. This high yield was also maintained when phenylboronic was replaced by p-tolylboronic acid (entry 11) or p-chloroboronic acid (entry 10), while a significant drop in yield was seen when thienylboronic acid was used (entry 12). The ortho-substituted aryl (entry 3) or heteroaryl chlorides (entries 8 and 9) also provided reasonably good yields with our catalyst. 3.2. Reusability and heterogeneity test The potential for recyclability of the catalyst was tested using pbromoanisole and phenylboronic acid (ESI: Fig. S7). The catalyst was found to exhibit recyclability up to three cycles without compromising with its activity, however from the fourth cycle a gradual decrease in product yield was noticed, and after the sixth cycle the conversion dropped to 92%. The ICP-AES analysis of the 6th time used catalyst shows only negligible loss in Pd content (b2%). SEM image of the used catalyst shows a significant decrease in the silica particle size compared to that of fresh catalyst (ESI: Fig. S2b). No aggregation of Pd NPs was observed by the TEM analysis of the used catalyst (Fig. 1b), although a slight decrease in sizes of the Pd-NPs was observed for the used catalyst. A similar type of decrease in Pd particle size has also been reported earlier [35]. Hot filtration test (ESI: Fig. S8) together with solid phase poisoning test (ESI: Section 3.4) confirm that the catalyst is truly heterogeneous in nature. 4. Conclusion In conclusion we have disclosed a convenient protocol for synthesizing a silica-supported Pd nanocatalyst without involving phosphine, and also without using any external reducing or stabilizing agent. The Pd nanocatalyst could efficiently catalyze the Suzuki-Miyaura crosscoupling reactions of aryl bromides in water at room temperature and aryl chlorides in aqueous-ethanol under relatively mild condition.

350

348

346

344

342

340

338

336

334

332

330

binding energy (eV) Fig. 2. High resolution XPS (Pd 3d) spectrum of the silica supported catalyst PdNP-NMe2@ SiO2.

Acknowledgements The CSIR), New Delhi, is gratefully acknowledged for financial support (Grant no: 02/0081/12). The UGC New Delhi was also acknowledged for SAP-DRS programme and for providing fellowship to D.S.

D. Sahu et al. / Catalysis Communications 86 (2016) 32–35

35

Table 1 Suzuki-Miyaura cross coupling reactions of aryl/heteroaryl bromides with arylboronic acids with PdNP-NMe2@SiO2 as catalyst. R–Br þ R0 –BðOHÞ2

PdNP‐NMe2 @SiO2 ð0:1 mol%Þ



K2 CO3 ; H2 O; RT‐50 ÅC

R–R0

Entry

R-Br

R′-B(OH)2

Time (h)

Yield (%)a

1 2 3 4 5 6 7 8 9 10 11 12 13 14

4-OCH3-C6H4 C6H5 4-CHO-C6H4 4-CH3-C6H4 2-CHO-C6H4 2,5-OCH3-C6H3 4-COPh-C6H4 5-Bromopyrimidine 2-Bromo-3-methylthiophene Methyl 2-bromobenzoate 5-Bromo-2-furaldehyde 4-OCH3-C6H4 4-OCH3-C6H4 C6H5

C6H5 C6H5 C6H5 C6H5 C6H5 C6H5 C6H5 C6H5 C6H5 C6H5 C6H5 4-Cl-C6H4 4-CH3-C6H4 2-Thenylboronic acid

4 4 4 4 6 6 6 8 8 8 8 7 7 8

98 97 85 96 80 87 83 82b 80b 86b 76b 88 92 63b

Reaction conditions: 0.5 mmol aryl bromide, 0.65 mmol arylboronic acid, 1.5 mmol K2CO3, 6 mL H2O. a Isolated yield. b Reactions are conducted at 50 °C.

Table 2 Suzuki-Miyaura cross coupling reactions of aryl chlorides with arylboronic acids with PdNP-NMe2@SiO2 as catalyst. R–Cl þ R0 −BðOHÞ2

PdNP‐NMe2 @SiO2 ð1:0 mol%Þ



K2 CO3 ; H2 O‐EtOH; 90 ÅC

R−R0

Entry

R-Cl

R′-B(OH)2

Time (h)

Yield (%)a

1 2 3 4 5 6 7 8 9 10 11 12

4-NO2-C6H4 3-NO2-C6H4 2-NO2-C6H4 C6H5 4-CHO-C6H4 4-COCH3-C6H4 4-OCH3-C6H4 3-Chloropyridine 2-Chlorothiophene 4-OCH3-C6H4 4-OCH3-C6H4 C6H5

C6H5 C6H5 C6H5 C6H5 C6H5 C6H5 C6H5 C6H5 C6H5 4-Cl-C6H4 4-CH3-C6H4 2-Thenylboronic acid

6 7 7 6 6 6 6 8 8 6 6 8

88 80 52 94 85 88 95 70 65 88 82 56

Reaction conditions: 0.5 mmol aryl chloride, 0.65 mmol arylboronic acid, 1.5 mmol K2CO3, 6 mL H2O-EtOH (1:1). a Isolated yield.

under BSR-RFSMS scheme. Thanks are also due to SAIF, NEHU, Shillong for the TEM analysis. Appendix A. Supplementary data Supplementary data to this article can be found online at http://dx. doi.org/10.1016/j.catcom.2016.08.005. References [1] N. Miyaura, A. Suzuki, Chem. Rev. 95 (1995) 2457–2483. [2] A. Fihri, M. Bouhrara, B. Nekoueishahraki, J.-M. Basset, V. Polshettiwar, Chem. Soc. Rev. 40 (2011) 5181–5203. [3] S. Paul, M.M. Islam, S.M. Islam, RSC Adv. 5 (2015) 42193–42221. [4] P. Das, L. Wolfgang, Coord. Chem. Rev. 311 (2016) 1–23. [5] A. Molnar, Chem. Rev. 111 (2011) 2251–2320. [6] M. Pérez-Lorenzo, J. Phys. Chem. Lett. 3 (2012) 167–174. [7] C. Deraedt, D. Astruc, Acc. Chem. Res. 47 (2014) 494–503. [8] N.J.S. Costa, L.M. Rossi, Nanoscale 4 (2012) 5826–5834. [9] K. Karimi, M. Ghasemi, N.H. Naeini, Catal. Commun. 38 (2013) 10–15. [10] A. Guarnizo, I. Angurell, M.D. Rossell, J. Llorca, G. Muller, M. Seco, O. Rossell, RSC Adv. 5 (2015) 91340–91348. [11] S. Jadhav, A. Jagdale, S. Kamble, A. Kumbhar, R. Salunkhe, RSC Adv. 6 (2016) 3406–3420. [12] L. Wan, C. Cai, Catal. Commun. 24 (2012) 105–108. [13] N.J.S. Costa, P.K. Kiyohara, A.L. Monteiro, Y. Coppel, K. Philippot, L.M. Rossi, J. Catal. 276 (2010) 382–389. [14] D.-H. Lee, J.-Y. Jung, M.-J. Jin, Green Chem. 12 (2010) 2024–2029. [15] S. Dhara, K. Sarkar, D. Srimani, S.K. Saha, P. Chattopadhyay, A. Bhaumik, Dalton Trans. 39 (2010) 6395–6402.

[16] H. Yang, X. Han, Z. Ma, R. Wang, J. Liu, X. Ji, Green Chem. 12 (2010) 441–451. [17] J. Sun, Y. Fu, G. He, X. Sun, X. Wang, Appl. Catal. Environ. 165 (2015) 661–667. [18] L. Zhong, A. Chokkalingam, S.S. Cha, K.S. Lakhi, X. Su, G. Lawrence, A. Vinu, Catal. Today 243 (2015) 195–198. [19] L. Zhang, C. Feng, S. Gao, Z. Wang, C. Wang, Catal. Commun. 61 (2015) 21–25. [20] M. Planellas, R. Pleixats, A. Shafir, Adv. Synth. Catal. 354 (2012) 651–662. [21] J.E. Camp, J.J. Dunsford, E.P. Cannons, W.J. Restorick, A. Gadzhieva, M.W. Fay, R.J. Smith, ACS Sustain. Chem. Eng. 2 (2014) 500–505. [22] D. Sahu, P. Das, RSC Adv. 5 (2015) 3512–3520. [23] P.R. Boruah, A.A. Ali, B. Saikia, D. Sarma, Green Chem. 17 (2015) 1442–1445. [24] I. Hoffmann, B. Blumenröder, S.O. Thumann, S. Dommer, J. Schatz, Green Chem. 17 (2015) 3844–3857. [25] S. Rostamnia, E. Doustkhah, B. Zeynizadeh, Microporous Mesoporous Mater. 222 (2016) 87–93. [26] J. Hoseini, V. Heidari, H. Nasrabadi, J. Mol. Catal. A Chem. 396 (2015) 90–95. [27] S.N. Jadhav, A.S. Kumbhar, C.V. Rode, R.S. Salunkhe, Green Chem. 18 (2016) 1898–1911. [28] D.V. Jawale, E. Gravel, C. Boudet, N. Shah, V. Geertsen, H. Li, I.N.N. Namboothiri, E. Doris, Catal. Sci. Technol. 5 (2015) 2388–2392. [29] N. Ohno, H.-J. Wang, H. Yan, N. Toshima, Polym. J. 33 (2001) 165–171. [30] P. Das, A.R. Silva, A.P. Carvalho, J. Pires, C. Freire, Colloids Surf. A Physicochem. Eng. Asp. 329 (2008) 190–197. [31] A.S. Roy, J. Mondal, B. Banerjee, P. Mondal, A. Bhaumik, S.M. Islam, Appl. Catal. Gen. 469 (2014) 320–327. [32] R. McCrindle, G. Ferguson, G.J. Arsenault, A.J. McAlees, J. Chem. Soc. Chem. Commun. (1983) 571–572. [33] J. Sheng, Y. Guo, J. Wu, Tetrahedron 69 (2013) 6495–6499. [34] S. Singha, M. Sahoo, K.M. Parida, Dalton Trans. 40 (2011) 7130–7132. [35] R. Bernini, S. Cacchi, G. Fabrizi, G. Forte, F. Petrucci, A. Prastaro, S. Niembro, A. Shafir, A. Vallribera, Green Chem. 12 (2010) 150–158.