Next Article in Journal / Special Issue
Guanidine Hydrochloride/ZnI2 as Heterogeneous Catalyst for Conversion of CO2 and Epoxides to Cyclic Carbonates under Mild Conditions
Previous Article in Journal
Preparation for Pt-Loaded Zeolite Catalysts Using w/o Microemulsion and Their Hydrocracking Behaviors on Fischer-Tropsch Product
Previous Article in Special Issue
Polymer Supported Triphenylphosphine-Palladium Acetate Complex PS-TPP-Pd(OAc)2 as a Heterogeneous and Reusable Catalyst for Indirect Reductive Amination of Aldehydes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Application of “Boomerang” Linear Polystyrene-Stabilized Pd Nanoparticles to a Series of C-C Coupling Reactions in Water

1
Department of Applied Chemistry, Faculty of Engineering, Osaka Institute of Technology, 5-16-1 Ohmiya, Asahi, Osaka 535-8585, Japan
2
Institute for Molecular Science (IMS), Higashiyama 5-1, Myodaiji, Okazaki 444-8787, Japan
3
Osaka Municipal Technical Research Institute, 1-6-50 Morinomiya, Joto, Osaka, 536-8553, Japan
4
Nanomaterials and Microdevices Research Center, Osaka Institute of Technology, 5-16-1 Ohmiya, Asahi, Osaka 535-8585, Japan
*
Author to whom correspondence should be addressed.
Catalysts 2015, 5(1), 106-118; https://doi.org/10.3390/catal5010106
Submission received: 19 December 2014 / Accepted: 3 February 2015 / Published: 9 February 2015
(This article belongs to the Special Issue Metal Catalysts Recycling and Heterogeneous/Homogeneous Catalysis)

Abstract

:
The application of a catch-and-release system for soluble Pd species between water (reaction medium) and polystyrene (polymer support) was examined in the Suzuki coupling reaction with 2-bromothiophene and the Heck reaction with styrene or bromobenzene. Although a slight increase in particle size was observed by TEM after re-stabilization of the Pd species on linear polystyrene, no agglomeration was observed.

Graphical Abstract

1. Introduction

Recently, homogeneous-heterogeneous switching catalysts have become of great interest to chemists, because they have the advantages of both homogeneous (high activity) and heterogeneous (recovery and recyclability) catalysts [1,2,3,4]. For example, Ikegami and Aoshima have developed homogeneous-heterogeneous switching catalysts by using thermosensitive polymers, such as poly(N-isopropylacrylamide) and poly[2-(2-ethoxy)ethoxyethyl vinyl ethers]. We have also reported a switching catalyst using a pH-sensitive polyion complex. A polymer supported “boomerang” catalyst that operates through the reversible release of a homogeneous catalyst into the solution phase from the polymer support is another example of a homogeneous-heterogeneous switching catalyst [5,6,7,8,9,10]. For instance, Reiser et al. have developed a method for the reversible immobilization of pyrene-tagged palladium N-heterocyclic carbene (NHC) complexes on highly magnetic, graphene-coated cobalt nanoparticles through π-π stacking interactions.
On the other hand, metal nanoparticles possess high catalytic activity in water [11,12,13,14,15,16,17,18,19] and have been used not only as semi-heterogeneous catalysts, but also as semi-heterogeneous supports [20,21,22,23]. Unfortunately, heterogenization of metal nanoparticles decreases catalytic activity. Additionally, the recovery and reuse of metal nanoparticles have been difficult to achieve due to a significant loss and/or morphology changes of the metal nanoparticles during reaction and workup [13,14,15]. Since leaching of soluble metal species from the support is a major cause of catalyst deactivation, efforts have been reported to develop an effective support that prevents leaching of metal species [16,17,18,19]. In contrast, leaching of Pd species has been observed directly and shows high catalytic activity in some systems [24,25,26]. Recently, palladium nanoparticles generated in situ for the Suzuki coupling reaction and aerobic alcohol oxidation in water were recovered using linear polystyrene, even in the presence of organic compounds [27]. These reports prompted efforts to develop recyclable “boomerang” nanoparticles that are able to overcome the problems mentioned above. Previously, we developed a catch-and-release system for soluble Pd species in water using linear polystyrene as an efficient reservoir [28]. This report describes the application of this system to a series of coupling reactions.

2. Results and Discussion

In the previous report, we confirmed that a recyclable homogeneous-heterogeneous switching system was constructed using polystyrene-stabilized PdO nanoparticles (PS-PdONPs) and aryl halide and that tetrabutylammonium bromide (TBAB) would act as a stabilizer for leached palladium species [28]. To evaluate the applicability of the homogeneous-heterogeneous catalytic system, the Suzuki coupling reaction of 2-bromothiophene with arylboronic acid was investigated (Table 1). When the coupling reaction of 2-bromothiophene with 4-methylphenylboronic acid was performed in the absence of TBAB, only 14% of the product was obtained (Entry 1). In contrast, the same reaction took place smoothly in water at 80 °C for 1 h in the presence of TBAB to afford 2-(4-methylphenyl)thiophene in a 78% yield (Entry 2). After washing with water and diethyl ether and subsequent drying, the recovered catalyst was used for the next run under the same conditions. The yield of the reaction did not decrease markedly, even after five consecutive runs (Entries 3–6). When the reaction solution was analyzed by ICP-AES to determine the extent of palladium leaching after the reaction, palladium was not detectable (<0.1 ppm) after any of the runs. However, 0.4% of palladium leached after heating the mixture of PS-PdONPs and 2-bromothiophene under reaction conditions. A slight increase in the size of palladium was observed by TEM after the recycling experiments (Figure 1). The 1s peak of N was not observed by XPS analysis, indicating the absence of tetrabutylammonium species in the recovered catalyst (Figure 2). These results suggest that leached palladium species are participating in the catalytic process and are redeposited onto linear polystyrene after the reaction is complete. Indeed, similar processes have been observed in organic solvent [29,30]. After the reaction, leached palladium species were re-stabilized onto linear polystyrene, probably due to the chelate effect [31]. The reaction in the presence of SDS gave a low yield (Entry 7) and no corresponding product was obtained in the case of 5-bromo-2-thiophenecarboxylic acid, suggesting that TBAB did not simply act as a phase-transfer catalyst (Entry 8). Both electron-rich and electron-deficient arylboronic acids were reactive, affording the desired coupling products in good yields (Entries 9 and 10). The Suzuki coupling reaction with 1-naphtylboronic acid and 2-thienylboronic acid gave the corresponding coupling products in 87 and 99% yields, respectively (Entries 11 and 12).
Table 1. Suzuki coupling reaction of thienyl halides with arylboronic acids a. Catalysts 05 00106 i001
Table 1. Suzuki coupling reaction of thienyl halides with arylboronic acids a. Catalysts 05 00106 i001
EntryArAdditiveYield (%) b
14-methylphenyl-14
24-methylphenylTBAB78
3 c4-methylphenylTBAB74
4 d4-methylphenylTBAB72
5 e4-methylphenylTBAB74
6 f4-methylphenylTBAB74
74-methylphenylSDS24
8 g4-methylphenyl-<1
94-methoxyphenylTBAB81
104-acetylphenylTBAB56
111-naphthylTBAB87
122-thienylTBAB99
a All reactions were performed with 2-bromothiophene (0.5 mmol), arylboronic acid (0.75 mmol), PS-PdONPs (1.5 mol% of Pd), 1.5 mol/L KOH aqueous solution (1 mL). b NMR yields. c 2nd use. d 3rd use. e 4th use. f 5th use. g 5-Bromo-2-thiophenecarboxylic acid was used as a substrate.
Figure 1. (a) TEM micrograph of PS-PdONPs (scale bar = 20 nm); (b) size distribution of PS-PdONPs; (c) TEM micrograph of recovered catalyst after fifth run (scale bar = 20 nm); (d) size distribution of recovered catalyst after fifth run.
Figure 1. (a) TEM micrograph of PS-PdONPs (scale bar = 20 nm); (b) size distribution of PS-PdONPs; (c) TEM micrograph of recovered catalyst after fifth run (scale bar = 20 nm); (d) size distribution of recovered catalyst after fifth run.
Catalysts 05 00106 g001
Figure 2. XPS spectrum of the recovered catalyst.
Figure 2. XPS spectrum of the recovered catalyst.
Catalysts 05 00106 g002
As we reported earlier [32], PS-PdONPs showed low catalytic activity for the Heck reaction with styrene or bromobenzene. Therefore, we also examined the applicability of our boomerang system in the Heck reaction. When the Heck reaction of iodobenzene with styrene was performed in 1.5 mol·L−1 KOH aqueous solution at 90 °C in the absence or presence of TBAB, only 14% or 17% of the product was obtained, respectively (Table 2, Entries 1 and 2). However, on using polystyrene-stabilized Pd nanoparticles (PS-PdNPs) as a catalyst, the coupling reaction proceeded efficiently (Entry 4). Based on the mechanism of the Heck reaction using Pd(II) species as the catalyst, the first step is the reduction of Pd(II) to Pd(0) with alkene. The difference in activity is thought to be caused by the rate of the reduction step (insertion of alkene). Indeed, the reduction of palladium was observed by XPS after treatment of PS-PdONPs with acrylic acid in 1.5 mol·L−1 KOH aqueous solution at 90 °C for 5 h. In contrast, no formation of Pd(0) was observed after treatment of PS-PdONPs with styrene (Figure 3). When the recyclability of the catalyst was examined under the same conditions, catalytic deactivation was observed in the second and third runs. However, similar yields of products were obtained after the third run as in the Suzuki coupling reaction of aryl chloride with arylboronic acid [28]. When the reaction solution was checked by ICP-AES, no palladium species was observed in any runs. In addition, the size of palladium nanoparticles in the recovered catalyst was slightly larger than that of PS-PdNPs (fresh catalyst, Figure 4). The Heck reaction using both electron-rich and electron-deficient aryl iodides also occurred to produce the desired product in good yields. On the other hand, coupling products were obtained in moderate yields from the reaction with 4-methoxystyrene (Entries 18–20).
Table 2. Heck reaction of aryl iodide with styrene a. Catalysts 05 00106 i002
Table 2. Heck reaction of aryl iodide with styrene a. Catalysts 05 00106 i002
EntryR1R2Yield (%) b
1 c,dHH14
2 cHH17
3 dHH35
4HH95
5 eHH88
6 fHH56
7 gHH52
8 hHH53
9 iHH20
10 jHH11
114-CH3-H95
124-CH3O-H95
134-CF3-H83
142-CH3-H98
15HCl99
164-CH3-Cl99
174-CH3O-Cl96
18HOMe74
194-CH3-OMe77
204-CH3O-OMe41
a All reactions were performed with aryl iodide (0.5 mmol), styrene (0.75 mmol), catalyst (1.5 mol% of Pd), 1.5 mol/L KOH aqueous solution (1 mL). b NMR yield. c PS-PdONPs was used as a catalyst. d Reaction was performed in the absence of TBAB for 20 h. e 2nd run. f 3rd run. g 4th run. h 5th run. I SDS was used as an additive. j 2-Iodophenol was used as a substrate.
Figure 3. XPS spectrum of the Pd 3d region of the catalyst: (a) PS-PdONPs; (b) after treatment with acrylic acid; (c) after treatment with styrene.
Figure 3. XPS spectrum of the Pd 3d region of the catalyst: (a) PS-PdONPs; (b) after treatment with acrylic acid; (c) after treatment with styrene.
Catalysts 05 00106 g003
Figure 4. (a) TEM micrograph of PS-PdNPs (scale bar = 20 nm); (b) size distribution of PS-PdNPs; (c) TEM micrograph of recovered catalyst after fifth run (scale bar = 20 nm); (d) size distribution of recovered catalyst after fifth run.
Figure 4. (a) TEM micrograph of PS-PdNPs (scale bar = 20 nm); (b) size distribution of PS-PdNPs; (c) TEM micrograph of recovered catalyst after fifth run (scale bar = 20 nm); (d) size distribution of recovered catalyst after fifth run.
Catalysts 05 00106 g004
When the Heck reaction of bromobenzene with acrylic acid was performed in 1.5 mol·L−1 KOH aqueous solution at 90 °C for 5 h, only 5% of the product was obtained in the absence of additive (Table 3, Entry 1). In contrast, in the presence of TBAB, as expected, an increased product yield was observed (Entry 2). Although a similar yield was obtained with the recovered catalyst, a broader size distribution was observed by TEM in the recovered catalyst (Figure 5). Both electron-rich and electron-deficient aryl bromides were reactive, affording the desired coupling products in high yields (Entries 3–6). 2-Bromotoluene and 1-bromonaphthalene also underwent the Heck reaction with acrylic acid under similar conditions to afford 2-methylcinnamic acid and 3-(1-naphthyl)propanoic acid in 78% and 96% yields, respectively (Entries 7 and 8).
Table 3. Heck reaction of aryl bromide with acrylic acid a. Catalysts 05 00106 i003
Table 3. Heck reaction of aryl bromide with acrylic acid a. Catalysts 05 00106 i003
EntryRAdditiveYield (%) b
1H-5
2HTBAB84
34-CH3-TBAB97
4 c4-CH3-TBAB94
54-CH3O-TBAB96
64-CF3-TBAB88
74-NO2-TBAB71
82-CH3-TBAB78
92-naphthylTBAB96
a All reactions were performed with aryl bromide (0.5 mmol), acrylic acid (0.75 mmol), PS-PdONPs (1.5 mol% of Pd), 1.5 mol/L KOH aqueous solution (1 mL). b NMR yield. c 2nd run.
Figure 5. (a) TEM micrograph of the recovered catalyst (scale bar = 10 nm); (b) size distribution of the recovered catalyst.
Figure 5. (a) TEM micrograph of the recovered catalyst (scale bar = 10 nm); (b) size distribution of the recovered catalyst.
Catalysts 05 00106 g005

3. Experimental Section

3.1. Preparation of PS-PdONPs

To a screw-capped vial with a stirring bar were added 9.0 mg of polystyrene (85 μmol of styrene unit), Pd(OAc)2 5.5 mg (25 μmol) and 1.5 M aqueous K2CO3 solution (3 mL). After stirring at 90 °C for 1 h, the reaction mixture was filtered with hot water. Subsequently, the polystyrene-stabilized PdO nanoparticles were washed with water (5 × 1.0 mL) and acetone (5 × 1.0 mL).

3.2. Preparation of PS-PdNPs

To a screw-capped vial with a stirring bar were added 1.8 mg of polystyrene (17 μmol of styrene unit), Pd(OAc)2 1.0 mg (5 μmol), phenylboronic acid 9.1 mg (0.75 mmol) and water (1 mL). After stirring at 90 °C for 1 h, the aqueous solution was decanted. Subsequently, the polystyrene-stabilized Pd nanoparticles were washed with water (5 × 1.0 mL) and acetone (5 × 1.0 mL).

3.3. Determination of Loading of the Palladium

PS-PdONPs (2.9 mg) was placed in a screw-capped vial and then 13 M nitric acid (5 mL) was added. The mixture was heated at 80 °C to dissolve completely. After cooling to room temperature, the solution was adjusted to 50 mL by water, and then, the amount of Pd metal was measured by ICP-AES analysis (15.3 ppm). The amount of Pd in PS-PdNPs was 15.5 ppm.
After the catalytic reaction, the aqueous phase was adjusted to 10 g by nitric acid, and then, the amount of Pd metal was measured by ICP-AES analysis.

3.4. Typical Procedures for the Suzuki Coupling Reaction

To a screw-capped vial with a stirring bar were added 2-bromothiophene (81.5 mg, 0.5 mmol), 4-methylphenylboronic acid (110 mg, 0.75 mmol), PS-PdONPs (2.9 mg, 1.5 mol% of Pd), TBAB (161 mg, 0.5 mmol) and 1.5 mol·L−1 aqueous KOH solution (1 mL). After stirring at 80 °C for 1 h, the reaction mixture was cooled to room temperature by immediately immersing the vial in water (~20 °C) for about 10 min. After separating the catalyst and the aqueous phase by centrifugation, the aqueous phase was decanted. Recovered catalyst was washed with H2O (5 × 3.0 mL) and diethyl ether (5 × 3.0 mL), which were then added to the aqueous phase. The aqueous phase was extracted eight times with diethyl ether. The combined organic extracts were dried over MgSO4 and concentrated under reduced pressure. The product was analyzed by 1H NMR. The recovered catalyst was dried in vacuo and reused. Furthermore, the amount of Pd metal in the aqueous phase determined by ICP-AES analysis was <0.1 ppm.
2-(p-Tolyl)thiophene: 1H NMR (CDCl3) δ 7.48 (d, J = 8.2 Hz, 2 H), 7.23 (d, J = 3.6 Hz, 1 H), 7.19 (d, J = 5.0 Hz, 1 H), 7.15 (d, J = 8.2 Hz, 2 H), 7.06 (dd, J = 5.0 Hz, 3.6 Hz, 1 H), 2.32 (s, 3 H): 13C NMR (CDCl3) δ 144.6, 137.0, 131.7, 129.5, 127.9, 125.8, 124.2, 122.6, 21.2. CAS registry number: 16939-04-1.
2-(4-Methoxyphenyl)thiophene: 1H NMR (CDCl3) δ 7.55 (d, J = 8.8 Hz, 2 H), 7.20 (m, 2 H), 7.05 (dd, J = 3.5 Hz, J = 5.0 Hz, 1 H), 6.92 (d, J = 8.8 Hz, 2 H), 3.83 (s, 3 H): 13C NMR (CDCl3) δ 159.1, 144.2, 127.9, 127.2, 127.1, 123.8, 122.0, 114.2, 55.3. CAS registry number: 42545-43-7.
2-(4-Acetylphenyl)thiophene: 1H NMR (CDCl3) δ 8.02–7.90 (m, 2 H), 7.73–7.60 (m, 2 H), 7.42 (d, J = 3.6 Hz, 1 H), 7.35 (d, J = 5.1 Hz, 1 H), 7.10 (dd, J = 5.1 Hz, J = 3.6 Hz, 1 H), 2.60 (s, 3 H): 13C NMR (CDCl3) δ 197.2, 143.0, 138.7, 135.5, 129.1, 128.3, 126.5, 125.5, 124.6, 26.7. CAS registry number: 35294-37-2.
2-(o-Tolyl)thiophene: 1H NMR (CDCl3) δ 7.42 (d, J = 7.2 Hz, 1 H), 7.35 (d, J = 5.0 Hz, 1 H), 7.20–7.30 (m, 3 H), 7.10 (dd, J = 5.0 Hz, J = 3.5 Hz, 1 H), 7.07 (d, J = 3.5 Hz, 1 H), 2.44 (s, 3 H): 13C NMR (CDCl3) δ 143.0, 136.1, 134.1, 130.7, 130.4, 127.8, 127.0, 126.5, 125.9, 125.0, 21.5. CAS registry number: 99846-56-7.
2-(1-Naphthyl)thiophene: 1H NMR (CDCl3) δ 8.20 (m, 1 H), 7.88 (m, 1 H), 7.85 (d, J = 8.4 Hz, 1 H), 7.56 (d, J = 7.2 Hz, 1 H), 7.55–7.48 (m, 3 H), 7.40 (d, J = 5.2 Hz, 1 H), 7.23 (d, J = 3.3 Hz, 1 H), 7.18 (dd, J = 5.2 Hz, J = 3.3 Hz, 1 H): 13C NMR (CDCl3) δ 141.7, 133.8, 132.4, 131.8, 128.4, 128.3, 128.2, 127.3, 127.2, 126.4, 126.0, 125.7, 125.6, 125.2. CAS registry number: 4632-51-3.
2,2'-Bithiophene: 1H NMR (CDCl3) δ 7.22–7.15 (m, 4 H), 7.05–7.00 (m, 2 H): 13C NMR (CDCl3) δ 137.4, 127.5, 124.3, 123.5. CAS registry number: 492-97-7.

3.5. Typical Procedures for the Heck Reaction

To a screw-capped vial with a stirring bar were added iodobenzene (102 mg, 0.5 mmol), styrene (78.1 mg, 0.75 mmol), PS-PdNPs (3.0 mg, 1.5 mol% of Pd), TBAB (161 mg, 0.5 mmol) and 1.5 mol·L−1 aqueous KOH solution (1 mL). After stirring at 90 °C for 15 h, the reaction mixture was cooled to room temperature by immediately immersing the vial in water (~20 °C) for about 10 min. After separating the catalyst and the aqueous phase by centrifugation, the aqueous phase was decanted. Recovered catalyst was washed with H2O (5 × 3.0 mL) and diethyl ether (5 × 3.0 mL), which were then added to the aqueous phase. The aqueous phase was extracted eight times with diethyl ether. The combined organic extracts were dried over MgSO4 and concentrated under reduced pressure. The product was analyzed by 1H NMR. The recovered catalyst was dried in vacuo and reused. Furthermore, the amount of Pd metal in the aqueous phase determined by ICP-AES analysis was <0.1 ppm.
Cinnamic acid: 1H NMR (CDCl3) δ 7.81 (d, J = 16.0 Hz, 1 H), 7.62–7.50 (m, 2 H), 7.42 (m, 3 H), 6.47 (d, J = 16.0 Hz, 1 H): 13C NMR (CDCl3) δ 172.7, 147.0, 134.0, 130.7, 128.9, 128.3, 117.3. CAS registry number: 140-10-3.
4-Methylcinnamic acid: 1H NMR (CDCl3) δ 7.78 (d, J = 15.9 Hz, 1 H), 7.46 (d, J = 8.1 Hz, 2 H), 7.20 (d, J = 8.1 Hz, 2 H), 6.40 (d, J = 15.9 Hz, 1 H), 2.38 (s, 3 H): 13C NMR (CDCl3) δ 172.6, 147.0, 141.0, 131.5, 129.8, 128.4, 116.2, 21.4. CAS registry number: 940-61-4.
4-Methoxycinnamic acid: 1H NMR (CDCl3) δ 7.75 (d, J = 15.9 Hz, 1 H), 7.52 (d, J = 8.7 Hz, 2 H), 6.92 (d, J = 8.7 Hz, 2 H), 6.32 (d, J = 15.9 Hz, 1 H), 3.85 (s, 3 H): 13C NMR (CDCl3) δ 172.2, 161.8, 146.5, 130.5, 126.8, 114.4, 114.2, 55.4. CAS registry number: 943-89-5.
4-Trifluoromethylcinnamic acid: 1H NMR (DMSO-d6) δ 7.90 (d, J = 8.1 Hz, 2 H), 7.74 (d, J = 8.1 Hz, 2 H), 7.65 (d, J = 15.9 Hz, 1 H), 6.67 (d, J = 15.9 Hz, 1 H): 13C NMR (DMSO-d6) δ 167.8, 142.2, 138.2, 130.7 (q, J = 25.4 Hz), 129.7, 126.6 (q, J = 3.1 Hz), 126.0, 125.8, 122.1. CAS registry number: 16642-92-5.
4-Nitrocinnamic acid: 1H NMR (DMSO-d6) δ 8.22 (d, J = 8.1 Hz, 2 H), 7.97 (d, J = 8.1 Hz, 2 H), 7.68 (d, J = 15.9 Hz, 1 H), 6.74 (d, J = 15.9 Hz, 1 H): 13C NMR (CDCl3) δ 167.7, 148.4, 141.8, 141.3, 129.7, 124.4, 123.8. CAS registry number: 619-89-6.
2-Methylcinnamic acid: 1H NMR (CDCl3) δ 8.00 (d, J = 15.6 Hz, 1 H), 7.60 (d, J = 7.8 Hz, 1 H), 7.30–7.20 (m, 3 H), 6.36 (d, J = 15.6 Hz, 1 H), 2.40 (s, 3 H): 13C NMR (CDCl3) δ 170.4, 144.2, 138.0, 133.8, 131.8, 131.2, 126.8, 118.8, 19.7. CAS registry number: 2373-76-4.
Stilbene: 1H NMR (CDCl3) δ 7.60 (dd, J = 8.4 Hz, J = 1.5 Hz, 4 H), 7.45 (tt, J = 7.5 Hz, J = 1.5 Hz, 4 H), 7.34 (tt, J = 7.2 Hz, J = 1.5 Hz, 2 H), 7.20 (s, 2 H): 13C NMR (CDCl3) δ 137.4, 128.8, 128.7, 127.7, 126.6. CAS registry number: 103-30-0.
4-Methylstilbene: 1H NMR (CDCl3) δ 7.52 (d, J = 7.2 Hz, 2 H), 7.43 (d, J = 8.1 Hz, 2 H), 7.35 (t, J = 7.5 Hz, 2 H), 7.26 (t, J = 7.2 Hz, 1 H), 7.18 (d, J = 8.1 Hz, 2 H), 7.10 (s, 2 H), 2.38 (s, 3 H): 13C NMR (CDCl3) δ 137.7, 137.5, 134.8, 129.4, 128.7, 127.7, 127.4, 126.5, 126.4, 21.4. CAS registry number: 1860-17-9.
4-Methoxystilbene: 1H NMR (CDCl3) δ 7.53–7.46 (m, 4 H), 7.37 (t, J = 7.5 Hz, 2 H), 7.25 (t, J = 7.2 Hz, 1 H), 7.10 (d, J = 16.2 Hz, 1 H), 6.94 (d, J = 8.7 Hz, 2 H), 3.85 (s, 3 H): 13C NMR (CDCl3) δ 159.4, 137.7, 130.2, 128.5, 127.8, 127.2, 126.4, 126.1, 114.2, 55.3. CAS registry number: 1694-19-5.
4-Trifluoromethylstilbene: 1H NMR (CDCl3) δ 7.62–7.56 (m, 4 H), 7.54 (d, J = 7.5 Hz, 2 H), 7.4–7.36 (m, 2 H), 7.32–7.28 (m, 1 H), 7.19 (d, J = 16.5 Hz, 1 H), 7.12 (d, J = 16.5 Hz, 1 H): 13C NMR (CDCl3) δ 140.8, 136.6, 131.2, 129.2 (q, J = 32.3 Hz), 128.8, 128.3, 127.1, 126.8, 126.5, 125.6 (q, J = 3.8 Hz), 124.2. CAS registry number: 1149-56-0.
2-Methylstilbene: 1H NMR (CDCl3) δ 7.66 (d, J = 7.2 Hz, 1 H), 7.58 (d, J = 7.5 Hz, 2 H), 7.45–7.35 (m, 3 H), 7.30–7.25 (m, 4 H), 7.05 (d, J = 16.2 Hz, 1 H), 2.48 (s, 3 H): 13C NMR (CDCl3) δ 137.7, 136.5, 135.8, 130.5, 130.0, 128.7, 127.5, 127.5, 126.6, 126.3, 125.3, 20.1. CAS registry number: 22257-16-5.
4-Chlorostilbene: 1H NMR (CDCl3) δ 7.52 (d, J = 7.2 Hz, 2 H), 7.43 (d, J = 8.4 Hz, 2 H), 7.40–7.20 (m, 5 H), 7.07 (s, 2 H): 13C NMR (CDCl3) δ 137.0, 135.9, 133.2, 129.3, 128.8, 128.7, 127.9, 127.7, 127.4, 126.6. CAS registry number: 1657-50-7.
4-Chloro-4'-methylstilbene: 1H NMR (CDCl3) δ 7.43 (d, J = 8.1 Hz, 2 H), 7.41 (d, J = 8.4 Hz, 2 H), 7.31 (d, J = 8.4 Hz, 2 H), 7.17 (d, J = 8.1 Hz, 2 H), 7.07 (d, J = 15.9 Hz, 1 H), 7.00 (d, J = 15.9 Hz, 1 H), 2.36 (s, 3 H): 13C NMR (CDCl3) δ 137.8, 136.0, 134.1, 132.9, 129.4, 129.2, 128.8, 127.5, 126.4, 126.3, 21.3. CAS registry number: 3041-83-6.
4-Chloro-4'-methoxystilbene: 1H NMR (CDCl3) δ 7.49–7.35 (m, 4 H), 7.33–7.20 (m, 2 H), 7.02 (d, J = 16.2 Hz, 1 H), 6.95–6.88 (m, 3 H), 3.83 (s, 3 H): 13C NMR (CDCl3) δ 159.6, 136.2, 132.5, 129.7, 128.9, 128.8, 127.8, 127.4, 125.3, 114.2, 55.4. CAS registry number: 18878-89-2.
4-Methoxy-4'-methylstilbene: 1H NMR (CDCl3) δ 7.45 (d, J = 8.7 Hz, 2 H), 7.40 (d, J = 8.1 Hz, 2 H), 7.16 (d, J = 7.8 Hz, 2 H), 7.04 (d, J = 16.2 Hz, 1 H), 6.96 (d, J = 16.2 Hz, 1 H), 6.90 (d, J = 8.7 Hz, 2 H), 3.83 (s, 3 H), 2.36 (s, 3 H): 13C NMR (CDCl3) δ 159.3, 137.2, 135.0, 130.5, 129.5, 127.7, 127.4, 126.7, 126.3, 114.3, 55.5, 21.4. CAS registry number: 37163-68-1.
4,4'-Dimethoxystilbene: 1H NMR (CDCl3) δ 7.44–7.41 (m, 4 H), 6.94–6.86 (m, 6 H), 3.83 (s, 6 H): 13C NMR (CDCl3) δ 159.3, 130.7, 127.7, 126.4, 114.4, 55.6. CAS registry number: 15638-14-9.

4. Conclusions

In summary, a “boomerang” catalytic system, which allows one to combine the benefits of homogeneous (high catalytic activity) and heterogeneous (easy separation and recycling) catalysts, was constructed by using PS-PdONPs and TBAB. The results with water-soluble substrates or another stabilizer suggested that TBAB did not simply act as a phase-transfer catalyst. A slight increase in the size of palladium was observed by TEM after re-stabilization of the Pd species on linear polystyrene. This system is applicable to the Suzuki coupling reaction of 2-bromothiophene with arylboronic acid and the Heck reaction with aryl styrene or bromide.

Acknowledgements

The authors are grateful to the Nanomaterials and Microdevices Research Center (NMRC) of OIT for financial and instrumental support.

Author Contributions

Atsushi Ohtaka and Toshiyuki Okagaki contributed to the experimental design and the article writing and revising. Go Hamasaka, Yasuhiro Uozumi, and Tsutomu Shinagawa contributed to the catalyst characterization. Osamu Shimomura and Ryoki Nomura contributed to a helpful discussion.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gruttadauria, M.; Giacalone, F.; Noto, R. “Release and Catch” Catalytic Systems. Green Chem. 2013, 15, 2608–2618. [Google Scholar] [CrossRef]
  2. Kanaoka, S.; Yagi, N.; Fukuyama, Y.; Aoshima, S.; Tsunoyama, H.; Tsukuda, T.; Sakurai, H. Thermosensitive Gold Nanoclusters Stabilized by Well-Defined Vinyl Ether Star Polymers: Reusable and Durable Catalysts for Aerobic Alcohol Oxidation. J. Am. Chem. Soc. 2007, 129, 12060–12061. [Google Scholar] [CrossRef] [PubMed]
  3. Hamamoto, H.; Suzuki, Y.; Yamada, Y.M.A.; Tabata, H.; Takahashi, H.; Ikegami, S. A Recyclable Catalytic System Based on a Temperature-Responsive Catalyst. Angew. Chem. Int. Ed. 2005, 44, 4536–4538. [Google Scholar] [CrossRef]
  4. Ohtaka, A.; Tamaki, Y.; Igawa, Y.; Egami, K.; Shimomura, O.; Nomura, R. Polyion Complex Stabilized Palladium Nanoparticles for Suzuki and Heck Reaction in Water. Tetrahedron 2010, 66, 5642–5646. [Google Scholar] [CrossRef]
  5. Wittmann, S.; Schätz, A.; Grass, R.N.; Stark, W.J.; Reiser, O. A Recyclable Nanoparticle-Supported Palladium Catalyst for the Hydroxycarbonylation of Aryl Halides in Water. Angew. Chem. Int. Ed. 2010, 49, 1867–1870. [Google Scholar] [CrossRef]
  6. Clavier, H.; Nolan, S.P.; Mauduit, M. Ionic Liquid Anchored “Boomerang” Catalysts Bearing Saturated and Unsaturated NHCs: Recyclability in Biphasic Media for Cross-Metathesis. Organometallics 2008, 27, 2287–2292. [Google Scholar] [CrossRef]
  7. Rix, D.; Caïjo, F.; Laurent, I.; Gulajski, L.; Grela, K.; Mauduit, M. Highly Recoverable Pyridinium-Tagged Hoveyda-Grubbs Pre-catalyst for Olefinmetathesis. Design of the Boomerang Ligand Toward the Optimal Compromise Between Activity and Reusability. Chem. Commun. 2007, 3771–3773. [Google Scholar] [CrossRef]
  8. Varray, S.; Lazaro, R.; Martinez, J.; Lamaty, F. New Soluble-Polymer Bound Ruthenium Carbene Catalysts: Synthesis, Characterization, and Application to Ring-Closing Metathesis. Organometallics 2003, 22, 2426–2435. [Google Scholar] [CrossRef]
  9. Jafarpour, L.; Nolan, S.P. Simply Assembled and Recyclable Polymer-Supported Olefin Metathesis Catalysts. Org. Lett. 2000, 2, 4075–4078. [Google Scholar] [CrossRef] [PubMed]
  10. Ahmed, M.; Arnauld, T.; Barrett, A.G.M.; Braddock, D.C.; Procopiou, P.A. Second Generation Recyclable “Boomerang” Polymer Supported Catalysts for Olefien Metathesis: Application of Arduengo Carbene Complexes. Synlett 2000, 1007–1009. [Google Scholar]
  11. Astruc, D. Nanoparticles and Catalysis; Wiley-VCH: Weinheim, Germany, 2008. [Google Scholar]
  12. Yan, N.; Xiao, C.; Kou, Y. Transition Metal Nanoparticle Catalysis in Green Solvents. Coord. Chem. Rev. 2010, 254, 1179–1218. [Google Scholar] [CrossRef]
  13. Dell’Anna, M.M.; Mali, M.; Mastrorilli, P.; Rizzuti, A.; Ponzoni, C.; Leonelli, C. Suzuki-Miyaura Coupling under Air in Water Promoted by Polymer Supported Palladium Nanoparticles. J. Mol. Catal. A 2013, 366, 186–194. [Google Scholar] [CrossRef]
  14. Beletskaya, I.P.; Kashin, A.N.; Khotina, I.A.; Khokhlov, A.R. Efficient and Recyclabe Catalyst of Palladium Nanoparticles Stabilized by Polymer Micelles Soluble in Water for Suzuki-Miyaura Reaction, Ostwald Ripening Process with Palladium Nanoparticles. Synlett 2008, 1547–1552. [Google Scholar] [CrossRef]
  15. Moldal, J.; Modak, A.; Bhaumik, A. One-pot Efficient Heck Coupling in Water Catalyzed by Palladium Nanoparticles Tethered into Mesoporous Organic Polymer. J. Mol. Catal. A 2011, 350, 40–48. [Google Scholar] [CrossRef]
  16. Boffi, A.; Cacchi, S.; Ceci, P.; Cirilli, R.; Fabrizi, G.; Prastaro, A.; Niembro, S.; Shafir, A.; Vallribera, A. The Heck Reaction of Allylic Alcohols Catalyzed by Palladium Nanoparticles in Water: Chemoenzymatic Synthesis of (R)-(−)-Rhododendrol. ChemCatChem 2011, 3, 347–353. [Google Scholar] [CrossRef]
  17. Yu, Y.; Hu, T.; Chen, X.; Xu, K.; Zhang, J.; Huang, J. Pd Nanoparticles on a Porous Ionic Copolymer: A Highly Active and Recyclable Catalyst for Suzuki-Miyaura Reaction under Air in Water. Chem. Commun. 2011, 47, 3592–3594. [Google Scholar] [CrossRef]
  18. Zhi, J.; Song, D.; Li, Z.; Lei, X.; Hu, A. Palladium Nanoparticles in Carbon Thin Film-lined SBA-15 Nanoreactors: Efficient Heterogeneous Catalysts for Suzuki-Miyaura Cross Coupling Reaction in Aqueous Media. Chem. Commun. 2011, 47, 10707–10709. [Google Scholar] [CrossRef]
  19. Ogasawara, S.; Kato, S. Palladium Nanoparticles Captured in Microporous Polymers: A Tailr-Made Catalyst for Heterogeneous Carbon Cross-Coupling Reactions. J. Am. Chem. Soc. 2010, 132, 4608–4613. [Google Scholar] [CrossRef] [PubMed]
  20. Astruc, D.; Lu, F.; Aranzaes, J.R. Nanoparticles as Recyclable Catalysts: The Frontier between Homogeneous and Heterogeneous Catalysis. Angew. Chem. Int. Ed. 2005, 44, 7852–7872. [Google Scholar] [CrossRef]
  21. Gallon, B.J.; Kojima, R.W.; Kaner, R.B.; Diaconescu, P.L. Palladium Nanoparticles Supported on Polyaniline Nanofibers as a Semi-Heterogeneous Catalyst in Water. Angew. Chem. Int. Ed. 2007, 46, 7251–7254. [Google Scholar] [CrossRef]
  22. Hu, J.; Wang, Y.; Han, M.; Zhou, Y.; Jiang, X.; Sun, P. A Facile Preparation of Palladium Nanoparticles Supported on Magnetite/s-graphene and Their Catalytic Application in Suzuki-Miyaura Reaction. Catal. Sci. Technol. 2012, 2, 2332–2340. [Google Scholar] [CrossRef]
  23. Schätz, A.; Reiser, O.; Stark, W.J. Nanoparticles as Semi-Heterogeneous Catalyst Supports. Chem. Eur. J. 2010, 16, 8950–8967. [Google Scholar] [CrossRef] [PubMed]
  24. Fang, P.-P.; Jutand, A.; Tian, Z.-Q.; Amatore, C. Au-Pd Core-Shell Nanoparticles Catalyze Suzuki-Miyaura Reactions in Water through Pd Leaching. Angew. Chem. Int. Ed. 2011, 50, 12184–12188. [Google Scholar] [CrossRef]
  25. MacQuarrie, S.; Horton, J.H.; Barnes, J.; McEleney, K.; Loock, H.-P.; Crudden, C.M. Visual Observation of Redistribution and Dissolution of Palladium during the Suzuki-Miyaura Reaction. Angew. Chem. Int. Ed. 2008, 47, 3279–3282. [Google Scholar] [CrossRef]
  26. Gniewek, A.; Trzeciak, A.M.; Ziółkowski, J.J.; Kępiński, L.; Wrzyszcz, J.; Tylus, W. Pd-PVP Colloid as Catalyst for Heck and Carbonylation Reactions: TEM and XPS Studies. J. Catal. 2005, 229, 332–343. [Google Scholar] [CrossRef]
  27. Ohtaka, A.; Kuroki, R.; Teratani, T.; Shinagawa, T.; Hamasaka, G.; Uozumi, Y.; Shimomura, O.; Nomura, R. Recovery of in situ-generated Pd Nanoparticles with Linear Polystyrene. Green Sus. Chem. 2011, 1, 19–25. [Google Scholar] [CrossRef]
  28. Ohtaka, A.; Sakaguchi, E.; Yamaguchi, T.; Hamasaka, G.; Uozumi, Y.; Shimomura, O.; Nomura, R. A Recyclable “Boomerang” Linear Polystyrene-Stabilized Pd Nanoparticles for the Suzuki Coupling Reaction of Aryl Chlorides in Water. ChemCatChem 2013, 5, 2167–2169. [Google Scholar] [CrossRef]
  29. Pryjomska-Ray, I.; Gniewek, A.; Trzeciak, A.M.; Ziółkowski, J.J.; Tylus, W. Homogeneous/Heterogeneous Palladium Based Catalytic System for Heck Rection. The Reversible Transfer of Palladium between Solution and Support. Top. Catal. 2006, 40, 173–184. [Google Scholar] [CrossRef]
  30. Zhao, F.; Murakami, K.; Shirai, M.; Arai, M. Recyclable Homogeneous/Heterogeneous Catalytic Systems for Heck Reaction through Reversible Transfer of Palladium Species between Solvent and Support. J. Catal. 2000, 194, 479–483. [Google Scholar] [CrossRef]
  31. Sun, W.; Zhou, S.; You, B.; Wu, L. Facile Fabrication and High Photoelectric Properties of Hierarchically Ordered Porous TiO2. Chem. Mater. 2012, 24, 3800–3810. [Google Scholar] [CrossRef]
  32. Ohtaka, A.; Yamaguchi, T.; Teratani, T.; Shimomura, O.; Nomura, R. Linear Polystyrene-Stabilized PdO Nanoparticle-Catalyzed Mizoroki-Heck Reactions in Water. Molecules 2011, 16, 9067–9076. [Google Scholar] [CrossRef] [PubMed]

Share and Cite

MDPI and ACS Style

Ohtaka, A.; Okagaki, T.; Hamasaka, G.; Uozumi, Y.; Shinagawa, T.; Shimomura, O.; Nomura, R. Application of “Boomerang” Linear Polystyrene-Stabilized Pd Nanoparticles to a Series of C-C Coupling Reactions in Water. Catalysts 2015, 5, 106-118. https://doi.org/10.3390/catal5010106

AMA Style

Ohtaka A, Okagaki T, Hamasaka G, Uozumi Y, Shinagawa T, Shimomura O, Nomura R. Application of “Boomerang” Linear Polystyrene-Stabilized Pd Nanoparticles to a Series of C-C Coupling Reactions in Water. Catalysts. 2015; 5(1):106-118. https://doi.org/10.3390/catal5010106

Chicago/Turabian Style

Ohtaka, Atsushi, Toshiyuki Okagaki, Go Hamasaka, Yasuhiro Uozumi, Tsutomu Shinagawa, Osamu Shimomura, and Ryôki Nomura. 2015. "Application of “Boomerang” Linear Polystyrene-Stabilized Pd Nanoparticles to a Series of C-C Coupling Reactions in Water" Catalysts 5, no. 1: 106-118. https://doi.org/10.3390/catal5010106

Article Metrics

Back to TopTop