Accepted Manuscript Title: Cellulose-supported N-Heterocyclic Carbene-palladium Catalyst: Synthesis and its applications in the Suzuki Cross-coupling Reaction Author: Xiaoxia Wang Peibo Hu Fengjun Xue Yuping Wei PII: DOI: Reference:

S0144-8617(14)00799-1 http://dx.doi.org/doi:10.1016/j.carbpol.2014.08.030 CARP 9167

To appear in: Received date: Revised date: Accepted date:

16-3-2014 23-7-2014 6-8-2014

Please cite this article as: Wang, X., Hu, P., Xue, F., and Wei, Y.,Cellulosesupported N-Heterocyclic Carbene-palladium Catalyst: Synthesis and its applications in the Suzuki Cross-coupling Reaction, Carbohydrate Polymers (2014), http://dx.doi.org/10.1016/j.carbpol.2014.08.030 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

1

Cellulose-supported N-Heterocyclic Carbene-palladium Catalyst:

2

Synthesis and its applications in the Suzuki Cross-coupling

3

Reaction

5

Xiaoxia Wang1, Peibo Hu1, Fengjun Xue, Yuping Weia,b,*

6

a

7

China

8

b

9

300072, P.R. China

ip t

4

us

cr

Department of Chemistry, School of Science, Tianjin University, Tianjin 300072, P.R.

an

Collaborative Innovation Center of Chemical Science and Engineering (Tianjin), Tianjin

10

1

11

*Corresponding Author. Tel and Fax: +86 22 27403475; E-mail: [email protected]

12

14

Abstract

Ac ce pt e

13

d

M

Both authors contributed equally to this work.

A cellulose-supported

N-methylimidazole-palladium catalyst

(Cell-NHC-Pd)

was

15

synthesized and used for Suzuki cross-coupling reactions between aryl halides and

16

phenylboronic acids to create the corresponding coupling products in good to excellent

17

yields. Moreover, the catalyst is easily recovered using only a few cycles of simple filtration.

18

Keywords: Cellulose, N-methylimidazole, Palladium, Suzuki cross-coupling reaction

19 20

1. Introduction

21

Palladium-catalyzed Suzuki cross-coupling reactions of aryl halides with arylboronic

22

acids are among the most efficient methods to selectively construct biaryl compounds in

1  Page 1 of 26

organic synthesis (Littke & Fu, 2002; Martin & Buchwald, 2008; Yin & Liebscher, 2007;

24

Old et al., 1998; Zhang et al., 1999; Lemo et al., 2005; Li et al., 2000). To date, Suzuki

25

reactions have been extensively used for the synthesis of natural products, herbicides,

26

pharmaceuticals, and other products (Corbet & Mignani, 2006; Hassan et al., 2002; Suzuki,

27

1999; Miyaura & Suzuki, 1995; Meijere & Diederich, 2004). Commonly, phosphine ligands

28

are employed to facilitate the corresponding transformations. However, a critical setback for

29

phosphine ligands is the fact that they are toxic and air-sensitive (Birkholz et al., 2009; Fu,

30

2008). Recently, N-heterocyclic carbenes (NHCs) have attracted many chemists’ attention as

31

transition metal ligands (Wang et al., 2012; Gao et al., 2012; Fleckenstein et al., 2007) after

32

they were first reported by Herrmann and co-workers (Herrmann et al., 1995; Herrmann et

33

al., 1998). Compared with phosphine ligands, NHCs have the same σ-donor and low

34

π-acceptor ability while they can form stronger bonds with the majority of metals. More

35

importantly, they are air and moisture stable. Nowadays, a wide range of NHC ligands are

36

commercially available which exhibit high activities in various important organic

37

transformations when combined with metallic pre-catalysts (Drge & Glorius, 2010; Dorta et

38

al., 2005; Marion & Nolan, 2008). Despite the significant progress that has been made in the

39

design of these NHC ligands, the desired products are always inevitably contaminated by

40

catalyst residues during these homogeneous reactions (Blaser, 2002; Cole-Hamilton, 2003),

41

which greatly limits the use of NHCs-related organic transformations in industrial

42

applications.

Ac ce pt e

d

M

an

us

cr

ip t

23

43

The immobilization of metal catalysts on solid supports has been considered an efficient

44

approach to address the aforementioned issues. Numerous solid supports have been utilized

2  Page 2 of 26

as anchors for NHC and their metal complex such as polystyrene (Kang et al., 2005; Lo &

46

Luo, 2012), polysilylethers (Lázaro et al., 2013) and mesoporous silica (Pozo et al., 2011).

47

However, the progress in the quality of catalysts on solid support has been relatively slow

48

compared to soluble homogeneous complexes because of the deactivation that results from

49

particle aggregation, metal leaching and matrix degradation (Molnár & Papp, 2014). Another

50

major problem is the environmental concern arising from the use of poisonous and often

51

costly materials applied in catalyst synthesis. In this regards, cellulose is an attractive and

52

promising candidate that could be used as an efficient, renewable, and biodegradable support

53

for catalysts (Jamwal et al., 2011). In addition, cellulose is stable in most organic solvents,

54

inexpensive, and has a high surface area. Typically, functional groups present in cellulose

55

backbone participate in bonding with metals. Compared with widely used PS support,

56

cellulose can easily be functionalized due to its abundant hydroxyl groups and can create

57

either soluble or insoluble supports through chemical modification. Cellulose and its

58

derivatives have also been widely used for coatings, optical films, pharmaceuticals, and

59

consumables (Klemm et al., 2005). The use of cellulose as an efficient support for Pd

60

nanoparticles in cross-coupling reactions has also been widely reported (Jamwal et al., 2011;

61

Reddy et al., 2006; Cirtiu et al., 2011; Xu et al., 2008). As demonstrated in a recently

62

accepted review publication (Molnár & Papp, 2014), the most studied catalysts are still

63

cellulose deposited Pd particles. In contrast, the use of immobilized Pd complexes has not

64

been fully exploited. According to the literature, only three examples for such support

65

materials have been synthesized: cellulose-tagged diphenylphosphinate (Du & Li, 2012; Du

66

& Li, 2011; Keshipour et al., 2013), triphenylphosphine (Xu et al., 2014) and

Ac ce pt e

d

M

an

us

cr

ip t

45

3  Page 3 of 26

67

ethylenediamine (Zheng et al., 2009). More versatile functionalization techniques should be

68

developed to achieve stable, immobilized cellulose metal complexes, allowing for the

69

improvement of catalytic activities. With all these considerations in mind, in this paper, we present the anchoring of N-

71

methylimidazole onto a cellulose backbone as a precursor, followed by coordination with

72

Pd(OAc)2 to prepare a novel cellulose-NHC-palladium catalyst (Cell-NHC-Pd). In addition,

73

we explore its catalytic activity and recycling performance using the Suzuki reaction as an

74

example.

us

cr

ip t

70

2. Experimental

77

2.1. Materials

M

76

an

75

Microcrystalline cellulose AVICEL PH-101 (Microcrystalline cellulose, degree of

79

polymerization = 180) was purchased from Shandong Liaocheng A Hua Pharmaceutical Co.,

80

Ltd. (Shandong Province, P.R. China). Triethylamine (Et3N), anhydrous lithium chloride

81

(LiCl), potassium carbonate (K2CO3), dimethylformamide (DMF), all of which are analytical

82

grade (AR), were obtained from Tianjin Guangfu Fine Chemical Research Institute (Tianjin,

83

P.R. China). N-methylimidazole was purchased from Sinopharm Chemical Reagent Beijing

84

Co., Ltd. Acetone was obtained from Tianjin Jiangtian Chemical Technology Co., Ltd. The

85

cellulose, anhydrous lithium chloride, and potassium carbonate were vacuum dried for 48 h

86

at 50 °C, 40 °C and 30 °C respectively, before use. Dimethylacetamide (DMAc),

87

dimethylsulfoxide (DMSO), dichloromethane (CH2Cl2) and DMF were desiccated over

88

CaH2 overnight, then filtered through silica gel; they were then vacuumed distilled and

Ac ce pt e

d

78

4  Page 4 of 26

89

stored in brown bottles with 4A molecular sieves under N2. Other reagents were only

90

redistilled.

91

2.2. Instruments NMR spectra were recorded on a Bruker Avance 400 MHz spectrometer in CDCl3 using

93

TMS as an internal standard. The particle sizes of the samples were observed using a Tecnai

94

G2 F20 transmission electron microscope (TEM). The elemental content of palladium was

95

determined by inductively coupled plasma atomic emission spectroscopy (ICP-AES) using a

96

Thermo Jarrell-Ash corporation, ICP-9000 (N+M) instrument. X-ray photoelectron

97

spectrometry (XPS) was conducted using a PHI1600 ESCA System (Perkin-Elmer, United

98

States). The content of sulfur and nitrogen were determined by a Vario Micro cube elemental

99

analyzer. Thermogravimetric analysis (TGA) was measured with a STA 409 PC thermal

100

analyzer (NETZSCH, Germany). Fourier transform infrared spectroscopy (FTIR) analysis

101

was performed on a BIO-RAD FTS3000 IR Spectrum Scanner.

102

2.3. Preparation of Cell-NHC-Cl

Ac ce pt e

d

M

an

us

cr

ip t

92

103

Cellulose p-toluenesulfonate (Cell-OTs) was prepared according to methods reported in

104

the literature (Rahn et al., 1996). Cell-OTs (1.91 g, 10.0 mmol) was then added to DMSO

105

(10 mL) in a 25 mL flask, with degree of substitution (DS) of 0.14, and the mixture was

106

stirred for 24 h at 60 °C followed by overnight stirring at room temperature to dissolve the

107

Cell-OTs completely. N-methylimidazole (0.86 g, 10.0 mmol) was then added to the reaction

108

mixture under nitrogen and the viscous black mixture was stirred for 12 h at 100 °C. After

109

cooling to room temperature, the mixture was added dropwise into acetone in a beaker and

110

stirred vigorously for approximately 6 h. The reaction mixture was filtered and the resulting

5  Page 5 of 26

[Cell-NHC][OTs]- was obtained. [Cell-NHC][OTs]- was then dissolved in a solution of NaCl

112

(1.15 g, 20.0 mmol) in acetone (10 mL) to replace OTs- with Cl-. The precipitate

113

(Cell-NHC-Cl) was collected via filtration. The obtained brown powder was vacuum dried

114

for 24 h at 40 °C. The DS of Cell-NHC was 0.08 determined on the base of nitrogen content.

115

2.4. Preparation of Cell-NHC-Pd complex

ip t

111

Under nitrogen, Cell-NHC-Cl (0.50 g) was added to a solution of Pd(OAc)2 (0.10 g, 0.45

117

mmol) in DMF (10 mL) and the mixture was stirred for 24 h at 60 °C. After cooling to room

118

temperature, the reaction mixture was filtered. The obtained solid black product was washed

119

carefully with distilled water (3×25 mL), absolute ether (2×25 mL) and absolute ethyl

120

alcohol (2×25 mL) successively then dried under reduced pressure at room temperature to

121

give the dark palladium complex (Cell-NHC-Pd).

122

2.5 General procedure for the Suzuki reaction of aryl halides with arylboronic acids

d

M

an

us

cr

116

The aryl halide (0.5 mmol), arylboronic acid (0.75 mmol), Cell-NHC-Pd (0.75 mol%),

124

K2CO3 (1 mmol) and DMF : H2O (1:1) (2.5 mL) were successively added into a 25 mL

125

round bottom flask. The mixture was stirred vigorously at 80 °C for the varying specific

126

length of time based on each different substrate. After reaction completion, the solid catalyst

127

was filtered off. The filtrate was extracted with ethyl acetate (3×5 mL), washed with water

128

and brine three times, and then dried over anhydrous MgSO4. The MgSO4 was then filtered

129

off. After removal of the solvent, the crude product was purified by preparative TLC (eluent:

130

petroleum ether/ethyl acetate, 20/1) to give the desired product and calculate the yield.

131

2.6 Separation of the catalyst and recycling tests

132

Ac ce pt e

123

After completion of the Suzuki–Miyaura reaction, the liquid phase was removed by

6  Page 6 of 26

133

centrifugation. The solid catalyst was simply recovered by filtration and washing. Without

134

drying, the recovered catalyst was directly used in next run with new portions of reactants.

136

3. Results and Discussion

137

3.1. Synthesis and characterization of Cell-NHC-Pd

ip t

135

Cellulose-supported NHC (Cell-NHC) was synthesized via the nucleophilic substitution of

139

Cell-OTs with N-methylimidazole as shown in Scheme 1. The tosylate acted as an effective

140

leaving group, allowing the SN2 reaction with N-methylimidazole to create Cell-NHC-OTs.

141

In order to improve the catalytic activity, OTs- was exchanged by Cl- (Kim et al., 2005). The

142

resulting Cell-NHC-Cl was then coordinated with palladium acetate in DMF to give the final

143

catalyst Cell-NHC-Pd complex. The Pd loading was determined to be 0.24 mmol/g by

144

inductively coupled plasma-atomic emission spectroscopy (ICP-AES), which was higher

145

than that of other polymer-supported NHC palladium complex reported in previous literature

146

(Lo & Luo, 2012; Kim et al., 2006).

Ac ce pt e

d

M

an

us

cr

138

147 148

Scheme 1. Reaction scheme for the formation of catalyst (Cell-NHC-Pd).

149

The FTIR spectra of cellulose, cellulose derivatives Cell-OTs, Cell-NHC-Cl, and

150

Cell-NHC-Pd are shown in Fig. 1. The bands at 1359 cm-1 and 1179 cm-1 corresponds to the

7  Page 7 of 26

O=S asymmetric and symmetric stretching vibration of Cell-OTs, respectively, which are

152

greatly weakened after substitution with N-methylimidazole as shown in Fig. 1b-c.

153

Correspondingly, the N-methylimidazole graft is visible from appearance of a new band at

154

1580 cm−1, which corresponds to the stretching vibration of N–sp2 C bond (Zhao et al., 1995;

155

Kurt et al., 2000). In addition, the absorption in the 1220 cm−1 region is interpreted as the

156

vibration of N–sp3 C bond (Kurt et al., 2000; Rodil et al., 2000). As shown in Fig. 1c-d, the

157

band at 632 cm−1 region corresponds to the vibration of the imidazole ring, and the intensity

158

of its absorption is reduced after reaction with palladium acetate, suggesting the formation of

159

the NHC-Pd complex (He & Cai, 2011).

160 161

Ac ce pt e

d

M

an

us

cr

ip t

151

Fig. 1. FT-IR spectra of Cell (a), Cell-OTs (b), Cell-NHC-Cl (c), and Cell-NHC-Pd (d).

162

As the Suzuki reaction usually requires heating, thermogravimetric analysis (TGA) was

163

performed to evaluate the thermal stability of the cellulose-supported catalyst in air and high

164

temperature. As shown in Fig. 2, TGA of the catalyst system showed high thermal stability

165

with decomposition at around 225 °C in air. The result demonstrated that cellulose was

8  Page 8 of 26

considerably stable in an oxygen-containing atmosphere as a catalyst support in

167

experimental conditions.

an

us

cr

ip t

166

169

M

168

Fig. 2. TGA traces of Cell-NHC and Cell-NHC-Pd catalyst under air flow. XPS spectroscopy was employed to investigate the elemental coordination state of Pd

171

species. As shown in Fig. 3a, the binding energies of Pd3d5/2 in the Cell-NHC-Pd catalyst

172

were 335.3 eV and 337.4 eV, respectively (Xu et al., 2008), which revealed that both Pd (0)

173

and Pd (II) were present. Based on previous literature reports (Wu et al., 2013; Xiong et al.,

174

2013), parts of Pd (II) species were likely reduced in the existence of cellulose backbone.

Ac ce pt e

d

170

9  Page 9 of 26

ip t cr us an Ac ce pt e

d

M

175

176

10  Page 10 of 26

ip t cr us an

177

Fig. 3. XPS spectrum of fresh Cell-NHC-Pd catalyst: Pd 3d (a), C 1s (b), N 1s (c).

M

178

As shown in Fig. 3b, the status of C in Cell-NHC-Pd can be also explored by XPS

180

analysis of C 1s. The C 1s peak was divided into four components at binding energies of

181

282.8, 284.8, 286.5 and 288.0 eV, which were attributed to C=C, C–C, C–O and carbene

182

carbon to Pd bonds, respectively (Khabashesku et al., 2000; Mao et al., 2011).

183

Correspondingly, XPS of N 1s (Fig. 3c) showed 401.3 eV as the major peak and 397.4 eV as

184

a minor peak, which were assigned to N–sp2C and N–sp3C bonds (Gammon et al., 2003; Li

185

et al., 2009). Combining these results with those obtained from FTIR, it can be speculated

186

that N-methylimidazole on cellulose backbone captured the Pd species successfully, as

187

shown by the model in scheme 1.

Ac ce pt e

d

179

188

Surface morphology of the cellulose fibers was investigated with transmission electron

189

microscopy (TEM). TEM images (Fig. 4) revealed the presence of palladium nanoparticles

190

of about 9 nm in size, which proved the excellent dispersion of palladium sites on the surface

11  Page 11 of 26

of cellulose.

193 194

Fig. 4. TEM images of the fresh Cell-NHC-Pd (a and c) and recovered (b and d).

Ac ce pt e

192

d

M

an

us

cr

ip t

191

195

3.2. Applying the catalyst to Suzuki cross-coupling reactions

196

3.2.1 Evaluation of its catalytic activity

197

Initial examination was carried out using p-bromoanisole (0.250 mmol, 1.0 equiv) and

198

phenylboronic acid (0.375 mmol, 1.5 equiv) as a model reaction in the presence of

199

Cell-NHC-Pd. Several variables such as the amount of catalyst, temperature, solvent and

200

base were screened to optimize the reaction conditions. Results were summarized in Table 1.

201

As shown in Table 1, the product yield increased with the increase of the amount of

202

catalyst (Table 1, entries 1-3). However, the yield decreased when the amount of catalyst

12  Page 12 of 26

increased from 0.75 to 1.25 mol% (Table 1, entries 3-5). Therefore, 0.75 mol% of the

204

catalyst (Table 1, entry 3) was sufficient to efficiently catalyze the reaction. The coupling

205

yield increased slightly from 60 °C to 80 °C (Table 1, entries 6 and 7). However, the yield

206

decreased when the temperature rose to 100 °C (Table 1, entry 8). We tried a variety of

207

solvent systems, going from pure water (Table 1, entry 12) to a mixture of an organic solvent

208

with water. The organic solvents tried were dimethylformamide (DMF), dimethylacetamide

209

(DMAc) and Ethanol (EtOH) (Table 1, entries 6-11, 13-17). Results showed that DMF : H2O

210

(1:1) was the most favorable system. Using potassium carbonate and potassium tert-butoxide

211

(Table 1, entries 7 and 16) both gave excellent yields. However, due to the hygroscopic

212

nature of potassium tert-butoxide and considering its higher cost, we opted for potassium

213

carbonate. When we increased the amount of K2CO3 to 3 equivalents, the yield rose to 98%

214

(Table 1, entry 13).

215

Ac ce pt e

d

M

an

us

cr

ip t

203

216

Table 1

217

Cell-NHC-Pd catalyzed Suzuki cross-coupling reaction.

218 Entry a

Amount of catalyst (mol%)

Temperature (°C)

Solvent

Base

Yield (%) b

0.25

80

DMF : H2O (1:1)

K2CO3

79

0.50

80

DMF : H2O (1:1)

K2CO3

85

3

0.75

80

DMF : H2O (1:1)

K2CO3

94

4

1.00

80

DMF : H2O (1:1)

K2CO3

92

5

1.25

80

DMF : H2O (1:1)

K2CO3

85

1 2

13  Page 13 of 26

222

60

DMF : H2O (1:1)

K2CO3

89

7

0.75

80

DMF : H2O (1:1)

K2CO3

93

8

0.75

100

DMF : H2O (1:1)

K2CO3

84

9

0.75

80

DMAc : H2O (1:1)

K2CO3

91

10

0.75

80

DMF : H2O (1:3)

K2CO3

76

11

0.75

80

EtOH : H2O (1:1)

K2CO3

86

12

0.75

80

H2O

K2CO3

22

13 c

0.75

80

DMF : H2O (1:1)

K2CO3

98

14

0.75

80

DMF : H2O (1:1)

Cs2CO3

88

15

0.75

80

DMF : H2O (1:1)

K3PO4

90

16

0.75

80

DMF : H2O (1:1)

t-BuOK

92

17

0.75

80

DMF : H2O (1:1)

Et3N

52

cr

us

an M

d

a

ip t

0.75

Reaction conditions: 4-bromoanisole (0.250 mmol), phenylboronic acid (0.375 mmol), base (0.5 mmol),

solvent (2.5 mL), heating in air for 2 h, detected by TLC. b Isolated yield. c K2CO3 was used for 3 equiv (0.75 mmol).

Ac ce pt e

219 220 221

6

Overall, the optimized conditions were as follows: phenylboronic acid (0.375 mmol),

223

4-bromoanisole (0.250 mmol), catalyst (0.75 mol%), K2CO3 (0.75 mmol) in DMF : H2O (1:1)

224

(2.5 mL) at 80 °C under air.

225

To extend the scope of the heterogeneous reaction, the Suzuki cross-coupling reaction of

226

aryl halides with various phenylboronic acids were examined as shown in Table 2. Most of

227

the aryl bromides were converted to the corresponding biaryls with good to excellent yields

228

in reasonable reaction times. Coupling of 4-bromoanisole with sterically hindered

229

phenylboronic acids proceeded with different yields (Table 2, entries 3-5 and 6-8). The

230

reaction of electron-deficient aryl bromides (4-bromobenzonitrile as a model substrate) with 14  Page 14 of 26

phenylboronic acid occurred smoothly to afford the biaryls with excellent yields within 4 h

232

at 80 °C (Table 2, entries 10-18). However, the reaction between benzene halides and

233

4-ethylphenylboronic acid was sluggish and gave only small amounts of products even with

234

prolonged reaction times (Table 2, entries 19-21). The reaction of electron-deficient

235

bromobenzene with 4-fluorine phenylboronic acid went smoothly. On the other hand, the

236

reaction of electron-rich bromobenzene with 4-fluorine phenylboronic acid was not

237

successful, with a yield of only 10%.

us

cr

ip t

231

238 Table 2

240

Suzuki Cross-coupling Reaction of Aryl halide with Phenylboronic acidsa.

241 Aryl halide

1

Br

3 4 5

B(OH)2

OCH 3

Ac ce pt e

2

Phenylboronic acid

Br

OCH 3

F3 C

Br

OCH 3

H3 CO

Br

OCH 3

B(OH)2

B(OH)2

7

OCH 3

Br

OCH 3

F

B(OH) 2

Br

OCH 3

10

Br

CN

78

OCH 3

1.5

85

1.5

80

1.5

66

5.0

91

OCH 3

5.0

81

OCH 3

5.0

45

5.0

93

2.0

99

OCH 3

OCH 3 OCH 3

F

OCH 3

B(OH) 2 F

F

Br

4.5

F

OCH 3

9

OCH 3

H 3 CO

B(OH)2

F

8

97

H 3CO

B(OH)2

OCH 3

Br

2.0

OCH 3

F3C

OCH3

6

Yield (%)c

B(OH)2

H3 CO

Br

Time (h)b

Product

d

Entry

M

an

239

C 2H 5

B(OH)2

C2 H5

B(OH)2

OCH 3

CN

15  Page 15 of 26

11

Br

CN

F3 C

B(OH)2

F3C

12

Br

CN

H3 CO

B(OH)2

H 3CO

13

Br

CN

CN

B(OH)2 H3 CO

CN

B(OH)2 OCH3

16

Br

CN

Br

F

F

B(OH) 2

Br

B(OH)2

19

B(OH) 2

CN

I

C 2H 5

B(OH)2

C2 H5

20

Br

C 2H 5

B(OH)2

C2 H5

21

Cl

C 2H 5

B(OH)2

CN

M

C2H5

C2 H5

F

Br

d

B(OH) 2

OCH 3

F

B(OH) 2

CH 3

242 243 244 245 246

a

96

95

83

2.3

98

4.0

61

4.0

53

4.0

33

3.5

76

6.0

10

OCH 3

F

Br

Ac ce pt e

23

2.6

F

B(OH)2

F

2.0

CN

C 2H 5

22

95

F

CN

CN

1.5 2.0

CN

F

18

97

OCH 3

F

17

4.0

ip t

CN

99

cr

Br

4.0

H3CO

CN

15

80

us

Br

CN

1.5

an

14

CN

CH 3

Unless otherwise specified, all the reactions were carried out using aryl halides (0.250 mmol), phenylboronic

acids (0.375 mmol), K2CO3 (0.75 mmol), Cell-NHC-Pd (0.75 mol%), DMF : H2O (1:1) (2.5 mL), under 80 °C in air. b Detected by TLC. c Isolated yield.

3.2.2 Separation of the catalyst and recycling tests.

247

The recyclability of the Cell-NHC-Pd catalyst was also investigated in the Suzuki

248

cross-coupling reaction of 4-bromoanisole with phenylboronic acid. After the reaction, the

249

liquid phase was removed by centrifugation and the catalyst was washed with ethyl acetate.

250

The catalyst was then used in the four additional reactions and the catalytic activity for the

251

five consecutive cycles was shown in Table 3. For the second and third reactions, only a

252

minor loss of catalytic activity was observed compared to the initial run. However, the 16  Page 16 of 26

reaction at the 4th and 5th recycled runs gave the products in 60% and 58% yields,

254

respectively. In addition, the filtrate resulting from the reaction mixture had no catalytic

255

activity under identical conditions. The total amount of palladium leaching after the 5th cycle

256

was 1.2% as detected by ICP-AES analysis (The amount of palladium is 2.55% before

257

reaction and 2.52% after 5 cycles). Therefore, the decrease of activity may be due to the

258

aggregation of palladium nanoparticles during the reaction as shown in TEM images of Fig.

259

4 (Narayanan & El-Sayed, 2003; Hu & Liu, 2005; Sin et al., 2010).

us

cr

ip t

253

260

266 267

an

Reusability of the Cell-NHC-Pd in Suzuki cross-coupling reactiona. 1st

2nd

3rd

Time (h) b

2.6

3

3

Yield (%) c

91

85

79

a

4th

M

Recycle

5th

3.5

4

60

58

d

263 264 265

Table 3

The reactions conditions: 4-bromoanisole (0.250 mmol), phenylboronic acid (0.375 mmol), K2CO3 (0.75

Ac ce pt e

261 262

mmol), Cell-NHC-Pd (0.75 mol%), DMF : H2O (1:1) (2.5 mL), under 80 °C in air. b Detected by TLC.

c

Isolated yield.

4. Conclusions

268

In conclusion, cellulose-supported N-methylimidazole was prepared by the nucleophilic

269

substitution reaction of N-methylimidazole with cellulose tosylate. The success of the

270

grafting of N-methylimidazole onto the cellulose backbone was confirmed by FT-IR and

271

XPS analysis. Cellulose-supported N-methylimidazole was then coordinated with Pd(OAc)2

272

in DMF to give the final cellulose-tagged N-methylimidazole palladium complex, which

273

showed high thermal ability from TG analysis. Nanopalladiums sites on the surface of the

274

cellulose support were very well distributed, as demonstrated by TEM images. Moreover, 17  Page 17 of 26

this catalyst efficiently catalyzed Suzuki cross-coupling reactions and gave the

276

corresponding products in good to excellent yields. Of particular note is that the

277

Cell-NHC-Pd catalyst can be recycled by simple filtration following a Suzuki cross-coupling

278

reaction. Further applications of cellulose-based N-methylimidazole in other organic

279

reactions are currently being investigated in our laboratory.

ip t

275

cr

280

Acknowledgment

282

This work was supported by National Natural Science Foundation of China (No. 20972109).

us

281

an

283

References

285

Birkholz, M. N.(née Gensow), Freixa, Z., & van Leeuwen, P. W. N. M. (2009). Bite angle

286

effects of diphosphines in C–C and C–X bond forming cross coupling reactions. Chem

287

Soc Rev 38, 1099–1118.

289

d

Ac ce pt e

288

M

284

Blaser, H. U. (2002). The Chiral Switch of (S)-Metolachlor: A Personal Account of an Industrial Odyssey in Asymmetric Catalysis. Adv Synth Catal, 344, 17–31.

290

Cirtiu, C. M., Dunlop-Brière, A. F., & Moores, A. (2011). Cellulose nanocrystallites as an

291

efficient support for nanoparticles of palladium: application for catalytic hydrogenation

292

and Heck coupling under mild conditions. Green Chem, 13, 288–291.

293 294 295 296

Cole-Hamilton, D. J. (2003). Homogeneous Catalysis-New Approaches to Catalyst Separation, Recovery, and Recycling. Science, 299, 1702–1706. Corbet, J. P., & Mignani, G. (2006). Selected Patented Cross-Coupling Reaction Technologies. Chem Rev, 106, 2651–2710.

18  Page 18 of 26

297

Dorta, R., Stevens, E. D., Scott, N. M., Costabile, C., Cavallo, L., Hoff, C. D., & Nolan, S. P.

298

(2005). Steric and Electronic Properties of N-Heterocyclic Carbenes (NHC): A Detailed

299

Study on Their Interaction with Ni(CO)4. J Am Chem Soc, 127, 2485–2495.

301

Drge, T., & Glorius, F. (2010). The Measure of All Rings—N-Heterocyclic Carbenes.

ip t

300

Angew Chem Int Ed, 49, 6940–6952.

Du, Q., & Li, Y. (2011). Air-stable, recyclable, and time-efficient diphenylphosphinite

303

cellulose-supported palladium nanoparticles as a catalyst for Suzuki-Miyaura reactions.

304

Beilstein J Org Chem, 7, 378–385.

us

cr

302

Du, Q., & Li, Y. (2012). Application of an air-and-moisture-stable diphenylphosphinite

306

cellulose-supported nanopalladium catalyst for a Heck reaction. Res Chem Intermed, 38,

307

1807–1817.

M

d

309

Fleckenstein, C., Roy, S., Leuthäußer, S., & Plenio, H. (2007). Sulfonated N-heterocyclic carbenes for Suzuki coupling in water. Chem Commun, 2870–2872.

Ac ce pt e

308

an

305

310

Fu, G. C. (2008). The Development of Versatile Methods for Palladium-Catalyzed Coupling

311

Reactions of Aryl Electrophiles through the Use of P(t-Bu)3 and PCy3 as Ligands.

312

Accounts Chem Res, 41, 1555–1564.

313

Gammon, W. J., Kraft, O., Reilly, A. C., & Holloway, B. C. (2003). Experimental

314

comparison of N(1s) X-ray photoelectron spectroscopy binding energies of hard and

315

elastic amorphous carbon nitride films with reference organic compounds. Carbon, 41,

316

1917–1923.

317 318

Gao,

T.

T.,

Jin,

A.

P.,

&

Shao,

L.

X.

(2012).

N-Heterocyclic

carbene–palladium(II)-1-methylimidazole complex catalyzed Mizoroki–Heck reaction of

19  Page 19 of 26

319

aryl chlorides with styrenes. Beilstein J Org Chem, 8, 1916–1919. Hassan, J., Sévignon, M., Gozzi, C., Schulz, E., & Lemaire, M. (2002). Aryl-Aryl Bond

321

Formation One Century after the Discovery of the Ullmann Reaction. Chem Rev, 102,

322

1359–1469.

ip t

320

Herrmann, W. A., Elison, M., Fischer, J., Köcher, C., & Artus, G. R. J. (1995). Metal

324

Complexes of N-Heterocyclic Carbenes—A New Structural Principle for Catalysts in

325

Homogeneous Catalysis. Angew Chem Int Ed Eng, 34, 2371–2374.

us

cr

323

Herrmann, W. A., Reisinger, C. P., & Spiegle, M. (1998). Chelating N-heterocyclic carbene

327

ligands in palladium-catalyzed heck-type reactions. J Organomet Chem, 557, 93–96.

328

He, Y., & Cai, C. (2011). A simple procedure for the polymer-supported N-heterocyclic

329

carbene–rhodium complex via click chemistry: a recyclable catalyst for the addition of

330

arylboronic acids to aldehydes. Chem Commun , 47, 12319–12321.

332

M

d

Hu, J., & Liu, Y. (2005). Pd Nanoparticle Aging and Its Implications in the Suzuki

Ac ce pt e

331

an

326

Cross-Coupling Reaction. Langmuir, 21, 2121–2123.

333

Jamwal, N., & Sodhi, R. K., Gupta, P., Paul, S., (2011). Nano Pd(0) supported on cellulose:

334

A highly efficient and recyclable heterogeneous catalyst for the Suzuki coupling and

335

aerobic oxidation of benzyl alcohols under liquid phase catalysis. Int J Bio Macromol, 49,

336

930–935.

337

Kang, T., Feng, Q., & Luo, M. (2005). An Active and Recyclable Polystyrene-Supported

338

N-Heterocyclic Carbene–Palladium Catalyst for the Suzuki Reaction of Arylbromides

339

with Arylboronic Acids under Mild Conditions. Synlett, 2305–2308.

340

Keshipour, S., Shojaei, S., & Shaabani, A. (2013). Palladium nano-particles supported on

20  Page 20 of 26

341

ethylenediamine functionalized cellulose as a novel and efficient catalyst for the Heck

342

and Sonogashira couplings in water. Cellulose, 20, 973–980.

344

Khabashesku, V. N., Zimmerman, J. L., & Margrave, J. L. (2000). Powder Synthesis and Characterization of Amorphous Carbon Nitride. Chem Mater, 12, 3264–3270.

ip t

343

Kim, J. H., Kim, J. W., Shokouhimehr, M., & Lee, Y. S. (2005). Polymer-Supported

346

N-Heterocyclic Carbene-Palladium Complex for Heterogeneous Suzuki Cross-Coupling

347

Reaction. J Org Chem, 70, 6714–6720.

us

cr

345

Kim, J. W., Kim, J. H., Lee, D. H., & Lee, Y. S. (2006). Amphiphilic polymer supported

349

N-heterocyclic carbene palladium complex for Suzuki cross-coupling reaction in water.

350

Tetrahedron Lett, 47, 4745–4748.

353 354

M

Biopolymer and Sustainable Raw Material. Angew Chem Int Ed, 44, 3358–3393.

d

352

Klemm, D., Heublein, B., Fink, H.-P., & Bohn, A. (2005). Cellulose: Fascinating

Kurt, R., Sanjines, R., Karimi, A., & Lévy, F. (2000). Structural and mechanical properties

Ac ce pt e

351

an

348

of CNx thin films prepared by magnetron sputtering. Diamond Relat Mater, 9, 566–572.

355

Lázaro, G., Fernández-Alvarez, F. J., Iglesias, M., Horna, C., Vispe, E., Sancho, R., Lahoz, F.

356

J., Iglesias, M., Pérez-Torrente, J. J., & Oro, L. A. (2014). Heterogeneous catalysts based

357

on supported Rh–NHC complexes: synthesis of high molecular weight poly(silyl ether)s

358

by catalytic hydrosilylation. Catal Sci Technol, 4, 62–70.

359 360 361 362

Lemo, J., Heuze, K., & Astruc, D. (2005) .Efficient Dendritic Diphosphino Pd(II) Catalysts for the Suzuki Reaction of Chloroarenes. Org Lett, 7, 2253–2256. Littke, A. F., & Fu, G. C. (2002). Palladium-Catalyzed Coupling Reactions of Aryl Chlorides. Angew Chem Int Ed, 41, 4176–4211.

21  Page 21 of 26

363

Li, X., Wang, H., Robinson, J. T., Sanchez, H., Diankov, G., & Dai, H. (2009).

364

Simultaneous Nitrogen Doping and Reduction of Graphene Oxide. J Am Chem Soc, 131,

365

15939–15944. Li, Y., Hong, X. M., Collard, D. M., & El-Sayed, M. A. (2000). Suzuki Cross-Coupling

367

Reactions Catalyzed by Palladium Nanoparticles in Aqueous Solution. Org Lett, 2,

368

2385–2388.

cr

ip t

366

Lo, H. K., & Luo, F. T. (2012). Synthesis of PS-supported NHC-Pd Catalyst Derived from

370

Theobromine and its Applications in Suzuki-Miyaura Reaction. J Chin Chem Soc, 59,

371

0000-0000.

an

us

369

Mao, P., Yang, L., Liu, X., Cai, Y., Zhao, C., & Song, M. (2011). Synthesis and Application

373

in Suzuki Reaction of SiO2-Supported Chiral Imidazole Thioether Ligand Palladium

374

Catalyst. Chin J Org Chem, 31, 1828–1834.

376

d

Marion, N., & Nolan, S. P. (2008). N-Heterocyclic carbenes in gold catalysis. Chem Soc Rev,

Ac ce pt e

375

M

372

37, 1776–1782.

377

Martin, R., & Buchwald, S. L. (2008). Palladium-Catalyzed Suzuki-Miyaura Cross-Coupling

378

Reactions Employing Dialkylbiaryl Phosphine Ligands. Accounts Chem Res, 41,

379

1461–1473.

380 381 382 383 384

Meijere, A. de., & Diederich, F. (2004). Metal-Catalyzed Cross-Coupling Reactions. Wiley-VCH, Weinheim, Germany. Miyaura, N., & Suzuki, A. (1995). Palladium-Catalyzed Cross-Coupling Reactions of Organoboron Compounds. Chem Rev, 95, 2457–2483. Molnár, Á., & Papp, A. (2014). The use of polysaccharides and derivatives in

22  Page 22 of 26

385

palladium-catalyzed coupling reactions. Catal Sci Technol, 4, 295. Narayanan, R., & El-Sayed, M. A. (2003). Effect of Catalysis on the Stability of Metallic

387

Nanoparticles: Suzuki Reaction Catalyzed by PVP-Palladium Nanoparticles. J Am Chem

388

Soc, 125, 8340–8347.

ip t

386

Old, D. W., Wolfe, J. P., & Buchwald, S. L., (1998). A Highly Active Catalyst for

390

Palladium-Catalyzed Cross-Coupling Reactions: Room-Temperature Suzuki Couplings

391

and Amination of Unactivated Aryl Chlorides. J Am Chem Soc, 120, 9722–9723.

us

cr

389

Pozo, C. d., Corma, A., Iglesias, M., & Sánchez, F. (2011). Recyclable mesoporous

393

silica-supported chiral ruthenium-(NHC)NN-pincer catalysts for asymmetric reactions.

394

Green Chem, 13, 2471–2481.

M

an

392

Rahn, K., Diamantoglou, M., Klemm, D., Berghmans, H., & Heinze, T. (1996).

396

Homogeneous synthesis of cellulose p-toluenesulfonates in N,N-dimethylacetamidel

397

LiCl solvent system. Die Angew Makromol Chem, 238, 143–163.

Ac ce pt e

d

395

398

Reddy, K. R., Kumar, N. S., Reddy, P. S., Sreedhar, B., & Kantam, M. L. (2006). Cellulose

399

supported palladium(0) catalyst for Heck and Sonogashira coupling reactions. J Mol

400

Catal A: Chem, 252, 12–16.

401

Rodil, S., Morrison, N. A., Milne, W. I., Robertson, J., Stolojan, V., & Jayawardane, D. N.

402

(2000). Deposition of carbon nitride films using an electron cyclotron wave resonance

403

plasma source. Diamond Relat Mater, 9, 524–529.

404 405 406

Sin, E., Yi, S., & Lee, Y. (2010). Chitosan-g-mPEG-supported palladium (0) catalyst for Suzuki cross-coupling reaction in water. J Mol Catal A Chem, 315, 99–104. Suzuki, A. (1999). Recent advances in the cross-coupling reactions of organoboron

23  Page 23 of 26

407 408

derivatives with organic electrophiles, 1995–1998. J Organomet Chem, 576, 147–168. Wang,

Z.

Y.,

Chen,

G.

Q.,

&

Shao,

L.

(2012).

Catalyzed

N-Heterocyclic

409

Carbene−Palladium(II)−1-Methylimidazole

Suzuki−Miyaura

410

Coupling of Aryl Sulfonates with Arylboronic Acids. J Org Chem, 77, 6608–6614.

ip t

Complex

X.

Wu, X., Lu, C., Zhang, W., Yuan, G., Xiong, R., & Zhang, X. (2013). A novel reagentless

412

approach for synthesizing cellulose nanocrystal-supported palladium nanoparticles with

413

enhanced catalytic performance. J Mater Chem A, 1, 8645–8652.

us

cr

411

Xiong, R., Lu, C., Wang, Y., Zhou, Z., & Zhang, X. (2013). Nanofibrillated cellulose as the

415

support and reductant for the facile synthesis of Fe3O4/Ag nanocomposites with catalytic

416

and antibacterial activity. J Mater Chem A, 1, 14910–14918.

M

an

414

Xu, Y., Wang, X., & Wei, Y. (2014). Heck coupling reactions of aryl halides with acrylates

418

catalyzed by cellulose-supported triphenylphosphine palladium nanoparticles. Chemistry

419

& Bioengineering, 1, 43-47.

421 422 423

Ac ce pt e

420

d

417

Xu, Y., Zhang, L., & Cui, Y. (2008). Catalytic Performance of Cellulose Supported Palladium Complex for Heck Reaction in Water. J Appl Polym Sci, 110, 2996–3000. Yin, L., & Liebscher, J. (2007). Carbon-Carbon Coupling Reactions Catalyzed by Heterogeneous Palladium Catalysts. Chem Rev, 107, 133–173.

424

Zhang, C., Huang, J., Trudell, M. L., & Nolan, S. P. (1999). Palladium-Imidazol-2-ylidene

425

Complexes as Catalysts for Facile and Efficient Suzuki Cross-Coupling Reactions of

426

Aryl Chlorides with Arylboronic Acids. J Org Chem, 64, 3804–3805.

427 428

Zhao, X. A., Ong, C. W., Tsang, Y. C., Wong, Y. W., Chan, P. W., & Choy, C. L. (1995). Reactive pulsed laser deposition of CNx films. Appl Phys Lett, 66, 2652–2654.

24  Page 24 of 26

429 430

Zheng, C., Li, Y., & Zheng, W. (2009). Ethylenediamine-Functionalized Cellulose Supported Nano-palladium for Suzuki Reaction. Chin J Org Chem, 29, 1983–1987.

Ac ce pt e

d

M

an

us

cr

ip t

431

25  Page 25 of 26

432 433 434 435 436 437

Highlights: NHC ligand was introduced onto cellulose backbone to explore an avenue for polysaccharide-based catalysts. Cellulose-supported NHC-Pd complex can efficiently promote Suzuki coupling reactions. Catalysts can be recovered by easy filtration and can be used several times.

ip t

431

Ac ce pt e

d

M

an

us

cr

438

26  Page 26 of 26

Cellulose-supported N-heterocyclic carbene-palladium catalyst: synthesis and its applications in the Suzuki cross-coupling reaction.

A cellulose-supported N-methylimidazole-palladium catalyst (Cell-NHC-Pd) was synthesized and used for Suzuki cross-coupling reactions between aryl hal...
2MB Sizes 1 Downloads 6 Views