Next Article in Journal
Microwave-Assisted Improved Synthesis of Pyrrolo[2,3,4-kl]acridine and Dihydropyrrolo[2,3,4-kl]acridine Derivatives Catalyzed by Silica Sulfuric Acid
Previous Article in Journal
Extraction of Antioxidant Components from Bidens pilosa Flowers and Their Uptake by Human Intestinal Caco-2 Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Efficient Microwave-Assisted Suzuki Reaction using a New Pyridine-Pyrazole/Pd(II) Species as Catalyst in Aqueous Media

Key Laboratory of Development and Application of Forest Chemicals of Guangxi, College of Chemistry and Chemical Engineering, Guangxi University for Nationalities, Nanning, 530006, China
*
Authors to whom correspondence should be addressed.
Molecules 2013, 18(2), 1602-1612; https://doi.org/10.3390/molecules18021602
Submission received: 27 December 2012 / Revised: 21 January 2013 / Accepted: 22 January 2013 / Published: 25 January 2013

Abstract

:
A new pyridine-pyrazole N–N ligand has been conveniently synthesized and characterized by 1H-, 13C-NMR, IR spectroscopies, HRMS and X-ray single-crystal crystallography analyses. The ligand adds to palladium(II) under basic conditions to give high yields of an air-stable and water-soluble complex that was fully characterized by NMR and HRMS. The complex was investigated as a catalyst for the Suzuki reaction in aqueous media under microwave irradiation. The compound proved to be an effective catalyst.

1. Introduction

In the past three decades, the palladium-catalyzed Suzuki reaction of aryl halides with aryl boronic acids has been one of the most important and efficient methods for the formation of unsymmetrical biaryl compounds [1,2], which are extensively found in a range of natural products [3], pharmaceuticals [4], ligands [5], herbicides and advanced materials [6]. Modern techniques have been developed in order to enable simpler, faster and cheaper versions of already known chemical transformations to meet the purposes of Green Chemistry [7]. Recently, the use of aqueous phases including water and water/organic mixtures as solvents for the Suzuki reaction has also received considerable attention, as water is cheap, environmental friendly, inflammable, and allows simple separation and catalyst recycling [8,9]. For these reasons, much effort has been directed to perform the Suzuki coupling in neat water. Casalnuovo [10] was the first to demonstrate that the palladium-catalyzed cross-coupling reactions could be carried out in aqueous solvents catalyzed by TPPMS/Pd(OAc)2, and since then, other hydrophilic ligands for aqueous-phase Suzuki cross-coupling reactions have been developed [11]. Recently the application of nitrogen-based ligands systems, such as Schiff bases, guanidine, aryloximes, arylimines, has also been found to produce highly active catalysts for Suzuki reaction in water [12,13,14,15,16,17]. Nowadays, microwave heating has been widely exploited in organic synthesis and its advantages over traditional heating lay in the reduction in reaction time from hours to minutes [18,19]. In addition, microwave heating is generally able to reduce side reactions, increase yields and improve reproducibility [20]. Microwave-assisted Suzuki reactions can be considered today as an efficient synthetic methodology [21,22,23,24,25].
To the best of our knowledge, Pd(II) complexes with the pyridine-pyrazole ligand have been examined for cytotoxic activity and few investigations have been carried out on their catalytic activity [26,27,28]. In this paper, we report carboxylated water-soluble pyridine-pyrazole ligands as supporting ligands for the Suzuki reaction in water and in aqueous phases in conjunction with microwave heating.

2. Results and Discussion

2.1. Synthesis and Characterization of the Ligand and Complee

The synthetic pathway for compounds 3 and 4 was shown in Scheme 1. 5-Hydroxy-1-pyridin-2-yl-1H-pyrazole-3-carboxylic acid methyl ester (1) was easily synthesized by the reaction of 2-hydrazinopyridine and dimethyl acetylenedicarboxylate (DMADC), with subsequent cyclization in methanolic NaOCH3 solution.
Scheme 1. Synthesis of pyridine-pyrazole ligand 3 and the Pd(II) complex 4.
Scheme 1. Synthesis of pyridine-pyrazole ligand 3 and the Pd(II) complex 4.
Molecules 18 01602 g003
The alkylation of 2 with iodomethane provided 5-methoxy-1-pyridin-2-yl-1H-pyrazole-3-carboxylic acid methyl ester (2). Compound 2 was hydrolyzed with LiOH in methanol at room temperature to afford 5-methoxy-1-pyridin-2-yl-1H-pyrazole-3-carboxylic acid (3) bearing a carboxylic acid group. Compound 3 has been characterized by 1H-NMR, 13C-NMR, IR and high resolution mass spectrometry (HRMS). The solid-state structure has been established by X-ray single-crystal crystallography [29] (Figure 1). The reaction of a methanol solution of ligand 3 with an equimolar amount of an aqueous solution of K2PdCl4 produced the palladium(II) complex (4) in 81% yield as yellow crystals. The mononuclear complex with a pyridine-pyrazole N,N-chelate has been characterized by IR, UV/vis (Figure 2), NMR and high resolution mass spectrometry.
Figure 1. The molecular structure of the compound 3 with atom-numbering scheme, displacement ellipsoids are drawn at the 30% probability level.
Figure 1. The molecular structure of the compound 3 with atom-numbering scheme, displacement ellipsoids are drawn at the 30% probability level.
Molecules 18 01602 g001
Figure 2. (a) UV-vis spectra of K2PdCl4 in water; (b) UV-vis spectra of complex 4 in water.
Figure 2. (a) UV-vis spectra of K2PdCl4 in water; (b) UV-vis spectra of complex 4 in water.
Molecules 18 01602 g002
We now observe the interaction of K2PdCl4 with the ligand 3 by UV/vis spectroscopy. In Figure 2, as can be seen, the spectrum of K2PdCl4 displays one major absorption band at 208 nm, that changes upon addition of ligand 3 (195 and 294 nm). This change in the UV/vis spectrum can be attributed to the coordination of Pd(II) to the ligand 3 [30].

2.2. Catalysis of the Suzuki Reaction

As a starting point for the development of our microwave-promoted methodology, the Suzuki reaction were performed using a scientific microwave apparatus [31], working on a 1 mmol scale in a 10 mL sealed glass vessel. The Suzuki reaction of different types of aryl halides using complex 4 as catalyst was then investigated under microwave irradiation. Initially, in order to optimize the reaction conditions, we employed the coupling reaction of 4'-bromoacetophenone with phenylboronic acid in water/ethanol as a model reaction to investigate the effect of different bases on the reaction. Water/ethanol as solvent facilitates solvation of the aryl halide in neat water [32]. For this purpose, this reaction was performed using different bases in the presence of 0.1 mol% of complex 4 as catalyst. Using a microwave power of 60 W, we ramped the temperature from 25 °C to 120 °C, and then held it at 120 °C for 2 min. The results are summarized in Table 1. The best results were obtained using KOH as the base (Table 1, entries 1–4), other bases, such as K3PO4, K2CO3 and Et3N were slightly less efficient. We next investigated the effect of different solvents for the same reaction (Table 1, 5–8). As can be seen in Table 1, the best result was obtained using aqueous EtOH/H2O as the solvent (Table 1), perhaps, attributable to the better solubility of the reagents and easier reduction of Pd2+ to Pd(0).
Table 1. Optimization of base and solvent for Suzuki cross-coupling reaction under microwave irradiation a.
Table 1. Optimization of base and solvent for Suzuki cross-coupling reaction under microwave irradiation a.
Molecules 18 01602 i001
EntryBaseSolventConversion (%) b
1KOHEtOH/H2O99
2K3PO4EtOH/H2O95.2
3K2CO3EtOH/H2O96.0
4Et3NEtOH/H2O92.3
5KOHDMF/H2O90.9
6KOHDMAC/H2O85.4
7KOHMeCN/H2O94.5
8KOHH2O85.0
a Reaction conditions: 1 mmol 4'-bromoacetophenone, 1.3 mmol phenylboronic, 2 mmol base, 2 mL solvent (EtOH/H2O = 1:1; MeCN/H2O = 1:1; DMF/H2O = 1:1; DMA/H2O = 1:1), Microwave irradiation = 60 W. and 0.1 mol% complex 4. DMF = N,N-Dimethylformamide; DMAC = N,N-Dimethylacetamide; b Determined by HPLC analysis.
Under the above optimized conditions, complex 4 was applicable to a wide range of aryl bromides, iodides, and less reactive aryl chlorides and the results are summarized in Table 2. Generally, the catalyst was very effective when electron poor or electron-rich aryl bromides were used (Table 2, entry 1–7). From Table 2, it can be seen that the electronic and steric characters of the aryl bromides have an effect on the Suzuki reactions under the optimized conditions. For this purpose we applied 2-, 3-, and 4-bromobenzaldehyde under the optimized reaction conditions (Table 2, entries 1–3). The Suzuki reaction of 4-bromobenzaldehyde was complete and gave excellent yields. Sterically demanding 2-bromobenzaldehyde resulted in decreased coupling. The reaction of 4-chlorobromobenzene with phenylboronic acid produced mono-substituted 4-chlorobiphenyl (Table 2, entry 8), the result indicates that the catalytic reaction has a good chemical selectivity. The coupling reaction using 4-methoxyphenylboronic acid was also performed, the corresponding products are obtained in good yields (Table 2, entry 14–26). We also investigated the scope of this method on aryl chlorides, and found that the conversion of activated aryl chloride was up to 89.0% (Table 2, entry 26) and unactivated aryl chlorides gave moderate yields.
Table 2. Suzuki coupling of aryl halides and aryl boronic acids in H2O/EtOH using complex 4 under optimized reaction conditions under microwave irradiation a.
Table 2. Suzuki coupling of aryl halides and aryl boronic acids in H2O/EtOH using complex 4 under optimized reaction conditions under microwave irradiation a.
Molecules 18 01602 i002
EntryXYZYield (%) b
1Br4-CHOH90.3
2Br3-CHOH85.1
3Br2-CHOH80.3
4Br4-OMeH86.7
5Br4-CH3H82.4
6Br4-OHH93
7Br4-COMeH92.6
8Br4-ClH90.8
9BrHH86.7
10IHH91.3
11ClHH63.8
12Cl2-COOHH75.5
13Cl4-NO2H88.3
14Br4-CHO4-OMe92.7
15Br3-CHO4-OMe87.7
16Br2-CHO4-OMe83.5
17Br4-OMe4-OMe88.1
18Br4-CH34-OMe85.3
19Br4-OH4-OMe93.2
20Br4-COMe4-OMe93.0
21Br4-Cl4-OMe91.2
22BrH4-OMe90.8
23IH4-OMe92.7
24ClH4-OMe67.2
25Cl2-COOH4-OMe78.4
26Cl4-NO24-OMe89.0
a Reaction conditions: 1 mmol 4'-bromoacetophenone, 1.3 mmol phenylboronic, 2 mmol KOH, 2 mL solvent (EtOH/H2O = 1:1), Microwave irradiation = 60 W, and 0.1 mol% complex 4; b Isolated yield after purification by column chromatography.

2.3. Catalyst Recycling

The potential recyclability of the catalysts derived from the pyridine-pyrazole/Pd system was examined using the model cross-coupling of 4-bromoacetophenone with phenylboronic acid. The reaction was carried out in 1:1 H2O/CH3CH2OH under microwave irradiation for 2 min. After extracting the products with dichloromethane the yields were determined by HPLC. The resulting aqueous solution was recharged with the same substrates for the next cycle. It was shown that the catalytic solution could be reused for third cycles, and by the fifth cycle, the yield had dropped to 21% (Table 3).
Table 3. Recycling of catalyst a.
Table 3. Recycling of catalyst a.
Cycles(n)12345
Conversion (%) b9997865321
a Reaction conditions: 1 mmol 4'-bromoacetophenone, 1.3 mmol phenylboronic, 2 mmol KOH, 120 °C under microwave irradiation 2 min; b Determined by HPLC analysis.

3. Experimental

3.1. General

All chemicals were of reagent grade and used as commercially purchased without further purification. All solvents were purified according to standard procedures. Melting points were determined using a WRS-1B apparatus and were uncorrected. The IR spectra were recorded in the range of 400–4000 cm−1 with a Magna 550 FT-IR spectrometer using KBr pellets. The 1H and 13C-NMR spectra were recorded on a Bruker AV600 spectrometer. High resolution mass (HRMS) spectra were obtained in ESI mode on a Finnigan MAT95XP HRMS system (Thermo Electron Corporation). HPLC analysis of the target compound was performed on a C18 reversed-phase column (5 µm, 200 × 4.6 mm) using water-methanol (20:80, v/v) as the mobile phase at a flow rate of 0.8 mL/min. Microwave reactions were carried out in a Biotage Initiator 60.

3.2. Synthesis and Characterization

5-Hydroxy-1-pyridin-2-yl-1H-pyrazole-3-carboxylic acid methyl ester (1) [33]. Dimethyl acetylenedicarboxylate (7.5 mL, 61 mmoL) in methanol (15 mL) was added dropwise to a solution of 2-pyridylhydrazine (6.0 g, 55 mmoL) in methanol (60 mL), and the mixture was stirred for 6.0 h at 0 °C. Sodium methoxide (4.6 g, 85.1 mmoL) was then added and stirring was continued for 30 min at 70 °C. After cooling, and quenched with water. The reaction mixture was filtered and washed with cold methanol, and dried in vacuo to give a white solid. Yield: 8.9 g (77.4%). M.p.: 142.2–143.0 °C; 1H-NMR (600 MHz, CDCl3) δ 3.95 (s, 3H), 6.10 (s, 1H), 7.27–7.31 (m, 1H), 7.91–7.98 (m, 1H), 8.11 (dd, J = 5.0, 4.2 Hz, 1H), 8.31–8.36 (m, 1H), 12.83(s, 1H). ppm. 13C-NMR (151 MHz, CDCl3) δ 52.4, 90.3, 111.3, 121.6, 140.4, 144.5, 145.4, 154.2, 157.1, 162.8 ppm; HRMS (M+H+) calcd. for C10H10N3O3 220.0722, found 220.0726. IR (KBr): 3471, 3030, 2957, 1744, 1612, 1588, 1452, 1254, 775 cm−1.
5-Methoxy-1-pyridin-2-yl-1H-pyrazole-3-carboxylic acid methyl ester (2). Potassium carbonate (1.23 g, 9.91 mmoL) was added to stirred solution of 1-(pyridin-2-yl)-5-hydroxy-1H-pyrazole-3-carboxylicacid methyl ester (1, 0.65 g, 2.97 mmoL) in acetone (25 mL) at 50 °C, methyl iodide (0.4 mL) was added under argon. The mixture allowed to be stirred at 50 °C for 12 h, the reaction mixture was filtered and extracted with CHCl3 three times, after evaporating the organic layers under pressure, the residue was purified by chromatography on silica gel to give a pale white solid. Yield: 0.53 g (76.6%). M.p.: 108.7–116.6 °C; 1H-NMR (600 MHz, CDCl3) δ 3.94 (s, 3H), 4.00 (s, 3H), 6.26 (s, 1H), 7.28–7.31 (m, 1H), 7.75–7.77 (m, 1H), 7.83–7.84 (m, 1H), 8.53–8.59 (m, 1H) ppm. 13C-NMR (151 MHz, CDCl3) δ 52.2, 59.7, 88.5, 117.8, 122.8, 138.5, 143.5, 148.7, 151.0, 156.8, 162.8 ppm; HRMS (M+H+) calcd. for C11H12N3O3 234.0878, found 234.0880. IR (KBr): 3159, 3086, 2947, 1737, 1595, 1564, 1480, 1473, 1230, 1067, 754 cm−1.
5-Methoxy-1-pyridin-2-yl-1H-pyrazole-3-carboxylic acid (3). Lithium hydroxide (190 mg, 4.5 mmoL) was added to a solution of 1-(pyridin-2-yl)-5-methyl ether-1H-pyrazole -3-carboxylic acid methyl ester (2, 0.35 g, 1.5 mmoL) in methanol (18 mL) and water (3 mL),The mixture was stirred at room temperature for 12 h. After completion of the reaction, extracted with chloroform (3 × 15 mL), and evaporating the organic layers under reduced pressure, the residue recrystallised from absolute ethyl alcohol to obtain the white product. Yield: 0.24 g (73%). M.p.: 201.1–203.5 °C; 1H-NMR (600 MHz, DMSO) δ 3.96 (s, 3H), 6.33 (s, 1H), 7.46 (ddd, J = 7.5, 4.9, 0.6 Hz, 1H), 7.67 (d, J = 8.1 Hz, 1H), 8.02 (td, J = 7.8, 1.9 Hz, 1H), 8.55 (dd, J = 4.8, 1.6 Hz, 1H), 12.93 (s, 1H) ppm. 13C-NMR (151 MHz, DMSO) δ 59.6, 88.4, 118.0, 123.4, 139.1, 143.6, 148.6, 150.3, 156.3, 163.0 ppm; HRMS (M+H+) calcd. for C10H10N3O3 220.0722, found 220.0726. IR (KBr): 3444, 3134, 3114, 2997, 1701, 1604, 1574, 1484, 1439, 1242, 793 cm−1.
Palladium(II) complex 4. An aqueous solution (4 mL) of K2[PdCl4] (66.3 mg, 0.2 mmoL) was added dropwise to a stirred methanol solution (20 mL) of ligand 3 (43.7 g, 0.2 mmol) at room temperature [34]. During the addition a solid product precipitated immediately from the reaction mixture. Stirring was continued for 30 min. The stable yellow complex was filtered off, washed with cold water, and dried; 68.3 mg yield (83.3%), m.p. > 300 °C; 1H-NMR (600 MHz, DMSO) δ 3.95 (s, 3H), 6.32 (s, 1H), 7.47 (m, 1H), 7.67 (d, J = 8.2 Hz, 1H), 8.00 (m, 1H), 8.55 (d, J = 4.9 Hz, 1H), 12.93 (s, 1H) ppm. 13C-NMR (151 MHz, DMSO) δ 59.5, 88.4, 118.2, 123.4, 139.2, 143.6, 148.7, 150.2, 156.4, 163.1 ppm; HRMS (M+H+) calcd. for C10H10Cl2N3O3Pd 395.9056, found 395.9055. IR (KBr): 3430, 3110, 2936, 1664, 1609, 1584, 1511, 1487, 1317, 771, 469 cm−1.

3.3. Crystallography

A suitable crystal was obtained from an ethanol solution by slow evaporation. Diffraction experiments for 3 was carried out on with Mo Ka radiation (λ = 0.71073 Å) using a Bruker SMART APEX CCD diffractometer at 296 K. Raw frame data were integrated with the SAINT program. The structures were solved by direct methods which gave the positions of all non-hydrogen atoms and refined with full-matrix least-squares on F2 using SHELXS-97 and SHELXL-97 [35]. The hydrogen atoms were set in the calculated positions and refined by riding model. The crystallographic and refinement data of 3 is listed in Table 4.

3.4.General Procedure for Aqueous Suzuki Coupling Reactions

Into a 10 mL glass vial were placed aryl halide (1.0 mmol), phenylboronic acid (1.2 mmol), base (2 mmol), complex 4 (0.4 mg, 0.001 mmol), ethanol (1 mL), water (1 mL), and a magnetic stir bar. The vessel was sealed by capping with a Teflon septum fitted in an aluminum crimp top and placed into the microwave cavity. Microwave irradiation of 60 W was used, the temperature being ramped from room temperature to 120 °C. Once 120 °C was reached, the reaction mixture was held at this temperature for 2 min. After the mixture was allowed to cool to room temperature, the reaction vessel was opened and the contents poured into a separating funnel. Water and dichloromethane (30 mL of each) were added, and the organic material was extracted and removed. After further extraction of the aqueous layer with dichloromethane, combining of the organic washings and drying over MgSO4, the dichloromethane was removed in vacuo, leaving the crude product. The crude material was flash chromatographed on a silica gel column. All of the compounds have been characterized by comparing 1H-NMR with the values found in the literature [35,36,37,38,39,40,41].
Table 4. Crystal data and refinement details for 3.
Table 4. Crystal data and refinement details for 3.
Compound reference3
Empirical formulaC10H5N3O3
Formula weight215.17
Crystal systemMonoclinic
Space groupP2(1)/n
Unit cell dimensionsa =7.144(3) Å, α =90.00;
b = 8.093(4)Å, β = 95.6380(10);
c = 16.662(3)Å, γ = 90.00
Volume958.6(6)Å3
Z4
Density(calculated)1.491 g/cm3
Absorption coefficient0.115 mm−1
F(000)440
Crystal size0.27 × 0.25 × 0.22 mm3
Theta range for data collection (°)2.46–25.00 °
Reflections collected4745
Completeness to θmax0.993
Data/restraints/parameters1678/0/146
Goodness-of-fit on F20.917
Final R indices [I > 2σ(I)] a,bR1 = 0.0612, wR2 = 0.1673
R indices (all data)R1 = 0.0940, wR2 = 0.2057
Largest diff. peak and hole0.309 and −0.294e Å−3
a R1 = ∑||Fo|-|Fc||/∑|Fo|; b wR2 = [∑w(Fo2-Fc2)2/∑w(Fo2)2]1/2, w = [2(Fo)2 + (0.1(max(0, Fo2) + 2Fc2)/3)2]−1.

3.5. Recycling of Catalyst

Into a 10 mL glass vial were placed 4-bromoacetophenone (1.0 mmol), phenylboronic acid (1.2 mmol), KOH (2 mmol), complex 4 (0.4 mg, 0.001 mmol), ethanol (1 mL), water (1 mL), and a magnetic stir bar. The reaction mixtures were run at 120 °C for 2 min under microwave irradiation. After the mixture was allowed to cool to room temperature, the reaction vessel was opened and the contents poured into a separating funnel. Diethyl ether (5 mL × 2) was added. The aqueous phase was transferred to another glass vial for the next reaction cycle. The yields were determined by HPLC.

4. Conclusions

In summary, a new and efficient catalyst system for microwave-mediated Suzuki coupling reaction was developed. To the best of our knowledge this is the first report of pyridine-pyrazole/Pd(II) complexes for catalysis of Suzuki reaction. The reactions were performed under microwave irradiation in water/EtOH co-solvent system and have shown the application of this to the synthesis of a number of biaryls. The method has the advantage of rapid reaction times, no need for anaerobic conditions and use of a nontoxic solvent. The catalysts could be used for up to five cycles. The methodology is a good synthetic route for biaryl synthesis.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/18/2/1602/s1.

Acknowledgments

This work was supported by National Natural Science Foundation of China (No. 21062002), the Scientific Research Program of Guangxi University for Nationalities (Project No. 2010QD018).
  • Sample Availability: Samples of most compounds are available from the authors.

References and Notes

  1. Miyaura, N.; Suzuki, A. Palladium-catalyzed cross-coupling reactions of organoboron compounds. Chem. Rev. 1995, 95, 2457–2483. [Google Scholar] [CrossRef]
  2. Martin, R.; Buchwald, S.L. Palladium-catalyzed Suzuki-Miyaura cross-coupling reactions employing dialkylbiaryl phosphine ligands. Acc. Chem. Res. 2008, 41, 1461–1473. [Google Scholar] [CrossRef]
  3. Baudoin, O.; Cesario, M.; Guénard, D.; Guéritte, F. Application of the palladium-catalyzed borylation/Suzuki coupling (BSC) reaction to the synthesis of biologically active biaryl lactams. J. Org. Chem. 2002, 67, 1199–1207. [Google Scholar] [CrossRef]
  4. Markham, A.; Goa, K.L. Valsartan. A review of its pharmacology and therapeutic use in essential hypertension. Drugs 1997, 54, 299–311. [Google Scholar] [CrossRef]
  5. Tomori, H.; Fox, J.M.; Buchwald, S.L. An improved synthesis of functionalized biphenyl-based phosphine ligands. J. Org. Chem. 2000, 65, 5334–5341. [Google Scholar] [CrossRef]
  6. Lightowler, S.; Hird, M. Monodisperse aromatic oligomers of defined structure and large size through selective and sequential Suzuki palladium-catalyzed cross-coupling reactions. Chem. Mater. 2005, 17, 5538–5549. [Google Scholar] [CrossRef]
  7. Anastas, P.T.; Kirchhoff, M.M. Origins, current status, and future challenges of green chemistry. Acc. Chem. Res. 2002, 35, 686–694. [Google Scholar] [CrossRef]
  8. Mondal, M.; Bora, U. An efficient protocol for palladium-catalyzed ligand-free Suzuki–Miyaura coupling in water. Green Chem. 2012, 14, 1873–1876. [Google Scholar] [CrossRef]
  9. Inés, B.; SanMartin, R.; Moure, M.J.; Domínguez, E. Insights into the role of New palladium pincer complexes as robust and recyclable precatalysts for Suzuki-Miyaura couplings in neat water. Adv. Synth. Catal. 2009, 351, 2124–2132. [Google Scholar] [CrossRef]
  10. Casalnuovo, A.L.; Calabrese, J.C. Palladium-catalyzed alkylations in aqueous media. J. Am. Chem. Soc. 1990, 112, 4324–4330. [Google Scholar] [CrossRef]
  11. Shaughnessy, K.H. Beyond TPPTS: New approaches to the development of efficient palladium-catalyzed aqueous-phase cross-coupling reactions. Eur. J. Org. Chem. 2006, 2006, 1827–1835. [Google Scholar] [CrossRef]
  12. Meise, M.; Haag, R. A highly active water-soluble cross-coupling catalyst based on dendritic polyglycerol N-heterocyclic carbene palladium complexes. ChemSusChem 2008, 1, 637–642. [Google Scholar] [CrossRef]
  13. Kostas, I.D.; Coutsolelos, A.G.; Charalambidis, G.; Skondra, A. The first use of porphyrins as catalysts in cross-coupling reactions: A water-soluble palladium complex with a porphyrin ligand as an efficient catalyst precursor for the Suzuki-Miyaura reaction in aqueous media under aerobic conditions. Tetrahedron Lett. 2007, 48, 6688–6691. [Google Scholar]
  14. Oertel, A.M.; Ritleng, V.; Chetcuti, M.J. Synthesis and catalytic activity in suzuki coupling of nickel complexes bearing n-butyl- and triethoxysilylpropyl-substituted NHC ligands: Toward the heterogenization of molecular catalysts. Organometallics 2012, 31, 2829–2840. [Google Scholar] [CrossRef]
  15. Zhao, D.B.; Fei, Z.F.; Geldbach, T.J.; Scopelliti, R.; Dyson, P.J. Nitrile-functionalized pyridinium ionic liquids: synthesis, characterization, and their application in carbon-carbon coupling reactions. J. Am. Chem. Soc. 2004, 126, 15876–15882. [Google Scholar]
  16. Liu, P.; Yan, M.; He, R. Bis(imino)pyridine palladium(II) complexes as efficient catalysts for the Suzuki-Miyaura reaction in water. Appl. Organometal. Chem. 2010, 24, 131–134. [Google Scholar]
  17. Cobo, I.; Matheu, M.I.; Castillón, S.; Boutureira, O.; Davis, B.G. Phosphine-free Suzuki-Miyaura cross-coupling in aqueous media enables access to 2-C-aryl-glycosides. Org. Lett. 2012, 14, 1728–1731. [Google Scholar]
  18. Adam, D. Microwave chemistry: out of the kitchen. Nature 2003, 421, 571–572. [Google Scholar] [CrossRef]
  19. Marx, V. Riding the microwave. Chem. Eng. News 2004, 82, 14–19. [Google Scholar] [CrossRef]
  20. Kappe, C.O. Controlled microwave heating in modern organic synthesis. Angew. Chem. Int. Ed. 2004, 43, 6250–6284. [Google Scholar] [CrossRef]
  21. Dawood, K.M. Microwave-assisted Suzuki–Miyaura and Heck-Mizoroki cross-coupling reactions of aryl chlorides and bromides in water using stable benzothiazole-based palladium(II) precatalysts. Tetrahedron 2007, 63, 9642–9651. [Google Scholar] [CrossRef]
  22. Nguyen, H.H.; Kurth, M.J. Microwave-assisted synthesis of 3-nitroindoles from N-aryl enamines via intramolecular arene–alkene coupling. Org. Lett. 2013, 15, 362–365. [Google Scholar] [CrossRef]
  23. Chang, W.; Chae, G.; Jang, S.R.; Shin, J.; Ahn, B.J. An efficient microwave-assisted Suzuki reaction using Pd/MCM-41 and Pd/SBA-15 as catalysts in solvent-free condition. J. Ind. Eng. Chem. 2012, 18, 581–585. [Google Scholar] [CrossRef]
  24. Arvela, R.K.; Leadbeater, N.E. Suzuki coupling of aryl chlorides with phenylboronic acid in water, using microwave heating with simultaneous cooling. Org. Lett. 2005, 7, 2101–2104. [Google Scholar] [CrossRef]
  25. Mehta, V.P.; van der Eycken, E.V. Microwave-assisted C–C bond forming cross-coupling reactions: An overview. Chem. Soc. Rev. 2011, 40, 4925–4936. [Google Scholar] [CrossRef]
  26. Budzisz, E.; Lorenz, I.P.; Mayer, P.; Paneth, P.; Szatkowski, L.; Krajewska, U.; Rozalski, M.; Miernicka, M. Synthesis, crystal structure, theoretical calculation and cytotoxic effect of new Pt(II), Pd(II) and Cu(II) complexes with pyridine-pyrazoles derivatives. New J. Chem. 2008, 32, 2238–2244. [Google Scholar]
  27. Adhikari, N.; Saha, N. Synthesis and spectroscopic characterization of palladium(II) and platinum(II) complexes with substituted pyrazoles. Asian J. Chem. 2008, 20, 521–529. [Google Scholar]
  28. Ciolkowski, M.; Paneth, P.; Lorenz, I.P.; Mayer, P.; Rozalski, M.; Krajewska, U.; Budzisz, E. Tautomeric forms study of 1H-(2'-pyridyl)-3-methyl-5-hydroxypyrazole and 1H-(2'-pyridyl)-3-phenyl-5-hydroxypyrazole. Synthesis, structure, and cytotoxicactivity of their complexes with palladium(II) ions. J. Enzym. Inhib. Med. 2009, 24, 1257–1268. [Google Scholar] [CrossRef]
  29. Crystallographic data for compound 3 in this paper have been deposited with the Cambridge Crystallographic Data Centre as supplementary publication number CCDC 916185. Copies of the data can be obtained, free of charge, on application to CCDC, 12 Union Road, CB2 1EZ, UK [fax: +44(1223)336033 or e-mail: [email protected]].
  30. Ornelas, C.; Ruiz, J.; Salmon, L.; Astruca, D. Sulphonated “click” dendrimer-stabilized palladium nanoparticles as highly efficient catalysts for olefin hydrogenation and Suzuki coupling reactions under ambient conditions in aqueous media. Adv. Synth. Catal. 2008, 350, 837–845. [Google Scholar] [CrossRef]
  31. Leadbeater, N.E.; Marco, M. Ligand-free palladium catalysis of the suzuki reaction in water using microwave heating. Org. Lett. 2002, 4, 2973–2976. [Google Scholar] [CrossRef]
  32. Zhou, J.; Guo, X.M.; Tu, C.Z.; Li, X.Y.; Sun, H.J. Aqueous Suzuki coupling reaction catalyzed by water-soluble diimine/Pd(II) systems. J. Organomet. Chem. 2009, 694, 697–702. [Google Scholar] [CrossRef]
  33. Shen, L.Q.; Huang, S.Y.; Diao, K.S.; Lei, F.H. Synthesis, crystal structure, computational study of 1-(6-chloro-pyridin-2-yl) -5-hydroxy-1H-pyrazole-3-carboxylic acid methyl ester and its 5-acetoxy analogs. J. Mol. Struct. 1021, 167–173. [Google Scholar]
  34. Watson, A.A.; House, D.A.; Steel, P.J. Chiral heterocyclic ligands. VIII syntheses and complexes of new chelating ligands derived from camphor. Aust. J. Chem. 1995, 48, 1549–1572. [Google Scholar] [CrossRef]
  35. Sheldrick, G.M. A short history of SHELX. Acta Cryst. A 2008, 64, 112–122. [Google Scholar] [CrossRef]
  36. Zhou, W.-J.; Wang, K.-H.; Wang, J.-X. Atom-efficient, palladium-catalyzed Stille coupling reactions of tetraphenylstannane with aryl iodides or aryl bromides in polyethylene glycol 400(PEG-400). Adv. Synth. Catal. 2009, 351, 1378–1382. [Google Scholar] [CrossRef]
  37. Zhou, W.-J.; Wang, K.-H.; Wang, J.-X. Pd(PPh3)4-PEG 400 catalyzed protocol for the atom-efficient Stille cross-coupling reaction of organotin with aryl bromides. J. Org. Chem. 2009, 74, 5599–5602. [Google Scholar] [CrossRef]
  38. Cahiez, G.; Chaboche, C.; Mahuteau-Betzer, F.; Ahr, M. Iron-catalyzed homo-coupling of simple and functionalized arylmagnesium reagents. Org. Lett. 2005, 7, 1943–1946. [Google Scholar] [CrossRef]
  39. Percec, V.; Golding, G.M.; Smidrkal, J.; Weichold, O. NiCl2(dppe)-catalyzed cross-coupling of aryl mesylates, arenesulfonates, and halides with arylboronic acids. J. Org. Chem. 2004, 69, 3447–3452. [Google Scholar] [CrossRef]
  40. Firouzabadi, H.; Iranpoor, N.; Gholinejad, M. 2-Aminophenyl diphenylphosphinite as an easily accessible ligand for heterogeneous palladium-catalyze SuzukieMiyura reaction in water in the absence of any organic co-solvent. J. Organomet. Chem. 2010, 695, 2093–2097. [Google Scholar] [CrossRef]
  41. Hanhan, M.E.; Senemoglu, Y. Microwave-assisted aqueous Suzuki coupling reactions catalyzed by ionic palladium(II) complexes. Transit. Metal. Chem. 2012, 37, 109–116. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Shen, L.; Huang, S.; Nie, Y.; Lei, F. An Efficient Microwave-Assisted Suzuki Reaction using a New Pyridine-Pyrazole/Pd(II) Species as Catalyst in Aqueous Media. Molecules 2013, 18, 1602-1612. https://doi.org/10.3390/molecules18021602

AMA Style

Shen L, Huang S, Nie Y, Lei F. An Efficient Microwave-Assisted Suzuki Reaction using a New Pyridine-Pyrazole/Pd(II) Species as Catalyst in Aqueous Media. Molecules. 2013; 18(2):1602-1612. https://doi.org/10.3390/molecules18021602

Chicago/Turabian Style

Shen, Liqun, Suyu Huang, Yuanmei Nie, and Fuhou Lei. 2013. "An Efficient Microwave-Assisted Suzuki Reaction using a New Pyridine-Pyrazole/Pd(II) Species as Catalyst in Aqueous Media" Molecules 18, no. 2: 1602-1612. https://doi.org/10.3390/molecules18021602

Article Metrics

Back to TopTop