Next Article in Journal
Biotransformation of Ginsenoside Rf to Rh1 by Recombinant β-Glucosidase
Previous Article in Journal
Pentacyclic Triterpene Distribution in Various Plants – Rich Sources for a New Group of Multi-Potent Plant Extracts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Synthesis of Some Perhydrobenzimidazolinium Salts and Their Application in Pd-Carbene Catalyzed Heck and Suzuki Reactions

Department of Chemistry, Faculty of Science and Art, Adıyaman University, 02040 Adıyaman, Turkey
Molecules 2009, 14(6), 2032-2042; https://doi.org/10.3390/molecules14062032
Submission received: 13 May 2009 / Revised: 27 May 2009 / Accepted: 3 June 2009 / Published: 5 June 2009

Abstract

:
Novel 1,3-dialkylperhydrobenzimidazolinium chloride salts were prepared as precursors of N-heterocyclic carbenes 3a-e by reacting N,N’-dialkylcyclohexandiamine, triethyl orthoformate and ammonium chloride. The salts were characterized spectroscopically and the complexes formed in situ from Pd(OAc)2 and 3 have been tested as catalysts in homogenous Heck and Suzuki reactions.

Graphical Abstract

1. Introduction

Palladium-catalyzed C-C bond formation is one of the most fundemental and important reactions in modern synthetic chemistry [1,2]. It represents the key the step in a wide range of preparative organic processes, from the synthesis of natural products [3] to supramolecular chemistry and material science [4]. Common methodologies used are the palladium mediated coupling of the organic halides or halide equvalents with Grignard reagents, organotin or organoboron reagents where monodentate phosphine are usually employed as ancillary ligands [5,6,7,8]. Recently, N-heterocyclic carbenes (NHCs) have been used as alternative or better ancillary ligands than phosphine. The major advantage of NHCs in catalysis is that an increased stability of the active catalyst is often observed, allowing for lower operational catalyst loadings, prolonged reaction times and tolerance of increased reaction temperatures [9,10,11]. The Heck and Suzuki couplings are fascinating reactions from a catalysis science perspective. Virtually all forms of palladium can be used as precatalysts for the simpler reactions, yet specifically designed catalysts are required for activation of bulky or electronically unactivated substrates. Additionally, in many cases, extremely small amounts of palladium (ppm or ppb levels) are sufficient to give very high turnover frequencies, whereas in others, 10% or higher precatalyst loadings are required to obtain adequate product yields [12]. Initially, water-soluble phosphines were used as ligands for the cross-coupling reactions in aqueous media [13], but in recent years, other hydrophilic phosphine-free systems [14] and soluble palladium nanoparticles [15,16,17] have also been found to be higly efficient catalysts for this transformation. The use of water as a solvent for chemical reactions clearly has both economical and environmental advantages because it is inexpensive, abundant, nontoxic, nonflammable and readily separable from organic compounds [18]. There have been a number of reports of the palladium-mediated Heck and Suzuki reaction being performed using water as solvent [19,20,21,22,23,24,25,26,27]. Due to the large number of Suzuki and other coupling reactions that are carried out in aqueous [19] or biphasic systems, there has been an increased interest in the development of water-soluble ligands for these reactions. Shaughnessy and co-workers utilized both sterically demanding water-soluble alkylphosphines [28] and triarylphosphines [29] in Suzuki and Heck couplings of aryl bromides, respectively. We have previously reported the use of the various catalytic systems, such as palladium catalyzed cross-coupling, ruthenium catalyzed hydrogenation and rhodium catalyzed arylation or hydrosilylation [30,31,32,33,34]. Although the nature of the NHC ligand on complexes has a tremendous influence on the rate of catalyzed reactions, the use of perhydrobenzimidazolinium ligands in Heck and Suzuki reactions is a neglected field. In order to find more efficient catalyst, we have prepared a series of new perhydrobenzimidazolinium chlorides, 3a-e (Scheme 1), containing a saturated benzimidazole ring and we report here an in situ palladium-carbene based catalytic system for Heck and Suzuki cross-coupling of aryl halides.

2. Results and Discussion

The general route to the p-substituent containing ligand precursors is shown in Scheme 1. The condensation of 1,2-diaminocyclohexane with aldehydes in ethanol gave the corresponding Schiff bases in high yields. Reduction of these Schiff bases with NaBH4 in methanol leads to N,N’-dialkyl-cyclohexan-1,2-diamines. The symmetrical 1,3-dialkylperhydrobenzimidazolinium salts 3 were easly synthesized in high yields from the N,N’-dialkylcyclohexan-1,2-diamines, triethyl ortoformate and ammonium chloride. After purification, the 1,3-dialkylperhydrobenzimidazolinium salts 3a-e were obtained in good yields (85-90%). The salts are soluble in the common polar solvents ethanol and dichloromethane and are stable under air and in the presence of moisture. The structures of 3 were determined by their spectroscopic data and elemental analyses (see Experimental section). The 13C- NMR spectra of 3a-e show only sharp high field signals [159.5 ppm (3a), 162.1 ppm (3b), 161.9 ppm (3c), 162.5 ppm (3d), 162.0 ppm (3e)] for the imino carbon and for a benzylic carbon [58.7 ppm (3a), 66.7 ppm (3b), 66.6 ppm (3c), 67.0 ppm (3d), 66.8 ppm(3e)]. The 1H-NMR spectra of the perhydro-benzimidazolinium salts further supported the assigned structures. The resonances of the C(2)-H were observed as sharp singlets at δ= 10.87, 10.96, 10.74, 11.10, 10.79 ppm for 3a-e, respectively. The benzylic protons appeared as a doublet at 4.36- 5.50 ppm. The IR data for perhydrobenzimidazolinium 3a-e salts clearly indicate the presence of the –C=N- group with a ν(C=N) vibration at 1,644, 1,592, 1,612, 1,587 and 1,600 cm-1 for 3a-e, respectively. These NMR and IR values were similar to those reproted for other 1,3-dialkylperhydrobenzimidazolinium and 1,3-dialkylimidazolinium salts [33,35].
Scheme 1. Synthesis of 1,3-dialkylperhydrobenzimidazolinium salts.
Scheme 1. Synthesis of 1,3-dialkylperhydrobenzimidazolinium salts.
Molecules 14 02032 g001
Reagents and conditions: (i) EtOH, 3h, reflux; (ii) MeOH, NaBH4, 5h, RT; (iii) HC(OEt)3, NH4CI, 12h, 110 °C.
The Pd-catalyzed Heck reaction is an efficient way to prepare styrene derivatives, which are important chemicals for many applications. The Heck couplings of styrene with aryl bromides to form alkenes were undertaken with Pd(OAc)2/3a-e as catalysts. The choice of bases and solvents are usually important in achieving an efficient cross coupling reaction. The widely used solvents such as DMF, DMSO, toluene, dioxane, THF and CH3CN had remarkable effects on the coupling reaction. For optimal reaction conditions, the Pd(OAc)2-catalyzed cross coupling of bromobenzene with styrene was employed as the model reaction using ligand 3a at 80 oC; Cs2CO3, K2CO3, K3PO4 and t-BuOK were tested as the base. The coupling reactions of aryl bromides and styrene were carried out in DMF/H2O (3:3 mL) with 1 mol% Pd(OAc)2, 2 mol% 3 and 2 equiv. K2CO3 for 2 h at 80oC. The results are summarized in Table 1.
Control experiments indicate that the coupling reactions did not occur in the absence of 3a. Under these reaction conditions, a wide range of aryl bromides bearing electron-donating or electron-withdrawing groups react with styrene affording the coupled products in excellent yields (Table 1, entries 3, 4, 8, 15, 18 and 22). Enhancements in activity, although less significant, are also observed employing 4-bromobenzaldehyde instead of 4-bromoacetophenone (entries 1–5 and 11-15, respectively).
Table 1. The Heck coupling reaction of aryl bromides with styrene. Molecules 14 02032 i001
Table 1. The Heck coupling reaction of aryl bromides with styrene. Molecules 14 02032 i001
EntryRSALTYield (%)b,c
1COCH33a96
2COCH33b95
3COCH33c98
4COCH33d98
5COCH33e96
6CHO3a95
7CHO3b94
8CHO3c96
9CHO3d91
10CHO3e97
11H3a94
12H3b93
13H3c90
14H3d92
15H3e95
16OCH33a92
17OCH33b88
18OCH33c93
19OCH33d91
20OCH33e87
21CH33a86
22CH33b89
23CH33c85
24CH33d81
25CH33e84
aReaction conditions: R-C6H4Br-p 1.0 mmol, styrene 1.5 mmol, K2CO3 2.0 mmol, Pd(OAc)2 1% (molar ratio), 3a-e 2% (molar ratio), water (3 mL)/DMF (3 mL), 80°C, 2 h. bIsolated yields are based on aryl bromide. cAll reactions were monitored by GC, and the compound purity was checked by NMR.
However, chloroarenes do not react under standard conditions, and yields are typically <5%. A systematic study on the substituent effect in the perhydrobenzimidazolinium salts 3a-e indicated that the introduction of a p-isopropylbenzyl substituent on the N-atoms notably increased the yield of product (Table 1, entries 3, 8, 13, 18, 23). Comparatively the perhydrobenzimidazolinium salts displayed better performance than benzimidazolinium salts in the Heck and Suzuki reactions [36,37].
Due to the palladium/imidazolium salt catalyst systems display high reactivity with aryl chlorides in the Suzuki cross-coupling reaction has been applied successfully to the synthesis of fenbuten and to a key intermediate in the synthesis of sartans [38]. The Suzuki coupling of phenylboronic acid with aryl chlorides to form biaryls were undertaken with Pd(OAc)2 /3a-e as catalysts.
Table 2. The Suzuki coupling reaction of aryl chlorides with phenylboronic acid. Molecules 14 02032 i002
Table 2. The Suzuki coupling reaction of aryl chlorides with phenylboronic acid. Molecules 14 02032 i002
EntryRSALTYield (%)b,c
1COCH33a94
2COCH33b95
3COCH33c97
4COCH33d98
5COCH33e92
6CHO3a93
7CHO3b97
8CHO3c98
9CHO3d94
10CHO3e91
11H3a87
12H3b89
13H3c86
14H3d94
15H3e90
16OCH33a84
17OCH33b86
18OCH33c88
19OCH33d89
20OCH33e82
21CH33a78
22CH33b75
23CH33c81
24CH33d77
25CH33e74
aReaction conditions: R-C6H4CI-p 1.0 mmol, phenylboronic acid 1.5 mmol, K2CO3 2.0 mmol, Pd(OAc)2 1% (molar ratio), 3a-e 2% (molar ratio), water (3 mL)/DMF (3 mL), 80°C, 1h. bIsolated yields are based on aryl chloride. cAll reactions were monitored by GC, and the compound purity was checked by NMR.
Similar reaction conditions were employed for the Suzuki reactions. The coupling reactions of aryl chlorides and phenylboronic acid were carried out in DMF/H2O (3:3 mL) with 1 mol% Pd(OAc)2, 2 mol% 3 and 2 equiv. K2CO3 for 1 h at 80oC. We started our investigation examining the coupling of 4-chloroacetophenone and phenylboronic in the presence of Pd(OAc)2 /3. The results are summarized in Table 2. It can be show these salts are an effective ligand precursor for the coupling of unactivated, activated and deactivated chlorides. These results are in agreement with other reports [38,39,40,41,42,43].

3. Conclusions

In conclusion, we have synthesized five 1,3-dialkylperhydrobenzimidazolinium chloride salts and have investigated their catalytic activity in the Heck and Suzuki coupling reactions. The procedure is simple and efficient toward various types of aryl halides and does not require an induction period. The advandage of the catalyst is that it has low-loading capabilities, and it is usable in air. Detailed investigations, focusing on imidazolidin-2-ylidene and benzimidazolin-2-ylidene substituent effects, functional group tolerance, and catalytic activity in this and other coupling reactions are ongoing.

4. Experimental

4.1. General

All reactions for the preparation of 1,3-dialkylperhydrobenzimidazolinium salts 3a-e were carried out under argon using standard Schlenk-type flasks. Heck and Suzuki coupling reactions were carried out in air. 1,2-Diaminocyclohexane (mixture of cis and trans), aldehydes and other reagents were purchased from Aldrich Chemical Co. (Turkey). All 1H- and 13C-NMR were recorded in CDCI3 using a Bruker AC300P FT spectrometer operating at 300.13 MHz (1H) or 75.47 MHz (13C). Chemical shifts (δ) are given in ppm relative to TMS, coupling constants (J) in hertz. FT-IR spectra were recorded as KBr pellets in the range 400-4000 cm-1 on a Mattson 1000 spectrophotometer (wavenumbers, cm-1). GC were measured by GC-FID on a Agilent 6890N gas chromatograph equipped with an HP-5 column of 30 m length, 0.32 mm diameter and 0.25 μm film thickness. Melting points were measured in open capillary tubes with an Electrothermal-9200 melting point apparatus and uncorrected. Elemental analyses were performed at TUBITAK (Ankara, Turkey) Microlab.

4.2. General procedure for preparation of the 1,3-dialkylperhydrobenzimidazolinium chlorides 3a-e

A mixture of N,N’-dialkylcyclohexan-1,2-diamine (6.80 mmol), NH4CI (6.80 mmol) and triethyl orthoformate (5 ml) was heated for 12 h at 110oC. Upon cooling to room temperature, colorless crystals of 3a-e were obtained. The crystals were filtered, washed with diethyl ether (3 x 15 mL) and dried under vacuum. The crude product was recrystallized from EtOH/Et2O.
1,3-bis(4-Methylbenzyl)perhydrobenzimidazolinium chloride (3a). Yield: 2.27 g, 90%; mp: 238oC; IR (cm-1) ν = 1,644 (-CH=N-); 1H-NMR (CDCI3) δ: 1.32-1.41 (m, 4H, NCHCH2CH2CH2CH2CHN), 1.66-1.69 (m, 4H, NCHCH2CH2CH2CH2CHN), 3.71-3.80 (m, 2H, NCHCH2CH2CH2CH2CHN), 2.29 (s, 6H, CH2C6H4CH3-p), 4.38 and 5.19 (d, 4H, J=11.1 Hz, CH2C6H4CH3-p), 7.13 and 7.24 (d, 8H, J=6.0 Hz, CH2C6H4CH3-p), 10.87 (s, 1H, 2-CH); 13C-NMR (CDCI3) δ: 21.3 (NCHCH2CH2CH2CH2CHN), 22.8 (NCHCH2CH2CH2CH2CHN), 49.8 (NCHCH2CH2CH2CH2CHN), 18.3 (CH2C6H4CH3-p), 58.7 (CH2C6H4CH3-p), 128.7, 128.8, 130.0, 139.0 (CH2C6H4CH3-p), 159.6 (2-CH); Anal. Calcd. for C23H29N2CI: C, 74.89; H, 7.86; N, 7.59. Found C, 74.62; H, 7.71; N, 7.67.
1,3-bis(4-Ethylbenzyl)perhydrobenzimidazolinium chloride (3b). Yield: 2.41 g; 85%, mp: 146oC; IR (cm-1) ν = 1,592 (-CH=N-); 1H-NMR (CDCI3) δ: 1.11-1.48 (m, 4H, NCHCH2CH2CH2CH2CHN), 1.73-2.09 (m, 4H, NCHCH2CH2CH2CH2CHN), 3.10-3.22 (m, 2H, NCHCH2CH2CH2CH2CHN), 1.21 (t, 6H, J=7.8 Hz, CH2C6H4CH2CH3-p), 2.64 (q, 4H, J=7.5 Hz, CH2C6H4CH2CH3-p), 4.41, 5.05 and 4.68, 5.26 (d, 4H, J=14.7 Hz, and J=15.0 Hz, CH2C6H4CH2CH3-p), 7.19 and 7.32 (d, 8H, J=7.8 Hz, CH2C6H4CH2CH3-p), 10.96 (s, 1H, 2-CH); 13C-NMR (CDCI3) δ: 23.5 (NCHCH2CH2CH2CH2CHN), 28.5 (NCHCH2CH2CH2CH2CHN), 50.6 (NCHCH2CH2CH2CH2CHN), 15.4 (CH2C6H4CH2CH3-p), 22.6 (CH2C6H4CH2CH3-p), 66.7 (CH2C6H4CH2CH3-p), 128.6, 130.1, 145.1, 159.4 (CH2C6H4CH2CH3-p), 162.1 (2-CH); Anal. Calcd. for C25H33N2CI: C, 75.66; H, 8.32; N, 7.06. Found C, 75.24; H, 8.54; N, 7.19.
1,3-bis(4-Isopropylbenzyl)perhydrobenzimidazolinium chloride (3c). Yield: 2.15 g; 87%, mp: 133-135oC; IR (cm-1) ν = 1,612 (-CH=N-); 1H-NMR (CDCI3) δ: 1.09-1.33 (m, 4H, NCHCH2CH2CH2CH2CHN), 1.78-2.04 (m, 4H, NCHCH2CH2CH2CH2CHN), 3.15-3.18 (m, 2H, NCHCH2CH2CH2CH2CHN), 1.20 (d, 12H, J=9 Hz, CH2C6H4CH(CH3)2-p), 2.85 (hept, 2H, J=6.9 Hz, CH2C6H4CH(CH3)2-p), 4.66 and 5.50 (d, 4H, J=15 Hz, CH2C6H4CH(CH3)2-p), 7.18 and 7.29 (d, 8H, J=8.1 Hz, CH2C6H4CH(CH3)2-p), 10.74 (s, 1H, 2-CH); 13C-NMR (CDCI3) δ: 23.8 (NCHCH2CH2CH2CH2CHN), 27.3 (NCHCH2CH2CH2CH2CHN), 50.5 (NCHCH2CH2CH2CH2CHN), 18.4 (CH2C6H4CH(CH3)2-p), 33.7 (CH2C6H4CH(CH3)2-p), 66.6 (CH2C6H4CH(CH3)2-p), 127.2, 128.5, 130.3, 149.5 (CH2C6H4CH(CH3)2-p), 161.9 (2-CH); Anal. Calcd. for C27H37N2CI: C, 76.32; H, 8.71; N, 6.59. Found C, 76.81; H, 8.93; N, 6.80.
1,3-bis(4-Biphenylbenzyl)perhydrobenzimidazolinium chloride (3d). Yield: 2.37 g; 86%, mp: 263-265oC; IR (cm-1) ν = 1,587 (-CH=N-); 1H-NMR (CDCI3) δ:1.15-1.28 (m, 4H, NCHCH2CH2CH2CH2CHN), 1.83-2.13 (m, 4H, NCHCH2CH2CH2CH2CHN), 3.26-3.29 (m, 2H, NCHCH2CH2CH2CH2CHN), 4.56, 5.14 and 4.85, 5.37 (d, 4H, J=15 Hz and J=14.7 Hz CH2C6H4C6H5), 7.33-7.62 (m, 18H, CH2C6H4C6H5), 11.10 (s, 1H, 2-CH); 13C-NMR (CDCI3) δ: 23.6 (NCHCH2CH2CH2CH2CHN), 27.5 (NCHCH2CH2CH2CH2CHN), 50.6 (NCHCH2CH2CH2CH2CHN), 67.0 (CH2C6H4C6H5), 127.0, 127.8, 128.8, 129.2, 132.0, 140.1, 141.6, 159.5 (CH2C6H4C6H5), 162.5 (2-CH); Anal. Calcd. for C33H33N2CI: C, 80.40; H, 6.70; N, 5.68. Found C, 80.21; H, 6.86; N, 5.54.
1,3-bis(4-Ethoxybenzyl)perhydrobenzimidazolinium chloride (3e). Yield: 1.84 g; 82%, mp: 226oC; IR (cm-1) ν = 1,600 (-CH=N-); 1H-NMR (CDCI3) δ: 1.05-1.20 (m, 4H, NCHCH2CH2CH2CH2CHN), 1.70-2.00 (m, 4H, NCHCH2CH2CH2CH2CHN), 3.05-3.07 (m, 2H, NCHCH2CH2CH2CH2CHN), 1.23 (t, 6H, J=9 Hz, CH2C6H4OCH2CH3-p), 3.91 (q, 4H, J=6.4 Hz, CH2C6H4OCH2CH3-p), 4.56 and 4.92 (d, 4H, J=10.8 Hz, CH2C6H4OCH2CH3-p), 6.78 and 7.26 (d, 8H, J =8.7 Hz, CH2C6H4OCH2CH3-p), 10.79 (s, 1H, 2-CH); 13C-NMR (CDCI3) δ: 23.7 (NCHCH2CH2CH2CH2CHN), 27.8 (NCHCH2CH2CH2CH2CHN), 50.5 (NCHCH2CH2CH2CH2CHN), 14.9 (CH2C6H4OCH2CH3-p), 63.6 (CH2C6H4OCH2CH3-p), 66.8 (CH2C6H4OCH2CH3-p), 115.2, 124.9, 130.1, 159.4 (CH2C6H4OCH2CH3-p), 162.0 (2-CH); Anal. Calcd. for C25H33N2O2CI: C, 70.01; H, 7.70; N, 6.53. Found C, 70.18; H, 7.54; N, 6.71.

4.3. General procedure for the Heck coupling reactions

Pd(OAc)2 (1.0 mmol%), the appropriate 1,3-dialkylperhydrobenzimidazolinium salt 3a-e (2 mmol%), aryl bromide (1.0 mmol), styrene (1.5 mmol), K2CO3 (2 mmol), water (3 mL) and DMF (3 mL) were added to a small Schlenk tube and the mixture was heated at 80 C for 2 h. At the conclusion of the reaction, the mixture was cooled, extracted with ethyl acetate–hexane (1 : 5), filtered through a pad of silica gel with copious washing, concentrated and purified by flash chromatography on silica gel. All reactions were monitored by GC. The purity of the compounds was checked by NMR and yields are based on aryl bromide.

4.4. General procedure for the Suzuki coupling reactions

Pd(OAc)2 (1.0 mmol%), the appropriate 1,3-dialkylperhydrobenzimidazolinium salt 3a-e (2 mmol%), aryl chloride (1.0 mmol), phenylboronic acid (1.5 mmol), K2CO3 (2 mmol), water (3 mL) and DMF (3 mL) were added to a small Schlenk tube and the mixture was heated at 80 C for 1 h. At the conclusion of the reaction, the mixture was cooled, extracted with ethyl acetate–hexane (1 : 5), filtered through a pad of silica gel with copious washing, concentrated and purified by flash chromatography on silica gel. All reactions were monitored by GC. The purity of the compounds was checked by NMR and yields are based on aryl chloride.

Acknowledgements

Financial support from Inonu University Research Fund (BAP 2007/29) is gratefully acknowledged.

References

  1. Trost, B.M.; Verhoeven, T.R. Organopalladium Compounds in Organic Synthesis and in Catalysis. In Comprehensive Organometallic Chemistry; Wilkinson, G., Stone, F.G.A., Abel, E.W., Eds.; Pergamon: Oxford, UK, 1982; Volume 8, pp. 799–938. [Google Scholar]
  2. Miyaura, N.; Suzuki, A. Palladium-Catalyzed Cross-Coupling Reactions of Organoboron Compounds. Chem. Rev. 1995, 95, 2457–2483. [Google Scholar] [CrossRef]
  3. Tietze, L.F.; Petersen, S. Stereoselective Synthesis of Novel 19-Nor-Steroids by a Double Heck Reaction. Eur. J. Org. Chem. 2001, 9, 1619–1624. [Google Scholar] [CrossRef]
  4. Cianga, I.; Yağci, Y. Synthesis and Characterization of Comb-Like Polyphenylenes via Suzuki Coupling of Polystyrene Macromonomers Prepared by Atom Transfer Radical Polymerization. Eur. Polymer J. 2002, 38, 695–703. [Google Scholar] [CrossRef]
  5. Busacca, C.A.; Eriksson, M.C.; Fiaschi, R. Cross Coupling of Vinyl Triflates and Alkyl Grignard Reagents Catalyzed by Nickel(0)-Complexes. Tetrahedron Lett. 1999, 40, 3101–3104. [Google Scholar] [CrossRef]
  6. Miller, J.A.; Farrell, R.P. Synthesis of Functionally Substituted Unsymmetrical Biaryls via a Novel Double Metal Catalyzed Coupling Reaction. Tetrahedron Lett. 1998, 39, 7275–7278. [Google Scholar] [CrossRef]
  7. Suzuki, A. Cross-Coupling Reactions via Organoboranes. J. Organomet. Chem. 2002, 653, 83–90. [Google Scholar] [CrossRef]
  8. Xu, L.; Chen, W.; Xiao, J. Palladium Catalyzed Regioselective Arylation of Electron-Rich Olefins by Aryl Halides. J. Mol. Catal. A 2012, 187, 189–193. [Google Scholar]
  9. Jafarpour, L.; Nolan, S.P. Transition-Metal Systems Bearing a NucleophilicCarbene Ancillary Ligand: From Thermochemistry to Catalysis. Adv. Organomet. Chem. 2000, 46, 181–222. [Google Scholar]
  10. Herrmann, W.A. N-Heterocyclic Carbenes: A New Concept in Organometallic Catalysis. Angew. Chem. Int. Ed. 2002, 41, 1290–1309. [Google Scholar] [CrossRef]
  11. Peris, E.; Crabtree, R.H. Recent Homogeneous Catalytic Applications of Chelate and Pincer N-Heterocyclic Carbenes. Coord. Chem. Rev. 2004, 248, 2239–2246. [Google Scholar] [CrossRef]
  12. Phan, N.T.S.; Sluys, M.V.; Jones, C.W. On the Nature of the Active Species in Palladium Catalyzed Mizoroki-Heck and Suzuki-Miyaura Couplings - Homogeneous or Heterogeneous Catalysis, A Critical Review. Adv. Synth. Catal. 2006, 348, 609–679. [Google Scholar] [CrossRef]
  13. Casalnuova, A.L.; Calabrese, J.C. Palladium-Catalyzed Alkylations in Aqueous Media. J. Am. Chem. Soc. 1990, 112, 4324–4330. [Google Scholar] [CrossRef]
  14. Churruca, F.; SanMartin, R.; İnes, B.; tellitu, I.; Dominquez, E. Hydrophilic CNC-Pincer Palladium Complexes: A Source for Highyl Efficient, Recyclable Homogeneous Catalysts in Suzuki-Miyaura Cross-Coupling. Adv. Synth. Catal. 2006, 348, 1836–1840. [Google Scholar] [CrossRef]
  15. Alimardanov, A.; Schmieder-Van de Vondervoort, L.; de Vries, A.H.M.; de Vries, J.G. Use of “Homeopathic” Ligand-Free Palladium as Catalyst for Aryl-Aryl Coupling Reactions. Adv. Synth. Catal. 2004, 346, 1812–1817. [Google Scholar] [CrossRef]
  16. de Vries, J.G. A Unifying Mechanism for all High-Temperature Heck Reactions. The Role of Palladium Colloids and Anionic Species. Dalton Trans. 2006, 421–429. [Google Scholar] [CrossRef]
  17. Saha, D.; Chattopadhyay, K.; Ranu, B.C. Aerobic Ligand-Free Suzuki Coupling Catalyzed by in Situ-Generated Palladium Nanoparticles in Water. Tetrahedron Lett. 2009, 50, 1003–1006. [Google Scholar] [CrossRef]
  18. Li, C.J.; Chan, T.H. Organic Reactions in Aqueous Media; John Wiley & Sons: New York, 1997. [Google Scholar]
  19. Genet, J.P.; Savignac, M. Recent Developments of Palladium(0) Catalyzed Reactions in Aqueous Medium. J. Organomet. Chem. 1999, 576, 305–317. [Google Scholar] [CrossRef]
  20. Sakurai, H.; Tsukuda, T.; Hirao, T. Pd/C as a Reusable Catalyst for the Coupling Reaction of Halophenols and Arylboronic Acids in Aqueous Media. J. Org. Chem. 2002, 67, 2721–2722. [Google Scholar] [CrossRef]
  21. Parisot, S.; Kolodziuk, R.; Henry, C.G.; Iourtchenko, A.; Sinou, D. Glucosamine-Based Phosphines. Efficient Ligands in the Suzuki Cross-Coupling Reaction in Water. Tetrahedron Lett. 2002, 43, 7397–7400. [Google Scholar] [CrossRef]
  22. Schönfelder, D.; Nuyken, O.; Weberskirch, R. Heck and Suzuki Coupling Reactions in Water Using Poly(2-Oxazoline)s Functionalized with Palladium Carbene Complexes as Soluble, Amphiphilic Polymer Supports. J. Organomet. Chem. 2005, 690, 4648–4655. [Google Scholar] [CrossRef]
  23. Zhou, J.; Guo, X.; Tu, C.; Li, X.; Sun, H. Aqueous Suzuki Coupling Reaction Catalyzed by Water-Soluble Diimine/Pd(II) Systems. J. Organomet. Chem. 2009, 694, 697–702. [Google Scholar] [CrossRef]
  24. Ma, J.; Cui, X.; Gao, L.; Wu, Y. FerrocenylimidazolinePalladacycles: Syntheses, Crystal Structures and Applications as Catalysts for Suzuki Cross Coupling Reaction in Water. Inorg. Chem. Commun. 2007, 10, 762–766. [Google Scholar] [CrossRef]
  25. Shang, Y.J.; Wu, J.W.; Fan, C.L.; Hu, J.S.; Lu, B.Y. Synthesis of 1,3-Bis-(5-Ferrocenylisoxazole-3-yl)Benzene-Derived Palladium(II)acetate Complex and its Application in Mizoroki–Heck Reaction in an Aqueous Solution. J. Organomet. Chem. 2008, 693, 2963–2966. [Google Scholar] [CrossRef]
  26. Vashchenko, V.; Krivoshey, A.; Knyazeva, I.; Petrenko, A.; Goodby, J.W. Palladium-Catalyzed Suzuki Cross-Coupling Reactions in a Microemulsion. Tetrahedron Lett. 2008, 49, 1445–1449. [Google Scholar] [CrossRef]
  27. Kostas, I.D.; Coutsolelos, A.G.; Charalambidis, G.; Skondrab, A. The First Use of Porphyrins as Catalysts in Cross-Coupling Reactions: A Water-Soluble Palladium Complex with a PorphyrinLigand as an Efficient Catalyst Precursor for the Suzuki-Miyaura Reaction in Aqueous Media under Aerobic Conditions. Tetrahedron Lett. 2007, 48, 6688–6691. [Google Scholar] [CrossRef]
  28. Shaughnessy, K.H.; Booth, R.S. Sterically Demanding, Water-Soluble Alkylphosphines as Ligands for High Activity Suzuki Coupling of Aryl Bromides in Aqueous Solvents. Org. Lett. 2001, 3, 2757–2759. [Google Scholar] [CrossRef]
  29. Moore, L.R.; Shaughnessy, K.H. Efficient Aqueous-Phase Heck and Suzuki Couplings of Aryl Bromides Using Tri(4,6-dimethyl-3- sulfonatophenyl)phosphine Trisodium Salt (TXPTS). Org. Lett. 2004, 6, 225–228. [Google Scholar] [CrossRef]
  30. Özdemir, İ.; Yiğit, M.; Çetinkaya, E.; Çetinkaya, B. Synthesis of Novel Palladium N-Heterocyclic-Carbene Complexes as Catalysts for Heck and Suzuki Cross-Coupling Reactions. Appl. Organometal. Chem. 2006, 20, 187–192. [Google Scholar] [CrossRef]
  31. Yiğit, M.; Yiğit, B.; Özdemir, İ.; Çetinkaya, E.; Çetinkaya, B. Active Ruthenium-(N-Heterocyclic Carbene) Complexes for Hydrogenation of Ketones. Appl. Organometal. Chem. 2006, 20, 322–327. [Google Scholar] [CrossRef]
  32. Yiğit, M.; Özdemir, İ.; Çetinkaya, E.; Çetinkaya, B. Novel Rhodium N-Heterocyclic Carbene Catalysed Arylation of Aldehydes with Phenylboronic acid. Trans. Met. Chem. 2007, 32, 536–540. [Google Scholar] [CrossRef]
  33. Yiğit, M.; Özdemir, İ.; Çetinkaya, E.; Çetinkaya, B. In Situ Generated Rhodium-Based Catalyst for Addition of Phenylboronic Acid to Aldehydes. Heteroatom Chem. 2005, 16, 461–465. [Google Scholar] [CrossRef]
  34. Yiğit, M.; Özdemir, İ.; Çetinkaya, B.; Çetinkaya, E. Novel N-heterocyclic-Carbene-Rhodium Complexes as Hydrosilylation Catalysts. J. Mol. Catal. A 2005, 241, 88–92. [Google Scholar] [CrossRef]
  35. Yaşar, S.; Özdemir, İ.; Çetinkaya, B.; Renaud , J.L.; Bruneau, C. Benzylic İmidazolidinium, 3,4,5,6-Tetrahydropyrimidinium and Benzimidazolium Salts: Applications in Ruthenium-Catalyzed Allylic Substitution Reactions. Eur. J. Org. Chem. 2008, 2142–2149. [Google Scholar]
  36. Huang, W.; Guo, J.; Xiao, Y.; Zhu, M.; Zou, G.; Tang, J. Palladium-Benzimidazolium Salt Catalyst Systems for Suzuki Coupling: Development of a Practical and Highly Active Palladium Catalyst System for Coupling of Aromatic Halides with Arylboronic Acids. Tetrahedron 2005, 61, 9783–9790. [Google Scholar] [CrossRef]
  37. Özdemir, İ.; Gök, Y.; Gürbüz, N.; Çetinkaya, B. Benzimidazolylidene Carbene Ligated Palladium Catalysis of the Heck Reaction in Aqueous Media. Turk. J. Chem. 2007, 31, 397–402. [Google Scholar]
  38. Grasa, G.A.; Viciu, M.S.; Huang, J.; Zhang, C.; Trudell, M.L.; Nolan, S.P. Suzuki-Miyaura Cross-Coupling Reactions Mediated by Palladium/İmidazolium Salt Systems. Organometallics 2002, 21, 2866. [Google Scholar] [CrossRef]
  39. Ozdemir, İ.; Demir, S.; Çetinkaya, B. Palladium-Catalyzed Heck Reaction of Aryl Bromides in Aqueous Media Using Tris(N-Heterocyclic Carbene) Ligands. Synlett. 2007, 6, 889–892. [Google Scholar] [CrossRef]
  40. Loch, J.A.; Albrecht, M.; Peris, E.; Mata, J.; Faller, J.W.; Crabtree, R.H. Palladium Complexes with Tridentate Pincer Bis-CarbeneLigands as Efficient Catalysts for C−C Coupling. Organometallics 2002, 21, 700–706. [Google Scholar] [CrossRef]
  41. Baillie, C.; Xiao, J.L. Palladium-Catalysed Synthesis of BiarylPhosphines. Tetrahedron 2004, 60, 4159–4168. [Google Scholar] [CrossRef]
  42. Zhang, C.M.; Huang, J.K.; Trudell, M.L.; Nolan, S.P. Palladium−Imidazol-2-ylidene Complexes as Catalysts for Facile and Efficient Suzuki Cross-Coupling Reactions of Aryl Chlorides with Arylboronic Acids. J. Org. Chem. 1999, 64, 3804–3805. [Google Scholar] [CrossRef]
  43. Altenhoff, G.; Goddart, R.; Lehmann, C.W.; Glorius, F. Sterically Demanding, Bioxazoline-Derived N-Heterocyclic CarbeneLigands with Restricted Flexibility for Catalysis. J. Am. Chem. Soc. 2004, 126, 15195–15201. [Google Scholar] [CrossRef]
  • Sample Availability: Samples of the compounds are available from the authors.

Share and Cite

MDPI and ACS Style

Yiğit, M. The Synthesis of Some Perhydrobenzimidazolinium Salts and Their Application in Pd-Carbene Catalyzed Heck and Suzuki Reactions. Molecules 2009, 14, 2032-2042. https://doi.org/10.3390/molecules14062032

AMA Style

Yiğit M. The Synthesis of Some Perhydrobenzimidazolinium Salts and Their Application in Pd-Carbene Catalyzed Heck and Suzuki Reactions. Molecules. 2009; 14(6):2032-2042. https://doi.org/10.3390/molecules14062032

Chicago/Turabian Style

Yiğit, Murat. 2009. "The Synthesis of Some Perhydrobenzimidazolinium Salts and Their Application in Pd-Carbene Catalyzed Heck and Suzuki Reactions" Molecules 14, no. 6: 2032-2042. https://doi.org/10.3390/molecules14062032

Article Metrics

Back to TopTop