Next Article in Journal
Synthesis and Field Evaluation of the Sex Pheromone Analogues to Soybean Pod Borer Leguminivora glycinivorella
Previous Article in Journal
Acetylated and Methylated β-Cyclodextrins as Viable Soluble Supports for the Synthesis of Short 2′-Oligodeoxyribo-nucleotides in Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Suzuki-Miyaura Reactions Catalyzed by C2-Symmetric Pd-Multi-Dentate N-Heterocyclic Carbene Complexes

1
State Key Laboratory of Fine Chemicals, Dalian University of Technology, Dalian 116024, China
2
College of Environmental and Chemical Engineering, Dalian University, Dalian 116622, China
*
Authors to whom correspondence should be addressed.
Molecules 2012, 17(10), 12121-12139; https://doi.org/10.3390/molecules171012121
Submission received: 20 September 2012 / Revised: 6 October 2012 / Accepted: 9 October 2012 / Published: 16 October 2012

Abstract

:
Suzuki-Miyaura coupling reactions are promoted by Pd complexes ligated with C2-symmetric multi-dentate N-heterocyclic carbenes derived in situ from Pd(OAc)2 and imidazolium salts. Good to excellent yields were obtained for aryl bromides as substrates. Turnover numbers of up to 105 could be achieved with 5 × 10−4 mol% of Pd(OAc)2/1 × 10−3 mol% NHC precatalyst in 24 h.

1. Introduction

The formation of C-C bonds catalyzed by transition metal complexes represents one of the most powerful tools in organic synthesis [1,2,3] and has found important applications in the synthesis of organic molecules such as pharmaceuticals [4], natural products [5], and polymers [6]. The Suzuki-Miyaura reaction, the C-C cross-coupling reaction between aryl halides and arylboronic acids, is an example of this kind of reaction [7,8,9]. As is well known, many Pd-phosphine complexes have been employed as the catalysts for this transformation [10,11], but most phosphine ligands, especially those displaying good catalytic properties, are expensive, toxic and air-sensitive. Accordingly, Pd-complexes overcoming these limitations are highly desirable. N-Heterocyclic carbenes (NHCs) have received increasing attention after the isolation of free NHC by Arduengo and co-workers in 1991 [12], and Herrmann’s seminal work demonstrating the catalysis of coupling reactions using Pd-NHC complexes [13]. The excellent σ-donor and lower π-acceptor characteristics of NHC, in combination with their good stability towards air and moisture, make them attractive as ligands in catalytic reactions [14,15,16]. Furthermore, substituents attached to the NHC framework could be easily modulated to tune their electronic as well as the steric properties. These results have shown NHCs to be an alternative for conventional phosphine ligands in homogeneous catalysis including olefin metathesis [17], hydrosilylation [18,19], hydrogenation [20,21], C-C coupling reactions, etc. [22].
Suzuki-Miyaura coupling reactions promoted by Pd-NHC complexes were initiated by Herrmann’s report in 1998 [23]. Since then, Pd-NHC complexes have been found as efficient catalysts for this kind of coupling reaction [24,25,26,27,28,29,30], and new NHCs or their precursors have been synthesized to confer more efficient catalytic properties or ease of operation [31,32,33,34]. Among these catalysts, most of them are derived from monodentated NHCs, and some are bidentate anionic ligands [26,27]. The chelating NHC ligands predominantly consist of two NHC moieties linked by a chain or a ring, e.g., Shi’s cis-chelating, bidentate NHC derived from binaphthyl-2,2′-diamine [35], and more importantly, hybrid NHC ligands. Typical examples of the hybrid NHC ligands used in palladium-catalyzed reactions include NHC,P chelating ligands derived from 1 and 2, and NHC,N chelating ligands in complexes 3, 4 and 5 (Figure 1) [36,37,38,39]. Nonetheless, little attention has been paid to hybrid NHC chelating ligands bearing weakly-coordinating O-atom as a potential coordination atom until now [40,41,42].
Figure 1. Typical Pd-NHC complexes and NHC precursors with a (potentially) coordinative heteroatom.
Figure 1. Typical Pd-NHC complexes and NHC precursors with a (potentially) coordinative heteroatom.
Molecules 17 12121 g001
The fact that Pd complexes containing both NHC and phosphine ligands, show higher activity than those with phosphine as the only ligand or Pd(NHC)2X2, was reasoned to be due to the strong Pd-NHC bond and relatively weak Pd-P bond, which results in the stabilization of Pd, easy dissociation of phosphine ligand and favorable oxidative addition of aryl halides to Pd. In metal-NHC complexes promoted reactions, metal complexes with both NHC and phosphine structural motif generally exhibit better activity and selectivity than those only with monodentate NHC ligands. Therefore, a second, relatively weak coordination atom in the chelating metal-NHC complex might be beneficial for improving the catalytic performance.
We have an interest in the applications of multidentate, C2-symmetric NHCs with one chelating carbene and two O- atoms as ligands in catalytic reactions, for which there have been few precedents. We envisioned that the coordinative ability of the heteroatom can be adjusted by changing its bonding. Previously, we have synthesized tridentate C2-symmetric imidazolinium salts 6 (Figure 2) and their NHC precursors, which have two oxygen atoms in the arms, and found that their copper complexes could catalyze the asymmetric conjugate addition of Et2Zn to cycloalkenones efficiently [43]. To continue this work, we designed 7, the unsaturated counterpart of 6, to explore their applications in Suzuki-Miyaura coupling reactions, as it has been reported that Pd coordinated with unsaturated NHCs show higher reactivity than those with saturated NHCs [44]. Herein, we wish to report the preparation and catalytic properties of the novel C2-symmetric tridentate imidazolium salts 7 in the Suzuki-Miyaura coupling reaction.
Figure 2. The structure of tridentate C2-symmetric imidazol(in)ium salts 6 and 7.
Figure 2. The structure of tridentate C2-symmetric imidazol(in)ium salts 6 and 7.
Molecules 17 12121 g002

2. Results and Discussion

The synthesis of the NHC precursor imidazolium salts 7a–f in 22-30% overall yield is shown in Scheme 1. The synthetic route to the NHC precursor imidazolium salts 7a–f started from (S)-ethyl lactate or (S)-ethyl mandelate by etherification with phenol or substituted phenols. The α-aryloxycarboxylates 8a–f were next reduced, yielding β-aryloxyalcohols 9a–f. Then the alcohol were converted into halides 11a–f, and reacted with imidazole giving imidazolium salts 7a–f, which contain β-aryloxyl groups in the side chains. They are quite stable towards air and moisture.
Scheme 1. Synthesis of the NHC precursor imidazolium salts 7a–f.
Scheme 1. Synthesis of the NHC precursor imidazolium salts 7a–f.
Molecules 17 12121 g005
The imidazolium salts 7c–f synthesized from (S)-ethyl lactate were optically pure. This can be deduced from the fact that only one set of NMR signals was observed in both 1H-NMR and 13C-NMR spectra of these compounds. Reaction of racemic bromides 11c–f with imidazole would yield racemic C2-symmetric (R,R), (S,S)-isomers of 7c–f, and Cs-symmetric meso-(R,S)-isomer of 7c–f. Even though the (R,R), (S,S)-isomers could not be distinguished by NMR, the signals of the racemic isomers and meso-isomer should be different. Therefore, the sets of NMR signals of 7 could be used to judge whether the products are optically pure.
On the other hand, use of (S)-ethyl mandelate yielded 7a and 7b as a mixture of diastereomers, since 1H-NMR and 13C-NMR signals for two sets or more than one set were observed. This indicates that racemization occurs using (S)-ethyl mandelate as the starting material, whose α-H is more acidic than that in ethyl lactate.
The catalytic activities of the Pd-complexes of the NHCs derived from imidazolium salts 7 in Suzuki-Miyaura coupling reactions were screened, using catalysts generated in situ from a mixture in a solvent of 7, Pd(OAc)2, and a base.
The reaction conditions were optimized on a model reaction between phenyl bromide and phenylboronic acid. First, the effect of solvents was examined using the catalyst generated from imidazolium salt 7a and Pd(OAc)2. The product biphenyl 12a was obtained in 95% yield in either toluene or DMF. Addition of H2O to DMF (V/V = 1:2) accelerated the reaction but no increase in the yield was observed. The reaction was completed in 1.5 h using toluene as the solvent. Moderate to low yields were obtained when polar solvents, including 1,4-dioxane, THF and EtOH, were used, even though 1,4-dioxane and EtOH have been reported in the literature as good solvents for this kind of reaction.
Using toluene as the solvent, bases were then optimized, and the results are shown in Table 1. Although Cs2CO3 is generally efficient for the Suzuki-Miyaura coupling reactions catalyzed by Pd-NHC or palladium-phosphine complexes, in our case, only a 32% yield of 12a was observed, whereas the use of K2CO3 resulted in an excellent yield (entry 2). With other bases like NaOH, KF and K3PO4, the reaction also proceeded well, but somewhat more slowly. Then the temperature was optimized. Phenyl bromide was quantitatively converted to 12a in 0.5 h at 110 °C, much faster than at 90 °C. At lower temperature, the reaction is very slow. Only a 53% yield of 12a was obtained after 21 h at 70 °C. Therefore, subsequent reactions were carried out using K2CO3 as the base and toluene as the solvent at 110 °C.
Table 1. Influence of the base onthe Suzuki-Miyaura reaction catalyzed by Pd(OAc)2/7a a.
Table 1. Influence of the base onthe Suzuki-Miyaura reaction catalyzed by Pd(OAc)2/7a a.
Molecules 17 12121 i001
EntryBaseTime (h)Yield b(%)
1Na2CO32412
2K2CO31.595
3Cs2CO32432
4NaOH2486
5KF2475
6K3PO42482
7CH3COONa244
a Reaction conditions: 0.5 mmol phenyl bromide, 0.75 mmol phenylboronic acid, 1.5 mmol base, 0.5 mol% Pd(OAc)2, 1 mol% 7a, 3.0 mL toluene, 90 °C. b GC yield.
Next, the catalytic abilities of 7a analogues were evaluated and the results are shown in Table 2. The most efficient catalyst was derived from 7a (entry 1). A comparable yield was obtained with the methyl substituted analogue 7c (entry 3). Introduction of a nitro group to the aryloxy substituent results in a ligand-Pd complex generating a lower yield of the coupling product (entry 4), which shows that reduction of the electron density of the phenyl ring and the resultant decrease in the coordination ability of the O-atom reduced the catalytic ability of 7. The presence of a 4-methyl group in the aryloxy substituent also led to a lower yield (entry 5), but higher than that using 7d. Yields of 86–87% were generated when the (substituted)-phenoxy substituent was changed to a 2-naphthoxy group (entries 2 and 6), which is more steric demanding and resulted in the O-atom being less coordinative. Two typical imidazolium salts IMes·HCl (IMes = 1,3-bis(2,4,6-trimethylphenyl)-imidazol-2-ylidene) and IPr·HCl (IPr = 1,3-bis(2,6-diisopropylphenyl)-imidazol-2-ylidene) catalyze the coupling reaction giving inferior results (entries 7, 8). In the absence of imidazolium salt or other ligand, a 72% yield of 12a was obtained in 1h using Pd(OAc)2 as the catalyst (entry 9). These results demonstrate that the ligand derived from imidazolium salts 7 which contains two potentially chelated oxygen atoms is more efficient than simple NHC ligands, e.g., IPr and IMes in the coupling reactions. This might be due to the presence of a NHC moiety and oxygen atoms in the hybrid NHC ligands. The strong coordinating of NHC can prevent the dissociate of ligand from palladium complex, and therefore, stablize the catalytically active palladium species. In the meanwhile, the weak coordinative oxygen atom(s) can easily dissociate from palladium complex and generate a coordination-unsaturated palladium species. This favors the oxidation of aryl halides or arylboronic acids toward the coordination-unsaturated palladium species and catalytic cycle. Indeed, formation of palladium balck, which is catalytic inert to the Suzuki reaction and an indication of ligands dissociate from palladium completely, is rarely observed in the coupling reaction using 7. The coupling reaction of phenylboric acid with phenyl bromide could be performed in air with a slightly reduced yield (78%).
Table 2. Evaluation of the NHCsfrom 7a–f in the Suzuki-Miyaura reaction a.
Table 2. Evaluation of the NHCsfrom 7a–f in the Suzuki-Miyaura reaction a.
Molecules 17 12121 i002
EntryImidazolium saltTime (h)Yield b(%)
17a0.599
27b187
37c196
47d185
57e190
67f186
7IPr·HCl175
8IMes·HCl170
9-172
a Reaction conditions: 0.5 mmol phenyl bromide, 0.75 mmol phenylboronic acid, 1.5 mmol K2CO3, 0.5 mol% Pd(OAc)2, 1 mol% 7, 3.0 mL toluene, 110 °C. b GC yield.
The reactions of various aryl halides with phenylboronic acids were then investigated under the optimized conditions, and the results are summarized in Table 3. Only 5 minutes were needed for the complete reactions of 4-nitrophenyl bromide and 4-acetophenyl bromide, which are electron-deficient and more reactive in this kind of reaction.
Table 3. Suzuki-Miyaura reaction of aryl halides (benzyl halides) with phenylboronic acid catalyzed by Pd(OAc)2/7a.
Table 3. Suzuki-Miyaura reaction of aryl halides (benzyl halides) with phenylboronic acid catalyzed by Pd(OAc)2/7a.
Molecules 17 12121 i003
EntryArX or BnXRTimeProductYield a (%)
1 Molecules 17 12121 i004H5 min12b95
2 Molecules 17 12121 i005H5 min12c98
3 Molecules 17 12121 i006H1 h12d98
4 Molecules 17 12121 i007H0.5 h12e99
5 Molecules 17 12121 i008H0.5 h12f96
6 Molecules 17 12121 i009H1 h12g93
7 Molecules 17 12121 i010H0.5 h12h95
8 Molecules 17 12121 i011H1 h12i92
9 Molecules 17 12121 i012H0.5 h12j91
10 Molecules 17 12121 i013H1 h12k85
11 Molecules 17 12121 i014H1 h12l90
12 Molecules 17 12121 i015H1 h12m96
13 Molecules 17 12121 i016H1 h12n97
14 Molecules 17 12121 i017H3 h12o90
15 Molecules 17 12121 i018H24 h12a35
16 Molecules 17 12121 i019H24 h12b70
17 Molecules 17 12121 i020H10 h12o71
18 Molecules 17 12121 i0213-Me0.5 h12p93
19 Molecules 17 12121 i0224-Me0.5 h12j85
20 Molecules 17 12121 i0232-Me2.5 h12k93
21 Molecules 17 12121 i0244-MeO2C24 h12q86
22 Molecules 17 12121 i0252-Me1 h12r21 b
23 Molecules 17 12121 i0262-Me11.5 h12r56 b,c
a Isolated yield except noted. b GC yield. c Using 0.05 mol% Pd(OAc)2, 0.1 mol% imidazolium salt 7a.
For those aryl halides with electron-donating groups (entries 7, 10, 13) or sterically hindered aryl bromides (entry 11), longer reaction times were needed. Generally, almost quantitative yields of coupling products could be obtained in 1 h. It is worth to note that 4-hydroxylphenyl bromide, which is very electron-rich in the phenyl ring and the C-Br bond, is highly reactive to coupling. As expected, a wide range of functional groups including keto, nitro, cyano, ester, amide, hydroxy and ether, were tolerated. The coupling reactions could be extended to benzyl bromide which contains a sp3-C atom bearing the halide, and therefore, a C-C coupling product between a sp3-C and a sp2-C was achieved (entry 14). The reaction of 4-chlorophenyl bromide gave 4-chlorobiphenyl as the only coupling product in high yield, and demonstrated the inertness of C-Cl in the presence of C-Br (entry 6). When phenyl chloride was used, only a 35% yield of biphenyl was obtained (entry 15). Introduction of an acetyl group to phenyl chloride led to an increase in the activity, and a 70% yield of the coupling product was achieved (entry 16). Gratifyingly, benzyl chloride also gave a reasonable yield of coupling product (entry 17), although in lower yield than that obtained with benzyl bromide. Palladium black was only observed unremarkably in reactions performed at 110 °C and after long reaction times.
In addition, reactions of various arylboronic acids with substituted phenyl bromides were performed. The reactions proceed smoothly in high yield using 2-tolylboronic acid. In contrast to aryl halides, the presence of an electron-withdrawing group results in a low coupling yield which may be attributed to the tendency of hydrolysis of the electron deficient arylboronic acids (Table 3, entry 21) [45]. Even though the presence of an ortho-Me group in either the aryl halide or the arylboronic acid did not affect the coupling yield, it was still difficult to obtain tri-ortho substituted biphenyl in high yields, a challenging reaction in literature, due to the large hindrance for our catalyst system. In the case of a very sterically hindered reaction (entry 22), the less bulky 2,2′-dimethylbiphenyl (the homocoupling product of two arylboronic acid molecules), was formed in preference to the more bulky 2,6,2′-trimethylbiphenyl product derived from the expected Suzuki-Miyaura coupling when 0.5 mol% of Pd catalyst was employed. This indicated the relative inertness of the 2,6-dimethylphenyl bromide. Therefore, the amount of the Pd salt was decreased to 0.05 mol% to depress the relative faster coupling reaction between the aryl groups in the boronic acid. In this case, the yield of the trimethylbiphenyl increased to 56% as judged by gas chromatography over a prolonged reaction time (entry 23). Similar results were observed for the reaction of 4-bromoacetophenone with phenylboronic acid (entries 1, 2 in Table 4 vs. entry 1 in Table 3).
Table 4. The activities of Pd(OAc)2/7a with variation of loading.
Table 4. The activities of Pd(OAc)2/7a with variation of loading.
Molecules 17 12121 i027
EntryRPd (mol%)Time(h)ProductYield a (%)TON
1COMe0.05112b1002000
2COMe0.005412b10020000
3COMe0.00052412b50100000
4H0.05212a951900
5H0.005212a387600
6Me0.051.512j1002000
7Me0.005412j9519000
a GC yield using diethylene glycol di-n-butyl ether as an internal standard.
The efficiencies of Pd(OAc)2/7a were further tested with different catalyst loadings, and the results are summarized in Table 4. 4-Acetylphenyl bromide could be completely converted to the coupling product using 5 × 10−2 mol% in 1 h and 5 × 10−3 mol% of Pd catalyst in 4 h, respectively (entries 1 and 2). The turnover numbers (TON) reached 2 × 104. Furthermore, a maximum TON of 105 was obtained at 5 × 10−4 mol% of catalyst loading in 24 h (entry 3), and the turnover frequency (TOF ) is ca. 4.17 × 103 h−1. A TON of up to 7,600 was recorded for the reaction of phenyl bromide (entry 5). Interestingly, for the less active 4-bromotoluene, TON and TOF values up to 1.9 × 104 and 4.75 × 103 h−1, respectively, could be achieved with 5 × 10−3 mol% Pd catalyst in 4 h (entry 7), which was better than the case of phenyl bromide. Details of the progress of reactions with different amount of catalyst and substrates are given in Figure 3 and Figure 4.
Figure 3. Reaction profiles for Suzuki-Miyaura reaction between 4-acetylphenyl bromide and phenylboronic acid catalyzed by different Pd/7a loading: 0.05 mol% (■), 0.005 mol% (○), 0.0005 mol% (▲).
Figure 3. Reaction profiles for Suzuki-Miyaura reaction between 4-acetylphenyl bromide and phenylboronic acid catalyzed by different Pd/7a loading: 0.05 mol% (■), 0.005 mol% (○), 0.0005 mol% (▲).
Molecules 17 12121 g003
Figure 4. Reaction profiles for Suzuki-Miyaura reaction of 4-methylphenyl bromide catalyzed by Pd/7a 0.05 mol% (●), 0.005 mol% (○), and phenyl bromide catalyzed by Pd/7a 0.05 mol% (▲), 0.0005 mol% (△).
Figure 4. Reaction profiles for Suzuki-Miyaura reaction of 4-methylphenyl bromide catalyzed by Pd/7a 0.05 mol% (●), 0.005 mol% (○), and phenyl bromide catalyzed by Pd/7a 0.05 mol% (▲), 0.0005 mol% (△).
Molecules 17 12121 g004

3. Experimental

3.1. General

All chemicals were purchased from Alfa Aesar Co., Ltd. (Tianjin, China) and Accela ChemBio Co., Ltd. (Shanghai, China), except arylboronic acids which were products of Ally Chemical Ltd. (Dalian, China). The solvents were freshly distilled prior to use. NMR spectra were recorded on a Varian 400 MHz spectrometer or on a Bruker DRX500 spectrometer, using TMS as an internal standard. IR spectra were recorded on a Nicolet 550 spectrometer. MS spectra were measured on a Hewlett-Packard HP-6890/5973 gas chromatography-mass spectrometer. HRMS were recorded on a Micromass UPLC/Q-Tof Micro spectrometer. The reaction mixtures were analyzed by gas chromatography (Shimadzu GC-2010, capillary column SE-54, 30 m × 0.32 mm × 4 μm; FID detector; N2 gas). Column chromatography was performed with silica gel (200–300 mesh).

3.2. Synthesis of Imidazolium Salts 7a–f (Exemplified by the Synthesis of 7f)

3.2.1. Ethyl 2-(2-Naphthoxyl)propanoate (8f)

β-Naphthol (1.586 g, 11.00 mmol), PPh3 (3.148 g, 12.00 mmol) and dichloromethane (10 mL) were added to a round-bottom flask, and the mixture was stirred for 5 min. (S)-Ethyl lactate (1.182 g, 10.00 mmol) was added to the flask, then the flask was cooled with an ice-water bath. DEAD (1.742 g, 10.00 mmol) was added slowly. The mixture was stirred for 30 min at 0 °C, and then 6 h at room temperature. After removal of the colorless precipitate by filtration, the filtrate was concentrated in vacuo. 8f (2.230 g, 91.2% yield) was isolated by column chromatography as colorless oil. Rf = 0.57 (EA: PE = 10: 90), Molecules 17 12121 i028 = +12.1 (c 1.46, EtOH). 1H-NMR (400 MHz, CDCl3) δ 7.72 (d, J = 9.0 Hz, 2H, Ar-H), 7.67 (d, J = 8.2 Hz, 1H, Ar-H), 7.40 (ddd, J = 8.2, 6.9, 1.2 Hz, 1H, Ar-H), 7.31 (ddd, J = 8.1, 6.9, 1.2 Hz, 1H, Ar-H), 7.19 (dd, J = 8.9, 2.6 Hz, 1H, Ar-H), 7.04 (d, J = 2.5 Hz, 1H, Ar-H), 4.87 (q, J = 6.8 Hz, 1H, CH), 4.27–4.13 (m, 2H, CH2), 1.66 (d, J = 6.8 Hz, 3H, CH3CH), 1.21 (t, J = 7.1 Hz, 3H, CH3CH2). 13C-NMR (101 MHz, CDCl3) δ 171.99 (C=O), 155.44, 134.21, 129.55, 129.22, 127.52, 126.74, 126.31, 123.85, 118.78, 107.60 (Ar-C), 72.51 (CH), 61.10 (CH2), 18.40 (CH3CH), 13.99 (CH3CH2).

3.2.2. 2-(2-Naphthoxyl)propan-1-ol (9f)

To an ice-water cooled round-bottom flask containing a solution of 8f (1.243 g, 5.09 mmol) in THF (10.0 mL) was added NaBH4 (0.380 g, 10 mmol) in portions, and the mixture was stirred for 30 min at 0 °C. Then the temperature was recovered to room temperature slowly, and the mixture was stirred at room temperature for 6 h. The volatiles were removed by evaporation in vacuo. Water was added to the residue, and the aqueous phase was extracted with dichloromethane. The combined organic phase was dried with Na2SO4 and purified by column chromatography, yielding 9f (0.85 g, 84.2% yield) as colorless oily liquid. Rf = 0.14 (EA:PE = 10:90), Molecules 17 12121 i029 = −14.5 (c 1.11, EtOH). 1H-NMR (400 MHz, CDCl3) δ 7.75–7.61 (m, 3H, Ar-H), 7.39 (dd, J = 8.1, 7.0 Hz, 1H, Ar-H), 7.30 (dd, J = 8.0, 7.0 Hz, 1H, Ar-H), 7.15 (d, J = 2.1 Hz, 1H, Ar-H), 7.11 (dd, J = 8.9, 2.4 Hz, 1H, Ar-H), 4.61–4.51 (m, 1H, CH), 3.77–3.69 (m, 2H, CH2), 2.88 (s, 1H, OH), 1.25 (d, J = 6.2 Hz, 3H, CH3). 13C-NMR (101 MHz, CDCl3) δ 155.50, 134.51, 129.61, 129.13, 127.64, 126.78, 126.42, 123.83, 119.51, 108.98 (Ar-C), 74.75 (CH), 66.11 (CH2), 15.73 (CH3).

3.2.3. 2-(2-Naphthoxyl)propyl 4-methylbenzenesulfonate (10f)

To a round-bottom flask, 9f (0.850g, 3.93 mmol) and dichloromethane (15 mL) was added, followed by tolylsulfonyl chloride (0.964 g, 5.06 mmol) and pyridine (0.396 g, 5.01 mmol). The mixture was stirred at room temperature for 24 h, and quenched by the addition of water. After phases separation and general workup, oily 10f (1.104g, 73.8% yield) was isolated. Rf = 0.28 (EA:PE = 20:80), Molecules 17 12121 i030 = +70.27 (c 1.04, CHCl3). 1H-NMR (400 MHz, CDCl3) δ 7.73 (t, J = 7.1 Hz, 3H, Ar-H), 7.65 (dd, J = 13.7, 8.6 Hz, 2H, Ar-H), 7.41 (t, J = 7.5 Hz, 1H, Ar-H), 7.31 (t, J = 7.5 Hz, 1H, Ar-H), 7.19 (d, J = 8.0 Hz, 2H, Ar-H), 7.01 (s, 1H, Ar-H), 6.98 (dd, J = 8.9, 2.4 Hz, 1H, Ar-H), 4.72–4.63 (m, 1H, CH), 4.20 (dd, J = 10.6, 5,8 Hz, 1H, CH-H), 4.14 (dd, J = 10.6, 4.4 Hz, 1H, CH-H), 2.31 (s, 3H, Ph-CH3), 1.33 (d, J = 6.3 Hz, 3H, CH3CH). 13C-NMR (101 MHz, CDCl3) δ 154.95, 144.93, 134.36, 132.82, 129.84, 129.58, 129.21, 127.90, 127.63, 126.81, 126.47, 123.98, 119.34, 108.76 (Ar-C), 71.74 (CH), 71.23 (CH2), 21.55 (Ph-CH3), 16.44 (CH3CH).

3.2.4. 1-Bromo-2-(2-naphthoxyl)propane (11f)

Compound 10f (1.10 g, 2.97 mmol), acetone (15 mL), LiBr (0.348 g, 4.01 mmol) and Li2CO3 (0.030 g, 0.45 mmol) were added to a flask in sequence. The mixture was refluxed for 48 h. After cooled to room temperature, the solvent was removed by evaporation in vacuo. The residue was separated by column chromatography, affording oily 11f (0.654 g, 82.6% yield). Rf = 0.69 (EA:PE = 20:80), Molecules 17 12121 i031 = −41.86 (c 1.01, CHCl3). 1H-NMR (400 MHz, CDCl3) δ 7.76 (dd, J = 8.3, 2.9 Hz, 2H, Ar-H), 7.72 (d, J = 8.3 Hz, 1H, Ar-H), 7.44 (ddd, J = 8.2, 6.9, 1.3 Hz, 1H, Ar-H), 7.35 (ddd, J = 8.1, 6.9, 1.2 Hz, 1H, Ar-H), 7.20–7.12 (m, 2H, Ar-H), 4.77–4.66 (m, 1H, CH), 3.62 (dd, J = 10.4, 4.8 Hz, 1H, CH-H), 3.48 (dd, J = 10.4, 6.2 Hz, 1H, CH-H), 1.52 (d, J = 6.1 Hz, 3H, CH3). 13C-NMR (101 MHz, CDCl3) δ 155.16, 134.50, 129.83, 129.38, 127.74, 126.89, 126.56, 124.08, 119.60, 109.17 (Ar-C), 73.50 (CH), 35.43 (CH2), 18.84 (CH3).

3.2.5. 1,3-Bis-[2-(2-naphthoxy)propyl]imidazolium Bromide (7f)

To a flask, 11f (0.653 g, 2.46 mmol), DMF (10 mL), imidazole (0.0816 g, 1.20 mmol) and K2CO3 (0.0552 g, 0.40 mmol) were added. The mixture was refluxed for 48 h, and quenched by the addition of water after cooled to room temperature. The separated aqueous phase was extracted with dichloromethane (3 × 10 mL). The combined organic phase was dried and concentrated. Oily 1,3-bis-[2-(2-naphthoxy)propyl]imidazolium bromide (7f) (0.400 g, 64.4%) was isolated by column chromatography. Overall yield 30.1%. Yellow oil. Molecules 17 12121 i032 = +41.2 (c 0.566, CHCl3). IR: υC=N 1602 cm−1, υC-O 1179 cm−1. 1H-NMR (400 MHz, CDCl3) δ 10.57 (s, 1H, N-CH=N), 7.71–7.61 (m, 6H, Ar-H), 7.50 (s, 2H, N-CH), 7.46–7.38 (m, 2H, Ar-H), 7.34 (dd, J = 11.0, 4.0 Hz, 2H, Ar-H), 7.13 (d, J = 2.2 Hz, 2H, Ar-H), 7.00 (dd, J = 8.9, 2.5 Hz, 2H, Ar-H), 5.01–4.96 (m, 2H, O-CH), 4.95 (dd, J = 15.0, 2.3 Hz, 2H, NCH-H), 4.44 (dd, J = 14.3, 8.1 Hz, 2H, NCH-H), 1.41 (d, J = 6.2 Hz, 6H, CH3). 13C-NMR (101 MHz, CDCl3) δ 153.80 (N-CH=N), 137.18 (N-CH=CH), 133.86, 129.50, 128.90, 127.23, 126.50, 126.24, 123.85, 122.79, 118.53, 108.88 (Ar-C), 72.07 (O-CH), 53.70 (CH2), 16.25 (CH3). HRMS: m/z 437.2233 (calculated 437.2229 for C29H29N2O2).

3.2.6. Compounds 7a–e were Similarly Prepared and Characterized

1,3-Bis-(2-phenoxy-2-phenylethyl)imidazolium Bromide (7a). Total yield 23.6%. Yellow solid. IR: υC=N 1602 cm−1, υC-O 1172 cm−1. 1H-NMR (500 MHz, CDCl3) δ 10.78 (s, 1H, N-CH=N), 10.77 (s, 1H, N-CH′=N), 7.51–7.45 (m, 8H, Ar-H), 7. 38 (d, J = 1.5 Hz, 2H, N-CH), 7.34 (d, J = 1.5 Hz, 2H, N-CH’), 7.33–7.27 (m, 12H, Ar-H), 7.18–7.10 (m, 8H, Ar-H), 6.92–6.86 (m, 4H, Ar-H), 6.78–6.77 (m, 8H, Ar-H), 5.72 (dd, J = 8.6, 2.8 Hz, 1H, O-CH), 5.64 (dd, J = 8.4, 2.8 Hz, 1H, O-CH), 4.93 (dd, J = 14.3, 2.9 Hz, 1H, NCH-H), 4.92 (dd, J = 14.2, 2.9 Hz, 1H, NCH-H′), 4.51 (dd, J = 14.3, 8.2 Hz, 2H, NCH-H), 4.49 (dd, J = 14.2, 8.4 Hz, 2H, NCH-H′). 13C-NMR (101 MHz, CDCl3) δ 156.53 (N-CH=N), 156.49 (N-C′H=N), 138.98 (N-CH=CH), 138.78 (N-C′H=CH), 135.85, 135.72, 129.65, 129.64, 129.21, 129.05, 129.02, 126.29, 126.17, 122.55, 122.50, 122.02, 122.00, 115.82, 115.72 (Ar-C), 78.03 (O-CH), 77.91 (O-C′H), 55.51 (CH2). HRMS: m/z 461.2222 (calcd. 461.2229 for C31H29N2O2). Two sets of NMR signals indicated the racemization of 7a.
1,3-Bis-[2-(2-naphthoxy)-2-phenylethyl]imidazolium Bromide(7b). Total yield 28.4%. Yellow oil. Molecules 17 12121 i033 = −9.80 (c 0.21, CHCl3). IR: υC=N 1602 cm−1, υC-O 1179 cm−1. 1H-NMR (400 MHz, CDCl3) δ 10.29 (s, 1H, N-CH=N), 7.61–7.50 (m, 5H, Ar-H), 7.50–7.38 (m, 7H, Ar-H, N-CH), 7.30–7.11 (m, 10H, Ar-H), 7.06 (ddd, J = 16.6, 9.0, 2.1 Hz, 2H, Ar-H), 6.95 (s, 2H, Ar-H), 5.87 (d, J = 7.9 Hz, 1H, NCH-H), 5.80 (d, J = 7.5 Hz, 1H, NCH-H), 4.97–4.77 (m, 2H, O-CH), 4.52 (d, J = 13.7, 8.2 Hz, 1H, NCH-H), 4.50 (d, J = 13.0, 8.4 Hz, 1H, NCH-H). 13C-NMR (101 MHz, CDCl3) δ 154.16 (N-CH=N), 137.81 (N-CH=CH), 135.72, 135.61, 133.85, 129.48, 129.03, 128.90, 128.73, 127.37, 126.77, 126.28, 126.16, 124.01, 122.72, 118.44, 118.36, 109.45 (Ar-C), 77.47 (O-CH), 55.04 (CH2). HRMS: m/z 561.2527 (calcd. 561.2542 for C39H33N2O2). More than one set of NMR signals indicated the racemization of 7b.
1,3-Bis-(2-phenoxypropyl)imidazolium Bromide (7c). Total yield 21.8%. Yellow oil. Molecules 17 12121 i034 = −20.55 (c 0.69, CHCl3). IR: υC=N 1602 cm−1, υC-O 1172 cm−1. 1H-NMR (400 MHz, CDCl3) δ 10.26 (s, 1H, N-CH=N), 7.63 (s, 2H, N-CH), 7.21 (t, J = 7.7 Hz, 4H, Ph-H), 6.94 (t, J = 7.2 Hz, 2H, Ph-H), 6.82 (d, J = 8.1 Hz, 4H, Ph-H), 4.89 (d, J = 14.0 Hz, 2H, NCH-H), 4.85–4.79 (m, 2H, O-CH), 4.44 (dd, J = 14.0, 7.8 Hz, 2H, NCH-H), 1.35 (d, J = 6.1 Hz, 6H, CH3). 13C-NMR (101 MHz, CDCl3) δ 156.07 (N-CH=N), 137.16 (N-CH=CH), 129.43, 122.86, 121.57, 115.68 (Ph-C), 72.17 (O-CH), 53.83 (CH2), 16.42 (CH3). HRMS: m/z 337.1920 (calcd. 337.1916 for C21H25N2O2).
1,3-Bis[2-(4-nitrophenoxy)propyl]imidazolium Bromide (7d). Total yield 23.2%. Yellow oil. Molecules 17 12121 i035 = −10.99 (c 0.56, CHCl3). IR: υC=N 1602 cm−1, υNO2 1508 cm−1, υC-O 1260 cm−1. 1H-NMR (400 MHz, CD3OD) δ 9.58 (s, 1H, N-CH=N), 8.07 (d, J = 2.1 Hz, 2H, N-CH), 8.05 (d, J = 2.0 Hz, 2H, Ph-H), 7.98 (d, J = 1.5 Hz, 2H, Ph-H), 7.16 (d, J = 2.1 Hz, 2H, Ph-H), 7.14 (d, J = 2.0 Hz, 2H, Ph-H), 5.23–5.15 (m, 2H, O-CH), 4.91 (dd, J = 14.3, 2.6 Hz, 2H, NCH-H), 4.74 (dd, J = 14.4, 8.4 Hz, 2H, NCH-H), 1.55 (d, J = 6.1 Hz, 6H, CH3). 13C-NMR (101 MHz, CD3OD) δ 163.13 (N-CH=N), 142.48 (Ph-C), 138.54 (N-CH=CH), 126.69, 124.49, 116.50 (Ph-C), 74.37 (O-CH), 54.80 (CH2), 16.76 (CH3). HRMS: m/z 427.1630 (calcd. 427.1618 for C21H23N4O6).
1,3-Bis[2-(4-methylphenoxy)propyl]imidazolium Bromide (7e). Total yield 22.7%. Yellow oil. Molecules 17 12121 i036 = −15.12 (c 1.025, CHCl3). IR: υC=N 1616 cm−1, υC-O 1092 cm−1. 1H-NMR (400 MHz, CDCl3) δ 10.04 (s, 1H, N-CH=N), 7.83 (d, J = 0.9 Hz, 2H, N-CH), 6.96 (d, J = 8.4 Hz, 4H, Ph-H), 6.71 (t, J = 5.9 Hz, 4H, Ph-H), 4.91 (dd, J = 13.8, 2.0 Hz, 2H, NCH-H), 4.77–4.70 (m, 2H, O-CH), 4.47 (dd, J = 13.9, 7.9 Hz, 2H, NCH-H), 2.22 (s, 6H, Ph-CH3), 1.33 (d, J = 6.2 Hz, 6H, CH-CH3). 13C-NMR (101 MHz, CDCl3) δ 153.54 (N-CH=N), 136.59 (N-CH=CH), 130.31, 129.33, 122.40, 115.28 (Ph-C), 72.02 (O-CH), 53.31 (CH2), 19.68 (CH-CH3), 15.95 (Ph-CH3). HRMS: m/z 365.2228 (calcd. 365.2229 for C23H29N2O2).

3.3. General Procedure for the Suzuki-Miyaura Coupling Reactions

Under an Ar atmosphere, Pd(OAc)2 (0.6 mg, 0.0025 mmol, 0.5 mol%), imidazolium salt (0.005 mmol, 1 mol%), arylboronic acid (0.75 mmol), aryl halide (0.5 mmol), K2CO3 (207.30 mg, 1.5 mmol), and toluene (3.0 mL) were added to a dried Schlenk tube in sequence. The mixture was stirred at 110 °C and the progress of the reaction was monitored by TLC and gas chromatography. Upon the consumption of aryl halide, the mixture was cooled to room temperature, and H2O (3.0 mL) was added to quench the reaction. The organic layer was separated, and the aqueous layer was back-extracted with CH2Cl2 (3.0 mL × 3). The organic phases were combined, dried over Na2SO4 and concentrated. The product was isolated by column chromatography with petroleum ether-ethyl acetate as the eluents or analyzed by gas chromatography using diethylene glycol di-n-butyl ether as an internal standard. The structures of the coupling products were confirmed by comparison of 1H-NMR, 13C-NMR with those reported in literature. All products, 12aq, showed molecular ionic peak in MS specta.
Biphenyl (12a): colorless solid, m.p. 69–70 °C (lit. [46], m.p. 69–70 °C).1H-NMR (400 MHz, CDCl3) δ 7.62–7.56 (m, 2H), 7.48–7.40 (m, 2H), 7.40–7.29 (m, 1H) ppm. 13C-NMR (101 MHz, CDCl3) δ 141.37, 128.89, 127.38, 127.30 ppm. The NMR data were in agreement with those reported in literature [46].
4-Acetylbiphenyl (12b): colorless solid, m.p. 122–124 °C (lit. [46], m.p. 120–121 °C). 1H-NMR (400 MHz, CDCl3) δ 8.04 (d, J = 8.3 Hz, 2H), 7.69 (d, J = 8.3 Hz, 2H), 7.65–7.60 (m, 2H), 7.48 (t, J = 7.4 Hz, 2H), 7.41 (d, J = 7.2 Hz, 1H), 2.64 (s, 3H) ppm. 13C-NMR (101 MHz, CDCl3) δ 197.62, 145.69, 139.82, 135.87, 128.96, 128.91, 128.24, 127.24, 127.17, 26.60 ppm. The NMR spectral data matched literature data [46,47].
4-Nitrobiphenyl (12c): yellow solid, m.p. 112–113 °C (lit. [46], m.p. 112–114 °C). 1H-NMR (400 MHz, CDCl3) δ 8.29 (d, J = 8.7 Hz, 2H), 7.73 (d, J = 8.7 Hz, 2H), 7.63 (d, J = 7.3 Hz, 2H), 7.49 (dt, J = 13.6, 7.0 Hz, 3H) ppm. 13C-NMR (101 MHz, CDCl3) δ 147.62, 147.06, 138.74, 129.21, 128.99, 127.82, 127.42, 124.14 ppm. The NMR spectral data matched literature data [46,47].
4-Cyanobiphenyl (12d): colorless solid, m.p. 86–87 °C (lit. [46], m.p. 85–86 °C). 1H-NMR (400 MHz, CDCl3) δ 7.73–7.60 (m, 4H), 7.59–7.52 (m, 2H), 7.50–7.33 (m, 3H) ppm. 13C-NMR (101 MHz, CDCl3) δ 145.55, 139.05, 132.55, 129.10, 128.66, 127.67, 127.18, 118.95, 110.81 ppm. The NMR spectral data matched literature data [46,47].
Methyl biphenyl-4-carboxylate (12e): colorless solid, m.p. 119–120 °C (lit. [46], m.p. 115–116 °C). 1H-NMR (400 MHz, CDCl3) δ 8.13 (d, J = 8.3 Hz, 2H), 7.65 (dd, J = 14.4, 7.8 Hz, 4H), 7.44 (dt, J = 26.9, 7.2 Hz, 3H), 3.95 (s, 3H) ppm. 13C-NMR (101 MHz, CDCl3) δ 166.96, 145.58, 139.94, 130.12, 128.94, 128.87, 128.16, 127.27, 127.03, 52.13 ppm. The NMR data are in agreement to those in literature [46,47].
N-Acetyl-4-aminobiphenyl (12f): colorless solid, m.p. 172–174 °C (lit. [46], m.p. 171–172 °C). 1H-NMR (400 MHz, CD3OD) δ 7.65–7.50 (m, 6H), 7.39 (t, J = 7.6 Hz, 2H), 7.28 (t, J = 7.3 Hz, 1H), 2.13 (s, 3H) ppm. 13C-NMR (101 MHz, CD3OD) δ 171.60, 141.81, 139.21, 138.13, 129.80, 128.20, 128.04, 127.64, 121.41, 23.84 ppm. The NMR spectral data matched literature data [46,48].
4-chlorobiphenyl (12g): colorless solid, m.p. 77–78 °C (lit. [46], m.p. 78–79 °C, lit. [49], m.p. 75–78 °C). 1H-NMR (400 MHz, CDCl3) δ 7.56 (dd, J = 14.8, 7.9 Hz, 4H), 7.50–7.35 (m, 5H) ppm. 13C-NMR (101 MHz, CDCl3) δ 140.09, 139.76, 133.48, 129.03, 129.01, 128.51, 127.71, 127.11 ppm. The NMR spectral data matched literature data [46,49].
4-Phenylphenol (12h): colorless solid, m.p. 164–166 °C (lit. [50], m.p. 162–164 °C, lit. [51], m.p. 167.2–168.0 °C). 1H-NMR (400 MHz, CD3OD) δ 7.48 (dd, J = 8.2, 1.0 Hz, 2H), 7.45–7.37 (m, 2H), 7.32 (t, J = 7.7 Hz, 2H), 7.24–7.16 (m, 1H), 6.92–6.83 (m, 2H) ppm. 13C-NMR (101 MHz, CD3OD) δ 157.93, 142.18, 133.74, 129.60, 128.96, 127.30, 127.28, 116.53, 49.43, 49.21, 49.00, 48.79, 48.57 ppm. The NMR spectral data matched literature data [50,51]. MS (EI): m/z = 170 [M+].
2-Butyoxyl-5-tert-butylbiphenyl (12i): colorless oil. 1H-NMR (400 MHz, CDCl3) δ 7.70–7.65 (m, 2H), 7.53–7.46 (m, 3H), 7.44–7.38 (m, 2H), 7.02 (d, J = 8.6 Hz, 1H), 4.04 (t, J = 6.4 Hz, 2H), 1.79 (dt, J = 14.3, 6.5 Hz, 2H), 1.52 (dd, J = 14.9, 7.5 Hz, 2H), 1.45 (s, 9H), 1.01 (t, J = 7.4 Hz, 3H) ppm. 13C-NMR (101 MHz, CDCl3) δ 154.00, 143.45, 139.37, 130.46, 129.80, 128.24, 127.90, 126.74, 125.24, 112.30, 77.48, 77.16, 76.84, 68.35, 34.28, 31.71, 31.50, 19.45, 13.91 ppm.
4-Methylbiphenyl (12j): colorless solid, m.p. 42–44 °C (lit. [52], m.p. 41–42 °C). 1H-NMR (400 MHz, CDCl3) δ 7.70–7.65 (m, 2H), 7.59 (d, J = 8.1 Hz, 2H), 7.51 (t, J = 7.6 Hz, 2H), 7.41 (dd, J = 11.6, 4.3 Hz, 1H), 7.33 (d, J = 7.9 Hz, 2H), 2.48 (s, 3H) ppm. 13C-NMR (101 MHz, CDCl3) δ 141.30, 138.50, 137.11, 129.60, 128.83, 127.12, 127.09, 21.22 ppm. The NMR spectral data matched literature data [52,53].
2-Methylbiphenyl (12k): colorless oil. 1H-NMR (400 MHz, CDCl3) δ 7.53–7.47 (m, 2H), 7.42 (d, J = 7.2 Hz, 3H), 7.39–7.30 (m, 4H), 2.37 (s, 3H) ppm. 13C-NMR (101 MHz, CDCl3) δ 142.06, 142.04, 135.44, 130.42, 129.91, 129.30, 128.18, 127.36, 126.87, 125.88, 20.61 ppm. The NMR spectral data matched literature data [52,54].
2,5-Dimethylbiphenyl (12l): colorless solid, m.p. 68–70 °C. 1H-NMR (400 MHz, CDCl3) δ 7.40–7.35 (m, 2H), 7.30 (d, J = 7.1 Hz, 3H), 7.15 (d, J = 7.9 Hz, 1H), 7.06 (d, J = 6.7 Hz, 2H), 2.33 (s, 3H), 2.22 (s, 3H) ppm. 13C-NMR (101 MHz, CDCl3) δ 142.21, 141.87, 135.28, 132.27, 130.66, 130.37, 129.29, 128.14, 128.06, 126.80, 21.06, 20.10 ppm. The NMR spectral data matched literature data [55,56].
2,6-Dimethylbiphenyl (12m): colorless oil. 1H-NMR (400 MHz, CDCl3) δ 7.37 (t, J = 7.5 Hz, 2H), 7.28 (t, J = 7.3 Hz, 1H), 7.15–7.03 (m, 5H), 1.99 (s, 6H) ppm. 13C-NMR (101 MHz, CDCl3) δ 142.01, 141.25, 136.17, 129.16, 128.54, 127.41, 127.15, 126.73, 20.97 ppm. The NMR spectral data matched literature data [52,57].
3,5-Di-tert-butylbiphenyl (12n): colorless solid, m.p. 63–64 °C (lit. [58], m.p. 62–63 °C). 1H-NMR (400 MHz, CDCl3) δ 7.69–7.64 (m, 2H), 7.55–7.45 (m, 5H), 7.40 (dd, J = 6.9, 1.1 Hz, 1H), 1.45 (d, J = 2.3 Hz, 18H) ppm. 13C-NMR (101 MHz, CDCl3) δ 151.25, 142.70, 140.87, 128.77, 127.62, 127.09, 121.87, 121.53, 35.14, 31.71 ppm. The NMR spectral data matched literature data [58].
Diphenylmethane (12o): colorless solid, m.p. 20–22 °C (lit. [59], m.p. 21–24 °C). 1H-NMR (400 MHz, CDCl3) δ 7.27 (d, J = 7.2 Hz, 4H), 7.18 (d, J = 6.5 Hz, 6H), 3.98 (s, 2H) ppm. 13C-NMR (101 MHz, CDCl3) δ 141.24, 129.06, 128.58, 126.19, 42.07 ppm. The NMR spectral data matched literature data [59,60].
3-Methylbiphenyl (12p): colorless oil. 1H-NMR (400 MHz, CDCl3) δ 7.48 (d, J = 7.2 Hz, 2H), 7.31 (t, J = 10.4 Hz, 4H), 7.22 (t, J = 7.1 Hz, 2H), 7.06 (d, J = 7.1 Hz, 1H), 2.31 (s, 3H) ppm. 13C-NMR (101 MHz, CDCl3) δ 141.46, 141.34, 138.42, 128.81, 128.79, 128.12, 128.09, 127.29, 124.39, 21.68 ppm. The NMR spectral data matched literature data [61].
Methyl 1-(4′-acetyl-biphenyl-4-yl)-carboxylate (12q): colorless solid, m.p. 165–168 °C (lit. [62], m.p. 164.5–166 °C). 1H-NMR (400 MHz, CDCl3) δ 8.14 (d, J = 8.4 Hz, 2H), 8.05 (d, J = 8.4 Hz, 2H), 7.71 (t, J = 8.8 Hz, 4H), 3.95 (s, 3H), 2.65 (s, 3H) ppm. 13C-NMR (101 MHz, CDCl3) δ 197.68, 166.87, 144.57, 144.32, 136.65, 130.34, 129.91, 129.10, 127.56, 127.35, 52.34, 26.80 ppm. The NMR spectral data matched literature data [62,63].

4. Conclusions

In conclusion, we have synthesized a range of novel C2-symmetric NHC precursor imidazolium salts containing side arms substituted with aryloxyl groups. The Suzuki-Miyaura coupling reaction could be catalyzed remarkably by the Pd/NHC catalysts formed in situ. Various functionalized and sterically hindered aryl halides and arylboronic acids could be used. TON of up to 105 was achieved with 5 × 10−4 mol% Pd catalyst.

Supplementary Materials

Supplementary materials can be accessed at: https://www.mdpi.com/1420-3049/17/10/12121/s1.

Acknowledgments

We are grateful to the NSFC for financial support (Grant No. 20672016, 20972020), to Huiying Chen for recording NMR spectra, and to Ally Chem. Ltd., for a generous gift of arylboronic acids.

Conflict of Interest

The authors declare no conflict of interest.
  • Sample Availability: Samples of the compounds 12aq are available from the authors.

References

  1. Suzuki, A. Recent advances in the cross-coupling reactions of organoboron derivatives with organic electrophiles, 1995–1998. J. Organomet. Chem. 1999, 576, 147–168. [Google Scholar] [CrossRef]
  2. Nicolaou, K.C.; Bulger, P.G.; Sarlah, D. Palladium-catalyzed cross-coupling reactions in total synthesis. Angew. Chem. Int. Ed. 2005, 44, 4442–4489. [Google Scholar] [CrossRef]
  3. Phan, N.T.S.; Sluys, M.V.D.; 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]
  4. Nicolaou, K.C.; Ramanjulu, J.M.; Natarajan, S.; Bräse, S.; Rübsam, F. A Suzuki coupling-macrolactamization approach to the AB-COD bicyclic system of vancomycin. Chem. Commun. 1997, 1899–1900. [Google Scholar]
  5. 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]
  6. Sandee, A.J.; Williams, C.K.; Evans, N.R.; Davies, J.E.; Boothby, C.E.; Köhler, A.; Friend, R.H.; Holmes, A.B. Solution-processible conjugated electrophosphorescent polymers. J. Am. Chem. Soc. 2004, 126, 7041–7048. [Google Scholar]
  7. Miyaura, N.; Suzuki, A. Stereoselective synthesis of arylated (E)-alkenes by the reaction of alk-1-enylboranes with aryl halides in the presence of palladium catalyst. J. Chem. Soc. Chem. Commun. 1979, 866–867. [Google Scholar] [CrossRef]
  8. Miyaura, N.; Yanagi, T.; Suzuki, A. The palladium-catalyzed cross-coupling reaction of phenylboronic acid with haloarenes in the presence of bases. Synth. Commun. 1981, 11, 513–519. [Google Scholar] [CrossRef]
  9. Miyaura, N.; Suzuki, A. Palladium-catalyzed cross-coupling reactions of organoboron compounds. Chem. Rev. 1995, 95, 2457–2483. [Google Scholar] [CrossRef]
  10. Netherton, M.R.; Dai, C.; Neuschütz, K.; Fu, G.C. Room-temperature alkyl–alkyl Suzuki cross-coupling of alkyl bromides that possess β-hydrogens. J. Am. Chem. Soc. 2002. [Google Scholar] [CrossRef]
  11. Stambuli, J.P.; Kuwano, R.; Hartwig, J.F. Unparalleled rates for the activation of aryl chlorides and bromides: Coupling with amines and boronic acids in minutes at room temperature. Angew. Chem. Int. Ed. 2002, 41, 4746–4748. [Google Scholar] [CrossRef]
  12. Arduengo, A.J., III; Harlow, R.L.; Kline, M. A stable crystalline carbene. J. Am. Chem. Soc. 1991, 113, 361–363. [Google Scholar]
  13. Herrmann, W.A.; Elison, M.; Fischer, J.; Köcher, C.; Artus, G.R.J. Metal complexes of N-heterocyclic carbenes—a new structural principle for catalysts in homogeneous catalysis. Angew. Chem. Int. Ed. Engl. 1995, 34, 2371–2374. [Google Scholar] [CrossRef]
  14. Herrmann, W.A. N-heterocyclic carbenes: A new concept in organometallic catalysis. Angew. Chem. Int. Ed. 2002, 41, 1290–1309. [Google Scholar] [CrossRef]
  15. Scott, N.M.; Clavier, H.; Mahjoor, P.; Stevens, E.D.; Nolan, S.P. Synthetic, structural, and thermochemical studies of N-heterocyclic carbene (NHC) and tertiary phosphine ligands in the [(L)2Ni(CO)2] (L = PR3, NHC) system. Organometallics 2008, 27, 3181–3186. [Google Scholar]
  16. Lee, M.-T.; Hu, C.-H. Density functional study of N-heterocyclic and diamino carbene complexes: Comparison with phosphines. Organometallics 2004, 23, 976–983. [Google Scholar]
  17. Colacino, E.; Martinez, J.; Lamaty, F. Preparation of NHC-ruthenium complexes and their catalytic activity in metathesis reaction. Coord. Chem. Rev. 2007, 251, 726–764. [Google Scholar]
  18. Gade, L.H.; César, V.; Bellemin-Laponnaz, S. A modular assembly of chiral oxazolinylcarbene-rhodium complexes: Efficient phosphane-free catalysts for the asymmetric hydrosilylation of dialkyl ketones. Angew. Chem. Int. Ed. 2004, 43, 1014–1017. [Google Scholar] [CrossRef]
  19. Song, C.; Ma, C.; Ma, Y.; Feng, W.; Ma, S.; Chai, Q.; Andrus, M.B. Bis-paracyclophane N-heterocyclic carbene-ruthenium catalyzed asymmetric ketone hydrosilylation. Tetrahedron Lett. 2005, 46, 3241–3244. [Google Scholar]
  20. Powell, M.T.; Hou, D.-R.; Perry, M.C.; Cui, X.; Burgess, K. Chiral imidazolylidine ligands for asymmetric hydrogenation of aryl alkenes. J. Am. Chem. Soc. 2001, 123, 8878–8879. [Google Scholar]
  21. César, V.; Bellemin-Laponnaz, S.; Gade, L.H. Chiral N-heterocyclic carbenes as stereodirecting ligands in asymmetric catalysis. Chem. Soc. Rev. 2004, 33, 619–636. [Google Scholar]
  22. Alonso, F.; Beletskaya, I.P.; Yus, M. Non-conventional methodologies for transition-metal catalysed carbonecarbon coupling: A critical overview. Part 2: The Suzuki reaction. Tetrahedron 2008, 64, 3047–3101. [Google Scholar] [CrossRef]
  23. Herrmann, W.A.; Reisinger, C.P.; Spiegler, M. Chelating N-heterocyclic carbene ligands in palladium-catalyzed Heck-type reactions. J. Organomet. Chem. 1998, 557, 93–96. [Google Scholar]
  24. Kim, J.-H.; Kim, J.-W.; Shokouhimehr, M.; Lee, Y.-S. Polymer-supported N-heterocyclic carbene-palladium complex for heterogeneous Suzuki cross-coupling reaction. J. Org. Chem. 2005, 70, 6714–6720. [Google Scholar]
  25. Schneider, S.K.; Herrmann, W.A.; Herdtweck, E. Active catalysts for the Suzuki coupling: Palladium complexes of tetrahydropyrimid-2-ylidenes. J. Mol. Catal. A Chem. 2006, 245, 248–254. [Google Scholar] [CrossRef]
  26. Navarro, O.; Kelly, R.A.; Nolan, S.P. A general method for the Suzuki-Miyaura cross-coupling of sterically hindered aryl chlorides: Synthesis of di- and tri-ortho-substituted biaryls in 2-propanol at room temperature. J. Am. Chem. Soc. 2003, 125, 16194–16195. [Google Scholar] [CrossRef]
  27. Marion, N.; Navarro, O.; Mei, J.; Stevens, E.D.; Scott, N.M.; Nolan, S.P. Modified (NHC)Pd(allyl)Cl (NHC=N-heterocyclic carbene) complexes for room-temperature Suzuki-Miyaura and Buchwald-Hartwig reactions. J. Am. Chem. Soc. 2006, 128, 4101–4111. [Google Scholar]
  28. Kantchev, E.A.B.; ÓBrien, C.J.; Organ, M.G. Palladium complexes of N-heterocyclic carbenes as catalysts for cross-coupling reactions—a synthetic chemist’s perspective. Angew. Chem. Int. Ed. 2007, 46, 2768–2813. [Google Scholar] [CrossRef]
  29. Marion, N.; Nolan, S.P. Well-defined N-heterocyclic carbenes-palladium(II) precatalysts for cross-coupling reactions. Acc. Chem. Res. 2008, 41, 1440–1449. [Google Scholar] [CrossRef]
  30. Jiang, L.; Li, Z.N.; Zhao, D.F. Progress in the Heck reaction and Suzuki reaction catalyzed by Pd-N-heterocyclic carbene complexes. Chin. J. Org. Chem. 2010, 30, 200–210. [Google Scholar]
  31. Türkmen, H.; Pelit, L.; Çetinkaya, B. Water-soluble cis-[(NHC)PdBr2(TPPTS)] catalysts and their applications in Suzuki-Miyaura coupling of aryl chlorides. J. Mol. Catal. A Chem. 2011, 348, 88–93. [Google Scholar] [CrossRef]
  32. Godoy, F.; Segarra, C.; Poyatos, M.; Peris, E. Palladium catalysts with sulfonate-functionalized-NHC ligands for Suzuki-Miyaura cross-coupling reactions in water. Organometallics 2011, 30, 684–688. [Google Scholar]
  33. Dible, B.R.; Cowley, R.E.; Holland, P.L. Remote substitution on N-heterocyclic carbenes heightens the catalytic reactivity of their palladium complexes. Organometallics 2011, 30, 5123–5132. [Google Scholar]
  34. Xu, X.; Xu, B.; Li, Y.; Hong, S.H. Abnormal N-heterocyclic carbene promoted Suzuki-Miyaura coupling reaction: A comparative study. Organometallics 2010, 29, 6343–6349. [Google Scholar] [CrossRef]
  35. Xu, Q.; Duan, W.-L.; Lei, Z.-Y.; Zhu, Z.B.; Shi, M. A novel cis-chelated Pd(II)-NHC complex for catalyzing Suzuki and Heck-type cross-coupling reactions. Tetrahedron 2005, 61, 11225–11229. [Google Scholar] [CrossRef]
  36. Shi, J.-C.; Yang, P.-Y.; Tong, Q.; Wu, Y.; Peng, Y. Highly efficient and stable palladium/imidazolium salt-phosphine catalysts for Suzuki-Miyaura cross-coupling of aryl bromides. J. Mol. Catal. A Chem. 2006, 259, 7–10. [Google Scholar] [CrossRef]
  37. Tulloch, A.A.D.; Danopoulos, A.A.; Tooze, R.P.; Cafferkey, S.M.; Kleinhenz, S.; Hursthouse, M.B. Pyridine functionalised N-heterocyclic carbene complexes of palladium. Chem. Commun. 2000, 1247–1248. [Google Scholar]
  38. Netland, K.A.; Krivokapic, A.; Schröder, M.; Boldt, K.; Lundvall, F.; Tilset, M. Synthesis, X-ray structures, and catalytic applications of palladium(II) complexes bearing N-heterocyclic iminocarbene ligands. J. Organomet. Chem. 2008, 693, 3703–3710. [Google Scholar] [CrossRef]
  39. Yang, C.; Lee, H.M.; Nolan, S.P. Highly efficient Heck reactions of aryl bromides with n-butyl acrylate mediated by a palladium/phosphine-imidazolium salt system. Org. Lett. 2001, 3, 1511–1514. [Google Scholar] [CrossRef]
  40. Wang, J.-W.; Meng, F.-H.; Zhang, L.-F. Suzuki coupling reaction of aryl halides catalyzed by an N-heterocyclic carbene-PdCl2 species based on a porphyrin at room temperature. Organometallics 2009, 28, 2334–2337. [Google Scholar] [CrossRef]
  41. Zhang, X.; Qiu, Y.; Rao, B.; Luo, M. Palladium(II)−N-heterocyclic carbene metallacrown ether complexes: Synthesis, structure, and catalytic activity in the Suzuki-Miyaura reaction. Organometallics 2009, 28, 3093–3099. [Google Scholar]
  42. Yuan, D.; Huynh, H.V. Dinuclear and tetranuclear palladium(II) complexes of a thiolato-functionalized, benzannulated N-heterocyclic carbene ligand and their activities toward Suzuki-Miyaura coupling. Organometallics 2010, 29, 6020–6027. [Google Scholar] [CrossRef]
  43. Shan, F.J.; Jiang, L.; Li, Z.N.; Zhao, D.F. Asymmetric conjugate addition to cyclic enone catalyzed by Cu-NHC complexes with C2 symmetry. Chin. J. Chem. 2011, 29, 973–977. [Google Scholar] [CrossRef]
  44. Navarro, O.; Kaur, H.; Mahjoor, P.; Nolan, S.P. Cross-coupling and dehalogenation reactions catalyzed by (N-heterocyclic carbene)Pd(allyl)Cl complexes. J. Org. Chem. 2004, 69, 3173–3180. [Google Scholar] [CrossRef]
  45. Watanabe, T.; Miyaura, N.; Suzuki, A. Synthesis of sterically hindered biaryls via the palladium-catalyzed cross-coupling reaction of arylboronic acids or their esters with haloarenes. Synlett 1992, 3, 207–210. [Google Scholar]
  46. 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]
  47. 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]
  48. 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]
  49. Kylmälä, T.; Kuuloja, N.; Xu, Y.; Rissanen, K.; Franzén, R. Synthesis of chlorinated biphenyls by Suzuki cross-coupling using diamine or diimine-palladium complexes. Eur. J. Org. Chem. 2008, 4019–4024. [Google Scholar]
  50. Schmidt, B.; Hölter, F. Suzuki-Miyaura cross coupling reactions with phenoldiazonium salts. Org. Biomol. Chem. 2011, 9, 4914–4920. [Google Scholar] [CrossRef]
  51. Hanada, S.; Yuasa, A.; Kuroiwa, H.; Motoyama, Y.; Nagashima, H. Hydrosilanes are not always reducing agents for carbonyl compounds, II: ruthenium-catalyzed deprotection of tert-butyl groups in carbamates, carbonates, esters, and ethers. Eur. J. Org. Chem. 2010, 6, 1021–1025. [Google Scholar]
  52. Fan, X.-H.; Yang, L.-M. NiII–(σ-Aryl) complex catalyzed Suzuki reaction of aryl tosylates with arylboronic acids. Eur. J. Org. Chem. 2010, 13, 2457–2460. [Google Scholar]
  53. Diebolt, O.; Braunstein, P.; Nolan, S.P.; Cazin, C.S.J. Room-temperature activation of aryl chlorides in Suzuki–Miyaura coupling using a [Pd(l-Cl)Cl(NHC)]2 complex (NHC = N-heterocyclic carbene). Chem. Commun. 2008, 3190–3192. [Google Scholar]
  54. Firouzabadi, H.; Iranpoor, N.; Gholinejad, M. 2-Aminophenyl diphenylphosphinite as an easily accessible ligand for heterogeneous palladium-catalyzed SuzukieMiyaura reaction in water in the absence of any organic co-solvent. J. Organomet. Chem. 2010, 695, 2093–2097. [Google Scholar]
  55. Kawai, H.; Kobayashi, Y.; Oi, S.; Inoue, Y. Direct C–H bond arylation of arenes with aryltin reagents catalysed by palladium complexes. Chem. Commun. 2008, 1464–1466. [Google Scholar]
  56. Liu, D.; Gao, W.; Dai, Q.; Zhang, X. Triazole-based monophosphines for Suzuki-Miyaura coupling and amination reactions of aryl chlorides. Org. Lett. 2005, 7, 4907–4910. [Google Scholar] [CrossRef]
  57. Wolfe, J.P.; Singer, R.A.; Yang, B.H.; Buchwald, S.L. Highly active palladium catalysts for Suzuki coupling reactions. J. Am. Chem. Soc. 1999, 121, 9550–9561. [Google Scholar] [CrossRef]
  58. Thomas, A.; Anilkumar, G.; Nair, V. Photolytic double decarbonylation route to highly substituted indenes and benzene derivatives. Tetrahedron 1996, 52, 2481–2488. [Google Scholar] [CrossRef]
  59. Karami, K.; Rizzoli, C.; Salah, M.M. Synthesis and application of ortho-palladated complex of (4-phenylbenzoylmethylene)triphenylphosphorane as a highly active catalyst in the Suzuki cross-coupling reaction. J. Organomet. Chem. 2011, 696, 940–945. [Google Scholar]
  60. Peña-López, M.; Ayán-Varela, M.; Sarandeses, L.A.; Sestelo, J.P. Palladium-catalyzed cross-coupling reactions of organogold(I) reagents with organic electrophiles. Chem. Eur. J. 2010, 16, 9905–9909. [Google Scholar] [CrossRef]
  61. Pal, A.; Ghosh, R.; Adarsh, N.N.; Sarkar, A. Pyrazole-tethered phosphine ligands for Pd(0): useful catalysts for Stille, Kumada and Hiyama cross-coupling reactions. Tetrahedron 2010, 66, 5451–5458. [Google Scholar] [CrossRef]
  62. Zhu, L.; Duquette, J.; Zhang, M. An improved preparation of arylboronates: Application in one-pot Suzuki biaryl synthesis. J. Org. Chem. 2003, 68, 3729–3732. [Google Scholar] [CrossRef]
  63. Wu, L.; Li, Z.-W.; Zhang, F.; He, Y.-M.; Fan, Q.-H. Air-stable and highly active dendritic phosphine oxide-stabilized palladium nanoparticles: preparation, characterization and applications in the carbon-carbon bond formation and hydrogenation reactions. Adv. Synth. Catal. 2008, 350, 846–862. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Jiang, L.; Shan, F.; Li, Z.; Zhao, D. Suzuki-Miyaura Reactions Catalyzed by C2-Symmetric Pd-Multi-Dentate N-Heterocyclic Carbene Complexes. Molecules 2012, 17, 12121-12139. https://doi.org/10.3390/molecules171012121

AMA Style

Jiang L, Shan F, Li Z, Zhao D. Suzuki-Miyaura Reactions Catalyzed by C2-Symmetric Pd-Multi-Dentate N-Heterocyclic Carbene Complexes. Molecules. 2012; 17(10):12121-12139. https://doi.org/10.3390/molecules171012121

Chicago/Turabian Style

Jiang, Lan, Fengjun Shan, Zhengning Li, and Defeng Zhao. 2012. "Suzuki-Miyaura Reactions Catalyzed by C2-Symmetric Pd-Multi-Dentate N-Heterocyclic Carbene Complexes" Molecules 17, no. 10: 12121-12139. https://doi.org/10.3390/molecules171012121

Article Metrics

Back to TopTop