Next Article in Journal
Synthesis and Biological Activity of trans-Tiliroside Derivatives as Potent Anti-Diabetic Agents
Next Article in Special Issue
Enhancement in the Catalytic Activity of Pd/USY in the Heck Reaction Induced by H2 Bubbling
Previous Article in Journal
Antioxidant Activities of Total Phenols of Prunella vulgaris L. in Vitro and in Tumor-bearing Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sonogashira Reaction of Aryl and Heteroaryl Halides with Terminal Alkynes Catalyzed by a Highly Efficient and Recyclable Nanosized MCM-41 Anchored Palladium Bipyridyl Complex

1
Institute of Organic and Polymeric Materials, National Taipei University of Technology, Taipei 106, Taiwan
2
Department of Chemistry, National Taiwan University, Taipei 106, Taiwan
*
Author to whom correspondence should be addressed.
Molecules 2010, 15(12), 9157-9173; https://doi.org/10.3390/molecules15129157
Submission received: 11 November 2010 / Revised: 6 December 2010 / Accepted: 9 December 2010 / Published: 10 December 2010
(This article belongs to the Special Issue Cross-Coupling Reactions in Organic Synthesis)

Abstract

:
A heterogeneous catalyst, nanosized MCM-41-Pd, was used to catalyze the Sonogashira coupling of aryl and heteroaryl halides with terminal alkynes in the presence of CuI and triphenylphosphine. The coupling products were obtained in high yields using low Pd loadings to 0.01 mol%, and the nanosized MCM-41-Pd catalyst was recovered by centrifugation of the reaction solution and re-used in further runs without significant loss of reactivity.

1. Introduction

The reaction of aryl halides or vinyl halides with terminal alkynes catalyzed by a Pd(II)/Cu(I) system is known as the Sonogashira coupling, and is one of the most powerful methods for the straightforward construction of sp2sp carbon–carbon bonds in synthetic chemistry [1,2,3,4,5,6,7]. This methodology has been widely applied to prepare biologically-active molecules [8,9,10,11,12,13], natural products [14,15,16,17], conducting polymers/engineering materials [18,19], and macrocycles with acetylene links [20,21].
The Sonogashira reaction is, in general, carried out in a homogeneous phase [22], and therefore the recovery of expensive palladium complexes, facile separation of catalysts and products, and industrial application are major aims for the benefit of both economy and the environment. For these reasons, heterogenization of the homogeneous Sonogashira reaction has become an aim of great interest to chemists in recent years. Choudary and co-workers described a layered double hydroxide-supported nanopalladium catalyst for the coupling of aryl chlorides and phenylacetylene [23], and Pd/C has been used to catalyze the Sonagashira reaction of aryl halides with acetylenes [24,25,26,27,28,29,30], while PVP-supported nanoparticle palladium metal can be employed for the coupling of aryl iodides and bromides with terminal alkynes [31]. Djakovitch and co-workers reported that microporous [Pd-Cu]/NaY [32], [Pd(NH3)4]2+/(NH4)Y [33], and [Pd(NH3)4]2+/NaY [34] systems can be applied in the Sonogashira reaction using 1–2 mol% of the Pd catalyst, and palladium can be also supported by silica in order to create a recyclable catalyst for use in the Sonogashira reaction [35].
Mesoporous silica is becoming more and more widely used as a solid support owing to its well-defined structure, uniform pore size, high surface area, and large number of silanol groups for the grafting of metal complexes [36,37,38,39,40,41,42,43,44]. Djakovitch’s group prepared a mesoporous [Pd]/SBA-15 catalyst to demonstrate that larger aryl halides such as bromoanthacene can be active in this catalytic system, whereas the microporous support [Pd(NH3)4]2+/NaY is inactive [45]. Cai and co-workers employed MCM-41-supported sulfur palladium [46], bidentate phosphine palladium [47], and thioether palladium [48] systems to catalyze the coupling of aryl iodides and terminal alkynes after reduction of the catalyst. Although most known heterogeneous catalysts have been demonstrated to be able to be recycled for use in further runs, the use of catalytic amounts of 0.2–5 mol% of Pd for the Sonagashira reaction is still too high for a single batch reaction when compared with homogeneous catalysts [49,50,51,52,53,54]. We have recently prepared a nanosized MCM-41 grafted palladium bipyridyl complex, NS-MCM-41-Pd (Figure 1), as a highly efficient and recyclable catalyst for the Mizoroki-Heck reaction [55], Kumada-Tamao-Corriu reaction [56], ketone formation [57], and ynone formation [58], which require a very low catalyst loading for a single batch reaction.
Figure 1. NS-MCM-41-Pd.
Figure 1. NS-MCM-41-Pd.
Molecules 15 09157 g001
The major advantage of this catalyst is that the short and highly-connective wormhole-like channels of nanosized MCM-41 lead to the easy exchange of reactants, salts and products throughout the nanochannels, avoiding saturation of activity. In this paper, we report the use of nanosized MCM-41-Pd to catalyze the coupling of aryl and heteroaryl halides with phenylacetylene and alkynols with high efficiency under Sonogashira reaction conditions using a catalyst loading as low as 0.01 mol%, with the ability to recycle the catalyst for further use (Scheme 1).
Scheme 1. NS-MCM-41-Pd-catalyzed Sonogashira reaction.
Scheme 1. NS-MCM-41-Pd-catalyzed Sonogashira reaction.
Molecules 15 09157 g003

2. Results and Discussion

2.1. Optimization of reaction conditions for the Sonogashira reaction catalyzed by NS-MCM-41-Pd

The procedure for the synthesis of the catalyst, NS-MCM-41-Pd, was presented in our previous reports. After the grafting of the palladium bipyridyl complex onto NS-MCM-41, the surface area and pore diameter decreased from 705 m2/g and 2.5 nm to 588 m2/g and 2.3 nm, respectively, and the amount of Pd complex anchored to the wall of NS-MCM-41 was quantified to be 0.15 mmol/g by ICP-MASS analysis. In order to optimize the conditions for this prepared heterogeneous catalyst, the solvent effect was first examined using iodobenzene (1a) and phenylacetylene (2a) as representative reactants. Reactions were carried out in the presence of 0.1 mol% catalyst, 0.2 mol% CuI, 0.2 mol% PPh3, and Et3N at 50 °C under N2 for 3 h. The results are summarized in Table 1, and it was found that Et3N was the best solvent for this reaction (Table 1, Entry 1).
Table 1. NS-MCM-41-Pd-catalyzed Sonogashira coupling reaction of iodobenzene 1a with phenylacetylene 2a.a
Table 1. NS-MCM-41-Pd-catalyzed Sonogashira coupling reaction of iodobenzene 1a with phenylacetylene 2a.a
EntryPd (mol%)CuI (mol%)PPh3 (mol%)SolventBaseYield (%)b
10.10.20.2Et3NEt3N97
20.10.20.2TolueneEt3Nc60
30.10.20.2DMFEt3Nc34
40.10.20.2DMSOEt3Nc26
50.10.20.2NMPEt3Nc5
60.10.20Et3NEt3N38
70.100.2Et3NEt3N0
80.100Et3NEt3N0
90.10.20.2TolueneKOHc0
100.10.20.2DMFKOHc0
110.10.20.2TolueneK2CO3c0
120.10.20.2DMFK2CO3c0
a Reaction conditions: [1a]:[2a]:[Pd] = 1000:1100:1, at 50 °C for 3 h. b Isolated yields. c 3 equiv based on 1a was used as a base.
The use of toluene, DMF, DMSO, and NMP as solvents resulting in lower rates of conversion (Entries 2–5). In the absence of triphenylphosphine, this Sonogashira coupling reaction still proceeded, but with less satisfactory yields (Entry 6). However, the co-catalyst CuI appeared to be necessary for the coupling reaction (Entries 7–8). Regarding the use of a base, inorganic bases such as KOH and K2CO3 were also examined, but under these conditions the desired product was not obtained due to the poor solubility of these salts in organic solvents (Entries 9–12) [55].

2.2. Sonogashira reaction of aryl halides with phenylacetylene

Following optimization of the reaction conditions, the reactions of 2a with various aryl halides were screened in the subsequent investigation (Table 2).
Table 2. Sonogashira reaction of aryl halides (1) with phenylacetylene (2a) catalyzed by nanosized MCM-41-Pd.a
Table 2. Sonogashira reaction of aryl halides (1) with phenylacetylene (2a) catalyzed by nanosized MCM-41-Pd.a
EntryAryl halidePd (mol%)Solvent/BaseT (°C)t (h)Yield (%)bTON
1C6H5I1a0.1Et3N/Et3N5033a, 97970
2C6H5I1a0.01Et3N/Et3N50123a, 989800
34-IC6H4CN1b0.1Et3N/Et3N5033b, 96960
44-IC6H4CN1b0.01Et3N/Et3N5093b, 969600
54-MeOC6H4I1c0.1Et3N/Et3N50243c, 87870
6C6H5Br1d0.1NMP/Bu3Nc140243a, 30300
7C6H5Br1d0.1Toluene/Bu3Nc100243a, 56560
84-BrC6H4CN1e0.1Et3N/Et3N9033b, 93930
94-MeCOC6H4Br1f0.1NMP/Et3Nc9063d, 98980
104-NO2C6H4Br1g0.01NMP/Et3Nc9063e, 999900
114-ClC6H4Br1h0.1NMP/Et3Nc90243f, 46460
124-MeOC6H4Br1i0.1NMP/Et3Nc90723c, 40400
132-Bromothiophene1j0.1NMP/Et3Nc90483g, 71710
143-Bromothiophene1k0.1NMP/Et3Nc90963h, 36360
152-Bromopyridine1l0.1NMP/Et3Nc9033i, 99990
163-Bromopyridine1m0.1NMP/Et3Nc90243j, 98980
a Reaction conditions: [1]:[2a] = 1:1.1, [Pd]:[CuI]:[PPh3] = 1:2:2. b Isolated yields. c 3 equiv based on 1 was used as a base.
Reactions of 2a with aryl iodides proceeded well with the use of 0.1 mol% NS-MCM-41-Pd at 50 °C (Table 2, Entries 1, 3, and 5), and it should be noted that the same good yields were also obtained in these reactions when using a lower amount of catalyst (0.01 mol%) (Entries 2 and 4). However, the Sonogashira reaction of bromobenzene (1d) and 2a under the same conditions did not afford any product, but replacing the Et3N solvent by NMP, a typical solvent for such coupling reactions, resulted in the formation of 3a in a 30% yield at an elevated temperature (Entry 6), and a yield of up to 56% could be achieved by performing the reaction in toluene at 100 °C for 24 h (Entry 7). Using activated aryl bromides such as 4-bromobenzonitrile (1e), 4-bromoacetophenone (1f), and 4-bromonitrobenzene (1g), better yields of the coupling reactions were observed (Entries 8–10). In the case of the coupling of 1h with 2a, the C–Cl bond was inert under the reaction conditions, while the product coupled through the C–Br bond was obtained in a 46% yield (Entry 11). Next, we studied the coupling of halothiophenes and halopyridines with 2a, and it appeared that 2-bromothiophene (1j) and 2-bromo-pyridine (1l) resulted in better yields than the corresponding bromides at the 3-position (Entries 13–16).

2.3. Sonogashira reaction of aryl halides with alkynols

Under similar conditions, NS-MCM-41-Pd-catalyzed Sonogashira coupling of a wide variety of aryl halides with 2-methyl-3-butyn-2-ol (4a) was also achieved (Table 3), and a reaction temperature of 90 °C was found to be optimal. Aryl iodides reacted with 4a to give the corresponding coupling products in good to excellent yields (Entries 1–3), whereas the use of deactivated bromides as substrates resulted in lower yields (Entry 4). Reactions of activated bromides delivered better conversion rates (Entries 5–9): for example, the catalyst had a turnover number (TON) of 9,600 for the coupling of 1e with 4a (Entry 7), and for the heteroaryl halides (Entries 11–15), the catalyst exhibited great activity, with the exception of the reactions of 1k (Entries 12).
Table 3. Sonogashira reaction of aryl halides 1 with alkynols 4 catalyzed by nanosized MCM-41-Pd.a
Table 3. Sonogashira reaction of aryl halides 1 with alkynols 4 catalyzed by nanosized MCM-41-Pd.a
EntryAryl halide Alkynyl alcohol Pd (mol%)t (h)Yield (%)bTON
1C6H5I1aHC≡CC(CH3)2OH4a0.135a, 94940
24-IC6H4CN1bHC≡CC(CH3)2OH4a0.135b, 98980
34-MeOC6H4I1cHC≡CC(CH3)2OH4a0.1725c, 61610
4C6H5Br1dHC≡CC(CH3)2OH4a0.1965a, 21210
54-BrC6H4CN1eHC≡CC(CH3)2OH4a0.135b, 98980
64-BrC6H4CN1eHC≡CC(CH3)2OH4a0.01125b, 969600
74-MeCOC6H4Br1fHC≡CC(CH3)2OH4a0.135d, 98980
84-NO2C6H4Br1gHC≡CC(CH3)2OH4a0.135e, 97970
94-ClC6H4Br1hHC≡CC(CH3)2OH4a0.1245f, 69690
104-MeOC6H4Br1iHC≡CC(CH3)2OH4a0.1965c, 20200
112-Bromothiophene1jHC≡CC(CH3)2OH4a0.1485g, 99990
123-Bromothiophene1kHC≡CC(CH3)2OH4a0.1965h, 59590
132-Bromopyridine1lHC≡CC(CH3)2OH4a0.135i, 99990
143-Bromopyridine1mHC≡CC(CH3)2OH4a0.165j, 98980
153-Bromopyridine1mHC≡CC(CH3)2OH4a0.01245j, 343400
16C6H5I1aHC≡CCH2OH4b0.1125k, 85850
17C6H5I1aHC≡CCH2OH4b0.01245k, 848400
184-IC6H4CN1bHC≡CCH2OH4b0.135l, 83830
19C6H5Br1dHC≡CCH2OH4b0.1965k, 10100
204-BrC6H4CN1eHC≡CCH2OH4b0.1245l, 71710
214-BrC6H4COMe1fHC≡CCH2OH4b0.135m, 98980
224-BrC6H4COMe1fHC≡CCH2OH4b0.01485m, 989800
234-BrC6H4NO21gHC≡CCH2OH4b0.1245n, 99990
244-BrC6H4Cl1hHC≡CCH2OH4b0.1725o, 15150
252-Bromothiophene1jHC≡CCH2OH4b0.1485p, 18180
263-Bromothiophene1kHC≡CCH2OH4b0.1965q, 880
272-Bromopyridine1lHC≡CCH2OH4b0.135r, 81810
283-Bromopyridine1mHC≡CCH2OH4b0.1485s, 65650
29C6H5I1aHC≡CCH2CH2OH4c0.165t, 78780
304-MeOC6H4I1cHC≡CCH2CH2OH4c0.1125u, 45450
314-BrC6H4CN1eHC≡CCH2CH2OH4c0.1125v, 99990
324-BrC6H4COMe1fHC≡CCH2CH2OH4c0.1125w, 89890
332-Bromothiophene1jHC≡CCH2CH2OH4c0.1125x, 32320
342-Bromopyridine1lHC≡CCH2CH2OH4c0.1125y, 64640
a Reaction conditions: [1]:[4a or 4c] = 1:1.1; [1]:[4b] = 1:1.5; [Pd]:[CuI]:[PPh3] = 1:2:2; Et3N was used as the solvent and base at 90 °C. b Isolated yields.
We also studied the reactivity of propargyl alcohol (4b) with aryl and heteroaryl halides in the presence of 0.1–0.01 mol% of catalyst (Entries 16–28). Generally, the reaction rates for the coupling of aryl and heteroaryl halides with 4b were slower than those for coupling with 4a, and reaction of aryl iodides with 4b at 90 °C gave the desired products in high yields (Entries 16–18). The use of 1d afforded only 10% of product under the reaction conditions described (Entry 19), while for electron-poor aryl bromides, excellent yields were obtained (Entries 20–23). In the case of 1f, using a 0.01 mol% catalyst loading, a TON of 9,800 was achieved, which is comparable with the reported efficiency of homogeneous catalysts [17b,f,g] (Entry 22). On the other hand, coupling of 1h with 4b gave a yield of only 15% (Entry 24). As for heteroaryl halides, the use of bromothiophenes 1j and 1k did not provide the products in good yields (Entries 25 and 26), while with halopyridines, the coupling products were obtained in good to excellent yields (Entries 27 and 28). The coupling of 3-butyn-1-ol (4c) with aryl and heteroaryl halides at a catalyst loading of 0.1 mol% was also screened, and the corresponding products were obtained in moderate to high yields (Entries 29–34).

2.4. Recycling and leaching studies of NS-MCM-41-Pd in the Sonogashira reaction

One of the purposes of designing this catalyst was to enable catalyst recycling for further use in subsequent reactions. In context, aryl iodides, activated aryl bromides, and several terminal alkynes were examined under optimized reaction conditions, and after completion of the initial cycle, the NS-MCM-41-Pd catalyst was extracted by centrifugation from the reaction mixture, washed successively with THF, H2O, and THF, and used for the next run with no regeneration treatment. The results using the recycled catalyst are shown in Table 4. In the case of aryl iodides, we found that the activity of the catalyst was completely retained after two recycled runs, giving an overall TON of between 2,450 and 2,960 (Table 4, Entries 1, 3, and 5). For the activated aryl bromides, the NS-MCM-41-Pd catalyst also exhibited high TONs in the recycled runs (Entries 2 and 4), but a gradual decrease in catalytic activity was observed in the reaction of 1f with 4b (Entry 6).
Table 4. Sonogashira coupling reaction catalyzed by recycled nanosized MCM-41-Pd.a
Table 4. Sonogashira coupling reaction catalyzed by recycled nanosized MCM-41-Pd.a
EntryAryl halideAlkyneSolvent/BaseT (°C)t (h)Yield, % b (TON)
Initial cycle1st recycle2nd recycle
11a2aEt3N/Et3N50399 (990)99 (990)98 (980)
21f2aNMP/Et3Nc90698 (980)93 (930)91 (910)
31a4aEt3N/Et3N90394 (940)90 (900)88 (880)
41f4aEt3N/Et3N90398 (980)99 (990)95 (950)
5d1b4bEt3N/Et3N90383 (830)82 (820)80 (800)
6d1f4bEt3N/Et3N90398 (980)88 (880)76 (760)
a Reaction conditions: [1]:[2 or 4]:[Pd]:[CuI]:[PPh3] = 1,000:1,100:1:2:2. b Isolated yields. c 3 equiv based on 1 was used as a base. d[1]:[4b]:[Pd]:[CuI]:[PPh3] = 1,000:1,500:1:2:2.
Several studies have successfully determined the amount of metal leaching using a hot-filtration technique, and this method was therefore used in this study to examine the activity of the catalyst with regards to metal leaching [45,46,47,48,56,57,58]. A reaction mixture of 1a with 2a in the above-described catalytic system was stirred at 50 °C for 30 min, resulting in a GC yield of 32%. The hot reaction mixture was then filtered through a dried Celite pad under nitrogen to remove the NS-MCM-41-Pd catalyst and any insoluble species, and the clear filtrate was introduced to another Schlenk tube at 50 °C. Further detection by GC demonstrated improvement of the yield to only 37% after 3 h, even in the presence of additional CuI and PPh3 (Figure 2). This result shows that no active species were dissolved in the solution to catalyze the coupling reaction. We further determined the Pd-content in the filtrate by ICP-MASS, and only 0.5 ppm of palladium was found in the solution, which indicated that the catalytic activity may mainly result from the grafted palladium complex. However, another pathway of the Sonogashira reaction is catalysis by a dissolved Pd-species that occurs inside the channels of NS-MCM-41, and this pathway cannot be excluded.
Figure 2. Plot of yield versus time with hot-filtration for 0.5 h of reaction at 50 °C (■) and a comparative reaction without hot-filtration (□). [1a]:[2a]:[Pd]:[CuI]:[PPh3] = 1000/1100/1/2/2.
Figure 2. Plot of yield versus time with hot-filtration for 0.5 h of reaction at 50 °C (■) and a comparative reaction without hot-filtration (□). [1a]:[2a]:[Pd]:[CuI]:[PPh3] = 1000/1100/1/2/2.
Molecules 15 09157 g002

3. Experimental

3.1. General

All reactions involving air- and moisture- sensitive conditions were carried out under a dry nitrogen atmosphere. N-Methylpyrrolidinone (NMP) was distilled under reduced pressure before use; Et3N and Bu3N were distilled from KOH; and toluene was distilled from sodium benzophenone ketyl. Aryl halides and terminal alkynes were purchased from ARCOS Co. Ltd and were used without further purification. 4,4’-Bis(bromomethyl)-2,2’-bipyridine [59,60], nanosized MCM-41 [61], and NS-MCM-41-Pd [55,56] were prepared according to the previously-published procedures. Melting points were recorded on melting point apparatus and were uncorrected. 1H- and 13C-NMR spectra were recorded in CDCl3 or C6D6 solution at 25 °C on a Varian 200 NMR spectrometer. GC analysis was performed on an SRI 8610C instrument equipped with a fused silica capillary column.

3.2. General procedure for the Sonogashira coupling

Under a nitrogen atmosphere, a mixture of NS-MCM-41-Pd (50 mg, 7.5 × 10-3 mmol-Pd), CuI (2.9 mg, 1.5 × 10-2 mmol), and PPh3 (3.9 mg, 1.5 × 10-2 mmol) in Et3N (15 mL) in a 50 mL Schlenk tube was charged with aryl halide (7.5 mmol) and terminal alkyne (8.3 mmol; in the case of 4b, 11.3 mmol was used), and the reaction mixture stirred at 50 °C or 90 °C. After cooling to room temperature, the resulting solution was passed through a short silica gel column with ethyl acetate as the eluent to remove ammonium salt. After evaporation of the solvent, column chromatography on silica gel afforded the desired product.
Diphenylacetylene (3a). White solid. m.p. 60–61 °C (lit.[47] 60–61 °C). 1H-NMR: δ 7.32–7.34 (m, 6H), 7.51–7.56 (m, 4H); 13C-NMR: δ 89.2 (2C), 122.9 (2C), 127.8 (4C), 127.9 (2C), 131.2 (4C).
4-(Phenylethynyl)benzonitrile (3b). Pale yellow solid. m.p. 108–110 °C (lit.[62] 106–108 °C). 1H-NMR: δ 7.35–7.38 (m, 3H), 7.51–7.55 (m, 2H), 7.60–7.61 (m, 4H); 13C-NMR: δ 87.6, 93.6, 111.1, 118.2, 121.8, 127.8, 128.1 (2C), 128.7, 131.3 (2C), 131.6 (2C), 131.6 (2C).
4-(Phenylethynyl)anisole (3c). Brown solid. m.p. 60–61 °C (lit.[62] 60.6 °C). 1H-NMR: δ 2.88 (s, 3H), 6.28–6.33 (m, 2H), 6.69–6.70 (m, 3H), 7.15–7.20 (m, 4H); 13C-NMR: δ 55.4, 87.9 (2C), 113.7 (2C), 115.0, 123.2, 127.5, 127.9 (2C), 131.0 (2C), 132.6 (2C), 158.9.
4-(Phenylethynyl)acetophenone (3d). Brown solid. m.p. 97–99 °C (lit.[47] 98–99 °C). 1H-NMR: δ 2.60 (s, 3H), 7.34–7.37 (m, 3H), 7.51–7.61 (m, 2H), 7.78–7.83 (m, 2H), 7.91–7.95 (m, 2H); 13C-NMR: δ 27.0, 88.5, 92.6, 122.3, 127.8, 127.9 (2C), 128.0 (2C), 128.4, 131.3 (2C), 132.0 (2C), 135.7, 196.4.
4-(Phenylethynyl)nitrobenzene (3e). Yellow solid. m.p. 116–117 °C (lit.[62] 114–116 °C). 1H-NMR: δ 7.34–7.39 (m, 3H) 7.52–7.56 (m, 2H), 7.63–7.66 (m, 2H), 8.19–8.22 (m, 2H); 13C-NMR: δ 87.4, 94.6, 121.7, 123.3, 128.1 (2C), 128.9 (2C), 129.8, 131.4 (2C), 131.8 (2C), 146.4.
Phenyl-(4-chlorophenyl)acetylene (3f). White solid. m.p. 82–83 °C (lit.[47] 82–83 °C). 1H-NMR: δ 7.30–7.36 (m, 5H), 7.43–7.46 (m, 2H), 7.51–7.53 (m, 2H); 13C-NMR: δ 88.1, 90.2, 121.4, 122.5, 128.0 (2C), 128.1, 128.3 (2C), 131.2 (2C), 132.4 (2C), 133.8.
2-(Theinylethynyl)benzene (3g). Pale yellow solid. m.p. 50–52 °C (lit.[63] 51–53 °C). 1H-NMR: δ 7.00–7.02 (m, 1H), 7.28–7.29 (m, 2H), 7.34–7.35 (m, 3H), 7.51–7.52 (m, 2H); 13C-NMR: δ 82.5, 92.9, 126.7, 126.8, 127.9 (2C), 128.0, 128.1, 131.0 (2C), 131.5, 132.0.
3-(Theinylethynyl)benzene (3h). Brown solid. m.p. 50–52 °C (lit.[64] 52–54 °C). 1H-NMR: δ 7.20–7.22 (m, 1H), 7.29–7.32 (m, 1H), 7.33–7.37 (m, 3H), 7.52–7.54 (m, 3H); 13C-NMR: δ 84.4, 88.8, 121.9, 122.8, 125.0, 127.8, 127.9 (2C), 128.2, 129.4, 131.1 (2C).
2-(Phenylethynyl)pyridine (3i) [65]. Colorless liquid. 1H-NMR: δ 7.15–7.18 (m, 1H), 7.29–7.32 (m, 3H), 7.45–7.47 (m, 1H), 7.54–7.62 (m, 3H), 8.55–8.57 (m, 1H); 13C-NMR: δ 88.4, 88.9, 121.7, 122.3, 126.6, 127.9, 128.4, 131.5, 135.6, 142.8, 149.3.
3-(Phenylethynyl)pyridine (3j). Yellow solid. m.p. 50–51 °C (lit.[66] 50–51 °C). 1H-NMR: δ 7.23–7.26 (m, 1H), 7.33–7.36 (m, 3H), 7.52–7.55 (m, 2H), 7.76–7.79 (m, 1H), 8.51–8.53 (m, 1H), 8.75–8.76 (m, 1H); 13C-NMR: δ 85.8, 92.4, 121.4, 120.0, 122.0, 122.6, 128.0 (2C), 128.3, 131.2 (2C), 137.8, 147.9, 151.6.
2-Methyl-4-phenyl-3-butyn-2-ol (5a). Yellow solid. m.p. 53–54 °C (lit.[67] 53.5–54.5 °C). 1H-NMR: δ 1.61 (s, 6H), 2.01 (s, 1H), 7.27–7.28 (m, 3H), 7.38–7.41 (m, 2H); 13C-NMR: δ 31.8 (2C), 65.7, 82.1, 104.1, 122.3, 127.8 (2C), 127.9, 131.2 (2C).
2-Methyl-4-(4’-cyano)phenyl-3-butyn-2-ol (5b) [68]. Yellow solid. m.p. 69–70 °C (lit.[69] 68.5–69.5 °C). 1H-NMR: δ 1.62 (s, 6H), 2.12 (s, 1H), 7.47 (d, J = 6.4 Hz, 2H), 7.57 (d, J = 6.4 Hz, 2H); 13C-NMR: δ 31.6 (2C), 65.7, 80.6, 98.0, 111.3, 118.1, 127.3, 131.5 (2C), 131.7 (2C).
2-Methyl-4-(4’-methoxy)phenyl-3-butyn-2-ol (5c) [70]. Yellow oil. 1H-NMR: δ 1.59 (s, 6H), 2.24 (s, 1H), 3.76 (s, 3H), 6.79 (d, J = 8.0 Hz, 2H), 7.31 (d, J = 8.0 Hz, 2H); 13C-NMR: δ 31.8 (2C), 55.3, 65.6, 81.8, 92.3, 113.5 (2C), 114.5, 132.5 (2C), 158.7.
2-Methyl-4-(4’-acetyl)phenyl-3-butyn-2-ol (5d) [70]. Yellow oil. 1H-NMR: δ 1.59 (s, 6H), 2.53 (s, 3H), 2.83 (s, 1H), 7.40 (d, J = 6.8 Hz, 2H), 7.81 (d, J = 6.8 Hz, 2H); 13C-NMR: δ 26.9, 31.5 (2C), 65.4, 81.0, 97.1, 127.3, 127.7 (2C), 131.2 (2C), 135.5, 196.6.
2-Methyl-4-(4’-nitro)phenyl-3-butyn-2-ol (5e). Brown solid. m.p. 100–102 °C (lit.[71] 102 °C). 1H-NMR: δ 1.62 (s, 6H), 2.09 (s, 1H), 7.49 (d, J = 7.6 Hz, 2H), 8.14 (d, J = 7.6 Hz, 2H); 13C-NMR: δ 31.5 (2C), 65.7, 80.4, 99.8, 123.1 (2C), 129.3, 131.9 (2C), 146.5.
2-Methyl-4-(4’-chloro)phenyl-3-butyn-2-ol (5f). White solid. m.p. 55–56 °C (lit.[72] 55–57 °C). 1H-NMR: δ 1.60 (s, 6H), 2.11 (s, 1H), 7.25 (d, J = 8.4 Hz, 2H), 7.31 (d, J = 8.4 Hz, 2H); 13C-NMR: δ 31.7 (2C), 65.7, 81.0, 94.5, 120.8, 128.2 (2C), 132.4 (2C), 133.8.
2-Methyl-4-(2-thienyl)-3-butyn-2-ol (5g). Off-white solid. m.p. 56–57 °C (lit.[73] 54 °C). 1H-NMR: δ 1.58 (s, 6H), 2.51 (s, 1H), 6.91–6.93 (m, 1H), 7.14–7.15 (m, 1H), 7.19–7.21 (m, 1H); 13C-NMR: δ 31.6 (2C), 65.7, 75.4, 97.3, 122.2, 126.5, 126.6, 131.5.
2-Methyl-4-(3-thienyl)-3-butyn-2-ol (5h). Brown solid. m.p. 54–56 °C (lit.[74] 56 °C). 1H-NMR: δ 1.61 (s, 6H), 2.12 (s, 1H), 7.08–7.09 (m, 1H), 7.24–7.26 (m, 1H), 7.41–7.42 (m, 1H); 13C-NMR: δ 31.8 (2C), 65.7, 77.3, 93.2, 124.9, 128.2, 129.4.
2-Methyl-4-(2-pyridyl)-3-butyn-2-ol (5i). Off-white solid. m.p. 60–61 °C (lit.[25] 61 °C). 1H-NMR: δ 1.59 (s, 6H), 2.95 (s, 1H), 7.11–7.15 (m, 1H), 7.29–7.32 (m, 1H), 7.52–7.57 (m, 1H), 8.47–8.49 (m, 1H); 13C-NMR: δ 31.4 (2C), 64.9, 80.9, 94.6, 122.4, 126.6, 135.5, 142.3, 148.9.
2-Methyl-4-(3-pyridyl)-3-butyn-2-ol (5j). Yellow solid. m.p. 55–56 °C (lit.[25] 53 °C). 1H-NMR: δ 1.58 (s, 6H), 2.00 (s, 1H), 7.18–7.22 (m, 1H), 7.64–7.67 (m, 1H), 8.44–8.45 (m, 1H), 8.71 (s, 1H); 13C-NMR: δ 31.6 (2C), 64.9, 78.1, 98.3, 120.0, 122.8, 138.4, 147.3, 151.3.
3-Phenyl-2-propyn-1-ol (5k) [75]. Yellow oil. 1H-NMR: δ 2.87 (s, 1H), 4.50 (s, 2H), 7.27–7.33 (m, 3H), 7.42–7.45 (m, 2H); 13C-NMR: δ 51.4, 85.3, 87.1, 123.1, 127.8 (2C), 127.9, 131.1 (2C).
3-(4’-Cyano)phenyl-2-propyn-1-ol (5l) [76]. Off-white solid. m.p. 89–91 °C (lit.[77] 87.5–88 °C). 1H-NMR: δ 2.01 (s, 1H), 4.50 (s, 2H), 7.48 (d, J = 6.4 Hz, 2H), 7.57 (d, J = 6.4 Hz, 2H); 13C-NMR: δ 51.6, 83.9, 91.6, 111.5, 118.0, 127.1, 131.5 (2C), 131.7 (2C).
4-(3-Hydroxy-1-propynyl)acetophenone (5m). Yellow solid. m.p. 80–81 °C (lit.[78] 80–81 °C). 1H-NMR: δ 2.55 (s, 3H), 2.80 (s, 1H), 4.90 (s, 2H), 7.43 (d, J = 8.4 Hz, 2H), 7.82 (d, J = 8.4 Hz, 2H); 13C-NMR: δ 26.9, 51.5, 84.4, 90.7, 127.1, 127.8 (2C), 131.2 (2C), 135.7, 196.8.
3-(4’-Nitro)phenyl-2-propyn-1-ol (5n). Yellow solid. m.p. 96–97 °C (lit.[79] 95–96.5 °C). 1H-NMR: δ 1.95 (s, 1H), 4.52 (s, 2H), 7.54 (d, J = 8.0 Hz, 2H), 8.15 (d, J = 8.0 Hz, 2H); 13C-NMR: δ 51.6, 83.7, 92.4, 123.2 (2C), 129.0, 131.9 (2C), 146.6.
3-(4’-Chlorophenyl)-2-propyn-1-ol (5o). Yellow solid. m.p. 77–79 °C (lit.[72] 78.5–79 °C). 1H-NMR: δ 2.11 (s, 1H), 4.47 (s, 2H), 7.25 (d, J = 8.8 Hz, 2H), 7.33 (d, J = 8.8 Hz, 2H); 13C-NMR: δ 51.7, 84.5, 88.0, 120.6, 128.2 (2C), 132.4 (2C), 134.1.
3-(2’-Thiophenyl)-2-propyn-1-ol (5p) [64]. Pale yellow oil. 1H-NMR: δ 2.17 (s, 1H), 4.50 (s, 2H), 6.95–6.97 (m, 1H), 7.20–7.21 (m, 1H), 7.25–7.26 (m, 1H); 13C-NMR: δ 51.8, 78.8, 91.0, 122.0, 126.6, 127.0, 131.9.
3-(3’-Thiophenyl)-2-propyn-1-ol (5q) [54]. Brown oil. 1H-NMR: δ 3.06 (s, 1H), 4.46 (s, 2H), 7.06–7.08 (m, 1H), 7.19–7.24 (m, 1H), 7.41–7.42 (m, 1H); 13C-NMR: δ 51.4, 80.6, 86.8, 121.1, 124.9, 128.6, 129.3.
3-(2-Pyridyl)-2-propyn-1-ol (5r). White solid. m.p. 83–84 °C (lit.[80] 82 °C). 1H-NMR: δ 2.43 (s, 1H), 4.53 (s, 2H), 7.19–7.22 (m, 1H), 7.39–7.41 (m, 1H), 7.60–7.64 (m, 1H), 8.50–8.51 (m, 1H); 13C-NMR: δ 51.3, 84.1, 88.3, 122.7, 126.8, 136.0, 142.2, 149.2.
3-(3-Pyridyl)-2-propyn-1-ol (5s). White solid. m.p. 101–102 °C (lit.[81] 99–100 °C). 1H-NMR: δ 2.08 (s, 1H), 4.49 (s, 2H), 7.23–7.36 (m, 1H), 7.70–7.74 (m, 1H), 8.47–8.48 (m, 1H), 8.74–8.75 (m, 1H); 13C-NMR: δ 51.0, 81.3, 92.0, 119.9, 122.9, 138.5, 147.6, 151.4.
4-Phenyl-3-butyn-1-ol (5t) [82]. Light brown oil. 1H-NMR: δ 1.84 (t, J = 6.2 Hz, 1H), 2.70 (t, J = 6.2 Hz, 2H), 3.82 (q, J = 6.2 Hz, 2H), 7.26–7.32 (m, 3H), 7.39–7.44 (m, 2H); 13C-NMR: δ 23.5, 60.9, 82.1, 86.5, 123.3, 127.7, 128.1 (2C), 131.5 (2C).
4-(4’-Methoxy)phenyl-3-butyn-1-ol (5u). Pale yellow solid. m.p. 58–59 °C (lit.[83] 61 °C). 1H-NMR: δ 1.90 (br, 1H), 2.65 (t, J = 6.1 Hz, 2H), 3.77 (t, J = 6.1 Hz, 2H), 6.80 (d, J = 8.6 Hz, 2H), 7.32 (d, J = 8.6 Hz, 2H); 13C-NMR: δ 23.6, 55.1, 61.1, 81.9, 84.8, 113.7 (2C), 115.4, 132.8 (2C), 159.1.
4-(4’-Cyano)phenyl-3-butyn-1-ol (5v) [53]. Pale yellow solid. m.p. 80–81 °C. 1H-NMR: δ 1.97 (br, 1H), 2.69 (t, J = 6.2 Hz, 2H), 3.81 (t, J = 6.2 Hz, 2H), 7.44 (d, J = 8.4 Hz, 2H), 7.54 (d, J = 8.4 Hz, 2H); 13C-NMR: δ 23.6, 60.6, 80.6, 91.8, 110.8, 118.3, 128.4, 131.7 (2C), 132.0 (2C).
4-(4-Hydroxy-1-butynyl)acetophenone (5w). Pale yellow solid. m.p. 75–77 °C (lit.[53] 74–76 °C). 1H-NMR: δ 2.00 (br, 1H), 2.56 (s, 3H), 2.69 (t, J = 6.2 Hz, 2H), 3.81 (t, J = 6.2 Hz, 2H), 7.45 (d, J = 8.2 Hz, 2H), 7.85 (d, J = 8.2 Hz, 2H); 13C-NMR: δ 23.7, 26.4, 60.8, 81.4, 90.5, 128.0 (2C), 128.4, 131.6 (2C), 135.7, 197.4.
4-(2-Thiophenyl)-3-butyn-1-ol (5x) [54]. Light brown oil. 1H-NMR: δ 1.93 (br, 1H), 2.69 (t, J = 6.2 Hz, 2H), 3.79 (t, J = 6.2 Hz, 2H), 6.92 (dd, J = 5.1, 3.6 Hz, 1H), 7.14 (d, J = 3.6 Hz, 1H), 7.18 (d, J = 5.2 Hz, 1H); 13C-NMR: δ 24.0, 60.9, 75.4, 90.5, 123.3, 126.2, 126.7, 131.4.
4-(2-Pyridyl)-3-butyn-1-ol (5y) [25]. Light brown oil. 1H NMR: δ 2.72 (t, J = 6.0 Hz, 2H), 3.06 (br, 1H), 3.87 (t, J = 6.0 Hz, 2H), 7.18–7.25 (m, 1H), 7.39 (d, J = 7.8 Hz, 1H), 7.64 (td, J = 7.8, 2.0 Hz, 1H), 8.54 (d, J = 5.0 Hz, 1H); 13C-NMR: δ 23.3, 59.9, 80.7, 88.4, 122.2, 126.4, 136.0, 142.7, 148.8.

3.3. General procedure for recycling of nanosized MCM-41-Pd

Under a nitrogen atmosphere, a 50 mL Schlenk tube was charged with NS-MCM-41-Pd (50 mg, 7.5 × 10-3 mmol-Pd), CuI (2.9 mg, 1.5 × 10-2 mmol), PPh3 (3.9 mg, 1.5 × 10-2 mmol), Et3N (15 mL), aryl halide (7.5 mmol), and terminal alkyne (8.3 mmol; in the case of 4b, 11.3 mmol was used). The mixture was stirred at 50 °C for 3 h (6 h for Entry 2) and then cooled to room temperature. Recovery of NS-MCM-41-Pd was achieved by centrifugation and successive washes with THF, H2O, and THF (2 × 40 mL each washing). The solid was then dried under vacuum overnight and used for the next run.

4. Conclusions

In conclusion, NS-MCM-41-Pd is a highly efficient and recyclable catalyst for the coupling of a wide variety of aryl and heteroaryl halides with terminal alkynes, requiring catalyst loadings as low as 0.01 mol% for a single run. The NS-MCM-41-Pd catalyst also exhibited excellent reusability when a catalyst loading of only 0.1 mol% was employed for the recycling studies. The results of this study demonstrate the usefulness of anchored palladium bipyridyl complex on mesoporous silica as a heterogeneous catalyst in cross-coupling reactions.

Acknowledgements

This research was financially supported by the National Science Council of Taiwan (NSC95-2113-M-027-001).
  • Sample Availability: Samples of the compounds are available from the authors.

References and Notes

  1. Sonogashira, K.; Tohda, Y.; Hagihara, N. A convenient synthesis of acetylenes: catalytic substitutions of acetylenic hydrogen with bromoalkenes, iodoarenes and bromopyridines. Tetrahedron Lett. 1975, 16, 4467–4470. [Google Scholar] [CrossRef]
  2. Moore, J.S. Shape-persistent molecular architectures of nanoscale dimension. Acc. Chem. Res. 1997, 30, 402–413. [Google Scholar] [CrossRef]
  3. Sonogashira, K. Metal-Catalyzed Cross-Coupling Reactions; Diederich, F., Stang, P.J., Eds.; Wiley-VCH: Weinheim, Germany, 1998; pp. 203–209. [Google Scholar]
  4. Brandsma, L.; Vasilevsky, S.F.; Verkruijsse, H.D. Application of Transition Metal Catalysts in Organic Synthesis; Springer: Berlin, Germany, 1998; pp. 179–225. [Google Scholar]
  5. Sonogashira, K. Handbook of Organopalladium Chemistry for Organic Synthesis; Negishi, E., de Meijere, A., Eds.; Wiley-VCH: New York, NY, USA, 2002; p. 493. [Google Scholar]
  6. Sonogashira, K. Development of Pd-Cu catalyzed cross-coupling of terminal acetylenes with sp2-carbon halides. J. Organomet. Chem. 2002, 653, 46–49. [Google Scholar] [CrossRef]
  7. Negishi, E.; Anastasia, L. Palladium-catalyzed alkynylation. Chem. Rev. 2003, 103, 1979–2017. [Google Scholar] [CrossRef]
  8. Taylor, E.C.; Dowling, J.E. Replacement of the 1′,4′-phenylene region in 5,10-dideaza-5,6,7,8-tetrahydrofolic acid (DDATHF) by 4,5,6,7-tetrahydrobenzo[c]thiophene and 4,5,6,7-tetrahydroisobenzofuran Nuclei. J. Org. Chem. 1997, 62, 1599–1603. [Google Scholar] [CrossRef]
  9. Nakamura, H.; Aizawa, M.; Takeuchi, D.; Murai, A.; Shimoura, O. Convergent and short-step syntheses of dl-Cypridina luciferin and its analogues based on Pd-mediated cross couplings. Tetrahedron Lett. 2000, 41, 2185–2188. [Google Scholar] [CrossRef]
  10. Liu, T.-Z.; Isobe, M. Synthesis of the H-I-J tricyclic fragment of ciguatoxin, a marine polyether toxin. Synlett 2000, 266–268. [Google Scholar]
  11. de Kort, M.; Correa, V.; Valentijn, A.R.P.M.; van der Marel, G.A.; Potter, B.V.L.; Taylor, C.W.; van Boom, J.H. Synthesis of potent agonists of the D-myo-inositol 1,4,5-trisphosphate receptor based on clustered disaccharide polyphosphate analogues of adenophostin A. J. Med. Chem. 2000, 43, 3295–3303. [Google Scholar] [CrossRef]
  12. Amiet, G.; Hügel, H.M.; Nurlawis, F. The synthesis of the kynurenamines K1 and K2, metabolites of melatonin. Synlett 2002, 495–497. [Google Scholar]
  13. Cosford, N.D.P.; Tehrani, L.; Roppe, J.; Schweiger, E.; Smith, N.D.; Anderson, J.; Bristow, L.; Brodkin, J.; Jiang, X.; McDonald, I.; Rao, S.; Washburn, M.; Varney, M.A. 3-[(2-Methyl-1,3-thiazol-4-yl)ethynyl]-pyridine: A potent and highly selective metabotropic glutamate subtype 5 receptor antagonist with anxiolytic activity. J. Med. Chem. 2003, 46, 204–206. [Google Scholar]
  14. Nicolaou, K.C.; Dai, W.-M. Chemistry and biology of the enediyne anticancer antibiotics. Angew. Chem., Int. Ed. Engl. 1991, 30, 1387–1416. [Google Scholar]
  15. Yoshimura, F.; Kawata, S.; Hirama, M. Synthetic study of kedarcidin chromophore: Atropselective construction of the ansamacrolide. Tetrahedron Lett. 1999, 40, 8281–8285. [Google Scholar] [CrossRef]
  16. Toyota, M.; Komori, C.; Ihara, M. A concise formal total synthesis of mappicine and nothapodytine B via an intramolecular hetero Diels-Alder reaction. J. Org. Chem. 2000, 65, 7110–7113. [Google Scholar] [CrossRef]
  17. Paterson, I.; Davies, R.D.M.; Marquez, R. Total synthesis of the callipeltoside Aglycon. Angew. Chem., Int. Ed. 2001, 40, 603–607. [Google Scholar]
  18. Kiehl, A.; Müllen, K. Catalysis in Precision Polymerization; Kobayashi, S., Ed.; Wiley-VCH: Chichester, England, 1997; pp. 162–169. [Google Scholar]
  19. Bunz, U.H.F. Poly(aryleneethynylene)s: Syntheses, properties, structures, and applications. Chem. Rev. 2000, 100, 1605–1644. [Google Scholar] [CrossRef]
  20. Höger, S.; Rosselli, S.; Ramminger, A.-D.; Enkelmann, V. A facile synthesis of large extraannular-functionalized phenyl-ethynyl macrocycles containing m-terphenyl units. Org. Lett. 2002, 4, 4269–4272. [Google Scholar] [CrossRef]
  21. Li, C.-J.; Slaven, W.T., IV; John, V.T.; Banerjee, S. Palladium catalysed polymerization of aryl diodides with acetylene gas in aqueous medium: A novel synthesis of areneethynylene polymers and oligomers. Chem. Commun. 1997, 1569–1570. [Google Scholar]
  22. Sonogashira, K. Comprehensive Organic Synthesis; Trost, B.M., Fleming, I., Eds.; Pergamon: New York, NY, USA, 1991; Volume 3, pp. 521–549. [Google Scholar]
  23. Choudary, B.M.; Madhi, S.; Chowdari, N.S.; Kantam, M.L.; Sreedhar, B. Layered double hydroxide supported nanopalladium catalyst for Heck-, Suzuki-, Sonogashira-, and Stille-type coupling reactions of chloroarenes. J. Am. Chem. Soc. 2002, 124, 14127–14136. [Google Scholar] [CrossRef]
  24. Heidenreich, R.G.; Köhler, K.; Krauter, J.G.E.; Pietsch, J. Pd/C as a highly active catalyst for Heck, Suzuki and Sonogashira reactions. Synlett 2002, 1118–1122. [Google Scholar]
  25. Novák, Z.; Szabó, A.; Répási, J.; Kotschy, A. Sonogashira coupling of aryl halides catalyzed by palladium on charcoal. J. Org. Chem. 2003, 68, 3327–3329. [Google Scholar] [CrossRef]
  26. Komáromi, A.; Novák, Z. Efficient copper-free Sonogashira coupling of aryl chlorides with palladium on charcoal. Chem. Commun. 2008, 4968–4970. [Google Scholar]
  27. Mori, S.; Yanase, T.; Aoyagi, S.; Monguchi, Y.; Maegawa, T.; Sajiki, H. Ligand-free sonogashira coupling reactions with heterogeneous Pd/C as the catalyst. Chem. Eur. J. 2008, 14, 6994–6999. [Google Scholar] [CrossRef]
  28. Reddy, E.A.; Barange, D.K.; Islam, A.; Mukkanti, K.; Pal, M. Synthesis of 2-alkynylquinolines from 2-chloro and 2,4-dichloroquinoline via Pd/C-catalyzed coupling reaction in water. Tetrahedron 2008, 64, 7143–7150. [Google Scholar] [CrossRef]
  29. Shang, H.; Hua, R.; Zheng, Q.; Zhang, J.; Liang, X.; Zhu, Q. An improved practical Pd/C-catalyzed Sonogashira cross-coupling reaction for the synthesis of liquid crystals of trans-cyclohexyltolans. Appl. Organomet. Chem. 2010, 24, 473–476. [Google Scholar]
  30. Duplais, C.; Forman, A.J.; Baker, B.A.; Lipshutz, B.H. UC Pd: A new form of PdVC for sonogashira couplings. Chem. Eur. J. 2010, 16, 3366–3371. [Google Scholar] [CrossRef]
  31. Li, P.; Wang, L.; Li, H. Application of recoverable nanosized palladium(0) catalyst in Sonogashira reaction. Tetrahedron 2005, 61, 8633–8640. [Google Scholar] [CrossRef]
  32. Chouzier, S.; Gruber, M.; Djakovitch, L. New hetero-bimetallic Pd-Cu catalysts for the one-pot indole synthesis via the Sonogashira reaction. J. Mol. Catal. A: Chem. 2004, 212, 43–52. [Google Scholar]
  33. Djakovitch, L.; Rollet, P. Sonogashira cross-coupling reactions catalysed by copper-free palladium zeolites. Adv. Synth. Catal. 2004, 346, 1782–1792. [Google Scholar] [CrossRef]
  34. Djakovitch, L.; Rollet, P. Sonogashira cross-coupling reactions catalysed by heterogeneous copper-free Pd-zeolites. Tetrahedron Lett. 2004, 45, 1367–1370. [Google Scholar] [CrossRef]
  35. Tyrrell, E.; Whiteman, L.; Williams, N. Sonogashira cross-coupling reactions and construction of the indole ring system using a robust, silica-supported palladium catalyst. Synthesis 2009, 829–835. [Google Scholar]
  36. Anwander, R. SOMC@PMS. Surface organometallic chemistry at periodic mesoporous silica. Chem. Mater. 2001, 13, 4419–4438. [Google Scholar] [CrossRef]
  37. Biffis, A.; Zecca, M.; Basato, M. Palladium metal catalysts in Heck C-C coupling reactions. J. Mol. Catal. A: Chem. 2001, 173, 249–274. [Google Scholar]
  38. He, X.; Antonelli, D. Recent advances in synthesis and applications of transition metal containing mesoporous molecular sieves. Angew. Chem. Int. Ed. 2002, 41, 215–229. [Google Scholar]
  39. De Vos, D.E.; Dams, M.; Sels, B.F.; Jacobs, P.A. Ordered mesoporous and microporous molecular sieves functionalized with transition metal complexes as catalysts for selective organic transformations. Chem. Rev. 2002, 102, 3615–3640. [Google Scholar] [CrossRef]
  40. Trong On, D.; Desplantier-Giscard, D.; Danumah, C.; Kaliaguine, S. Perspectives in catalytic applications of mesostructured materials. Appl. Catal. A: Gen. 2003, 253, 545–602. [Google Scholar]
  41. Thomas, J.M.; Raja, R. Catalytic significance of organometallic compounds immobilized on mesoporous silica: Economically and environmentally important examples. J. Organomet. Chem. 2004, 689, 4110–4124. [Google Scholar] [CrossRef]
  42. Li, C. Chiral synthesis on catalysts immobilized in microporous and mesoporous materials. Catal. Rev. 2004, 46, 419–492. [Google Scholar] [CrossRef]
  43. Taguchi, A.; Schüth, F. Ordered mesoporous materials in catalysis. Microporous Mesoporous Mater. 2005, 77, 1–45. [Google Scholar] [CrossRef]
  44. Hoffmann, F.; Cornelius, M.; Mprell, J.; Fröba, M. Silica-based mesoporous organic-inorganic hybrid materials. Angew. Chem., Int. Ed. 2006, 45, 3216–3251. [Google Scholar]
  45. Rollet, P.; Kleist, W.; Dufaud, V.; Djakovitch, L. Copper-free heterogeneous catalysts for the Sonogashira cross-coupling reaction: Preparation, characterisation, activity and applications for organic synthesis. J. Mol. Catal. A: Chem. 2005, 241, 39–51. [Google Scholar]
  46. Cai, M.; Xu, Q.; Wang, P. A novel MCM-41-supported sulfur palladium(0) complex catalyst for Sonogashira coupling reaction. J. Mol. Catal. A: Chem. 2006, 250, 199–202. [Google Scholar]
  47. Cai, M.; Sha, J.; Xu, Q. MCM-41-supported bidentate phosphine palladium(0) complex: a highly active and recyclable catalyst for the Sonogashira reaction of aryl iodides. Tetrahedron 2007, 63, 4642–4647. [Google Scholar] [CrossRef]
  48. Cai, M.; Xu, Q.; Sha, J. Copper-free Sonogashira coupling reaction catalyzed by MCM-41-supported thioether palladium(0) complex in water under aerobic conditions. J. Mol. Catal. A: Chem. 2007, 272, 293–297. [Google Scholar]
  49. Alonso, D.A.; Nájera, C.; Pacheco, M.C. A copper- and amine-free Sonogashira-type coupling procedure catalyzed by oxime palladacycles. Tetrahedron Lett. 2002, 43, 9365–9368. [Google Scholar] [CrossRef]
  50. McGuinness, D.S.; Cavell, K.J. Donor-Functionalized Heterocyclic Carbene Complexes of Palladium(II): Efficient Catalysts for C-C Coupling Reactions. Organometallics 2000, 19, 741–746. [Google Scholar] [CrossRef]
  51. Buchmeiser, M.R.; Schareina, T.; Kempe, R.; Wurst, K. Bis(pyrimidine)-based palladium catalysts: Synthesis, X-ray structure and applications in Heck-, Suzuki-, Sonogashira-Hagihara couplings and amination reactions. J. Organomet. Chem. 2001, 643, 39–46. [Google Scholar]
  52. Feuerstein, M.; Berthiol, F.; Doucet, H.; Santelli, M. Palladium-tetraphosphine complex: An efficient catalyst for the coupling of aryl halides with alkynes. Org. Biomol. Chem. 2003, 1, 2235–2237. [Google Scholar] [CrossRef]
  53. Feuerstein, M.; Doucet, H.; Santelli, M. Coupling reactions of aryl bromides with 1-alkynols catalysed by a tetraphosphine/palladium catalyst. Tetrahedron Lett. 2004, 45, 1603–1606. [Google Scholar] [CrossRef]
  54. Feuerstein, M.; Doucet, H.; Santelli, M. Sonogashira cross-coupling reactions with heteroaryl halides in the presence of a tetraphosphine-palladium catalyst. Tetrahedron Lett. 2005, 46, 1717–1720. [Google Scholar] [CrossRef]
  55. Tsai, F.-Y.; Wu, C.-L.; Mou, C.-Y.; Chao, M.-C.; Lin, H.-P.; Liu, S.-T. Palladium bipyridyl complex anchored on nanosized MCM-41 as a highly efficient and recyclable catalyst for Heck reaction. Tetrahedron Lett. 2004, 45, 7503–7506. [Google Scholar] [CrossRef]
  56. Tsai, F.-Y.; Lin, B.-N.; Chen, M.-J.; Mou, C.-Y.; Liu, S.-T. Anchored palladium bipyridyl complex in nanosized MCM-41: a recyclable and efficient catalyst for the Kumada-Corriu reaction. Tetrahedron 2007, 63, 4304–4309. [Google Scholar] [CrossRef]
  57. Chen, J.-Y.; Chen, S.-C.; Tang, Y.-J.; Mou, C.-Y.; Tsai, F.-Y. Coupling of acyl chlorides with triarylbismuths catalyzed by palladium bipyridyl complex anchored on nanosized MCM-41: A recyclable and atom-efficient catalytic process for the synthesis of diaryl and alkyl aryl ketones. J. Mol. Catal. A: Chem. 2009, 307, 88–92. [Google Scholar]
  58. Chen, J.-Y.; Lin, T.-C.; Chen, S.-C.; Chen, A.-J.; Mou, C.-Y.; Tsai, F.-Y. Highly-efficient and recyclable nanosized MCM-41 anchored palladium bipyridyl complex-catalyzed coupling of acyl chlorides and terminal alkynes for the formation of ynones. Tetrahedron 2009, 65, 10134–10141. [Google Scholar] [CrossRef]
  59. Oki, A.R.; Morgan, R.J. An efficient preparation of 4,4'-dicarboxy-2,2'-bipyridine. Synth. Commun. 1995, 25, 4093–4097. [Google Scholar] [CrossRef]
  60. Will, G.; Boschloo, G.; Rao, S.N.; Fitzmaurice, D. Potentiostatic modulation of the lifetime of light-induced charge separation in a heterosupermolecule. J. Phys. Chem. B 1999, 103, 8067–8079. [Google Scholar]
  61. Lin, H.-P.; Tsai, C.-P. Synthesis of Mesoporous Silica Nanoparticles from a Low-concentration CnTMAX-Sodium Silicate Components. Chem. Lett. 2003, 32, 1092–1093. [Google Scholar] [CrossRef]
  62. Elangovan, A.; Wang, Y.-H.; Ho, T.-I. Sonogashira coupling reaction with diminished homocoupling. Org. Lett. 2003, 5, 1841–1844. [Google Scholar] [CrossRef]
  63. Kabalka, G.W.; Wang, L.; Pagni, R.M. Sonogashira coupling and cyclization reactions on alumina: A route to aryl alkynes, 2-substituted-benzo[b]furans and 2-substituted-indoles. Tetrahedron 2001, 57, 8017–8028. [Google Scholar] [CrossRef]
  64. Van den Hoven, B.G.; Alper, H. The first regioselective hydroformylation of acetylenic thiophenes catalyzed by a zwitterionic rhodium complex and triphenyl phosphate. J. Org. Chem. 1999, 64, 9640–9645. [Google Scholar] [CrossRef]
  65. Sørensen, U.S.; Pombo-Villar, E. Copper-free palladium-catalyzed sonogashira-type coupling of aryl halides and 1-aryl-2-(trimethylsilyl)acetylenes. Tetrahedron 2005, 61, 2697–2703. [Google Scholar] [CrossRef]
  66. Novák, Z.; Nemes, P.; Kotschy, A. Tandem Sonogashira coupling: An efficient tool for the synthesis of diarylalkynes. Org. Lett. 2004, 6, 4917–4920. [Google Scholar] [CrossRef]
  67. Pal, M.; Subramanian, V.; Parasuraman, K.; Yeleswarapu, K.R. Palladium catalyzed reaction in aqueous DMF: Synthesis of 3-alkynyl substituted flavones in the presence of prolinol. Tetrahedron 2003, 59, 9563–9570. [Google Scholar] [CrossRef]
  68. Kayaki, Y.; Yamamoto, M.; Ikariya, T. Stereoselective formation of α-alkylidene cyclic carbonates via carboxylative cyclization of propargyl alcohols in supercritical carbon dioxide. J. Org. Chem. 2007, 72, 647–649. [Google Scholar] [CrossRef]
  69. Pourjavadi, A.; Marandi, G.B. Preparation of conjugated enynes and arylacetylenic compounds from arylalkynols using alumina in dry media. J. Chem. Res. 2002, 11, 552–555. [Google Scholar] [CrossRef]
  70. Adjabeng, G.; Brenstrum, T.; Frampton, C.S.; Robertson, A.J.; Hillhouse, J.; McNulty, J.; Capretta, A. Palladium complexes of 1,3,5,7-tetramethyl-2,4,8-trioxa-6-phenyl-6- phosphaadamantane: Synthesis, crystal structure and use in the Suzuki and Sonogashira reactions and the α-arylation of ketones. J. Org. Chem. 2004, 69, 5082–5086. [Google Scholar]
  71. Kondo, K.; Fujitani, T.; Ohnishi, N. Synthesis and non-linear properties of disubstituted diphenylacetylene and related compounds. J. Mater. Chem. 1997, 7, 429–433. [Google Scholar] [CrossRef]
  72. Alonso, D.A.; Najera, C.; Pacheco, M.C. C(sp2)-C(sp) and C(sp)-C(sp) coupling reactions catalyzed by oxime-derived palladacycles. Adv. Synth. Catal. 2003, 345, 1146–1158. [Google Scholar] [CrossRef]
  73. Sarkar, A.; Talwar, S.S. Heteroaryl functionalised diacetylenes: Preparation and solid-state reactivity. J. Chem. Soc., Perkin Trans. 1 1998, 4141–4146. [Google Scholar]
  74. Sarkar, A.; Manjunath, S.P.; Kamath, B.; Bhagwat, L.; Babu, K.N.; Rajalakshmi, K.; Talwar, S.S. A convenient method for the preparation of thienylacetylnes. Indian J. Chem. Sect. B 1991, 30, 360–362. [Google Scholar]
  75. Ishikawa, T.; Mizuta, T.; Hagiwara, K.; Aikawa, T.; Kudo, T.; Saito, S. Catalytic alkynylation of ketones and aldehydes using quaternary ammonium hydroxide base. J. Org. Chem. 2003, 68, 3702–3705. [Google Scholar] [CrossRef]
  76. Bernini, R.; Cacchi, S.; Fabrizi, G.; Forte, G.; Petrucci, F.; Prastaro, A.; Niembro, S.; Shafir, A.; Vallribera, A. Alkynylation of aryl halides with perfluoro-tagged palladium nanoparticles immobilized on silica gel under aerobic, copper- and phosphine-free conditions in water. Org. Biomol. Chem. 2009, 7, 2270–2273. [Google Scholar] [CrossRef]
  77. Bumagin, N.A.; Ponomaryov, A.B.; Beletskaya, I.P. A convenient synthesis of substituted propargyl alcohols and terminal acetylenes. Synthesis 1984, 9, 728–729. [Google Scholar]
  78. Batey, R.A.; Shen, M.; Lough, A.J. Carbamoyl-substituted N-heterocyclic carbene complexes of palladium(II): Application to Sonogashira cross-coupling reactions. Org. Lett. 2002, 4, 1411–1414. [Google Scholar] [CrossRef]
  79. Harris, M.A.; McMillan, I.; Nayler, J.H.C.; Osborne, N.F.; Pearson, M.J.; Southgate, R. Syntheses based on 1,2-secopenicillins. Part II. Preparation of 4-(3-substituted prop-2-ynylthio)azetidin-2-ones. J. Chem. Soc., Perkin Trans. 1 1976, 1612–1615. [Google Scholar]
  80. Al-Arnaout, A.; Courtois, G.; Miginiac, L. Synthèse régiosélective par voie organométallique de pyridines, 4-picolines et 3,5-lutidines substituées en 2 par un groupe insaturé et/ou fonctionnel. J. Organomet. Chem. 1987, 333, 139–154. [Google Scholar]
  81. Bleicher, L.S.; Cosford, N.D.P.; Herbaut, A.; McCallum, J.S.; McDonald, I.A. A practical and efficient synthesis of the selective neuronal acetylcholine-gated ion channel agonist (S)-(-)-5-ethynyl-3-(1-methyl-2-pyrrolidinyl)pyridine maleate (SIB-1508Y). J. Org. Chem. 1998, 63, 1109–1118. [Google Scholar] [CrossRef]
  82. Kim, I.S.; Dong, G.R.; Jung, Y.H. Palladium(II)-catalyzed isomerization of olefins with tributyltin hydride. J. Org. Chem. 2007, 72, 5424–5426. [Google Scholar] [CrossRef]
  83. Collins, C.J.; Hanack, M.; Stutz, H.; Auchter, G.; Schoberth, W. Vinyl cations. 41. Influence of 4-aryl and 4-alkyl substituents on the π-route solvolyses of homopropargyl esters. J. Org. Chem. 1983, 48, 5260–5268. [Google Scholar]

Share and Cite

MDPI and ACS Style

Lin, B.-N.; Huang, S.-H.; Wu, W.-Y.; Mou, C.-Y.; Tsai, F.-Y. Sonogashira Reaction of Aryl and Heteroaryl Halides with Terminal Alkynes Catalyzed by a Highly Efficient and Recyclable Nanosized MCM-41 Anchored Palladium Bipyridyl Complex. Molecules 2010, 15, 9157-9173. https://doi.org/10.3390/molecules15129157

AMA Style

Lin B-N, Huang S-H, Wu W-Y, Mou C-Y, Tsai F-Y. Sonogashira Reaction of Aryl and Heteroaryl Halides with Terminal Alkynes Catalyzed by a Highly Efficient and Recyclable Nanosized MCM-41 Anchored Palladium Bipyridyl Complex. Molecules. 2010; 15(12):9157-9173. https://doi.org/10.3390/molecules15129157

Chicago/Turabian Style

Lin, Bo-Nan, Shao-Hsien Huang, Wei-Yi Wu, Chung-Yuan Mou, and Fu-Yu Tsai. 2010. "Sonogashira Reaction of Aryl and Heteroaryl Halides with Terminal Alkynes Catalyzed by a Highly Efficient and Recyclable Nanosized MCM-41 Anchored Palladium Bipyridyl Complex" Molecules 15, no. 12: 9157-9173. https://doi.org/10.3390/molecules15129157

Article Metrics

Back to TopTop