Next Article in Journal
HS-SPME-GC-MS Analyses of Volatiles in Plant Populations—Quantitating Compound × Individual Matrix Effects
Next Article in Special Issue
Advances in Magnetic Nanoparticles-Supported Palladium Complexes for Coupling Reactions
Previous Article in Journal
Data-Driven Exploration of Selectivity and Off-Target Activities of Designated Chemical Probes
Previous Article in Special Issue
Development of Conjugate Addition of Lithium Dialkylcuprates to Thiochromones: Synthesis of 2-Alkylthiochroman-4-ones and Additional Synthetic Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Efficient Suzuki–Miyaura C-C Cross-Couplings Induced by Novel Heterodinuclear Pd-bpydc-Ln Scaffolds

1
Faculty of Chemical Engineering, Shenyang University of Chemical Technology, Shenyang 110142, China
2
Institute of Organic Chemistry, Romanian Academy, Spl. Independentei 202B, P.O.Box 35-108, 060023 Bucharest, Romania
*
Authors to whom correspondence should be addressed.
Molecules 2018, 23(10), 2435; https://doi.org/10.3390/molecules23102435
Submission received: 23 August 2018 / Revised: 16 September 2018 / Accepted: 20 September 2018 / Published: 23 September 2018
(This article belongs to the Special Issue Organometallic Catalysis in Organic Synthesis)

Abstract

:
An easy access to a series of previously unreported heterodinuclear Pd-Ln compounds, Pd-bpydc-La, Pd-bpydc-Ce and Pd-bpydc-Nd (bpydc = 2,2′-bipyridine-5,5′-dicarboxylate) has been developed. The Pd-Ln hybrid networks were effectively applied as catalysts in Suzuki–Miyaura C-C cross-coupling reactions of 4-bromoanisole and 4-bromobenzonitrile with phenylboronic acid, under mild conditions. A systematic investigation revealed Pd-bpydc-Nd as the most active catalyst. In all cases, reaction yields varied with the base, catalyst loading and substantially augmented with temperature (from 30 to 60 °C). Substituent effects were operative when changing from 4-bromoanisole to 4-bromobenzonitrile. The key role played by the lanthanides, aromatic substrate and base, in modulating the Pd-catalytic cycle has been highlighted. Importantly, the new catalysts proved to be stable in air and vs. functionalities and are quite efficient in Suzuki–Miyaura carbon-carbon bond formation conducted in protic solvents.

Graphical Abstract

1. Introduction

Palladium-catalyzed C-C bond construction [1,2,3,4,5,6,7,8] that makes use of readily available substrates and proceeds with high chemoselectivity has established itself as a flexible synthetic protocol for obtaining substituted arene and heteroarene derivatives [9,10,11,12,13] (e.g., symmetrical or unsymmetrical biaryls) [14]. This synthesis approach has revolutionized the production of advanced materials [15,16,17,18], pharmaceuticals [19,20,21], agrochemicals [21,22], liquid crystals [23], and natural or biologically active compounds [24,25,26,27,28], etc. Numerous attractive strategies have been developed to address the ample scope of this catalytic process including the utilization of palladium nanoparticles [29,30,31,32,33,34,35], or palladium immobilized on magnetic nanoparticles [36] and natural supports [37], use of nucleophilic carbene ligands (mostly NHCs) [38,39,40,41,42,43,44,45,46], Schiff bases [47,48], water-soluble ligands like poly(ethylene glycol)-functionalized N-heterocyclic carbenes [49], thiourea [50] or phosphines [51,52], induction by microwave (MW) acceleration [53,54], use of “greener” solvents, such as water [55,56,57,58,59,60,61,62,63,64,65] (activated by MW [66,67] or in catalysis under micellar conditions [68] for enhancing the solubility of the aromatic halide), water–DMF [69] or ionic liquids [70]. The reaction has long been the subject of vast research in transition-metal catalysis. Various transition metal complexes [71,72,73,74,75,76,77,78,79] (e.g., Ni, Au, Co, Cu, etc., as such or combined in pairs as heterobimetallic catalysts), or nonconventional Pd(II)-complexes with intricate organic ligands containing S or Se have been intensively explored [80,81]. Efficient transition metal-free catalysts have also been reported for Suzuki–Miyaura cross-coupling reactions [82].
Significantly, within the last few years, a new generation of heterogeneous catalysts based on metal–organic frameworks (MOFs) with an Ln-Pd heterobimetallic construction, has been developed and effectively applied to C-C cross-couplings. In such multifunctional configurations, the lanthanide nodes are interconnected by appropriate organic linkers to form rigid porous networks which attach PdCl2, prone to catalyze C-C bond formation. Thus, Jin and coworkers [83] ingeniously combined Pd nitrophilic metal cations, which coordinate diimine groups from the 2,2′-bipyridine-5,5′-dicarboxylate (bpydc) linker, with the Ln oxophilic rare earth metals (Ln = Sm, Eu, Gd or Tb) to build-up MOFs displaying a good heterogeneous catalytic activity for C-C cross-coupling reactions. With Ln(bpydc)PdCl2 embedded MOFs they performed the Suzuki–Miyaura reaction of selected aryl halides with arylboronic acid in toluene at 95 °C reporting appreciable yields in biaryls (82–97% yields, 4 h). More recently, we have developed a novel set of air- and water-stable heterobimetallic coordination complexes [Ln2Pd3(BPDC)2(HBPDC)2(μ2-O)Cl4(H2O)6·nH2O]m (Ln = Pr, Gd, Tb) binding the rare earth elements through the carboxylate groups of the heteroleptic organic ligand, 2,2′-bipyridine-4,4′-dicarboxylic acid and coordinating the nitrophilic Pd cations to the diimine groups of the difunctional organic linker [84]. By properly engineering the Pd network through the inclusion of lanthanides by intermediacy of a totally different organic vector, the newly created hybrid materials exhibited a reasonably robust and stable structural pattern with enhanced physical and chemical features which is dependent on the nature of the lanthanide incorporated. This type of catalyst demonstrated a high performance in eco-friendly and cost-effective Suzuki–Miyaura, Heck and Sonogashira reactions. Subsequently, we have extended this concept to the synthesis of heterobimetallic coordination complexes, [Ln2Pd2(BPDC)5/2(H2O)2·4H2O]n (Ln = Nd, Sm, Eu, Dy) employing the above bifunctional organic linker 2,2′-bipyridine-4,4′-dicarboxylic acid to associate Pd ions with rare earth elements [85]. The new class of Ln/Pd heterobimetallic coordination polymers offered a user-friendly methodology for application as catalysts in Heck and Suzuki–Miyaura cross-couplings under green reaction conditions (DMF-H2O, 1:1). In both chemical transformations, the new class of complexes displayed an excellent catalytic activity (yields of 99% in Heck reaction and 98% in Suzuki–Miyaura reaction) as compared to state of the art. Applying an alternative metallolig and strategy for MOF construction [86,87], a new set of heterobimetallic Pd/Ln MOFs with Sm, Eu, Tb, Dy, as lanthanide nodes, interconnected by a distinct bifunctional organic linker, 1,1′-di(p-carboxybenzyl)-2,2′-diimidazole, has been prepared [88]. This novel class of MOFs showed excellent stability in air and water and displayed high catalytic activity in Suzuki–Miyaura reactions conducted both homogeneously and heterogeneously, in neat water, ethanol or water–ethanol. Phase switchability was demonstrated when changing from water (homogeneous reaction) to ethanol (heterogeneous reaction) with beneficial consequences for practical applications. Such complexes were found to be pH-responsive, in a reversible way, enabling convenient recovery from acidic water solutions in view of recycling, as well as for environmentally friendly, final separation of metal residues.
With the aim of disclosing new catalytic abilities of rare-earth elements associated with palladium when they are incorporated within heterodinuclear Pd networks by means of 2,2′-bipyridine-5,5′-dicarboxylate (bpydc) as the heteroleptic organic linker, we have synthesized three novel dinuclear Pd-Ln MOF-type catalysts [Pd-bpydc-La (1), Pd-bpydc-Ce (2), Pd-bpydc-Nd (3)] by a simple one-pot procedure and successfully applied them in Suzuki–Miyaura reaction. We took advantage of the oxophilic propensity of the lanthanides to properly bind to the carboxylate groups of the bifunctional organic ligand to create a rigid network tethering Pd chloride at its nucleophilic nitrogen sites. Additionally, the electronic properties of the dipodal organic linker, 2,2′-bipyridine-5,5′-dicarboxylate, will influence the charge transfer between lanthanide and palladium active centers, tailoring, thus, the catalytic properties of the metal–organic framework. The final aim of these investigations lies in the evaluation of the catalytic properties of compounds 13 in heterogeneous cross-coupling of two model aryl bromides, i.e., 4-bromoanisole and 4-bromobenzonitrile, with phenylboronic acid, in methanol as a solvent, to give the corresponding biphenyl derivatives (Scheme 1).

2. Experimental Section

2.1. Materials and Synthesis Methodology

All reagents and solvents employed were of commercial grade, purchased from the suitable suppliers and were used without further purification. Powder XRD patterns were recorded with CuKα radiation by using a Bruker D8 Advance X-ray diffractometer. Elemental analyses were performed on a Perkin-Elmer elemental analyzer. Conversions were determined using Agilent 7890 GC-MSD. IR spectra were measured using a Nicolet IR-470 spectrometer for samples in KBr pellets.

2.2. Synthesis and Characterization of Pd-bpydc-Ln (13).

The heteronuclear bimetallic palladium compounds 13 have been synthesized in a one-pot procedure, according to the previously published methodology [88,89,90]. A mixture of K2PdCl4 (0.29 mmol), 2,2′-bipyridine-5,5′-dicarboxylic acid (0.29 mmol), Ln(NO3)3·6H2O (0.18 mmol) and water (10 mL) was stirred for 20 min in open air. The mixture was then transferred to a 23 mL Teflon reactor and kept at 100 °C for 24 h under autogenous pressure, thereafter, cooled to room temperature at a rate of 5 °C/h. The resulted yellow crystals were characterized by elemental analysis, powder XRD and infrared spectroscopy (KBr). Anal. Calcd. for C24 H35N4O19Cl4La1Pd2 (1): C, 24.49; H, 3.00; N, 4.76. Found: C, 24.41; H, 3.08; N, 4.80. IR (KBr)(cm−1): 3445, 3062, 1694, 1612, 1592, 1555, 1410, 1389, 1294, 1257, 1143, 1054, 858, 835, 778, 699, 661, 599, 415. Anal. Calcd for C24H35N4O19Ce1Cl4Pd2 (2): C, 24.46; H, 2.99; N, 4.75. Found: C, 24.38; H, 3.10; N, 4.71. IR (KBr)(cm−1): 3448, 3061, 1698, 1613, 1557, 1392, 1296, 1142, 1057, 1037, 858, 893, 776, 700, 661, 527, 419. Anal. Calcd. for C24H35N4O19Cl4Nd1Pd2 (3): C, 24.38; H, 2.98; N, 4.74. Found: C, 24.30; H, 3.10; N, 4.81. IR (KBr)(cm−1): 3458, 3071, 1711, 1623, 1567, 1402, 1302, 1161, 1065, 1042, 861, 888, 779, 706, 683, 568, 421. X-Ray analysis indicated that compounds 1–3 are isostructural and crystallize in a triclinic crystal system with space group P-1. Powder X-ray diffraction analyses of compounds 13 have been performed at room temperature. The patterns for 13 are in good agreement with the calculated data obtained from the single-crystal structures, confirming that the purities equal those of the single crystal samples.

2.3. General Procedure for Suzuki–Miyaura Reaction

A mixture of 4-bromoanisole (0.5 mmol) or 4-bromobenzonitrile (0.5 mmol), phenylboronic acid (0.6 mmol), specified base (0.5 mmol), methanol (1 mL), 0.2–0.5 mol% of catalyst was stirred at 30 °C or 60 °C in air. The progress of the reaction was monitored by withdrawing samples periodically which were analyzed by gas chromatography–mass spectrometry (GC-MS). GC-calculated yields were based on the amount of 4-bromoanisole or 4-bromobenzonitrile employed. At the end of the reaction, the catalyst was separated by simple filtration. The mixture was extracted with diethyl ether (20 mL), washed with water and dried over anhydrous Na2SO4. The solvent from the extract was completely removed with a rotary evaporator to obtain the product, which was further purified by column chromatography on silica gel. All coupling products were identified by Agilent 7890A-5975C GC-MS and 1H NMR spectroscopy.

3. Results and Discussion

With the new set of Pd-Ln compounds in hand, we focused first on the optimization studies of the cross-coupling reaction between 4-bromoanisole and phenylboronic acid as special model substrates (Scheme 1) striving to identify the best catalytic system and reaction conditions. On this line, catalysts 13 and various combinations of temperatures, reaction times and catalyst/substrate ratios were investigated (Table 1).
For properly surveying and comparing the catalytic performance of our three Pd-bpydc-Ln compounds (13), we used a range of bases and methanol to ensure good solubility of the reaction partners and intermediates (best for the sodium tert-butoxide–methanol pair). Catalytic runs were carried out in the 30–60 °C temperature range, with catalyst/substrate molar ratios corresponding to different catalyst loadings (varying from 0.5 mol % to 0.2 mol %). Notably, low catalyst loadings allowed high yields (93–95%) and conversions (73–99%) to be attained with catalyst 3, at the optimized reaction time and temperature. For almost all sets of reaction parameters, the recorded conversions and yields follow the order: Pd-bpydc-Nd (3) > Pd-bpydc-Ce (2) > Pd-bpydc-La (1). An increase in the product yield with temperature is the general trend for the three catalysts. This result is in accordance with the fact that palladium-catalyzed cross-coupling in organic solvents is conveniently performed at higher temperature [83]. Under all tested conditions, Pd-bpydc-Nd (3) is clearly the best catalyst (Table 1, entries 9–12) affording highest yields (Figure 1) and selectivities. Remarkably, these favorable yields were obtained at 30 °C as well as at 60 °C, which are milder than other currently used temperatures in cross-couplings occurring in common organic solvents.
Noteworthy, we found that the nature of the lanthanide highly matters. As will be seen below, the lanthanides play a significant role in tailoring in several ways the catalytic properties of the palladium complexes. Working either at 30 °C or 60 °C and at low catalyst loadings the activity of Pd-bpydc-Nd (3) (yield: 95%; Table 1, entry 9 and 11) surpasses that of the previously reported Pd(diimine)Cl2 embedded heterobimetallic (Pd-Ln) heterogeneous catalysts [83] which have a related structure but contain another lanthanide (i.e., Ln = Sm). A sharp variation in product yield and catalyst activity with temperature was found for Pd-bpydc-Ce (2); this catalyst proved to be highly active at 60 °C, even for a low catalyst loading (Table 1, entry 7 and 8) but displayed lower yields at 30 °C (Table 1, entry 5 and 6) under the employed reaction conditions (8 h reaction time, 0.2 mol % or 0.5 mol % catalyst loading). Most intriguingly, under best conditions catalyst Pd-bpydc-La (1) provided only moderate yields (max. 55%, Table 1, entry 3), as compared to Nd and Ce. This totally distinct behavior of the La complex is still not well understood. Since compounds 13 having as rare-earth elements La, Ce, Nd are isostructural, we assume that the observed dissimilarities in activity should not stem from the crystal structure but originate from the nature of the lanthanides and their mode of coordination into the metal–organic framework. The varying Lewis acid character and electropositivity of the three rare-earth elements may determine the stability and activity of the coordinated Pd sites with consequences on the reaction outcome. To rationalize this different behavior of the Pd-Ln catalysts, we consider that the lanthanides would build a specific electronic environment at the palladium site through the organic linker, as a function of their electropositive propensity and are, therefore, responsible for the activity and stability of the catalyst. Furthermore, the differences in electropositivity and reducing the power of the three lanthanides, La, Ce, and Nd, may affect otherwise the palladium catalytic cycle and particularly the reductive elimination step by favoring the generation of Pd(0) species able to resume a new catalytic cycle (Scheme 2).
In our case, Nd3+ endowed with a small ionic radius and high availability of the f-electrons (Table 2) seems to have a strong influence on the Pd site through charge transfer, mediated by the flexible organic linker, conducive to an enhancement of the reductive elimination step of the Pd catalytic cycle. This process occurs via synergistic cooperation between the lanthanide and the aromatic linker in transferring charge density to the nitrogen-Pd bond.
The lanthanides, having multiple coordination abilities and an enhanced oxophilic character, enable the formation of a 3D framework anchoring PdCl2 and yielding, thus, a highly active and stable Pd catalyst.
Another significant observation is that low catalyst loadings result in a slight drop in the yield and conversion. Obviously, in our experiments, the yields are consistently higher when the catalyst amount is 0.5 mol %. Moreover, Table 1 convincingly demonstrates that, under the promotion of our dinuclear catalytic systems, the reaction in methanol can lead to advantageous results. The rationale behind the excellent performance in methanol likely lies in the fact that it can dissolve most of the substrate and base ensuring a favorable concentration of these compounds in the reaction medium. Furthermore, methanol is cheap and readily available; however, running catalytic tests in this solvent at more elevated temperatures is hindered by its relatively low boiling point. Gratifyingly, from the above data, Pd-bpydc-Nd catalyst (3) in methanol as the solvent is an excellent choice for efficient application in the Suzuki–Miyaura cross-coupling.
Variation of the base (Table 3) showcased that sodium and potassium hydroxide, as well as sodium tert-butoxide, are the most appropriate bases affording high yields (Table 3, entries 1, 3 and 5), even at lower catalyst loading (0.2 mol %) and reaction temperature (30 °C).
Interestingly, the other tested inorganic bases (e.g., sodium and potassium carbonates) gave rather good yields proving that the employed bases provide a sustainable outcome in the biaryl product (yields of 92–95%), thus, widening the scope of our catalytic process.
As commonly accepted, an important issue in the Suzuki reaction is the nature of the aryl halide, R-X. In this respect, the above-investigated parameters that affect the reaction course have been conducted on 4-bromoanisole as a model substrate. A different R-X substrate, namely 4-bromobenzonitrile, was additionally examined in cross-couplings with phenylboronic acid (Table 4), under the same reaction conditions as given in Table 3, allowing, thus, an accurate comparison of the two aryl substrates.
Studies on the substrate scope unequivocally indicated that, for each base, conversions and yields pertaining to reactions with 4-bromobenzonitrile (Table 4) are superior to those obtained with 4-bromoanisole (Table 3). These results are reasonable taking into account the well-established fact that oxidative addition [Pd(0) to RPd(II)X] is the rate-determining step in the Pd catalytic cycle [5,7,90,93] and, therefore, aryl halides activated by an electron-withdrawing group, as is the case of 4-bromobenzonitrile, are more reactive in the oxidative addition step as compared with substrates bearing electron-donating groups (i.e., 4-bromoanisole). The increase in reactivity of 4-bromobenzonitrile could, thus, explain the obvious difference observed for the carbonates for the two substrates, these weak bases becoming quite effective in C-C couplings with the new aryl bromide (entries 2–4 Table 4 vs. entries 2–4 in Table 3, Figure 2).
Furthermore, the high level of activity maintained by all bases examined by us may be rationalized by considering the same mechanism of action exerted by the base in activating the organoboron compound in the transmetalation step of the Pd catalytic cycle (Scheme 2) [93,94,95]. Inherent differences could be determined by the nature of the counter cations and solubility pattern of the base in the protic medium.
Recyclability of the Pd-bpydc-Nd (3) catalyst has been demonstrated in the reaction of 4-bromoanisole with phenylboronic acid, in consecutive runs, working under our mildest conditions. It was found that the catalyst could be reused three times without significant loss of the catalytic activity. Nonetheless, the structure of the catalyst after three runs could not be determined so far and further investigations are to be conducted in this respect.

4. Conclusions

In summary, the heterodinuclear compounds, Pd-bpydc-La, Pd-bpydc-Ce, and Pd-bpydc-Nd (bpydc = 2,2′-bipyridine-5,5′-dicarboxylate), synthesized in a one-pot procedure, provided advantageous access to biphenyl derivatives by cross-coupling of aryl bromideswithphenylboronic acid. The attractive feature of our approach lies in the association of palladium, the most active and traditionally employed transition metal for the catalyzed Suzuki–Miyaura reaction, with rare earth metals, i.e., La, Nd or Ce, that together with the organic linker build MOFs with good chemical, thermal and mechanical stability. The beneficial cooperativity between the electropositive Ln ions and the electrophilic Pd ions, providing a distinct electronic environment at the Pd active sites, seems to be favored by the bifunctional organic linker, finally leading to a highly active Pd catalyst.
Both high conversions and selectivities have been attained, under mild conditions. It was found that the Pd-bpydc-Nd catalyst performs best. Use of methanol as the solvent and of alkaline hydroxides or carbonates as the base led to optimum outcomes. Results suggest that the rare earth elements, by influencing the activity of palladium, play an important role in attaining the excellent selectivities observed especially for the combination Pd-Nd. Moreover, this catalyst could be recycled in three consecutive runs. Further efforts to extend the application of the here disclosed catalytic systems to other cross-coupling transformations are underway in our laboratory.

Author Contributions

For The authors contributed equally to the work presented in this paper as follows: Y.L, Y.X., Y.D., D.W. and Y.Z. performed the experiments and discussed results; F.D., I.D., V.D. and K.W. designed the experiments, evaluated the results and wrote the manuscript All authors discussed and approved the final version.

Funding

This research was funded by the Natural Science Foundation of China (No.21071100), the LNET (LJQ 2011038) and theDoctor Scientific Startup Foundation ofLiaoning Province (No. 20111046).

Acknowledgments

This work was conducted in the framework of a project sponsored by the Natural Science Foundation of China, the LNET and the Doctor Scientific Startup Foundation of Liaoning Province. I.D. and V.D. acknowledge support from the Romanian Academy, Institute of Organic Chemistry C.D. Nenitescu.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Trusova, M.E.; Rodriguez-Zubiri, M.; Kutonova, K.V.; Jung, N.; Bräse, S.; Felpin, F.-X.; Postnikov, P.S. Ultra-fast Suzuki and Heck reactions for the synthesis of styrenes and stilbenes using arenediazonium salts as super-electrophiles. Org. Chem. Front. 2018, 5, 41–45. [Google Scholar] [CrossRef] [Green Version]
  2. Kostas, I.D. (Ed.) Suzuki–Miyaura Cross-Coupling Reaction and Potential Applications; MDPI: Basel, Switzerland, 2017. [Google Scholar]
  3. Martina, K.; Manzoli, M.; Calcio Gaudino, E.; Cravotto, G. Eco-Friendly Physical Activation Methods for Suzuki–Miyaura Reactions. Catalysts 2017, 7, 98. [Google Scholar] [CrossRef]
  4. Colacot, T. (Ed.) New Trends in Cross-Coupling: Theory and Applications; RSC Publishing: Cambridge, UK, 2015. [Google Scholar]
  5. De Meijere, A.; Bräse, S.; Oestreich, M. (Eds.) Metal-Catalyzed Cross-Coupling Reactions and More; Wiley-VCH Verlag GmbH & Co. KgaA: Weinheim, Germany, 2014; pp. 1–1511. [Google Scholar]
  6. Johansson Seechurn, C.C.C.; Kitching, M.O.; Colacot, T.J.; Snieckus, V. Palladium-catalyzed cross-coupling: A historical contextual perspective to the 2010 Nobel Prize. Angew. Chem. Int. Ed. 2012, 51, 5062–5085. [Google Scholar] [CrossRef] [PubMed]
  7. Suzuki, A. Cross-coupling reactions of organoboranes: An easy way to construct C-C bonds (Nobel Lecture). Angew. Chem. Int. Ed. 2011, 50, 6722–6737. [Google Scholar] [CrossRef] [PubMed]
  8. Molnar, A. Efficient, Selective, and Recyclable Palladium Catalysts in Carbon−Carbon Coupling Reactions. Chem. Rev. 2011, 111, 2251–2320. [Google Scholar] [CrossRef] [PubMed]
  9. Rivera, R.P.; Ehlers, P.; Ohlendorf, L.; Rodriguez, E.T.; Villinger, A.; Langer, P. Chemoselective Synthesis of Arylpyridines through Suzuki–Miyaura Cross-Coupling Reactions. Eur. J. Org. Chem. 2018, 8, 990–1003. [Google Scholar] [CrossRef]
  10. Liu, N.; Liu, C.; Jin, Z. Green synthesis of fluorinated biaryl derivatives via thermoregulated ligand/palladium-catalyzed Suzuki reaction. J. Organomet. Chem. 2011, 696, 2641–2647. [Google Scholar] [CrossRef]
  11. Lopušanskaja, E.; Paju, A.; Järving, I.; Lopp, M. Synthesis of Cyclic 3-Aryl-Substituted 1,2-Dicarbonyl Compounds via Suzuki Cross-Coupling Reactions. Synthesis 2018, 50, 1883–1890. [Google Scholar] [CrossRef]
  12. Dai, J.J.; Liu, J.H.; Luo, D.F. Pd-catalysed decarboxylative Suzuki reactions and orthogonal Cu-based O-arylation of aromatic carboxylic acids. Chem. Commun. 2011, 47, 677–679. [Google Scholar] [CrossRef] [PubMed]
  13. Ramakrishna, V.; Rani, M.J.; Reddy, N.D. A Zwitterionic Palladium(II) Complex as a Precatalyst for Neat Water Mediated Cross-Coupling Reactions of Heteroaryl, Benzyl and Aryl Acid Chlorides with Organoboron Reagents. Eur. J. Org. Chem. 2017, 48, 7238–7255. [Google Scholar] [CrossRef]
  14. Polshettiwar, V.; Decottignies, A.; Len, C.; Fihri, A. Suzuki–Miyaura Cross-Coupling Reactions in Aqueous Media: Green and Sustainable Syntheses of Biaryls. ChemSusChem 2010, 3, 502–522. [Google Scholar] [CrossRef] [PubMed]
  15. Drozdov, F.V.; Cherkaev, G.V.; Muzafarov, A.M. The Suzuki modification of functional polydimethylsiloxanes. Mendeleev. Commun. 2017, 27, 570–571. [Google Scholar] [CrossRef]
  16. Wu, W.B.; Wang, C.; Zhong, C.; Ye, C.; Qiu, G.; Qin, J.; Li, Z. Changing the shape of chromophores from “H-type” to “star-type”: Increasing the macroscopic NLO effects by a large degree. Polym. Chem. 2013, 4, 378–396. [Google Scholar] [CrossRef]
  17. Liu, M.; Su, S.-J.; Jung, M.-C.; Qi, Y.B.; Zhao, W.-M.; Kido, J. Hybrid Heterocycle-Containing Electron-Transport Materials Synthesized by Regioselective Suzuki Cross-Coupling Reactions for Highly Efficient Phosphorescent OLEDs with Unprecedented Low Operating Voltage. Chem. Mater. 2012, 24, 3817–3827. [Google Scholar] [CrossRef]
  18. Soloducho, J.; Cabaj, J.; Olech, K.; Data, P.; Lapkowski, M. Advanced Heterocyclic Branched Semiconducting Units—Highly Efficient Synthesis and Physicochemical Characteristic. Curr. Org. Chem. 2013, 17, 283–295. [Google Scholar] [CrossRef]
  19. Martinez-Amezaga, M.; Delpiccolo, C.M.L.; Mata, E.G. Immobilized boronic acid for Suzuki–Miyaura coupling: Application to the generation of pharmacologically relevant molecules. RSC Adv. 2017, 7, 34994–35003. [Google Scholar] [CrossRef]
  20. Magano, J.; Dunetz, J.R. Large-scale applications of transition metal-catalyzed couplings for the synthesis of pharmaceuticals. Chem. Rev. 2011, 111, 2177–2250. [Google Scholar] [CrossRef] [PubMed]
  21. Torborg, C.; Beller, M. Recent Applications of Palladium-Catalyzed Coupling Reactions in the Pharmaceutical, Agrochemical, and Fine Chemical Industries. Adv. Synth. Catal. 2009, 351, 3027–3043. [Google Scholar] [CrossRef]
  22. De Vries, J.G. Palladium-Catalysed Coupling Reactions. In Organometallics as Catalysts in the Fine Chemical Industry; Beller, M., Blaser, H.-U., Eds.; Springer-Verlag: Berlin/Heidelberg, Germany, 2012; pp. 1–34. [Google Scholar]
  23. Kitney, S.P.; Cheng, F.; Khan, S.; Hope, C.N.; McNab, W.; Kelly, S.M. Synthesis of liquid crystals using Suzuki–Miyaura coupling reactions using mesoporous, ligand-free Pd/Si3N4 catalysts in aqueous media. Liq. Cryst. 2011, 38, 1027–1033. [Google Scholar] [CrossRef]
  24. Kal Koshvandi, A.T.; Heravi, M.M.; Momeni, T. Current Applications of Suzuki–Miyaura Coupling Reaction in the Total Synthesis of Natural Products: An update. Appl. Organomet. Chem. 2018, 32, e4210. [Google Scholar] [CrossRef]
  25. Zheng, S.; Laraia, L.; O′Connor, C.J.; Sorrell, D.; Tan, Y.S.; Xu, Z.; Venkitaraman, A.R.; Wu, W.; Spring, D.R. Synthesis and biological profiling of tellimagrandin I and analogues reveals that the medium ring can significantly modulate biological activity. Org. Biomol. Chem. 2012, 10, 2590–2593. [Google Scholar] [CrossRef] [PubMed]
  26. Lopopolo, G.; Fiorella, F.; de Candia, M.; Nicolotti, O.; Martel, S.; Carrupt, P.A.; Altomare, C. Biarylmethoxy isonipecotanilides as potent and selective inhibitors of blood coagulation factor Xa. Eur. J. Pharm. Sci. 2011, 42, 180–191. [Google Scholar] [CrossRef] [PubMed]
  27. Xie, J.; Okano, A.; Pierce, J.G.; James, R.C.; Stamm, S.; Crane, C.M.; Boger, D.L. Total Synthesis of [Ψ[C(═S)NH]Tpg4]Vancomycin Aglycon, [Ψ[C(═NH)NH]Tpg4]Vancomycin Aglycon, and Related Key Compounds: Reengineering Vancomycin for Dual d-Ala-d-Ala and d-Ala-d-Lac Binding. J. Am. Chem. Soc. 2012, 134, 1284–1297. [Google Scholar] [CrossRef] [PubMed]
  28. Edwankar, C.R.; Edwankar, R.V.; Deschamps, J.R.; Cook, J.M. Nature-inspired stereospecific total synthesis of P-(+)-dispegatrine and four other monomeric sarpagine indole alkaloids. Angew. Chem. Int. Ed. 2012, 51, 11762–11765. [Google Scholar] [CrossRef] [PubMed]
  29. Fujiki, K.; Tanaka, K. Ligand: Self-Assembly of Palladium Nanoparticles and Catalytic Reactivity on C–C. Bond Formation. Synthsis 2017. [Google Scholar] [CrossRef]
  30. Deraedt, C.; Astruc, D. “Homeopathic” palladium nanoparticle catalysis of cross carbon-carbon coupling reactions. Acc. Chem. Res. 2014, 47, 494–503. [Google Scholar] [CrossRef] [PubMed]
  31. Sydnes, M.O. Use of nanoparticles as catalysts in organic synthesis—Cross-coupling reactions. Curr. Org. Chem. 2014, 18, 312–326. [Google Scholar] [CrossRef]
  32. Heinrich, F.; Keßler, M.T.; Dohmen, S.; Singh, M.; Prechtl, M.H.G.; Mathur, S. Molecular Palladium Precursors for Pd Nanoparticle Preparation by Microwave Irradiation: Synthesis, Structural Characterization and Catalytic Activity. Eur. J. Inorg. Chem. 2012, 36, 6027–6033. [Google Scholar] [CrossRef]
  33. Zhi, J.; Song, D.; Li, Z. Palladium nanoparticles in carbon thin film-lined SBA-15 nanoreactors: Efficient heterogeneous catalysts for Suzuki–Miyaura cross coupling reaction in aqueous media. Chem. Commun. 2011, 47, 10707–10709. [Google Scholar] [CrossRef] [PubMed]
  34. Venkatesan, P.; Santhanalakshmi, J. Synthesis, characterization and catalytic activity of trimetallic nanoparticles in the Suzuki C–C coupling reaction. J. Mol. Catal. A-Chem. 2010, 326, 99–106. [Google Scholar] [CrossRef]
  35. Fujiki, K.; Tanaka, K. Bis[N,N′-(2-indanolyl)]-1,5-diazacyclooctane as Unique Metal Ligand: Self-Assembly of Palladium Nanoparticles and Catalytic Reactivity on C–C Bond Formation. Synthesis 2017, 49, 1097–1104. [Google Scholar]
  36. Pourjavadi, A.; Habibi, Z. Palladium nanoparticle-decorated magnetic pomegranate peel-derived porous carbon nanocomposite as an excellent catalyst for Suzuki–Miyaura and Sonogashira cross-coupling reactions. Appl. Organomet. Chem. 2018, e4480. [Google Scholar] [CrossRef]
  37. Baran, T.; Sargin, I.; Kaya, M.; Menteş, A.; Ceter, T. Design and application of sporopollenin microcapsule supported palladium catalyst: Remarkably high turnover frequency and reusability in catalysis of biaryls. J. Colloid Interface Sci. 2017, 486, 194–203. [Google Scholar] [CrossRef] [PubMed]
  38. Liu, Q.-X.; Hu, Z.-L.; Yu, S.-C.; Zhao, Z.-X.; Wei, D.-C.; Li, H.-L. NHC Pd(II) and Ag(I) Complexes: Synthesis, Structure, and Catalytic Activity in Three Types of C−C. Coupling Reactions. ACS Omega 2018, 3, 4035–4047. [Google Scholar] [CrossRef]
  39. Akkoç, M.; İmik, F.; Yaşar, S.; Dorcet, V.; Roisnel, T.; Bruneau, C.; Özdemir, I. An Efficient Protocol for Palladium N-Heterocyclic Carbene-Catalysed Suzuki-Miyaura Reaction at room temperature. Chem. Sel. 2017, 2, 5729–5734. [Google Scholar] [CrossRef]
  40. Chartoire, A.; Nolan, S.P. Chapter. 4: Advances in C-C and C-X Coupling Using Palladium N-Heterocyclic Carbene (Pd-NHC) Complexes. In New Trends in Cross-Coupling: Theory and Applications; Colacot, T., Ed.; RSC Publishing: Cambridge, UK, 2015. [Google Scholar]
  41. Izquierdo, F.; Corpet, M.M.A.E.; Nolan, S.P. The Suzuki–Miyaura Reaction Performed Using a Palladium–N-Heterocyclic Carbene Catalyst and a Weak Inorganic Base. Eur. J. Org. Chem. 2015, 9, 1920–1924. [Google Scholar] [CrossRef]
  42. Guisado-Barrios, G.; Hiller, J.; Peris, E. Pyracene-Linked Bis-Imidazolylidene Complexes of Palladium and Some Catalytic Benefits Produced by Bimetallic Catalysts. Chem.-Eur. J. 2013, 19, 10405–10411. [Google Scholar] [CrossRef] [PubMed]
  43. Gu, P.; Xu, Q.; Shi, M. Synthesis of Novel N-Heterocyclic Carbene-Oxazoline Palladium Complexes and Their Applications in Suzuki–Miyaura Cross-Coupling Reaction. Synlett 2013, 24, 1255–1259. [Google Scholar] [CrossRef]
  44. Szulmanowicz, M.S.; Gniewek, A.; Gil, W.; Trzeciak, A.M. Palladium(II) Complexes with Small N-Heterocyclic Carbene Ligands as Highly Active Catalysts for the Suzuki–Miyaura Cross-Coupling Reaction. ChemCatChem. 2013, 5, 1152–1160. [Google Scholar] [CrossRef]
  45. DeAngelis, A.; Colacot, T.J. Chapter. 2: Prominent Ligand Types in Modern Cross-Coupling Reactions. In New Trends in Cross-Coupling: Theory and Applications; Colacot, T., Ed.; RSC Publishing: Cambridge, UK, 2015. [Google Scholar]
  46. Stradiotto, M. Chapter. 5: Ancillary Ligand Design in the Development of Palladium Catalysts for Challenging Selective Monoarylation Reactions. In New Trends in Cross-Coupling: Theory and Applications; Colacot, T., Ed.; RSC Publishing: Cambridge, UK, 2015. [Google Scholar]
  47. Gogoi, G.; Bora, U.; Borah, G.; Gogoi, P.K. Nanosilica-anchored Pd(II)-Schiff base complex as efficient heterogeneous catalyst for activation of aryl halides in Suzuki–Miyaura cross-coupling reaction in water. Appl. Organomet. Chem. 2017, 31, e3686. [Google Scholar] [CrossRef]
  48. Sobhani, S.; Mesbah Falatooni, Z.M.; Asadi, S.; Honarmand, M. Palladium-Schiff Base Complex Immobilized Covalently on Magnetic Nanoparticles as an Efficient and Recyclable Catalyst for Heck and Suzuki Cross-Coupling Reactions. Catal. Lett. 2016, 146, 255–268. [Google Scholar] [CrossRef]
  49. Liu, N.; Liu, C.; Jin, Z. Poly(ethylene glycol)-functionalized imidazolium salts–palladium-catalyzed Suzuki reaction in water. Green Chem. 2012, 14, 592–597. [Google Scholar] [CrossRef]
  50. Chen, W.; Lu, X.-Y.; Xu, B.-H.; Yu, W.-G.; Zhou, Z.-N.; Hu, Y. The Rational Design and Synthesis of Water-Soluble Thiourea Ligands for Recoverable Pd-Catalyzed Aerobic Aqueous Suzuki–Miyaura Reactions at Room Temperature. Synthesis 2018, 50, 1499–1510. [Google Scholar] [CrossRef]
  51. Johansson Seechurn, C.C.C.; Li, H.; Colacot, T.J. Chapter. 3: Pd-Phosphine Precatalysts for Modern Cross-Coupling Reactions. In New Trends in Cross-Coupling: Theory and Applications; Colacot, T., Ed.; RSC Publishing: Cambridge, UK, 2015. [Google Scholar]
  52. Li, H.B.; Johansson Seechurn, C.C.C.; Colacot, T.J. Development of Preformed Pd Catalysts for Cross-Coupling Reactions, Beyond the 2010 Nobel Prize. ACS Catal. 2012, 2, 1147–1164. [Google Scholar] [CrossRef]
  53. Kumar, S.; Ahmed, N. A facile approach for the synthesis of novel 1-oxa- and 1-aza-flavonyl-4-methyl-1H-benzo[d][1,3]oxazin-2(4H)-ones by microwave enhanced Suzuki–Miyaura coupling using bidentate chromen-4-one-based Pd(II)–diimine complex as catalyst. RSC Adv. 2015, 5, 77075–77087. [Google Scholar] [CrossRef]
  54. Zhou, Z.G.; Shi, J.C.; Hu, Q.S.; Xie, Y.R.; Du, Z.Y.; Zhang, S.Y. Suzuki cross-coupling reactions of aryl chlorides using [Cl2Pd(COD)]/piperazine derivative under microwave conditions. Appl. Organomet. Chem. 2011, 25, 616–619. [Google Scholar] [CrossRef]
  55. Kong, S.; Malik, A.U.; Qian, X.; Shu, M.; Xiao, W. C-C Coupling Reactions in Water Catalyzed by Palladium(II) Supported on Microporous Organic Polymer. Chin. J. Org. Chem. 2018, 38, 432–442. [Google Scholar] [CrossRef]
  56. Fareghi-Alamdari, R.; Golestanzadeh, M.; Bagheri, O. An efficient and recoverable palladium organocatalyst for Suzuki reaction in aqueous media. Appl. Organomet. Chem. 2017, 31, e3698. [Google Scholar] [CrossRef]
  57. Aktaş, A.; Celepci, D.B.; Gök, Y.; Aygün, M. 2-Hydroxyethyl-Substituted Pd-PEPPSI Complexes: Synthesis, Characterization and the Catalytic Activity in the Suzuki-Miyaura Reaction for Aryl Chlorides in Aqueous Media. ChemistrySelect 2018, 3, 9974–9980. [Google Scholar] [CrossRef]
  58. Chatterjee, A.; Ward, T.R. Recent advances in the palladium catalyzed Suzuki–Miyaura cross-coupling reaction in water. Catal. Lett. 2016, 146, 820–840. [Google Scholar] [CrossRef]
  59. Liang, Q.B.; Xing, P.; Huang, Z.G.; Dong, J.J.; Sharpless, K.B.; Li, X.X.; Jiang, B. Palladium-Catalyzed Ligand-Free Suzuki Reaction In Water Using Aryl Fluorosulfates. Org. Lett. 2015, 17, 1942–1945. [Google Scholar] [CrossRef] [PubMed]
  60. Zhong, R.; Pöthig, A.; Feng, Y.; Riener, K.; Herrmann, W.A.; Kühn, F.E. Facile-prepared sulfonated water-soluble PEPPSI-Pd-NHC catalysts for aerobic aqueous Suzuki–Miyaura cross-coupling reactions. Green Chem. 2014, 16, 4955–4962. [Google Scholar] [CrossRef] [Green Version]
  61. Mao, S.L.; Sun, Y.; Yu, G.A.; Zhao, C.; Han, Z.J.; Yuan, J.; Zhu, X.; Yang, Q.; Liu, S.H. A highly active catalytic system for Suzuki–Miyaura cross-coupling reactions of aryl and heteroaryl chlorides in water. Org. Biomol. Chem. 2012, 10, 9410–9417. [Google Scholar] [CrossRef] [PubMed]
  62. Li, L.; Wang, J.; Zhou, C.; Wang, R.; Hong, M. pH-Responsive chelating N-heterocyclic dicarbene palladium(II) complexes: Recoverable precatalysts for Suzuki–Miyaura reaction in pure water. Green Chem. 2011, 13, 2071–2077. [Google Scholar] [CrossRef]
  63. Wu, X.F.; Neumann, H.; Beller, M. Suzuki Coupling of Benzyl Halides with Potassium Aryltrifluoroborates in Aqueous Media. Adv. Synth. Catal. 2011, 353, 788–792. [Google Scholar] [CrossRef]
  64. Crisóstomo-Lucas, C.; Toscano, R.A. Synthesis and characterization of new potentially hydrosoluble pincer ligands and their application in Suzuki–Miyaura cross-coupling reactions in water. Tetrahedron Lett. 2013, 54, 3116–3119. [Google Scholar]
  65. Conelly-Espinosa, P.; Morales-Morales, D. [Pd(HQS)2] (HQS = 8-hydroxyquinoline-5-sulfonic acid) a highly efficient catalyst for Suzuki-Miyaura cross couplings in water. Inorg. Chim. Acta 2010, 363, 1311–1315. [Google Scholar] [CrossRef]
  66. Basauri-Molina, M.; Hernández-Ortega, S.; Morales-Morales, D. Microwave-assisted C-C and C-S couplings catalysed by organometallic Pd-SCS or coordination Ni-SNS pincer complexes. Eur. J. Inorg. Chem. 2014, 2014, 4619–4625. [Google Scholar] [CrossRef]
  67. Arvela, R.K.; Leadbeater, N.E. Suzuki coupling of aryl chlorides with phenylboronic acid in water, using microwave heating with simultaneous cooling. Org. Lett. 2005, 7, 2101–2104. [Google Scholar] [CrossRef] [PubMed]
  68. Lipshutz, B.H.; Taft, B.R.; Abela, A.R.; Ghorai, S.; Krasovskiy, A.; Duplais, C. Catalysis in the Service of Green Chemistry: Nobel Prize-Winning Palladium-Catalysed Cross-Couplings, Run in Water at Room Temperature: Heck, Suzuki–Miyaura and Negishi Reactions Carried Out in the Absence of Organic Solvents, Enabled by Micellar Catalysis. Platin. Met. Rev. 2012, 56, 62–74. [Google Scholar] [CrossRef] [PubMed]
  69. Liu, C.; Ni, Q.; Bao, F.; Qiu, J. A simple and efficient protocol for a palladium-catalyzed ligand-free Suzuki reaction at room temperature in aqueous DMF. Green Chem. 2011, 13, 1260–1266. [Google Scholar] [CrossRef]
  70. Karimi, B.; Zamani, A. SBA-15-functionalized palladium complex partially confined with ionic liquid: An efficient and reusable catalyst system for aqueous-phase Suzuki reaction. Org. Biomol. Chem. 2012, 10, 4531–4536. [Google Scholar] [CrossRef] [PubMed]
  71. Malineni, J.; Jezorek, R.L.; Zhang, N.; Percec, V. NiIICl(1-Naphthyl)(PCy3)2, An Air-Stable σ-NiII Precatalyst for Quantitative Cross-Coupling of Aryl C–O Electrophiles with Aryl Neopentylglycolboronates. Synthesis 2016, 48, 2808–2815. [Google Scholar] [CrossRef]
  72. Malineni, J.; Jezorek, R.L.; Zhang, N.; Percec, V. An Indefinitely Air-Stable σ-NiII Precatalyst for Quantitative Cross-Coupling of Unreactive Aryl Halides and Mesylates with Aryl Neopentylglycolboronates. Synthesis 2016, 48, 2795–2807. [Google Scholar] [CrossRef]
  73. Jezorek, R.L.; Zhang, N.; Leowanawat, P.; Brunner, M.H.; Gutsche, N.; Pesti, A.K.R.; Olsen, J.T.; Percec, V. Air-Stable Nickel Precatalysts for Fast and Quantitative Cross-Coupling of Aryl Sulfamates with Aryl Neopentylglycolboronates at Room Temperature. Org. Lett. 2014, 16, 6326–6329. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, N.; Hoffman, D.J.; Gutsche, N.; Gupta, J.; Percec, V. Comparison of Arylboron-Based Nucleophiles in Ni-Catalyzed Suzuki–Miyaura Cross-Coupling with Aryl Mesylates and Sulfamates. J. Org. Chem. 2012, 77, 5956–5964. [Google Scholar] [CrossRef] [PubMed]
  75. Leowanawat, P.; Zhang, N.; Percec, V. Nickel Catalyzed Cross-Coupling of Aryl C–O Based Electrophiles with Aryl Neopentylglycolboronates. J. Org. Chem. 2012, 77, 1018–1025. [Google Scholar] [CrossRef] [PubMed]
  76. Baghbanzadeh, M.; Pilger, C.; Kappe, C.O. Rapid Nickel-Catalyzed Suzuki−Miyaura Cross-Couplings of Aryl Carbamates and Sulfamates Utilizing Microwave Heating. J. Org. Chem. 2011, 76, 1507–1510. [Google Scholar] [CrossRef] [PubMed]
  77. Hanley, P.S.; Ober, M.S.; Krasovskiy, A.L.; Whiteker, G.T.; Kruper, W.J. Nickel- and Palladium-Catalyzed Coupling of Aryl Fluorosulfonates with Aryl Boronic Acids Enabled by Sulfuryl Fluoride. ACS Catal. 2015, 5, 5041–5046. [Google Scholar] [CrossRef]
  78. You, L.-X.; Liu, H.-J.; Cui, L.-X.; Ding, F.; Xiong, G.; Wang, S.-J.; Ren, B.-Y.; Dragutan, I.; Dragutan, V.; Sun, Y.-G. The synergistic effect of cobalt on a Pd/Co catalyzed Suzuki–Miyaura cross-coupling in water. Dalton Trans. 2016, 45, 18455–18458. [Google Scholar] [CrossRef] [PubMed]
  79. Kamijo, S.; Yamamoto, Y. A New Pd0–CuI Bimetallic Catalyst for the Synthesis of Indoles from Isocyanates and Allyl Carbonates. Angew. Chem. Int. Ed. 2002, 41, 3230–3233. [Google Scholar] [CrossRef]
  80. Kumar, A.; Rao, G.K.; Kumar, S.; Singh, A.K. Organosulphur and related ligands in Suzuki–Miyaura C–C coupling. Dalton Trans. 2013, 42, 5200–5223. [Google Scholar] [CrossRef] [PubMed]
  81. Sharma, K.N.; Joshi, H.; Sharma, A.K.; Prakash, O.; Singh, A.K. Selenium-Containing N-Heterocyclic Carbenes and Their First Palladium(II) Complexes: Synthesis, Structure, and Pendent Alkyl Chain Length Dependent Catalytic Activity for Suzuki–Miyaura Coupling. Organometallics 2013, 32, 2443–2451. [Google Scholar] [CrossRef]
  82. He, K.; Song, F.; Sun, H.; Huang, Y. Transition-Metal-Free Suzuki–Type Cross-Coupling Reaction of Benzyl Halides and Boronic Acids via 1,2-Metalate Shift. J. Am. Chem. Soc. 2018, 140, 2693–2699. [Google Scholar] [CrossRef] [PubMed]
  83. Huang, S.L.; Jia, A.Q.; Jin, G.X. Pd(diimine)Cl2 embedded heterometallic compounds with porous structures as efficient heterogeneous catalysts. Chem. Commun. 2013, 49, 2403–2405. [Google Scholar] [CrossRef] [PubMed]
  84. You, L.; Zong, W.; Xiong, G.; Ding, F.; Wang, S.; Ren, B.; Dragutan, I.; Dragutan, V.; Sun, Y.-G. Cooperative effects of lanthanides when associated with palladium in novel, 3D Pd/Ln coordination polymers. Sustainable applications as water-stable, heterogeneous catalysts in carbon–carbon cross-coupling reactions. Appl. Catal. A Gen 2016, 511, 1–10. [Google Scholar] [CrossRef]
  85. You, L.; Zhu, W.; Wang, S.; Xiong, G.; Ding, F.; Ren, B.; Dragutan, I.; Dragutan, V.; Sun, Y.-G. High Catalytic Activity in Aqueous Heck and Suzuki–Miyaura Reactions Catalyzed by Novel Pd/Ln Coordination Polymers Based on 2,2′-Bipyridine-4,4′-dicarboxylic Acid as a Heteroleptic Ligand. Polyhedron 2016, 115, 47–53. [Google Scholar] [CrossRef]
  86. Wu, X.-R.; Yao, S.-Y.; Wei, L.-Q.; Li, L.-P.; Ye, B.-H. Construction of Heterometallic M2Pd3 Supramolecular Cages via a Metalloligand Strategy as Heterogeneous Catalyst for Suzuki–Miyaura Coupling Reaction. Inorg. Chim. Acta 2018, 482, 605–611. [Google Scholar] [CrossRef]
  87. Xiong, G.; Chen, X.-L.; You, L.-X.; Ren, B.-Y.; Ding, F.; Dragutan, I.; Dragutan, V.; Sun, Y.-G. La-Metal–Organic Framework Incorporating Fe3O4 Nanoparticles, Post-Synthetically Modified with Schiff Base and Pd. A Highly Active, Magnetically Recoverable, Recyclable Catalyst for C–C Cross-Couplings at Low Pd Loadings. J. Catal. 2018, 361, 116–125. [Google Scholar] [CrossRef]
  88. You, L.-X.; Cui, L.-X.; Zhao, B.-B.; Xiong, G.; Ding, F.; Ren, B.-Y.; Shi, Z.-L.; Dragutan, I.; Dragutan, V.; Sun, Y.-G. Tailoring the structure, pH sensitivity and catalytic performance in Suzuki–Miyaura cross-couplings of Ln/Pd MOFs based on the 1,1′-di(p-carboxybenzyl)-2,2′-diimidazole linker. Dalton Trans. 2018, 47, 8755–8763. [Google Scholar] [CrossRef] [PubMed]
  89. Sun, Y.-G.; Li, J.; You, J.; Guo, Y.; Xiong, G.; Ren, B.Y.; You, L.-X.; Ding, F.; Wang, S.-J.; Verpoort, F.; et al. Bis-imidazole coordination polymers controlled by oxalate as an auxiliary ligand. J. Coord. Chem. 2015, 68, 1199–1212. [Google Scholar] [CrossRef]
  90. Sun, Y.-N.; Xiong, G.; Dragutan, V.; Dragutan, I.; Ding, F.; Sun, Y.-G. Novel luminescent heterobimetallic Ln-Cu(I) 3D coordination polymers based on 5-(4-pyridyl) isophthalic acid as heteroleptic ligand. Synthesis and structural characterization. Inorg. Chem. Commun. 2015, 62, 103–106. [Google Scholar] [CrossRef]
  91. Greenwood, N.N.; Earnshaw, A. Chemistry of the Elements, 2nd ed.; Butterworth-Heinemann: Oxford, UK, 1997; pp. 1230–1242. ISBN 0-08-037941-9. [Google Scholar]
  92. Cotton, S. Lanthanide and Actinide Chemistry; John Wiley & Sons Ltd.: Hoboken, NJ, USA, 2006; ISBN 9780470010051. [Google Scholar]
  93. Amatore, C.; LeDuc, G.; Jutand, A. Mechanism of Palladium-Catalyzed Suzuki–Miyaura Reactions: Multiple and Antagonistic Roles of Anionic “Bases” and Their Countercations. Chem. Eur. J. 2013, 19, 10082–10093. [Google Scholar] [CrossRef] [PubMed]
  94. Amatore, C.; Jutand, A.; Le Duc, G. Mechanistic Origin of Antagonist Effects of Usual Anionic Bases (OH, CO32−) as Modulated by their Countercations (Na+, Cs+, K+) in Palladium-Catalyzed Suzuki–Miyaura Reactions. Chem. Eur. J. 2012, 18, 6616–6625. [Google Scholar] [CrossRef] [PubMed]
  95. Lima, C.F.R.A.C.; Rodrigues, A.S.M.C.; Silva, V.L.M.; Silva, A.M.S.; Santos, M.N.B.F. Role of the Base and Control of Selectivity in the Suzuki-Miyaura Cross-Coupling Reaction. ChemCatChem 2014, 6, 1291–1302. [Google Scholar] [CrossRef]
Sample Availability: No samples of the compounds are available from the authors.
Scheme 1. The Suzuki–Miyaura cross-couplings of 4-bromoanisole and 4-bromobenzonitrile with phenylboronic acid promoted by Pd-bpydc-Ln catalysts (13).
Scheme 1. The Suzuki–Miyaura cross-couplings of 4-bromoanisole and 4-bromobenzonitrile with phenylboronic acid promoted by Pd-bpydc-Ln catalysts (13).
Molecules 23 02435 sch001
Figure 1. Product yields in Suzuki–Miyaura reaction of 4-bromoanisole with phenylboronic acid catalyzed by 13.
Figure 1. Product yields in Suzuki–Miyaura reaction of 4-bromoanisole with phenylboronic acid catalyzed by 13.
Molecules 23 02435 g001
Scheme 2. Influence of lanthanides (Ln = La, Ce, Nd), substrate, base and solvent on the Pd catalytic cycle in the Suzuki–Miyaura cross-coupling reaction.
Scheme 2. Influence of lanthanides (Ln = La, Ce, Nd), substrate, base and solvent on the Pd catalytic cycle in the Suzuki–Miyaura cross-coupling reaction.
Molecules 23 02435 sch002
Figure 2. Product yields in Suzuki–Miyaura reaction of 4-bromobenzonitrile with phenylboronic acid catalyzed by 3.
Figure 2. Product yields in Suzuki–Miyaura reaction of 4-bromobenzonitrile with phenylboronic acid catalyzed by 3.
Molecules 23 02435 g002
Table 1. Suzuki–Miyaura cross-coupling of 4-bromoanisole with phenylboronic acid in methanol-induced by Pd-bpydc-Ln catalysts 13.
Table 1. Suzuki–Miyaura cross-coupling of 4-bromoanisole with phenylboronic acid in methanol-induced by Pd-bpydc-Ln catalysts 13.
EntryCatalystT (°C)Time (h)Yield (%) a,b,c
1Pd-bpydc-La (1)30835.0 a
230815.0 b
360455.0 a
460414.9 b
5Pd-bpydc-Ce (2)30825.0 a
630825.9 b
760491.9 a
860488.9 b
9Pd-bpydc-Nd (3)30895.0 a
1030893.4 b
1160495.0 a
1260494.9 b
a Reaction conditions: 4-bromoanisole (0.5 mmol), phenylboronic acid (0.6 mmol), sodium tert-butoxide (0.5 mmol), methanol (1 mL),catalyst (0.5 mol %); b Catalyst (0.2 mol %).; c Reaction yield (%) for total conversion determined from the internal standard yield, at the specified reaction time, based on gas chromatography (GC), using hexadecane as internal standard.
Table 2. Ln3+ radius (pm) and electron configuration of La, Ce, and Nd.
Table 2. Ln3+ radius (pm) and electron configuration of La, Ce, and Nd.
LanthanideLaCeNd
Ln3+ radius (pm) a10310298.3
Ln3+ electron configuration b4f04f14f3
a [91]; b [92].
Table 3. Suzuki–Miyaura cross-coupling of 4-bromoanisole with phenylboronic acid in the presence of Pd-bpydc-Nd (3) when various bases are employed a.
Table 3. Suzuki–Miyaura cross-coupling of 4-bromoanisole with phenylboronic acid in the presence of Pd-bpydc-Nd (3) when various bases are employed a.
EntryCatalystBaseYield (%) b
1Pd-bpydc-Nd
(3)
KOH95.0
2K2CO391.9
3NaOH92.0
4Na2CO392.0
5NaOtBu95.0
a Reaction conditions: 4-bromoanisole (0.5 mmol), phenylboronic acid (0.6 mmol), base (0.5 mmol), methanol (1 mL); reaction time 8 h; temperature: 30 °C, catalyst: 0.2 mol %. b Reaction yield (%) for total conversion determined from the internal standard yield, at the specified reaction time, based on the GC, using hexadecane as internal standard.
Table 4. Suzuki–Miyaura cross-coupling of 4-bromobenzonitrile with phenylboronic acid in the presence of Pd-bpydc-Nd (3) when various bases are used a.
Table 4. Suzuki–Miyaura cross-coupling of 4-bromobenzonitrile with phenylboronic acid in the presence of Pd-bpydc-Nd (3) when various bases are used a.
EntryCatalystBaseYield (%) b
1Pd-bpydc-Nd
(3)
KOH94,2
2K2CO394.5
3NaOH97.6
4Na2CO395.9
5NaOtBu90.8
a Reaction conditions: 4-bromobenzonitrile (0.5 mmol), phenylboronic acid (0.6 mmol), base (0.5 mmol), methanol (1 mL); reaction time 8 h; temperature: 30 °C, catalyst: 0.2 mol %. b Reaction yield (%) for total conversion determined from the internal standard yield, at the specified reaction time, based on the GC, using hexadecane as internal standard.

Share and Cite

MDPI and ACS Style

Ding, F.; Li, Y.; Yan, P.; Deng, Y.; Wang, D.; Zhang, Y.; Dragutan, I.; Dragutan, V.; Wang, K. Efficient Suzuki–Miyaura C-C Cross-Couplings Induced by Novel Heterodinuclear Pd-bpydc-Ln Scaffolds. Molecules 2018, 23, 2435. https://doi.org/10.3390/molecules23102435

AMA Style

Ding F, Li Y, Yan P, Deng Y, Wang D, Zhang Y, Dragutan I, Dragutan V, Wang K. Efficient Suzuki–Miyaura C-C Cross-Couplings Induced by Novel Heterodinuclear Pd-bpydc-Ln Scaffolds. Molecules. 2018; 23(10):2435. https://doi.org/10.3390/molecules23102435

Chicago/Turabian Style

Ding, Fu, Yanli Li, Pingxuan Yan, Yan Deng, Dongping Wang, Yajing Zhang, Ileana Dragutan, Valerian Dragutan, and Kangjun Wang. 2018. "Efficient Suzuki–Miyaura C-C Cross-Couplings Induced by Novel Heterodinuclear Pd-bpydc-Ln Scaffolds" Molecules 23, no. 10: 2435. https://doi.org/10.3390/molecules23102435

Article Metrics

Back to TopTop