首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 78 毫秒
1.
Gold surface-bound hyperbranched polyethyleneimine (PEI) films decorated with palladium nanoparticles have been used as efficient catalysts for a series of Suzuki reactions. This thin film-format demonstrated good catalytic efficiency (TON up to 3.4 × 103) and stability. Incorporation into a quartz crystal microbalance (QCM) instrument illustrated the potential for using this approach in lab-on-a-chip-based synthesis applications.

Gold surface-bound hyperbranched polyethyleneimine (PEI) films decorated with palladium nanoparticles have been used as efficient catalysts for a series of Suzuki reactions in a lab-on-a-chip format.

Metal and metal-oxide nanostructures and nanoparticles are often key features in materials used in separation technology, biomedical devices, and catalysis.1,2 Poor mechanical strength and a tendency to aggregate, thus reducing the surface to volume ratio, both limit their use in real-time applications.3,4 To overcome these shortcomings, nanoparticles are often dispersed in a matrix, such as sol–gel, biopolymers, and carbon-based materials without disturbing their innate properties at nano-scale level.5,6 Functionalized polymers have been explored extensively in this regard due to their robustness, low cost, ease of handling, and their amenability for chemical modification.7–9Recently, we have demonstrated the immobilization of hyperbranched polyethyleneimine (PEI) hydrogels on metallic surfaces for use as antifouling surfaces.10 Owing to their hydrogelation properties, polymers of this type have proven useful in therapeutic applications, for e.g. bio-mineralization, and the formation of nanostructures.11,12 PEI is a hyperbranched polymer containing very high densities of primary, secondary and tertiary amines, in a 1 : 2 : 1 ratio, and are capable of coordination with transition metals to form nanostructures.13,14 Recent reports on amine functionalized polymer-supported heterogeneous catalysts for C–C coupling reactions (Table 1-SI) prompted us to deploy PEI as a matrix for synthesizing surface-bound palladium nanoparticles for use as catalysts of C–C coupling reactions. For catalytic applications, a miniaturized microfluidic setup, such as a lab-on-a-chip device, can significantly reduce the consumption of chemicals, catalysts, process time and be beneficial for on-site analysis.15,16In this study, we demonstrate the use of the PEI surface-supported nanoparticles for as a support for catalysis of the Suzuki reaction the possibility for using of these Pd-nanoparticle immobilized PEI-derivatized gold surfaces for performing Suzuki reactions in a microfluidics device. Catalytic surface fabrication (Scheme 1) was performed using gold sputtered quartz surfaces (Au/quartz) that were functionalized with 11-mercaptoundecanoic acid (MuDA), and then activated and derivatization with PEI. The polymer attachment was carried out at high ionic strength (150 mM NaCl), which has been found to enhance the thickness and growth of PEI brush-like structures.10 Optimization of the Pd nanoparticle synthesis procedure was performed by varying incubation times and Pd(OAc)2 concentrations (Table 1-SI). A quartz crystal microbalance (QCM) was used to monitor the amount of Pd deposited on the PEI coated Au/quartz resonators.Open in a separate windowScheme 1Palladium immobilization on polyethyleneimine coated Au/quartz surfaces.The energy dispersive X-ray (EDX) spectrum confirmed the presence of Pd in the PEI film with a distinct band at 2.8 keV (Fig. 1A). XPS spectra of the Pd-bound PEI surfaces revealed the presence of the anticipated proleptic elements (C1s, N1s, O1s, Pd3d and Au4f) (Fig. 1A-SI). Deconvoluted peaks differentiated the amine N of PEI (399.8 eV) from that of the amide N (–*N–C Created by potrace 1.16, written by Peter Selinger 2001-2019 O–, 400.8 eV) (Fig. 1B-SI).17,18 Bands around 335 and 340 eV in the survey spectra correspond to the 3d5/2 and 3d3/2 states of the surface bound Pd.19 The deconvoluted bands at 335.3 and 338.2 demonstrated the presence of Pd(0) and Pd(ii), respectively (Fig. 1B). Importantly, peak integration showed the immobilized Pd to be predominantly in the Pd(0) state, with less than 5% present as Pd(ii).19Open in a separate windowFig. 1(A) Energy dispersive X-ray (EDX) analysis and (B) X-ray photoelectron spectra (XPS) of the palladium nanoparticle immobilized PEI coated Au/quartz surface.RAIR spectra confirmed the presence of the PEI on the gold surface based on the discernible vibrational bands of the –N–H–, –CH– and –CN– bending modes of the adsorbed PEI film (Fig. 2). Subtle differences can be observed in the RAIR spectra of the PEI before and after Pd immobilization. The band corresponding to –N–H– bending mode has been significantly red shifted emphasizing the interaction of the Pd particles with the amine moieties of the PEI film (Fig. 2, inset). This, together with the XPS data, provides evidence for the reduction of Pd(ii) to Pd(0) and its incorporation as nanoparticles into the PEI brush layer.Open in a separate windowFig. 2RAIR spectra of PEI coated Au/quartz surface before and after Pd immobilization.SEM images (Fig. 3) showed uniform long-range coating of the palladium nanoparticles on the PEI immobilized surface (PEI/Pd). The crystallinity of the Pd coated PEI surface was evaluated with powder diffraction measurements which showed characteristic peaks for Pd(111), Pd(200) and Pd(220) with Miller indices of 40, 47.2 and 68.3° (JCPDS No. 98-004-7386), respectively, again confirming the presence of Pd nanoparticles in the (0) oxidation state (curve a, Fig. 2-SI).Open in a separate windowFig. 3Scanning electron micrographs of (A) gold surface, (B) polyethyleneimine (PEI) film and (C) PEI-supported palladium particles fabricated on Au/quartz surfaces.The possibility of using these surfaces for catalysis of the Suzuki reaction was explored using a series of phenylboronic acids and substituted aryl halides (). The amount of catalytic nanopalladium loaded was determined by QCM. Au/quartz surfaces coated with PEI/Pd were immersed in the reaction mixtures at 55 °C for 12 h (Section 1.7-SI). Reactions of aryl halides with a series of arylboronic acids offered the corresponding products in good to excellent yield (). This negligible effect on Pd catalyst poisoning confirms the nature of the catalytic Pd exists predominantly in the (0) oxidation state.20 No biphenyl product has been observed when the Suzuki reaction was performed in presence of unmodified Au/quartz surface (without PEI and Pd).Suzuki cross-coupling reactions of aryl halides with arylboronic acids using PEI/Pd as catalystsa
EntryR1XR2Isolated yieldTON × 103
1HIH93%3.1 × 103
2HBrH95%3.4 × 103
3HI2-CH382%2.2 × 103
4HI3-OCH357%1.5 × 103
5HI4-OCH384%2.4 × 103
6HI2-CN15%0.4 × 103
7HI4-CN95%2.8 × 103
84-CH3BrH88%1.5 × 103
94-OCH3BrH95%1.2 × 103
10HI3-NH2n.r.
11HClH94%10.0 × 103
124-OCH3ClH80%5.3 × 103
134-COCH3ClHn.r.
Open in a separate windowaGeneral procedure: 1.0 mmol of aryl halide, 1.2 mmol of arylboronic acid, 2.0 mmol of K2CO3, in H2O/EtOH. Turnover number TON = mol product/mol Pd. n.r. = no reaction.The Suzuki reaction of phenyl iodide and phenyl boronic acid was also performed using shorter reaction times (6 h and 2 h), with the shorter time providing the product in an acceptable 84% and 78% yield respectively (Tables 2 and 3-SI).To assess the stability of the surfaces and the potential for their reuse, the surfaces were removed from the reaction mixtures and the amount of residual Pd was determined by QCM. XPS measurements revealed no significant change on the nature of the immobilized Pd (335.26 and 340.58 eV) (Fig. 3-SI), and as reflected in the ratio of the Pd3d and Au4f bands measured before (0.45) and after (0.44) the reaction () with peaks for Pd(111), Pd(200) and Pd(220) comparable to those before reaction. The SEM and EDX measurements revealed that the nanopalladium had remained immobilized on the PEI matrix (Fig. 4-SI). The amount of residual Pd was again determined by QCM after use of the surfaces in a series of reactions (
EntryRunConc. of Pd, μgIsolated yieldTON
11st±2.393%4.2 × 103
22nd±2.2289%4.1 × 103
33rd±2.2285%4.0 × 103
44th±2.1580%3.8 × 103
Open in a separate windowaGeneral procedure: 1.0 mmol of aryl halide, 1.2 mmol of arylboronic acid, 2.0 mmol of K2CO3, in H2O/EtOH. TON = mol product/mol.The long-term stability of the catalyst surfaces was studied by storing freshly prepared surfaces in water (Milli-Q grade water, 18.2 MΩ, UHP grade N2) for three months after which catalytic activity was determined (93%).We then explored the use of these surfaces in a lab-on-a-chip format by using them in a QCM instrument fitted with a microfluidics liquid delivery system, where the Suzuki reaction of phenyl iodide and phenylboronic acid was examined under flow injection analysis conditions. The reactants were introduced into the 2 μL volume microreactor using the instrument''s peristaltic pump, and the temperature of the reaction was maintained at 40 °C. Formation of the biphenyl product was confirmed by HPLC analysis of the effluent (Fig. 5-SI).In comparison with other reported heterogeneous catalysts for the Suzuki reaction (Table 4-SI), the Pd/PEI surfaces presented here have significantly lower Pd loadings (>0.001 mol%) though comparable performance, highlighting the role of the PEI matrix for facilitating mass transfer of substrates and products to and from the catalytic centers, and for avoiding aggregation of the Pd.  相似文献   

2.
Palladium catalyst immobilized on functionalized microporous organic polymers for C–C coupling reactions     
Wei Xu  Cijie Liu  Dexuan Xiang  Qionglin Luo  You Shu  Hongwei Lin  Yangjian Hu  Zaixing Zhang  Yuejun Ouyang 《RSC advances》2019,9(59):34595
Two microporous organic polymer immobilized palladium (MOP-Pd) catalysts were prepared from benzene and 1,10-phenanthroline by Scholl coupling reaction and Friedel–Crafts reaction, respectively. The structure and composition of the catalyst were characterized by FT-IR, TGA, N2 sorption, SEM, TEM, ICP-AES and XPS. MOP-Pd catalysts were found to possess high specific surface areas, large pore volume and low skeletal bone density. Moreover, the immobilized catalyst also had advantages, such as readily available raw materials, chemical and thermal stability, and low synthetic cost. The Pd catalyst is an effective heterogeneous catalyst for carbon–carbon (C–C) coupling reactions, such as the Heck reaction and Suzuki–Miyaura reaction, affording good to high yields. In these reactions, the catalyst was easily recovered and reused five times without significant activity loss.

Two microporous organic polymers were prepared from 1,10-phenanthroline by Scholl coupling reaction and Friedel–Crafts reaction, and applied to Heck reaction and Suzuki–Miyaura reaction as heterogeneous catalysts.

Carbon–carbon (C–C) coupling reactions have become one of the most versatile and utilized reactions for the selective construction of C–C bonds for the formation of functionalised aromatics,1 natural products,2 pharmaceuticals,3 polymers4 and advanced materials.5 Many transition metals have been used as catalysts in these reactions, aided by a great variety of ligands ranging from simple, commercial phosphines to complex custom-made molecules.6 Among these transition metals, palladium plays a significant role in various cross-coupling reactions, such as Suzuki,7 Heck,8 Sonogashira,9 Stille,10 and Ullmann coupling reactions,11 due to their strong electrical and chemical properties.12 Over the past decades, various homogeneous catalytic systems have been developed for organic transformations,13 which often progress smoothly under the inert atmosphere in organic solvents, for example, toluene or tetrahydrofuran in the presence of soluble palladium complexes as catalysts. However, most homogeneous palladium catalysts suffer from drawbacks such as high-cost of phosphine ligands, use of various additives, difficult separation, metal leaching, recovery, recyclability, and the toxicity of phosphine ligands.Heterogeneous catalysis have attracted increasing attention as they have been proven to be useful for different organic reactions owning to their unique properties, such as high reactivity, stability, easy separation, purification and recyclability.14 Many active heterogeneous palladium catalysts have been developed and widely applied in the C–C coupling reactions.15 Palladium has been immobilized on various solid supporting materials, such as zeolite,16 silica,17 metal organic frameworks,18 and functionalized graphene oxide.19 However, a substantial decrease in activity and selectivity of the heterogeneous palladium catalysts is frequently observed because of their long diffusion pathway to catalytic sites and the difference of electron density on active sites. To address these problems, materials with larger interface and more active site are employed to support palladium as heterogeneous catalysts, such as palladium immobilized on hyper-crosslinked polymers were high activity in Suzuki–Miyaura coupling reaction.20Microporous organic polymers (MOPs) consists of purely organic elements have recently emerged as versatile platforms for heterogeneous catalysts thanks to their unique properties, including superior chemical, thermal and hydrothermal stability, synthetic diversity, low skeletal density and high surface area.20,21 More importantly, the bottom–up approach of MOPs provides an opportunity for the design of polymer frameworks with a range of functionalities into the porous structure to use as catalysts or ligands.22 Recently, Kaskel reported the incorporation of a thermally fragile imidazolium moiety into MOPs resulted in a heterogeneous organocatalyst active in carbene-catalyzed Umpolung reaction.23 Wang designed photocatalysts with microporous via the copolymerization from pyrene and dibenzothiophene-S,S-dioxide building blocks and tested the effect of the photocatalytic hydrogen evolution.24 Xu described the synthesis of microporous with N-heterocyclic carbenes by an external cross-linking reaction and applied it in Suzuki reaction.25 Zhou demonstrated for the first time that the microporous structure has a positive effect on controlling selectivities in the hydrosilylation of alkynes.26 Recently, we also reported three pyridine-functionalized N-heterocyclic carbene–palladium complexes and its application in Suzuki–Miyaura coupling reactions.271,10-phenanthroline is an ideal candidate of ligands due to its structural features such as two N-atom placed in juxta position to provide binding sites for metal cations.28 To utilize the unique structure feature, we employed it in the construction of MOPs via Scholl and Friedel–Crafts reaction, respectively. Therefore, this paper presents our recent studies on the synthesis of two heterogeneous palladium catalysts supported on MOPs through a simple and low-cost procedure. These catalysts displayed remarkable catalytic activity in C–C coupling reactions, including Suzuki–Miyaura reaction and Heck coupling reaction. The properties of simple preparation, wide application of this catalyst and good performance in C–C coupling reactions and adaptability with various substrates make it perfect catalytic option for C–C coupling reactions.The microporous network with 1,10-phenanthroline functional groups and incorporation of Pd metal were confirmed by Fourier transform infrared (FT-IR) spectroscopy. The FT-IR spectra of MOPs and MOPs-Pd (Fig. 1) displayed a series of bands around 2800–3100 cm−1, which were assigned to the C–H stretching band and in-of-plane bending vibrations of the aryl rings. The bands around 1550–1750 cm−1 were attributed to the –C Created by potrace 1.16, written by Peter Selinger 2001-2019 N- stretching band. The bands around 1400–1450 and 850–700 cm−1 were corresponded to the benzene and 1,10-phenanthroline skeletal stretching and the C–H out-of-plane bending vibrations of the aryl rings, respectively. The bond around 1495 cm−1 in MOPs-I and MOPs-Pd-I is assigned to in-of-plane bending vibrations of CH2, which indicated that 1,10-phenanthroline and benzene were linked by CH2.Open in a separate windowFig. 1FT-IR spectra of MOPs and MOPs-Pd.The X-ray photoelectron spectroscopy (XPS) analysis of the MOPs-Pd is performed to investigate the coordination states of palladium species (Fig. 2). In Fig. 2, the Pd 3d XPS spectra of the MOPs-Pd-I catalysts reveal that Pd is present in the +2 oxidation state rather than in the metallic state. This is corresponding to the binding energy (B.E.) of 337.4 eV and 342.4 eV, which are assigned to be Pd 3d5/2 and 3d3/2 of Pd (+2), respectively. Compared with the PdCl2 (337.9 eV and 343.1 eV), the Pd2+ binding energy in the MOPs-Pd-I catalyst shifts negatively by 0.5 eV and 0.7 eV. This can be attributed to the effect of the coordination with 1,10-phenanthroline in microporous networks. The results show that Pd2+ can be immobilized successfully on the MOPs by coordinating to 1,10-phenanthroline rather than by physical adsorption of Pd2+ on the surface. XPS graphs of MOPs-Pd-II also reveal that Pd2+ is immobilized on MOPs materials.Open in a separate windowFig. 2XPS spectra of the MOPs-Pd.The surface area and pore structure of the MOPs and MOPs-Pd were investigated by nitrogen adsorption analyses at 77.3 K. In Fig. 3, the MOPs-Pd exhibits type I adsorption–desorption isotherms, which is similar to the isotherms exhibited by the parent MOPs polymers. The result implies that these microporous organic polymers and metalized polymers consist of both micropores and mesopores. The apparent Brunauer–Emmett–Teller surface areas (SBET) of MOPs-Pd are smaller than those of the non-metallized parent networks (Open in a separate windowFig. 3N2 adsorption–desorption isotherms and corresponding pore size distributions of MOPs and MOPs-Pd.Physical properties of MOPs and MOPs-Pd
Sample S BET a [m2 g−1] S Micro b [m2 g−1]VMicroc [m3 g−1][Pd]d [wt%]
MOPs-I7614470.211
MOPs-Pd-I7444220.1992.5
MOPs-II6645060.225
MOPs-Pd-II6235020.2252.4
Open in a separate windowaSurface area calculated from the nitrogen adsorption isotherm using the BET method.bThe micropore volume derived using a t-plot method based on the Halsey thickness equation.cTotal pore volume at P/P0 = 0.99.dData were obtained by inductively coupled plasma mass spectrometry (ICP-AES).The thermal stability of the MOPs and MOPs-Pd was also assessed by TGA. The TGA traces obtained from MOPs and MOPs-Pd are shown in Fig. 4. The data analysis has been performed, and results are also shown in Fig. 4. These results show that MOPs and MOPs-Pd exhibit good thermal stability in nitrogen. It is obvious to see that the T5% and T10% of the MOPs-II and MOPs-Pd-II are lower compared with MOPs-I and MOPs-Pd-I. This is because of the large amount CH2 in MOPs-I and MOPs-Pd-I.Open in a separate windowFig. 4TGA curves of MOPs and MOPs-Pd.MOPs and MOPs-Pd were subjected to SEM and TEM analyses, and the results are shown in Fig. 5 and and6.6. We can see a large number of pores in MOPs and MOPs-Pd from the SEM imagines, and uniformly distributed Pd nanoparticles in MOPs-Pd from the TEM images. No remarkable change in terms of the morphology of the materials occurs after loading the palladium species. Then, scanning electron microscopy elemental mapping was employed to investigate the composition of MOPs-Pd. The results are shown in ESI (Section IV). Obviously, the metal Pd in MOPs-Pd-I and II are distributed in the support with a high degree of dispersion. Meanwhile, C, N, Pd and Cl are observed from these images, implying those are the major elements to construct the MOPs-Pd catalyst.Open in a separate windowFig. 5SEM image of MOPs and MOPs-Pd.Open in a separate windowFig. 6TEM image of MOPs and MOPs-Pd.Then, we investigated the activities of the MOPs-Pd catalysts to determine the potential relationships between the structure and catalyst activity. To check the catalytic activity of the MOPs-Pd in the Heck coupling reaction, iodobenzene 1a and ethyl acrylate 2a were taken as the model substrate in presence of MOPs-Pd catalyst for optimization of the reaction condition. First, the reaction of 1a with 2a was carried out in the present of Et3N with MOPs-Pd-I as catalysis in EtOH under reflex to afford 3a in 75% yield. Then, a series of experiments was carried out to screen the reaction conditions, including catalysis, base, solvent, and reaction temperature. The optimal results were obtained when the reaction of 1a with 2a was carried out in the present of Et3N with MOPs-Pd-I as catalysis in DMF at 120 °C for 1.5 h to afford 3a in 96% yield. Under the optimal conditions, we carried out a series of reactions of 1 with 2 aiming to determine its scope. As shown in EntryR1R2R33Yieldb (%)1HHEt3a9624-MeHEt3b9834-MeOHEt3c9744-ClHEt3d9554-NO2HEt3e9364-CNHEt3f9373-MeHEt3g9483,5-(Me)2HEt3h979HHMe3i98103-MeHMe3j9511HHBu3k97123-MeHBu3l9413HHH3m94144-MeHH3n95154-MeOMeMe3o90Open in a separate windowaReaction conditions: 1a (2.5 mmol), 2a (3.7 mmol), Et3N (3.7 mmol), MOPs-Pd-I (50 mg, 0.28 mol%), DMF (10 mL), 120 °C, 1.5 h.bIsolated yields.To determine the active catalyst, we did two experiments in the same condition as 29 Thus, we investigated the recycling performance in Heck reaction, and the results are shown in Fig. 7. The reaction was conducted in the present of Et3N with MOPs-Pd as catalysis in DMF at 120 °C for 1.5 h. Then, the catalyst was recovered by filtering, washing with water and ethyl acetate. Finally, the recovered catalyst was dried in an oven for 2.0 h. After six runs, the reused MOPs-Pd-I and II are still capable of catalyzing the reaction in 93% and 91% yield, respectively. This clearly reveals a slight decrease in catalytic activity and product yield. In addition, the morphology of recovered catalyst was analyzed by the SEM (see ESI, Section IV), and the results show that there is no remarkable change in terms of the morphology of the materials. The contents of Pd in recovered MOPs-Pd-I and II were 2.3% and 2.1% by ICP-AES, implying a slight leaching of palladium species.Open in a separate windowFig. 7Recycle test of MOPs-Pd in Heck reaction.To extend the utility of MOPs-Pd in the carbon–carbon coupling reactions, we examined other organic reactions. Suzuki–Miyaura reaction is an important palladium-catalyzed cross coupling in organic synthesis. Therefore, we examined the MOPs-Pd catalysts in Suzuki–Miyaura reaction. First, the reaction of 1a with 4a was put together in the present of K3PO4 and MOPs-Pd-I in MeOH under reflex. As monitored by TLC, the reaction proceeded smoothly and the yield of 5a reached 91%. Then we investigated the optimization of the reaction conditions, including catalysis, base, solvent, and reaction temperature. A series of experiments revealed that EtOH/H2O (VEtOH/VH2O = 2 : 1) was effective for the synthesis of 5a. The yield of 5a reached 97% when the reaction of 1a with 4a was performed in the present of K3PO4 with MOPs-Pd-I as a catalyst at 80 °C for 1.0 hour. In this reaction, the MOPs-Pd-I catalyst also can be reused for 5 times with no significant decrease in activity and the Pd content of the recovered catalyst is 2.36% by ICP-AES.Under the optimal conditions, we carried out a series of reactions of 1 with 4 aiming to determine its scope. In EntryR1XR45Yieldb (%)1HIH5a972HI2-Me5b963HI3-Me5c994HI4-Me5d985HI2 F5e966HI3 F5f957HI4 F5g978HI4-CN5h9594-MeIH5d98104-OMeIH5i99114-CNIH5h9412cHBrH5a92Open in a separate windowaReaction conditions: 1a (2.5 mmol), 4a (3.0 mmol), K3PO4 (5.0 mmol), MOPs-Pd-I (50 mg, 0.28 mol%), EtOH/H2O (10 mL), 80 °C, 1.0 h.bIsolated yields.cThe reaction time was 3.0 h.In summary, a simple and low-cost method for synthesis of palladium complexes supported on microporous organic polymers was described. The MOPs-Pd catalysts were constructed based on highly stable microporous material, and characterized by FT-IR, TGA, SEM, TEM, N2 sorption, XPS and ICP. These heterogeneous catalysts displayed outstanding catalytic activities in Heck reaction and Suzuki coupling reaction. In these reactions, the MOPs-Pd catalyst was easily recovered and reused without loss of catalytic activity. The potential utilization and application of these heterogeneous catalysts are currently under investigation in our laboratory.  相似文献   

3.
Enantioselective bromination of axially chiral cyanoarenes in the presence of bifunctional organocatalysts     
Yuuki Wada  Akira Matsumoto  Keisuke Asano  Seijiro Matsubara 《RSC advances》2019,9(54):31654
Enantioselective bromination of axially chiral cyanoarenes bearing high intrinsic rotational barriers via dynamic kinetic resolution using bifunctional organocatalysts is reported. Sequential addition of a brominating reagent in several portions at an optimized temperature was effective in accomplishing high enantioselectivities.

Enantioselective bromination of axially chiral cyanoarenes bearing high intrinsic rotational barriers via dynamic kinetic resolution using bifunctional organocatalysts is reported.

Axially chiral biaryls are privileged structures in pharmaceuticals,1 asymmetric catalysts,2 functional materials,3etc. Thus, the development of efficient methods for their synthesis is desirable to advance research in these scientific fields. Among recent accomplishments on catalytic atroposelective transformations4 toward the synthesis of densely substituted axially chiral biaryls, powerful strategies include organocatalytic dynamic kinetic resolution involving ortho-functionalization of existing biaryls via the introduction of additional rotational barriers.5 In this method, substrates should in principle have rotational barriers low enough to enable fast rotation about the biaryl axis, leading to their rapid racemization. Therefore, it is difficult to employ biaryl substrates bearing intrinsic rotational barriers, which impede their racemization. In 2015, the Miller group reported an elegant example of this type of dynamic kinetic resolution via bromination of 3-arylquinazolin-4(3H)-ones, the rotational barrier of which is ∼19 kcal mol−1, by slow addition of a brominating agent.5e Here, we present enantioselective bromination of axially chiral cyanoarenes bearing intrinsic rotational barriers exceeding 18 kcal mol−1 (Scheme 1). To the best of our knowledge, there has been no report of a catalytic asymmetric reaction affording axially chiral cyanoarenes,6 despite their prevalence in bioactive agents7 and the rich chemistry of cyano compounds as synthetic intermediates.8Open in a separate windowScheme 1Enantioselective bromination of axially chiral cyanoarenes using bifunctional organocatalysts. 9 Other catalysts 3d and 3e, bearing a cyclohexanediamine framework, and 3f, bearing a binaphthyl framework, resulted in poor enantioselectivities (10 Using 3a and 3c, lower reaction temperatures were investigated (Fig. 1) were also investigated; NBA (4a) still afforded the best enantioselectivities ( EntryCatalystBrominating reagentSolventTemp. (°C)Yieldb (%)ee (%)13aNBA (4a)CH2Cl225812623bNBA (4a)CH2Cl22587633cNBA (4a)CH2Cl225891843dNBA (4a)CH2Cl22583653eNBA (4a)CH2Cl22585363fNBA (4a)CH2Cl22579373aNBA (4a)CH2Cl2−40822083aNBA (4a)CH2Cl2−60<1—93cNBA (4a)CH2Cl2−408341103cNBA (4a)CH2Cl2−60<1—113cNBA (4a)CHCl3−402154123cNBA (4a)Toluene−4018−10133cNBA (4a)THF−4013−2143cNBA (4a)Et2O−4046−20153cNBA (4a)EtOAc−4068−4163cNBA (4a)EtOH−40<5—173cDBH (4b)CH2Cl2−40791183cNBS (4c)CH2Cl2−408220193cNBP (4d)CH2Cl2−4079−520c,d3cNBA (4a)CH2Cl2−40854921c,e3cNBA (4a)CH2Cl2−40515922c,f3cNBA (4a)CH2Cl2−40336623c,g3cNBA (4a)CH2Cl2−401771 Open in a separate windowaReactions were run using 1a (0.10 mmol), the brominating reagent (0.30 mmol), and the catalyst (0.010 mmol) in the solvent (10 mL).bIsolated yields.cReactions were run using 3c (0.0050 mmol).dReaction was run for 48 h.eReaction was run for 24 h.fReaction was run for 12 h.gReaction was run for 6 h.Open in a separate windowFig. 1Brominating reagents.Subsequently, to improve the efficiency of dynamic kinetic resolution by retarding the enantiodetermining bromination,5e4a was added sequentially in five portions (Fig. 2).11 Although the procedure hardly affected the results at −40 °C, the enantioselectivity was greatly improved for reactions carried out at −20 °C and −30 °C. Such effects were smaller at temperatures above −10 °C.Open in a separate windowFig. 2Investigations of temperatures and procedures. Blue bar: reactions were run with 4a added in 1 portion. Red bar: reactions were run with 4a added in 5 portions. Green values represent yields of 2a isolated after silica gel column chromatography. At 0, −10, −20, and −30 °C, reactions were run for 24 h; at −40 °C, reactions were run for 72 h.Next, at −30 °C and −40 °C, respectively, the relationships between enantioselectivity and yield were investigated (Fig. 3). Reactions were carried out using various amounts of 4a. At both temperatures, the enantioselectivity decreased as the yield increased; however, the quantitative reactions also exhibited some enantioselectivity (−30 °C: 99% yield, 50% ee; −40 °C: 99% yield, 47% ee), implying the presence of the characteristics of dynamic kinetic resolution. In addition, although the enantioselectivity was better at −40 °C than −30 °C when the yield was low, the relationship became reversed as the yield increased; hence, the efficiency of dynamic kinetic resolution was revealed to be better at −30 °C than at −40 °C. Furthermore, when 1.5 equiv. of 4a were used at −30 °C affording 2a in 22% yield with 65% ee, the ortho-monobrominated product 1a-Br (Fig. 4) was also obtained with 75% ee (see Scheme S1 in the ESI for details). It shows that the bromination at one of the ortho-positions introduces a rotational barrier high enough to set the chiral axis, which is consistent with the rotational barriers calculated at the M06-2X/6-311++G(2d,3p)//B3LYP/6-31+G(d,p) level of theory (Fig. 4).Open in a separate windowFig. 3Relationships between ee and yield. Red line: reactions were run at −30 °C for 24 h with 4a added in 5 portions. Blue line: reactions were run at −40 °C for 72 h with 4a added in 5 portions. Red and blue values represent amounts of 4a used for each reaction.Open in a separate windowFig. 4Rotational barriers of substrate, intermediate, and product calculated at the M06-2X/6-311++G(2d,3p)//B3LYP/6-31+G(d,p) level of theory.Under the conditions of using 3c as the catalyst at −30 °C with 3 equiv. of 4a added sequentially in five portions, other substrates bearing substituted phenols were also investigated (Scheme 2).12 First, substrates 1b–1e bearing a substituent at the meta-position were investigated. While the electron-deficient substrate 1b resulted in poor enantioselectivity, 1c and 1d bearing aliphatic substituents gave improved enantioselectivities; however, substrate 1e with a methoxy group resulted in low enantioselectivity. In addition, substrates 1f–1i bearing substituents at the para-positions of the biaryl axis were then examined; phenol 2h bearing a methyl group resulted in higher enantioselectivities than phenols 2f and 2g bearing electron-withdrawing groups and 2i bearing a methoxy group. These results suggest that aliphatic substituents might efficiently facilitate the racemization of 1 during bromination, leading to dynamic kinetic resolution with greater enantioselectivity. Utilizing this methodology with the characteristics of dynamic kinetic resolution, the reactions of 1c and 1g were also carried out using a sub-stoichiometric amount of 4a (Scheme 3); higher enantioselectivities were accomplished albeit with lower yields.13 The absolute configuration of 2c was determined by X-ray crystallography (see the ESI for details), and the configurations of all other products were assigned analogously.Open in a separate windowScheme 2Reactions of substrates with substituted phenols. aReaction was run for 72 h.Open in a separate windowScheme 3Reactions with a substoichiometric amount of 4a.In summary, we present enantioselective bromination of axially chiral cyanoarenes bearing high intrinsic rotational barriers via dynamic kinetic resolution using bifunctional organocatalysts. The sequential addition of 4a in several portions at the optimized temperature was effective in improving the enantioselectivity. Although the enantioselectivities are still moderate using the current catalytic system, the guidelines for designing catalytic asymmetric syntheses of axially chiral cyanoarenes were established. Further studies on the additional optimization and application of this methodology to the construction of densely substituted axially chiral biaryls are currently underway.  相似文献   

4.
Enantioselective conjugate hydrosilylation of α,β-unsaturated ketones     
Huan Yang  Guanglin Weng  Dongmei Fang  Changjiang Peng  Yuanyuan Zhang  Xiaomei Zhang  Zhouyu Wang 《RSC advances》2019,9(21):11627
Enantioselective conjugate hydrosilylation of β,β-disubstituted α,β-unsaturated ketones was realized. In the presence of a chiral picolinamide–sulfonate Lewis base catalyst, the reactions provided various chiral ketones bearing a chiral center at the β-position in up to quantitative yields with moderate enantioselectivities.

Enantioselective conjugate hydrosilylation of β,β-disubstituted α,β-unsaturated ketones was realized.

Chiral ketones are important intermediates for the synthesis of natural products or chiral drugs, and some themselves are useful chiral drugs.1 Asymmetric conjugate reduction of α,β-unsaturated carbonyl compounds is an attractive and challenging transformation for the construction of chiral ketones. During the past few decades, many groups have devoted considerable efforts to this area and made great improvement, for instance, transition metal catalyzed asymmetric hydrogenation of α,β-unsaturated carbonyl compounds, including palladium,2 iridium,3 copper,4 ruthenium,5 rhodium6 and cobalt.7 Besides, some organocatalyzed asymmetric conjugate transfer hydrogenations using Hantzsch ester8 or pinacolborane9 as the hydride source have also been reported. However, very few examples about chiral Lewis base catalyzed conjugate hydrosilylation of α,β-unsaturated carbonyl compounds have been reported and only chiral phosphine oxide Lewis base catalysts were used in these reactions (Scheme 1).10 Therefore, exploration of the application of other kinds of chiral Lewis base catalysts in this reaction is still highly desirable.Open in a separate windowScheme 1Enantioselective conjugate hydrosilylation of α,β-unsaturated ketones by chiral Lewis base catalysts.Recently, we developed a kind of easily accessible chiral picolinamide–sulfonate Lewis base catalysts and used them in asymmetric hydrosilylation of α-acyloxy-β-enamino esters,11 one of them exhibiting excellent reactivity, diastereoselectivity and enantioselectivity. Herein we present the enantioselective conjugate hydrosilylation of β,β-disubstituted α,β-unsaturated ketones using this kind of Lewis base as catalysts, leading to various chiral ketones bearing a chiral center at β-position (Fig. 1).Open in a separate windowFig. 1Three typical chiral drugs containing chiral ketone moiety.First, various chiral Lewis base catalysts 2 were screened in the enantioselective hydrosilylation of (E)-1,3-diphenylbut-2-en-1-one 1a in acetonitrile at 0 °C. As shown in Fig. 2, l-piperazine-2-carboxylic acid derived N-formamide 2a,12aR-(+)-tert-butylsulfinamide derived catalyst 2b12b and picolinamide 2c12c were found to be totally inactive for the reaction. Meanwhile picolinamide–tosylate catalyst 2d was highly active to afford the product with excellent yield in moderate ee value. Afterwards, several picolinamide–tosylate catalysts 2f–2h bearing electron-withdrawing group in 4-position of pyridine were employed in the reaction and 4-bromo picolinamide 2f delivered a slightly higher ee value. 5-Methoxy picolinamide 2i gave the product with the same ee value as that of 2d but in much lower yield. Lower enantioselectivities were observed with 4-phenyl picolinamide 2j and 3-methyl picolinamide 2k. When (R)-1-(2-aminonaphthalen-1-yl)naphthalen-2-ol derived catalyst 2l was used, no product was observed. When (1S,2R)-1-amino-2,3-dihydro-1H-inden-2-ol derived catalyst 2m gave the product in both poor yield and ee value. Moreover, two picolinamide–sulfonamide catalysts 2n and 2o were also used in the reaction and almost racemic products were obtained. Hence, 2f was determined as the optimal catalyst and was used through out our study.Open in a separate windowFig. 2Evaluation of the chiral Lewis base catalysts 2 in conjugate hydrosilylation of (E)-1,3-diphenylbut-2-en-1-one 1a. Unless otherwise specified, the reactions were carried out with 1a (0.1 mmol), trichlorosilane (0.2 mmol) and catalyst 2 (0.02 mmol) in 1 mL of acetonitrile at 0 °C for 24 hours. Isolated yield based on 1a. The ee values were determined by using chiral HPLC.Subsequently, the other reaction conditions were optimized. The results are summarized in EntryaSolvent T (°C)Time [h]Yieldb [%]eec,d [%]1CH3CN0247751 (R)2CH3CH2CN0248450 (R)3C6H5CN0248324 (R)4THF0249211 (R)51,4-Dioxane024635 (R)6CHCl30249225 (S)7CH2Cl20249638ClCH2CH2Cl0249010 (R)9CCl40249535 (S)10Toluene0249555 (S)11Xylene024N.R.—12Mesitylene024N.R.—13C6H5CF30249531 (S)14Toluene−10489764 (S)15Toluene−20609265 (S)16Toluene−4072N.R.—Open in a separate windowaUnless otherwise specified, the reactions were carried out with 1a (0.1 mmol), trichlorosilane (0.2 mmol) and catalyst 2f (0.02 mmol) in 1 mL of solvent.bIsolated yield based on 1a.cThe ee values were determined by using chiral HPLC.dThe absolute configuration of 3a was determined by comparison of the retention times of the two enantiomers on the stationary phase with those in the literatures.With the optimized conditions in hand, the scope and limitations of the reaction were explored. The results are summarized in Fig. 3. In the presence of 20 mol% of chiral Lewis base catalyst 2f and 2 equivalents of trichlorosilane, various α,β-unsaturated ketones were hydrosilylated. We first tested the effect of various 3-aryl groups of 1. 3-(4-Methoxy-phenyl) substrate 3e (Fig. 3) and 3-(naphthalen-2-yl) substrate 3i (Fig. 3) underwent the reaction to give the products with good yields in enantioselectivities close to 3a, while lower ee values were observed with 3b–3d, 3h, 3k and 3l (Fig. 3). No product was obtained with 3-(2-chloro-phenyl) substrate 3f and 3-(naphthalen-1-yl) substrate 3j, perhaps due to the high steric hindrance (Fig. 3). When 3-(2-methoxy-phenyl) substrate 3g was used, by prolonging the reaction time to 60 hours, the product was obtained with good yield but very poor enantioselection (Fig. 3). Trace amount of the product was detected with 3-(pyridin-2-yl) substrate 3m (Fig. 3). Next, some 3-phenyl-but-2-en-1-ones with different 1-aryl groups were also employed in the reaction. The 4-substituted or 3-substituted substrates delivered good yields of the products with similar or slightly lower enantioselectivities (Fig. 3), while much lower ee value was observed with 2-substituted substrate 3s (Fig. 3). For some other 3-alkyl chalcones, moderate to good yields and moderate ee values were obtained (Fig. 3), except for the bulkier tertiary butyl substituted 3w that gave trace amount of the product (Fig. 3). Reaction of 3y with two different 3,3-aryl groups afforded the product with moderate yield in very low enantioselectivity (Fig. 3). Finally, cyclic substrate 3z was subjected in the reaction and provided the product with excellent yield but in poor ee value (Fig. 3).Open in a separate windowFig. 3Substrate scope of the reaction. Unless otherwise specified, the reactions were carried out with 1 (0.1 mmol), trichlorosilane (0.2 mmol) and catalyst 2f (0.02 mmol) in 1 mL of toluene at −10 °C for 48 hours. Isolated yield based on 1. The ee values were determined by using chiral HPLC. The absolute configuration of 3a was determined by comparison of the retention times of the two enantiomers on the stationary phase with the literatures. The absolute configurations of other products were determined in analogy. aThe reaction time was 60 hours. bThe reaction time was 72 hours.Although detailed structural and mechanistic studies remain to be carried out, based on the absolute configuration of the product 3a, we propose a mechanism shown in Scheme 2. First, the nitrogen atom of the pyridine ring and the carbonyl oxygen atom of catalyst 2f are coordinated to Cl3SiH to create an activated hydrosilylation species. Substrate 1a may approache the Cl3SiH-catalyst complex to generate two transition states A and B. In transition state A, the N–H of catalyst 2f activates the carbonyl group of 1a through H-bonding. In addition, there could be π–π stackings between the two aromatic systems of the catalyst and the substrate. Then Si-face conjugate attack of the hydride to 1a generate (S)-product 3a. On the contrary, in transition state B through which (R)-product will be obtained, the carbonyl group of 1a can not be connected with the N–H of catalyst 2f. Thus the fact that (S)-enriched product 3a was obtained is consistent with the suggestion that the hydrosilylation predominantly proceeds through the pathway involving transition state A rather than transition state B.Open in a separate windowScheme 2A plausible reaction mechanism for hydrosilylation of 1a catalyzed by 2f.In conclusion, we have developed a facile, metal-free and mild enantioselective conjugate hydrosilylation of β,β-disubstituted α,β-unsaturated ketones. By using chiral picolinamide–sulfonate Lewis base as catalyst, the reactions provided various optically active ketones bearing a chiral center at β-position with moderate to good yields in moderate enantioselectivities. Comparing with the chiral phosphine oxide Lewis base catalysts, the chiral picolinamide–sulfonate is cheaper and easier accessible. The absolute configuration of one product was determined by comparison of the retention times of the two enantiomers on the stationary phase with those in the literature.  相似文献   

5.
Mesoporous PbO nanoparticle-catalyzed synthesis of arylbenzodioxy xanthenedione scaffolds under solvent-free conditions in a ball mill     
Trimurti L. Lambat  Ratiram G. Chaudhary  Ahmed A. Abdala  Raghvendra Kumar Mishra  Sami H. Mahmood  Subhash Banerjee 《RSC advances》2019,9(54):31683
A protocol for the efficient synthesis of arylbenzodioxy xanthenedione scaffolds was developed via a one-pot multi-component reaction of aromatic aldehydes, 2-hydroxy-1,4-naphthoquinone, and 3,4-methylenedioxy phenol using mesoporous PbO nanoparticles (NPs) as a catalyst under ball milling conditions. The synthesis protocol offers outstanding advantages, including short reaction time (60 min), excellent yields of the products (92–97%), solvent-free conditions, use of mild and reusable PbO NPs as a catalyst, simple purification of the products by recrystallization, and finally, the use of a green process of dry ball milling.

An efficient one-pot multicomponent protocol was developed for the synthesis of arylbenzodioxy xanthenedione scaffolds using mesoporous PbO nanoparticles as reusable catalyst under solvent-free ball milling conditions.

Recently, the ball milling technique has received great attention as an environmentally benign strategy in the context of green organic synthesis.1a The process of “ball milling” has been developed by adding mechanical grinding to the mixer or shaker mills. The ball milling generates a mechanochemical energy, which promotes the rupture and formation of the chemical bonds in organic transformations.1b Subsequently, detailed literature1c and books on this novel matter have been published.2a,b Several typical examples include carbon–carbon and carbon–heteroatom bond formation,2c organocatalytic reactions,2d oxidation by using solid oxidants,2e dehydrogenative coupling, asymmetric, and peptide or polymeric material synthesis, which have been reported under ball milling conditions.2e Hence, the organic reactions using ball milling activation carried out under neat reaction environments, exhibit major advantages,2f including short reaction time, lower energy consumption, quantitatively high yields and superior safety with the prospective for more improvement than the additional solvent-free conditions and clear-cut work-up.3–5On the other hand, the organic transformations using metal and metal oxide nanoparticles6 are attracting enormous interest due to the unique and interesting properties of the NPs.7,8,9a Particularly, PbO NPs9b provide higher selectivity in some organic reactions9c and find applications in various organic reactions, like Paal–Knorr reaction,10 synthesis of diethyl carbonate,11 phthalazinediones,12 disproportionation of methyl phenyl carbonate to synthesize diphenyl carbonate,13 the capping agent in organic synthesis, and selective conversion of methanol to propylene.14 In addition, the PbO NPs are also used in many industrial materials.15,16However, till date, PbO NPs have not been explored in MCRs leading to biologically important scaffolds. Among others, the xanthene scaffolds17 are one of the important heterocyclic compounds18 and are extensively used as dyes, fluorescent ingredients for visual imaging of the bio-molecules, and in optical device technology because of their valuable chemical properties.19 The xanthene molecules have conjointly been expressed for their antibacterial activity,20 photodynamic medical care, anti-inflammatory drug impact, and antiviral activity. Because of their various applications, the synthesis of these compounds has received a great deal of attention.21 Similarly, vitamin K nucleus22,23 shows a broad spectrum of biological properties, like anti-inflammatory, antiviral, antiproliferative, antifungal, antibiotic, and antipyretic.24a As a consequence, a variety of strategies24b have been demonstrated in the literature for the synthesis of xanthenes and their keto derivatives, like rhodomyrtosone-B,25a rhodomyrtosone-I,25b and BF-6 25c as well as their connected bioactive moieties. Few biologically active xanthene scaffolds are shown in (Fig. 1).Open in a separate windowFig. 1Some biologically important xanthenes and their keto derivatives.Due to the significance of these compounds, the synthesis of xanthenes and their keto derivatives using green protocols is highly desirable. Reported studies reveal that these scaffolds are synthesized by three-component condensations using p-TSA26 and scolecite27 as catalysts. However, these methods suffer from the use of toxic acidic catalysts like p-TSA, long reaction times (3 h), harsh refluxing26 or microwave reaction conditions,27 and tedious work-up procedures. The previously reported methods for the synthesis of xanthenediones are shown in Scheme 1.Open in a separate windowScheme 1Previous protocol for the synthesis of xanthenedione derivatives.Herein, we report an economical and facile multicomponent protocol, using ball milling, for the synthesis of 7-aryl-6H-benzo[H][1,3]dioxolo[4,5-b]xanthene-5,6(7H)-dione using PbO NPs as a heterogeneous catalyst (Scheme 2). The PbO NPs are non-corrosive, inexpensive, and easily accessible.Open in a separate windowScheme 2General reaction scheme of PbO NP-catalyzed synthesis of the xanthenedione scaffolds under ball milling conditions.In our protocol,28 the PbO NPs were initially prepared by mixing sodium dodecyl sulphate (2.5 mmol) and sodium hydroxide (10 mL, 0.1 N) with an aqueous methanolic solution of lead nitrate (2 mmol) under magnetic stirring at 30 °C by continuing the reaction for 2 h. Then, the obtained white polycrystalline product was filtered, washed with H2O, and dried at 120 °C, followed by calcination at 650 °C for 2 h. During this step, the white PbO NPs turned pale yellow in colour. Eventually, the synthesized PbO was then characterized by spectroscopic and analytical techniques.The powder X-ray diffraction (XRD) pattern revealed the crystalline nature of the PbO NPs as the diffraction peaks corresponding to (131), (311), (222), (022), (210), (200), (002), and (111) crystal planes were identified (Fig. 2). The XRD outline of the synthesized PbO NPs was further established for the formation of space group Pca2129 with a single orthorhombic structure (JCPDS card number 76-1796). The sharp diffraction peaks indicated good crystallinity, and the average particle size of the PbO NPs was estimated to be 69 nm, as calculated using the Debye–Scherer equation.Open in a separate windowFig. 2The powder XRD pattern of PbO NPs.The surface morphology of the PbO NPs was analyzed by scanning electron microscopy (SEM), and the SEM image revealed the discrete and spongy appearance of the PbO NPs (Fig. 3).Open in a separate windowFig. 3The SEM image of PbO NPs.Moreover, the elemental composition obtained from energy dispersive X-ray (EDX) analysis confirmed that the material contains Pb and O elements, and no other impurity was present (Fig. 4).Open in a separate windowFig. 4The EDAX spectrum of crystalline PbO NPs.The transmission electron microscopy (TEM) image shown in Fig. 5 indicated the formation of orthorhombic crystallites of PbO with several hexagon-shaped particles. The dark spot in the TEM micrograph further confirmed the synthesis of PbO NPs, as the selected area diffraction pattern associated with such spots reveals the occurrence of the PbO NPs in total agreement with the X-ray diffraction data (Fig. 6). The average size of the PbO nanocrystals by TEM was approximated to be around 20 nm.Open in a separate windowFig. 5The TEM image of nanocrystalline PbO NPs.Open in a separate windowFig. 6The SAED image of nanocrystalline PbO NPs.The Fourier transform infrared (FT-IR) spectrum (ESI, S6) of the PbO NPs displayed peaks at 575, 641, and 848 cm−1, which corresponds to the Pb–O vibrations. Furthermore, the absorption band at ∼3315 cm−1 was due to the presence of the hydroxyl group (–OH) in the NPs.The N2 adsorption–desorption isotherms of the PbO nanoparticles shown in Fig. 7 was consistent with type IV adsorption–desorption isotherms with H1 hysteresis corresponding to the cylindrical mesoporous structure. Moreover, the surface area, pore-volume, and BJH pore diameter were found to be 32.0 m2 g−1, 0.023 cm3 g−1, and 30.9 Å, respectively.Open in a separate windowFig. 7BET surface area and pore size of nanocrystalline PbO catalyst.The catalytic activity of the synthesized PbO NPs was tested in a one-pot multicomponent synthesis of arylbenzodioxoloyl xanthenedione derivative under ball milling condition according to the reaction scheme 2a, with 3,4-dimethoxybenzaldehyde (166.2 mg, 1.0 mmol), 3,4-methylenedioxyphenol (138.0 mg, 1.0 mmol), and 2-hydroxy-1,4-naphthoquinone (174.0 mg, 1.0 mmol) as reactants. The reaction conditions, the ball milling parameters (speed, time, and ball to solids ratio), and the PbO nanocatalyst amount were first optimized to produce the highest yield using experimental design as shown in EntryConditionsRotation (rpm)Catalyst (mol%)Time (min)Yield (%)a1Ball milling4000050212Ball milling4001050483Ball milling4001560544Ball milling4002070595Ball milling5001050626Ball milling5001550657Ball milling5002060678Ball milling6001070719Ball milling60015507710Ball milling60020608211Ball milling60005709012Ball milling600105091 13 Ball milling b 600 15 60 97 14Ball milling60020709715No ball millingc—1560—Open in a separate windowaIsolated yield; model reaction: 3,4-dimethoxybenzaldehyde (166.2 mg, 1.0 mmol), 3,4-methylenedioxyphenol (138.1 mg, 1.0 mmol), 2-hydroxy-1,4-naphthoquinone (174.1 mg, 1.0 mmol) under ball milling.bOptimized reaction conditions.cThe reaction was performed under stirring condition in a RB flask.Next, by utilizing the general experimental procedure (ESI for detail experimental procedure; S2) and the aforementioned optimized conditions (29 we also investigated the possible scopes of the reactants as revealed in 26 These data are available in S4 (see ESI for the spectroscopic data). The aromatic aldehydes comprising both electron-withdrawing (e.g., nitro group) and electron-donating (e.g., –OMe, –OH, –Cl, –Me, and –Br) groups participated proficiently in the reaction without including any electronic effects. The aromatic aldehyde with electron-donating groups (e.g., –OMe, –OH, –Cl, –Me, and –Br) increased the product yield, while in the case of aryl aldehyde having an electron-withdrawing group (e.g., –NO2), both the product yield as well as the reaction rate decreased. These findings are depicted in Scope of the PbO NP-catalyzed synthesis of arylbenzodioxoloyl xanthenedione derivatives
Open in a separate windowFollowing a previously reported mechanism,26 a possible mechanism for the synthesis of arylbenzodioxoloyl xanthenedione derivative under ball milling at 600 rpm for 60 min is shown in Scheme 3. It is speculated that in the first step, the surface of the PbO NPs having free –O–H groups facilitated the carbon–carbon bond formation by activating aromatic aldehyde 1a to react with 2-hydroxy-1,4-naphthoquinone 1b leading to the intermediate B, which further undergoes dehydration, followed by the addition of 3,4-methylenedioxyphenol 1c, which upon cyclization leads to the formation of the product 2a with the recovery of the catalyst, PbO NPs.Open in a separate windowScheme 3Plausible mechanism of PbO NP-catalyzed synthesis of arylbenzodioxoloyl xanthenedione (2a).Further, to signify the advantages of the current methodology, a comparative study of known methods is provided in Sr. no.CatalystReaction conditionsYield (%)Time (min)Reusable?1 p-TSA26EtOH/90–120 °C85–90180No2Scolecites27EtOH/80 oC90–924–15 MWYes up to 3 cycles3 aPbO NPsAmbient temperature92–9760Yes up to 8 cyclesOpen in a separate windowaPresent work.Next, we investigated the reusability of the PbO nanocatalyst for the synthesis of 7-(3,4-dimethoxyphenyl)-6H-benzo[H][1,3]dioxolo[4,5-b]xanthene-5,6(7H)-dione (2a) as a model reaction. After the reaction, PbO NPs were separated from the reaction mixture by centrifugation, washed consecutively with aqueous ethanol, dried, and reused for the next run. As shown in Fig. 8, the reaction yield was reduced by only 12% after eight consecutive runs. This slight decrease in the yield was observed due to the loss of PbO NPs (∼10 wt%) during the recycling process.Open in a separate windowFig. 8Reusability of PbO NPs for the synthesis of 7-(3,4-dimethoxyphenyl)-6H-benzo[H][1,3]dioxolo[4,5-b]xanthenes-5,6 (7H)-dione as a model reaction.The fate of the recycled PbO NPs was analyzed by performing SEM and TEM studies after the 8th run, and considerable agglomeration of NPs was observed. However, interestingly the particle size of the NPs reduced to ∼15 nm compared to fresh PbO NPs during the ball milling process (Fig. 9).Open in a separate windowFig. 9(a) SEM and (b) TEM images of the recycled PbO NPs after 8th run.In conclusion, we demonstrated a facile and efficient method for the synthesis of 7-aryl-6H-benzo[H][1,3]dioxolo[4,5-b]xanthene-5,6(7H)-dione using PbO NPs as a catalyst. The entire synthesis process was very clean and provided very high yields (86–97%) of xanthenedione derivatives (2a–l) via mild ball milling. Moreover, the present protocol has demonstrated significant development in terms of higher isolated yields, faster rate of reaction (1 h), and most importantly, it is environment-friendly. Moreover, the use of solvent-free ball milling conditions allows simple isolation and purification of the products, with no column chromatography, as well as the mild PbO NPs as a reusable catalyst made the current synthetic method more suitable and environmentally benign in nature.  相似文献   

6.
Correction: Base recognition by l-nucleotides in heterochiral DNA     
Shuji Ogawa  Shun-ichi Wada  Hidehito Urata 《RSC advances》2019,9(17):9692
Correction for ‘Base recognition by l-nucleotides in heterochiral DNA’ by Shuji Ogawa et al., RSC Adv., 2012, 2, 2274–2275.

The authors regret that some of the data in the original article were presented incorrectly. Some of the oligonucleotide sequences in the Graphical Abstract, Fig. 2 and Fig. 2 and Open in a separate windowFig. 2Effects of base pair mismatch of d- (a–d) and l-nucleotide (e–h) on duplex stability. Samples contained 6 mM duplex in 10 mM MgCl2, 100 mM NaCl, and 70 mM MOPS (pH 7.1). Yellow bars denote Tm values of fully matched duplexes, and blue bars denote Tm values of mismatched duplexes.UV-melting points of homo- and heterochiral duplexesa
DuplexTemplate strandComplementary strand T m (°C)ΔTmb (°C)
Homochiral strand
1d(AAATCTGCG)d(CGCAGATTT)42.1
Heterochiral strand
2d(AAlATCTGCG)d(CGCAGATTT)33.6−8.5
3d(AAATCTlGCG)d(CGCAGATTT)32.6−9.5
4d(AAATlCTGCG)d(CGCAGATTT)38.2−3.9
5d(AAAlTCTGCG)d(CGCAGATTT)33.9−8.2
Open in a separate windowaSamples contained 6 μM duplex in 10 mM MgCl2, 100 mM NaCl, and 70 mM MOPS (pH 7.1).bMelting temperature difference from the homochiral duplex.The Royal Society of Chemistry apologises for these errors and any consequent inconvenience to authors and readers.  相似文献   

7.
Highly dispersed Ni nanoparticles on mesoporous silica nanospheres by melt infiltration for transfer hydrogenation of aryl ketones     
Hyemin Kweon  Sanha Jang  Akerke Bereketova  Ji Chan Park  Kang Hyun Park 《RSC advances》2019,9(25):14154
Nickel-based catalysts have been applied to the catalytic reactions for transfer hydrogenation of carbonyl compounds. In the present work, highly dispersed nickel particles located at the pores of mesoporous silica spheres (Ni@mSiO2) were prepared via an optimized melt infiltration route. The nickel nanoparticles of 10 wt% in the Ni@mSiO2 catalyst could be uniformly loaded with high dispersion of 36.3%, resulting excellent performance for catalytic transfer hydrogenation of aryl ketones.

Nickel-based catalysts have been applied to the catalytic reactions for transfer hydrogenation of aryl ketones.

The reduction of carbonyl compounds has received steady interest as a common transformation in organic synthetic chemistry. Corresponding alcohols are important building blocks for the manufacture of chemicals, pharmaceuticals, and cosmetics.1,2 The transfer hydrogenation (TH) reaction of ketones has lots of advantages owing to its low cost, simple process, easy handling, and mild conditions.3 Isopropyl alcohol both as a solvent and hydrogen donor provides safe reaction conditions.4 Until now, many catalysts have been developed for TH reactions.5Among the catalysts for TH reactions, nickel-based catalysts are representative, due to the metallic nickel surface can absorbs hydrogen and easily activates hydrogen in the atomic state.6–11 In addition, nickel compounds could be the significant catalyst candidate because of the low price, accessibility, and high reactivity.12,13 Therefore, several works used nickel nanocatalysts for TH reactions.14–17,44–51 However, developing the new supported catalyst with the high metal dispersion and narrow particle size distribution is still required, because it can enhance the reactant conversion and improve product yield.18 Recently, melt infiltration process has been exploited to prepare supported nanocatalysts as a fast and convenient route with no solvent use.19–21The method which uses highly porous supports with regular porosity and hydrated metal salts with mild melting temperatures, allowed the metal salts to permeate inside the porous support via capillary forces. High performance nanocatalysts based on well-dispersed nanoparticles could be obtained through uniform infiltration of metal precursors and sequential thermal treatment.22 The obtained nanoparticles have narrow size distributions as well as small sizes, enlarging active sites.23,24Mesoporous silica nanosphere (mSiO2) as a porous support can be an outstanding platform for the synthesis of metal nanoparticle, owing to the good thermal stability, high surface area, and large pore volume as well as uniform pore structures.25–29 Herein, we report a new catalyst with highly dispersed Ni nanoparticles at mesoporous silica sphere nanostructures (Ni@mSiO2) for catalytic TH reactions, which was synthesized through a melt infiltration of hydrated nickel salts and a sequential thermal reduction (Scheme 1). First, mSiO2 supports could be prepared by a sol–gel method based on a modified Stöber process. More details were described in the ESI. The mesopores were formed in the silica nanospheres using the cationic surfactant C16TAB as a self-assembly template. The TEM images show mesoporous silica spheres with an average diameter of 417 ± 20 nm (Fig. 1a and b). Ordered mesoporous channels were generated by removing the long carbon chains of C16TAB by calcination at 500 °C. The corresponding Fourier-transform (FT) pattern showed a ring pattern of polycrystalline (inset of Fig. 1c). The inverse fast Fourier-transform (IFFT) image, calculated by Micrograph TM Gatan software, also demonstrated the existence of hexagonal lattice planes (Fig. 1c).Open in a separate windowScheme 1A brief synthetic scheme of Ni@mSiO2 nanocatalyst.Open in a separate windowFig. 1(a and b) TEM images, (c) FT pattern for (b) (inset of c) and IFFT image, (d) HRTEM image, (e) N2 adsorption/desorption isotherms, and (f) pore size distribution diagrams of mSiO2. The bars represent 1 mm (a), 100 nm (b), and 20 nm (d).The brightest of the rings originated from a two-dimensional hexagonal (p6mm) structure with d100 spacing. The interplanar spacing computed from the brightest FFT diffraction ring was approximately 3 nm, corresponding to the pore sizes shown in the TEM image (Fig. 1d).N2 sorption experiments for the mSiO2 nanosphere exhibited type IV adsorption–desorption hysteresis with H4 type hysteresis loop (Fig. 1e). The Brunauer–Emmett–Teller (BET) surface area and pore volume of the mSiO2 nanosphere were 1501 m2 g−1 and 0.70 cm3 g−1, respectively. The average pore size was estimated to be approximately 2 nm from the adsorption/desorption branches of N2 isotherms by using the Barrett–Joyner-Halenda (BJH) method (Fig. 1f). Scheme 1 illustrated the simple procedure for the synthesis of the Ni@mSiO2 nanocatalyst. Based on the small pore size (3 nm) of mSiO2 nanospheres, very tiny nickel nanoparticles (∼2 nm) could be generated via a melt-infiltration process and thermal treatment under hydrogen flow. Using a 0.55 gnickel salt/gsilica support ratio, the hydrated nickel salt (melting point = 56.7 °C) was successfully incorporated in the mSiO2 nanosphere during the melt infiltration process, driven by the capillary forces. The Ni-loading content after the final thermal treatment was calculated to be nominally 10 wt% on the basis of Ni converted from the nickel nitrate salt. First, the hydrated nickel salt was melt-infiltrated into the mesoporous silica pores of the mSiO2 nanospheres by physical mixing at room temperature with subsequent aging at 60 °C for 24 h in a tumbling oven. Then, tiny nickel nanoparticles were generated by thermal reduction of the confined hydrated Ni salt in the small pores at 500 °C under hydrogen flow.The low-resolution TEM images of the Ni@mSiO2 nanostructure show Ni nanoparticles as black dots (Fig. 2a and b). The high-angle annular dark-field scanning transmission electron microscopy (HAADF-STEM) image showed bright dots of a very small size (average 2.0 nm), which indicates uniform incorporation of the Ni nanoparticles in the porous silica (Fig. 2c and S1 in the ESI). In the elemental mapping of silica (green color) and nickel (red color), high dispersion of nickel nanoparticles was observed (Fig. 2d). HRTEM analysis showed lattices of a single Ni nanoparticles in the mSiO2 pores (Fig. 2e). The inner particle size was observed to be around 2 nm. The Fourier-transform pattern represented a single crystal of metallic nickel with a distance of 0.2 nm between neighboring fringes, which corresponded to the (111) planes of face centered-cubic nickel (Fig. 2e).Open in a separate windowFig. 2(a and b) TEM and (c) HAADF TEM images, (d) scanning TEM image with elemental mapping and (e) HRTEM images, and (f) XRD spectrum of Ni@mSiO2 nanostructure. The bars represent 200 nm (a), 20 nm (b–d), and 2 nm (e).The X-ray diffraction (XRD) spectrum of Ni@mSiO2 nanocatalyst has a broad peak at 2θ = 44.5°, which is assigned to the reflections of the (111) plane in the fcc-nickel phase (Fig. 2f, JCPDS No. 04-0850). The average crystal size of the nickel particle was estimated to be 1.8 nm, from the broadness of the (111) peak using the Debye–Scherrer equation, which is well-matched with that observed in the TEM images. The other intense peak around 23° of the Ni@mSiO2 nanocatalyst indicates the presence of amorphous silica.Using a H2 chemisorption experiment, the active nickel surface area and average nickel particle size of the Ni@mSiO2 nanocatalyst could be analyzed, and measured 242 m2 g−1 and 2.78 nm, respectively. The obtained Ni dispersion was very high, and was calculated to be 36.3%.The XPS spectrum analysis was carried out to probe the chemical states of Ni@mSiO2 (Fig. S2). The Ni 2p peak of Ni@mSiO2 showed NiO and Ni peak. The four peaks at 853.9, 855.6, 860.7, and 863.5 eV are assigned to NiO, and the other peak at 852.1 eV is assigned metallic Ni peak. The metallic Ni surface of the small nanoparticles (∼2 nm) was easily oxidized under an ambient condition.N2 sorption experiments for the Ni@mSiO2 nanocatalyst showed a type IV isotherm with type H4 hysteresis (Fig. 3a). The BET surface area was calculated to be 636 m2 g−1. The total pore volume was found to be 0.3 cm3 g−1, which is about 43% of the initial mSiO2 nanosphere (0.7 cm3 g−1). The significant decrease in pore volume was attributed to inner nickel nanoparticle occupancy. Applying the BJH method, small pore sizes were obtained by the adsorption/desorption branches (Fig. 3b). Because of the occupancy of tiny nickel particles in the pristine silica pores, the pore size of the Ni@mSiO2 nanocatalyst was also slightly decreased, and was observed to be 1.8 nm.Open in a separate windowFig. 3(a) N2 adsorption/desorption isotherms and (b) pore size distribution diagrams of Ni@mSiO2 nanocatalyst.The Ni@mSiO2 nanocatalyst was applied to the hydrogen-transfer reaction of acetophenone. Acetophenone is an ideal substrate in the hydrogen transfer reaction, and is generally used as a hydrogen acceptor.30 Among the various products of acetophenone reduction, only 1-phenylethanol resulted from the catalyzed reaction. The metallic Ni on the catalyst surface is the active species and the reaction was promoted by base.17 The dihydride species referred in this catalyst system to make alcohol from the transfer of the two hydrogen atoms of the donor to the surface of the metal.31 However, small amount of various byproduct including hemiacetals and condensation products which is occurred between ketone and alcohol were produced due to basic conditions.32 For optimization, the reaction parameters, such as amount of catalyst and base, temperature, and time, are adjusted. The TH reaction results of acetophenone are summarized in EntryCat. (mol%)Temp. (°C)Time (min)Base (eq.)Conv.b (%)Yieldb (%)TON1111045110095952111060176737331100602100999941100601100929251100600.562595961906011009696718045158555581806017268689180751100979710180901676565110.58075110096192120.25807511095380130.1807516260603140.05100602806262Open in a separate windowaRxn. condition: acetophenone (2 mmol), i-PrOH (solvent, 10 mL), base (NaOH).bDetermined by GC-MS spectroscopy.cnickel-aluminium alloy purchased from Lancaster (10034177) was applied.First, we studied the effect of the base and determined the amount of base required (Entries 3–6, 33,34 Although no reaction was observed without bases, a small amount of base was sufficient to trigger the reaction (Entry 5, 6,15,35 The optimized conditions were applied to extend the scope.To verify the general applicability of Ni@mSiO2, various aromatic ketones were tested ().36–43Catalytic transfer hydrogenation reactions of various aromatic carbonyl compounds with Ni@mSiO2a
EntryCompoundConv. (%)Sel. (%)Yield (%)
1 78b, 100c, 10033b, 75c, 6526b, 75c, 65
2 54, 68d, 44e92, 97d, 100e50, 66d, 44e
3 55, 64d57, 55d31, 35d
4 738864
5 8, 35f88, 97f7, 34f
6 40, 45e97, 93e39, 42e
7 34, 56d93, 93d32, 52d
8 55, 67e76, 57e42, 38e
9 556134
10 78, 99e96, 97e75, 96e
11 69, 93e95, 100e66, 96e
12 75, 95e64, 84e48, 79e
13 71, 95e78, 97e56, 92e
14 63, 91e64, 53e40, 48e
Open in a separate windowaCat. (0.25 mol%), base NaOH (1 eq.), rxn. temp. 80 °C, rxn. time 75 min, determined by GC-MS spectroscopy.bRxn. time 15 min.cRxn. time 30 min.dRxn. time 2 h.eRxn. time 3 h.fRxn. temp. 100 °C.In the TEM image of the recovered Ni@mSiO2 nanocatalyst shown totally collapsed silica structure and sintering nickel nanoparticles (Fig. S5a and b). In the stability of structure, silica was occurred procreated due to the steam during the reaction. In the XRD data of the recovered Ni@mSiO2 nanocatalyst showed sharp peaks (Fig. S5c). It means reflected the increased crystal size. In addition, metallic nickel phase changed nickel silicide and nickel carbide during the reaction.The recycle reaction of the recovered Ni@mSiO2 catalyst showed a low conversion (79%) compared high conversion (100%) of fresh Ni@mSiO2 for 1 h. In the recycle reaction, the catalyst exhibited lower performance due to metallic nickel was changed no active site catalyst such as nickel carbide and nickel silicide.  相似文献   

8.
Continuous flow synthesis of aryl aldehydes by Pd-catalyzed formylation of phenol-derived aryl fluorosulfonates using syngas     
Manuel Kckinger  Paul Hanselmann  Guixian Hu  Christopher A. Hone  C. Oliver Kappe 《RSC advances》2020,10(38):22449
This communication describes the palladium-catalyzed reductive carbonylation of aryl fluorosulfonates (ArOSO2F) using syngas as an inexpensive and sustainable source of carbon monoxide and hydrogen. The conversion of phenols to aryl fluorosulfonates can be conveniently achieved by employing the inexpensive commodity chemical sulfuryl fluoride (SO2F2) and base. The developed continuous flow formylation protocol requires relatively low loadings for palladium acetate (1.25 mol%) and ligand (2.5 mol%). Good to excellent yields of aryl aldehydes were obtained within 45 min for substrates containing electron withdrawing substituents, and 2 h for substrates containing electron donating substituents. The optimal reaction conditions were identified as 120 °C temperature and 20 bar pressure in dimethyl sulfoxide (DMSO) as solvent. DMSO was crucial in suppressing Pd black formation and enhancing reaction rate and selectivity.

Pd-catalyzed formylation of aryl fluorosulfonates by using syngas as an atom efficient and sustainable source of carbon monoxide and hydrogen.

Aryl aldehydes are ubiquitous intermediates and building blocks in the chemical and pharmaceutical industry. As such, their convenient and cost-efficient synthesis is of high interest. One of the simplest ways of forming aldehydes is through the Pd-catalyzed formylation of Ar–X (where X = I, Br, or OTf) using syngas (CO and H2).1 The main limitations in the case of aryl iodides and bromides are the price and availability of the starting materials. In the case of triflates, trifluoromethanesulfonic anhydride (Tf2O) is the reagent most commonly employed to prepare triflates from alcohols, but it is relatively expensive, prone to hydrolysis and atom uneconomic.2 Fluorosulfonates (–OSO2F) have been proposed as an alternative leaving group to triflates because their reactivity is considered largely the same,2,3 or placed in-between bromides and chlorides.4 The fluorosulfonate leaving group is an emerging chemical motif5 that has been used in many types of chemical transformations including reduction,6 metal catalyzed cross-coupling,3,5,7 deoxyfluorination,8 amination,9 and methoxycarbonylation.10 Aryl fluorosulfonates can be conveniently and inexpensively prepared by treating readily available phenols with the commodity chemical sulfuryl fluoride (SO2F2) and base.Previously, Beller and co-workers successfully demonstrated the Pd-catalyzed formylation of aryl triflates with syngas as a sustainable and cost effective reagent under batch conditions (Scheme 1a).11 We were interested in the synthesis of aryl aldehydes from aryl fluorosulfonates, which could be derived from their corresponding phenols. Our group previously reported the continuous flow Pd-catalyzed formylation of (hetero)aryl bromides within a continuous flow system.12 Carbonylation reactions benefit from the highly efficient mixing, enhanced mass and heat transfer characteristics, precise residence time, accessibility to high pressure and temperature regimes, and the high operational safety of continuous flow reactors.13–16 Herein, we describe the development of a continuous flow protocol for the synthesis of aryl aldehydes from their corresponding aryl fluorosulfonates (Scheme 1b).Open in a separate windowScheme 1Pd-catalyzed reductive carbonylation using syngas for (a) triflates (Beller)11 and (b) fluorosulfonates (this work) as leaving groups.We commenced our investigation by evaluating the catalytic system palladium(ii) acetate (Pd(OAc)2) and 1,3-bis(diphenylphosphino)propane (dppp), which was utilized by Beller and co-workers for the batch formylation of aryl triflates.11,17 The reaction was optimized within a continuous flow reactor for 4-methoxyphenyl sulfurofluoridate (1a) as model substrate. We used a Uniqsis FlowSyn flow system consisting of two HPLC pumps and a heated reactor coil (32 mL). Two sample loops were used to deliver the substrate solution (2 mL) and catalyst feed solution (3 mL). The gases, H2 and CO, were introduced into a four-way mixer by using two mass flow controllers (MFCs). The flow system was pressurized to 10 bar by using a back pressure regulator (BPR). Substrate 1a was mixed with triethylamine (Et3N) as base and diphenyl ether (Ph2O) as internal standard for preparation of the substrate feed. Pd(OAc)2 and dppp were used as the catalyst feed. Initially, we evaluated toluene (PhMe), tetrahydrofuran (THF), acetonitrile (MeCN) and dimethylformamide (DMF) as solvents for the carbonylation reaction ( EntrySolventTemp. [°C]Conv. 1ab [%]Sel.b [%]Yieldb2a [%]1PhMe1007.7002PhMe12017.12.20.43THF1009.66.90.74THF12017.04.90.85MeCN10020.73.05.96MeCN12042.75.93.47DMF10067.119.112.88DMF12079.122.117.59DMFc12050.322.611.410DMFd12036.759.021.611DMFd,e12032.850.016.4Open in a separate windowaGeneral conditions: feed 1: 0.2 M 4-methoxy sulfurofluoridate (1a), 1.0 equiv. Et3N and 0.15 equiv. Ph2O in solvent; feed 2: 5 mol% Pd(OAc)2 and 10 mol% dppp in solvent. Feed 1/feed 2/H2/CO = 0.1 : 0.1 : 5 : 5 mL min−1 resulting in a 27 min residence time (tres). 10 bar system pressure.bAnalyzed by GC-FID.c13.5 min tres.d1 equiv. of pyridine as base.eFlow rates for feed 1/feed 2/H2/CO = 0.3 : 0.3 : 1.87 : 1.87 mL min−1, tres 35 min.Higher conversion was typically observed when using 120 °C instead of 100 °C. Using PhMe and THF as solvent gave only minimal conversion and trace amount of product (Fig. 1). The drop in yield and black output solution indicates a well-known phenomenon of Pd0 – agglomerating and forming clusters.18 These clusters then irreversibly precipitate as Pd black particles. These Pd black particles coat the reactor wall and can catalyze further Pd black formation.12,19 In order to tackle this issue, we looked for an additive or a method to prevent or slow the rate of Pd black agglomeration.Open in a separate windowFig. 1Drop in 4-methoxybenzaldehyde (2a) GC yield after repeated reaction runs without wash runs in-between. Conditions used are provided in 20 When applied to our reaction system using 4-chloro sulfurofluoridate (1b) as substrate, the use of DMSO as co-solvent alleviated the issues associated with Pd black formation (Fig. 2). Furthermore, we could drastically increase the yield of desired product 2b, due to reduced catalyst decomposition. The only side product in all cases was the hydrogenated product 3b, with the loss of the fluorosulfonate group (-OSO2F). Increased pressure (20 bar) and base (1.5 equiv.) resulted in higher selectivity to the aldehyde product 2b (Fig. S1). The fraction of DMSO solvent could be increased to no more than 80% owing to catalyst solubility.Open in a separate windowFig. 2Influence of DMSO on conversion and yield for Pd-catalyzed formylation of 4-chloro sulfurofluoridate (1b) with syngas measured by GC-FID (Ph2O as internal standard). For 0–80 vol% DMSO, 0.2 M 4-chloro sulfurofluoridate (1b), 1.0 equiv. pyridine and 0.15 equiv. Ph2O in DMF/DMSO; feed 2: 5 mol% Pd(OAc)2 and 10 mol% dppp in DMF/DMSO. 120 °C temperature and 10 bar system pressure. Flow rates for feed 1/feed 2/H2/CO = 0.3 : 0.3 : 1.87 : 1.87 mL min−1, corresponding to tres 35 min. 100 vol% DMSO experiment was performed using 1.25 mol% Pd(OAc)2 and 2.5 mol% dppp.The reaction conditions in terms of residence time and catalyst loading were then re-optimized for using DMSO as solvent. The catalyst loading was lowered to 1.25 mol% Pd(OAc)2 and 2.5 mol% dppp to ensure full dissolution of the catalyst system. A further decrease in loading did not provide satisfactory results, with a drop in yield observed (Fig. S2). Under these conditions 40 min of residence time was sufficient to achieve a high yield of 4-chlorobenzaldehyde (2b) (Fig. S3). With these new conditions (Scheme 2a), a long run was successfully operated which was stable for 4 hours with no apparent drop in yield, thus demonstrating no or minimal catalyst decomposition (Scheme 2b).Open in a separate windowScheme 2(a) Continuous flow setup for the Pd-catalyzed formylation of 4-chloro sulfurofluoridate (1b) long run. (b) 4-Chlorobenzaldehyde (2b) GC yield at 30 min intervals over 4 h operation time. 0.15 equiv. Ph2O was used as an internal standard.The applicability of the optimized conditions, shown in Scheme 2a, was demonstrated on a range of substrates. In most cases, substrates bearing an electron withdrawing group (EWG) in the para position underwent full conversion and the corresponding aryl aldehyde was formed in moderate to excellent yields (Fig. 3). The hydrogenated product was the sole side product observed. However, the reaction of substrates 1e and 1f resulted in the formation of a mixture of double formylation and hydrogenated products. The formylation of 4-nitro sulfurofluoridate (1 m) was problematic because of deactivation of the catalyst.Open in a separate windowFig. 3Substrate scope for aryl fluorosulfonates containing electron withdrawing groups. Yields were determined by GC-FID using Ph2O as internal standard. Yields in parentheses were isolated yield. Conditions are the same as given in Scheme 2a.Less reactive substrates, 1k and 1l, were only fully converted when modified reaction conditions were applied. In these cases, the liquid flow rates were lowered, which prolonged the residence time to approximately 2 h (Fig. 4a). Moreover, the H2 flow rate was modified to deliver 4.2 equiv. The same modified conditions were also used for substrates containing electron donating groups (EDG). The formation of the hydrogenated side product was not a significant problem for substrates containing EDG substituents with the aldehyde products obtained in good to excellent yields (Fig. 4b). Gratifyingly, 6-methoxy-2-naphthalaldehyde (2p) could be isolated in 82% yield after purification by column chromatography. Substrate 2b is a potential precursor to naproxen, an important non-steroidal anti-inflammatory active pharmaceutical ingredient.Open in a separate windowFig. 4(a) Unreactive substrates containing electron withdrawing groups; (b) substrates containing electron donating groups. For 43 min conditions, see Scheme 2a. Conditions: feed 1: 0.2 M fluorosulfonate, 1.5 equiv. pyridine and 0.15 equiv. Ph2O in DMSO; feed 2: 1.25 mol% Pd(OAc)2 and 2.5 mol% dppp in DMSO. 120 °C temperature and 10 bar system pressure. Feed 1/feed 2/H2/CO = 0.08 : 0.08 : 1.5 : 0.5 mL min−1, corresponding to a tres = 123 min.Substrates bearing a substituent in ortho position to the fluorosulfonate group afforded the corresponding aldehyde either in low yield, for 1r and 1u, or no yield, for 1d and 1o. The drop in yield could be caused by steric hindrance from the ortho-substituent. The steric block on the catalytic center will favor the addition of the much smaller hydrogen molecule instead of a larger CO molecule, resulting in the hydrogenated product.In conclusion, we have described a continuous flow method for the Pd-catalyzed synthesis of valuable aryl aldehyde building blocks from aryl fluorosulfonates and syngas as an inexpensive, atom-economic, and environmentally friendly source of CO and H2. The continuous flow approach enabled the precise addition of gas by using mass flow controllers. Meta and para substituted aryl fluorosulfonates could be converted to their corresponding aldehydes in good to excellent yields. Catalyst decomposition was successfully avoided by using DMSO as solvent. DMSO coordinates to Pd0 and facilitates the re-oxidation to PdII. Additionally, reaction rates and selectivity could be enhanced by the use of DMSO. Starting materials for the reductive carbonylation could be conveniently derived in excellent yields from readily-available phenols and the commodity chemical sulfuryl fluoride. The developed process is especially appealing for the chemical industry, where reagent cost and availability are important factors in process feasibility.  相似文献   

9.
Palladium catalysed carbonylation of 2-iodoglycals for the synthesis of C-2 carboxylic acids and aldehydes taking formic acid as a carbonyl source     
Ajaz Ahmed  Nazar Hussain  Monika Bhardwaj  Anuj Kumar Chhalodia  Amit Kumar  Debaraj Mukherjee 《RSC advances》2019,9(39):22227
Pd catalyzed carbonylative reaction of 2-iodo-glycals has been developed taking formic acid as a carbonyl source for the synthesis of 2-carboxylic acids of sugars by the hydroxycarbonylation strategy. The methodology was successfully extended to the synthesis of 2-formyl glycals by using a reductive carbonylation approach. Both ester and ether protected glycals undergo the reaction and furnished sugar acids in good yield which is otherwise not possible by literature methods. The C-2 sugar acids were successfully utilized for the construction of 2-amido glycals, 2-dipeptido-glycal by Ugi reaction and C-1 and C-2 branched glycosyl esters.

Pd catalyzed carbonylative reaction of 2-iodo-glycals has been developed taking formic acid as a carbonyl source for the synthesis of 2-carboxylic acids of sugars by the hydroxycarbonylation strategy.

Sugar acids constitute a diverse family of carbohydrates1 which play a crucial role in cell–cell recognition, cellular adhesion, and virus–host recognition processes, for protection of cells from pathogen attachment, and in the synthesis of biologically active natural products.2 α,β-Unsaturated sugar acids such as zanamivir and ianinamivir (Fig. 1) are subjects of particular interest because of their application as inhibitors of different glycoproteins such as hemagglutinin (HA) and neuraminidase (NA),3 the major glycoproteins expressed by influenza viruses. While several reports dealing with the synthesis of carboxylic acids at C-6 and C-1 positions of sugars exist,4a,b there is no established procedure for the synthesis of C-2 carboxylic acids. In glycals accessing carboxylic group at C-1 position required t-butyl lithium and carbon dioxide treatment at −78 °C.4c There is only one report available in the literature in which carboxylic group was introduced to the C-2 position of glycals by Furstner et al.4c,d where the C-2 carboxylic acid was derived from the Pinnick oxidation of 2-formyl glycal obtained by classical Vilsmeier–Haack reaction5 and thereafter utilized in the total synthesis of bioactive natural orevactaene (exhibits HIV-1 inhibitory property). This strategy has certain drawbacks like long reaction times with cocktails of oxidants and limited substrate specificity. For example, it works only with ether protected sugars like tri-O-benzyl-d-glycal and fails with other base labile and silyl protecting groups. Further, recovery of 2-formyl glycals after base workup is rather low in our hand. Our experience with glycals6af encouraged us to formulate an attractive way to launch carboxylic acid at C-2 position of glycals as shown in Scheme 1 and apply them in the synthesis of C-2 glycoconjugates.Open in a separate windowFig. 1Glycal based acids in drugs.Open in a separate windowScheme 1Art of launching carboxyl group in sugars.This is pertinent to mention that carbonylation reactions of C-2 glycals have been successfully carried out by using metal carbonyl for the synthesis of C-2 branched glycoconjugates.7a,b In these reaction stoichiometric amount of costly Mo(CO)6 is required for such transformation that too ends up with some non-carbonylative side products. CO surrogates8,9 such as formic acid, formamide, chloroform and anhydride have been explored in recent times obviating metal carbonyls and CO gas. Among all formic acid is an attractive candidate for insertion of CO in an organic molecule,10 because it liberates one water molecule after releasing one CO molecule thereby making the process environmentally benign. We felt that palladium catalyzed hydroxycarbonylation of stable glycal halides, which are conveniently accessed from glycals in good yield, may prove to be the most effective and environmentally benign method to prepare such molecules. With our continuous interest in synthesis of C-2 branched sugars,11 this time we developed a reagent system for the direct synthesis of C-2 sugar carboxylic acids from 2-iodo glycals using formic acid as carbonyl source. Further, we transformed the synthesized acid for the synthesis of different C-2 branched glycoconjugates.Preliminary experiments were conducted by using 2-iodoglycal 1a, HCOOH as carbonyl source, N,N′-dicyclohexylcarbodiimide, (DCC) as an activator and xantphos as a ligand. When 1a was reacted with 5 mol% Pd(OAc)2, 10 mol% of xantphos, 1 equiv. of DCC, 2 equiv. of formic acid and 2 equiv. of triethyl amine as base in DMF at 90 °C for 16 h the desired product 3a was obtained along with 3a′ in 60 : 40 ratio with overall 63% yield ( EntryPd sourceLigandTime (h)Conversion (%)(3a : 3a′)Yieldb (overall)%1Pd(OAc)2L1169060 : 40632Pd(OAc)2L21699>99 : 1723Pd(OAc)2L2699>99 : 180 4 Pd(OAc) 2 L2 1.5 99 >99 : 1 81 5Pd(OAc)2L2190>99 : 1736Pd(OAc)2L3164045 : 55237Pd(OAc)2L4163030 : 70158Pd(OAc)2L51610—Traces9Pd(OAc)2L61610—Traces10Pd(PPh)3L2275>99 : 12111Pd(TFA)2L2235>99 : 12712PdCl2L2223>99 : 111Open in a separate windowaReaction conditions: 1a (0.18 mmol), 2a (0.36 mmol), Pd(OAc)2 (0.009 mmol), L2 (0.018 mmol), N,N′-dicyclohexylcarbodiimide (DCC) (0.18 mmol), triethylamine (0.36 mmol) at 90 °C for 2 h.bYield of isolated product. Pd 5 mol% and ligand 10 mol% were used. Ratio of 3a and 3a′ and conversion were determined through 1H NMR.Utilising the optimised reaction condition (l-rhamnal 1b was tested to get a better yield of the product 3b, (85%). The galactal substrate also furnished the desired 2-carboxyl galactal 3c in good yield (76%). In order to broaden the substrate scope the reactivity of glycals protected with different protecting groups was next investigated. Gratifyingly, 2, 3-acetonide protected 2-iodo-d-galactal 1d survived under the reaction condition and yielded the product 3d in good yield (75%). Tri-O-ethyl-2-iodoglucal 1e also reacted well and formed the respective acid derivative 3e in (76%) yield. Next we utilized different glycals having silicon based protection or ester protection. Tri-O-acetyl-2-iodoglucal 1f and di-O-acetyl-2-iodoxylal 1g were also well tolerated under the reaction condition and gave the desired products 3f–3g in reasonable yields (73–75%). 2-Bromo-glucal was next tested under the optimized reaction condition and the desired compound 3a was obtained in good yield although it takes 3 h to complete the reaction.Substrate scopea
EntrySubstrateProductTime (h)Yieldb (%)
1 1.581
2 1.585
3 1.576
4 1.575
5 276
6 2.573
7 2.575
Open in a separate windowaReaction conditions: 1 (1 equiv.), 2a (2 equiv.), Pd(OAc)2 (5 mol%), L2 (10 mol%), DCC (1 equiv.), triethylamine (2 equiv.) at 90 °C for 1.5 to 2.5 h in 3 mL of DMF.bYield of isolated product.TBS protected sugars acids are used in the total synthesis of various bioactive natural products by activating the anomeric carbon which is otherwise not possible with other protecting groups. In the literature in order to get such types of acids multiple steps are required along with poor overall yield. By utilizing carbonylation strategy under Pd catalysis we were able to synthesis the TBS protected acids in just 2 h when substrate 1h was reacted with HCOOH under Pd catalysis to generate product 3h in good yield.After successful execution of our strategy for hydroxycarbonylation of 2-iodoglycals, we became interested to apply the same carbonylative approach for the synthesis of 2-formyl glycals. In the literature formylation at C-2 of glycals has been carried out either by Vilsmeier–Haack or XtalFluor-E catalyzed reactions (Scheme 3).5,12 We utilized the reductive carbonylation approach employing triethylsilyl hydride (1.2 equiv.) as a hydride source keeping other reagents same as for acid synthesis. To our delight, the required 2-formyl glycals were obtained in good to excellent yields (Scheme 3, 3aa–3ae).Open in a separate windowScheme 3Synthesis of 2-formyl glycals.Open in a separate windowScheme 2Synthesis of silyl protected C-2 carboxylic acid.To test the utility of sugar acids in the synthesis of C-2 glycoconjugates, sugar acid 3a was successfully utilized for the synthesis of C-2 linked dipeptides via Ugi reaction (Scheme 4, 5a). The simple amide 5b was also synthesised from 3a using thionyl chloride and ammonium hydroxide; the product is otherwise difficult to synthesize via the existing methods. Where substituted amides are synthesized. The acid chloride of 3a could be successfully coupled with 8-aminoquinoline, leading to the amide 5c in good yield. In order to test the reactivity of the acid 3a, we used it as a glycosyl acceptor in glycosylation reaction by treating with a suitable glycosyl donor like 6, when the pseudodisaccharide 7 was isolated in good yield (63%) with a mixture of anomers. On the other hand, treatment with aryne precursor 8 resulted in the formation of compound 9 in excellent yield (84%) via coupling of aryne with the sugar acid. When sugar acid 3a was treated with thiophenol thioester 11 was isolated in excellent yield (Scheme 5).Open in a separate windowScheme 4Synthesis of amides using 3a. Reaction conditions: (a) 3a (1.5 equiv.), aniline (1 equiv.), anisaldehyde (1 equiv.), cyclohexyl isocyanate (1 equiv.) in ethanol at rt for 48 h; (b) 3a (1.0 equiv.), SOCl2 (1.5 equiv.), NH4OH (37%, 2 mL) in THF for 2 h; (c) 3a (1.0 equiv.), PCl5 (1.2 equiv.), pyridine (6 equiv.), 8-aminoquinoline (1.2 equiv.) in DCM for 5 h.Open in a separate windowScheme 5Utilization of 3a for establishing ester linkages.A plausible reaction mechanism has been proposed for hydroxycarbonylation of 2-iodoglycals (Scheme 6). Initially Pd(0) is generated in situ in the presence of ligand Ln. The catalytic cycle then starts with oxidative addition of Pd(0) to 2-iodoglycal A which produces the pallado complex B.Open in a separate windowScheme 6Plausible mechanism of hydroxycarbonylation reaction.Coordination and insertion of carbon monoxide generated in situ by the combination of DCC and formic acid leads to the formation of acyl Pd(ii) complex C. Formic acid attack on complex Cvia transmetallation affords the intermediate D with release of HI. Pd(0) could be regenerated for the next catalytic cycle after reductive elimination from complex D with formation of anhydride E. Decomposition of the anhydride with release of one molecule of CO generates the desired sugar acid F.In conclusion we have developed an efficient and mild Pd catalysed synthetic strategy for hydroxycarbonylation of 2-iodoglycals using the cheap reagent formic acid as CO source and 1 equiv. of DCC as an activator. The methodology was successfully extended to various glycals with different protecting groups like acetonide, ether, ester and silicon based ones. 2-Formyl glycals were also synthesised by using reductive carbonylation approach. The synthesised sugar acid could be used in the synthesis of glycoconjugates, pseudodisaccharides, for aryl ester and thioester.  相似文献   

10.
DBU mediated one-pot synthesis of triazolo triazines via Dimroth type rearrangement     
Ab Majeed Ganai  Tabasum Khan Pathan  Nisar Sayyad  Babita Kushwaha  Narva Deshwar Kushwaha  Andreas G. Tzakos  Rajshekhar Karpoormath 《RSC advances》2022,12(4):2102
Herein we report an efficient one-pot synthesis of [1,2,4]triazolo[1,5 a][1,3,5]triazines from commercially available substituted aryl/heteroaryl aldehydes and substituted 2-hydrazinyl-1,3,5-triazines via N-bromosuccinimide (NBS) mediated oxidative C–N bond formation. Isomerisation of [1,2,4]triazolo[4,3-a][1,3,5]triazines to [1,2,4]triazolo[1,5-a][1,3,5]triazines is driven by 1,8-diazabicyclo[5.4.0]undec-7-ene (DBU) affording both isomers with good to excellent yields (70–96%).

We demonstrate a simple yet efficient one-pot synthesis of two triazolotriazine isomers via DBU mediated Dimroth type rearrangement with excellent yields.

Purines are nitrogen-containing heterocycles and are structural motifs in the nucleobases adenine and guanine of DNA as well as RNA. Purine nucleotides (ATP, GTP, cAMP, cGMP, NAD, FAD) also act as co-factors, substrates, or mediators in the functioning of numerous proteins.1 Therefore, bioisosteres of purines are widely explored and exploited by pharmaceutical chemists in developing new drug entities. Heterocycles containig the 1,3,5-triazine ring act as bioisosteres of purine, which exist as two isomers, namely [1,2,4]triazolo[4,3-a][1,3,5]triazine and [1,2,4]triazolo[1,5-a][1,3,5]triazine (Fig. 1), which have been extensively studied as adenosine receptor antagonists,2,3 as well as for other pharmacological activities1,4,5 (Fig. 2). Literature reports suggest that [1,2,4]triazolo[1,5-a][1,3,5]triazine has been exploited extensively in drug discovery as compared to its corresponding isomer. Further, from the literature it is evident that symmetrical disubstituted triazines,6–9 especially morpholine10–12 substituted, have displayed broad pharmacological activities.Open in a separate windowFig. 1Structure of isomers of triazolo–triazine.Open in a separate windowFig. 2Biologically active molecules of [1,2,4]triazolo[4,3-a][1,3,5]triazine (1) and [1,2,4]triazolo[1,5-a][1,3,5]triazine (2, 3, 4).Various synthetic methods have been reported for the C–N bond formation by employing different starting materials13–15via oxidative cyclisation,16 high temperature condition17–19 and/or metal-catalysed reactions.20 However, the current reported protocols were environmentally unfriendly as they suffered from drawbacks such as, multistep, and tedious procedures, use of carcinogenic solvents, high-temperature, expensive and toxic metal-catalysts, and other hazardous reagents.Furthermore, there is very less reported research on these heterocycles, and that can be because of the unavailability of the efficient and cheaper methods. Thus, there is a need to develop new versatile synthetic method for the synthesis of disubstituted triazolotriazine heterocycles of pharmacological interest.In 1970, for the first time Kobe21et al. (scheme a) reported the synthesis of [1,2,4]triazolo[4,3-a][1,3,5]triazine utilizing lead tetraacetate in benzene under reflux conditions (Fig. 3). However, no further Isomerization was carried out and resulted low to moderate yield. Deshpande22et al. reported (scheme b) the reaction of 2-hydrazinyl-1,3,5-triazine with various substituted benzoic acids. The product formed was treated with P2O5, refluxed in xylene for 10h to yield [1,2,4]triazolo[4,3-a][1,3,5]triazine. Further, Isomerization of the resulting product was carried out in 2% methanolic-NaOH solution resulting in poor yields. Recently, Stefano23et al. (scheme c) reported, a multistep protocol by reacting the intermediate with bis(methyl-sulfanyl)methylenecyanamide at 180 °C under N2 for 3h resulting in low yield due to the formation of several side products. In addition no isomerization studies were carried out, and only the [1,2,4]triazolo[1,5-a][1,3,5]triazine analogs were reported.Open in a separate windowFig. 3Different approaches for the synthesis of triazolo triazines.Herein, we report an economical one-pot synthesis of [1,2,4]triazolo[1,5-a][1,3,5]triazine analogs via Dimroth type rearrangement of [1,2,4]triazolo[4,3-a][1,3,5]triazine derivatives. This one pot novel methodology was carried out by reacting readily available, inexpensive starting materials such as substituted aryl/heteroaryl benzaldehydes and substituted 2-hydrazinyl-1,3,5-triazine in methanol (a mild solvent)24 at room temperature giving excellent yields of the desired product. For cyclization reaction, an eco-friendly reagent NBS25 was used and the resulting product was treated with DBU to yield its corresponding desired isomer. To the best of our knowledge this is the first report for the greener synthesis of symmetric disubstituted triazolotriazine heterocycles via Dimroth type rearrangement. We believe that this simple, yet novel methodology could be further exploited by the researchers in pharmaceutical industries and academics settings in drug discovery.Formation of the Schiff base was initiated reacting 4,4′-(6-hydrazinyl-1,3,5-triazine-2,4-diyl)dimorpholine 1a and unsubstituted benzaldehyde as shown in Entry no.SolventOxidantBaseBase equiv.Time (hour)Yield%1EtOHNBSDBU1.016682DCMNBSDBU1.01603DMFNBSDBU1.01604MeOHNBSDBU1.016725i-PrOHNBSDBU1.016Trace6H2ONBSDBU1.01607MeOHNBS——1608MeOHNBS2%NaOH1.016Trace9MeOHNBSK2CO31.016010MeOHNBSTEA1.016011MeOHNBSDABCO1.016012MeOHNBSDBU1.528513MeOHNBSDBU2.01.58014MeOHNCSDBU1.525615MeOHNISDBU1.526116MeOHIBDDBU1.526317MeOHKI/I2DBU1.525318MeOH/H2ONBSDBU1.5270Open in a separate windowaConditions: 1a (1 mmol), benzaldehyde (1 mmol), solvent, rt, 20 min, then oxidant (1 mmol), 5 min, rt, then DBU, stir at rt till completion of the reaction.Meanwhile, equivalents of DBU were adjusted (entry 12, 13) to attain the highest yield %. Reaction with 1.5 eq. showed drastic improvement in yields in 2 h whereas 2.0 eq. resulted in good yields with less reaction time. Considering the yield factor (entry 12), the oxidant optimization was achieved (entry 14–17). Using N-chlorosuccinimide (NCS) and N-iodosuccinimide (NIS) (entry 14, 15) offered 56% and 61% yield respectively. When the reaction was performed using phenyliodine(iii) diacetate (PIDA) and KI/I2 (entry 16, 17) it provided 2a in 63% and 53% yields, respectively. Finally, methanol–water and ethanol–water systems in 3 : 1 were used, which gave approximately 70% yield, concluding that entry 12 gives the best result, demanding alcoholic solvents especially methanol as a key factor for Dimroth type rearrangement. Additionally, it was supported by the observation that reaction goes well in methanol, a little worse in ethanol (Fig. 4. The reaction involves a Schiff base formation (II) by 1a and benzaldehydes, the addition of NBS results in oxidative cyclization reaction to produce isomer 1. The addition of DBU in isomer-1 initiates a famous Dimroth type rearrangement, protonation of III results in ring-opening with the formation of unstable intermediate V. It undergoes tautomerism by 1,3 proton shift, bond rotation and proton abstraction by methoxide ion facilitating the intramolecular cyclization to afford the isomer 2. This reaction mechanism is supported by the formation of the Schiff base, isomer-1, and its conversion to isomer-2, which were easily monitored by TLC, isolated, and characterized (ESI). In addition to that, a single crystal of compound 2e was obtained, which further supports the Dimorth type of rearrangement (ESI).Open in a separate windowFig. 4Plausible mechanism of the reaction pathway.Having in hand the optimized conditions, the substrate scope was further explored by using different aldehydes (Fig. 5). The reaction was carried out using 1a with benzaldehydes having electron-donating groups (2a-f) followed by NBS and DBU additions. The reaction was allowed to stir at room temperature till reaction completion, as monitored by TLC, which would approximately take 2 h. The reaction could smoothly give final products in excellent reaction yields (79 to 96%). Electron withdrawing groups such as chlorobenzaldehydes (2g-i) despite the position of substitution gave excellent yields (93–95%). With the 4-bromo and different fluoro-benzaldehydes, the reaction resulted in 2j (89%) and 2k-m (84–90%) with very good yields. Also, reaction afforded 70% and 77% yields with heterocyclic aldehydes such as 3-pyridinecarboxyaldehyde (2n) and thiophene-2-carbaldehyde (2o), respectively.Open in a separate windowFig. 5Substrate scope for different aldehydes.The gram scale reaction was performed to further validate this synthetic procedure. The reaction was carried out using 1a (1 g) with benzaldehyde (0.38 g) and stirred for 30 min. The NBS (1.26 g, 1 eq.) was added slowly and stirred till new spot appeared on TLC followed by the slow addition of DBU (1.5 eq.) resulted in isomerisation to give 2a in good yields (1.1 g, 84%) (ESI).Encouraged from the potency of the reaction to generate in good to excellent yields [1,2,4]triazolo[1,5-a][1,3,5]triazine analogues, we shifted our investigation to isolate [1,2,4]triazolo[4,3-a][1,3,5]triazine derivatives (isomer-1) as described in Fig. 6. Different aldehydes were reacted with hydrazinyl-1,3,5-triazine analogs (1a, 3a, and 4a) and followed by the addition of NBS, stirred at room temperature till the completion of reaction as monitored by TLC. Compound 1a on reaction with 4-bromobenzaldehyde and NBS resulted in 1b with 90% yield. Similarly, 3a with 4-bromobenzaldehydes and thiophene-2-carbaldehyde afforded 3b and 3c with 84% and 83% yields, respectively. The same reaction was carried out with 4a and 4-bromobenzaldehyde bearing an electron-withdrawing group gave 3d with 87% yield. Further, 4a with aldehydes bearing electron-donating groups resulted 4c and 4d, with 88% and 84% yields, respectively (Fig. 7).Open in a separate windowFig. 6Substrate scope for [1,2,4]triazolo[4,3-a][1,3,5]triazine analogues.Open in a separate windowFig. 7Substrate scope for different disubstituted triazinyl-hydrazine and aldehydes.With these promising results, we explored the scope of aldehydes and 2-hydrazinyl-1,3,5-triazine analogs for the synthesis of [1,2,4]triazolo[1,5-a][1,3,5]triazine derivatives. Compound 3a, when reacted with different aldehydes and employing the optimized protocols gave 5a-e with excellent yields (86–92%) in 2–4 h. Similarly compounds 6a, 6b and 6c, synthesized from 4a also afforded excellent yields 88%, 86% and 90% respectively. Further, substituted cinnamaldehydes were also explored and reacted with 1a followed by the addition of NBS and DBU resulted in 7a and 7b in appreciable yields of 84% and 82%, respectively.Finally, we investigated both the isomers (3b and 5d) spectroscopically, to elucidate the change in proton chemical shifts during rearrangement. Interestingly, it was observed that the proton chemical shift for 5d at the 7th position (Fig. 6) was not affected. However, 5th position of 5d suffers a downfield shift due to change in its electronic environment as compared to its corresponding isomer 3b. In general, the proton chemical shifts for the substitution at 5th position and for the aromatic region was found to be more for isomer-2 as compared to isomer-1.  相似文献   

11.
Ligand-free Pd/Ag-mediated dehydrogenative alkynylation of imidazole derivatives     
Fabio Bellina  Matteo Biagetti  Sara Guariento  Marco Lessi  Mattia Fausti  Paolo Ronchi  Elisabetta Rosadoni 《RSC advances》2021,11(41):25504
A variety of 2-alkynyl(benzo)imidazoles have been synthesized by dehydrogenative alkynylation of (benzo)imidazoles with terminal alkyne in NMP under air in the presence of Ag2CO3 as the oxidant and Pd(OAc)2 as the catalyst precursor. The data obtained in this study support a reaction mechanism involving a non-concerted metalation deprotonation (n-CMD) pathway.

The regioselective synthesis of 2-alkynyl(benz)imidazoles was successfully achieved by Pd(ii)/Ag(i)-mediated dehydrogenative alkynylation of the corresponding (benz)imidazoles with terminal alkynes in an open vessel.

The development of synthetic protocols that enable the direct and selective functionalization of Csp2–H bonds are of primary importance in the decoration of the imidazole scaffold.1–10 The presence of two nitrogen atoms in the pentatomic structure of this important azole in fact introduces a differentiation in the chemical behaviour of the three C–H bonds, allowing their selective involvement in carbon–carbon bond-forming reactions as long as appropriate experimental conditions are identified.During our studies dedicated to the synthesis and investigation of azole-based fluorophores featuring a π-conjugated backbone end-capped with electron-donating (EDG) and electron-acceptor (EWG) groups (the so-called “push–pull” systems),11–15 taking into account the structural characteristics and synthetic versatility of a triple carbon–carbon bond16–23 that make it an excellent π-spacer, we recently decided to evaluate the possibility of obtaining alkynyl-substituted imidazoles by transition metal-catalyzed Csp2–Csp bond-forming reactions. Among the synthetic procedures that allow the selective alkynylation of imidazoles and other five-membered heteroarenes,5,24 there is no doubt that the possibility of forming the new σ carbon–carbon bond through the double activation of two distinct carbon–hydrogen bonds through a cross-dehydrogenative alkynylation (CDA) represents the best synthetic approach in terms of atom economy and functional group tolerance (eqn (1), Scheme 1).25–34 In fact, when compared with the transition metal-catalyzed direct Csp2–H alkynylation protocols involving 1-haloalkynes (the so-called “inverse Sonogashira coupling”) (eqn (2), Scheme 1),35–45 hypervalent iodine reagents (eqn (3), Scheme 1),46–48 or α,β-ynoic acids (decarboxylative direct arylation) (eqn (4), Scheme 1),26,49,50 the CDA methodology makes it possible to use terminal alkynes without the need for preliminary activation.5,24Open in a separate windowScheme 1Synthetic protocols for the transition metal-catalyzed Csp2–H alkynylation of heteroarenes.Despite several papers having been dedicated to the dehydrogenative alkynylation of pyrroles,27 indoles,27,31 oxazoles,27,30,32–34 benzoxazoles,26,29,30,32,34 thiazoles,27,30 benzothiazoles,26,29,31,32,34 pyrazoles,25,27 1,3,4-oxadiazoles,33 imidazopyridines,27,34 and N-substituted sydnones,28 to the best of our knowledge a study specifically devoted to the dehydrogenative alkynylation of imidazoles has never been reported. However, careful reading the literature where alkynylation reactions involving azoles other than imidazole were described, it emerged that two papers reported the C-2 alkynylation of N-benzylimidazole and N-benzylbenzimidazole with phenylacetylene. In a study mainly devoted to the dehydrogenative alkynylation of imidazo[1,5-a]pyridine, Shihabara, Dohke and Murai employed the commercially unavailable Pd(ii) complex bi(4-nitropyridinyl)Pd(OAc)2, Ag2CO3 as the oxidant and AcOH as additive, in a 95 : 5 mixture of DMF and DMSO for 2 h at 120 °C under argon.34 Due to the propensity of alkynes to give Glaser-type homocoupling side-products in the presence of silver(i) salts,51 phenylacetylene was slowly added over 90 min to the reaction mixture. In 2015, Likhar, Kantam and co-workers used a synthetic Pd(ii) carbene complex. In this case Ag2O was used as the oxidant, Cs2CO3 base, performing the coupling in DMF at 85 °C under air.26Encouraged by these results but with the intention of developing a simplified procedure that would allow the use of a commercial Pd(ii) pre-catalyst in the absence of any palladium ligands, we decided to review the conditions of reaction and, in this work, we are pleased to summarize the results obtained in the synthesis of 2-alkynylimidazole derivatives by dehydrogenative alkynylation with terminal alkynes (Scheme 2). In particular, a careful screening allowed us to show how the reactivity of Csp2–H bond of imidazole derivatives and other azoles can be enhanced simply by performing the alkynylation using NMP as the reaction solvent, under air in non-anhydrous conditions, and without the need for palladium ligands.Open in a separate windowScheme 2Pd/Ag-Promoted dehydrogenative alkynylation of imidazole derivatives.We decided to start our synthetic investigations by trying a cross-dehydrogenative coupling between N-methylbenzimidazole (1) and phenylacetylene (2a), chosen as typical coupling partners, under conditions very similar to those proposed by Murai and co-workers for the regioselective C-3 dehydrogenative alkynylation of 1-alkynyl-3-arylimidazo[1,5-a]-pyridines.34Hence, to a mixture of 1.0 mmol 1, 5 mol% Pd(OAc)2, 1.5 equiv. Ag2CO3, and 1.0 equiv. AcOH in a 95 : 5 (v/v) mixture of DMF and DMSO under argon 3.0 equiv. of 2a were added dropwise over 90 min. However, after 2 h at 120 °C a 56% GLC conversion of 1 was recorded, and the required 2-phenylethynyl-N-methylbenzimidazole (3a) was obtained in only 34% GLC yield (entry 1, EntryPd cat. (mol%)Ag(i) salt (equiv.)RCOOHSolventTemp./react. timeb (°C/h)1 GLC conversionc (%)3a GLC yieldd (%)3a:4 AP%1ePd(OAc)2 (5)Ag2CO3 (1.5)AcOHDMF/DMSO (95 : 5)120/2563457 : 432Pd(OAc)2 (5)Ag2CO3 (1.5)AcOHDMF/DMSO (95 : 5)100/5.5676632 : 683Pd(OAc)2 (5)Ag2CO3 (2.0)AcOHDMF/DMSO (95 : 5)100/2737034 : 664Pd(OAc)2 (5)—AcOHDMF/DMSO (95 : 5)100/3.5NR——5Pd2(dba)3 (2.5)Ag2CO3 (2.0)AcOHDMF/DMSO (95 : 5)100/3.5736930 : 706Pd(acac)2 (5)Ag2CO3 (2.0)AcOHDMF/DMSO (95 : 5)100/3.5676327 : 737PdCl2 (5)Ag2CO3 (2.0)AcOHDMF/DMSO (95 : 5)100/3.5726835 : 658Pd(OAc)2 (5)Ag2CO3 (2.0)—DMF/DMSO (95 : 5)100/3.5532372 : 289Pd(OAc)2 (5)Ag2CO3 (2.0)—AcOH100/3.50—0 : 10010Pd(OAc)2 (2.5)Ag2CO3 (2.0)AcOHDMF/DMSO (95 : 5)100/3.5766930 : 7011Pd(OAc)2 (2.5)AgOAc (2.0)AcOHDMF/DMSO (95 : 5)100/3.5554825 : 7512Pd(OAc)2 (2.5)Ag2O (2.0)AcOHDMF/DMSO (95 : 5)100/3.5666228 : 7213Pd(OAc)2 (2.5)Ag2CO3 (2.0)EtCOOHDMF/DMSO (95 : 5)100/3.5807135 : 6514Pd(OAc)2 (2.5)Ag2CO3 (2.0)PivOHDMF/DMSO (95 : 5)100/3.5454522 : 7815Pd(OAc)2 (2.5)Ag2CO3 (2.0)AcOHDMF/DMSO (95 : 5)80/3.5434017 : 8316Pd(OAc)2 (2.5)Ag2CO3 (2.0)AcOHDMF100/3.5757034 : 6617Pd(OAc)2 (2.5)Ag2CO3 (2.0)AcOHDMA100/3.58175(68)33 : 67 18 Pd(OAc) 2 (2.5) Ag 2 CO 3 (2.0) AcOH NMP 100/3.5 85 81(69) 36 : 64 19fPd(OAc)2 (2.5)Ag2CO3 (2.0)AcOHNMP100/3.58266(51)40 : 6020gPd(OAc)2 (2.5)Ag2CO3 (2.0)AcOHNMP100/3.5838036 : 64Open in a separate windowaReaction conditions: 2a (3.0 mmol) in the selected solvent (2.0 mL) was added dropwise over 45 min to a mixture of 1 (1.0 mmol), Pd cat., Ag(i) salt, RCOOH (1.0 mmol) in the selected solvent (2.0 mL) under air, unless otherwise reported.bAfter the reported reaction time the GLC conversion of 2a was quantitative.cGLC conversion of 1vs. biphenyl.dGLC yield vs. biphenyl. In parentheses isolated yield.eThis reaction was carried out under an argon atmosphere.fThis reaction was carried out on a 10 mmol scale.gThis reaction was carried out in the presence of TEMPO (1.0 equiv.) as radical scavenger.Taking note of this unexpected negative result, we decided to re-examine the reaction conditions, starting from the consideration that even if Ag2CO3 can theoretically serve as the oxidant, run the reaction in air (open flask) may be critical.28 Gratifyingly, when the coupling was performed under air there was a marked increase in GLC yield, which went from 34% to 66% (entry 2, i) salts different from Ag2CO3 gave slightly worse results (entries 11 and 12, 31 the use of the right buffer system may be crucial for the deprotonation of the terminal alkyne, and also for the reoxidation of Pd(0) to Pd(ii).With our delight the chemical yield went up to 69% when the DMF/DMSO mixture was replaced by NMP (entry 18, 34 it should be highlighted that the slow addition of phenylacetylene never avoided in our hands the formation of Glaser side-product 4.Finally, when the reaction was performed in the presence of 1.0 equiv. of TEMPO (2,2,6,6-tetramethylpiperidin-1-yl)oxyl as radical scavenger, it still proceeded smoothly to afford 2-alkynylazole 3a in 80% GLC yield (entry 20, Scheme 3).Open in a separate windowScheme 3Ligandless Pd/Ag-promoted dehydrogenative alkynylation of 1-methylimidazole 5 or 1-methylbenzimidazole 1 with terminal alkynes 2a–h.Noteworthy, the reaction outcome was not influenced by the electronic nature of the aromatic moiety on the terminal alkyne. More importantly, several functional groups on the phenyl ring resulted well tolerated, including formyl and hydroxy groups that are potentially sensitive to the oxidative conditions required by dehydrogenative couplings.While good results were obtained when arylacetylenes were used as coupling partners, worse results were observed when TIPS-acetylene 2i or 1-octyne (2j) were reacted with 5. In fact, the expected 2-alkynyl substituted imidazoles 6i and 6j were isolated in 30% and 11% yields, respectively (Scheme 4).25,31 The low reactivity of these two terminal alkynes in respect to their aryl analogues could be attributed, as already noted,31 to the decreased electrophilic nature of the corresponding alkynyl-palladium(ii) intermediates being generated during the reaction. As a further proof of their lower reactivity, the slow addition of 2i and 2j was found to be unnecessary.Open in a separate windowScheme 4Ligandless Pd/Ag-promoted dehydrogenative alkynylation of 1-methylimidazole 5 with terminal alkynes 2i and 2j.Gratifyingly, the optimized Pd(OAc)2/Ag2CO3-promoted coupling conditions proved to be useful also for the C-2 dehydrogenative alkynylation of 4,5-diphenyl-1-methylimidazole 7. In fact, this disubstituted azole gave the expected C-2 alkynylation product 8 in 70% isolated yield when reacted with phenylacetylene (2a) (Scheme 5).Open in a separate windowScheme 5Ligandless Pd/Ag-promoted dehydrogenative alkynylation of 4,5-disubstituted 1-methylimidazole 7 or 9 with terminal alkynes 2a or 2d.One of the main advantages of dehydrogenative cross-coupling reactions is the tolerance, among others, of carbon–halogen bonds that, hence, are available for subsequent transformations. We were please to confirm this tolerance; in fact, 4,5-dibromo-1-methyl-1H-imidazole 9 was efficiently reacted with alkynes 2a and 2d, giving rise to the required 2-alkynyl-4,5-dibromoimidazoles 10a,b in 68 and 75% isolated yields, respectively (Scheme 5). Notably, no conventional Sonogashira-type coupling side-products were observed.Similar satisfactory yields were obtained when 1-benzyl- and 1-phenyl-1H-imidazoles 11 and 12 when reacted with 2a. The corresponding 2-phenylethynylimidazoles 13 and 14 were recovered in 68% and 52% isolated yields, respectively (Scheme 6). In contrast, the C-2 alkynylation of thiazole (15) and oxazole (16) with 2a gave the required products 17 and 18 in somewhat lower yields (48 and 42%, respectively). It is noteworthy, however, that the observed reactivity of 1-substituted imidazole 6, 11 and 12, i.e. 1-methyl > 1-benzyl > 1-phenyl, and that found for the three 1,3-azoles 6, 13 and 14, i.e. 1-methylimidazole > thiazole > oxazole, parallels that reported in the literature for classical SEAr.52,53 As regards the regioselectivity, when the reaction involved 1-substituted imidazoles 5, 11, and 12, thiazole 15 and oxazole 16 a clean C-2 alkynylation was observed, while the corresponding C-5 or even the less probable C-4 alkynyl-substituted regioisomers were never detected in the crude reaction mixtures.54 However, it is worth mentioning that when the alkynylation was tempted on 1,2-dimethyl-1H-imidazole 19, the C-5 alkynylated product 20 was obtained in 42% isolated yield (eqn (1)).1Open in a separate windowScheme 6Ligandless Pd/Ag-promoted dehydrogenative alkynylation of 1-substituted imidazoles 5, 11, 12, thiazole (15), and oxazole (16) with phenylacetylene (2a).Based on earlier reports and our observations, a possible reaction mechanism for this Pd/Ag-promoted dehydrogenative alkynylation of imidazoles is summarized in Fig. 1, using 1-methylimidazole 5 as an example. Initial transmetallation involving a Pd(ii) complex and a (presumed) silver(i) acetylide generates the alkynyl Pd(ii) species A, which then affords the imidazole–Pd–alkyne complex E through a sequence of base-assisted carbanion generation from an azole–Pd complex (B → C), followed by a C-palladation step via carbene D. This particular C–H palladation mechanism, known as non-concerted metalation-deprotonation (n-CMD) pathway, was initially proposed by Hoarau and co-workers to justify the observed C-2 regioselectivity in the copper-free Pd-catalyzed direct arylation of oxa(thia)zoles-4-carboxylate with aryl bromides.55–57Open in a separate windowFig. 1Proposed n-CMD mechanistic pathway for the Pd(OAc)2/Ag2CO3-promoted dehydrogenative alkynylation of 1-methylimidazole 5.Although the comparison of the reactivity of 1-methylimidazole 5 with both the other two 1-substituted imidazoles 11 and 12, and with thiazole (15) and oxazole (16) suggested a SEAr mechanism (hence via a classic Wheland intermediate),58–60 it should be noted that the most reactive position should have been C-5, and not the position C-2 that instead has the most acidic C–H bond. Similarly, a pure concerted metalation–deprotonation (CMD) pathway can reasonably be excluded considering that DFT calculations have shown that in the case of 1,3-azoles position C-5 is, again, the most reactive.61 The n-CMD mechanism hypothesizes that the deprotonation occurs through the formation of an azole–Pd(ii) complex and, therefore, justifies the observed reactivity by the most acidic C–H bond, i.e. the one in position 2 on the azole nucleus. On the other hand, one key step of this mechanism is the formation of N3–Pd complex C, which is the more effective the more the terminal alkyne is electron-poor (hence having the Csp–H bond more acid). That the steps of the mechanism from A to E are certainly critical for the success of the reaction is also proved by the beneficial effect resulting from the slow addition of the terminal alkyne 2 to the reaction mixture. Finally, reductive elimination involving complex E gave the coupling product 6, and reoxidation of Pd(0) to Pd(ii) by Ag(i) or air closed the catalytic cycle.  相似文献   

12.
Alkyl coupling in tertiary amines as analog of Guerbet condensation reaction     
Yage Zhou  Dan Wu  Willinton Yesid Hernndez  Changru Ma  Huangyang Su  Vitaly Ordomsky 《RSC advances》2019,9(17):9845
We report here that C–C coupling in tertiary amines for the synthesis of long chain and hindered amines might be efficiently performed over Pt and Pd catalysts. The mechanism study confirms similarity with the Guerbet reaction through dehydrogenation of the alkyl group and subsequent attack of the α-carbon atom by an alkyl group of another molecule. Finally, secondary amines and tertiary amines with longer alkyl chains are formed.

C–C coupling in tertiary amines for the synthesis of fatty and hindered amines proceeds efficiently over supported Pt and Pd catalysts.

Tertiary amines are important products in the modern chemical industry. Usually applications of tertiary amines include quaternary derivatives, amine oxides and betaines which are used in household, industrial, and institutional cleaners and disinfectants, wood treatment, personal care, oil field, and water treatment end-use markets.1Different methods have been reported for the synthesis of tertiary amines, for example, (1) reductive alkylation of aldehydes with secondary amines over metal catalysts,2 (2) N-alkylation of amines or nitroarenes with alkylhalides or alcohols,3 (3) hydroamination of olefins with amines4 and (4) amination of arylhalides.5 The main disadvantage of these routes is complex multistep processes and environmental pollution due to the usage of aggressive chemicals like halides. At the same time, these methods often provide low selectivities to desired tertiary amines.There are several classes of tertiary amines, which are especially important. The linear tertiary amines containing alkyl chain between 6 and 20 carbon atoms with two other short chain alkyl or piperidine ring are the commonly referred to fatty tertiary amines which are used for the synthesis of surfactants.6 Hindered tertiary amines containing iso-alkyl groups is another important class of tertiary amines which is hard to produce.7 These amines are used as a non-nucleophilic base and as a stabilizer for polymers.8The problem of increase of the chain length for alcohols and synthesis of isomerized alcohols have been solved by Marcel Guerbet in 1899.9 The Guerbet reaction involves coupling of two (or more) alcohol molecules through intermediate aldehyde formation, aldol condensation, dehydration of aldol product and hydrogenation of allylic aldehyde. The reaction proceed over bifunctional catalysts containing usually Cu, Ni metal for dehydrogenation/hydrogenation and base sites like MgO for C–H bond activation and coupling reaction.10Depending on the types of used alcohols in the reaction like long chain or short chain branched or unbranched products could be formed, respectively.11 This route would be an interesting opportunity for the synthesis of long chain or hindered amines. We have found that transformation of tertiary amines over Pd and Pt metallic catalyst proceed through C–C coupling with formation of heavier amines depending on the type of used tertiary amine.The transformation of non-symmetric tertiary amines over Pd black has been described in 1978 by Murahashi et al.12 The authors observed fast exchange of alkyls with formation of the mixture of tertiary amines. The reaction has been explained by insertion of palladium into a carbon–hydrogen bond adjacent to the nitrogen, leading to a highly active intermediate complex of an iminium ion. Afterwards a lot of reports have been devoted to C–C coupling of tertiary amines with different compounds like acrylates13 and alkynes14 usually activated by electro and photo-energy.15 The self-alkylation of triethylamine over noble metal catalysts in CO atmosphere has been described in the patent,16 however, no mechanism or further applications have been proposed. As far as we know, there are no scientific studies devoted to C–C coupling between tertiary amines. The reaction proceed by transfer of alkyl group from one molecule to another with formation of new C–C bond and increase of the chain length in one tertiary amine and formation of secondary amine from another molecule (Fig. 1).Open in a separate windowFig. 1Scheme of alkyl coupling of tertiary amines.Triethylamine (TEA) has been chosen as a model reagent for the reaction like ethanol in Guerbet reaction. Catalytic results of transformation of TEA over different catalysts (2 g TEA, p(N2) = 5 bar, 0.1 g catalyst, 0.24 mol% Pd, 5 h)
Catalyst T, °CTOF, h−1Conv.,%Selectivity to amines, mol%
CouplingDealkylation
123456
2001
Pd/C2006228920133261117
Pt/C20040062275344128
Pd/Al2O32001644741514292
Pd/Al2O3a20094273924211
Pd/Al2O3b20012837472483
Pd/Al2O3c2008324444491
Ru/Al2O32001434849
Ru/C200251338145522
Rh/Al2O32002
Pd/Al2O315013443532
Pd/Al2O3250881474302611
Pd black200684743
Ni/Al2O32006216514344
Open in a separate windowaH2 gas phase.bCyclohexane has been used as a solvent.cN2 pressure was 30 bar.The catalytic activity in transformation of TEA in inert atmosphere depends on the type of the metal, dispersion and support. The catalysts have been characterized by TEM and CO adsorption to determine the metal surface area (Table S1, Fig. S6–S7, ESI). TOF numbers over the same supports decrease in the row Pd > Pt > Ni > Ru ≈ Rh. It should be noted that activity is higher over carbon support in comparison with alumina. This catalytic behaviour correlates with activity of the metals in hydrogenation/dehydrogenation reactions. Indeed, Pt, Pd and Ni are well known as highly active hydrogenation catalysts in comparison with Rh and Ru which are good catalysts for hydrogenolysis reactions.17 This is why amination of aldehydes or alcohols over Pt and Pd lead to the synthesis of secondary and tertiary amines and Rh and Ru to selective synthesis of primary amines.18 It has to be noted that Pd black provides very low activity in transformation of TEA, which indicates on size, and electronic effect of Pd nanoparticles on intrinsic activity of TEA transformation. It explains mainly exchange activity of this catalyst in tertiary amines reported in the work of Murahashi.12Analysis of the dependence of the selectivity versus conversion over Pd/Al2O3 shows that selectivity to butyldiethylamine and diethylamine is close to 50% at low conversion and decreases with increase of the conversion (Fig. 2). The selectivity to the heavier products like dibutylethylamine and hexyldiethylamine is increasing with increase of the conversion. At the same time, we observe appearance of new dealkylation products like ethylamine, butylethylamine and dibutylamine. Thus, reaction takes consecutive character with transformation of initially formed butyldiethylamine further into the heavier products. The selective disproportionation to butyldiethylamine and diethylamine at reasonable conversion might be attained using organic solvent (Open in a separate windowFig. 2Selectivity to the products of TEA transformation depending on the conversion of TEA over Pd/Al2O3 (5 wt%) catalyst (T = 200 °C, 2 g TEA, p(N2) = 5 bar, 0.1 g catalyst, 0.24 mol% Pd, 1–17 h).This distribution of the products could be explained by combination of condensation and decomposition reactions of tertiary amines. However, analysis of the gas phase during transformation of TEA at the reaction conditions (200 °C) shows absence of hydrocarbons which means that reaction does not proceed through dealkylation of tertiary amines and most probably takes place by direct transfer of alkyl group from one tertiary amine to another one. Also carbon balance analysis shows that the molar amount of tertiary amines multiplied by amount of added C2 units corresponds to amount of amines with lost ethyl groups (Table S2, ESI).The key parameters for the reaction have been found to be reaction temperature and gas phase. At 150 °C there is almost no conversion of TEA (10 In order to verify that this rule works for tertiary amines we have performed reaction of transformation of tri-n-propylamine (TPA). The main product of the reaction over Pd/Al2O3 in comparison with TEA was tertiary amine with isomerized alkyl chain – 2-methyl-N,N-dipropylpentane-1-amine (). The identification and assignment of the products by GC-MS and NMR is given in the Fig. S8–S11, ESI. It indicates that propyl group attaches to the secondary β carbon atom of propyl group of another tertiary amine. In the case of TEA it leads to the synthesis of linear butyl chain.Catalytic results of transformation of TPA over Pd/Al2O3 (5 wt%) (2 g TPA, p(N2) = 5 bar, 0.1 g catalyst, 0.34 mol% Pd, 5 h)
Catalyst T, °CConv., %Selectivity to amines, mol%
78
Pd/Al2O320095444
Pd/Al2O3250696230
Open in a separate windowThe activity of TPA transformation was lower in comparison with TEA most probably due to steric restrictions in the condensation of alkyl groups. It has to be noted that stability of TPA toward decomposition at 250 °C was lower than for TEA which resulted in appearance of light hydrocarbons (propane, hexane) and increase of the contribution of dipropylamine and propylamine ().One of the main disadvantages of this route of the synthesis of tertiary amines is that at high conversion reaction does not proceed in the direction of increase of only one alkyl chain. All groups of tertiary amine participate in the condensation. The possible solution of this problem could be in transformation of tertiary amines containing only one active group for coupling reaction. The alkyl groups might be not active in the coupling reaction, for example, when the role of two alkyls is playing aliphatic ring like in the case of N-ethylpiperidine (). The identification and assignment of the products by GC-MS is given in the Fig. S12–S13, ESI.Catalytic results of transformation of N-ethylpiperidine over Pd/Al2O3 (5 wt%) (2 g N-ethylpiperidine, p(N2) = 5 bar, 0.1 g catalyst, 0.27 mol% Pd, 5 h)
Catalyst T, °CConv., %Selectivity to amines, mol%
9101112
Pd/Al2O320036542164
Open in a separate windowThe mechanism of the reaction of C–C coupling in tertiary amines has been studied by FTIR spectroscopy. Fig. 3 demonstrates results of adsorption of TEA over Pd/Al2O3 with subsequent heating. Adsorbed TEA demonstrates usual set of bands attributed to asymmetric and symmetric C–H stretching in CH2 and CH3 groups (2972, 2882 and 2940, 2827 cm−1), CH2 and CH3 bending (1460 and 1385 cm−1) and C–N stretching (1206 cm−1) vibrations. Gradual heating to 100 °C results in appearance of the band at 1580 cm−1 assigned to olefin double bond with significant decrease of the bands related to CH3 groups. Thus, the first step is dehydrogenation of TEA to enamines on the metal surface (Fig. 4). Enamine–imine tautomerization should lead to negative charge of terminal CH2 group and double C Created by potrace 1.16, written by Peter Selinger 2001-2019 N bond which explains disappearance of C–N stretching. The formation of palladium–iminium complex has been identified earlier.19,20Open in a separate windowFig. 3FTIR spectra during transformation of TEA over Pd/Al2O3.Open in a separate windowFig. 4Scheme of the mechanism of the reaction.Further heating of the sample to 200 °C leads to appearance of the new bands at 1643 and 1530 cm−1 which might be explained by formation of conjugated double bonds in smaller molecules. Enamines are well known as strong nucleophiles and have been used earlier in the reactions of alkylation, acylation etc.21 Similar to Guerbet condensation mechanism it should lead to attack of α-carbon attached to nitrogen by CH2 with formation of C–C bond (Fig. 4). The bond of N with tertiary carbon should not be stable and splits with removal of secondary amine like water in Guerbet condensation. In infrared cell at low pressure these amines will be in dehydrogenated state which leads to appearance of high frequency C Created by potrace 1.16, written by Peter Selinger 2001-2019 C group vibration. In the reactor subsequent hydrogenation of the formed product should result in the formation of butyldiethylamine and diethylamine as the main products of the reaction (Fig. 4).The formation of intermediate diamine compounds from TEA could be observed during analysis of the reaction kinetic by 13C and 1H NMR. The presence of weak signals of protons shifted down-field due to higher electronic density in –CH2– groups and signal of carbon shifted high-field due to presence of tertiary carbon indicates on the presence of diamine compounds in the reaction products (Fig. S3 and S4, ESI). The NMR analysis also shows presence of traces of olefins with 1H shift 7 and 3.6 ppm in the product of the reaction which indicates on hydrogen borrowing mechanism of the reactions (Fig. S5, ESI). GC-MS analysis also shows the presence of additional peak at high retention time, which according to MS spectrum, might be assigned to diamine compound (Fig. S1 and S2, ESI).The stability of intermediate diamine product should be higher in the case of delocalization of nitrogen electrons into the pi system of the benzene ring. In order to prove formation of intermediate diamine species, N,N-diethylaniline has been converted over Pd/Al2O3 catalyst (Catalytic results of transformation of N,N-diethylaniline over Pd/Al2O3 (5 wt%) (2 g N,N-diethylaniline, p(N2) = 5 bar, 0.1 g catalyst, 0.36 mol% Pd, 5 h)
Catalyst T, °CConv., %Selectivity to amines, mol%
131415
Pd/Al2O32005364832
Open in a separate windowUsually Guerbet reaction requires besides metal also acid or base sites for condensation of aldehydes by activation of C–H bond.10,11 In the case of tertiary amines condensation proceed even over carbon support (Fig. 5 demonstrates results of CO adsorption over parent Pd/Al2O3 catalyst. The peaks at 2090 cm−1 and 1977 cm−1 with shoulder at 1920 cm−1 are observed which might be assigned to linear and bridged bonded CO, respectively.22 The presence of TEA and products on the surface of metal leads to significant broadening of the CO peaks and shift to the lower frequencies. This result indicates on broad distribution of electronic states of Pd and higher electronic density on the metal leading to transfer from d-orbital of the metal to anti-bonding orbital of CO.23Open in a separate windowFig. 5FTIR CO adsorption with desorption in vacuum over Pd/Al2O3 before and after treatment of TEA.The imine form of adsorbed dehydrogenated TEA is highly polarized molecule and should interact strongly with Pd through double bond C Created by potrace 1.16, written by Peter Selinger 2001-2019 N and CH2 group (Fig. 5). The fact that the highest activity is observed over small size supported metal nanoparticles might be explained by strong stabilization of imine over defected Pd sites which provides high activity in the coupling reaction.To conclude, we uncovered that tertiary amines in the presence of Pd and Pt catalysts can transfer alkyl groups by C–C coupling with formation of tertiary amines with longer chains and secondary amines. The proposed mechanism based on FTIR results is similar to Guerbet condensation reaction and involves dehydrogenation with subsequent attack of α-carbon atom by carbanion with elimination of secondary amine. This method provides tool for the synthesis of long chain or hindered tertiary amines.  相似文献   

13.
Simple and efficient synthesis of bicyclic enol-carbamates: access to brabantamides and their analogues     
Ondrej Zborský   udmila Petrovi ov  Jana Doh&#x;o&#x;ov  Jn Moncol  Rbert Fischer 《RSC advances》2020,10(12):6790
A novel synthetic approach towards the formation of the unusual bicyclic enol-carbamates, as found in brabantamides A–C, is reported. The bicyclic oxazolidinone framework was obtained in very good yield and with high E/Z selectivity from a readily available β-ketoester under mild reaction conditions using Tf2O and 2-chloropyridine tandem. The major E isomer was used in the synthesis of the brabantamide A analogue.

A simple and short synthesis of bicyclic enol-carbamates with high E/Z selectivity and the synthesis of brabantamide A analogue are presented.

Brabantamides A–C (1–3) (Fig. 1) were first isolated in 2000 from the culture extracts of Pseudomonas fluorescens.1 They displayed nanomolar inhibitory activity towards lipoprotein-associated phospholipase A2 (Lp-PLA2) and therefore they could be used in the treatment of the inflammatory diseases such as atherosclerosis.1–3 It was found that the sugar moiety is not necessary for the biological activity against Lp-PLA2. In contrast, deglycosylated brabantamides A and C showed improved inhibitory activity compared to their natural counterparts.3 Moreover, simplified brabantamide analogues with amide functional group and long alkyl side chains showed even higher inhibitory effect.4 Brabantamides A–C also exhibit significant activity against Gram-positive bacteria, fungi, and oomycetes.5,6 A recent study confirmed that the bicyclic scaffold and the long lipophilic side chain are essential for the antibacterial activity.7Open in a separate windowFig. 1Structures of brabantamides A–C (1–3). Rha = rhamnose.Surprisingly, only five reports concerning the synthesis of the bicyclic oxazolidinone with an exocyclic double bond have been reported to date.In 2000, Pinto et al. prepared a series of analogues of brabantamide A where the enol-carbamate 5 was first obtained by direct iodocyclization from acetylene derivative 4 and then converted into key ester 6 by carbonylation of the corresponding vinyl iodide in the presence of 2-(trimethylsilyl)ethanol and PdCl2 (Scheme 1a).4 Another synthesis of the related esters 9 has been reported by Snider et al. in 2006, employing Wittig reaction between stabilized ylides and bicyclic oxazolidindione 8 synthesized from l-proline 7 (Scheme 1b).8 Shortly after, bicyclic enol-carbamate 10 with an exocyclic methylene group was synthesized in one step from acetylene derivative 4 using gold(i) catalysts (Scheme 1c).9 Very recently, Witte et al. reported the synthesis of series of Z-analogues of brabantamides 12 by cyclization of β-ketoamides 11 using CDI (Scheme 1d).7 As a part of our ongoing research program aimed at the utilization of trifluoromethanesulfonic anhydride (Tf2O)/2-halopyridine (2-XPy) tandem in the synthesis of bioactive natural products and their analogues, we envisioned that similar reaction conditions could be effectively used to form the bicyclic oxazolidinone framework as found in brabantamides.Open in a separate windowScheme 1Literature syntheses of bicyclic enol-carbamates and method proposed herein.The combination of Tf2O/2-XPy was extensively and successfully used in the amide activation10,11 as well as in generating isocyanate species from N-Boc and N-Cbz protected amines.12 We anticipated that the N-Boc-protected β-ketoester 13 could react in its enolate form with in situ generated isocyanate ion by intramolecular 5-endo-dig cyclization to give bicyclic enol-cyclocarbamate 14 (Scheme 1e).At the start of our investigation, model β-ketoester 15, prepared from N-Boc-l-proline,7 was chosen as the model substrate to identify optimal reaction conditions (12c the initial using of 1.5 equivalents of Tf2O and 3 equivalents of 2-ClPy led to a full conversion of the substrate 15 in 15 minutes (monitored by TLC) (13 Any variation of the amount of 2-ClPy did not have any positive impact on the reaction ( EntryTf2OBaseTime (min)16a : 16baYieldb (%)11.5—60—–c21.5Et3N60—–c31.5DMAP60—–c41.5Pyridine60—–c51.5d2,6-Lutidine60—–c61.5dDBU6090 : 1041e,f71.52-ClPy1593 : 753e 8 1.1 2-ClPy 15 85 : 15 80 e 91.12-ClPy (1.5 equiv.)4085 : 1575e101.12-ClPy (5 equiv.)1587 : 1364e111.12-FPy1589 : 1171e121.12-BrPy7086 : 1473e131.12-IPy9086 : 1468e14Witte''s protocolgOvernight50 : 5036eOpen in a separate windowaRatio determined by 1H NMR of the crude reaction mixture.bIsolated combined yield.cTraces of products.dReactions performed with 1.1 equiv. of Tf2O did not lead to full conversion of ester 15.eReactions were performed on 1 mmol of ester 15.fReaction mixture contained a large amount of unidentified by-products.gReaction conditions: (1) TMSOTf (2 equiv.), CH2Cl2, 0 °C, 1 h. (2) CDI (1.5 equiv.), CH2Cl2, 0 °C – rt, overnight.7It is noteworthy that the reaction can be performed on a gram scale without affecting the yield and both isomers are easily separable by FCC (see the ESI).The 1H and 13C NMR data of the major E isomer 16a were consistent with those published previously.8 Possible racemization in the course of the reaction was dismissed based on the comparing specific optical rotation with the published data for 16a ([α]22D = −261.1 (c 1.01, MeOH); ref. 8: [α]22D = −207 (c 1.0, MeOH)). Most importantly, X-ray crystallographic analysis of 16a (Fig. 2; see the ESI for further details)14 confirmed its absolute configuration on the C-7a carbon atom.Open in a separate windowFig. 2Molecular structure of the enol-carbamate E-16a confirmed by X-ray crystallographic analysis.The minor Z isomer 16b was isolated for the first time as the pure compound and was fully characterized. Its structure was assigned on the basis of its 1H, 13C, COSY, HSQC, and HMBC NMR spectra.A plausible mechanism of the cyclization of β-ketoester 15 was based upon previous works12ac and it is depicted in Scheme 2. Isocyanate cation II, as a key intermediate, can be formed directly from iminium triflate I (path A) or through the formation of carbamoyl triflate III with subsequent elimination of triflate ion spontaneously (path B). Ester enolate moiety IV then reacts as O-nucleophile via 5-endo-dig cyclization and leads predominantly to the formation of the enol-carbamate 16a.Open in a separate windowScheme 2Plausible mechanism of the cyclization β-ketoester 15.Next, the optimized conditions were briefly applied in the synthesis of the brabantamide A analogue 21 (Scheme 3). Starting β-ketoester 18 was synthetized in two steps in a 70% yield using both commercially available N-Boc-d-proline 17 and 2-(trimethylsilyl)ethanol. It ought to be mentioned that previously examined hydrolysis of the corresponding methyl ester 16a under acidic as well as basic conditions failed due to the instability of the bicyclic enol-carbamate.3,8Open in a separate windowScheme 3Synthesis of the brabantamide A analogue 21.Subsequent cyclization of ester 18 using optimized reaction conditions afforded enol-carbamate 19 in 76% yield as a mixture of E and Z isomers in a ratio of 89 : 11. After isolation of the major isomer E-19a, it was treated with TBAF, providing free acid 20 in 89% yield. Finally, an amidation of 20 with tetradecylamine in the presence of EDCI gave amide 21 in moderate 40% yield. Both free acid 20 and amide 21 were fully characterized for the first time and their structures were assigned on the basis of its 1H, 13C, COSY, HSQC, and HMBC NMR spectra. Moreover, their structures were unambiguously confirmed by X-ray crystallographic analysis (Fig. 3; see ESI for further details).14Open in a separate windowFig. 3Molecular structures of acid 20 (top) and amide 21 (bottom) confirmed by X-ray crystallographic analysis.  相似文献   

14.
Catalytic enantioselective bromohydroxylation of cinnamyl alcohols     
Jing Li  Yian Shi 《RSC advances》2021,11(22):13040
This work describes an effective enantioselective bromohydroxylation of cinnamyl alcohols with (DHQD)2PHAL as the catalyst and H2O as the nucleophile, providing a variety of corresponding optically active bromohydrins with up to 95% ee.

Optically active bromohydrins are obtained with up to 95% ee via asymmetric bromohydroxylation of cinnamyl alcohols with H2O as nucleophile.

Electrophilic halogenation of olefins allows installation of two stereogenic centers onto the C–C double bond and is one of the most important transformations in organic chemistry.1 Optically active halogen containing products resulting from asymmetric halogenation would serve as versatile chiral building blocks for organic synthesis. As a result, extensive efforts have been devoted to the development of asymmetric halogenation process. In recent years, great progress has been made in both intramolecular2,3 and intermolecular4,5 reaction processes with various types of olefins and nucleophiles. However, there are still challenges remaining to be addressed. In many cases, the developed catalytic systems often only apply to certain ranges of substrates and the reaction reactivity as well as selectivity can''t be rationally adjusted. The substrate scope is also often difficult to be logically extended and requires much experimentation, largely due to the complexity of the reaction systems and the lack of clear understanding of the reaction mechanisms.Halohydroxylation of olefins simply with H2O as nucleophile is a classic electrophilic addition reaction in organic chemistry and produces synthetically useful halohydrins (Scheme 1). Asymmetric version of such process has been challenging with only a few reports.6,7 As part of our general intertest in asymmetric halogenation,8 recently we have been investigating the intermolecular asymmetric reaction processes, particularly with unfunctionalized olefins, which has been a long standing challenging problem. During such studies, we have found that up to 92% ee could be achieved for the bromoesterification of unfunctionalized olefins with (DHQD)2PHAL (Scheme 2, eqn (a)).9 This work represents an early example of asymmetric halogenation for unfunctionalized olefins with high enantioselectivity. To our delight, high enantioselectivity can also be achieved for bromohydroxylation with H2O upon further investigation, giving optically active bromohydrins with up to 98% ee (Scheme 2, eqn (b)).10 In our efforts to expand the reaction scope of the asymmetric bromohydroxylation, we have found that cinnamyl alcohols are effective substrates, giving the corresponding bromohydrins with up to 95% ee. Herein, we report our preliminary studies on this subject.Open in a separate windowScheme 1Asymmetric halohydroxylation of olefins.Open in a separate windowScheme 2Asymmetric oxybromination of olefins.Initial studies were carried out with (E)-3-(4-bromophenyl)prop-2-en-1-ol (1a) as substrate. Several bromine reagents were examined with 10 mol% (DHQD)2PHAL (3a) (Fig. 1) as the catalyst and 10 mol% (−)-camphorsulfonic acid (CSA) as additive in acetone/H2O (10 : 1) at −30 °C (Open in a separate windowFig. 1Selected examples of catalyst examined.Studies on reaction conditionsa
EntryCat.Br sourceAdditiveSolventYieldb (ee)c %
13aNBS(−)-CSAAcetone/H2O (10 : 1)79 (65)
23aDBDMH(−)-CSAAcetone/H2O (10 : 1)76 (62)
33aTBCO(−)-CSAAcetone/H2O (10 : 1)55 (7)
43aMeCONHBr(−)-CSAAcetone/H2O (10 : 1)48 (67)
53aPhCONHBr(−)-CSAAcetone/H2O (10 : 1)59 (76)
63bPhCONHBr(−)-CSAAcetone/H2O (10 : 1)18 (6)
73cPhCONHBr(−)-CSAAcetone/H2O (10 : 1)9 (0)
83dPhCONHBr(−)-CSAAcetone/H2O (10 : 1)35 (−57)
93e (quinidine)PhCONHBr(−)-CSAAcetone/H2O (10 : 1)31 (0)
103aPhCONHBr(−)-CSACH3CN/H2O (10 : 1)70 (83)
113aPhCONHBr(−)-CSAEtOAc/H2O (10 : 1)16 (67)
123aPhCONHBr(−)-CSATFE/H2O (10 : 1)43 (51)
133aPhCONHBr(−)-CSADCM/H2O (10 : 1)13 (70)
14d3aPhCONHBr(−)-CSACH3CN/H2O (5 : 1)66 (82)
15e3aPhCONHBr(−)-CSACH3CN/H2O (20 : 1)68 (81)
163aPhCONHBrCH3CN/H2O (10 : 1)36 (77)
173aPhCONHBr(+)-CSACH3CN/H2O (10 : 1)63 (82)
183aPhCONHBrPhCO2HCH3CN/H2O (10 : 1)34 (77)
193aPhCONHBr1-NapCO2HCH3CN/H2O (10 : 1)32 (77)
203aPhCONHBr p-TsOHCH3CN/H2O (10 : 1)68 (80)
213aPhCONHBrAlCl3CH3CN/H2O (10 : 1)39 (57)
22f3aPhCONHBr(−)-CSACH3CN/H2O (10 : 1)49 (85)
Open in a separate windowaReactions were carried out with substrate 1a (0.30 mmol), catalyst (0.030 mmol), additive (0.030 mmol), and Br source (0.36 mmol) in solvent/H2O (10 : 1) (3.0 mL + 0.3 mL) at −30 °C for 72 h unless otherwise noted.bIsolated yield.cDetermined by chiral HPLC analysis.dCH3CN/H2O (5 : 1) (3.0 mL + 0.6 mL).eCH3CN/H2O (20 : 1) (3.0 mL + 0.15 mL).fAt −40 °C for 168 h.With the optimized reaction conditions in hand, the substrate scope was subsequently investigated with 10 mol% (DHQD)2PHAL (3a), N-bromobenzamide (1.2 eq.), and 10 mol% (−)-CSA in CH3CN/H2O (10 : 1) at −30 °C. As shown in EntrySubstrateProductYieldb (%)eec (%) 1X = p-Br, 1a2a (X-ray)70832X = p-Cl, 1b2b64803X = p-F, 1c2c75764dX = p-Ph, 1d2d87905X = p-Me,1e2e (X-ray)76826X = m-Me,1f2f71627X = o-Me,1g2g77708X = H,1h2h4655 9X = Br, 1i2i708010X = Cl, 1j2j718011X = F, 1k2k848012X = Me, 1l2l7382 13X = Br, 1m2m729514X = Cl, 1n2n789415X = F, 1o2o839116X = Ph, 1p2p849417X = Me, 1q2q879018e 759019 3180Open in a separate windowaReactions were carried out with substrate 1 (0.50 mmol), (DHQD)2PHAL (0.050 mmol), (−)-CSA (0.050 mmol), and PhCONHBr (0.60 mmol) in CH3CN (5.0 mL) and water (0.50 mL) at −30 °C for 72 h unless otherwise noted.bIsolated yield.cDetermined by chiral HPLC analysis. For entry 1, the absolute configuration was determined by comparing the optical rotation of the corresponding epoxide with the reported one11 upon treatment with K2CO3 in acetone (Scheme 3). For others, the absolute configurations were tentatively assigned by analogy.dThe reaction was carried out at −40 °C for 168 h.eMeOH was used as nucleophile.The absolute configuration of bromohydrin 2a was determined by converting it to the corresponding epoxide 4 with K2CO3 (Scheme 3) and comparing the optical rotation of the epoxide with the reported one.11 The bromohydroxylation reaction can also be carried out on a relatively large scale. For example, 1.1341 g of bromohydrin 2m was obtained in 70% yield with 95% ee (Scheme 4). As shown in Scheme 5, bromohydrin 2m can be converted to bromoacetal 5 in 86% yield without loss of the ee. Sulfide 6 was obtained in 65% yield and 95% ee when 2m was reacted with sodium thiophenolate.Open in a separate windowScheme 3Determination of absolute configuration of bromohydrin 2a.Open in a separate windowScheme 4Bromohydroxylation on gram scale.Open in a separate windowScheme 5Synthetic transformations of bromohydrin 2m.Optically active bromoether like 2r could also serve as useful intermediates for further transformations (Scheme 6). Treating 2r with NaN3 in DMF at 80 °C gave azide 7 in 50% yield and 90% ee with inversion of configuration. The bromide of 2r could also be converted to chloride 8 in 90% ee while the yield was somewhat low. Epoxide 9 was obtained in 87% yield and 90% ee by treatment of 2r with NaOH in dioxane and water. When 2r was reacted with PhSNa in DMF at 80 °C, sulfide 10 was isolated in 73% yield and 90% ee. The reaction likely proceeded via epoxide 9. The synthetic application is further illustrated in Scheme 7. Azide 11 and chloride 12 were obtained from 9 in 80% and 78% yield, respectively, without erosion of the optical activity.12Open in a separate windowScheme 6Synthetic transformations of bromoether 2r.Open in a separate windowScheme 7Synthetic transformations of epoxide 9.A precise understanding of the reaction mechanism awaits further study. As previously described,10 two possible transition state models are outlined in Fig. 2. The substrate is likely docked in the chiral pocket through π,π-stacking with quinoline of the catalyst. Such π,π-interaction appeared to be enhanced by the substituents on the phenyl groups, consequently leading to the significant increase of the enantioselectivity. In model A, N-bromobenzamide was activated by both the tertiary amine of the catalyst and additive (−)-CSA to increase its electrophility toward the double bond of the reacting substrate. In model B, the tertiary amine of the catalyst could first be protonated by additive (−)-CSA, and N-bromobenzamide would subsequently be activated by the resulting quaternary ammonium salt via hydrogen bonding.Open in a separate windowFig. 2Two possible transition state models.  相似文献   

15.
Base mediated spirocyclization of quinazoline: one-step synthesis of spiro-isoindolinone dihydroquinazolinones     
Rapolu Venkateshwarlu  V. Narayana Murthy  Krishnaji Tadiparthi  Satish P. Nikumbh  Rajesh Jinkala  Vidavalur Siddaiah  M. V. Madhu babu  Hindupur Rama Mohan  Akula Raghunadh 《RSC advances》2020,10(16):9486
A novel approach for the spiro-isoindolinone dihydroquinazolinones has been demonstrated from 2-aminobenzamide and 2-cyanomethyl benzoate in the presence of KHMDS as a base to get moderate yields. The reaction has been screened in various bases followed by solvents and a gram scale reaction has also been executed under the given conditions. Based on the controlled experiments a plausible reaction mechanism has been proposed. Further the substrate scope of this reaction has also been studied.

A novel approach for the spiro-isoindolinone dihydroquinazolinones has been demonstrated from 2-aminobenzamide and 2-cyanomethyl benzoate in the presence of KHMDS as a base to get moderate yields.

Because of their strong potent biologically activity, heterocyclic compounds have been a constant source of inspiration for the invention of new drugs especially for pharmaceutical and agro chemical industries.1 Indeed, investigation of novel methods for the synthesis of various natural products and heterocycles has always been a challenging task in modern organic chemistry. Amid all, spiro based scaffolds have been found to be very interesting because of their structural diversity. In spite of their intrinsic structures and immense biological activity there is a tremendous demand for the chemistry of spiro-isoindolinone dihydroquinazolinones.2 Indeed, nitrogen containing heterocyclic compounds like spiro-oxindole, spiro-isoindoline, spiro-isoindolinone are playing a significant role in medicinal chemistry and synthetic transformations.3 Moreover these compounds present in many natural products as a core unit like Lennoxamine, Zopiclone, Taliscanine and Pazinaclone (Fig. 1). In addition, many unnatural spiro-isoindolinones show significant biological activities acting as anti HIV-1, antiviral, antileukemic, anesthetic and antihypertensive agents.4,5 Notably, the spiro-isoindolinone dihydroquinazolinone unit has been found to be a combination of two potent pharmacophore units of dihydroquinazolinone and spiro-isoindolinone. Inspite of their remarkable biological activity afore mentioned, various methods have been developed for their synthesis like lithiation approaches, base mediated protocols, Diels–Alder and Wittig reactions, electrophilic and radical cyclization, metal-catalysed reactions and various electrochemical procedures.6Open in a separate windowFig. 1Some biologically active spiro-isoindolinone and quinazolinone units.Previously, a number of metal catalyzed reactions have also been reported for the spiroannulations.7 Among all, Nishimura et al. developed an Ir(i) catalyzed [3 + 2] annulation of benzosultam and N-acylketimines with 1,3-dienes via C–H activation for the synthesis of aminocyclopentene derivatives. Further, Xingwei Li et al. developed a Rh(iii)-catalysed [3 + 2] annulation of cyclic N-sulfonyl or N-acyl ketimines with activated alkenes for the preparation of various spirocyclic compounds.8 Recently Yangmin Ma et al. developed a one pot nano cerium oxide catalyzed synthesis of spiro-oxindole dihydroquinazolinone derivatives (Scheme 1).5c However, development of these type of novel compounds is always challenging and more attractive. Indeed, to the best of our knowledge there are no reports for the synthesis of spiro-isoindolinone dihydroquinazolinones. This led us to give more attention to study these compounds.Open in a separate windowScheme 1Different strategies for the synthesis of quinazolinone units.In continuation of our earlier efforts9 for the synthesis of various dihydroquinazolinones, herein we would like to report KHMDS mediated synthesis of novel spiro-isoindolinone dihydroquinazolinones. We envisioned the retro synthetic pathway for these compounds, as depicted in Scheme 2. Accordingly these compounds could be synthesised from 2-aminobenzamide and methyl-2-cyanobenzoate or ethyl-2-cyanobenzoate.Open in a separate windowScheme 2Retro synthetic approach for the synthesis of quinazolinone unit.Indeed, in order to understand the reaction conditions, we have commenced the reaction by taking 2-amino-N-hexyl-benzamide (2) and methyl-2-cyanobenzoate (3) as a model substrates. However, in the initial phase of reaction optimisation, we have screened the reaction in different bases (Fig. 2) and to our delight amongst all the bases KHMDS, LiHMDS and NaHMDS were amenable to get moderate yields. However the reaction had not progressed at low temperatures (5 °C) and could improve the yield at room temperature. Moreover the reaction underwent complete conversion with 1.5 equivalents of base. Further, the reaction was also executed with ethyl-2-cyanobenzoate and could replicate the same yield. Incontinuation, the reaction was futile when the reaction was carried out in DIPEA, DBU, K2CO3 and Cs2CO3. Subsequently, the reaction in NaOMe and tBuOK produced exclusively the hydrolysis product (4) of methyl-2-cyanobenzoate (Open in a separate windowFig. 2Base screening in 1,4-dioxane.Optimisation of the base-mediated spiroannulationa
EntryBaseSolvent1jb (%)Cyano benzoic acid (4)(%)
1KHMDS1,4-Dioxane604
2NaHMDS1,4-Dioxane583
3NaOMe1,4-Dioxane60
4 t BuOK1,4-Dioxane60
5KHMDSTHF405
6KHMDS1,2-DME605
7LiHMDS1,4-Dioxane555
Open in a separate windowaReaction conditions: KHMDS (1 M, 1.5 mmol), 2-amino-benzamide (1 mmol) and methyl-2-cyanobenzoate (1.5 mmol) in 1,4-dioxane (10 mL).bIsolated yield.Gratifyingly, among all the solvents 1,4-dioxane, THF and 1,2-dimethoxyethane were found to get good to moderate yields. Whereas, other solvents like DCM ended up with non-polar spots where as in toluene unknown polar impurity was observed. However, there is no reaction progress observed in the presence of trifluoroacetic acid as well as in BF3·Et2O as a solvent (Fig. 3).Open in a separate windowFig. 3Solvent screening in the presence of KHMDS.With the optimized conditions in hand, we have explored the applicability of our reaction with various substrates by taking various groups like alkyl, cyclopropyl, cyclohexyl, cycloheptyl, benzyl, naphthyl, furan and to our delight all the substrates were well tolerated under the aforementioned optimal conditions ( Open in a separate windowaReaction conditions: KHMDS (1 M, 1.5 mmol), 2-amino-benzamide (1 mmol) and methyl/ethyl-2-cyanobenzoate (1.5 mmol) in 1,4-dioxane (10 mL).Based on the aforementioned studies and the literature reports, a plausible mechanism for this reaction has been predicted (Scheme 3). Indeed, to gain insight into the mechanism a series of control experiments have been executed under the similar reaction conditions. Initially the reaction has been carried out without base and both the starting materials were intact. Further the reaction without 2-aminobenzamide resulted hydrolysis product. To explore further, the reaction has also been executed on a 10 gram-scale for the synthesis of 1j and has successfully been demonstrated under the aforementioned optimized conditions.Open in a separate windowScheme 3Plausible mechanisms for the synthesis of spiro-isoindolinone dihydroquinazolinones.The Scheme 3 describes a plausible mechanism for the preparation of compound 1. Initially, KHMDS will abstract N–H proton of amide and nucleophilic nitrogen will attack the cyanobenzoate to get imine intermediate 6 and 7, which on subsequent cyclization lead to the formation of 8 and 9. Finally, the compounds 8 and 9 underwent cyclization to get the spiro-isoindolinone dihydroquinazolinone 1.  相似文献   

16.
A formal intermolecular [4 + 1] cycloaddition reaction of 3-chlorooxindole and o-quinone methides: a facile synthesis of spirocyclic oxindole scaffolds     
Chao Lin  Qi Xing  Honglei Xie 《RSC advances》2021,11(30):18576
Herein, we developed an efficient and straightforward method for the rapid synthesis of spirocyclic oxindole scaffolds via the [4 + 1] cyclization reaction of 3-chlorooxindole with o-quinone methides (o-QMs), which were generated under mild conditions. The products could be obtained in excellent yields with numerous types of 3-chlorooxindole. This methodology features mild reaction conditions, high atom-economy and broad substrate scope.

Herein, we developed an efficient and straightforward method for the rapid synthesis of spirocyclic oxindole scaffolds via the [4 + 1] cyclization reaction of 3-chlorooxindole with o-quinone methides (o-QMs), which were generated under mild conditions.

The structural diversity of spirocyclic oxindole scaffolds is a reason for their frequent occurrence in many relevant natural products and medicinal agents (Fig. 1).1 In particular, natural spirocyclic-2-oxindole scaffolds have been proven to exhibit a broad range of biological activities and have attracted increasing attention in the synthetic field. For instance, XEN 907 is a novel pentacyclic spirooxindole with excellent activities as sodium channel blockers.2 Due to their unique structure and intriguing biological activity, numerous methodologies have been developed for the construction of these privileged frameworks.3 For example, in the past few years, transition-metal catalyzed or organocatalytic [3 + 2] cycloaddition reactions have been developed for the synthesis of spirocyclic oxindole scaffolds.4 Despite the emergence of these elegant approaches, exploiting new strategies for the construction of spirocyclic oxindole derivatives is still highly desirable.Open in a separate windowFig. 1Examples of biologically active spirocyclic oxindole scaffolds. Ortho-quinone methides (o-QMs) as highly reactive versatile intermediates have been of great interest to the chemical and biological community.5o-QMs react with various classes of reagents by three typical reaction pathways: 1,4-addition of nucleophiles, [4 + 2] cycloaddition with dienophiles and oxa-6π-electrocyclization.6 Because most o-QMs are unstable, these reactions generally depend on the reaction conditions used for the generation of o-QMs in situ. Rokita et al. reported that o-silylated phenols when exposed to fluoride could also produce o-QMs under mild reaction conditions.7Because of the dual nature (nucleophilic/electrophilic) of the C-3 position, 3-chlorooxindole serves as a highly reactive starting material in the synthesis of spirocyclic oxindole scaffolds. The introduction of a chloro group at the C-3 position of indoles serves as an excellent leaving group in favour of the subsequent cyclization. In addition, this also increases the acidity of the C–H bond for directly entering the C-3 quaternary centers.8 Inspired by this reactivity profile, 3-chlorooxindoles have been successfully utilized for [2 + 1]9 and [4 + 1]10 cyclization to synthesize spirocyclic oxindole scaffolds (Fig. 2).Open in a separate windowFig. 2Representation of the synthesis and applications of 3-chlorooxindoles.We designed an efficient and straightforward method for the rapid synthesis of spirocyclic oxindoles via the [4 + 1] cyclization reaction of 3-chlorooxindole with o-QMs, which were generated under mild conditions. In this study, using TBAF as the fluoride source and base ensures that the one-pot domino reaction will occur in mild reaction conditions, with high atom-economy and broad substrate scope.Initially, we carried out optimization studies by examining the reaction between O-silylated phenol 2a and 3-chlorooxindole 1a. Indeed, when TBAF was employed as the fluoride source, a smooth [4 + 1] cyclization reaction occurred, affording the spirocyclic oxindole product 3a with 75% yield (entry 1, EntryF source X Y SolventTemp (°C)Yieldb1TBAFc1.24.0DCMrt75%2TBAF1.54.0DCMrt87%3TBAF2.04.0DCMrt85%4CsF1.54.0DCMrt11%5TBAF1.53.0DCMrt80%6TBAF1.54.0CHCl3rt83%7TBAF1.54.0THFrt94%8TBAF1.54.0Toluenert90%9TBAF1.54.0DMFrt72%10TBAF1.54.0MeCNrt84%11TBAF1.54.0MeOHrtNDOpen in a separate windowaReaction conditions: 1a (0.3 mmol), solvent (3.0 mL), 6 h.bIsolated yield.cTBAF (1 M in THF solution).With the optimal conditions known, we next investigated the substrate scope of substituted 3-chlorooxindole 1 using O-silylated phenol 2a as a representative ( Open in a separate windowaReaction conditions: 1 (0.3 mmol), 2a (1.5 eq.), TBAF (4.0 eq.), THF (3.0 mL), 6 h.bIsolated yields.Next, we also explored the substrate scope of substituted 3-chlorooxindole 1 using O-silylated phenols 2a′ ( Open in a separate windowaReaction conditions: 1 (0.3 mmol), 2a′ (1.5 eq.), TBAF (4.0 eq.), THF (3.0 mL), 6 h.bIsolated yields.On the basis of above-mentioned results, a plausible mechanism for this formal [4 + 1] cycloaddition reaction is depicted in Scheme 1. Initially, the highly reactive o-QMs are generated via the desilylation/elimination reaction. Then, 3-chlorooxindole 1a as a nucleophile attacks the external carbon of o-QMs, affording zwitterion Int-1. Finally, the zwitterion Int-1 loses one molecular HCl through a nucleophilic attack, yielding the spirocyclic oxindole product 3a.Open in a separate windowScheme 1Possible mechanism.  相似文献   

17.
Access to 1-amino-3,4-dihydroisoquinolines via palladium-catalyzed C–H bond aminoimidoylation reaction from functionalized isocyanides     
Zhuang Xiong  Panyuan Cai  Yingshuang Mei  Jian Wang 《RSC advances》2019,9(72):42072
Efficient access to 1-amino-3,4-dihydroisoquinolines, through palladium-catalyzed intramolecular C–H bond aminoimidoylation of α-benzyl-α-isocyanoacetates, has been developed. Consecutive isocyanide insertion and C–H bond activation were realized with C–N and C–C bonds formation in one step under redox neutral conditions, employing O-benzoyl hydroxylamines as electrophilic amino sources.

Efficient and practice access to 1-amino-3,4-dihydroisoquinolines, through palladium-catalyzed intramolecular C–H bond aminoimidoylation of α-benzyl-α-isocyanoacetates, has been developed.

As a class of cyclic amidine compounds, 1-amino-3,4-dihydroisoquinolines are essential six-membered heterocycles which can be found in numerous natural products, functional chemicals and pharmaceutical agents (Scheme 1).1 For example, piperazine tethered dihydroisoquinoline and benzocyclic amide I was found to be an efficient cardiovascular agent.1a Phenylalanine-derived and arylamine or cyclic aliphatic-amine-containing 1-amino-3,4-dihydroisoquinolines (II–V) show outstanding BACE-1 inhibition, kinase VEGFR-2 inhibition, phosphodiesterase IV inhibition and hydrogen ion-sodium exchanger NHE-3 inhibition activities, respectively.1be Traditional preparation of such 1-amino-3,4-dihydroisoquinoline scaffolds is mainly based on nucleophilic substitution of readily prepared 3,4-dihydroisoquinolines2 and Bischler-Napieralski reaction starting from urea.3 Most of these methods suffer from multiple-step synthesis, harsh conditions or the use of toxic reagents. Thus, developing an efficient and practical method for the construction of 1-amino-3,4-dihydroisoquinoline derivatives remain an attractive synthetic task.Open in a separate windowScheme 1(a) Representative bioactive molecules containing the 1-amino-3,4-dihydroisoquinoline moiety. (b) Traditional synthetic methods.In recent decades, isocyanides have been extensively studied in multicomponent reactions,4 imidoyl radical-involved transformations5 and transition-metal-catalyzed insertion reactions6 due to their diverse and unique reactivities. A variety of acyclic imine derivatives were synthesized via palladium-catalyzed imidoylation reactions.7 For the construction of biorelevant nitrogen-containing heterocycles via Pd-catalyzed non-functionalized isocyanide insertion reactions, a nucleophilic group is usually preinstalled to the substrates containing C–X (C–H) bonds, followed by Pd-catalyzed oxidative addition (C–H bond activation)/isocyanide insertion/ligand substitution/reductive elimination sequence.8 For example, Maes and coworkers developed an efficient palladium-catalyzed synthesis of 4-aminoquinazolines from N-(2-bromoaryl)amidines using amidine as an internal nucleophile.8a Access to the same scaffold through palladium-catalyzed amidine-directed C–H bond imidoylation/cyclization reaction was reported by Zhu group.8c Isocyanide acted as a C1 synthon in these cyclization reactions, with the terminal carbon of isocyanide being used in the ring formation.9 Considering the diversity of R group on the nitrogen atom of isocyanide, the so-called functionalized isocyanide strategy was widely developed as well, with various N-heterocycles such as oxazoles,10 indoles,11 phenanthridines,12 9H-pyrido[3,4-b]indoles13 and nonaromatic azepines14 being synthesized efficiently. Zhu and coworkers developed the first asymmetric C–H bond imidoylation reaction from easily accessible α,α-dibenzyl-α-isocyanoacetates with chiral quaternary-carbon-containing 3,4-dihydroisoquinolines being generated (Scheme 2a).15 Recently, a catalytic system controlled regioselective C–H bond imidoylation was realized for the selective synthesis of 5 or 6-membered cyclic imines (Scheme 2b).16 Aryl halides were mostly used in these transformations while N–O compounds were rarely reported as electrophiles in Pd-catalyzed isocyanide insertion reactions.17 Zhu group reported the synthesis of phenanthridin-6-amine derivatives from O-benzoyl hydroxylamines and commonly used biaryl isocyanides.17c However, a methyl or methoxyl carbonyl group was needed at the ortho position of isocyano group, which critically limited the application of the reaction. Herein, we developed an efficient and practical synthesis of 1-amino-3,4-dihydroisoquinolines, through palladium-catalyzed intramolecular C–H bond aminoimidoylation reaction of α-benzyl-α-isocyanoacetates. Consecutive isocyanide insertion and C–H bond activation were realized with C–N and C–C bonds being constructed in one step, employing O-benzoyl hydroxylamines as electrophilic amino sources.Open in a separate windowScheme 2Palladium-catalyzed imidoylative cyclization of α-benzyl-ethylacetates.The reaction conditions were optimized with ethyl 2-benzyl-2-isocyano-3-phenylpropanoate (1a) and morpholino benzoate (2a) as model substrates, and ethyl 3-benzyl-1-morpholino-3,4-dihydroisoquinoline-3-carboxylate (3a) was obtained in 34% yield in the presence of Pd(OAc)2 (1.5 equiv.), PPh3 (20 mol%), Cs2CO3 (1.0 equiv.) and PivOH (0.6 equiv.) in toluene at 80 °C (entry 1, EntryCatalystLigandSolventYieldb1Pd(OAc)2PPh3Toluene342Pd(OAc)2PPh3THFTrace3Pd(OAc)2PPh3DCE294Pd(OAc)2PPh3Dioxane765Pd(OAc)2PPh3MeCNTrace6Pd(OAc)2PPh3DMSOTrace7Pd(OPiv)2PPh3Dioxane318Pd(TFA)2PPh3Dioxane269PdCl2(MeCN)2PPh3Dioxane3811Pd(PPh3)4Dioxane5612Pd(OAc)2BINAPDioxane3013Pd(OAc)2dppbDioxane2614Pd(OAc)2dpppDioxane2815Pd(OAc)2XantPhosDioxane4216Pd(OAc)2SPhosDioxane3417Pd(OAc)2XPhosDioxane2418cPd(OAc)2PPh3Dioxane84 (80)d19cNonePPh3Dioxane020cPd(OAc)2NoneDioxane5621c,ePd(OAc)2PPh3Dioxane022c,fPd(OAc)2PPh3Dioxane4123c,gPd(OAc)2PPh3Dioxane2424c,hPd(OAc)2PPh3Dioxane3125c,iPd(OAc)2PPh3Dioxane61Open in a separate windowaReaction conditions: 1a (0.1 mmol), 2a (0.15 mmol), Pd-catalyst (0.01 mmol, 10 mol%), ligand (0.02 mmol, 20 mol%), Cs2CO3 (0.1 mmol, 1.0 equiv.), PivOH (0.06 mmol, 0.6 equiv.), 80 °C, Ar. A solution of 1a was added via a syringe pump within 1 h.bNMR yield with 1-iodo-4-methoxybenzene as an internal standard.c T = 110 °C.dIsolated yield.eWithout Cs2CO3 and PivOH.fCsOPiv was used as the base.gK2CO3 was used as the base.hNa2CO3 was used as the base.i5 mol% of catalyst was used.With the optimized conditions in hand, the scope of functionalized isocyanide was investigated first, using morpholino benzoate 2a as the coupling partner (Scheme 3). Both electron rich and withdrawing substituents such as Me, t-Bu, F, Cl, CF3, CO2Me and NO2 at the para position of aromatic ring can tolerate the reaction conditions, and generate the corresponding products in good to excellent yields (3a–3h). Isocyanides bearing ortho substituents on the phenyl ring underwent the reaction smoothly, generating the products 3i and 3j in 58% and 82% yields respectively. Even substrates bearing two substituents were well suited to the reaction conditions, affording the aminoimidoylation products without loss of efficiency (3k–3l). Compound 3m was delivered with excellent regioselectivity in 63% yield when m-Cl substituted starting material was employed in the reaction, which indicated the hindrance sensitivity of this reaction. Modifying the ester moiety in 1 to a methyl ester or an amide did not impede the reaction, giving 3n and 3o with excellent yields. Changing one of the benzyl groups to alkyl substituents didn''t decrease the reaction efficiency, generating products 3p and 3q in moderate yields.Open in a separate windowScheme 3Scope of isocyanides. Reaction conditions: 1 (0.1 mmol), 2a (0.15 mmol), Pd(OAc)2 (0.01 mmol, 10 mol%), PPh3 (0.02 mmol, 20 mol%), Cs2CO3 (0.1 mmol, 1.0 equiv.), PivOH (0.06 mmol, 0.6 equiv.), 110 °C, Ar. A solution of 1 was added via a syringe pump within 1 h. Isolated yields.Then we moved our attention to the investigation of substrate scope of aminating partners (Scheme 4). It was found that various O-benzoyl hydroxylamines derived from piperidines could be smoothly employed in the reaction (4a–4e). Substituents such as Me, CO2Et, OMe and ketal group on the piperidines were well tolerated (4b–4e). Aminating reagent derived from Boc-protected piperazine was also compatible with this reaction, generating the product 4f in 99% yield. Fortunately, acyclic aminated product 4g was obtained as well in moderate yield.Open in a separate windowScheme 4Substrate scope of aminating agents. Reaction conditions: 1a (0.1 mmol), 2 (0.15 mmol), Pd(OAc)2 (0.01 mmol, 10 mol%), PPh3 (0.02 mmol, 20 mol%), Cs2CO3 (0.1 mmol, 1.0 equiv.), PivOH (0.06 mmol, 0.6 equiv.), 110 °C, Ar. A solution of 1a was added via a syringe pump within 1 h. Isolated yields.When ethyl 2-isocyano-2,3-diphenylpropanoate was conducted in the reaction with 2a, the 3,4-dihydroisoquinoline product 6 was generated exclusively, albeit in only 37% yield (Scheme 5a). To examine if isoindoline could be afforded by this aminoimidoylation reaction, ethyl 2-isocyano-4-methyl-2-phenylpentanoate (7) was selected as the substrate with 2a (Scheme 5b). However, no desired product was generated under the standard conditions, showing different site-selectivity with the previous imidoylation reaction developed by Zhu and coworkers.16Open in a separate windowScheme 5Reaction conditions: 5(7) (0.1 mmol), 2a (0.15 mmol), Pd(OAc)2 (0.01 mmol, 10 mol%), PPh3 (0.02 mmol, 20 mol%), Cs2CO3 (0.1 mmol, 1.0 equiv.), PivOH (0.06 mmol, 0.6 equiv.), 110 °C, Ar. A solution of 5(7) was added via a syringe pump within 1 h. Isolated yields.Gram-scale preparation of 3a was carried out smoothly under standard conditions (Scheme 6a). The reduction reaction of 3a was carried out using LiAlH4 and the corresponding product 8 was generated in 70% yield (Scheme 6b). The ester group was successfully transformed to tertiary alcohol (9) in moderate yield with Grignard reagent.Open in a separate windowScheme 6Gram-scale preparation and diversifications of 3a.In order to get further insight into the reaction mechanism, a radical trapping reaction with 0.1 equivalents of TEMPO was carried out (Scheme 7a). The product 3a was generated without loss of efficiency, which indicated that the radical-involved pathway could be ruled out. A plausible mechanism was then proposed in Scheme 7b. The reaction was initiated with oxidative addition of O-benzoyl hydroxylamine 2a to the in situ generated Pd(0) species. Migratory isocyanide insertion to intermediate B afforded aminoimidoyl Pd(ii) intermediate C. The following intramolecular C–H bond activation and reductive elimination viaE yielded the final product 3a and regenerated Pd(0) to complete the catalytic cycle.Open in a separate windowScheme 7Proposed mechanism.In conclusion, we have developed an efficient and practical method to prepare 1-amino-3,4-dihydroisoquinolines via palladium-catalyzed intramolecular C–H bond aminoimidoylation of α-benzyl-α-isocyanoacetates under redox-neutral conditions. Consecutive isocyanide insertion and C–H bond activation were realized with C–N and C–C bonds formation in one step, employing O-benzoyl hydroxylamines as electrophilic amino sources. A broad range of substrates were applicable to the reaction in good to excellent yields.  相似文献   

18.
One-pot reductive amination of carbonyl compounds with nitro compounds over a Ni/NiO composite     
Yusuke Kita  Sayaka Kai  Lesandre Binti Supriadi Rustad  Keigo Kamata  Michikazu Hara 《RSC advances》2020,10(54):32296
Easily prepared Ni/NiO acts as a heterogeneous catalyst for the one-pot reductive amination of carbonyl compounds with nitroarenes to afford secondary amines with H2 as a hydride source. This catalytic system does not require a special technique to avoid air-exposure, in contrast to the common heterogeneous Ni catalysts.

Easy-to-prepare Ni/NiO acts as an efficient heterogeneous catalyst for one-pot reductive amination of carbonyl compounds with nitroarenes.

Amines are among the most important organic compounds for the chemical, materials, pharmaceutical and agrochemical industries.1 In particular, aromatic and heteroaromatic amines occupy a privileged position in medicinal chemistry,1c,2 as exemplified in top-selling drugs such as atorvastatin, hydrochlorothiazide, furosemide and acetaminophen.3 The general methods to prepare aryl and heteroaryl amines are the reductive amination of carbonyl compounds,4 direct alkylation of amines with alkyl halides,5 Buchwald–Hartwig amination6 and Ullman-type C–N bond formation.7 These amination systems utilize aniline derivatives which are usually prepared in advance by hydrogenation of nitroarenes.8 Direct amine synthesis using nitroarenes is attractive because it eliminates the hydrogenation step, which saves time, energy and cost. Among the amination reactions, reductive amination has been actively investigated due to its high atom economy and ease of industrial application (Scheme 1);9 however, reductive amination has frequently been accompanied by unwanted side products due to the reduction of carbonyl compounds to alcohols and/or over-alkylation of product amines. One-pot reductive amination with nitro compounds has been achieved by precious metal catalysis.10 Recently, precious metal alloy nanoparticles were demonstrated to exhibit high catalytic performance for one-pot reductive amination with nitro compounds, even under mild reaction conditions.11 In the context of economic efficiency and a ubiquitous element strategy, the replacement of precious metals with earth-abundant metals has gained much attention. Although some catalytic systems based on Fe,12 Cu,13 Co 14 and Mo 15 have been reported using H2 as a reductant, severe reaction conditions are typically required (Table S1, ESI). It is noteworthy that the nitrogen-doped carbon supported cobalt catalysts were reported to be active for one-pot reductive amination using nitroarenes using formic acid or CO/H2O as a reductant though high temperature was required (Table S2, ESI).16Open in a separate windowScheme 1One-pot reductive amination using nitroarenes.To develop active catalysts for one-pot reductive amination using nitro compounds, we have focused on nickel catalysts, which are known as active catalysts for many transformation reactions.17,18 The active species is typically metallic nickel, which is easily oxidized in air and becomes covered with NiO.19 Therefore, special techniques to avoid air-exposure (i.e., pre-reduction in the reaction vessel, glovebox) are required to achieve high catalytic performance for liquid phase reactions. Herein we report an easily prepared Ni/NiO composite as a heterogeneous catalyst for one-pot reductive amination using nitro compounds. This Ni catalyst can be handled under an air atmosphere, even though the supposedly active species, metallic nickel, is oxidized by air-exposure.Ni/NiO was prepared by the partial reduction of NiO with H2 in the temperature range from 200 to 500 °C (Ni/NiO-X: X = reduction temperature). X-ray diffraction (XRD) patterns of the prepared Ni catalysts after exposure to air are summarized in Fig. 1(A). When NiO is treated in H2 at 200 °C, the dominant phase produced is still nickel oxide. Metallic nickel was formed by reduction at ≥250 °C, which is consistent with the H2-temperature programmed reduction (H2-TPR) profile of NiO in which the H2 consumption peak begins to increase at around 250 °C.20 The ratio of Ni to NiO, determined by Rietveld analysis, increased with increasing reduction temperature and no peaks attributed to metallic nickel were evident for Ni/NiO-500 () and crystallite diameter estimated from (111) diffraction lines using Scherrer''s equation (Table S3, ESI). Fig. 1(C) shows an SEM image of Ni/NiO-300, where the particle size was estimated to be 0.5–3 μm, and the layered structure of NiO was maintained after the reduction treatment (see also Fig. S4, ESI).Open in a separate windowFig. 1(A) XRD patterns for Ni/NiO-X; (a) Ni/NiO-500, (b) Ni/NiO-400, (c) Ni/NiO-350, (d) Ni/NiO-300, (e) Ni/NiO-250, and (f) Ni/NiO-200 (♦: Ni, ○: NiO), and (B) SEM images of Ni/NiO-300.Specific surface areas and weight ratios of Ni to NiO
EntryCatalystSpecific surface areaa (m2 g−1)Weight ratiob (%)
NiNiO
1Ni/NiO-20082100
2Ni/NiO-25078298
3Ni/NiO-300414357
4Ni/NiO-350256634
5Ni/NiO-400108515
6Ni/NiO-500<5100
Open in a separate windowaSpecific surface areas were obtained from BET measurements.bWeight ratios were obtained by Rietveld analysis.The one-pot reductive amination of benzaldehyde (2a) with nitrobenzene (1a) was evaluated with the prepared Ni catalyst using molecular hydrogen as the reductant (). For comparison of Ni/NiO with simple supported Ni catalysts, one-pot reductive amination was conducted over Ni catalysts supported on simple metal oxides of Nb2O5, TiO2, SiO2 and ZrO2 (entries 9–12). Ni/SiO2 exhibited comparable activity to Ni/NiO (entry 9); however, significant leaching (3.9%) of the Ni species was observed in the reaction mixture, as opposed to that with Ni/NiO-300 (0.1%). Although RANEY® Ni is known to be active for reductive amination,213aa was obtained only in 22% yield under the present reaction conditions (entry 13). No desired product was observed using unreduced NiO and Ni(OH)222 (entries 14 and 15). In the case of the reactions with low material balances such as Ni/NiO-400, Ni/Nb2O5 and Ni/ZrO2, we observed benzyl alcohol as a main byproduct, suggesting that the hydrogenation of aldehydes is the competitive side reaction. Because of the high selectivity of Ni/NiO-300, NiO support can contribute to suppress the hydrogenation of aldehyde. Further optimization was then conducted (Table S2, ESI). The yield of 3aa was finally increased to 92% under higher concentration conditions using 1.5 equivalents of 2a (entry 4). Reductive amination proceeded even under lower hydrogen pressure (0.5 MPa), although a longer reaction time was required (entry 5). The lower loading of Ni/NiO (15 mg) was accomplished by prolonging the reaction time, affording 3aa in 90% yield (entry 10, Table S4).Catalyst screeninga
EntryCatalystConv. of 1a (%)Yield of 3aa (%)Yield of 4aa (%)
1Ni/NiO-20024
2Ni/NiO-250224
3Ni/NiO-300>99775
4bNi/NiO-300>9992
5cNi/NiO-300>9989
6Ni/NiO-350>99622
7Ni/NiO-40082930
8Ni/NiO-500166
9Ni/Nb2O5>991251
10Ni/TiO23627
11Ni/SiO2>997317
12Ni/ZrO2>992533
13dRANEY® Nie>9922
14NiO23
15Ni(OH)220
Open in a separate windowaReaction conditions: catalyst (0.05 g), 1a (1 mmol), 2a (1 mmol), toluene (5 mL), H2 (1 MPa), 80 °C, 20 h. Conversion and yield were determined by GC analysis.b1.2 mmol of 2a and 1 mL of toluene were used.cRun at 0.5 MPa H2 pressure for 50 h.dMethanol was used as a solvent.eGenerated by the treatment of RANEY® alloy with NaOH.The reusability of Ni/NiO-300 was examined next. The used Ni/NiO-300 could be recovered from the reaction mixture by simple filtration, washing with methanol, and drying at 90 °C. The XRD patterns and the ratio of Ni to NiO were not changed during one-pot reductive amination (Fig. S2, ESI). The SEM images of the recovered catalyst indicates that morphological changes were negligible (Fig. S6, ESI). The recovered Ni/NiO catalyst was reactivated by pre-treatment (150 °C, 1 h under H2 flow), by which the catalytic activity was maintained at a high level without obvious decline, even after the 3rd reuse (Fig. 2(A)).Open in a separate windowFig. 2(A) Reuse experiments. Reaction conditions: Ni/NiO-300 (0.05 g), 1a, (1 mmol), 2a, (1.2 mmol), toluene (1 mL), H2 (1 MPa), 80 °C, 20 h. (B) Time course of the one-pot reductive amination over Ni/NiO-300. Reaction conditions: Ni/NiO-300 (0.05 g), 1a (1 mmol), 2a (1.2 mmol), toluene (1 mL), H2 (1 MPa), 80 °C. (C) Reaction pathway for one-pot reductive amination over Ni/NiO-300.A time-course analysis was then conducted under the optimized conditions (Fig. 2(B)). Aniline was generated by the hydrogenation of nitrobenzene, and imine 4aa was then gradually formed through the dehydrative coupling of 2a with aniline. Hydrogenation of 4aa subsequently afforded the secondary amine 3aa. To determine which step is key for the one-pot reductive amination, each step was examined using Ni/NiO-300 and Ni/SiO2, respectively. Imine formation step proceeded smoothly even without catalyst (Table S5, ESI). For the hydrogenation of 1a, Ni/NiO and Ni/SiO2 gave the comparable results (Table S6, ESI). For the imine hydrogenation step, we observed the higher activity of Ni/NiO than Ni/SiO2 (Fig. 3 and Table S7). With these results, the high hydrogenating ability of Ni/NiO for imine hydrogenation is the key for the high activity on one-pot reductive amination.Open in a separate windowFig. 3Hydrogenation of 4aa over Ni/NiO and Ni/SiO2. Reaction conditions: catalyst (0.05 g), 4aa (1 mmol), toluene (1 mL), H2 (1 MPa), 80 °C.Ni/NiO-300 could be applied to the one-pot reductive amination of other substrates (). The electron-withdrawing group on the benzaldehyde derivative retarded reductive amination, although it could facilitate both imine formation and imine hydrogenation (entry 6). Aliphatic aldehyde was also applicable though the yield was low due to the decomposition of aldehyde (entry 9).Substrate scopea
Entry12Isolated yield of 3 (%)
1b 2a93
2c 2a74
3 2a88
4 2a82
5c1a 62
6b,c1a 94
7b1a 98
8d1a 75
9d1a 21e
Open in a separate windowaReaction conditions: Ni/NiO-300 (0.05 g), 1a (1 mmol), 2a (1.2 mmol), toluene (1 mL), H2 (1 MPa), 80 °C, 20 h.bRun at 100 °C.cRun for 40 h.dRun at 120 °C for 96 h.eNMR yield.The surface of metallic Ni is easily oxidized in air; therefore, the surface states of Ni/NiO-300 were analysed by X-ray photoelectron spectroscopy (XPS) and the results are shown in Fig. 3. In the Ni 2p region of the spectrum for Ni/NiO-300, the main 2p3/2 peak is observed at 856.1 eV, which is assignable to Ni(OH)2.23 The catalyst was exposed to ambient conditions; therefore, hydroxylation of the surface was inevitable.24 Similarly, NiO and RANEY® Ni were also covered with Ni(OH)2 after exposure to ambient conditions (Fig. 4). Considering the lack of activity for NiO and the low activity of RANEY® Ni for one-pot reductive amination (25 in the XPS spectrum suggests that Ni2+ species covers the catalyst surface. This is the reason why Ni/NiO can be handled under an air atmosphere, in contrast to common heterogeneous Ni catalysts.Open in a separate windowFig. 4Ni 2p XPS spectra for (a) Ni/NiO-300, (b) NiO, and (c) RANEY® Ni after exposure to ambient conditions.In summary, Ni/NiO acts as a catalyst for one-pot reductive amination with nitro compounds to afford the secondary amines. No special technique (pre-reduction in the reaction vessel or glovebox) is required in the reaction setup. The reaction could proceed under milder conditions than those reported for typical catalytic systems. Ni/NiO could be reused without any significant loss of activity. This catalytic system could be applied to a variety of substrates that bear functional groups. Mechanistic studies suggested that Ni(OH)2 on metallic nickel can exhibit catalytic activity toward the present one-pot reductive amination. These results provide new insights into the development of heterogeneous Ni catalysts.  相似文献   

19.
Boron nitride nanochannels encapsulating a water/heavy water layer for energy applications     
Farzaneh Shayeganfar  Javad Beheshtian  Rouzbeh Shahsavari 《RSC advances》2019,9(11):5901
Water interaction and transport through nanochannels of two-dimensional (2D) nanomaterials hold great promises in several applications including separation, energy harvesting and drug delivery. However, the fundamental underpinning of the electronic phenomena at the interface of such systems is poorly understood. Inspired by recent experiments, herein, we focus on water/heavy water in boron nitride (BN) nanochannels – as a model system – and report a series of ab initio based density functional theory (DFT) calculations on correlating the stability of adsorption and interfacial properties, decoding various synergies in the complex interfacial interactions of water encapsulated in BN nanocapillaries. We provide a comparison of phonon vibrational modes of water and heavy water (D2O) captured in bilayer BN (BLBN) to compare their mobility and group speed as a key factor for separation mechanisms. This finding, combined with the fundamental insights into the nature of the interfacial properties, provides key hypotheses for the design of nanochannels.

Single layer water (SLW) on BN layer and encapsulated between bilayer BN (BLBN) as nanochannel.

Hexagonal boron nitride (h-BN) is a synthetic material, similar to graphene in structure and layered form,1,2 but with alternating B and N atoms. Among others properties, h-BN is considered for its hydrophobicity, remarkable chemical and thermodynamic stability (air stable up to 1000 °C),3 and exceptional mechanical4 and optical properties.5 Thin BN sheets have been fabricated by high-energy electron beam irradiation,6 ultrasonication,7 Lewis acid–base,8 hydrolysis of lithium based materials,9 liquid alloys of alkali metals at room temperature10 and chemical vapor deposition2 methods.Nanochannels by design is an appealing idea due to their potential applications in filtration and separation as well as molecular sieving.11,12 Recently, Agrawal et al.13 revealed the water phase transition encapsulated within carbon nanotubes of different radius. Specifically, fabricating artificial capillaries for molecular transport led to emergence of nanofluidics.14 Shultz et al.15 studied the local water structure by investigation of local and long-range response of hydrogen-bond network of water with surface. Water is polar due to the high electronegativity of oxygen atom and two weakly electropositive hydrogen atoms besides the buckled structure.16 Graphene-based materials such as graphene oxide (GO) membranes can have nanometer pores and tunable sieving of ions,12 indicating low frictional water flow. Sun et al.17 studied on electro/magneto modulated ion transport through GO membranes. While several experimental works demonstrate promising applications of nanochannels, the fundamental subatomic and electronic phenomena at the heart of such nanochannels are poorly understood.Herein, we present a comprehensive ab initio study to decode the complex interfacial interactions of water and heavy water encapsulated in bilayer BN (BLBN) as a model system. Our findings demonstrated that 2D BN can be used as nanochannels with tunable interlayer distance that is determined by in/out-plane hydrogen bonds of water layers. To discover the mechanism for separating water and heavy water, we further investigate phonon spectrum for water layers between BLBN and for heavy water, deuterium oxide D2O. This nanometer-scale BN capillaries is a flagship architecture in fabricating atomically channels for nanofluidic technology.18BN nanosheet could be a good candidate for water cleaning and purification as experimentally has been reported by Lei, et al.19 Moreover, Hao et al. in their paper20 reported the properties of atomic layer deposition (ALD) boron nitride nanotube, which is relevant for water purification. Li et al.21 reported that porous boron nitride nanosheets show an excellent performance for water purification.More recently Falin et al.22 synthesized single crystalline of mono/few layer BN nanosheets. They experimentally measured the Young''s modulus of high quality 1–9 layer BN.21 Mechanical properties of BN nanosheets placed them among the strongest insulating materials, indicating the fracture strength of 70.5 ± 5.5 GPa.21To obtain fundamental insights on the interfacial interactions, a series of DFT calculations were performed on single layer water (SLW) molecules and double layer water (DLW) molecules settled on BN layers and encapsulated in BLBN (see Methods). We first examined the molecular models of SLW settled on BN layers with and without external field. Fig. 1a and b show SLW/BN, including (H2O)16, representing 16 water molecules and the substrate containing 120 B, N atoms. In order to take into account the effect of structural boundary conditions and compare the interfacial properties of SLW and DLW on BN surface, we employed the energy minimized structures for the SLW and DLW on BN layer (Fig. 1d and e), containing 32H2O molecules. The electrical dipole moment (Open in a separate windowFig. 1Optimized single layer water (SLW)/BN and its electronic band structure for different external electric field, (a) E = 0 V Å−1, (b) Total density of states (DOS) and (c) projected DOS for SLW/BN. Optimized double layer water (DLW)/BN and its electronic band structure for different external electric field, (d) E = 0, (e) E = 0.1 V Å−1, (f) total DOS for nanochannels of (a) and (b).DFT results of interfacial properties of optimized SLW and DLW on BN
SLW/BNDLW/BN
External electric field Eext (V Å−1)0.00.10.20.30.00.10.20.3
E int (eV)−2.41−2.57−2.73−2.95−2.11−2.23−2.37−2.53
d O–H (Å)0.971.01.021.040.991.01.011.02
Deformation (BN) (Å)0.10.130.150.170.0330.0980.110.14
Electrical dipole (pz) (debye)3.179.311.425811
Pressure (GPa)0.170.130.10.080.50.410.240.16
Bandgap (eV)4.84.94.74.854.9354.9
Water net charge (|e|)−0.12−0.15−0.18−0.2−0.1−0.13−0.18−0.2
Open in a separate window Fig. 1 shows the electronic band structure and total density of states (DOS) of these systems. Fig. 1b shows that DOS of SLW/BN has an energy shift while the presence of SLW does not significantly affect the electron density, confirming physisorption of SLW/BN.23 This finding suggests that the interaction of water with monolayer BN is weak, consistent with the hydrophobicity of BN. The interfacial SLW can infuse positive charges on the monolayer BN, i.e., some electrons from the valence band (VB) of BN are transferred to SLW because of the existence of OH groups. This charge induction cause the DOS of SLW/BN system to be shifted by 0.15 eV below the Fermi energy of the isolated BN, indicating p-doping of the contacting BN.The electron density of states are experimentally measured by using scanning tunneling microscopy (STM). Therefore, to analyze orbital contribution of designed nanochannel, we plotted projected density of state (PDOS) in Fig. 1c for SLW/BN. This figure (PDOS) reveals that the interfacial interaction of water and BN layer is dominated by P-orbitals.An analogous DOS information can be obtained for DLW/BN. Applying Eext in the system enhances the distance between H and O atoms dO–H. For instance, when Eext is 0.1 V Å−1, dO–H elongates to 1.0 Å, or 1.02 Å for Eext = 0 : 2 (Fig. 2 with and without applying Eext. In order to clarify the influence of BLBN on the interfacial properties of water layers, we are presenting electronic band structure and DOS for SLW (DLW)@BLBN in Fig. 2. DFT results of the interfacial properties of these systems are presented in Open in a separate windowFig. 2Optimized single layer water (SLW)@BLBN and its electronic band structure for different external electric field, (a) E = 0 V Å−1, (b) total density of states (DOS) and (c) projected DOS for nanochannels of (a). Optimized double layer water (DLW)@BLBN and its electronic band structure for different external electric field, (d) E = 0, (e) E = 0.1 V Å−1, (f) total DOS for nanochannels of (a) and (b).DFT results of the structural and interfacial properties of optimized SLW and DLW between BLBN
SLW/BLBNDLW/BLBN
External electric field Eext (V Å−1)0.00.10.20.30.00.10.2
E int (eV)−1.96−2.04−2.18−2.78−0.6−1.43−1.62
d O–H (Å)0.970.991.01.021.01.011.03
Deformation (BN) (Å)0.130.150.150.140.80.970.96
Electrical dipole (pz) (debye)4.58.312.328.30.074.69.6
Pressure (GPa)0.240.180.120.090.420.280.09
Bandgap (eV)4.34.234.34.34.53.43
Water net charge (|e|)−0.19−0.22−0.3−0.39−0.11−0.19−0.28
Open in a separate windowComparing the intermolecular energy (Eint) and vdW pressure of SLW or DLW @BLBN in Fig. 2b describes the total DOS, while Fig. 2c shows the orbital contribution of SLW@BLBN. Again, one can observe that P-orbital is dominating over S-orbital for BLBN and confirms our obtained electronic properties. For instance, SLW@BLBN system shows 4.3 eV band gap smaller than SLW/BN (4.8 eV), arising from hybridization of S-orbital and P-orbital and creating π bond in BLBN weaker than SLW/BN. This weaker π bond yields to correlation of electrons in conduction and valance band, creating smaller band gap. Fig. 2f shows a few Fermi energy shift for 2D water networks encapsulated in BLBN with Eext and small differences in their electronic properties. The band structure and DOS associated with the SLW@BLBN (Fig. 2a–c) support the adsorption of SLW, suggesting that the weakly dispersed states in the BN can be linked to the hydrogen networks between the water molecules. Fig. 2d and e. Dispersed bands in conduction bands (CB) (Fig. 2e) indicate increasing group velocity and electron mobility, imparting the polar character in these structures and band gap reduction, as reported previously.24 In particular, a slight displacement of the VB and CB toward Fermi level region is observed with increasing the electric field (Fig. 2f).25Finally, we study the double layer heavy water (deuterium oxide) (DLD) between BLBN (Fig. 3). Phonon dispersion plots of isolated DLW (as a reference) and DLW encapsulated in BLBN are depicted in Fig. 3a and b, respectively. Fig. 3b shows that variation of the vibrational phonon modes of DLW@ BLBN includes additional details on the intermolecular interaction.Open in a separate windowFig. 3Phonon spectrum for (a) isolated double layer water (DLW), (b) DLW@BLBN, (c) isolated deuterium oxide (heavy water) (DLD) and (d) DLD@BLBN.First and foremost, no negative phonon vibrations exist for both cases. Second, we observe that splitting occurs for long wavelengths of DLW@BLBN, as compared to isolated DLW. The origin of this splitting is the interior electric filed created by charge transfer in such system, in line with previous reports for ice systems.22 Simply put, the charge transfer between DLW and BLBN induces interior electric field, arising from ionic displacements. Then, the long range Coulomb interactions between DLW water with BLBN cause splitting of the long wavelength. Third, there is some stop bands (bandgaps) in frequency spectra due to non-unity mass ratio of DLW.26 The dispersion of bands is strongly influenced by intermolecular interaction of DLW trapped inside BLBN, while the stop bands are shifted up. Note that the acoustic phonon modes (small frequency extent) are dispersed for DLW@BLBN due to the interactions of DLW with localized electrons of BN.These plots confirm that disperse bands lead to higher carrier mobility in contrast to localized carriers presented by flat bands.27 A common trend in these phonon spectra is that the contribution of optical phonons (large frequency extent) to the intermolecular interaction is negligible due to the flat bands, considering the low phonon group velocity. Fig. 3d presents our results for phonon dispersion of double layer deuterium oxide (heavy water) (DLD)@BLBN, and Fig. 3c extends our findings for isolated DLD. For isolated DLD, we observe the expected phonon spectra data as experimentally reported. Interestingly, comparison of Fig. 3c and d indicate that the phonon modes differ for isolated DLD and DLW. Different trends can be secured from these phonon plots.The first significant result lies in comparison of isolated DLD (Fig. 3c) and phonon spectra of DLD between BLBN (Fig. 3d), which is not drastically affected. The perceived reason can be considered as greater closed shell character of deuterium oxide than water, causing weaker interaction between DLD@BLBN than DLW@BLBN. The number of flat bands increases for large frequencies, indicating more localized carrier.28 Another consequence of phonon spectra is the distribution of stop bands in Fig. 3d, which differs from that of isolated DLD where the stop bands are shifted up.Our results for deuterium oxide and water interaction are in line with experimental reports on scalable methods for hydrogen isotopes enrichment, due to the high level energy barrier of deuterons compared to proton.29 For instance, it is reported that the mobility of hydrogen is greater than deuterium ions in parent gases30 or deuterons permeate through BN nanostructures more slowly than protons, achieving a separation factor of ∼10.28 Exclusive properties of water encapsulated between 2D materials such as BN could lead to wide applications of nanochannels (e.g., separation, drug delivery, energy harvesting) by making nanometer-sized capillaries. Overall, these fundamental insights could be useful to understand, correlate and predict experimental results.In summary, we investigated the interfacial and electronic properties of single/double layer water and heavy water settled on monolayer and double layer BN using DFT calculations. The formation of DLW as a network reduces the extent of OH deformation, and their interfacial interactions on BN become less prevalent. Thus, SLW/BN is energetically more stable than DLW/BN. The formation of several directional hydrogen bonds between the water molecules in DLW compensates the stability of the water layer on the BN substrate. The phonon spectra of DLW@BLBN have strong dispersion related to their high mobility and group speed, supporting higher rate of water interaction through BLBN, in line with experiments.28 Considering deuterium oxide, the phonon bands are less dispersed than isolated water spectra, suggesting that slower DLD mobility than water, thus an effective separation tool for D2O and H2O through such 2D materials. Moreover, phonon spectra of DLW@BLBN reveals that high mobility of water inside BLBN is originated from the weak interaction of water with BLBN. Broadly, the findings and methods of this work can have implications on fundamental understanding and designing effective filtration tools using a host of other 2D mono- and multi-layer atomic sheets (e.g. molybdenum disulfide, niobium diselenide, layered double hydroxides, etc.) for separation, drug delivery, and energy harvesting.  相似文献   

20.
Efficient synthesis of spirooxindolyl oxazol-2(5H)-ones via palladium(ii)-catalyzed addition of arylboronic acids to nitriles     
Hao Song  Na Cheng  Li-Qin She  Yi Wu  Wei-Wei Liao 《RSC advances》2019,9(50):29424
A versatile synthesis of spirooxindolyl oxazol-2(5H)-ones via palladium(ii)-catalyzed addition of arylboronic acids to nitriles is described. A wide range of spirooxindolyl oxazol-2(5H)-ones and other spirocyclic frameworks incorporating the oxazol-2(5H)-one unit can be readily prepared in good to high yields under the optimal conditions.

A versatile synthesis of spirooxindolyl oxazol-2(5H)-ones and derivatives via palladium(ii)-catalyzed addition of arylboronic acids to nitriles is described.

Rapid and efficient construction of pharmaceutical and biologically relevant compounds plays a very important role in modern organic synthesis, and constitutes the original impetus for the development of various novel synthetic approaches. The efficient construction of spirocyclic frameworks has been a topic of great relevance in organic synthesis due to their inherent three-dimensional architectures and the pronounced biological activities.1 In particular, the spirocyclic oxindoles have emerged as attractive synthetic targets because of their prevalence in numerous natural and unnatural products.2 Notably, the enhanced biological activities have been observed by the incorporation of a spiro five-membered azaheterocyclic ring at the C3 position of the oxindole core (Fig. 1).3 Thereby, a variety of synthetic methods have been developed to access analogous compounds possessing such privileged structure moieties.4Open in a separate windowFig. 1Examples of spiro oxindoles containing natural products and biological relevant compounds.As one of the important N–O heterocyclic compounds, oxazolidinones and their derivatives have been widely used not only as synthetic building blocks,5 but also as pharmaceuticals6 and agrochemicals,7 owing to a diverse range of biological activities.8 Although great contributions have been made to access these valuable scaffolds,9 the construction of structurally diverse spirooxindolyl oxazol-2(5H)-ones, characterized by a spiro ring fusion at the C3 position of the oxindole core with oxazol-2(5H)-one motif, has received less attention from synthetic community,10 despite the fact that these spirocyclic heterocycles could be promising candidates possessing biological responses. In 2017, He and co-workers reported a formal [3 + 2] cycloaddition reaction of in situ generated azaoxyallyl cation with cyclic ketones for the synthesis of spiro-4-oxazolidinones.11a In 2018, Alla and co-workers described a copper-catalyzed one-pot multicomponent protocol for the synthesis of spiro(indoline-3,5′-oxazolidine)-2,2′-diones starting from ketones, arylacetylenes and isocyanates.11bRecently, the transition-metal-catalyzed addition of organoboron reagents to nitriles has received remarkable progress,12 since the elegant works on the addition of arylpalladium species to the cyano group reported by Larock and Lu et al.,13 in which nitriles served as C building blocks and provided aryl ketones. In virtue of palladium-catalyzed tandem addition of organoboron reagents to nitriles/cyclization protocol, this approach enabled the combination of organoboron reagents and nitriles to construct a diversity of nitrogen-containing heterocycles such as 2-aminobenzophenones, benzofurans, and indoles, in which nitrile serves as C–N synthon instead and is incorporated into heterocyclic frameworks in an atom-economical fashion.14 However, the development of transition-metal-catalyzed tandem sequence involving the addition of organoboron reagents to nitriles to construct structural novel three-dimensional architectures such as spirocyclic systems is still undeveloped.We have recently developed both intramolecular and intermolecular cyclization approaches to prepare indole and thiophene fused polycyclic derivatives via Pd-catalyzed direct C–H bond addition to nitriles.15 Given the promising biological activities of spirooxindoles-containing molecules in medicinal chemistry and our ongoing interest in the development of efficient catalytic processes to prepare diverse aza-heterocyclic frameworks, herein, we report an efficient synthetic approach to prepare spirooxindolyl oxazol-2(5H)-ones via palladium(ii)-catalyzed addition of arylboronic acids to nitriles.As functionalized nitriles, cyanohydrins which are readily prepared from ketones and aldehydes have demonstrated considerable synthetic potential as useful building blocks.16 We chose the Pd(ii)-catalyzed reaction of 3-cyano-1-methyl-2-oxoindolin-3-yl ethyl carbonate 1a, which is readily prepared from isatin and ethyl cyanoformate, and phenylboronic acid 2a as a model reaction for the optimization of the reaction conditions (ii) catalyst proven to be essential to this transformation since no reaction happened without them (). In addition, the reaction also was evaluated with Ni(ii) catalyst system. However, Ni(ii)-catalyzed reaction gave the inferior to that of Pd(ii) catalytic system ().Effects of reaction parametersa
EntryCat.LigandSolvent t (h)Yieldb (%)
1Pd(OAc)2L1Toluene2427
2Pd(OAc)2L1THF2477
3Pd(OAc)2L1DMF2479
4Pd(OAc)2L1DMSO2471
5Pd(OAc)2L1NMP2482
6Pd(TFA)2L1NMP2477
7Pd(acac)2L1NMP2488
8Pd(acac)2L2NMP2486
9Pd(acac)2L3NMP2473
10Pd(acac)2L4NMP2488
11c,dPd(OAc)2L1NMP3691
12c,ePd(OAc)2L1NMP3683
13eL1NMP24nd
14ePd(OAc)2NMP24nd
15dNMP24nd
16c,d,fPd(OAc)2L1NMP3679
17gNi(acac)2L2MTBE2467
Open in a separate windowaReaction conditions: 1a (0.2 mmol), 2a (0.6 mmol), catalyst (10 mol%), ligand (12 mol%) and HOAc (10 equiv.) in solvent (1 mL) at 80 °C.bIsolated yields.cPd(OAc)2 (5 mol%) and bpy (6 mol%) were used.dHOAc (5.0 equiv.) was used.eWithout HOAc.f2a (0.4 mmol) was used.gNi(acac)2 (10 mol%), L2 (12 mol%) and Cs2CO3 (20 mol%) in MTBE (1 mL) at 110 °C. L1: 2,2′-bipyridine; L2: 4,4′-dimethyl-2,2′-bipyridine; L3: 5,5′-dimethyl-2,2′-bipyridine; L4: 1,10-phenanthroline.With the optimized reaction conditions in hand, the generality of the Pd-catalyzed addition/cyclization sequence for the preparation of spirooxindolyl oxazol-2(5H)-ones was evaluated by employing various isatin based-O-ethoxycarbonyl cyanohydrins 1 and phenylboronic acid 2a first (Scheme 1). Other than N-methyl substrate 1a, cyanohydrin analogues 1 bearing different N-substituents such as phenyl, benzyl, p-methoxybenzyl and p-nitrobenzyl can give the desired products 3ba–3ea in high yields. The substitution pattern at the benzene ring of cyanohydrins 1 has little influence on the results, and high yields could be obtained (3fa–3ia). In addition, the reactions between substrates possessing both electron-donating (MeO and Me) and electron-withdrawing (NO2, Br, Cl and I) substituents at the benzene ring and phenylboronic acid 2a proceeded well, and gave the corresponding products with excellent yields (3ja–3oa). The structures of spirooxindolyl oxazol-2(5H)-ones were unambiguously confirmed by the exemplification of X-ray crystal structural analysis of product 3aa.17Open in a separate windowScheme 1Substrate scope for preparation of spirooxindolyl oxazol-2(5H)-onesa. aReaction conditions: 1 (0.3 mmol), 2a (0.9 mmol), Pd(OAc)2 (5 mol%), bpy (6 mol%), HOAc (5 equiv.) in NMP (1.5 mL) at 80 °C for 36 h. Yields shown are of isolated products. PMB = p-methoxybenzyl; PNB = p-nitrobenzyl.Next, the substrate scope with respect to arylboronic acids was also investigated, the results of which are summarized in Scheme 2. Arylboronic acids bearing both electron-donating (3ab–3ad) and electron-withdrawing substituents (3ae–3ag) at the benzene ring were tolerated, affording the desired products in good to high yields, exception for strong electron-withdrawing substituent such as nitro group (3ah), which did not react with 3-cyano-1-methyl-2-oxoindolin-3-yl ethyl carbonate 1a. It is noteworthy that the reaction also proceeded smoothly when a substituent was situated at the ortho position of the arylboronic acid, albeit with the slightly decreased yield (3ai). As expected, meta- and di-substituted analogues afforded products (3aj and 3ak) in high yields. Additionally, aryl boronic acids with fused ring also gave their corresponding products with high yields. For examples, treatment of both α-naphthyl and β-naphthyl boronic acids with 1a can deliver the corresponding products (3al–3am) in high yields under the optimized reaction conditions, while 9-phenanthreneboronic acid gave spirooxindolyl product 3an in 83% yield. However, hetero-aromatic boronic and alkyl boronic acid did not provided any desired products (3ao–3ap).Open in a separate windowScheme 2Substrate scope with respect to boronic acidsa. aReaction conditions: 1a (0.3 mmol), 2 (0.9 mmol), Pd(OAc)2 (5 mol%), bpy (6 mol%), HOAc (5 equiv.) in NMP (1.5 mL) at 80 °C for 36 h. Yields shown are of isolated products.In addition, besides spirooxindolyl oxazol-2(5H)-one frameworks, this approach is also applicable to the construction of other spirocyclic frameworks incorporating oxazol-2(5H)-one unit via palladium-catalyzed tandem sequence (Scheme 3). For example, treatment of 1-cyanocyclopentyl ethyl carbonate 5a with 2a can furnish 4-phenyl-1-oxa-3-azaspiro[4.4]non-3-en-2-one 6a in 84% yield, while six-membered-ring analogues delivered the corresponding six-membered ring fused spiro-products (6b–6c) in high yields. Cyanohydrin 5d derived from 2-indanone can also serve as a suitable substrate for this tandem sequence, and provided the desired spiro-product 6d in 81% yield.Open in a separate windowScheme 3Preparation of other spirocyclic frameworksa. aReaction conditions: 5 (0.3 mmol), 2a (0.9 mmol), Pd(OAc)2 (5 mol%), bpy (6 mol%), HOAc (5 equiv.) in NMP (1.5 mL) at 80 °C for 36 h. Yields shown are of isolated products.Finally, the synthetic utility of this Pd-catalyzed cyclization was demonstrated (Scheme 4). The reduction of 3aa by using BH3·SMe2 readily gave spirooxindolyl product 7 bearing the oxazolidine unit in good yield with an excellent diastereoselectivity.Open in a separate windowScheme 4Synthetic transformation.On the basis of these results and other processes involving the addition of arylpalladium species to nitrile,14,15 a plausible mechanism was illustrated in Scheme 5. First, the transmetalation of arylboronic acid by Pd(ii) catalyst A generates arylpalladium species B. Then coordination of the nitrile provides intermediate C, which undergoes a carbopalladation of the cyano group to result in formation of the corresponding ketimine Pd(ii) complex D. The intramolecular cyclization of the intermediate D to form palladium complex E which affords product and regenerates the Pd(ii) catalyst.Open in a separate windowScheme 5Proposed mechanism.In summary, we have demonstrated an efficient protocol for the synthesis of spirooxindolyl oxazol-2(5H)-ones via Pd(ii)-catalyzed addition of arylboronic acids to nitriles. A diversity of functionalized spirooxindolyl oxazol-2(5H)-ones can be prepared in good to high yields under the optimal conditions. Furthermore, by the virtue of this Pd-catalyzed sequence, other five- and six-membered ring fused spiro-oxazol-2(5H)-ones can be readily prepared in good yields. Further studies on the application of this synthetic method are currently under investigation.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号