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 15 毫秒
1.
Three novel phosphine-free Ru-alkylidenes (7a–7c) have been synthesized and utilized as efficient catalysts for ring closing metathesis (RCM) reaction. Spectroscopic data, i.e. NMR and HRMS, along with single crystal X-ray diffraction analysis, were used to confirm their chemical structures. The tosylated carbenoid 7b showed the highest efficiency in cyclizing different acyclic diene substrates. RCM of various (un)substituted N,N-diallylaniline derivatives and stereoselective RCM of different macromolecular dienes were well tolerated using only a catalytic amount (0.5–2.0 mol%) of the additive catalyst (7b) as compared to the well-known Grubbs (II) and Hoveyda–Grubbs (II) catalysts.

Three novel phosphine-free Ru-alkylidenes (7a–7c) have been synthesized and utilized as efficient catalysts for ring closing metathesis (RCM) reaction.  相似文献   

2.
The rapid oxidation of carbon black (CB) is a major drawback for its use as a catalyst support in polymer electrolyte fuel cells. Here, we synthesize poly[2,2′-(4,4′-bipyridine)-5,5′-bibenzimidazole] (BiPyPBI) as a conducting polymer and use it to functionalize the surface of CB and homogenously anchor platinum metal nanoparticles (Pt-NPs) on a CB surface. The as-prepared materials were confirmed by different spectroscopic techniques, including nuclear magnetic resonance spectroscopy, energy-dispersive X-ray, thermal gravimetric analysis, and scanning-transmittance microscopy. The as-fabricated polymer-based CB catalyst showed an electrochemical surface area (ECSA) of 63.1 cm2 mgPt−1, giving a catalyst utilization efficiency of 74.3%. Notably, the BiPyPBI-based CB catalyst exhibited remarkable catalytic activity towards oxygen reduction reactions. The onset potential and the diffusion-limiting current density reached 0.66 V and 5.35 mA cm−2, respectively. Furthermore, oxidation stability testing showed a loss of only 16% of Pt-ECSA for BiPyPBI-based CB compared to a 36% loss of Pt-ECSA for commercial Pt/CB after 5000 potential cycles. These improvements were related to the synergetic effect between the nitrogen-rich BiPyPBI polymer, which promoted the catalytic activity through the structural nitrogen atoms, and demolished the degradation of CB via the wrapping process.

The rapid oxidation of carbon black (CB) is a major drawback for its use as a catalyst support in polymer electrolyte fuel cells.  相似文献   

3.
2,2′,2′′,4,4′,4′′,6,6′,6′′-Nonanitro-1,1′:3′,1′′-terphenyl (NONA) is currently recognized as an excellent heat-resistant explosive. To improve the atomistic understanding of the thermal decomposition paths of NONA, we performed a series of reactive force field (ReaxFF) molecular dynamics simulations under extreme conditions of temperature and pressure. The results show that two distinct initial decomposition mechanisms are the homolytic cleavage of the C–NO2 bond and nitro–nitrite (NO2 → ONO) isomerization followed by NO fission. Bimolecular and fused ring compounds are found in the subsequent decomposition of NONA. The product identification analysis under finite time steps showed that the gaseous products are CO2, N2, and H2O. The amount of CO2 is energetically more favorable for the system at high temperature or low density. The carbon-containing clusters are a favorable growth pathway at low temperatures, and this process was further demonstrated by the analysis of diffusion coefficients. The increase of the crystal density accelerates the decomposition of NONA judged by the analysis of reaction kinetic parameters and activation barriers. In the endothermic and exothermic stages, a 20% increase in NONA density increases the activation energies by 3.24 and 0.48 kcal mol−1, respectively. The values of activation energies (49.34–49.82 kcal mol−1) agree with the experimental data in the initial decomposition stage.

The bimolecular and fused ring compounds are found in the high-temperature pyrolysis of NONA using ReaxFF molecular dynamics simulations.  相似文献   

4.
Metal complexes used as sensitisers in dye-sensitised solar cells (DSCs) are conventionally constructed using a push–pull strategy with electron-releasing and electron-withdrawing (anchoring) ligands. In a new paradigm we have designed new DπA ligands incorporating diarylaminophenyl donor substituents and phosphonic acid anchoring groups. These new ligands function as organic dyes. For two separate classes of DπA ligands with 2,2′-bipyridine metal-binding domains, the DSCs containing the copper(i) complexes [Cu(DπA)2]+ perform better than the push–pull analogues [Cu(DD)(AA)]+. Furthermore, we have shown for the first time that the complexes [Cu(DπA)2]+ perform better than the organic DπA dye in DSCs. The synthetic studies and the device performances are rationalised with the aid of density functional theory (DFT) and time-dependent DFT (TD-DFT) studies.

Two homoleptic copper(i) complexes with [Cu(DπA)2]+ design have been studied as sensitisers in DSCs and are superior to the DπA ligands and related heteroleptic complexes as dyes.  相似文献   

5.
A periodic mesoporous organosilica (PMO) containing 2,2′-bipyridine groups (BPy-PMO) has been shown to possess a unique pore wall structure in which the 2,2′-bipyridine groups are densely and regularly packed. The surface 2,2′-bipyridine groups can function as chelating ligands for the formation of metal complexes, thus generating molecularly-defined catalytic sites that are exposed on the surface of the material. We here report the construction of a heterogeneous water oxidation photocatalyst by immobilizing several types of tris(2,2′-bipyridine)ruthenium complexes on BPy-PMO where they function as photosensitizers in conjunction with iridium oxide as a catalyst. The Ru complexes produced on BPy-PMO in this work were composed of three bipyridine ligands, including the BPy in the PMO framework and two X2bpy, denoted herein as Ru(X)-BPy-PMO where X is H (2,2′-bipyridine), Me (4,4′-dimethyl-2,2′-bipyridine), t-Bu(4,4′-di-tert-butyl-2,2′-bipyridine) or CO2Me (4,4′-dimethoxycarbonyl-2,2′-bipyridine). Efficient photocatalytic water oxidation was achieved by tuning the photochemical properties of the Ru complexes on the BPy-PMO through the incorporation of electron-donating or electron-withdrawing functionalities. The reaction turnover number based on the amount of the Ru complex was improved to 20, which is higher than values previously obtained from PMO systems acting as water oxidation photocatalysts.

Ruthenium complex photosensitizer fixed on 2,2′-bipyridine bridged-periodic mesoporous organosilica with iridium oxide exhibits an efficient photocatalytic water oxidation.  相似文献   

6.
Recently, carbon nanostructures have attracted interest because of their unique properties and interesting applications. Here, CoC@SiO2-850 (3) and CoC@SiO2-600 (4) cobalt–carbon/silica nanocomposites were prepared by solid-state pyrolysis of anthracene with Co(tph)(2,2′-bipy)·4H2O (1) complex in the presence of silica at 850 and 600 °C, respectively, where 2,2′-bipy is 2,2′-bipyridine and tph is the terephthalate dianion. Moreover, Co(μ-tph)(2,2′-bipy) (2) was isolated and its X-ray structure indicated that cobalt(ii) has a distorted trigonal prismatic coordination geometry. 2 is a metal–organic framework consisting of one-dimensional zigzag chains within a porous grid network. 3 and 4 consist of cobalt(0)/cobalt oxide nanoparticles with a graphitic shell and carbon nanotubes embedded in the silica matrix. They were characterized by scanning electron microscopy (SEM), transmission electron microscopy (TEM), powder X-ray diffraction (XRD), Brunauer–Emmett–Teller (BET), Raman spectroscopy, and X-ray photoelectron spectroscopy (XPS). XPS revealed that the nanocomposites are functionalized with oxygen-containing groups, such as carboxylic acid groups. In addition, the presence of metallic cobalt nanoparticles embedded in graphitized carbon was verified by XRD and TEM. The efficiency of 3 for adsorption of crystal violet (CV) dye was investigated by batch and column experiments. At 25 °C, the Langmuir adsorption capacity of 3 for CV was 214.2 mg g−1 and the fixed-bed column capacity was 36.3 mg g−1. The adsorption data were well fitted by the Freundlich isotherm and pseudo-second-order kinetic model. The adsorption process was spontaneous and endothermic.

A cobalt–carbon@silica nanocomposite was synthesized from a cobalt 2,2′-bipyridine terephthalate complex and its adsorption behavior towards crystal violet dye was tested using batch and column techniques.  相似文献   

7.
Multidentate, soft-Lewis basic, complexant scaffolds have displayed significant potential in the discrete speciation of the minor actinides from the neutron-absorbing lanthanides resident in spent nuclear fuel. Efforts to devise convergent synthetic strategies to targets of interest to improve liquid–liquid separation outcomes continue, but significant challenges to improve solubility in process-relevant diluents to effectively define meaningful structure–activity relationships remain. In the current work, a synthetic method to achieve the challenging 2,2′-bipyridine bond of the bis-1,2,4-triazinyl-2,2′-bipyridine (BTBP) complexant class leveraging a Pd-catalyzed Ullman-type coupling is reported. This convergent strategy improves upon earlier work focused on linear synthetic access to the BTBP complexant moiety. Method optimization, relevant substrate scope and application, as well as a preliminary mechanistic interrogation are reported herein.

Access to functionalized BTBP complexants through a reductive coupling strategy decreases linearity of common synthetic strategies towards these relevant materials for separations.  相似文献   

8.
In this study, a material (DLRMG) was synthesized by modifying Ca2+ and manganite (γ-MnOOH) on red mud granules (RMG), which were the main raw materials derived from industrial alumina. Moreover, a series of experiments were conducted on the adsorption of Fe2+ and Mn2+ in underground water. The prepared samples were analyzed by X-ray diffraction (XRD), thermogravimetric analysis-differential thermal analysis (TG-DTA), zeta potential analysis, BET and scanning electron microscopy (SEM); the concentration of the effluent was found to be of acceptable standard after the treatment. DLRMG continued to treat fluoride wastewater even after the saturated adsorption of Fe2+ and Mn2+, and the results clearly showed that the treatment was effective. Overall, the problems of red mud stockpile and pollution in China would be effectively controlled by DLRMG.

The use of the waste of aluminum industry to prepare effective polluted materials for the treatment of underground water.  相似文献   

9.
A series of ruthenium(ii) complexes with N-heterocyclic carbene ligands were successfully synthesized by transmetalation reactions between silver(i) N-heterocyclic carbene complexes and [RuCl2(p-cymene)]2 in dichloromethane under Ar conditions. All new compounds were characterized by spectroscopic and analytical methods. These ruthenium(ii)–NHC complexes were found to be efficient precatalysts for the transfer hydrogenation of ketones by using 2-propanol as the hydrogen source in the presence of KOH as a co-catalyst. The antibacterial activity of ruthenium N-heterocyclic carbene complexes 3a–f was measured by disc diffusion method against Gram positive and Gram-negative bacteria. Compounds 3d exhibited potential antibacterial activity against five bacterial species among the six used as indicator cells. The product 3e inhibits the growth of all the six tested microorganisms. Moreover, the antioxidant activity determination of these complexes 3a–f, using 2,2-diphenyl-1-picrylhydrazyl (DPPH) and 2,2′-azinobis-3-ethylbenzothiazoline-6-sulphonic acid (ABTS) as reagent, showed that compounds 3b and 3d possess DPPH and ABTS antiradical activities. From a concentration of 1 mg ml−1, these two complexes presented a similar scavenging activity to that of the two used controls gallic acid (GA) and butylated hydroxytoluene (BHT). From a concentration of 10 mg ml−1, the percentage inhibition of complexes 3b and 3d was respectively 70% and 90%. In addition, these two Ru–NHC complexes exhibited antifungal activity against Candida albicans. Investigation of the anti-acetylcholinesterase activity of the studied complexes showed that compounds 3a, 3b, 3d and 3e exhibited good activity at 100 μg ml−1 and product 3d is the most active. In a cytotoxicity study the complexes 3 were evaluated against two human cancer cell lines MDA-MB-231 and MCF-7. Both 3d and 3e complexes were found to be active against the tested cell lines showing comparable activity with examples in the literature.

A series of ruthenium(ii) complexes with N-heterocyclic carbene ligands were successfully synthesized by transmetalation reactions between silver(i) N-heterocyclic carbene complexes and [RuCl2(p-cymene)]2 in dichloromethane under Ar conditions.  相似文献   

10.
3′-N-(2-Thio-1,3,2-oxathiaphospholane) derivatives of 5′-O-DMT-3′-amino-2′,3′-dideoxy-ribonucleosides (NOTP-N), that bear a 4,4-unsubstituted, 4,4-dimethyl, or 4,4-pentamethylene substituted oxathiaphospholane ring, were synthesized. Within these three series, NOTP-N differed by canonical nucleobases (i.e., AdeBz, CytBz, GuaiBu, or Thy). The monomers were chromatographically separated into P-diastereomers, which were further used to prepare NNPSN′ dinucleotides (3), as well as short P-stereodefined oligo(deoxyribonucleoside N3′→O5′ phosphoramidothioate)s (NPS-) and chimeric NPS/PO- and NPS/PS-oligomers. The condensation reaction for NOTP-N monomers was found to be 5–6 times slower than the analogous OTP derivatives. When the 5′-end nucleoside of a growing oligomer adopts a C3′-endo conformation, a conformational ‘clash’ with the incoming NOTP-N monomer takes place, which is a main factor decreasing the repetitive yield of chain elongation. Although both isomers of NNPSN′ were digested by the HINT1 phosphoramidase enzyme, the isomers hydrolyzed at a faster rate were tentatively assigned the RP absolute configuration. This assignment is supported by X-ray analysis of the protected dinucleotide DMTdGiBuNPSMeTOAc, which is P-stereoequivalent to the hydrolyzed faster P-diastereomer of dGNPST.

Separated P-diastereomers of 3′-N-(2-thio-1,3,2-oxathiaphospholane) derivatives of 5′-O-DMT-3′-amino-2′,3′-dideoxy-ribonucleosides were used to prepare P-stereodefined NNPSN′ dinucleotides and short NPS-, NPS/PO- and NPS/PS-oligomers.  相似文献   

11.
In the present contribution, new binuclear ternary complexes; [M2(bpy)4L](ClO4)4 (M = Co(ii) (1) and Ni(ii) (2); bpy = 2,2′-bipyridine; L = 1,1′-(hexane-1,6-diyl)bis[2-(pyridin-2-yl)1H-benzimidazole] and [Cu2(bpy)2(OH2)2L](BF4)4 (3) were synthesized, characterized and screened for their antimicrobial activity and cytotoxicity against human liver carcinoma cells (HepG-2) as well as non-malignant human embryonic kidney cells (HEK-293). The structural studies were complemented by density functional theory (DFT) calculations. DNA binding of 1–3 was spectrophotometrically studied. The DNA cleavage ability of 1–3 towards the supercoiled plasmid DNA (pBR322 DNA) was examined through gel electrophoresis. Compound 3 has the highest cytotoxic activity (IC50 = 3.5 μg mL−1) against HepG-2 among the investigated complexes and is non cytotoxic to noncancerous HEK-293. Complexes (1 and 2) exhibited toxicity to HEK-293 with IC50 values of 30.3 and 23.5 μg mL−1 in that order. While compound 1 showed antifungal activity against Cryptococcus neoformans, complex 2 exhibited its toxicity against Candida albicans.

The binuclear ternary transition metal complexes have the ability to bind to lysozyme and DNA as well as DNA cleavage. Complexes exhibited cytotoxicity to only the cancerous cells (HepG-2) and is safe to the non-malignant HEK-293.  相似文献   

12.
While sulfadiazine (HLSZ) is extensively used to elaborate complexes of intriguing biological applications (e.g. topical antibiotic silvadene; silver sulfadiazine), the molecular structure modification of sulfadiazine or even other sulfa drugs by coordination to either η6-cymene Ru(ii) or η5-Cp* Rh(iii) motif has not been investigated. Here, half-sandwich organoruthenium(ii) and organorhodium(iii) compounds of the type [(η6-p-cymene)Ru(LSZ)2] (1) and [(η5-C5Me5)Rh(LSZ)2] (2) are synthesized, characterized and evaluated for their potential antimicrobial activity. Spectroscopic and single crystal X-ray analysis showed that LSZ is coordinated to Rh(iii) via both the sulfonamide and pyrimidine nitrogen atoms forming “piano-stool” geometry. In 2, the NMR equivalence clearly pointed to participation of two LSZ molecules in a fluxional process in which the third bond of the base of the stool is oscillating between two equivalent sulfonamide nitrogen atoms. While 1 was biologically inactive, complex 2 was potent against Gram-positive bacteria, Candida albicans and Cryptococcus neoformans. Hen white egg lysozyme (HEWL), a model protein, reacted covalently with 2via the loss of one LSZ molecule, while compound 1 decomposed during the interaction with that protein.

Binding of certain metal complexes to proteins may cause cytotoxicity. While [(η6-p-Cym)Ru(LSZ)2] decomposed during the reaction with hen white egg lysozyme, a Rh(iii) analogue was covalently bio-conjugated via the elimination of a sulfadiazine.  相似文献   

13.
Star-shaped 2,4,6-tris(4′,4′′,4′′′-trimethylphenyl)-1,3,5-triazine molecules self-assemble at the solid–liquid interface into a compact hexagonal nanoarchitecture on graphite. High resolution scanning tunneling microscopy (STM) images of the molecules reveal intramolecular features. Comparison of the experimental data with calculated molecular charge density contours shows that the molecular features in the STM images correspond to molecular LUMO+2.

Intramolecular contrast in the STM images of 2,4,6-tris(4′,4′′,4′′′-trimethylphenyl)-1,3,5-triazine molecules recorded at room-temperature and at the liquid–solid interface.  相似文献   

14.
The reaction of Ph2PCH2OH with PhPCl2 and PCl3 in the presence of Et3N afforded new phosphonite compounds PhP(OCH2PPh2)21 and P(OCH2PPh2)32, respectively. The reaction between 1 and [NiCl2(DME)] in dichloromethane gave the five-coordinate complex [NiCl2(1-κ3P,P,P)] 3. Conversely, 1 reacts with [NiCl2(DME)] in the presence of NH4PF6 in dichloromethane to yield the four coordinate ionic complex [NiCl(1-κ3P,P,P)][PF6] 4. The reactions between 1, [NiCl2(DME)] and KPF6 in the presence of RNC (R = Xylyl, tBu and iPr) in dichloromethane yielded the five coordinate monocationic [NiCl(1-κ3P,P,P)(RNC)][PF6] (R = Xylyl) and dicationic [Ni(1-κ3P,P,P)(RNC)2][PF6]2 (R = tBu and iPr) complexes, respectively. The analogous reaction of 2 with [NiCl2(DME)] in the presence of KPF6 gave complex [NiCl(2-κ4P,P,P,P)][PF6], 8. The structures of all complexes were determined by single crystal X-ray diffraction studies and supported by spectroscopic methods. To demonstrate their catalytic application, N-alkylation reactions between primary aryl amines, benzyl and 4-methoxy benzyl alcohols were found to proceed smoothly in the presence of 2.5 mol% of complexes bearing ligand 1 and <0.5 mmol of KOBut in toluene at 140 °C. The C–N coupled products were formed in very good yields. Its substrate scope includes sterically encumbered, heterocyclic amines and aliphatic alcohol.

A series of nickel(ii) complexes supported by the new tridentate P3 and tetradentate P4 ligands act efficiently as catalysts for the N-alkylation of primary amines with alcohols.  相似文献   

15.
Microdialysis was applied to sample the unbound drug concentration in the extracellular fluid in brain and muscle of rats given zalcitabine (2′,3′-dideoxycytidine; n = 4) or BEA005 (2′,3′-dideoxy-3′-hydroxymethylcytidine; n = 4) (50 mg/kg of body weight given subcutaneously). Zalcitabine and BEA005 were analyzed by high-pressure liquid chromatography with UV detection. The maximum concentration of zalcitabine in the dialysate (Cmax) was 31.4 ± 5.1 μM (mean ± standard error of the mean) for the brain and 238.3 ± 48.1 μM for muscle. The time to Cmax was found to be from 30 to 45 min for the brain and from 15 to 30 min for muscle. Zalcitabine was eliminated from the brain and muscle with half-lives 1.28 ± 0.64 and 0.85 ± 0.13 h, respectively. The ratio of the area under the concentration-time curve (AUC) (from 0 to 180 min) for the brain and the AUC for muscle (AUC ratio) was 0.191 ± 0.037. The concentrations of BEA005 attained in the brain and muscle were lower than those of zalcitabine, with Cmaxs of 5.7 ± 1.4 μM in the brain and 61.3 ± 12.0 μM in the muscle. The peak concentration in the brain was attained 50 to 70 min after injection, and that in muscle was achieved 30 to 50 min after injection. The half-lives of BEA005 in the brain and muscle were 5.51 ± 1.45 and 0.64 ± 0.06 h, respectively. The AUC ratio (from 0 to 180 min) between brain and muscle was 0.162 ± 0.026. The log octanol/water partition coefficients were found to be −1.19 ± 0.04 and −1.47 ± 0.01 for zalcitabine and BEA005, respectively. The degrees of plasma protein binding of zalcitabine (11% ± 4%) and BEA005 (18% ± 2%) were measured by microdialysis in vitro. The differences between zalcitabine and BEA005 with respect to the AUC ratio (P = 0.481), half-life in muscle (P = 0.279), and level of protein binding (P = 0.174) were not statistically significant. The differences were statistically significant in the case of the half-life in the brain (P = 0.032), clearance (P = 0.046), volume of distribution (P = 0.027) in muscle, and octanol/water partition coefficient (P = 0.019).  相似文献   

16.
To develop the biomimetic chemistry of [NiFe]-H2ases, the first azadithiolato-bridged NiFe model complexes [CpNi{(μ-SCH2)2NR}Fe(CO)(diphos)]BF4 (5, R = Ph, diphos = dppv; 6, 4-ClC6H4, dppv; 7, 4-MeC6H4, dppv; 8, CO2CH2Ph, dppe; 9, H, dppe) have been synthesized via well-designed synthetic routes. Thus, treatment of RN[CH2S(O)CMe]2 with t-BuONa followed by reaction of the resulting intermediates RN(CH2SNa)2 with (dppv)Fe(CO)2Cl2 or (dppe)Fe(CO)2Cl2 gave the N-substituted azadithiolato-chelated Fe complexes [RN(CH2S)2]Fe(CO)2(diphos) (1, R = Ph, diphos = dppv; 2, 4-ClC6H4, dppv; 3, 4-MeC6H4, dppv; 4, CO2CH2Ph, dppe). Further treatment of 1–4 with nickelocene in the presence of HBF4·Et2O afforded the corresponding N-substituted azadithiolato-bridged NiFe model complexes 5–8, while treatment of 8 with HBF4·Et2O resulted in formation of the parent azadithiolato-bridged model complex 9. While all the new complexes 1–9 were characterized by elemental analysis and spectroscopy, the molecular structures of model complexes 6–8 were confirmed by X-ray crystallographic study. In addition, model complexes 7 and 9 were found to be catalysts for H2 production with moderate icat/ip and overpotential values from TFA under CV conditions.

The first azadithiolato-bridged NiFe model complexes with a general formula [CpNi{(μ-SCH2)2NR}Fe(CO)(diphos)]BF4 have been synthesized, characterized, and for some of them found to be catalysts for proton reduction to H2 under CV conditions.  相似文献   

17.
The TW17 ribozyme, a catalytic RNA selected from a pool of artificial RNA, is specific for the Zn2+-dependent hydrolysis of a phosphorothiolate thiolester bond. Here, we describe the organic synthesis of both guanosine α-thio-monophosphate and the substrates required for selecting and characterizing the TW17 ribozyme, and for deciphering the catalytic mechanism of the ribozyme. By successively substituting the substrate originally conjugated to the RNA pool with structurally modified substrates, we demonstrated that the TW17 ribozyme specifically catalyzes phosphorothiolate thiolester hydrolysis. Metal titration studies of TW17 ribozyme catalysis in the presence of Zn2+ alone, Zn2+ and Mg2+, and Zn2+ and [Co(NH3)6]3+ supported our findings that Zn2+ is absolutely required for ribozyme catalysis, and indicated that optimal ribozyme catalysis involves the presence of outer-sphere and one inner-sphere Mg2+. A survey of the TW17 ribozyme activity at various pHs revealed that the activity of the ribozyme critically depends on the alkaline conditions. Moreover, a GNRA tetraloop-containing ribozyme constructed with active catalysis in trans provided catalysis and multiple substrate turnover efficiencies significantly higher than ribozymes lacking a GNRA tetraloop. This research supports the essential roles of Zn2+, Mg2+, and a GNRA tetraloop in modulating the TW17 ribozyme structure for optimal ribozyme catalysis, leading also to the formulation of a proposed reaction mechanism for TW17 ribozyme catalysis.

Zn(ii) and Mg(ii) and GAGA tetraloop in the ion atmosphere of the TW17 ribozyme is critical to optimal ribozyme catalysis at alkaline pH.  相似文献   

18.
Two amide–imine conjugates, viz. 3-methyl-benzoic acid (4-diethylamino-2-hydroxy-benzylidene)-hydrazide (L1) and 3-methyl-benzoic acid (2-hydroxy-naphthalen-1-ylmethylene)-hydrazide (L2), have been prepared and used for a further synthesis of Mo(vi) complexes (M1 and M2, respectively). Single crystal X-ray diffraction analysis confirmed their structures. Interestingly, M1 selectively recognizes Y3+ and Pb2+ at two different wavelengths, whereas M2 selectively interacts with Y3+ with a significantly high binding constant, 1.3 × 105 M−1. The proposed sensing mechanism involves the displacement of Mo(vi) by Y3+/Pb2+ from respective Mo(vi) complexes. The TCSPC experiment also substantiates the “turn-on” fluorescence process. A logic gate has been constructed utilizing the fluorescence recognition of cations by M1. DFT studies corroborated the cation–probe interactions and allowed exploring the orbital energy parameters.

Two amide–imine conjugates and their Mo(vi) complexes (M1 and M2) were characterized by single crystal X-ray diffraction analysis. While M1 recognises both Y3+ and Pb2+ at two different wavelengths, M2 recognises only Y3+ with significantly high binding constant.  相似文献   

19.
An in-depth theoretical study on the Pt(ii)/Pt(iv)–bisphenylpyridinylmethane complexes was carried out, which focused on the geometric/electronic structures, excitation procedures, on–off phosphorescence mechanisms, and structure–optical performance relationships. The key roles of the linkages (LK) connected in the middle of phenylpyridines were carefully investigated using multiple wavefunction analysis methods, such as non-covalent interaction (NCI) visualizations and natural bond orbital (NBO) studies. The phosphorescence-off phenomenon was considered by hole–electron analysis and visualizations, spin–orbit coupling (SOC) studies, and NBO analysis. Through these investigations, the relationship of the substituents in LK and the optical performances were revealed, as well as the fundamental principles of the phosphorescence-quenching mechanism in Pt(iv) complexes, which pave the way for further performance/structural renovation works. In addition, an intuitive visualization method was developed using a heatmap to quantitatively express the SOC matrix elementary (SOCME), which is helpful for big data simplification for phosphorescence analysis.

An in-depth theoretical study on the Pt(ii)/Pt(iv)–bisphenylpyridinylmethane complexes was carried out, which focused on the structures, excitation procedures, on–off phosphorescence mechanisms, and structure–optical performance relationships.  相似文献   

20.
Divergent and versatile synthetic routes to flavones and flavanones via efficient Pd(ii) catalysis are disclosed. These Pd(ii) catalyses expediently provide a variety of flavones and flavanones from 2′-hydroxydihydrochalcones as common intermediates, depending on oxidants and additives, via discriminate oxidative cyclization sequences involving dehydrogenation, respectively, in a highly atom-economic manner.

Divergent and versatile synthetic routes to flavones and flavanones via efficient Pd(ii) catalysis are disclosed.

Flavones and flavanones are widely occurring natural flavonoids produced by various medicinal plants and their synthetic derivatives, featuring a main 15-carbon skeleton possessing 2 phenyl rings and 1 oxacycle.1 In recent decades, they have been considered privileged structures exhibiting various biological activities,2 such as anti-inflammatory,3 anti-cancer,4 neuroprotective,5 and estrogen-related functions6 (Fig. 1). As routes towards such privileged flavonoids, many strategies, including the Allan–Robinson reaction for flavones and intramolecular conjugate addition of 2′-hydroxychalcones for flavanones, have been reported.7 Among these, intramolecular cyclization of 2′-hydroxychalcone intermediates was conventionally used for the synthesis of flavanones due to the readily available properties of the substrates through the condensation of 2′-hydroxy-acetophenones and corresponding aldehydes.8 In addition, such flavanones can be readily converted into flavones via further oxidation processes (Scheme 1).9 Thus, these types of flavonoid synthesis involving intramolecular conjugate addition have been generally used for flavonoid synthesis,10 but they generally have drawbacks of requiring harsh conditions such as acidic or basic reflux conditions for cyclization, indicating that chemically labile compounds may in some cases not be tolerable in reaction conditions.11 Additional oxidation steps in which flavanones are transformed into flavones often require the use of strong oxidants such as I2 that enable side reactions to occur.12 In this regard, novel synthetic routes featuring transformative and compatible reaction conditions toward flavonoids have been pursued in recent decades. In particular, given the increasing importance of constructing a privileged chemical library for probing biological systems, efficient and divergent synthesis of flavonoids is necessary.13Open in a separate windowFig. 1Examples of bioactive flavonoids.Open in a separate windowScheme 1Synthetic strategy for flavonoids pursued in this study.Taken together, it is important to develop a novel and versatile synthetic transformation of common substrates into flavones and flavanones under mild conditions and such a transformation is also expected to enable straightforward reactions for the construction of a privileged flavonoid library. As mentioned above, 2′-hydroxychalcones have been conventionally used as main precursors in flavonoid synthesis, but they are reactive and unmanageable in some cases due to their reactive α,β-unsaturated carbonyl and enol ether moieties. In addition, their use is generally involved with several disadvantages, as mentioned above.14 Thus, we think that the use of common intermediates instead of 2′-hydroxychalcones that is traditionally used but somewhat problematic is required. Recently, direct β-functionalization of simple ketones via several catalysts using transition metals and light was reported,15 indicating that simple ketones can be useful and potential surrogates of α,β-unsaturated ketones for further oxidative catalysis. In addition, we reported that chromanones, simple ketones, can be converted into flavones and flavanones via palladium(ii)-catalyzed dehydrogenation-mediated coupling sequences with arylboronic esters and acids, respectively, and also recently reported that N-substituted azaflavanone can be synthesized from N-substituted aminodihydrochalcones via Pd(ii)-catalyzed oxidative cyclization.16 Inspired by these previous works, we sought to find novel and divergent synthetic routes to a variety of privileged flavonoids, particularly flavones and flavanones using common starting materials.Herein, we report palladium(ii)-catalyzed oxidative cyclization of 2′-hydroxydihydroxychalcones that are novel and tolerable isosteres of 2′-hydroxychalcones as an efficient and divergent route for the synthesis of flavones and flavanones. To find an optimal reaction condition where flavone and flavanone were synthesized via Pd(ii)-catalyzed oxidative cyclization from the common intermediate in independent manners, simple 2′-hydroxydihydrochalcone was selected as the model compound to screen the reaction conditions (ii)-catalyzed dehydrogenation to occur with Pd(TFA)2 and DMSO under an O2 atmosphere, which is eco-friendly oxidant.17 The use of this reaction condition resulted in the formation of flavanone 4a as the major product (31% yield, entry 1) and flavone 3a with 12% yield, along with a small amount of 2′-hydroxychalcone 2a (14%). This result implied that 2′-hydroxydihydrochalcone could be dehydrogenated and slightly transformed into desired flavonoids under the reaction conditions featuring Pd(ii) catalysis. Based on the results, we tried to optimize the reaction conditions by using additives such as heterocyclic amines or inorganic bases as Pd(ii) ligands. In the presence of most of the heterocyclic amines and K2CO3, the overall yields of the reactions for the synthesis of flavonoids were significantly increased (entries 2–5). Interestingly, the catalytic system featuring 2,2′-bipyridine (bpy) as a ligand provided flavone 3a in 55% yield as a major product, along with flavanone 4a in 10% yield, indicating that bidentate amine was better than the other monodentate amines, including pyridine as a Pd(ii) ligand, under incorporating conditions (entry 6). Notably, the use of 5-nitro-1,10-phenanthroline as a ligand enables the reaction to exclusively yield flavone 3a with the highest yield of 81% and flavanone 4a in 2% yield (entry 7).Optimization of the reaction conditionsa
EntryAdditiveOxidantYieldb [%]
2a3a4a
1O2141231
2K2CO3O282820
3DMAPO262936
4PyridineO254137
5PyrimidineO274832
6bpyO255510
75-NO2 phenO23812
8Benzoquinone7112
9K2S2O811419
10AgOAc121329
11Cu(OAc)233644
12PhenanthrolineCu(OAc)2232734
13PyridineCu(OAc)2141333
14DMAPCu(OAc)2252235
15K2CO3Cu(OAc)27116
16AcOHCu(OAc)2201142
17HCO2HCu(OAc)29643
18 p-TsOHCu(OAc)2152127
19cCu(OAc)230355
20c,dCu(OAc)281079
21c,eCu(OAc)216
Open in a separate windowaReactions were carried out in the presence of 0.3 mmol of 1a, 10 mol% Pd(TFA)2, 20 mol% additive, molecular oxygen or 1.0 equiv. oxidant and DMSO 1 mL at 100 °C for 15–48 h.bIsolated yield.cDMSO 3 mL.dAddition of 2 N HCl 20 mL and ethyl acetate 10 mL for 24 h after 1a was consumed.eCu(OAc)2 2 equiv.Next, we focused on optimizing the conditions where flavanone, rather than flavone, is converted from 2′-hydroxydihydrochalcone that is a common starting material. In the view of the oxidation process, we supposed that the use of a stoichiometric amount of oxidant rather than molecular oxygen can result in a stepwise conversion of 2′-hydroxydihydrochalcone into 2′-hydroxychalcone, leading to the flavanone through the concurrent intramolecular 1,4-addition, rather than to the flavone for which an additional oxidation process may be needed. With this insight, we tested stoichiometric amounts of various oxidants instead of molecular oxygen for obtaining flavanone exclusively based on the entry 1 conditions (entries 8–11). As anticipated, in most of the conditions, flavanone was synthesized as a major product, but the yields were generally poor. When Cu(OAc)2 was used as the oxidant, the yield of flavanone was increased to 44%, along with the formation of flavone at a yield of 6% (entry 11). In addition, we screened various bases and acids as additives or potential ligands for improving the reaction under the conditions where Cu(OAc)2 was used as the oxidant.Heterocyclic ligands, including monodentates and bidentates, slightly decrease the yields of the reaction for providing flavanones (entries 12–14). This trend was also observed for the addition of potassium carbonate or acids (entries 15–18). Thus, unlike for flavone synthesis, additives did not significantly improve the reaction efficiency. With the entry 11 condition (no additive), the concentration was diluted to 0.1 M for facilitating intramolecular cyclization, and the reaction afforded flavanone 4a in 55% yield (entry 19). Next, with the conditions of entry 19, an additional 2 N HCl was added to cyclize the remaining chalcone to flavanone under acidic conditions when 2′-hydroxydihydrochalcone 1a was completely consumed. Under these conditions, flavanone 4a was successfully obtained with the highest yield of 79% (entry 20). The use of 2 equivalents of Cu(OAc)2 generated flavone 3a as a detectable major product in lower yield (16%) than expected, indicating that excess Cu(OAc)2 is less effective than molecular oxygen for flavonoid synthesis.With the optimized conditions in hand (Scheme 2). A variety of aryl A ring derivatives were successfully synthesized under these conditions. The reaction was well-tolerated and proceeded with the 2′-hydroxydihydrochalcone derivatives with electron-withdrawing groups such as halogens, triflate, pivalate and nitro groups (3b–f, 3h, and 3v–y), regardless of the A or B ring. In the case of using the substrates possessing electron-donating substituents such as methyl (3g), hydroxy (3i, 3o, and 3p), methoxy (3j, 3l and 3n), benzyloxy (3k), dimethoxy (3 m, 3q and 3r) and trimethoxy groups (3s and 3t), the reactions progressed well in moderate to good yields. In particular, the highest yield of 95% was obtained for 3′-hydroxyflavone 3p. 2-(Naphthalen-2-yl)-4H-chromen-4-one 3u was formed in 72% yield. However, 3-methylchromenone 3z was obtained from a corresponding 2′-hydroxydihydrochalcone in a relatively lower yield (30%). The substrates with heterocycles such as N-methyl indol-3-yl, thiophen-2-yl, and furan-2-yl, as B ring, were also converted to corresponding flavones 3aa, 3ab, and 3ac in 32%, 54%, and 15% yields.Open in a separate windowScheme 2Reaction scope of the flavone synthesis. Reaction conditions: 1 (1.0 equiv.), Pd(TFA)2 (10 mol%), 5-NO2-1,10-phenanthroline (20 mol%) and DMSO (0.3 M) at 100 °C under O2 for 48 h. a Yield determined by 1H NMR analysis. b 1.2 g scale reaction.Next, under the optimized conditions (Scheme 3). The desired flavanone products possessing electron-deficient substituents such as halogens and ester groups (4b–c, and 4o) as well as electron-rich substituents such as hydroxy and alkoxy moieties regardless of the A or B ring (4d–m) were readily obtained in moderate to good yields. In particular, flavanones with naturally abundant oxygen-containing moieties (4e and 4g–m) were generated effectively in the reaction. Flavanone 4n with a naphthyl moiety was also formed well in 72% yield. A substrate containing a thiophene-2-yl group as B ring was also converted to flavanone 4r in 29% yield. 3-Methylchromanone 4q was obtained from a corresponding 2′-hydroxydihydrochalcone in a lower yield (18%) compared to the other flavanones, which is similar to the case of 3-methylchromenone 3z. In the case of reactions for preparing 3z and 4q, lots of 1-(2-hydroxyphenyl)butan-1-one, a starting material, remained not to be reacted. A few flavones such as 3ac and flavanones such as 4p were not obtained in good yield, along with several side products. Collectively, these results indicated that 2′-hydroxydihydrochalcones that are chemically compatible substrates can be obviously transformed into flavones and flavanones in divergent and efficient manners, respectively.Open in a separate windowScheme 3Reaction scope of the flavanone synthesis. Reaction conditions: 1 (1.0 equiv.), Pd(TFA)2 (10 mol%), Cu(OAc)2 (1.0 equiv.) and DMSO (0.1 M) at 100 °C under Ar for 15 h; then addition of 2 N HCl 20 mL and ethyl acetate 10 mL for 24 h at 100 °C. a No addition of 2 N HCl and ethyl acetate. b 1.0 g scale reaction.To further investigate the divergent utility of our synthetic methodology, we applied it to the synthesis of natural flavonoids (Scheme 4). Under the optimal conditions, geraldone dimethyl ether 5a (ref. 18) and butin trimethyl ether 5b,19 the biologically active natural flavonoids, were successfully transformed from common substrate 1aa in moderate yields (88% and 48%, respectively, Scheme 4). Tithonine 5c, an anti-inflammatory flavone known as a selective COX (cyclooxygenase)-1 inhibitor,20 was also synthesized well under our reaction conditions.Open in a separate windowScheme 4Application for the synthesis of natural products.Next, to investigate the mechanism of our synthetic methodology, we carried out kinetic analysis of the reaction using high-pressure liquid chromatography (HPLC) to examine the time-dependent conversion of 2′-hydroxy-4-methoxydihydrochalcone 1n to the corresponding flavone 3n or flavanone 4h under optimized reaction conditions. For flavone synthesis, it was observed that the amount of 2′-hydroxy-4- methoxydihydrochalcone 1n gradually decreased over time, and 2′-hydroxy-4-methoxychalcone 2n was increasingly formed in a dramatic manner with the concurrent formation of flavone 3n within 5 h (Scheme 5a). Then, 2n was gradually diminished, and 3n was significantly formed, indicating that 2n may be consumed for the synthesis of the desired flavone 3n. In addition, in the conditions of flavanone synthesis (Scheme 5b), it was observed that the reaction proceeded rapidly at first (Fig. S1), resulting in the drastic formation of 2′-hydroxy-4-methoxychalcone 2n with the concurrent formation of flavonoids. After 15 h, the reaction was observed to reach the plateau state. Upon adding 2 N aqueous HCl, flavanone 4h was immediately formed with a significant disappearance of 2n, indicating that 2n may also be consumed for the synthesis of the desired flavanone 4h.Open in a separate windowScheme 5Kinetic experiments for flavone and flavanone synthesis from 2′-hydroxy-4-methoxydihydrochalcone 1n. (a) Flavone synthesis condition. (b) Flavanone synthesis condition.Based on the kinetic results, a plausible mechanism for this divergent synthesis is proposed, as shown in Fig. 2. First, Pd(ii)-catalyzed dehydrogenation may result in the oxidative conversion of 2′-hydroxydihydrochalcone 1a into 2′-hydroxychalcone 2a.16 Then, common intermediate 2a can be transformed into the desired flavanone 4a through Michael addition and flavone 3a through oxidative Heck coupling, respectively, depending on the optimized reaction conditions that are used in the synthesis. During oxidation processes such as dehydrogenation and oxidative cyclization, Pd(0) species regenerated in the reaction would be reoxidized into Pd(ii) via the [O] process, where O2 or Cu(OAc)2 acts as the main oxidant.21Open in a separate windowFig. 2Plausible mechanism.  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