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 328 毫秒
1.
The reactivity of Amberlite (IRA-67) base "heterogeneous" resin in Sonogashira cross-coupling of 8-bromoguanosine 1 with phenylacetylene 3 to give 2 has been examined. Both 1 and 2 coordinate to Pd and Cu ions, which explains why at equivalent catalyst loadings, the homogeneous reaction employing triethylamine base is poor yielding. X-ray photo-electron spectroscopy (XPS) has been used to probe and quantify the active nitrogen base sites of the Amberlite resin, and postreaction Pd and Cu species. The PdCl(2)(PPh(3))(2) precatalyst and CuI cocatalyst degrade to give Amberlite-supported metal nanoparticles (average size ~2.7 nm). The guanosine product 2 formed using the Amberlite Pd/Cu catalyst system is of higher purity than reactions using a homogeneous Pd precatalyst, a prerequisite for use in biological applications.  相似文献   

2.
Four Co(III)-, Cu(II)-, Zn(II)-, and Pd(II)-based potent antibacterial complexes of formula K3[Co(ox)3].3H2O (I), [Cu(bpy)2Cl]Cl.5H2O (II), [Zn(bpy)3]Cl2 (III), and [Pd(bpy)2](NO3)2 (IV) (where ox is oxalate and bpy is 2,2′-bipyridine) were synthesized. They were characterized by elemental analyses, molar conductance measurements, UV–Vis, FTIR, 1H NMR, and 13C NMR spectra. These metal complexes were ordered in three combination series of I + II, I + II + III, and I + II + III + IV. Antibacterial activity was tested for each of these four metal complexes and their combinations against Gram-positive and Gram-negative bacteria. All compounds were more potent antibacterial agents against the Gram-negative than those of the Gram-positive bacteria. The four metal complexes showed antibacterial activity in the order I > II > III > IV and the activity of their combinations followed the order of I + II + III + IV > I + II + III > I + II. CT-DNA binding studies of complex I and its three combinations were carried out using UV–vis spectral titration, displacement of ethidium bromide (EB), and electrophoretic mobility assay. The results obtained from UV–vis studies indicated that all series interact effectively with CT-DNA. Fluorescence titration revealed that the complexes quench DNA-EB strongly through the static quenching procedures. The binding constant (Kb), the Stern–Volmer constant (Ksv), and the number of binding sites (n) were determined at different temperatures of 293, 300, and 310 K, respectively. The calculated thermodynamic parameters supported that hydrogen binding and Van der Waals forces play a major role in association of each series of metal complexes with CT-DNA and follow the above-binding affinity order for the series.  相似文献   

3.
Abstract

In this study, four Co(III)-, Cu(II)-, Zn(II)- and Pd(II)-based potent antibacterial complexes of formula K3[Co(ox)3]·3H2O (I), [Cu(phen)2Cl]Cl·6.5H2O (II), [Zn(phen)3]Cl2 (III) and [Pd(phen)2](NO3)2 (IV) (where ox is oxalato and phen is 1,10-phenanthroline) were synthesized. They were characterized by elemental analysis, molar conductivity measurements, UV–vis, Fourier transform infrared (FT-IR) and proton nuclear magnetic resonance (1H-NMR) techniques. These metal complexes were ordered in three combination series of I+II, I+II+III and I+II+III+IV. Antibacterial screening for each metal complex and their combinations against Gram-positive and Gram-negative bacteria revealed that all compounds were more potent antibacterial agents against the Gram-negative than those of the Gram-positive bacteria. The four metal complexes showed antibacterial activity in the order I > II > III > IV, and the activity of their combinations followed the order of I+II+III+IV > I+II+III > I+II. The DNA-binding properties of complex (I) and its three combinations were studied using electronic absorption and fluorescence (ethidium bromide displacement assay) spectroscopy. The results obtained indicated that all series interact effectively with calf thymus DNA (CT-DNA). The binding constant (Kb), the number of binding sites (n) and the Stern–Volmer constant (Ksv) were obtained based on the results of fluorescence measurements. The calculated thermodynamic parameters supported that hydrogen bonding and van der Waals forces play a major role in the association of each series of metal complexes with CT-DNA and follow the above-binding affinity order for the series.

Communicated by Ramaswamy H. Sarma  相似文献   

4.
A number of 2H-1,3-benzoxazin-4(3H)-one derivatives containing indole or benzofuran moieties were synthesized by using Pd/C–Cu mediated coupling-cyclization strategy as a key step. The o-iodoanilides or o-iodophenol were coupled with 3-{2-(prop-2-ynyloxy)ethyl}-2H-benzo[e][1,3]oxazin-4(3H)-one using 10%Pd/C–CuI–PPh3 as a catalyst system and Et3N as a base to give the target compounds. All the synthesized compounds were tested for their PDE4B inhibitory potential in vitro using a cell based cAMP reporter assay. Some of them showed fold increase of the cAMP level when tested at 30 μM. A representative compound showed encouraging PDE4B inhibitory properties that were supported by its docking results.  相似文献   

5.
In this paper we report a new nanometallic, self-activating catalyst, namely, Ni-supported Pd nanoparticles (PdNPs/Ni) for low temperature ammonia cracking, which was prepared using a novel approach involving the transfer of nanoparticles from the intermediate carrier, i.e. nano-spherical SiO2, to the target carrier technical grade Ni (t-Ni) or high purity Ni (p-Ni) grains. The method that was developed allows a uniform nanoparticle size distribution (4,4±0.8 nm) to be obtained. Unexpectedly, the t-Ni-supported Pd NPs, which seemed to have a surface Ca impurity, appeared to be more active than the Ca-free (p-Ni) system. A comparison of the novel PdNPs/Ni catalyst with these reported in the literature clearly indicates the much better hydrogen productivity of the new system, which seems to be a highly efficient, flexible and durable catalyst for gas-phase heterogeneous ammonia cracking in which the TOF reaches a value of 2615 mmolH2/gPd min (10,570 molNH3/molPd(NP) h) at 600°C under a flow of 12 dm3/h (t-Ni).  相似文献   

6.
A series of palladium complexes of the type [Pd(phPS2)(PAr3)] (phPS2) = [PhP(C6H4-2-S)2]2− have been synthesized in good yields and their crystal structures determined. Heck coupling reactions were carried out using the [Pd(phPS2)(PPh3)] (1), [Pd(phPS2){P(C6H4-4-Cl)3}] (2), [Pd(phPS2){P(C6H4-4-F)3}] (3), [Pd(phPS2){P(C6H4-4-CF3)3}] (4), [Pd(phPS2){P(C6H4-4-Me)3}] (5) and [Pd(phPS2){P(C6H4-4-OMe)3}] (6) complexes as catalyst precursors in order to examine the potential effect of the para-substituted triarylphosphines in the reaction of bromobenzene and styrene.  相似文献   

7.
The recovery of a homogeneous catalysts can be performed effectively by means of a zeolite membrane, which in many cases has a crystalline structure with pores smaller than transition metal catalysts but larger than the products. This concept has been successfully applied to the Diels-Alder reaction between acroleine and cyclopentadiene, catalyzed by a cationic dinuclear Pd(II) complex. A reaction mixture containing 5-norbornene-2-carboxyaldehyde and the catalyst employed for its synthesis [Pd(μ-Cl)(Ph2P-(CH2)4-PPh2)](CF3SO3)2 shows close to 100% recovery of the metallic compound.  相似文献   

8.
An efficient green Heck reaction protocol was performed using a triethanolammonium acetate ionic liquid–palladium(II) catalytic system. The ionic liquid used acts as a reaction medium, base, precatalyst-precursor, and mobile support for the active Pd species. Our experimental investigation indicates that performing the Heck reaction in ionic liquid is superior to the same procedure carried out in triethanolamine. The mechanism of the reaction of triethanolammonium acetate with PdCl2 was examined using density functional theory (M06 method). It was found that two Pd(II) complexes are formed, one of which acts further as a precatalyst yielding catalytically active Pd(0) complex. The calculated activation energies are in agreement with our experimental findings.  相似文献   

9.
Abstract

Herein, we demonstrate that the degrees of catalytic performance of M-CeO2-based catalysts (M=Mn, Cu, Ru or Zr) for an ammonia selective catalytic reduction (NH3-SCR) of nitric-oxide (NO) can be estimated using three theoretical terms; (i) an oxygen vacancy formation energy of a catalyst, (ii) an adsorption energy of NO and (iii) an adsorption energy of NH3. Those terms predict the trend of the catalytic performance as the order; Mn–CeO2 > Cu–CeO2 > Ru–CeO2 > Zr–CeO2 > CeO2. To verify the theoretical prediction, the catalysts were synthesized and tested their performances on the NH3-SCR of NO reaction. The normalized NO conversion rates at low temperatures (100–200 °C) were measured for Mn–CeO2, Cu–CeO2, Ru–CeO2, Zr–CeO2 and CeO2 as 2.61–7.46, 1.30–6.82, 0.73–3.02, 0.81–3.31 and 1.55–2.33 mol s?1 m?2, respectively. In addition, a concept of a structure-activity relationship analysis shows a strong relationship between theoretical and experimental results. Consequently, an application of predicting the catalytic performance of catalysts from theoretical calculations prior the catalyst synthesis is useful in catalyst design and screening that can reduce time and cost.  相似文献   

10.
Novel as well as known C,N-palladacyclic neutral complexes were tested as precatalyst in the Heck reaction between bromobenzene and styrene under aerobic conditions. The catalytic system Pd(II) complex/K2CO3/EGME-H2O (EGME = ethylene glycol monomethyl ether) showed to be highly efficient. Best performances of the catalysts were achieved by controlling the amount of water: generally a H2O content within the 25-50% v/v range resulted in the highest conversion of the substrates into trans-stilbene. As a matter of fact, bromobenzene and styrene can be converted quantitatively using only 0.01 mol% of precatalyst with very high regioselectivity for the trans product and with a TOF of 10 000 h−1. In the absence of water, all complexes were less efficient and differences in their activity were found, while such a differentiation disappeared when water was added.  相似文献   

11.
Two novel copper (II) substituted thiosemicarbazone Schiff base complexes [Cu(L1)(µ-SCN)]n(NO3)2 (1) and [Cu2(µ-SCN)(SCN)(L2)2](NO3) (2) have been synthesized by condensing substituted thiosemicarbazides like 4-methyl-3-thiosemicarbazide or 4-ethyl-3-thiosemicarbazide with 2-acetylpyridine. Both the metal complexes 1 and 2 are characterized using different spectroscopic techniques like IR, UV-Vis, ESR spectroscopy followed by elemental analysis, cyclic voltammetric measurement and single crystal X-ray structure analysis. X-ray crystal structure analysis reveal that complex 1 is polymeric while complex 2 is dimeric in nature. The coordination geometry around Cu(II) are square pyramidal in which thiosemicarbazone Schiff base ligand coordinate to the central Cu(II) atom in tridentate fashion. The prominent interaction patterns of 1 and 2 with CT-DNA were examined by employing electronic absorption and emission spectral titrations, cyclic voltammetry and viscosity measurements. All the results show that CT-DNA binds with both copper (II) complexes 1 and 2. Furthermore, protein binding ability in vitro of complexes 1 and 2 with both BSA and HSA were carried out using multispectroscopic techniques and a static quenching pattern was observed in both cases. Molecular docking study was employed to ascertain the exact mechanism of action of 1 and 2 with DNA and protein molecules (BSA and HSA). In vitro cytotoxicity activity of complexes 1 and 2 toward AGS and A549 was evaluated using MTT assay which demonstrates that both complexes 1 and 2 have superior prospectus to act as anticancer agents.

Communicated by Ramaswamy H. Sarma  相似文献   


12.
Three bimetallic palladium(II) complexes were generated by cyclopalladation of N-methyl-N-phenylbenzamide derivatives, substrates known to undergo oxidative intramolecular cross-coupling via palladium catalysis. These isolable Pd-complexes were characterized by X-ray crystallography. Stoichiometric and catalytic experiments with [(3-methoxy-N-methyl-N-phenylbenzamide)Pd(μ-TFA)]2 were investigated, and this palladium complex was found to be an effective precatalyst for oxidative cross-coupling.  相似文献   

13.
3,5-bis(diphenylphosphino)benzoic acid (LH) reacted with Pd(COD)Cl2 to give [Pd(μ-LH)Cl2]2 (1) in which the Cl ligands are trans. Solventless grinding of 1 and ferric nitrate yielded an insoluble networked Pd-phosphine complex [LPdCl2Fe2/3] (2), in which Fe atom bonds to the carboxylate group of LH. The networked complex 2 has been characterized by using FT-IR, X-ray photoelectron spectroscopy (XPS), scanning electron microscope (SEM), and BET surface area analysis. Complex 2 efficiently catalyzed the Suzuki coupling reactions and could be reused under ambient atmosphere.  相似文献   

14.
Complex (S,S)-[Pd{C6H4(CH2CHNH2CO2CH2CH3)}(μ-Br)]2 (3) was prepared following the method by Vicente and Saura-Llamas (Organometallics 26:2768–2776, 2007), by the reaction of l-ethylphenylalanate and Pd(OAc)2 in 1:1 molar ratio under acetonitrile heating conditions and subsequently treating with NaBr. In addition, the cleavage of halogeno-bridge of the complex 3 via nucleophilic attack of some neutral ligands such as triphenylphosphine, pyridine, 2,4,6-trimethylpyridine and piperidine were investigated and the corresponding complexes (S)-[Pd{C6H4(CH2CHNH2CO2CH2CH3)(Y)(Br)}] (4a–f) were obtained in moderate yields. The six-member orthopalladated complexes were characterized by 1H-NMR, 31P-NMR, FT-IR and elemental analysis techniques.  相似文献   

15.
The bis(2-methoxyethyl)dithiocarbamate complexes [M{S2CN(CH2CH2OMe)2}2] (M = Ni, Cu, Zn, Pd) are readily prepared and the three lighter complexes have been crystallographically characterised. Disproportionation of [Cu{S2CN(CH2CH2OMe)2}2] upon addition of Cu(ClO4)2 · 6H2O affords the copper(III) complex [Cu{S2CN(CH2CH2OMe)2}2][ClO4] which has also been crystallographically characterised. Unlike other copper(III) dithiocarbamate salts, there are no intermolecular cation-cation or cation-anion interactions.  相似文献   

16.
Bacteria can fabricate platinum group metal (PGM) catalysts cheaply, a key consideration of industrial processes and waste decontaminations. Biorecovery of PGMs from wastes is promising but PGM leachates made from metallic scraps are acidic. A two‐step biosynthesis ‘pre‐seeds’ metallic deposits onto bacterial cells benignly; chemical reduction of subsequent metal from acidic solution via the seeds makes bioscaffolded nanoparticles (NPs). Cells of Escherichia coli were seeded using Pd(II) or Pt(IV) and exposed to a mixed Pd(II)/Pt(IV) model solution under H2 to make bimetallic catalyst. Its catalytic activity was assessed in the reduction of Cr(VI), with 2 wt% or 5 wt% preloading of Pd giving the best catalytic activity, while 1 wt% seeds gave a poorer catalyst. Use of Pt seeds gave less effective catalyst in the final bimetallic catalyst, attributed to fewer and larger initial seeds as shown by electron microscopy, which also showed a different pattern of Pd and Pt deposition. Bimetallic catalyst (using cells preloaded with 2 wt% Pd) was used in the hydrogenation of soybean oil which was enhanced by ~fourfold using the bimetallic catalyst made from a model waste solution as compared to 2 wt% Pd preloaded cells alone, with a similar selectivity to cis C18:1 product as found using a Pd‐Al2O3 commercial catalyst.  相似文献   

17.
The syntheses of five new complexes of the 2-formylpyridinethiosemicarbazone ligand (HFpyTSC) with Pd(II) and Rh(III) ions are described, viz., [Pd(FpyTSC)(PPh3)]PF6, [Pd(FpyTSC)(SCN)], [Pd(FpyTSC)Br], [Pd(HFpyTSC)2]Br2 and [Rh(FpyTSC)(PPh3)2Cl]ClO4. The formulation of the complexes is discussed in terms of their elemental analyses and IR, Raman, NMR (1H, 13C and 31P), mass and electronic spectra. The X-ray crystal structures of [Pd(FpyTSC)(PPh3)]PF6 and [Pd(FpyTSC)(SCN)] show that the HFpyTSC ligand coordinates to the central Pd(II) ion in a planar conformation through the pyridyl nitrogen, the azomethine nitrogen and the deprotonated thiol sulphur atom. Thus, HFpyTSC is a versatile ligand that usually acts as a mononegative tridentate ligand bonding through Npy, NCN and C-S while, in the case of [Pd(HFpyTSC)2]Br2, it behaves as a neutral bidentate ligand via NCN and CS.  相似文献   

18.
《Inorganica chimica acta》2004,357(15):4568-4576
The synthesis of palladacyclic derivatives with the hybrid pyridylphosphine ligands Py(CH2)OPPh2 (a) and PyNHPPh2 (b) in a neutral P,N-chelating coordination mode has been achieved. Treatment of selected chloride-bridged cyclometallated precursors [Pd(CN)(μ-Cl)]2 [CN = 2-pyridinin-phenyl Phpy, I-compounds; 7,8-benzoquinolyl Bzq, II-compounds; phenylazophenyl Azb, III-compounds or 2-(2-oxazolinyl)phenyl Phox, IV-compounds] with a or b in the presence of stoichiometric KPF6 gave the mononuclear derivatives Ia-IVa and Ib-IVb. The crystal structures of compounds [Pd(Azb)(Ph2POCH2Py-P,N)][PF6] (IIIa) and [Pd(Phpy)(Ph2PNHPy-P,N)][PF6] (Ib) have been determined. The new palladacyclopentadiene precursor [Pd{C4COOMe4}(CH3CN)2] (V) has been prepared starting from the polymeric complex [Pd{C4COOMe4}]n. Its usefulness in the preparation of new derivatives has been tested by means of the straightforward reaction with ligands (a) or (b) to give mononuclear compounds [Pd{C4(COOMe)4}(Ph2POCH2Py-P,N)] (Va) and [Pd{C4(COOMe)4}(Ph2PNHPy-P,N)] (Vb). The reactions of hydroxo-bridged precursors [Pd(CN)(μ-OH)]2 or [Pd2{C4(COOMe)4}2 (μ-OH)2][NBu4]2 with PyNHPPh2 afforded mononuclear complexes Ic-Vc in which a less common anionic P,N-binding mode is forced as a result of ligand deprotonation. The new complexes were characterised by partial elemental analyses and spectroscopic methods (IR, FAB, 1H and 31P{1H} NMR).  相似文献   

19.
Excessive amounts of oxy-functional groups in unprocessed bio-oil vitiate its quality as fuel; therefore, it has to be channelized to upgrading processes, and catalytic hydrodeoxygenation is one of the most suitable routes for the upgrading of crude bio-oil. In this computational work, catalytic hydrodeoxygenation (HDO) of guaiacol, which is an important phenolic compound of crude bio-oil, has been carried out using density functional theory (DFT) over a Pd(111) catalyst. The Pd(111) catalyst surface does not endorse direct eliminations of functional groups of guaiacol; however, it is found to perform excellently in stepwise dehydrogenation reactions of oxy-functionals of guaiacol according to present DFT results. The catechol product, formed through dehydrogenation of the methoxy group, followed by elimination of CH2 and association of the hydrogen atom, has been identified as one of the major products. The overall reaction rate is controlled by scission of CH2 from 2-methylene-oxy-phenol with an activation energy demand of 23.06 kcal mol–1. Further, the kinetic analysis of each reaction step involved in HDO of guaiacol over the Pd(111) catalyst surface has also been carried out at atmospheric pressure and at a wide range of temperatures from 473 to 673 K, with temperature intervals of 50 K. In the kinetic analysis part, various kinetic parameters, such as forward and reverse reaction rate constants, Arrhenius constants, and equilibrium rate constants, are reported. The kinetic modeling of the dominating reaction steps has revealed that even a lower temperature of 473 K provides a favorable reaction environment; and the temperature increment further improves the reaction favorability.  相似文献   

20.
The hydrogen evolution reaction (HER) on a noble metal surface in alkaline media is more sluggish than that in acidic media due to the limited proton supply. To promote the reaction, it is necessary to transform the alkaline HER mechanism via a multisite catalyst, which has additional water dissociation sites to improve the proton supply to an optimal level. Here, this study reports a top‐down strategy to create a multisite HER catalyst on a nano‐Pd surface and how to further fine‐tune the areal ratio of the water dissociation component to the noble metal surface in core/shell‐structured nanoparticles (NPs). Starting with Pd/Fe3O4 core/shell NPs, electrochemical cycling is used to tune the coverage of iron (oxy)hydroxide on a Pd surface. The alkaline HER activity of the core/sell Pd/FeOx (OH)2?2x NPs exhibits a volcano‐shaped correlation with the surface Fe species coverage. This indicates an optimum coverage level where the rates of both the water dissociation step and the hydrogen formation step are balanced to achieve the highest efficiency. This multisite strategy assigns multiple reaction steps to different catalytic sites, and should also be extendable to other core/shell NPs to optimize their HER activity in alkaline media.  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