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 31 毫秒
1.
 Dipeptides and tripeptides AcMet-aaH containing N-acetyl methionine, in which the group aaH is GlyH, AlaH, ValH, or Gly-GlyH, undergo hydrolytic cleavage of the Met-aaH peptide bond in the presence of the following complexes of palladium(II): cis-[Pd(en)(H2O)2]2+, cis-[Pd(tn)(H2O)2]2+, cis-[Pd(en)(CH3OH)2]2+, cis-[Pd(S,N-MetH)(H2O)2]2+, cis-[Pd(S,N-Met-GlyH)(H2O)2]2+, and cis-[Pd(S,N-Met-AlaH)(H2O)2]2+. These mononuclear complexes are precursors of binuclear palladium(II) complexes containing the substrates AcMet-aaH as bridging thioether ligands. The rate constant for cleavage is higher when the bidentate ligand in the precursor complex is ethylenediamine (which is completely displaced) than S,N-methionine (of which only the amino group is displaced), because the number of aqua ligands available for cleavage is greater in the former than in the latter case. The demonstrated dependence of the rate constant on the steric bulk (volume) of the leaving group, aaH, points the way toward achieving a degree of sequence selectivity in cleavage of peptide bonds by palladium(II) aqua complexes. One equivalent of cis-[Pd(en)(H2O)2]2+ cleaves as many as ten equivalents of AcMet-GlyH, but the rate constant decreases as the molar excess of the dipeptide over the catalyst increases. This demonstration of catalytic turnover points the way to our ultimate goal – artificial metallopeptidases. Received: 13 June 1997 / Accepted: 24 September 1997  相似文献   

2.
The increasing demand and limited natural resources for industrially important platinum‐group metal (PGM) catalysts render the recovery from secondary sources such as industrial waste economically interesting. In the process of palladium (Pd) recovery, microorganisms have revealed a strong potential. Hitherto, bacteria with the property of dissimilatory metal reduction have been in focus, although the biochemical reactions linking enzymatic Pd(II) reduction and Pd(0) deposition have not yet been identified. In this study we investigated Pd(II) reduction with formate as the electron donor in the presence of Gram‐negative bacteria with no documented capacity for reducing metals for energy production: Cupriavidus necator, Pseudomonas putida, and Paracoccus denitrificans. Only large and close‐packed Pd(0) aggregates were formed in cell‐free buffer solutions. Pd(II) reduction in the presence of bacteria resulted in smaller, well‐suspended Pd(0) particles that were associated with the cells (called “bioPd(0)” in the following). Nanosize Pd(0) particles (3–30 nm) were only observed in the presence of bacteria, and particles in this size range were located in the periplasmic space. Pd(0) nanoparticles were still deposited on autoclaved cells of C. necator that had no hydrogenase activity, suggesting a hydrogenase‐independent formation mechanism. The catalytic properties of Pd(0) and bioPd(0) were determined by the amount of hydrogen released in a reaction with hypophosphite. Generally, bioPd(0) demonstrated a lower level of activity than the Pd(0) control, possibly due to the inaccessibility of the Pd(0) fraction embedded in the cell envelope. Our results demonstrate the suitability of bacterial cells for the recovery of Pd(0), and formation and immobilization of Pd(0) nanoparticles inside the cell envelope. However, procedures to make periplasmic Pd(0) catalytically accessible need to be developed for future nanobiotechnological applications. Biotechnol. Bioeng. 2010;107: 206–215. © 2010 Wiley Periodicals, Inc.  相似文献   

3.
A novel catalytic activity of palladium [Pd(0)]-coated cells of Desulfovibrio desulfuricans ATCC 29577 ["bio-Pd(0)"] is demonstrated. Reduction of 700 microM Cr(VI) occurred within 24 h using formate (25 mM) or hydrogen (1 atm) as the electron donor, under conditions whereby cells lacking bound Pd(0), or palladium metal manufactured via chemical reduction of soluble Pd(II), did not reduce Cr(VI). The biomass-bound Pd(0) also functioned in the continuous removal of 400 microM Cr(VI) from a 1 mM solution under H(2) (flow residence time approximately 5 h), where chemically prepared Pd(0) was ineffective. This demonstrates a new type of active bioinorganic catalysis, whereby the presence of biomass bound to Pd(0) confers a novel catalytic capability not seen with Pd base metal or biomass alone.  相似文献   

4.
The nature of bonding at the interface between deposited silver/palladium and clean Al-terminated (001) surface of α-Al2O3 has been investigated using a periodic ab initio method. Substantial inter-planar relaxations within the alumina were found at both the interfaces and the bulk. The periodic calculation with both Ag and Pd deposition shows that 10% of loading on alumina results maximum stability. Surface energy and work function calculations were performed to propose the stability for the metals on the studied surfaces. The deposited Ag forms a three-dimensional (3-D) cluster on top of the alumina surface. The Pd cluster formed on the alumina surface is two-dimensional (2-D) and is distorted to accommodate the Ag cluster in its domain. A further low index calculation can explain the reason for a higher stability of the membrane generated over alumina support with silver and palladium. The results are discussed in view of the existing experimental data and models of metal-oxide interface and a reason for the difference of activity of the metal interaction with alumina surface is postulated.  相似文献   

5.
The sterically hindered thiolate complexes [Pd(SR)2(TMEDA)] SR = SC6F5 (2), SC6F4-4-H (3), SC6H4-2-SiPh3 (4) were easily synthesized by metathetical reactions of the corresponding palladium chloro compound [Pd(Cl)2(TMEDA)] (1) and the lead salt of the corresponding thiol. The identity of the three species being unequivocally determined by single crystal X-ray diffraction techniques. The catalytic activity of the palladium species [Pd(SR)2(TMEDA)] was explored in the Suzuki-Miyaura cross coupling reactions of different p-substituted bromobenzenes.  相似文献   

6.
N-(2-Pyridyl)acetamide (aapH) complexes of palladium(II), cobalt(II), nickel(II), and copper(II) have been studied by means of magnetic susceptibilities, and infrared, electronic, and PMR spectra. In the octahedral complexes M(aapH)2X2(M = Co, Ni, Cu; X = Cl, Br, NCS, NO3), bidentate aapH is chelated through the pyridine-N and amid-O atomes, whereas in the square-planar Pd(aapH)2X2 (X = Cl, Br) unidentate aapH is coordinated through the pyridine-N atom alone. Under alkaline conditions aapH is deprotonated in the presence of palladium(II) to form Pd(aap)2·4H2O, aap being an anionic bidentate ligand and chelating through the pyridine-N and amide-O atoms.  相似文献   

7.
In this work, a pair of new palladium(II) complexes, [Pd(Gly)(Phe)] and [Pd(Gly)(Tyr)], (where Gly is glycine, Phe is phenylalanine, and Tyr is tyrosine) were synthesized and characterized by UV–Vis, FT-IR, elemental analysis, 1H-NMR, and conductivity measurements. The detailed 1H NMR and infrared spectral studies of these Pd(II) complexes ascertain the mode of binding of amino acids to palladium through nitrogen of -NH2 and oxygen of -COO? groups as bidentate chelates. The Pd(II) complexes have been tested for in vitro cytotoxicity activities against cancer cell line of K562. Interactions of these Pd(II) complexes with CT-DNA and human serum albumin were identified through absorption/emission titrations and gel electrophoresis which indicated significant binding proficiency. The binding distance (r) between these synthesized complexes and HSA based on Forster?s theory of non-radiation energy transfer were calculated. Alterations of HSA secondary structure induced by complexes were confirmed by FT-IR measurements. The results of emission quenching at three temperatures have revealed that the quenching mechanism of these Pd(II) complexes with CT-DNA and HSA were the static and dynamic quenching mechanism, respectively. Binding constants (Kb), binding site number (n), and the corresponding thermodynamic parameters were calculated and revealed that the hydrogen binding and hydrophobic forces played a major role when Pd(II) complexes interacted with DNA and HSA, respectively. We bid that [Pd(Gly)(Phe)] and [Pd(Gly)(Tyr)] complexes exhibit the groove binding with CT-DNA and interact with the main binding pocket of HSA. The complexes follow the binding affinity order of [Pd(Gly)(Tyr)] > [Pd(Gly)(Phe)] with CT-DNA- and HSA-binding.  相似文献   

8.
Although peptide-enabled synthesis of nanostructures has garnered considerable interest for use in catalytic applications, it has so far been achieved mostly via Fmoc based solid phase peptide synthesis. Consequently, the potential of longer peptides in nanoparticle synthesis have not been explored largely due to the complexities and economic constraints of this chemical synthesis route. This study examines the potential of a 45-amino acid long peptide expressed as fusion to green fluorescence protein (GFPuv) in Escherichia coli for use in palladium nanoparticle synthesis. Fed-batch fermentation with E. coli harboring an arabinose-inducible plasmid produced a product containing three copies of Pd4 peptide fused to N-terminus of GFPuv ((Pd4)3-GFPuv). Using the intrinsic fluorescence of GFPuv, expression and enrichment of the fusion product was easily monitored. Crude lysate, desalted lysate, and an ion-exchange enriched fraction containing (Pd4)3-GFPuv were used to test the hypothesis that high purity of the biologic material used as the nanoparticle synthesis template may not be necessary. Nanoparticles were characterized using a variety of material science techniques and used to catalyze a model Suzuki–Miyaura coupling reaction. Results demonstrated that palladium nanoparticles can be synthesized using the soluble cell extract containing (Pd4)3-GFPuv without extensive purification or cleavage steps, and as a catalyst the crude mixture is functional.  相似文献   

9.
We improved the hydrogen yield from glucose using a genetically modified Escherichia coli. E. coli strain SR15 (ΔldhA, ΔfrdBC), in which glucose metabolism was directed to pyruvate formate lyase (PFL), was constructed. The hydrogen yield of wild-type strain of 1.08 mol/mol glucose, was enhanced to 1.82 mol/mol glucose in strain SR15. This figure is greater than 90 % of the theoretical hydrogen yield of facultative anaerobes (2.0 mol/mol glucose). Moreover, the specific hydrogen production rate of strain SR15 (13.4 mmol h−1 g−1 dry cell) was 1.4-fold higher than that of wild-type strain. In addition, the volumetric hydrogen production rate increased using the process where cells behaved as an effective catalyst. At 94.3 g dry cell/l, a productivity of 793 mmol h−1 l−1 (20.2 l h−1 l−1 at 37 °C) was achieved using SR15. The reported productivity substantially surpasses that of conventional biological hydrogen production processes and can be a trigger for practical applications.  相似文献   

10.

The acidophilic, Fe(III)-reducing heterotrophic bacteria Acidocella aromatica PFBCT and Acidiphilium cryptum SJH were utilized to produce palladium (Pd) bionanoparticles via a simple 1-step microbiological reaction. Monosaccharide (or intracellular NADH)-dependent reactions lead to visualization of intra/extra-cellular enzymatic Pd(0) nucleation. Formic acid-dependent reactions proceeded via the first slow Pd(0) nucleation phase and the following autocatalytic Pd(II) reduction phase regardless of the presence or viability of the cells. However, use of active cells (with full enzymatic and membrane protein activities) at low formic acid concentration (5 mM) was critical to allow sufficient time for Pd(II) biosorption and the following enzymatic Pd(0) nucleation, which consequently enabled production of fine, dense and well-dispersed Pd(0) bionanoparticles. Differences of the resultant Pd(0) nanoparticles in size, density and localization between the two bacteria under each condition tested suggested different activity and location of enzymes and membrane “Pd(II) trafficking” proteins responsible for Pd(0) nucleation. Despite the inhibitory effect of leaching lixiviant and dissolved metal ions, Pd(0) bionanoparticles were effectively formed by active Ac. aromatica cells from both acidic synthetic Pd(II) solutions and from the actual spent catalyst leachates at equivalent 18–19 nm median size with comparable catalytic activity.

  相似文献   

11.
The surfaces of cells of Desulfovibrio desulfuricans,Desulfovibrio vulgaris and a new strain, Desulfovibrio sp. `Oz-7' were used to manufacturea novel bioinorganic catalyst via the reduction of Pd(II) to Pd(0) at the cell surface usinghydrogen as the electron donor. The ability of the palladium coated (palladised) cells to reductivelydehalogenate chlorophenol and polychlorinated biphenyl species was demonstrated. Dried, palladisedcells of D. desulfuricans, D. vulgaris and Desulfovibrio sp. `Oz-7'were more effective bioinorganic catalysts than Pd(II) reduced chemically under H2 orcommercially available finely divided Pd(0). Differences were observed in the catalyticactivity of the preparations when compared with each other. Negligible chloride release occurredfrom chlorophenol and polychlorinated biphenyls using biomass alone.  相似文献   

12.
The strong affinity of chitosan for metal ions and more specifically for precious metals such as palladium and platinum has focused the interest on using this biopolymer as a support for catalytic metals. The manufacturing of hollow chitosan fibers, softly cross-linked with glutaraldehyde, followed by palladium sorption at pH 2 in HCl solutions and further reduction using hydrogen gas, opened the route for the design of a new continuous catalytic system. This material was used for the hydrogenation of nitrotoluene, which was converted into o-toluidine, in methanol solutions. The substrate was circulated inside the lumen of the fiber, while the hydrogen donor (hydrogen gas) was maintained at constant pressure in the outlet compartment of the reactor. Several parameters (substrate concentration, metal content in the fiber, and flow rate) have been tested for their impact on catalytic performance, measured by the turnover frequency (TOF), conversion yield or o-toluidine production, using a surface response methodology for the optimization of the process. Metal content in the fiber revealed a critical parameter; the influence of this parameter was extensively studied through the structural characterization of the fibers using XPS analysis (oxidation state of Pd), X-ray diffraction analysis (size of Pd crystals), TEM analysis (size and distribution of Pd crystals), and diffusion profiles (porosity) in order to correlate catalytic performance to fiber characterization.  相似文献   

13.
A new water-soluble palladium(II) complex, [Pd(bpy)(pyr-Ac]NO3 in which bpy = 2,2′-bipyridine and pyr-Ac is 1-pyrrolacetato, has been synthesized and characterized by spectroscopic methods (1H NMR, FT-IR, and UV-Vis), molar conductivity measurements, and elemental analysis. The results obtained from elemental analysis and conductivity measurements confirmed the stoichiometry of ligand and its complex while the characteristic peaks in UV-Vis and FT-IR and resonance peaks in 1H NMR spectra confirmed the formation of ligand frameworks around the palladium ion. The 50% cytotoxic concentration (Ic50) of new synthesized Pd(II) complex was determined by using MTT assay against human breast cancer cell line, T47D. The interaction between the Pd(II) complex with calf thymus DNA was studied at different temperatures by using absorption spectroscopy, fluorescence titration spectra, ethidium bromide displacement, and gel chromatography studies. The results obtained by absorption spectroscopy revealed that the Pd(II) complex can bind to DNA cooperatively at low concentrations. Several binding parameters in the above interaction were calculated by the fluorescence quenching method. The quenching mechanism was suggested to be the static quenching. The thermodynamic parameters: enthalpy change (ΔH °), entropy change (ΔS °), and Gibbs free energy (ΔG °), showed that van der Waals and hydrogen binding are predominant intermolecular forces between Pd(II) complex and DNA. These results were also consistent with the results obtained from Scatchard's plots.  相似文献   

14.
Ligand binding studies on carrier proteins are crucial in determining the pharmacological properties of drug candidates. Here, a new palladium(II) complex was synthesized and characterized. The in vitro binding studies of this complex with two carrier proteins, human serum albumin (HSA), and β-lactoglobulin (βLG) were investigated by employing biophysical techniques as well as computational modeling. The experimental results showed that the Pd(II) complex interacted with two carrier proteins with moderate binding affinity (Kb ≈ .5 × 104 M?1 for HSA and .2 × 103 M?1 for βLG). Binding of Pd(II) complex to HSA and βLG caused strong fluorescence quenching of both proteins through static quenching mechanism. In two studied systems hydrogen bonds and van der Waals forces were the major stabilizing forces in the drug-protein complex formation. UV–Visible and FT-IR measurements indicated that the binding of above complex to HSA and βLG may induce conformational and micro-environmental changes of two proteins. Protein–ligand docking analysis confirmed that the Pd(II) complex binds to residues located in the subdomain IIA of HSA and site A of βLG. All these experimental and computational results suggest that βLG and HSA might act as carrier protein for Pd(II) complex to deliver it to the target molecules.  相似文献   

15.
The application of Mg for hydrogen storage is hindered due to the slow absorption of hydrogen in Mg films. Herein, the hydrogenation process is explored theoretically using density functional theory calculations, and energy barriers are compared for hydrogen diffusion through Pd nanoparticle/Mg film interfaces and their variations, i.e., Pd(H)/Mg(O). Decomposing the mechanism into basic steps, it is shown that Pd undergoes a strain‐induced crystallographic phase transformation near the interface, and indicated that hydrogen saturation of Pd nanoparticles enhances their efficiency as nanoportals. Using energetic arguments, it is explained why hydrogen diffusion is practically prohibited through native Mg oxide and seriously suppressed through existing hydride domains. Hydrogen flux is experimentally investigated through the nanoportals in Pd‐nanoparticle decorated Mg films by pressure‐composition isotherm measurements. An r ≈ t1/3 relationship is theoretically calculated for the radial growth of hemispherical hydride domains, and this relationship is confirmed by atomic force microscopy. The diffusion constant of hydrogen in Mg films is estimated as DHfilm ≈ 8 × 10?18 m2 s?1, based on transmission electron microscopy characterization. The unique nanoportal configuration allows direct measurement of hydride domain sizes, thus forming a model system for the experimental investigation of hydrogenation in any material.  相似文献   

16.
Tritonia diomedea uses the Earth’s magnetic field as an orientation cue, but little is known about the neural mechanisms that underlie magnetic orientation behavior in this or other animals. Six large, individually identifiable neurons in the brain of Tritonia (left and right Pd5, Pd6, Pd7) are known to respond with altered electrical activity to changes in earth-strength magnetic fields. In this study we used immunochemical, electrophysiological, and neuroanatomical techniques to investigate the function of the Pd5 neurons, the largest magnetically responsive cells. Immunocytochemical studies localized TPeps, neuropeptides isolated from Pd5, to dense-cored vesicles within the Pd5 somata and within neurites adjacent to ciliated foot epithelial cells. Anatomical analyses revealed that neurites from Pd5 are located within nerves innervating the ipsilateral foot and body wall. These results imply that Pd5 project to the foot and regulate ciliary beating through paracrine release. Electrophysiological recordings indicated that, although both LPd5 and RPd5 responded to the same magnetic stimuli, the pattern of spiking in the two cells differed. Given that TPeps increase ciliary beating and Tritonia locomotes using pedal cilia, our results are consistent with the hypothesis that Pd5 neurons control or modulate the ciliary activity involved in crawling during orientation behavior.  相似文献   

17.
The versatility of palladium(II) as a template for Mannich-type macrocyclization is illustrated. Reaction of (bis(3-aminopropyl)piperazine)palladium(II) with formaldehyde and nitroethane in basic aqueous solution yields the ‘reinforced’ macrocycle 7-methyl-7-nitro-1,5,9,13-tetraazabicyclo[11.2.2]heptadecane as its palladium(II) complex. The crystal structure shows the palladium ion lies in a slightly tetrahedrally distorted square plane of four nitrogen donors, with distances to the two tertiary donors [av. 2.059(3) Å] slightly shorter than those to the secondary amines [av. 2.066(3) Å]. The 3-methyl-3-nitro-1,5,9,13-tetraazacyclohexadecane as its palladium(II) complex was prepared by an analogous route. In a separate reaction based on the [Pd(en)(chxn)]2+ (en = ethane-1,2-diamine; chxn = cyclohexane-1,2-diamine) intermediate, an unsymmetrical macrocycle with a fused cyclohexane ring, 4,11-dimethyl-4,11-dinitro-2,6,9,13-tetraazabicyclo[12.4.0]octadecane was isolated as its palladium(II) complex. Accessibility to an isolable mixed-ligand precursor is a key to this reaction, provided by using palladium(II) as the templating metal. Reaction of (4,8-diazaundecane-1,11-diamine)palladium(II) with formaldehyde and diethyl malonate in basic aqueous solution yields, with ester hydrolysis and decarboxylation, the carboxylate-pendant macrocycle 1,5,9,13-tetraazacyclohexadecane-3-carboxylic acid as its palladium(II) complex. The crystal structure is comprised of hydrogen-bonded dimers {[Pd(L)][Pd(L-H)]}3+ where the pair of inversion related square-planar complexes share a single proton between their pendant carboxylates. Bis(3-aminopropyl)(piperazine)palladium(II) yields the macrocyclic complex ion (1,5,9,13-tetraazabicyclo[11.2.2]heptadecane-7-carboxylic acid)palladium(II), in a similar reaction.  相似文献   

18.
Biomass of Desulfovibrio desulfuricans was used to recover Au(III) as Au(0) from test solutions and from waste electronic scrap leachate. Au(0) was precipitated extracellularly by a different mechanism from the biodeposition of Pd(0). The presence of Cu2+ (∼2000 mg/l) in the leachate inhibited the hydrogenase-mediated removal of Pd(II) but pre-palladisation of the cells in the absence of added Cu2+ facilitated removal of Pd(II) from the leachate and more than 95% of the Pd(II) was removed autocatalytically from a test solution supplemented with Cu(II) and Pd(II). Metal recovery was demonstrated in a gas-lift electrobioreactor with electrochemically generated hydrogen, followed by precipitation of recovered metal under gravity. A 3-stage bioseparation process for the recovery of Au(III), Pd(II) and Cu(II) is proposed.Victoria S. Baxter-Plant – Deceased  相似文献   

19.
 The palladium(II) aqua complexes [Pd(H2O)4]2+, cis-[Pd(en)(H2O)2]2+, and cis-[Pd(dtco-OH)(H2O)2]2+ effect hydrolytic cleavage of horse myoglobin in aqueous solution. The conditions were optimized with the third complex. Its structure was determined by X-ray crystallographic analysis of its precursor, the square-planar complex cis-[Pd(dtco-OH)Cl2], in which the chelating ligand adopts a boat-chair conformation. A weak interaction between the hydroxyl group and the palladium(II) atom seems to improve the stability of the reagent. The yield of cleavage after a 24-h incubation at 60  °C increases from 39% to 85% as the pH decreases from 6.2 to 3.2. The protein fragments are separated by SDS-PAGE electrophoresis and HPLC separation methods, and identified by ESIMS and MALDI-TOF mass spectrometric methods and by determination of terminal amino-acid sequences. Most of the 13 cleavage sites are clustered around the methionine, arginine, and some of the histidine residues, whose side chains can bind to palladium(II). Cleavage tends to occur at the peptide bonds one to three positions removed from the binding residues; the scissile bonds usually lie on the amino-terminal side, seldom on the carboxy-terminal side, of the binding residues. Removal of the heme and unfolding of the protein do not drastically alter the pattern of cleavage. The ability of palladium(II) aqua complexes to cleave proteins at relatively few sites, with explicable selectivity, with good to very good yield, and in weakly acidic and nearly neutral solutions, bodes well for their future use in biochemical and bioanalytical practice. Received: 23 December 1997 / Accepted: 14 April 1998  相似文献   

20.
The syntheses of nine palladium(II) complexes of type [Pd(phen)(AA)]+ (where AA is an anion of glycine, L-alanine, L-leucine, L-phenylalanine, L-tyrosine, L-tryptophan, L-valine, L-proline, or L-serine) have been achieved. These palladium(II) complexes have been characterized by ultraviolet-visible, infrared, and 1H NMR spectroscopy. The binding studies of several complexes [M(NN)(AA)]+ (where M is Pd(II) as Pt(II), NN is 2,2'-bipyridine or 1,10-phenanthrodine, and AA is an anion of amino acid) with calf thymus DNA have been carried out using UV difference absorption and fluorescence spectroscopy. The mode of binding of the above complexes to DNA suggests the involvement of the hydrogen bonding between them. Several complexes [M(phen)(AA)]+ (where M is Pd(II) or Pt(II) and AA is an anion of amino acid) have also been screened for cytotoxicity in P388 lymphocytic cells. Of them, only two complexes, [Pd(Phen)(Gly)]+ and [Pd(phen)(Val)]+, show comparable cytotoxicity, as cisplatin does.  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