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.
《Inorganica chimica acta》2004,357(9):2561-2569
Ni(II), Cu(II), Zn(II) and Cd(II) complexes of an N4-donor Schiff base, containing (CH2)2 as spacer, have been prepared. The X-ray crystal structures of monohelical Ni(ETs) · H2O and the homochirally crystallised Δ-Cu(ETs), as well as the meso-helicate Zn2(ETs)2 · MeCN [H2ETs: N,N-bis(2-tosylaminobenzylidene)-1,2-diaminoethane] have been solved. In the latter, the ligand behaves as bis-bidentate, displaying a “C”-type arrangement, instead of the typical “S”-type fashion present in bis-helical dinuclear complexes.  相似文献   

2.
The activation properties of Kv1.2 channels are highly variable, with reported half-activation (V1/2) values ranging from ∼−40 mV to ∼+30 mV. Here we show that this arises because Kv1.2 channels occupy two distinct gating modes (“fast” and “slow”). “Slow” gating (τact = 90 ± 6 ms at +35 mV) was associated with a V1/2 of activation of +16.6 ± 1.1 mV, whereas “fast” gating (τact = 4.5 ± 1.7 ms at +35 mV) was associated with a V1/2 of activation of −18.8 ± 2.3 mV. It was possible to switch between gating modes by applying a prepulse, which suggested that channels activate to a single open state along separate “fast” and “slow” activation pathways. Using chimeras and point mutants between Kv1.2 and Kv1.5 channels, we determined that introduction of a positive charge at or around threonine 252 in the S2-S3 linker of Kv1.2 abolished “slow” activation gating. Furthermore, dialysis of the cytoplasm or excision of cell-attached patches from cells expressing Kv1.2 channels switched gating from “slow” to “fast”, suggesting involvement of cytoplasmic regulators. Collectively, these results demonstrate two modes of activation gating in Kv1.2 and specific residues in the S2-S3 linker that act as a switch between these modes.  相似文献   

3.
This study describes an analysis of different treatments that influence the relative content and the midpoint potential of HP Cyt b559 in PS II membrane fragments from higher plants. Two basically different types of irreversible modification effects are distinguished: the HP form of Cyt b559 is either predominantly affected when the heme group is oxidized (“O-type” effects) or when it is reduced (“R-type” effects). Transformation of HP Cyt b559 to lower potential redox forms (IP and LP forms) by the “O-type” mechanism is induced by high pH and detergent treatments. In this case the effects consist of a gradual decrease in the relative content of HP Cyt b559 while its midpoint potential remains unaffected. Transformation of HP Cyt b559 via an “R-type” mechanism is caused by a number of exogenous compounds denoted L: herbicides, ADRY reagents and tetraphenylboron. These compounds are postulated to bind to the PS II complex at a quinone binding site designated as QC which interacts with Cyt b559 and is clearly not the QB site. Binding of compounds L to the QC site when HP Cyt b559 is oxidized gives rise to a gradual decrease in the Em of HP Cyt b559 with increasing concentration of L (up to 10 Kox(L) values) while the relative content of HP Cyt b559 is unaffected. Higher concentrations of compounds L required for their binding to QC site when HP Cyt b559 is reduced (described by Kred(L)) induce a conversion of HP Cyt b559 to lower potential redox forms (“R-type” transformation). Two reaction pathways for transitions of Cyt b559 between the different protein conformations that are responsible for the HP and IP/LP redox forms are proposed and new insights into the functional regulation of Cyt b559 via the QC site are discussed.  相似文献   

4.
Cobalt involvement in chemical and metallobiological processes entails largely unknown reactivity pathways with a variety of ligands. Such ligands include phosphonate and carboxylate-containing metal ion binders. In an attempt to investigate the nature and properties of species arising from aqueous interactions between Co(II) and N,N-bis(phosphonomethyl)-glycine (H5NTA2P), reactions between the two led to an assembly of species in (NH4)4[Co(H2O)6][(H2O)2Co(HNTA2P)Co(NH3)2(H2O)3]2[Co(NTA2P)(H2O)2]2 · 10H2O · 1.36CH3CH2OH (1) at pH ∼ 5.5. The analytical, spectroscopic and X-ray data on 1 reveal mononuclear and dinuclear complexes of Co(II) surrounded by oxygens, belonging to terminal carboxylates, phosphonates and bound water molecules, and nitrogen atoms from coordinated ammonia and HxNTA2Pq (x = 1, q = 4; x = 0, q = 5) ligands. Worth noting is the variable protonation state of the bound diphosphonate ligand and its ability to bridge two Co(II) centers with ostensibly differing coordination spheres. The assembly of three Co(II) species of variable nuclearity and composition attests to the importance of pH-specific conditions, under which “capturing” of more than one species can be achieved for a given Co(II):H5NTA2P stoichiometry in the presence of ammonia. Collectively, 1 provides a rare glimpse of a “slice” of the aqueous speciation of the binary Co(II)-H5NTA2P system, while its lattice composition projects key structural features in Co(II)-carboxyphosphonate materials.  相似文献   

5.
Three new iron(II) N6 tripodal complexes provide information on the role of ligand conformation on spin crossover behavior. The ligands (generated in situ) are the Schiff base condensate of tris(2-aminoethyl)amine (tren) with three equivalents of 4-methyl-5-imidazolecarboxaldehyde, H3(1), and the condensates of tris(2-aminoethyl)methylammonium ion (N(Me)tren+) with three equivalents of 4-methyl-5-imidazolecarboxaldehyde, N(Me)H3(1)+, or with 2-imidazole carboxaldehyde, N(Me)H3(3)+. The structures of [FeH3(1)](ClO4)2, [FeN(Me)H3(1)](ClO4)3 and [FeN(Me)H3(1)](ClO4)3 are reported. The central tren nitrogen atom in these complexes exhibits three different geometries, pyramidal with the nitrogen pointed toward the iron (“N in”, Fe-N distance of 3.050 Å), planar (Fe-N distance of 3.527 Å), and pyramidal with the nitrogen pointed away from the iron atom (“N out”, Fe-N distance of 3.921 Å). With iron(II) the “N in” geometry is high spin while the planar and “N out” geometries are low spin. [FeH3(1)](ClO4)2 exhibits spin crossover behavior between room temperature and 77 K as determined by Mössbauer spectroscopy and also exhibits a conformational change from “N in” to planar over this same temperature range. The structures of [FeN(Me)H3(1)](ClO4)3 and [FeN(Me)H3 (3)](ClO4)3 are locked into the “N out” geometry due to the quaternary nitrogen atom and are low spin even at room temperature. The LS planar and “N out” conformations place a strain on the bond angles of the aliphatic arms of the ligand, which are more pronounced in the “N out” case. The HS “N in” geometry lacks this strain.  相似文献   

6.
In many cytochrome c oxidases glutamic acid 242 is required for proton transfer to the binuclear heme a3/CuB site, and for proton pumping. When present, the side chain of Glu-242 is orientated “down” towards the proton-transferring D-pathway in all available crystal structures. A nonpolar cavity “above” Glu-242 is empty in these structures. Yet, proton transfer from Glu-242 to the binuclear site, and for proton-pumping, is well established, and the cavity has been proposed to at least transiently contain water molecules that would mediate proton transfer. Such proton transfer has been proposed to require isomerisation of the Glu-242 side chain into an “up” position pointing towards the cavity. Here, we have explored the molecular dynamics of the protonated Glu-242 side chain. We find that the “up” position is preferred energetically when the cavity contains four water molecules, but the “down” position is favoured with less water. We conclude that the cavity might be deficient in water in the crystal structures, possibly reflecting the “resting” state of the enzyme, and that the “up/down” equilibrium of Glu-242 may be coupled to the presence of active-site water molecules produced by O2 reduction.  相似文献   

7.
Nilaparvata lugens Stål (brown planthopper, BPH), is one of the major insect pests of rice (Oryza sativa L.) in the temperate rice-growing region. In this study, ASD7 harboring a BPH resistance gene bph2 was crossed to a susceptible cultivar C418, a japonica restorer line. BPH resistance was evaluated using 134 F2:3 lines derived from the cross between “ASD7” and “C418”. SSR assay and linkage analysis were carried out to detect bph2. As a result, the resistant gene bph2 in ASD7 was successfully mapped between RM7102 and RM463 on the long arm of chromosome 12, with distances of 7.6 cM and 7.2 cM, respectively. Meanwhile, both phenotypic selection and marker-assisted selection (MAS) were conducted in the BC1F1 and BC2F1 populations. Selection efficiencies of RM7102 and RM463 were determined to be 89.9% and 91.2%, respectively. It would be very beneficial for BPH resistance improvement by using MAS of this gene.  相似文献   

8.
A novel “off-On” electrogenerated chemiluminescence (ECL) biosensor has been developed for the detection of mercury(II) based on molecular recognition technology. The ECL mercury(II) biosensor comprises two main parts: an ECL substrate and an ECL intensity switch. The ECL substrate was made by modifying the complex of Ruthenium(II) tris-(bipyridine)(Ru(bpy)32+)/Cyclodextrins-Au nanoparticles(CD-AuNps)/Nafion on the surface of glass carbon electrode (GCE), and the ECL intensity switch is the single hairpin DNA probe designed according to the “molecular recognition” strategy which was functionalized with ferrocene tag at one end and attached to Cyclodextrins (CD) on modified GCE through supramolecular noncovalent interaction. We demonstrated that, in the absence of Hg(II) ion, the probe keeps single hairpin structure and resulted in a quenching of ECL of Ru(bpy)32+. Whereas, in the presence of Hg(II) ion, the probe prefers to form the T-Hg(II)-T complex and lead to an obvious recovery of ECL of Ru(bpy)32+, which provided a sensing platform for the detection of Hg(II) ion. Using this sensing platform, a simple, rapid and selective “off-On” ECL biosensor for the detection of mercury(II) with a detection limit of 0.1 nM has been developed.  相似文献   

9.
10.
Four lead(II) complexes with substituted 2,2′-bipyridine adducts and β-diketonates ligands, [Pb(5,5′-dm-2,2′-bpy)(tfpb)2]21, [Pb(4,4′-dmo-2,2′-bpy)(tfpb)2]22, [Pb(4,4′-dm-2,2′-bpy)(tfnb)2]23 and [Pb(5,5′-dm-2,2′-bpy)(tfnb)2]24, (“4,4′-dm-2,2′-bpy”, “5,5′-dm-2,2′-bpy”, “4,4′-dmo-2,2′-bpy”, “Htfpb” and “Htfnb” are the abbreviations of 4,4′-dimethyl-2,2′-bipyridine, 5,5′-dimethyl-2,2′-bipyridine, 4,4′-dimethoxy-2,2′-bipyridine, 4,4,4-trifluoro-1-phenyl-1,3-butanedione and 4,4,4-trifluoro-1-(2-naphthyl)-1,3-butanedione, respectively) have been synthesized and characterized by elemental analysis, IR, 1H NMR and 13C NMR spectroscopy and also studied by thermal and electrochemical as well as X-ray crystallography. The supramolecular features in these complexes are guided/controlled by weak directional intramolecular interactions.  相似文献   

11.
Although it is well known that the so-called “equivalent solution” or “effective” solvent permittivity (dielectric constant) in proteins and nucleic acids is lower than in bulk water, this fact is commonly neglected in (bioinorganic) studies of such compounds. Using domain 5 of the group II intron ribozyme Sc.ai5γ, we describe here the influence of 1,4-dioxane-d8 on the structure and magnesium(II)-binding properties of this catalytic domain. Applying one- and two-dimensional NMR, we observe distinct structural changes in the functionally important bulge region following a decrease in solvent permittivity. Concomitantly, an increase by a factor of 1.5 in the affinity of Mg2+ towards the individual-binding sites in the catalytic core domain is observed upon addition of 1,4-dioxane-d8. This has led to the detection of a new metal ion coordination site near the GU wobble pair in the catalytic triad. Our results show that solvent permittivity is an important factor in the formation of intrinsic RNA structures and affects their metal ion-binding properties. Hence, solvent permittivity should be taken into account in future studies.  相似文献   

12.
The dreaded pathogen Staphylococcus aureus is one of the causes of morbidity and mortality worldwide. Glyceraldehyde-3-phosphate dehydrogenase (GAPDH), one of the key glycolytic enzymes, is irreversibly oxidized under oxidative stress and is responsible for sustenance of the pathogen inside the host. With an aim to elucidate the catalytic mechanism and identification of intermediates involved, we describe in this study different crystal structures of GAPDH1 from methicillin-resistant S. aureus MRSA252 (SaGAPDH1) in apo and holo forms of wild type, thioacyl intermediate, and ternary complexes of active-site mutants with physiological substrate d-glyceraldehyde-3-phosphate (G3P) and coenzyme NAD+. A new phosphate recognition site, “new Pi” site, similar to that observed in GAPDH from Thermotoga maritima, is reported here, which is 3.40 Å away from the “classical Pi” site. Ternary complexes discussed are representatives of noncovalent Michaelis complexes in the ground state. d-G3P is bound to all the four subunits of C151S.NAD and C151G.NAD in more reactive hydrate (gem-di-ol) form. However, in C151S + H178N.NAD, the substrate is bound to two chains in aldehyde form and in gem-di-ol form to the other two. This work reports binding of d-G3P to the C151G mutant in an inverted manner for the very first time. The structure of the thiaocyl complex presented here is formed after the hydride transfer. The C3 phosphate of d-G3P is positioned at the “Ps” site in the ternary complexes but at the “new Pi” site in the thioacyl complex and C1-O1 bond points opposite to His178 disrupting the alignment between itself and NE2 of His178. A new conformation (Conformation I) of the 209-215 loop has also been identified, where the interaction between phosphate ion at the “new Pi” site and conserved Gly212 is lost. Altogether, inferences drawn from the kinetic analyses and crystal structures suggest the “flip-flop” model proposed for the enzyme mechanism.  相似文献   

13.
The interactions of monofunctional [MCl(chelate)] compounds (M = Pt(II), Pd(II) or Au(III) and chelate = diethylenetriamine, dien or 2,2′,2″-terpyridine, terpy) with the C-terminal finger of the HIV nucleocapsid NCp7 zinc finger (ZF) were studied by mass spectrometry and circular dichroism spectroscopy. In the case of [M(dien)] species, Pt(II) and Pd(II) behaved in a similar fashion with evidence of adducts caused by displacement of Pt-Cl or Pd-Cl by zinc-bound thiolate. Labilization, presumably under the influence of the strong trans influence of thiolate, resulted in loss of ligand (dien) as well as zinc ejection and formation of species with only Pd(II) or Pt(II) bound to the finger. For both Au(III) compounds the reactions were very fast and only “gold fingers” with no ancillary ligands were observed. For all terpyridine compounds ligand scrambling and metal exchange occurred with formation of [Zn(terpy)]2+. The results conform well to those proposed from the study of model Zn compounds such as N,N′-bis(2-mercapto-ethyl)-1,4-diazacycloheptanezinc(II), [Zn(bme-dach)]2. The possible structures of the adducts formed are discussed and, for Pt(II) and Pd(II), the evidence for possible expansion of the zinc coordination sphere from four- to five-coordinate is discussed. This observation reinforces the possibility of change in geometry for zinc in biology, even in common “structural” sites in metalloenzymes. The results further show that the extent and rate of zinc displacement by inorganic compounds can be modulated by the nature (metal, ligands) of the reacting compound.  相似文献   

14.
The projection structures of complex I and the I + III2 supercomplex from the C4 plant Zea mays were determined by electron microscopy and single particle image analysis to a resolution of up to 11 Å. Maize complex I has a typical L-shape. Additionally, it has a large hydrophilic extra-domain attached to the centre of the membrane arm on its matrix-exposed side, which previously was described for Arabidopsis and which was reported to include carbonic anhydrase subunits. A comparison with the X-ray structure of homotrimeric γ-carbonic anhydrase from the archaebacterium Methanosarcina thermophila indicates that this domain is also composed of a trimer. Mass spectrometry analyses allowed to identify two different carbonic anhydrase isoforms, suggesting that the γ-carbonic anhydrase domain of maize complex I most likely is a heterotrimer. Statistical analysis indicates that the maize complex I structure is heterogeneous: a less-abundant “type II” particle has a 15 Å shorter membrane arm and an additional small protrusion on the intermembrane-side of the membrane arm if compared to the more abundant “type I” particle. The I + III2 supercomplex was found to be a rigid structure which did not break down into subcomplexes at the interface between the hydrophilic and the hydrophobic arms of complex I. The complex I moiety of the supercomplex appears to be only of “type I”. This would mean that the “type II” particles are not involved in the supercomplex formation and, hence, could have a different physiological role.  相似文献   

15.
We present the application of a novel isotope dilution method, named Alternate Isotope-Coded Derivatization Assay (AIDA), to the quantitative analysis of hydrazone derivatives of malondialdehyde (MDA) and 4-hydroxynonenal (4-HNE) in exhaled breath condensate (EBC) samples using liquid chromatography–tandem mass spectrometry. AIDA is based on the alternate derivatization of the analyte(s) with reagents that are available in two pure isotopic forms, respectively “light” and “heavy”, by using light-derivatized standards for the quantification of the heavy-derivatized analytes, and vice versa. To this purpose, 2,4-dinitro-3,5,6-trideuterophenylhydrazine (d3-DNPH) has been synthesized and used as “heavy” reagent in combination with commercial “light” DNPH. Using the AIDA method, any unknown concentration of the analyte in the matrix can be calculated without the need of a calibration curve. An external calibration method has been also investigated for comparative purpose. The stability of DNPH and d3-DNPH derivatives was verified by excluding any exchange of hydrazones with each other. In the range of concentrations of biological interest, e.g., 2–40 nM for MDA and 0.5–10 nM for 4-HNE, the derivatization reactions of MDA and 4-HNE with DNPH and d3-DNPH showed overlapping kinetics and comparable yields. The MS response of both DNPH and d3-DNPH hydrazones was similar. The precision of AIDA, calculated as %RSD, was within 3.2–8% for MDA and 4.5–11% for 4-HNE. Accuracy was tested by analyzing a spiked EBC pool sample and acceptable results (accuracy within 98–108% for MDA and 93–114% for 4-HNE) were obtained by AIDA after subtraction of the blank, which was not negligible. The results of quantitative analysis of MDA and 4-HNE in EBC samples obtained by AIDA assay with four analyses per sample were in good agreement with those obtained by external calibration method on the same samples.  相似文献   

16.
Recent evidence indicates that winter-red leaf phenotypes in the mastic tree (Pistacia lentiscus) are more vulnerable to chronic photoinhibition during the cold season relative to winter-green phenotypes occurring in the same high light environment. This was judged by limitations in the maximum quantum yield of photosystem II (PSII), found in previous studies. In this investigation, we asked whether corresponding limitations in leaf gas exchange and carboxylation reactions could also be manifested. During the cold (“red”) season, net CO2 assimilation rates (A) and stomatal conductances (gs) in the red phenotype were considerably lower than in the green phenotype, while leaf internal CO2 concentration (Ci) was higher. The differences were abolished in the “green” period of the year, the dry summer included. Analysis of A versus Ci curves indicated that CO2 assimilation during winter in the red phenotype was limited by Rubisco content and/or activity rather than stomatal conductance. Leaf nitrogen levels in the red phenotype were considerably lower during the red-leaf period. Consequently, we suggest that the inherently low leaf nitrogen levels are linked to the low net photosynthetic rates of the red plants through a decrease in Rubisco content. Accordingly, the reduced capacity of the carboxylation reactions to act as photosynthetic electron sinks may explain the corresponding loss of PSII photon trapping efficiency, which cannot be fully alleviated by the screening effect of the accumulated anthocyanins.  相似文献   

17.
18.
A recombinant inbred line (RIL) population bred from a cross between a javanica type (cv. D50) and an indica type (cv. HB277) rice was used to map seven quantitative trait loci (QTLs) for thousand grain weight (TGW). The loci were distributed on chromosomes 2, 3, 5, 6, 8 and 10. The chromosome 3 QTL qTGW3.2 was stably expressed over two years, and contributed 9–10% of the phenotypic variance. A residual heterozygous line (RHL) was selected from the RIL population and its selfed progeny was used to fine map qTGW3.2. In this “F2” population, the QTL explained about 23% of the variance, rising to nearly 33% in the subsequent “F2:3” generation. The physical location of qTGW3.2 was confined to a ~ 556 kb region flanked by the microsatellite loci RM16162 and RM16194. The region also contains other factors influencing certain yield-related traits, although it is also possible that qTGW3.2 affects these in a pleiotropic fashion.  相似文献   

19.
Hydrocorals of the genus Millepora are abundant skeleton-forming inhabitants of coral reefs around the world. These species are popularly known as “fire corals” since contact with them causes severe pain, skin eruptions and blisters as a result of the release of unidentified toxins. Millepora species associate with photosynthetic dinoflagellates of the genus Symbiodinium (“zooxanthellae”), and up to now the role of these symbionts in the toxic effects induced by the “fire corals” is unknown. In this study, we compared the hemolytic, vasoconstrictor, and phospholipase A2 (PLA2) activities of the crude aqueous extracts prepared from normal and bleached specimens of two hydrocorals collected in the Mexican Caribbean, Milleporaalcicornis and Millepora complanata. Electrophoretic analysis revealed some differences between the protein profiles of the extracts prepared from normal and bleached specimens. Bleaching decreased, but not abolished, the hemolytic effect induced by the hydrocorals extracts and the phospholipase A2 activity of M. complanata extract. Furthermore, it did not modify the enzymatic activity of M. alcicornis extract and vasoconstriction elicited by both extracts. Our results suggest that the presence of the symbionts does not importantly influence the pharmacological and toxic effects induced by Millepora ssp. extracts, and indicate that cnidarians are the main source of the bioactive compounds.  相似文献   

20.
Abundant isolated remains of stylophoran echinoderms (cornutes and mitrates) are reported for the first time in the late Tremadocian (Asaphellus Zone) Tumugol Formation of Korea. Mitrate remains include numerous adorals of Kirkocystidae. Several new important anatomical features have been observed on these adorals, as an internal calcitic layer that is associated to s2 and possibly also to the palmar complex. This observation suggests that the palmar complex would be present not only in mitrocystitid mitrates, but also in peltocystitids. For the first time, several morphometric analyses have been undertaken based on isolated kirkocystid adorals, so as to explore the morphological diversity displayed by Korean adorals, but also in order to compare their morphology with that of other Gondwanan kirkocystids. Morphometric analyses indicate the occurrence of two contrasted morphologies within Korean adorals (morphotypes A and B), and of three distinct morphologies within European and North African forms (“Anatifopsis”, “Balanocystites”, and “escandei” morphotypes). Comparison of Korean adorals with those from Europe and North Africa shows that: (1) morphotypes B and “Anatifopsis” are equivalent; consequently, the two Korean specimens referred to morphotype B are assigned to the genus Anatifopsis; (2) morphology of most Korean adorals, which belong to morphotype A, is clearly distinct from that of all other described kirkocystids from Europe and North Africa. The small size, juvenile morphology, and great morphological variability observed in the morphotype A of the Korean adorals are suggestive of possible heterochronic processes (peramorphosis).  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