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.
The hydrolyzed α-[Ru(azpy)2Cl2] (azpy is 2-(phenylazo)pyridine; α indicates that the isomer in which the coordinating pairs Cl, N(py), and N(azo) are cis, trans, and cis, respectively) binding to guanine (G), adenine (A), methionine (Met), and histidine (His) residues were investigated by using density functional theory. Reactant complexes (RC), product complexes (PC), and transition states (TS) involved were fully characterized. The calculated energy profiles showed that the activation free energies for the substitutions of hydrolyzed α-[Ru(azpy)2Cl2] with Met was apparently lower than those of guanine and adenine. This indicate that the hydrolyzed α-[Ru(azpy)2Cl2] compounds may preferentially bind to the sulfur-containing amino acids residues in vivo. Moreover, the natural orbital population analysis (NPA) showed that the Ru atom gained the greatest negative charges in the reactions of hydrolyzed α-[Ru(azpy)2Cl2] with Met, which may contribute to their remarkably low activation free energies partially.  相似文献   

2.
Cytochromes c' are heme proteins found in photosynthetic and denitrifying bacteria, where they are presumably involved in electron transport. The cytochrome c' isolated from the bacterium Rhodocyclus gelatinosus (RGCP) forms a homodimer with each polypeptide containing 129 residues. It has been crystallised in ammonium sulfate at pH?6. Crystals belong to space group P3121 with cell parameters a?=?70.2?Å and c?=?126.8?Å, which corresponds to a dimer in the asymmetric unit (VM?=?3.5?Å3?/?Da). The crystal structure of RGCP was solved by the molecular replacement method and refined using data to 2.5-Å resolution. The final crystallographic R factor was 17.9% for all reflections (above 2?σ) in the resolution range 27.4 to 2.5?Å. The refined model includes 1876 non-hydrogen protein atoms and 56 water molecules. As typical of c–type cytochromes, the heme group is covalently bound to Cys-X-Y-Cys-His through thio-ether bonds, and His123 occupies the fifth axial coordination position. On the vacant "distal" site, Phe16 blocks the direct access to the sixth coordination site, which is in a predominantly hydrophobic environment. In spite of the low sequence homology among cytochromes c' the overall fold is similar. The monomer structure consists of 4 anti-parallel α-helices and has random coils in the loops between the helices, and at the N- and C-termini. The subunits cross each other to form an X shape.  相似文献   

3.
Serum albumin, a protein naturally abundant in blood plasma, shows remarkable ligand binding properties of numerous endogenous and exogenous compounds. Most of serum albumin binding sites are able to interact with more than one class of ligands. Determining the protein‐ligand interactions among mammalian serum albumins is essential for understanding the complexity of this transporter. We present three crystal structures of serum albumins in complexes with naproxen (NPS): bovine (BSA‐NPS), equine (ESA‐NPS), and leporine (LSA‐NPS) determined to 2.58 Å (C2), 2.42 Å (P61), and 2.73 Å (P212121) resolutions, respectively. A comparison of the structurally investigated complexes with the analogous complex of human serum albumin (HSA‐NPS) revealed surprising differences in the number and distribution of naproxen binding sites. Bovine and leporine serum albumins possess three NPS binding sites, but ESA has only two. All three complexes of albumins studied here have two common naproxen locations, but BSA and LSA differ in the third NPS binding site. None of these binding sites coincides with the naproxen location in the HSA‐NPS complex, which was obtained in the presence of other ligands besides naproxen. Even small differences in sequences of serum albumins from various species, especially in the area of the binding pockets, influence the affinity and the binding mode of naproxen to this transport protein. Proteins 2014; 82:2199–2208. © 2014 Wiley Periodicals, Inc.  相似文献   

4.
Cavalier MC  Kim SG  Neau D  Lee YH 《Proteins》2012,80(4):1143-1153
The molecular basis of fructose‐2,6‐bisphosphatase (F‐2,6‐P2ase) of 6‐phosphofructo‐2‐kinase/fructose‐2,6‐bisphosphatase (PFKFB) was investigated using the crystal structures of the human inducible form (PFKFB3) in a phospho‐enzyme intermediate state (PFKFB3‐P?F‐6‐P), in a transition state–analogous complex (PFKFB3?AlF4), and in a complex with pyrophosphate (PFKFB3?PPi) at resolutions of 2.45, 2.2, and 2.3 Å, respectively. Trapping the PFKFB3‐P?F‐6‐P intermediate was achieved by flash cooling the crystal during the reaction, and the PFKFB3?AlF4 and PFKFB3?PPi complexes were obtained by soaking. The PFKFB3?AlF4 and PFKFB3?PPi complexes resulted in removing F‐6‐P from the catalytic pocket. With these structures, the structures of the Michaelis complex and the transition state were extrapolated. For both the PFKFB3‐P formation and break down, the phosphoryl donor and the acceptor are located within ~5.1 Å, and the pivotal point 2‐P is on the same line, suggesting an “in‐line” transfer with a direct inversion of phosphate configuration. The geometry suggests that NE2 of His253 undergoes a nucleophilic attack to form a covalent N? P bond, breaking the 2O? P bond in the substrate. The resulting high reactivity of the leaving group, 2O of F‐6‐P, is neutralized by a proton donated by Glu322. Negative charges on the equatorial oxygen of the transient bipyramidal phosphorane formed during the transfer are stabilized by Arg252, His387, and Asn259. The C‐terminal domain (residues 440–446) was rearranged in PFKFB3?PPi, implying that this domain plays a critical role in binding of substrate to and release of product from the F‐2,6‐P2ase catalytic pocket. These findings provide a new insight into the understanding of the phosphoryl transfer reaction. Proteins 2012; © 2011 Wiley Periodicals, Inc.  相似文献   

5.
The interaction of water molecules with copper in wild-type azurin and different site-directed mutants of the coordinated residues is studied by nuclear magnetic relaxation dispersion. Different degrees of solvent accessibility are found. The low relaxivity of wild-type azurin agrees with a solvent-protected copper site in solution, the closest water being found at a distance of more than 5?Å from the copper. This low relaxivity contrasts with the relatively large relaxivity of the His46Gly and His117Gly azurin mutants, which shows clear evidence of copper-coordinated water. The data on the latter mutants are best analyzed in terms of one and two water molecules coordinated to the copper in His46Gly and His117Gly, respectively. The Met121His azurin mutant shows an intermediate behavior. The data are analyzed in terms of an increased solvent accessibility with respect to the wild-type azurin, resulting in semi-coordination of water at low pH. These different modes of coordination lead to different geometries, ranging from the trigonal type 1 site of wild-type azurin to the tetragonal type 2 copper sites of the His117Gly and His46Gly azurin mutants through a so-called type 1.5 site of the Met121His mutant. A correlation is found between the relaxation time (τs) of the unpaired electron of copper(II) and the geometry of the metal site: as the tetragonal character decreases the relaxation becomes significantly faster. τs values of ≤1?ns are found for the tetrahedrally distorted type 1 and type 1.5 sites and of 5–15?ns for the tetragonal type 2 sites.  相似文献   

6.
Three coordination compounds of formula {M(bmim)2Cl2} were synthetised (M?=?Co, Zn, and Hg) and fully characterised. Each complex incorporates 1-benzyl-2-methylimidazole (bmim) as ligand. The coordination polyhedron around the metal center for all complexes has a quasi-regular tetragonal geometry. Density functional theory calculations were carried out on the title compounds and as well on hypothetical complexes (Cu, Ni), in order to elucidate their electronic and molecular structure. The calculations reproduced the Co, Zn, and Hg experimental structures and could predict stable complexes in the case of Ni(II) and Cu(II) ions. The carbonic anhydrase (CA, EC 4.2.1.1) inhibitory effects of the three complexes were investigated. Only compound {Hg(bmim)2Cl2} (3) exhibited a modest inhibitory effect against hCA I, probably due to the affinity of Hg(II) for His residues at the entrance of the active site cavity.  相似文献   

7.
《Inorganica chimica acta》1988,151(4):243-248
The interactions of dimeric complex bis-[μ-chloro-chlorotricarbonylruthenium(II)], [Ru(CO)3Cl2]2, and the polymeric complex poly-[μ-dichlorodicarbonylruthenium(II)], [Ru(CO)2Cl2]x, with nucleosides (Nucl) in a 1:1 Ru:Nucl molar ratio for the dimer and 1:2 Ru:Nucl for the polymer, resulted in formation of the monomeric mononucleoside [Ru(CO)3(Nucl)Cl2] and bis-nucleoside [Ru(CO)2(Nucl)2Cl2] complexes, respectively. The dimer [Ru(CO)3Cl2]2 also gave the ionic bis-nucleoside complexes [Ru(CO)3(Nucl)2Cl]Cl in the molar ratio 1:2 Ru:Nucl. The mononucleoside complexes are stable in solution while the bis-nucleoside complexes tend to lose one nucleoside in strong complexing solvents, probably by solvent substitution. The complexes [Ru(CO)3(Nucl)Cl2] and [Ru(CO)2(Nucl)2Cl2] with one N(1)H ionizable imino proton undergo ionization in alkaline solution and the complexes [Ru(CO)3(NuclH+)Cl] and [Ru(CO)2(NuclH+)2], respectively, were isolated. In these deprotonated complexes the nucleosides behave as bidentate ligands, while in the protonated ones they act as monodentate. All Complexes were characterized by elemental analyses and various spectroscopic methods.  相似文献   

8.
The crystal structures of dithionite-reduced bovine Cu(I),Zn superoxide dismutase and of its adducts with the inorganic anions azide and thyocyanide have been determined in a C2221 crystal form obtained at pH?5.0. This crystal form is characterized by a high solvent content (72%) and by having the two Cu,ZnSOD monomers (A and B) in different crystal environments. One of them (B) is involved in few intermolecular crystal contacts so that it is in a more "solution like" environment, as indicated by average temperature factors which are about twice those of the other monomer. The differences in crystal packing affect the active site structures. While in the A monomer the Cu(I) is coordinated to all four histidine residues, in the B monomer the bridging His61 side chain is found disordered, implying partial detachment from copper. The same effect occurs in the structures of the anion complexes. The inorganic anions are found bound in the active site cavity, weakly interacting with copper at distances ranging from 2.5 to 2.8?Å. The copper site in the A subunit of the native enzyme structure displays significant electron density resembling a diatomic molecule, bound side-on at about 2.8?Å from the metal, which cannot be unambiguously interpreted. The crystallographic data suggest that the existence of the His61 bridge between copper and zinc is dominated by steric more than electronic factors and that the solution state favors the His61 detachment. These structures confirm the existence of an energetically available state for Cu(I) in Cu,ZnSOD where the histidinato bridge to zinc is maintained. This state appears to be favored by tighter crystal contacts. The binding of the anions in the active site cavity is different from that observed in the oxidized enzyme and it appears to be dominated by electrostatic interactions within the cavity. The anion binding mode observed may model the substrate interaction with the reduced enzyme during catalysis.  相似文献   

9.
A hypothesis is proposed to explain the function of Cl- in activating the oxygenevolving complex (OEC) of photosystem II (PS II), based on the results of recent 35Cl-NMR studies. The putative mechanism involves Cl- binding to two types of sites. An intrinsic site is suggested to be composed of three histidyl residues (His 332 and His 337 from D1 and His 337 D2). It is proposed that Cl- binding to this site accelerates the abstraction of H+ from water by raising the pKa's of the histidine imidazole groups. Cl- binding also stimulates the transfer of H+ from this intrinsic site to a set of extrinsic sites on the 33 kD extrinsic polypeptide. The extrinsic Cl- binding sites are suggested to involve four protein domains that are linked together by salt-bridge contacts. Chloride and H+ donated from the intrinsic site attack these intramolecular salt-bridges in a defined sequence, thereby exposing previously inaccessible Cl- and H+ binding sites and stimulating the oxidation of water. This hypothesis also proposes a possible structure for the Mn active site within the D1/D2 complex. Specific amino-acid residues that are likely to participate as Mn lignads are identified on the lumenal portions of the D1 and D2 proteins that are different from those in the L and M subunits of photosynthetic bacteria; the choice of these residues is based on the metal coordination chemistry of these residues, their location within the polypeptide chain, the regularity of their spacing, and their conservation through evolution. The catalytic Mn-binding residues are suggested to be D-61, E-65, E-92, E-98, D-103; D-308, E-329, E-342 and E-333 in D1, and H-62, E-70, H-88, E-97, D-101; E-313, D-334, E-338 and E-345 in D2. Finally, this hypothesis identifies sites on both D2 and the 33 kD extrinsic polypeptide that might be involved in high- and low-affinity Ca2+ binding.To whom correspondence should be addressed  相似文献   

10.
Difference Fourier maps of azide, thiocyanate and perchlorate complexes of methemerythrin from Themiste dyscritum have been calculated at 4 Å, 3.5 Å and 3.5 Å resolution, respectively, with phases from a refined model. N?3 and SCN? bind to the Fe complex in each subunit, indicating the mode of oxygen binding and suggesting a possible route followed by the anions in reaching the complex. ClO?4 binds in two different locations on the non-crystallographic 2-fold axes near cysteine 9 and cysteine 50, apparently being held to the protein by peptide amides and lysine side-chains. Binding of ClO?4 at these locations away from the Fe complex is observed to have some effect on the active site structure.  相似文献   

11.
Equilibrium constants for substitution of Cl? by cytidine in PdCl42? and Pd(en)Cl2 in aqueous solution were determined. Equilibrium constants for substitution of Cl? by pyridine in Pd(en)Cl2 were also measured. For unit ionic strength at 25°C, log K values for the successive substitutions of two chlorides are: PdCl42? -cytidine, log K1 = 4.49, log K2 = 3.45; Pd(en)Cl2-cytidine, log K1 = 3.32, log K2 = 2.56; Pd(en)Cl2-pyridine, log K1 = 4.31, log K2 = 3.15. Comparison of equilibrium constants indicates that intramolecular hydrogen bonding and steric interactions of the exocyclic substituents ortho to the binding site of cytidine do not contribute significantly to the stability of the palladium-nucleoside complexes. Coordination equilibria, of deprotonated cytidine occur in alkaline solutions of PdCl42? and the ligand.  相似文献   

12.
Intercalation complexes of daunomicin(+1) with tetramer duplexes in DNA are studied with the theoretically determined intercalation sites (I, ?0.4), (II, ?0.4), and (III, ?1.4). These sites occur with base pairs separated by 6.76 Å for helical angles of 26°, 22°, and 8° about the intercalation site. Site I is preferred, and this is in agreement with experimental unwinding angles. Optimum binding positions and conformations are established, and these are in agreement with experimental results from crystal structures. A systematic procedure is devised to study base-pair and base-sequence specificity, which results in the demonstration that the most stable sequences are mainly ↑BP1, T·A, DAUN, A·T, BP4↓ and ↑BP1, T·A, DAUN, G·C, BP4↓, i.e., with the TpA and CpG (pyrimidine)p(purine) sequences about the intercalation site. These 32 possible sequences are found among the 40 most stable complexes. These theoretical calculations of intercalation complexes with daunomicin(+1) provide the first example in which a drug specifically selects the base pair T·A and prefers it in a particular sequence about the intercalation site. This specificity is in agreement with some experimental results. Problems associated with the interpretation of specificity are discussed in terms of the base, base-pair, and base-sequence resulting from the DNA site and the DNA–drug interactions. T·A specificity is rationalized by noting that the 2′deoxyribo-5′-monophosphate backbone attached to A is slightly more negative than that on the other nucleotides. Hence, a preference exists for binding to the protonated daunosamine (+1) groups. Stereographic projections of daunomycinone and daunomycin(+1) in a bond model and in a space-filling model with steric contours illustrate the results.  相似文献   

13.
The binding of VIVO2+ to human serum transferrin (hTF) at the FeIII binding sites is addressed. Geometry optimization calculations were performed for the binding of VIVO2+ to the N-terminal lobe of hTF (hTFN), and indicate that in the presence of CO3 2? or HCO3 ?, VIV is bound to five atoms in a distorted geometry. The structures of VIVO–hTFN species optimized at the semiempirical level were also used to calculate the 51V and 14N A tensors by density functional theory methods, and were compared with the reported experimental values. Globally, of all the calculated VIVO–hTF structures, the one that yields the lowest calculated heats of formation and minimum deviations from the experimental values of the 51V and 14N A tensor components is the structure that includes CO3 2? as a synergistic anion. In this structure the V=O bond length is approximately 1.6 Å, and the vanadium atom is also coordinated to the phenolate oxygen atom of Tyr188 (at approximately 1.9 Å), the aspartate oxygen atom of Asp63 (at approximately 1.9 Å), the His249 Nτ atom (at approximately 2.1 Å), and a carbonate oxygen atom (at approximately 1.8 Å). The Tyr95 phenolic ocygen atom is approximately 3.3 Å from the metal center, and thus is very weakly bound to VIV. All of these oxygen atoms are able to establish dipolar interactions with groups of the protein.  相似文献   

14.
A density functional theory (DFT) investigation into the structures and bonding characteristics of [(B3O3H3)nM]+(n?=?1, 2;M?=?Cu, Ag, Au) complexes was performed. DFT calculations and natural bond orbital (NBO) analyses indicate that the ΙB metal complexes of boroxine exhibit intriguing bonding characteristics, different from the typical cation–π interactions between ΙB metal-cations and benzene. The complexes of [B3O3H3M]+ and [(B3O3H3)2?M]+ (M?=?Cu, Ag, and Au) favor the conformation of perfectly planar structures with the C2v and D2h symmetry along one of the threefold molecular axes of boroxine, respectively. Detailed natural resonance theory (NRT) and canonical molecular orbitals (CMOs) analyses show that interaction between the metal cation and the boroxine in [B3O3H3M]+ (M?=?Cu, Ag, and Au) is mainly ionic, while the ΙB metal-cations←π donation effect is responsible for the binding site. In these complexes, boroxine serves as terminals η1-B3O3H3 with one O atom of the B3O3 ring. The infra-red (IR) spectra of [B3O3H3M]+ were simulated to facilitate their future experimental characterization. The complexes all give two IR active modes at about 1,300 and 2,700 cm?1, which are inactive in pure boroxine. Simultaneously, the B–H stretching modes of the complexes are red-shifted due to the interaction between the metal-cation and boroxine. To explore the possibility of the structural pattern developed in this work forming mesoporous materials, complexes [(B3O3H3M)6]6+ (M?=?Cu, Ag, and Au) were also studied, which appear to be unique and particular interesting: they are all true minima with D6h symmetries and pore sizes ranging from 12.04 Å to 13.65 Å.
A density functional theory investigation on the structures and bonding characteristics of [(B3O3H3)nM]+(n?=?1, 2;M?=?Cu, Ag, Au) complexes has been performed. Detailed natural resonance theory (NRT) and canonical molecular orbitals (CMOs) analyses show that the interactions between the metal cation and the boroxine in [B3O3H3M]+ (M?=?Cu, Ag, and Au) are mainly ionic, while the IB metal-cations?←?π donation effect is responsible for the binding site. The [(B3O3H3M)6]6+ (M?=?Cu, Ag, and Au) appear to be unique and are particular interesting: they are all true minima with D6h symmetries and pore sizes ranging from 12.04 Å to 13.65 Å and may be extended to form mesoporous materials.  相似文献   

15.
Abstract

The octahedral Ru(II) complexes containing the 2(2,6-dimethoxypyridine-3-yl)-1H-imidazo(4,5-f)[1, 10]phenanthroline ligand of type [Ru(N-N)2(L)]2+, where N-N?=?phen (1,10-phenanthroline) (1), bpy (2,2'-bipyridine) (2), and dmb (4,4'-dimethyl-2,2'-bipyridine) (3); L(dmpip) = (2(2,6-dimethoxypyridine-3-yl)1Himidazo(4,5-f)[1, 10]phenanthroline), have been synthesized and characterized by UV–visible absorption, molar conductivity, elemental analysis, mass, IR, and NMR spectroscopic techniques. The physicochemical properties of the Ru(II) complexes were determined by UV–Vis absorption spectroscopy. The DNA binding studies have been explored by UV–visible absorption, fluorescence titrations, and viscosity measurements. The supercoiled pBR322 DNA cleavage efficiency of Ru(II) complexes 1–3 was investigated. The antimicrobial activity of Ru(II) complexes was done against Gram-positive and Gram-negative microorganisms. The in vitro anticancer activities of all the complexes were investigated by cell viability assay, apoptosis, cellular uptake, mitochondrial membrane potential detection, and semi-quantitative PCR on HeLa cells. The result indicates that the synthesized Ru(II) complexes probably interact with DNA through an intercalation mode of binding with complex 1 having slightly stronger DNA binding affinity and anticancer activity than 2 and 3.  相似文献   

16.
Gamma-aminobutyric type A receptor (GABAAR) is a member of the Cys-loop family of pentameric ligand gated ion channels (pLGICs). It has been identified as a key target for many clinical drugs. In the present study, we construct the structure of human 2α12γ2 GABAAR using a homology modeling method. The structures of ten benzodiazepine type drugs and two non-benzodiazepine type drugs were then docked into the potential benzodiazepine binding site on the GABAAR. By analyzing the docking results, the critical residues His102 (α1), Phe77 (γ2) and Phe100 (α1) were identified in the binding site. To gain insight into the binding affinity, molecular dynamics (MD) simulations were performed for all the receptor–ligand complexes. We also examined single mutant GABAAR (His102A) in complexes with the three drugs (flurazepam, eszopiclone and zolpidem) to elucidate receptor–ligand interactions. For each receptor–ligand complex (with flurazepam, eszopiclone and zolpidem), we calculated the average distance between the Cα of the mutant residue His102A (α1) to the center of mass of the ligands. The results reveal that the distance between the Cα of the mutant residue His102A (α1) to the center of flurazepam is larger than that between His102 (α1) to flurazepam in the WT type complex. Molecular mechanic-generalized Born surface area (MM-GBSA)-based binding free energy calculations were performed. The binding free energy was decomposed into ligand-residue pairs to create a ligand-residue interaction spectrum. The predicted binding free energies correlated well (R 2?=?0.87) with the experimental binding free energies. Overall, the major interaction comes from a few groups around His102 (α1), Phe77 (γ2) and Phe100 (α1). These groups of interaction consist of at least of 12 residues in total with a binding energy of more than 1 kcal mol?1. The simulation study disclosed herein provides a meaningful insight into GABAAR–ligand interactions and helps to arrive at a binding mode hypothesis with implications for drug design.  相似文献   

17.
The reaction of ruthenium carbonyl polymer ([Ru(CO)2Cl2]n) with azopyridyl compounds (2,2′-azobispyridine; apy or 2-phenylazopyridine; pap) generated new complexes, [Ru(azo)(CO)2Cl2] (azo = apy, pap). [Ru(apy)(CO)2Cl2] underwent photodecarbonylation to give a chloro-bridged dimer complex, whereas the corresponding pap complex ([Ru(pap)(CO)2Cl2]) was not converted to a dimer. The reactions of the chloro-bridged dimer containing the bpy ligand (bpy = 2,2′-bipyridine) with either apy or pap resulted in the formation of mixed polypyridyl complexes, [Ru(azo)(bpy)(CO)Cl]+. The novel complexes containing azo ligands were characterized by various spectroscopic measurements including the determination of X-ray crystallographic structures. Both [Ru(azo)(CO)2Cl2] complexes have two CO groups in a cis position to each other and two chlorides in a trans position. The azo groups are situated cis to the CO ligand in [Ru(azo)(bpy)(CO)Cl]+. All complexes have azo N-N bond lengths of 1.26-1.29 Å. The complexes exhibited azo-based two-electron reduction processes in electrochemical measurements. The effects of introducing azopyridyl ligands to the ruthenium carbonyl complexes were examined by ligand-based redox potentials, stretching frequencies and force constants of CO groups and bond parameters around Ru-CO moieties.  相似文献   

18.
A site-directed mutant of spinach plastocyanin, Pc(Tyr83-His), has been modified by covalent attachment of a photoactive [Ru(bpy)2(im)]2+ complex to the His83 residue. The residue is surface exposed and located about 10–12?Å from the copper ion at the entrance of a proposed natural electron transfer pathway from cytochrome f. Electron transfer within the Ru-Pc complex has been studied with time-resolved optical spectroscopy using two different approaches. In the first, the fully reduced [Cu(I), Ru(II)] protein was photoexcited and subsequently oxidized by an external quencher, forming the [Cu(I), Ru(III)] protein. This was followed by an electron transfer from reduced Cu(I) to Ru(III). In the second method, the initially oxidized Cu(II) ion acted as an internal quencher for excited Ru(II) and the photoinduced reduction of the Cu(II) ion was followed by a thermal recombination with the Ru(III) ion. The reoxidation of the Cu ion, which has an estimated driving force of 0.56?eV, occured with a rate constant k et?=?(9.5±1.0)×106?s–1, observed with both methods. The results suggest a strong electronic coupling (H DA>0.3?cm–1) along the Ru-His(83)-Cys(84)-Cu pathway.  相似文献   

19.
A 6Åresolution electron density map of crystals of penicillopepsin, an acid protease from Penicillium janthinellum, has been computed from multiple isomorphous replacement phases determined from two heavy metal derivatives, K2PtCl6 and UO2Cl2. The mean figure of merit of the map is 0.939. The boundaries of the molecules, of which there are four per unit cell, are readily discernible. The molecule is highly asymmetric with approximate dimensions 60Å× 40Å× 30Å. The molecule consists of two distinct lobes separated by a deep cleft, which is probably the extended substrate binding site.  相似文献   

20.
Ruthenium compounds are highly regarded as metallo-drug candidates. Many studies have focused their attention on the interaction between ruthenium complexes with their possible biological targets. The interaction of ruthenium complexes with transport proteins, enzymes and peptides is of great importance for understanding their biodistribution and mechanism of action, therefore, the development of an anti-cancer therapy involving ruthenium complexes has recently shifted from DNA targeting towards protein targeting. With the aim of gaining insight into possible interactions between ruthenium complexes with biologically relevant proteins, we have studied the interaction of cis-dichlorobis(2,2′-bipyridyl-4,4′-dicarboxylic acid)ruthenium(II) complex [Ru(II)(dcbpy)2Cl2], which previously showed good potency in photo-dynamic chemotherapy, with bovine serum albumin (BSA), phospholipase A2 (PLA2) and glutathione (GSH). Binding constants and possible number of binding sites to mentioned proteins and peptide are investigated by ultraviolet–visible spectroscopy and Matrix-Assisted Laser Desorption Ionization Mass Spectrometry (MALDI TOF MS). The complex binding affinities were in the following order: PLA2 > BSA > GSH. Moreover, genotoxic profile of the complex, tested on peripheral blood lymphocytes as a model system, was also promising.  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