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.
The NADPH-dependent HC-toxin reductases (HCTR1 and 2) encoded by enzymatic class of disease resistance homologous genes (Hm1 and Hm2) protect maize by detoxifying a cyclic tetrapeptide, HC-toxin, secreted by the fungus Cochliobolus carbonum race 1(CCR1). Unlike the other classes'' resistance (R) genes, HCTR-mediated disease resistance is an inimitable mechanism where the avirulence (Avr) component from CCR1 is not involved in toxin degradation. In this study, we attempted to decipher cofactor (NADPH) recognition and mode of HC-toxin binding to HCTRs through molecular docking, molecular dynamics (MD) simulations and binding free energy calculation methods. The rationality and the stability of docked complexes were validated by 30-ns MD simulation. The binding free energy decomposition of enzyme-cofactor complex was calculated to find the driving force behind cofactor recognition. The overall binding free energies of HCTR1-NADPH and HCTR2-NADPH were found to be −616.989 and −16.9749 kJ mol−1 respectively. The binding free energy decomposition revealed that the binding of NADPH to the HCTR1 is mainly governed by van der Waals and nonpolar interactions, whereas electrostatic terms play dominant role in stabilizing the binding mode between HCTR2 and NADPH. Further, docking analysis of HC-toxin with HCTR-NADPH complexes showed a distinct mode of binding and the complexes were stabilized by a strong network of hydrogen bond and hydrophobic interactions. This study is the first in silico attempt to unravel the biophysical and biochemical basis of cofactor recognition in enzymatic class of R genes in cereal crop maize.  相似文献   

2.
The delivery of drugs to the brain is complicated by the multiple factors including low blood–brain barrier (BBB) passive permeability, active BBB efflux systems, and plasma protein binding. Thus, a detailed understanding of the transport of the new potent substances through the membranes is vitally important and their physico-chemical characteristics should be analyzed at first. This work presents an evaluation of drug likeness of eight 7-O-arylpiperazinylcoumarin derivatives with high affinity towards serotoninergic receptors 5-HT1A and 5-HT2A with particular analysis of the requirements for the CNS chemotherapeutics. The binding constants to human serum albumin (HSA) were determined at physiological pH using fluorescence spectroscopy, and then their mode of action was explained by analysis of theoretical HSA complexes. Dynamic simulation of systems allowed for reliable evaluation of the interaction strength. The analyzed coumarins were able to pass BBB, and they present good drug likeness properties. They showed high affinities to HSA (log KQ = 5.3–6.0 which corresponds to −8.12 to −7.15 kcalmol−1 of Gibbs free energy). The changes of the emission intensity upon binding to HSA were scrutinized showing the different mode of action for 4-phenylpiperazinylcoumarins. The values of computed Gibbs free energy and determined on the basis of experimentally obtained binding constants log KQ coincide suggesting a good quality of the theoretical model. Overall the 8-acetyl-7-O-arylpiperazinyl-4-methylcoumarin derivatives represent valuable lead compounds to be further tested in various preclinical assays as a possible chemotherapeutics against CNS diseases. Studied coumarins can be metabolized by cytochrome P450 to aldehydes and hydroxy derivatives. The existence of other binding sites inside HSA than Sudlow’s site 1 was postulated. The longer aliphatic linker between coumarin and piperazine moieties favored binding to HSA in other than Sudlow site 1 pocket.  相似文献   

3.
To prepare a new antiviral duplex drug linking Zidovudine (AZT) and Foscarnet (PFA) via a lipophilic octadecylglycerol residue we condensed 1-O-4-monomethoxytrityl-3-O-octadecyl-sn-glycerol-2-hydrogenphosphonate obtained from 3-O-octadecyl-sn-glycerol with AZT by the phosphonate method. The purified condensation product was de-tritylated resulting in 3′-azido-3′-deoxythymidylyl-(5′  2-O)-3-O-octadecyl-sn-glycerol, followed by treatment with (ethoxycarbonyl)phosphoric dichloride. The resulting 3′-azido-3′-deoxy-thymidylyl-(5′  2)-3-O-octadecyl-sn-glycerol-1-O-(ethoxycarbonyl)phosphonate was purified by preparative RP-18 column chromatography. The antiviral duplex drug 3′-azido-3′-deoxythymidylyl-(5′  2-O)-3-O-octadecyl-sn-glycerol-1-O-phosphonoformate trisodium salt (AZT–lipid–PFA) was obtained after alkaline cleavage of the phosphonoformate ethylester residue. The overall yield of the five step synthesis performed at gram scale was about 30%. According to a supposed pathway AZT–lipid–PFA could be cleaved to yield a mixture of different antiviral compounds such as AZT, AZT-5′-monophosphate, octadecylglycerol–AZT, PFA and octadecylglycerol–PFA, possibly producing additive and/or synergistic antiviral effects. In vitro studies showed that the duplex drug exhibits antiviral activities against HIV and especially against drug-resistant strains and clinical isolates of HSV and HCMV. The E50 values of AZT–lipid–PFA against HIV ranged between 170 and 200 nM. The half-maximal inhibitory doses (IC50) against highly acyclovir (ACV)-resistant HSV isolates determined by a plaque reduction assay ranged between 1.87 and 4.59 μM. Using ganciclovir (GCV)-sensitive, GCV resistant and drug cross-resistant HCMV strains the IC50-values of AZT–lipid–PFA were between 2.78 and 1.18 μM. With regard to PFA, the IC50-value of AZT–lipid–PFA determined on a multi-drug-resistant HCMV strain was about 90-fold lower than that of PFA, demonstrating the superior antiviral effect of the duplex-drug.  相似文献   

4.
Leishmaniasis is one of the most important neglected tropical diseases (NTDs) that are especially common among low-income populations in developing regions of Africa, Asia, and the Americas. Many natural products, particularly alkaloids, have been reported to have inhibitory activity against arginase, the key enzyme in the pathology caused by Leishmania sp. In this way, piperidine alkaloids (–)-cassine (1), (–)-spectaline (2), (–)-3-O-acetylcassine (3), and (–)-3-O-acetylspectaline (4) were isolated from Senna spectabilis flowers. These compounds (1/2 and 3/4) initially present as homologous mixtures were separated by high performance liquid chromatography and evaluated against the promastigote phase of Leishmania amazonensis. In addition, molecular docking simulations were implemented in order to probe the binding modes of the ligands 14 to the amino acids in the active site of L. amazonensis arginase. Alkaloid 2 (IC50 15.81?μg?mL?1) was the most effective against L. amazonensis. Compounds 2 and 4, with larger side chain, were more effective against the parasite than compounds 1 and 3. The cell viability test on Vero cells revealed that compound 2 (CC50 66.67?μg?mL?1) was the most toxic. The acetyl group in the 3-O position of the parent structures reduced the leishmanicidal activity and the toxicity of the alkaloids. Further, molecular docking suggested that Asn143 is essential for arginase to interact with (–)-spectaline-derived compounds, which agreed with the IC50 measurements. Our findings revealed that S. spectabilis is an important source of piperidine alkaloids with leishmanicidal activity. Moreover, the natural compound 3 has been isolated for the first time. Experimental investigation combined with theoretical study advances knowledge about the enzyme binding site mode of interaction and contributes to the design of new bioactive drugs against Leishmania infection.  相似文献   

5.
Three covalent anthocyanin–flavonol complexes (pigments 1–3) were extracted from the violet-blue flower of Allium ‘Blue Perfume’ with 5% acetic acid-MeOH solution, in which pigment 1 was the dominant pigment. These three pigments are based on delphinidin 3-glucoside as their deacylanthocyanin and were acylated with malonyl kaempferol 3-sophoroside-7-glucosiduronic acid or malonyl-kaempferol 3-p-coumaroyl-tetraglycoside-7-glucosiduronic acid in addition to acylation with acetic acid.By spectroscopic and chemical methods, the structures of these three pigments 1–3 were determined to be: pigment 1, (6I-O-(delphinidin 3-O-(3I-O-(acetyl)-β-glucopyranosideI)))(2VI-O-(kaempferol 3-O-(2II-O-(3III-O-(β-glucopyranosylV)-β-glucopyranosylIII)-4II-O-(trans-p-coumaroyl)-6II-O-(β-glucopyranosylIV)-β-glucopyranosideII)-7-O-(β-glucosiduronic acidVI))) malonate; pigment 2, (6I-O-(delphinidin 3-O-(3I-O-(acetyl)-β-glucopyranosideI)))(2VI-O-(kaempferol 3-O-(2II-O-β-glucopyranosylIII)-β-glucopyranosideII)-7-O-(β-glucosiduronic acidVI))); and pigment 3, (6I-O-(delphinidin 3-O-(3I-O-(acetyl)-β-glucopyranosideI)))(2VI-O-(kaempferol 3-O-(2II-O-(3III-O-(β-glucopyranosylV)-β-glucopyranosylIII)-4II-O-(cis-p-coumaroyl)-6II-O-(β-glucopyranosylIV)-β-glucopyranosideII)-7-O-(β-glucosiduronic acidVI))) malonate.The structure of pigment 2 was analogous to that of a covalent anthocyanin–flavonol complex isolated from Allium schoenoprasum where delphinidin was observed in place of cyanidin. The three covalent anthocyanin–flavonol complexes (pigment 1–3) had a stable violet-blue color with three characteristic absorption maxima at 540, 547 and 618 nm in pH 5–6 buffer solution. From circular dichroism measurement of pigment 1 in the pH 6.0 buffer solution, cotton effects were observed at 533 (+), 604 (−) and 638 (−) nm. Based on these results, these covalent anthocyanin–flavonol complexes were presumed to maintain a stable intramolecular association between delphinidin and kaempferol units closely related to that observed between anthocyanin and hydroxycinnamic acid residues in polyacylated anthocyanins. Additionally, an acylated kaempferol glycoside (pigment 4) was isolated from the same flower extract, and its structure was determined to be kaempferol 3-O-sophoroside-7-O-(3-O-(malonyl)-β-glucopyranosiduronic acid).  相似文献   

6.
The title compound, methyl 2-methoxy-7-(4-methylbenzoyl)-4-oxo-6-p-tolyl-4H-furo[3,2-c]pyran-3-carboxylate (C25H20O7), was prepared and characterized by IR and single-crystal X-ray diffraction (XRD). The compound crystallizes in the triclinic space group P ?1 with a?=?8.9554(9) Å, b?=?10.0018(10) Å, c?=?12.7454(13) Å, α?=?67.678(7)°, β?=?89.359(8)° and γ?=?88.961(8)°. In addition to the molecular geometry from X-ray experiment, the molecular geometry and vibrational frequencies of the title compound in the ground state have been calculated using semiempirical AM1 and PM3 methods, as well as Hartree-Fock (HF) and density functional (B3LYP) levels of theory with 6–31G(d) basis set. To determine conformational flexibility, molecular energy profile of the title compound was obtained by semi-empirical (AM1) calculations with respect to two selected degrees of torsional freedom, which were varied from ?180° to +180° in steps of 10°. Besides, frontier molecular orbitals (FMO) analysis and thermodynamic properties of the title compound were performed by the B3LYP/6–31G(d) method.  相似文献   

7.
Phytochemical research of Pteris multifida Poir. led to the isolation of fifteen compounds, including six flavonoids (16) and nine sesquiterpenoids (715). Their structures were characterized by NMR, MS, ORD and CD data. Compounds kaempferol 3-O-α-L-rhamnoside-7-O-β-D-glucoside (1), myricetin 3-O-β-D-glucoside (2), kaempferol 3-O-β-D-glucoside (4), luteolin-7-O-β-D-rutinoside (5), quercetin-3-O-α-L-rhamnopyranoside (6), (2S,3S)-12-hydroxypterosin Q (7), (2S,3S)-pterosin Q (8), 2-hydroxypterosin C (9) and (2S)-12-hydroxypterosin A (10) were first isolated from P. multifida, and compounds 12 and 10 were first isolated from the family Pteridaceae. Furthermore, the chemotaxonomic significance of the isolates was discussed.  相似文献   

8.
3-O-Benzyl-1,2-O-isopropylidene-α-D-xylo-pentodialdo-1,4-furanose (1) was found to give, with nitromethane under catalysis by sodium methoxide, 3-O-benzyl-6-deoxy-1,2-O-isopropylidene-6-nitro- α-D-glucofuranose (2) as the kinetically favored product. Subsequent, spontaneous epimerization led to a 2:1 mixture of 2 and its β-L-ido isomer (3), from which crystalline 3 was isolated. The free nitro hexoses (4 and 5) obtained by deacetonation of 2 and 3 were subjected to barium hydroxide-catalyzed cyclization (internal Henry reaction) to give mixtures of O-benzyldeoxynitroinositols. Under conditions of kinetic control, the α-D-gluco derivative 4 furnished 6-O-benzyl-3-deoxy-3-nitro-muco-inositol (6) and optically active 4-O-benzyl-1-deoxy-1-nitro-L-myo-inositol (L-7) in a ratio of 3:1. The β-L-ido derivative 5 gave the enantiomer (D-7) of the myo compound and 4-O-benzyl-1-deoxy-1-nitro-scyllo-inositol (8) in a similar ratio. Slow, thermodynamically controlled epimerization led from each individual nitro inositol to mixtures of the same composition, with 17–18% of 6, 68–69% of DL-7, and 11–12% of 8. All of the nitroinositol benzyl ethers were isolated crystalline and characterized further as crystalline tetraacetates (6a–8a). The muco isomer 6 gave a di-O-isopropylidene derivative (6b).  相似文献   

9.
Designing selective protein kinase B (PKB/Akt) inhibitor is an area of intense research to develop potential anticancer drugs. In the present study, the molecular basis governing PKB-selective inhibition has been investigated using molecular dynamics simulation. The binding free energies calculated by MM/PBSA gave a good correlation with the experimental biological activity and a good explanation of the activity difference of the studied inhibitors. The decomposition of free energies by MM/GBSA indicates that the ethyl group on pyrrolo[2,3-d]pyrimidine ring of inhibitor Lig1 (N-{[(3S)-3-amino-1-(5-ethyl-7H-pyrrolo[2,3-d]pyrimidin-4-yl)pyrrolidin-3-yl]-methyl}-2,4-difluoro-benzamide) is an important contributor to its PKBα selectivity due to its hydrophobic interaction with the side chain of Thr291 in PKBα. The substituted groups on the pyrrolidine ring of Lig1 also show a strong tendency to mediate protein-ligand interactions through the hydrogen bonds formed between the amino or amide groups of Lig1 and the carboxyl O atoms of Glu234, Glu278, and Asp292 of PKBα. It was reported that there are only three key amino acid differences between PKBα (Thr211, Ala230, Met281) and PKA (Val104, Val123, Leu173) within the clefts of ATP-binding sites. These differences propel a drastic conformational change in PKA, weakening its binding interactions with inhibitor. The impact was also confirmed by MD simulated interaction modes of inhibitor binding to PKBα mutants with the in silico mutations of the three key amino acids, respectively. We expect that the results obtained here could be useful for future rational design of specific ATP-competitive inhibitors of PKBα.  相似文献   

10.
Four new flavonoid glycosides, curcucomosides A–D (14), three known flavonoid glycosides, 57, and four known diarylheptanoids, 811, were isolated from the ethanol extract of the aerial parts of Curcuma comosa. The structures of the new compounds were established as rhamnazin 3-O-α-l-arabinopyranoside (1), rhamnocitrin 3-O-α-l-arabinopyranoside (2), rhamnazin 3-O-α-l-rhamnopyranosyl-(1→2)-O-α-l-arabinopyranoside (3), and rhamnocitrin 3-O-α-l-rhamnopyranosyl-(1→2)-O-α-l-arabinopyranoside (4) by spectroscopic analysis and chemical reactions whereas those of the known compounds were identified by spectral comparison with those of the reported values.  相似文献   

11.
The sliding and hopping models encapsulate the essential protein-DNA binding process for binary complex formation and dissociation. However, the effects of a cofactor protein on the protein-DNA binding process that leads to the formation of a ternary complex remain largely unknown. Here we investigate the effect of the cofactor Sox2 on the binding and unbinding of Oct1 with the Hoxb1 control element. We simulate the association of Oct1 with Sox2-Hoxb1 using molecular dynamics simulations, and the dissociation of Oct1 from Sox2-Hoxb1 using steered molecular dynamics simulations, in analogy to a hopping event of Oct1. We compare the kinetic and thermodynamic properties of three model complexes (the wild-type and two mutants) in which the Oct1-DNA base-specific interactions or the Sox2-Oct1 protein-protein interactions are largely abolished. We find that Oct1-DNA base-specific interactions contribute significantly to the total interaction energy of the ternary complex, and that nonspecific Oct1-DNA interactions are sufficient for driving the formation of the protein-DNA interface. The Sox2-Oct1 protein-protein binding interface is largely hydrophobic, with remarkable shape complementarity. This interface promotes the formation of the ternary complex and slows the dissociation of Oct1 from its DNA-binding site. We propose a simple two-step reaction model of protein-DNA binding, called the tethered-hopping model, that explains the importance of the cofactor Sox2 and may apply to similar ternary protein-DNA complexes.  相似文献   

12.
A new soluble cyclodextrin derivative 6-O-(2-hydroxybutyl)-β-cyclodextrin (6-HB-β-CD) was prepared. Its molecular binding and recognition ability were investigated with the comparison of β-cyclodextrin (β-CD), 2-O-(2-hydroxypropyl)-β-cyclodextrin (2-HP-β-CD), 6-O-(2-hydroxypropyl)-β-cyclodextrin (6-HP-β-CD), and 2-O-(2-hydroxybutyl)-β-cyclodextrin (2-HB-β-CD). The relationship between the complex stability constants and the possible structures of inclusion compounds was discussed with the interaction of hosts and guests, including the weak hydrophobic interactions, the size/shape matching, the steric hindrance, and the hydrophilic property.  相似文献   

13.
Magnetic core/shell Fe3O4/Au nanoparticles were used in the determination of drug binding to bovine serum albumin (BSA) using a fluorescence spectroscopic method. The binding constants and number of binding sites for protein with drugs were calculated using the Scatchard equation. Because of their superparamagnetic and biocompatible characteristics, magnetic core/shell Fe3O4/Au nanoparticles served as carrier proteins for fixing proteins. After binding of the protein to a drug, the magnetic core/shell Fe3O4/Au nanoparticles–protein–drug complex was separated from the free drug using an applied magnetic field. The free drug concentration was obtained directly by fluorescence spectrometry and the proteins did not influence the drug determination. So, the achieved number of binding sites should be reliable. The binding constant and site number for ciprofloxacin (CPFX) binding to BSA were 2.055 × 105 L/mol and 31.7, and the corresponding values for norfloxacin (NOR) binding to BSA were 1.383 × 105 L/mol and 38.8. Based on the achieved results, a suitable method was proposed for the determination of binding constants and the site number for molecular interactions. The method was especially suitable for studies on the interactions of serum albumin with the active ingredients of Chinese medicine. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

14.
Abstract

Dengue virus (DENV) is one of the most dangerous mosquito-borne human pathogens known to the mankind. Currently, no vaccines or standard therapy is avaliable to treate DENV infection. This makes the drug development against DENV more significant and challenging. The MTase domain of DENV RNA RdRp NS5 is a promising drug target, because this domain hosts the RNA capping process of DENV RNA to escape from human immune system. In the present study, we have analysed the RNA intervention mechanism exerted by flavoniod molecules against NS5 MTase RNA capping site by using molecular docking, molecular dynamics simulation and the binding free energy calculations. The results from the docking analysis confirmed that the RNA intervention mecanism is exerted by the quercetagetin (QGN) molecule with all necessary intermolecular interactions and high binding affinity. Notably, QGN forms strong hydrogen bonding interactions with Asn18, Leu20 and Ser150 residues and π???π stacking interaction with Phe25 residue. The apo and QGN bound NS5 MTase and QGN-NS5 MTase complex were used for MD simulation. The results of MD simulation reveal that the RMSD and RMSF values of QGN-MTase complex have increased on comparing the apo protein due to the effect of ligand binding. The binding free energy calulation includes prediction of total binding free energy of ligand-protein complex and per-residue free energy decomposition. The QGN binding to NS5 MTase affects it’s native motion, this result is found from Principal component analysis.

Communicated by Ramaswamy H. Sarma  相似文献   

15.
We report a series of new 9-oxime ether non-ketolides, including 3-hydroxyl, 3-O-acyl and 3-O-alkyl clarithromycin derivatives, and thiophene-containing ketolides 1b–1d. Unlike previously reported ketolide 1a, none of them is comparable to telithromycin. A molecular modeling study was performed to gain insight into the binding mode of alkylides 1720 with bacterial rRNA and to rationalize the great disparity of their SAR. The 3-O-sidechains of 19 and 20 point to the so-called hydrophilic side of the macrolide ring, as seen in clarithromycin. In contrast, the 3-O-sidechains of 17 and 18 bend to the backside, the so-called hydrophobic side of the macrolide ring. The results clearly indicated the alkylides with improved antibacterial activity might possess a novel binding mode, which is different from clarithromycin and the alkylides with poor activity.  相似文献   

16.
Heparan sulfate (HS) plays a crucial role in the fibrosis associated with chronic allograft dysfunction by binding and presenting cytokines and growth factors to their receptors. These interactions critically depend on the distribution of 6-O-sulfated glucosamine residues, which is generated by glucosaminyl-6-O-sulfotransferases (HS6STs) and selectively removed by cell surface HS-6-O-endosulfatases (SULFs). Using human renal allografts we found increased expression of 6-O-sulfated HS domains in tubular epithelial cells during chronic rejection as compared with the controls. Stimulation of renal epithelial cells with TGF-β induced SULF2 expression. To examine the role of 6-O-sulfated HS in the development of fibrosis, we generated stable HS6ST1 and SULF2 overexpressing renal epithelial cells. Compared with mock transfectants, the HS6ST1 transfectants showed significantly increased binding of FGF2 (p = 0.0086) and pERK activation. HS6ST1 transfectants displayed a relative increase in mono-6-O-sulfated disaccharides accompanied by a decrease in iduronic acid 2-O-sulfated disaccharide structures. In contrast, SULF2 transfectants showed significantly reduced FGF2 binding and phosphorylation of ERK. Structural analysis of HS showed about 40% down-regulation in 6-O-sulfation with a parallel increase in iduronic acid mono-2-O-sulfated disaccharides. To assess the relevance of these data in vivo we established a murine model of fibrosis (unilateral ureteric obstruction (UUO)). HS-specific phage display antibodies (HS3A8 and RB4EA12) showed significant increase in 6-O-sulfation in fibrotic kidney compared with the control. These results suggest an important role of 6-O-sulfation in the pathogenesis of fibrosis associated with chronic rejection.  相似文献   

17.
p-Nitrophenyl 2-O-α-d-galactopyranosyl-α-d-mannopyranoside and p-nitrophenyl 2-O-α-d-glucopyranosyl-α-d-mannopyranoside were synthesized and the interactions of these disaccharides with concanavalin A (con A) were characterized. The kinetics of binding of the galactopyranosyl-containing disaccharide to con A were found to be similar to those observed with monosaccharides in that monophasic time dependencies for binding were observed. The glucopyranosyl-containing disaccharide, however, exhibited biphasic time dependencies which were similar to those previously observed for the binding of p-nitrophenyl 2-O-α-d-mannopyranosyl-α-d-mannopyranoside to con A. These results support a model wherein the α-(1→2)-linked disaccharides which exhibit biphasic binding kinetics must be able to bind to con A in two different and mutually exclusive orientations. The ability to bind to con A in two orientations is shared by α-(1→2)-linked disaccharides in which both glycosyl residues can interact separately with the primary glycosyl binding site of con A. According to the model, the initial fast phase of the biphasic reaction reflects binding of the ligand in two orientations so that two complexes are formed in amounts determined by the relative values of the rate constants for formation of each complex. The subsequent slow phase is proposed to reflect a slow equilibration of the less stable complex to the thermodynamically more stable one. In the more stable complex, the glycosyl residue at the reducing end of the disaccharide occupies the primary glycosyl binding site. The added stability of this complex is attributed to extended interactions between con A and groups on the second glycosyl residue. An axial orientation of OH-2 of the second glycopyranosyl residue appears to be the most important determinant for the extended interaction.  相似文献   

18.
Heme–copper oxidases (HCuOs) are the terminal components of the respiratory chain in the mitochondrial membrane or the cell membrane in many bacteria. These enzymes reduce oxygen to water and use the free energy from this reaction to maintain a proton-motive force across the membrane in which they are embedded. The heme–copper oxidases of the cbb3-type are only found in bacteria, often pathogenic ones since they have a low Km for O2, enabling the bacteria to colonize semi-anoxic environments. Cbb3-type (C) oxidases are highly divergent from the mitochondrial-like aa3-type (A) oxidases, and within the heme–copper oxidase family, cbb3 is the closest relative to the most divergent member, the bacterial nitric oxide reductase (NOR). Nitric oxide reductases reduce NO to N2O without coupling the reaction to the generation of any electrochemical proton gradient. The significant structural differences between A- and C-type heme–copper oxidases are manifested in the lack in cbb3 of most of the amino acids found to be important for proton pumping in the A-type, as well as in the different binding characteristics of ligands such as CO, O2 and NO. Investigations of the reasons for these differences at a molecular level have provided insights into the mechanism of O2 and NO reduction as well as the proton-pumping mechanism in all heme–copper oxidases. In this paper, we discuss results from these studies with the focus on the relationship between proton transfer and ligand binding and reduction. In addition, we present new data, which show that CO binding to one of the c-type hemes of CcoP is modulated by protein–lipid interactions in the membrane. These results show that the heme c-CO binding can be used as a probe of protein–membrane interactions in cbb3 oxidases, and possible physiological consequences for this behavior are discussed.  相似文献   

19.
Three anthocyanins (13) and eight flavonols (411) were isolated from the flowers of Amherstia nobilis endemic to Myanmar. Anthocyanins were identified as cyanidin 3-O-glucoside (1), 3-O-xyloside (2), and peonidin 3-O-glucoside (3). On the other hand, flavonols were identified as isorhamnetin 3-O-glucoside (4), 7-O-glucoside (5), 3,7-di-O-glucoside (6) and 3-O-rutinoside (7), quercetin 3-O-rutinoside (8) and 3-O-glucoside (9), and kaempferol 3-O-rutinoside (10) and 3-O-glucoside (11). Although an anthocyanin, pelargonidin 3-O-pentoside, has been reported from the flowers of A. nobilis, it was not found in this survey. The presence of flavonols in A. nobilis was reported in this survey for the first time. Flavonoid composition of Amherstia was chemotaxonomically compared with those of phylogenetically related genera Cynometra and Brownea.  相似文献   

20.
3-(4-Fluorophenyl)-N-((4-fluorophenyl)sulphonyl)acrylamide (FFSA) is a potential tubulin polymerisation inhibitor. In this article, a theoretical study of the binding between FFSA and tubulin in colchicine site was carried out by molecular docking, molecular dynamics (MD) simulation and binding free energy calculations. The docking calculations preliminarily indicate that there are three possible binding modes 1, 2 and 3; MD simulations and binding free energy calculations identify that binding mode 2 is the most favourable, with the lowest binding free energy of ? 29.54 kcal/mol. Moreover, our valuable results for the binding are as follows: the inhibitor FFSA is suitably located at the colchicine site of tubulin, where it not only interacts with residues Leu248β, Lys254β, Leu255β, Lys352β, Met259β and Val181a by hydrophilic interaction, but also interacts with Val181α and Thr179α by hydrogen bond interaction. These two factors are both essential for FFSA strongly binding to tubulin. These theoretical results help understanding the action mechanism and designing new compounds with higher affinity to tubulin.  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