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 46 毫秒
1.
The biochemical mechanism of action of human neutrophil peptide-1 (HNP-1) against Mycobacterium tuberculosis H37Ra was studied. Mycobacteria grown in the presence of a subinhibitory concentration (IC50) of HNP-1 showed a significant decrease in the biosynthesis of vital macromolecules, as shown by the incorporation of various radiolabeled precursors. Mycobacterial cells grown in the presence of HNP-1 exhibited surface changes, as was evident from the increased number of binding sites for l-anilinonaphthalene 8-sulfonate. Permeability studies carried out with spheroplasts showed a significantly high permeability to a fluorescent probe, N-phenyl naphthylamine, in the presence of HNP-1. Significant changes in the cell wall and cell membrane were observed when HNP-1-grown cells were analysed by transmission electron microscopy. Our results suggest the mycobacterial cell wall/membrane to be the major target(s) of HNP-1. Received: 7 January 1999 / Accepted: 15 February 1999  相似文献   

2.
To establish a novel strategy of P. aeruginosa control, we acquired recombination HNP-3 by gene recombination. Then we examined HNP-3 bio-activities and the influences of antimicrobial peptide on the efflux of ATP. Consequently, we obtained target protein with a molecular mass of 3,000 D consistent with the Anticipation. FIC index of Ciprofloxacin added HNP-3 was less than 0.5 and HNP3 synergistically cooperated with CIP to suppress P. aeruginosa colony formation revealed that there was significant synergy. ATP efflux was however up-regulated by low concentrations of HNP-3, although CIP did not exert any influence on ATP efflux. Conclusively, recombination protein HNP-3 displayed antimicrobial and synergic effects. HNP-3 enhanced the antimicrobial activity of CIP by promoting ATP efflux from P. aeruginosa cells and decreasing efflux of the drugs, which could have useful clinical applications.  相似文献   

3.
The galactose‐specific lectin LecA from Pseudomonas aeruginosa is a target for the development of new anti‐infectious compounds. Sugar based molecules with anti‐adhesive properties present great potential in the fight against bacterial infection and biofilm formation. LecA is specific for oligosaccharides with terminal α‐galactoside residues and displays strong affinity for melibiose (αGal1‐6Glc) with a Kd of 38.8 µM. The crystal structure of LecA/melibiose complex shows classical calcium‐bridged binding of αGal in the primary binding site but also revealed a secondary sugar binding site with glucose bound. This sugar binding site is in close proximity to the galactose binding one, is independent of calcium and mainly involves interactions with a symmetry‐related protein. This discovery would help to the design of new potent inhibitors targeting both binding sites. Proteins 2014; 82:1060–1065. © 2013 Wiley Periodicals, Inc.  相似文献   

4.
BceRS and PsdRS are paralogous two‐component systems in Bacillus subtilis controlling the response to antimicrobial peptides. In the presence of extracellular bacitracin and nisin, respectively, the two response regulators (RRs) bind their target promoters, PbceA or PpsdA, resulting in a strong up‐regulation of target gene expression and ultimately antibiotic resistance. Despite high sequence similarity between the RRs BceR and PsdR and their known binding sites, no cross‐regulation has been observed between them. We therefore investigated the specificity determinants of PbceA and PpsdA that ensure the insulation of these two paralogous pathways at the RR–promoter interface. In vivo and in vitro analyses demonstrate that the regulatory regions within these two promoters contain three important elements: in addition to the known (main) binding site, we identified a linker region and a secondary binding site that are crucial for functionality. Initial binding to the high‐affinity, low‐specificity main binding site is a prerequisite for the subsequent highly specific binding of a second RR dimer to the low‐affinity secondary binding site. In addition to this hierarchical cooperative binding, discrimination requires a competition of the two RRs for their respective binding site mediated by only slight differences in binding affinities.  相似文献   

5.
The DNA binding of amphiphilic iron(III) 2,17‐bis(sulfonato)‐5,10,15‐tris(pentafluorophenyl)corrole complex (Fe–SC) was studied using spectroscopic methods and viscosity measurements. Its nuclease‐like activity was examined by using pBR322 DNA as a target. The interaction of Fe–SC with human serum albumin (HSA) in vitro was also examined using multispectroscopic techniques. Experimental results revealed that Fe–SC binds to ct‐DNA via an outside binding mode with a binding constant of 1.25 × 104 M–1. This iron corrole also displays good activity during oxidative DNA cleavage by hydrogen peroxide or tert‐butyl hydroperoxide oxidants, and high‐valent (oxo)iron(V,VI) corrole intermediates may play an important role in DNA cleavage. Fe–SC exhibits much stronger binding affinity to site II than site I of HSA, indicating a selective binding tendency to HSA site II. The HSA conformational change induced by Fe–SC was confirmed by UV/Vis and CD spectroscopy. Copyright © 2015 John Wiley & Sons, Ltd.  相似文献   

6.
The transmissive plasmid R64 (IncI1) performs an antirestriction function, reducing the efficiency of EcoKI-dependent restriction in Escherichia coli K12 cells approximately fivefold. The R64 ardA gene has been cloned and sequenced. The ArdA proteins specifically inhibit type I restriction–modification enzymes. R64 ArdA is highly homologous to ColIb-P9 ArdA: only 4 out of 166 amino acid residues differ. While ColIb-P9 inhibits both endonuclease and methylase activities of the type I restriction–modification enzyme EcoKI (R2M2S), R64 ArdA inhibits only its endonuclease activity. It has been assumed that R64 ArdA suppresses the binding of unmodified DNA with the R subunit, which is responsible for DNA translocation and cleavage. ColIb-P9 ArdA suppresses DNA binding not only with the R, but also with the S subunit, which contacts the sK site containing target adenines. The binding of ArdA with the specific site inhibits both endonuclease and methylase activities; the binding of ArdA with the nonspecific site of the R subunit inhibits only the endonuclease activity ofEcoKI (R2M2S).  相似文献   

7.
Fur (f erric u ptake r egulator) is the master regulator of iron homeostasis in many bacteria, but how it responds specifically to Fe(II) in vivo is not clear. Biochemical analyses of Bacillus subtilis Fur (BsFur) reveal that in addition to Fe(II), both Zn(II) and Mn(II) allosterically activate BsFur–DNA binding. Dimeric BsFur co‐purifies with site 1 structural Zn(II) (Fur2Zn2) and can bind four additional Zn(II) or Mn(II) ions per dimer. Metal ion binding at previously described site 3 occurs with highest affinity, but the Fur2Zn2:Me2 form has only a modest increase in DNA binding affinity (approximately sevenfold). Metallation of site 2 (Fur2Zn2:Me4) leads to a ~ 150‐fold further enhancement in DNA binding affinity. Fe(II) binding studies indicate that BsFur buffers the intracellular Fe(II) concentration at ~ 1 μM. Both Mn(II) and Zn(II) are normally buffered at levels insufficient for metallation of BsFur site 2, thereby accounting for the lack of cross‐talk observed in vivo. However, in a perR mutant, where the BsFur concentration is elevated, BsFur may now use Mn(II) as a co‐repressor and inappropriately repress iron uptake. Since PerR repression of fur is enhanced by Mn(II), and antagonized by Fe(II), PerR may co‐regulate Fe(II) homeostasis by modulating BsFur levels in response to the Mn(II)/Fe(II) ratio.  相似文献   

8.
The interaction between DNA and ionen polymers, -[N+(CH3)2(CH2)mN+(CH3)2(CH2)n], with m-n of 3–3, 6–6, and 6–10 were examined in order to know how the binding behavior of cationic polymers with DNA depends on the charge density of polycation. The ionen polymer has no bulky side chain and the binding forces with DNA would be attributed mainly to electrostatic interaction. When 3–3 ionen polymers were added to DNA solution, precipitable complexes with the ratio of cationic residue to DNA phosphate (+/?) of 1/1 and the free DNA molecules were segregated, while 6–6 and 6–10 ionen polymers formed soluble complexes with DNA molecules up to (+/?) = 0.5. This suggests that 3–3 ionen polymers bind cooperatively with DNA while 6–6 and 6–10 ionen polymers bind noncooperatively. The cooperative binding of 3–3 ionen polymer and the noncooperative binding of 6–6 ionen polymer were also supported by the thermal melting and recooling profiles from the midpoint between first and second meltings. It was concluded that the charge density of DNA phosphate is a critical value determining whether the ionen polymers bind to DNA by a cooperative or by a noncooperative binding, since the distance between successive cationic charges of 3–3 ionen polymer is shorter than that between successive phosphate charges on DNA double helix and those of 6–6 and 6–10 ionen polymers are longer.  相似文献   

9.
Two aminoglycosides, micronomicin (MN), and tobramycin (TB), binding with DNA were studied using various spectroscopic techniques including fluorescence, UV–Vis, FT-IR, and CD spectroscopy coupled with relative viscosity and molecular docking. Studies of fluorescence quenching and time-resolved fluorescence spectroscopy all revealed that MN/TB quenching the fluorescence of DNA–EB belonged to static quenching. The binding constants and binding sites were obtained. The values of ΔH, ΔS, and ΔG suggested that van der Waals force or hydrogen bond might be the main binding force. FT-IR and CD spectroscopy revealed that the binding of MN/TB with DNA had an effect on the secondary structure of DNA. Binding mode of MN/TB with DNA was groove binding which was ascertained by viscosity measurements, CD spectroscopy, ionic strength, melting temperature (Tm), contrast experiments with single stranded (ssDNA), and double stranded DNA (dsDNA). Molecular docking analysis further confirmed that the groove binding was more acceptable result.  相似文献   

10.
Gene expression can be altered by small molecules that target DNA; sequence as well as shape selectivities are both extremely important for DNA recognition by intercalating and groove‐binding ligands. We have characterized a carbohydrate scaffold (1) exhibiting DNA “shape readout” properties. Thermodynamic studies with 1 and model duplex DNAs demonstrate the molecule's high affinity and selectivity towards B* form (continuous AT‐rich) DNA. Isothermal titration calorimetry (ITC), circular dichroism (CD) titration, ultraviolet (UV) thermal denaturation, and Differential Scanning Calorimetry were used to characterize the binding of 1 with a B* form AT‐rich DNA duplex d[5′‐G2A6T6C2‐3′]. The binding constant was determined using ITC at various temperatures, salt concentrations, and pH. ITC titrations were fit using a two‐binding site model. The first binding event was shown to have a 1:1 binding stoichiometry and was predominantly entropy‐driven with a binding constant of approximately 108 M?1. ITC‐derived binding enthalpies were used to obtain the binding‐induced change in heat capacity (ΔCp) of ?225 ± 19 cal/mol·K. The ionic strength dependence of the binding constant indicated a significant electrolytic contribution in ligand:DNA binding, with approximately four to five ion pairs involved in binding. Ligand 1 displayed a significantly higher affinity towards AT‐tract DNA over sequences containing GC inserts, and binding experiments revealed the order of binding affinity for 1 with DNA duplexes: contiguous B* form AT‐rich DNA (d[5′‐G2A6T6C2‐3′]) >B form alternate AT‐rich DNA (d[5′‐G2(AT)6C2‐3′]) > A form GC‐rich DNA (d[5′‐A2G6C6T2‐3′]), demonstrating the preference of ligand 1 for B* form DNA. © 2013 Wiley Periodicals, Inc. Biopolymers 101: 720–732, 2014.  相似文献   

11.
12.
13.
Summary We have previously identified a unique site, pac, from which packaging of precursor concatameric viral DNA into proheads starts during the maturation process of bacteriophage CP-T1. The direction of this packaging was determined from restriction enzyme cleavage patterns of CP-T1 DNA. A restriction enzyme generated fragment containing pac was cloned and the surrounding DNA region sequenced. Analysis of the nucleotide sequence revealed numerous repeat regions related to the consensus sequence PuagttGAT.AAT.aa.t. Within the sequenced region an open reading frame encoding a 12260 Mr protein was also identified. This protein appears to share homology with the binding domains of known DNA binding proteins and may represent a putative Pac terminase possessing the specific endonuclease activity required for cleavage at the pac site. Minicell analysis of deletion derivatives of the pac-containing clone revealed a protein of approximately 12900 Mr encoded within this same region, confirming that this Pac protein is phage encoded.  相似文献   

14.
The binding of tryptophan repressor (TrpR) to its operators was examined quantitatively usingin vitroandin vivomethods. DNA sequence requirements for 1:1 and tandem 2 :1 (TrpR : DNA) binding in various sequence contexts were studied. The results indicate that the optimal half-site sequence for recognition by one helix-turn-helix motif of one TrpR dimer is3′ CNTGA 5′5′ GNACT 3′, consistent with contacts observed by X-ray diffraction analysis of cocrystalline 1:1 and 2 :1 complexes. Half-sites can be paired to form a palindrome either by direct abutment, forming the nucleation site for a tandem 2 :1 complex, or with an 8-base-pair spacer, forming a 1:1 target. Dimethylsulfate (DMS) methylation-protection footprintingin vitroof 1:1 and 2 :1 complexes formed sequentially on the two unequal half-site pairs of thetrpEDCBA operator fromSerratia marcescensindicated an obligate hierarchy of site occupancy, with one half-site pair serving as the nucleation site for tandem binding. DMS footprinting ofEscherichia colioperatorsin vivoshowed that, over a wide range of intracellular TrpR concentration, thetrpEDCBA operator is occupied by three repressor dimers,aroH is occupied by two dimers, and the 1:1 binding mode is used on thetrpR operator. The coexistence of these distinct occupancy states implies that changes in protein concentration affect only the fractional occupancy of each operator rather than the binding mode, which is determined by the number of half-site sequences present in the operator region. Cooperativity of tandem complex formation measured by gel retardation using a symmetrized synthetic operator containing identical, optimal sites spaced as in natural operators was found to be modest, implying a maximum coupling free energy of ∼−2 kcal/mol. On other sequences the apparent degree of cooperativity, as well as the apparent affinity, varied with sequence and sequence context in a manner consistent with the structural models and which suggests compensation between affinity and cooperativity as a mechanism that allows tolerance of operator sequence variation.  相似文献   

15.
Abstract: The selective serotonin (5-HT) agonist 8-hydroxydipropylaminotetralin (8-OH-DPAT) has been extensively used to characterize the physiological, biochemical, and behavioral features of the 5-HT1A receptor. A further characterization of this receptor subtype was conducted with membrane preparations from rat cerebral cortex and hippocampus. The saturation binding isotherms of [3H]8- OH-DPAT (free ligand from 200 pM to 160 nM) revealed high-affinity 5-HT1A receptors (KH= 0.7–0.8 nM) and lowaffinity (KL= 22–36 nM) binding sites. The kinetics of [3H]8-OH-DPAT binding were examined at two ligand concentrations, i.e., 1 and 10 nM, and in each case revealed two dissociation rate constants supporting the existence of high- and low-affinity binding sites. When the high-affinity sites were labeled with a 1 nM concentration of [3H]8- OH-DPAT, the competition curves of agonist and antagonist drugs were best fit to a two-site model, indicating the presence of two different 5-HT1A binding sites or, alternatively, two affinity states, tentatively designated as 5-HT1AHIGH and 5-HT1ALOW. However, the low correlation between the affinities of various drugs for these sites indicates the existence of different and independent binding sites. To determine whether 5-HT1A sites are modulated by 5′-guanylylimidodiphosphate, inhibition experiments with 5-HT were performed in the presence or in the absence of 100 μM 5′-guanylylimidodiphosphate. The binding of 1 nM [3H]8-OH-DPAT to the 5-HT1AHIGH site was dramatically (80%) reduced by 5′-guanylylimidodiphosphate; in contrast, the low-affinity site, or 5-HT1ALOW, was seemingly insensitive to the guanine nucleotide. The findings suggest that the high-affinity 5-HT1AHIGH site corresponds to the classic 5-HT1A receptor, whereas the novel 5-HT1ALOW binding site, labeled by 1 nM [3H]8-OH-DPAT and having a micromolar affinity for 5-HT, may not belong to the G protein family of receptors. To further investigate the relationship of 5-HT1A sites and the 5-HT innervation, rats were treated with p-chlorophenylalanine or with the neurotoxin p-chloroamphetamine. The inhibition of 5-HT synthesis by p-chlorophenylalanine did not alter either of the two 5-HT1A sites, but deafferentation by p-chloroamphetamine caused a loss of the low-affinity [3H]8-OH- DPAT binding sites, indicating-that these novel binding sites may be located presynaptically on 5-HT fibers and/or nerve terminals.  相似文献   

16.
The fucosterol has been reported numerous biological activities. In this study, the activity in vitro of the fucosterol from Sargassum horridum as potential human acetylcholinesterase inhibitor was evaluated. The structural identification was obtained by nuclear magnetic resonance (NMR) spectroscopy and based on experimental data, we combined docking and molecular dynamics simulations coupled to the molecular-mechanics-generalized-born-surface-area approach to evaluating the structural and energetic basis for the molecular recognition of fucosterol and neostigmine at the binding site of acetylcholinesterase (AChE). In addition, the Lineweaver–Burk plot showed the nature of a non-competitive inhibition. The maximum velocity (Vmax) and the constant of Michaelis–Menten (Km) estimated for fucosterol (0.006 µM) were 0.015 1/VoA/h and 6.399 1/[ACh] mM?1, respectively. While, for neostigmine (0.14 µM), the Vmax was 0.022 1/VoA/h) and Km of 6.726 1/[ACh] mM?1, these results showed a more effective inhibition by fucosterol respect to neostigmine. Structural analysis revealed that neostigmine reaches the AChE binding site reported elsewhere, whereas fucosterol can act as a no-competitive and competitive acetylcholinesterase inhibitor, in agree with kinetic enzymatic experiments. Binding free energy calculations revealed that fucosterol reaches the acetylcholinesterase binding site with higher affinity than neostigmine, which is according to experimental results. Whereas the per-residue decomposition free energy analysis let us identify crucial residues involved in the molecular recognition of ligands by AChE. Results corroborate the ability of theoretical methods to provide crucial information at the atomic level about energetic and structural differences in the binding interaction and affinity from fucosterol with AChE.

Communicated by Ramaswamy H. Sarma  相似文献   


17.
18.
Chlorophyll-free plasma membranes of the unicellular green alga Chlamydomonas reinhardtii Dangeard were purified from a microsomal fraction using an aqueous polymer two-phase system of 6.5% (w/w) dextran T500, 6·5% (w/w) polyethylene glycol 3350, 60 mM NaCI, 0 33 M sucrose and 5 mM potassium phosphate (pH 7·8). The plasma membrane fraction contained only 2·4% of the microsomal membrane protein. Specific activity of the plasma membrane marker enzyme, K*, Mg2+-ATPase (EC 3.6.1.3). was enriched 9-fold over the microsomal fraction, and 22% of total activity was recovered in the upper, polyethylene glycol-rich phase. Contamination from intracellular membranes was minimal. K*, Mg2+-ATPase showed a pH optimum at about 6·5, and addition of 0·05% (w/v) Triton X-100 stimulated the activity 3-fold. [3H]-Nimodipinc was employed to characterize 1,4-dihydropyridine-specific membrane receptors. Two apparent binding sites with different affinities to nimodipine were found in the crude microsomal fraction. The separation of plasma membranes from intracellular membranes revealed that one binding site with higher affinity (KD= 9 nM) was located on the plasma membrane and a second binding site with lower affinity (KD= 36 nM) on an intracellular membrane The apparent dissociation constants determined from the association and dissociation rate constants in kinetic experiments were comparable to those determined by equilibrium experiments. The maximum number of binding sites of the plasma membrane fraction and the intracellular membrane fraction was Bmax= 440 and 470 fmol (mg protein)-1, respectively. [3H]-Nimodipinc binding was inhibited by (±) verapamil and stimulated by D-cis-diltiazem in both fractions. Moreover, ethyle-neglycol-bis(2-aminoethylcther)-N, N'-tetraacctic acid (EGTA) inhibited [3H]-nimo-dipinc binding in the plasma membrane fraction but not in the intracellular membrane fraction This effect was cancelled by the addition of CaCl2.  相似文献   

19.
A comparative study of secondary specificities of enteropeptidase and trypsin was performed using peptide substrates with general formula A-(Asp/Glu) n -Lys(Arg)--B, where n = 1-4. This was the first study to demonstrate that, similar to other serine proteases, enteropeptidase has an extended secondary binding site interacting with 6-7 amino acid residues surrounding the peptide bond to be hydrolyzed. However, in the case of typical enteropeptidase substrates containing four negatively charged Asp/Glu residues at positions P2-P5, electrostatic interaction between these residues and the secondary site Lys99 of the enteropeptidase light chain is the main factor that determines hydrolysis efficiency. The secondary specificity of enteropeptidase differs from the secondary specificity of trypsin. The chromophoric synthetic enteropeptidase substrate G5DK-F(NO2)G (k cat/K m = 2380 mM–1·min–1) is more efficient than the fusion protein PrAD4K-P26 (k cat/K m = 1260 mM–1·min–1).  相似文献   

20.
A series of complexes of Au(III) with nucleosides and nucleotides and their methyl derivatives in different stoichiometry have been prepared. Ultraviolet, visible, ir, and nmr studies have been performed to determine the site of binding of these ligands with the metal ion. In (1:4) Au(III): guanosine complex, N7 is the binding site, whereas at 1:1 complex, a bidentate type of chelation through C6O and N7 is observed. C6-NH2 is favored over N1 as coordinating site at all stoichiometry in the adenosine complex. Inosine binds through N1 at r = 1. In cytidine, N1 is the binding site, whereas thymidine reacts only at high pH. In the case of nucleotides a bidentate type of chelation through the phosphate and the ring nitrogen occurs. The phosphate binding ability of Au(III) was further confirmed by studying the interaction of Au(III) with dimethyl phosphate—a conformational analog of the phosphate backbone in DNA chain.  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