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 morphology of q = 0.5 fast-tumbling bicelles prepared with three different acyl chain lengths has been investigated by NMR. It is shown that bicelles prepared with DLPC (12 C) and DHPC are on average larger than those containing DMPC or DPPC (14 and 16 C) and DHPC, which may be due to a higher degree of mixing between DLPC and DHPC. The fast internal mobility of the lipids was determined from natural abundance carbon-13 relaxation. A similar dynamical behaviour of the phospholipids in the three different bicelles was observed, although the DPPC lipid acyl chain displayed a somewhat lower degree of mobility, as evidenced by higher generalized order parameters throughout the acyl chain. Carbon-13 relaxation was also used to determine the effect of different model transmembrane peptides, with flanking Lys residues, on the lipid dynamics in the three different bicelles. All peptides had the effect of increasing the order parameters for the DLPC lipid, while no effect was observed on the longer lipid chains. This effect may be explained by a mismatch between the hydrophobic length of the peptides and the DLPC lipid acyl chain.  相似文献   

2.
Summary A rapid and sensitive 2D approach is presented for measuring amide proton exchange rates and the NOE interaction between amide protons and water. The approach is applicable to uniformly 13C/15N-enriched proteins and can measure magnetization exchange rates in the 0.02 to >20s–1 range. The experiments rely on selective excitation of the water resonance, coupled with purging of underlying H resonances, followed by NOESY-or ROESY-type transfer to amide protons, which are dispersed by the amide 15N frequencies in an HSQC-type experiment. Two separate but interleaved experiments, with and without selective inversion of the H2O resonance, yield quantitative results. The method is demonstrated for a sample of the calcium-binding protein calcineurin B. Results indicate rapid amide exchange for the five calcineurin B residues that are analogous to the five rapidly exchanging residues in the central helix of the homologous protein calmodulin.  相似文献   

3.
When manganese stabilizing protein (MSP) was treated with 0.5 mM N-succinimidyl propionate (NSP), the rebinding ability and oxygen-releasing capabilities of the modified MSP were not altered, in spite of changes of MSP surface Lys residues. Furthermore, far-ultraviolet circular dichroism and intrinsic fluorescence spectra analysis revealed that 0.5 mM NSP-modified MSP retained most of its native secondary and tertiary structure. Mapping of the sites of NSP modification by Staphylococcus V8 protease digestion of the modified protein, as well as analysis by matrix-assisted laser desorption ionization-time of flight mass spectrometry, indicated that seven Lys residues were modified. The results suggested that these residues are not absolutely essential to the structure and function of MSP. However, when the NSP concentration was increased to 4 mM, the modified MSP was unable to bind photosystem Ⅱ and completely lost its reactivating capability. Both far-ultraviolet circular dichroism and intrinsic fluorescence spectra analysis revealed a clear conformational change in MSP after 4 mM NSP treatment, suggesting that some Lys residues are involved in maintaining the structure and function of MSP. Analysis by matrix-assisted laser desorption ionization-time of flight mass spectrometry indicated that another six Lys residues, namely Lys20, Lys 101, Lys196, Lys207, Lysl30 (or Lys137) and Lys66 (or Lys76), were modified by 4 mM NSP. Therefore, these six Lys residues are crucial in maintaining the structure and function of soluble MSP.  相似文献   

4.
5.
Abstract

The interaction of 6-hydroxyflavone (6HF) with hen egg white lysozyme (HEWL) has been executed using multi-spectroscopic and computational methods. Steady state fluorescence studies indicated that static quenching mechanism is involved in the binding of 6HF with HEWL, which was further supported by excited state lifetime and UV–vis absorption studies. The binding constant (Kb) of the HEWL–6HF complex was observed to be 6.44?±?0.09?×?104 M?1 at 293?K, which decreases with the increase in temperature. The calculation of the thermodynamic quantities showed that the binding is exothermic in nature with a negative enthalpy change (ΔH = ?11.91?±?1.02?kJ mol?1) along with a positive entropy change (ΔS = +51.36?±?2.43 J K?1 mol?1), and the major forces responsible for the binding are hydrogen bonding and hydrophobic interactions. The possibility of energy transfer from tryptophan (Trp) residue to the 6HF ligand was observed from Fo¨rster’s theory. The inclusion of 6HF within the binding site of HEWL induces some micro-environmental changes around the Trp residues as indicated by synchronous and three-dimensional (3D) fluorescence studies. The changes in secondary structural components of HEWL are observed on binding with 6HF along with a reduction in % α-helical content. Computational studies correlate well with the experimental finding, and the ligand 6HF is found to bind near to Trp 62 and Trp 63 residues of HEWL. Altogether, the present study provides an insight into the interaction dynamics and energetics of the binding of 6HF to HEWL.

Communicated by Ramaswamy H. Sarma  相似文献   

6.
The formation of a ligand-protein complex oftentimes results in significant chemical shift changes. These changes may occur not only in the binding pocket but also in distal regions of the protein target. Therefore the reassignment of the backbone resonances in the complex is frequently a time consuming challenge. Here we present a suite of resolution-enhanced N(z)-exchange NMR experiments useful for rapidly assigning backbone (1)H and (15)N amide resonances of the ligand-bound form of a protein in slow exchange with its free state. Incorporation of semi-constant time frequency labeling periods into 3D N(z)-exchange experiments in combination with the collection of resolution-enhanced 2D N(z)-exchange difference spectra leads to a powerful set of tools for analyzing protein-ligand complexes. This allows for both the assignment of the bound state and the rapid assessment of the protein binding interface. The proposed methodology is demonstrated on the complex formed by the dimerization-docking domain of the c-AMP-dependent protein kinase and the tethering domain of the dual-binding A-kinase anchoring protein (AKAP).  相似文献   

7.
Ion-channel activity of a series of gramicidin A analogues carrying charged amino-acid sequences on the C-terminus of the peptide was studied on planar bilayer lipid membranes and liposomes. It was found that the analogue with the positively charged sequence GSGRRRRSQS forms classical cationic pores at low concentrations and large unselective pores at high concentrations. The peptide was predominantly in the right-handed beta(6.3)-helical conformation in liposomes as shown by circular dichroism spectroscopy. The single-channel conductance of the large pore was estimated to be 320pS in 100mM choline chloride as judged from the fluctuation analysis of the multi-channel current. The analogue with the negatively charged sequence GSGEEEESQS exhibited solely classical cationic channel activity. The ability of a peptide to form different type of channels can be used in the search for broad-spectrum antibiotics.  相似文献   

8.
A new malbrancheamide analogue, isomalbrancheamide B (3), along with three known compounds, malbrancheamide (1), isomalbrancheamide (2), and premalbrancheamide (4), were isolated in higher yields from the alkaloid fraction of the fungus Malbranchea aurantiaca. The interaction of the alkaloids 1–4 with calmodulin (CaM) was analyzed using different enzymatic, fluorescence, spectroscopic, nuclear magnetic resonance (NMR), and molecular modelling techniques. On the basis of the enzymatic and fluorescence experiments, malbrancheamides 1–3 are classical CaM inhibitors. Compound 4, however, did not quench the extrinsic fluorescence of the CaM biosensor indicating that it could be a functional inhibitor. Circular dichroism, NMR, and molecular modelling studies revealed that 1 binds to CaM in the same hydrophobic pocket than the chlorpromazine and trifluoperazine, two classical CaM inhibitors. Thus, malbrancheamide and related monochlorinated analogues are compounds with a high potential for the development of new therapeutic agents, involving CaM as their molecular target.  相似文献   

9.
To determine the contribution of charged amino acids to binding with the photosystem II complex (PSII), the amino or carboxyl groups of the extrinsic 18 kDa protein were modified with N- succinimidyl propionate (NSP) or glycine methyl ester (GME) in the presence of a water-soluble carbodiimide, respectively. Based on isoelectric point shift, 4-10 and 10-14 amino groups were modified in the presence of 2 and 4 mM NSP, respectively. Similarly, 3-4 carboxyl groups were modified by reaction with 100 mM GME. Neutralization of negatively charged carboxyl groups with GME did not alter the binding activity of the extrinsic 18 kDa protein. However, the NSP-modified 18 kDa protein, in which the positively charged amino groups had been modified to uncharged methyl esters, failed to bind with the PSII membrane in the presence of the extrinsic 23 kDa protein. This defect can not be attributed to structural or conformational alterations imposed by chemical modification, as the fluorescence and circular dichroism spectra among native, GME- and NSP-modified extrinsic 18 kDa proteins were similar. Thus, we have concluded that the positive charges of lysyl residues in the extrinsic 18 kDa protein are important for its interaction with PSII membranes in the presence of the extrinsic 23 kDa protein. Furthermore, it was found that the negative charges of carboxyl groups of this protein did not participate in binding with the extrinsic 23 kDa protein associated with PSII membranes.  相似文献   

10.
Aggregation, orientation and dynamics of transmembrane helices are of relevance for protein function and transmembrane signaling. To explore the interactions of transmembrane helices and the interdependence of peptide structure and lipid composition of the membranes, β‐peptides were explored as model transmembrane domains. Various hydrophobic β‐peptide sequences were synthesized by solid phase peptide synthesis. Conformational analyses of β‐peptide helices were performed in organic solvents (methanol and 2,2,2‐trifluoroethanol) and in large unilamellar liposomes (dimyristoylphosphatidylcholine, dipalmitoylphosphatidylcholine and dioleoylphosphatidylcholine) indicating 12‐ and 14‐helix conformations, depending on β3‐amino acid sequences. The intrinsic tryptophan fluorescence of β3‐homotryptophan units inserted in the center or near the end of the sequence was used to verify the membrane insertion of the β‐peptides. A characteristic blue shift with peripheral β3‐homotryptophan compared with β‐peptides with central tryptophan served as indication for a transmembrane orientation of the β‐peptides within the lipid bilayer. Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
The interaction of four benzothiazole tethered triazole analogues (MS43, MS70, MS71, and MS78) with human serum albumin (HSA) was investigated using various spectroscopic techniques (ultraviolet–visible (UV–vis) light absorption, fluorescence, circular dichroism (CD), molecular docking and density functional theory (DFT) studies). Fluorescence quenching constants (~1012) revealed a static mode of quenching and binding constants (Kb ~104) indicating the strong affinity of these analogues for HSA. Further alteration in the secondary structure of HSA in the presence of these analogues was also confirmed by far UV–CD spectroscopy. The intensity loss in CD studied at 222 nm indicated an increase in random coil/β‐sheet conformations in the protein. Binding energy values (MS71 (?9.3 kcal mol?1), MS78 (?8.02 kcal mol?1), MS70 (?7.16 kcal mol?1) and MS43 (?6.81 kcal mol?1)) obtained from molecular docking revealed binding of these analogues with HSA. Molecular docking and DFT studies validated the experimental results, as these four analogues bind with HSA at site II through hydrogen bonding and hydrophobic interactions.  相似文献   

12.
Peroxiredoxin 6 (Prdx6) is a bifunctional protein with glutathione peroxidase and phospholipase A(2) (PLA(2)) activities, and it alone among mammalian peroxiredoxins can hydrolyze phospholipids. After identifying a potential catalytic triad (S32, H26, D140) from the crystal structure, site-specific mutations were used to evaluate the role of these residues in protein structure and function. The S32A mutation increased Prdx6 alpha-helical content, whereas secondary structure was unchanged by mutation to H26A and D140A. Lipid binding by wild-type Prdx6 to negatively charged unilamellar liposomes showed an apparent rate constant of 11.2 x 10(6) M(-1) s(-1) and a dissociation constant of 0.36 microM. Both binding and PLA(2) activity were abolished in S32A and H26A; in D140A, activity was abolished but binding was unaffected. Overoxidation of the peroxidatic C47 had no effect on lipid binding or PLA(2) activity. Fluorescence resonance energy transfer from endogenous tryptophanyls to lipid probes showed binding of the phospholipid polar head in close proximity to S32. Thus, H26 is a site for interfacial binding to the liposomal surface, S32 has a key role in maintaining Prdx6 structure and for phospholipid substrate binding, and D140 is involved in catalysis. This putative catalytic triad plays an essential role for interactions of Prdx6 with phospholipid substrate to optimize the protein-substrate complex for hydrolysis.  相似文献   

13.
The H2A/H2B heterodimer is a component of the nucleosome core particle, the fundamental repeating unit of chromatin in all eukaryotic cells. The kinetic folding mechanism for the H2A/H2B dimer has been determined from unfolding and refolding kinetics as a function of urea using stopped-flow, circular dichroism and fluorescence methods. The kinetic data are consistent with a three-state mechanism: two unfolded monomers associate to form a dimeric intermediate in the dead-time of the SF instrument (approximately 5 ms); this intermediate is then converted to the native dimer by a slower, first-order reaction. Analysis of the burst-phase amplitudes as a function of denaturant indicates that the dimeric kinetic intermediate possesses approximately 50% of the secondary structure and approximately 60% of the surface area burial of the native dimer. The stability of the dimeric intermediate is approximately 30% of that of the native dimer at the monomer concentrations employed in the SF experiments. Folding-to-unfolding double-jump experiments were performed to monitor the formation of the native dimer as a function of folding delay times. The double-jump data demonstrate that the dimeric intermediate is on-pathway and obligatory. Formation of a transient dimeric burst-phase intermediate has been observed in the kinetic mechanism of other intertwined, segment-swapped, alpha-helical, DNA-binding dimers, such as the H3-H4 histone dimer, Escherichia coli factor for inversion stimulation and E.coli Trp repressor. The common feature of a dimeric intermediate in these folding mechanisms suggests that this intermediate may accelerate protein folding, when compared to the folding of archael histones, which do not populate a transient dimeric species and fold more slowly.  相似文献   

14.
Dystroglycan (DG) is an adhesion complex, expressed in a wide variety of tissues, formed by an extracellular and a transmembrane subunit, alpha-DG and beta-DG, respectively, interacting noncovalently. Recently, we have shown that the recombinant ectodomain of beta-DG, beta-DG(654-750), behaves as a natively unfolded protein, as it is able to bind the C-terminal domain of alpha-DG, while not displaying a defined structural organization. We monitored the effect of a commonly used denaturing agent, the anionic detergent sodium dodecylsulphate (SDS), on beta-DG(654-750) using a number of biophysical techniques. Very low concentrations of SDS (< or =2 mM) affect both tryptophan fluorescence and circular dichroism of beta-DG, and significantly perturb the interaction with the alpha-DG subunit as shown by solid-phase binding assays and fluorescence titrations in solution. This result confirms, as recently proposed for natively unfolded proteins, that beta-DG(654-750) exists in a native state, which is crucial to fulfill its biological function. Two-dimensional NMR analysis shows that SDS does not induce any evident conformational rearrangement within the ectodomain of beta-DG. Its first 70 amino acids, which show a lower degree of mobility, interact with the detergent, but this does not change the amount of secondary structure, whereas the highly flexible and mobile C-terminal region of beta-DG(654-750) remains largely unaffected, even at a very high SDS concentration (up to 50 mM). Our data indicate that SDS can be used as a useful tool for investigating natively unfolded proteins, and confirm that the beta-DG ectodomain is an interesting model system.  相似文献   

15.
Drug and protein interaction provides a structural guideline in the rational drug designing and in the synthesis of new and improved drugs with greater efficacy. We have examined here the interaction tendency and mechanism of nintedanib (NTB), an anticancer drug (tyrosine kinase inhibitor) with bovine serum albumin (BSA), by spectroscopic techniques. The decline in Stern–Volmer quenching constants and binding constant with the temperature rise suggests that BSA forms a complex with NTB. Binding constant obtained by modified Stern–Volmer equation at 3 temperatures was realized to be of the order of ~104?M?1. Negative ΔG (~?5.93?kcal?mol?1), ΔH (?3.74?kcal?mol?1), and ΔS (?1.50?kcal?mol?1) values exhibited a spontaneous and exothermic reaction between BSA and NTB. NTB molecule interacts with BSA by forming hydrogen bonds, as elucidated by fluorescence results. Moreover, a minor increment in the helical conformation of BSA upon its binding to NTB was observed by circular dichroism spectroscopy. The modification in protein’s symmetry and a decline in hydrodynamic radii were observed in the presence of NTB (from ~3.6 to ~3?nm) as obtained by the dynamic light scattering measurement results.  相似文献   

16.
Isoprenaline hydrochloride is a potential cardiovascular drug helps in the smooth functioning of the heart muscles. So, we have performed the binding study of ISO with BSA. This study was investigated by UV absorption, fluorescence, synchronous fluorescence, circular dichroism, etc. The analysis of intrinsic fluorescence data showed the low binding affinity of ISO. The binding constant Kb was 2.8 × 103 M-1 and binding stoichiometry (n) was approximately one and the Gibb’s free energy change at 310 K was determined to be -8.69 kcal mol?1. Negative Gibb’s free energy change shows the spontaneity of the BSA and ISO interaction. We have found ISO-induced alternation in the UV absorption, synchronous fluorescence and CD spectra in the absence and presence of the quencher indicates the complex formation. In synchronous fluorescence, red shift was obtained because of the complex formation of BSA and ISO. The distance (r) between the BSA (donor) and ISO (acceptor) was 2.89 nm, determined by FRET. DLS measurements interpreted complex formation due to the reduction in hydrodynamic radii of the protein in the presence of the drug. The binding site of ISO was found to be nearer to Trp 134 with the help of molecular docking and the ΔG° was found to be –10.2 kcal mol?1. The esterase activity result suggests that ISO acts as competitive inhibitor. Thus, this study would help to determine the binding capacity of the drug to the protein which may indicate the efficiency of diffusion of ISO into the blood for the treatment of heart diseases.  相似文献   

17.
Intermolecular nuclear Overhauser effects (NOEs) were measured between the protons of various small solvent or gas molecules and the non-specific lipid transfer protein (ns-LTP) from wheat. Intermolecular NOEs were observed with the hydrophobic pocket in the interior of wheat ns-LTP, which grew in intensity in the order cyclopropane (saturated solution) < methane (140 bar) < ethane (40 bar) < acetonitrile (5% in water) < cyclohexane (saturated solution) < benzene (saturated solution). No intermolecular NOEs were observed with dioxane (5% in water). The intermolecular NOEs were negative for all of the organic molecules tested. Intermolecular NOEs between wheat ns-LTP and water were weak or could not be distinguished from exchange-relayed NOEs. As illustrated by the NOEs with cyclohexane versus dioxane, the hydrophobic pocket in wheat ns-LTP preferably binds non-polar molecules. Yet, polar molecules like acetonitrile can also be accommodated. The pressure dependence of the NOEs between methane and wheat ns-LTP indicated incomplete occupancy, even at 190 bar methane pressure. In general, NOE intensities increased with the size of the ligand molecule and its vapor pressure. NMR of the vapor phase showed excellent resolution between the signals from the gas phase and those from the liquid phase. The vapor concentration of cyclohexane was fivefold higher than that of the dioxane solution, supporting the binding of cyclohexane versus uptake of dioxane.  相似文献   

18.
The combination of several drugs is necessary, especially during long-term therapy. A competitive binding of the drugs can cause a decrease in the amount of drugs actually bound to the protein and increase the biologically active fraction of the drug. Here, the interaction between 4,4′-Diisothiocyano-2,2′-stilbenedisulfonic acid (DIDS) and 2,4-Dinitrophenol (DNP) with Hemoglobin (Hb) was investigated by different spectroscopic and molecular modeling techniques. Fluorescence analysis was used to estimate the effect of the DIDS and DNP on Hb as well as to define the binding properties of binary and ternary complexes. The distance r between donor and acceptor was obtained by the FRET and found to be 2.25 and 2.13 nm for DIDS and DNP in binary and 2.08 and 2.07 nm for (Hb–DNP) DIDS and (Hb–DIDS) DNP complexes in ternary systems, respectively. Time-resolved fluorescence spectroscopy confirmed static quenching for Hb in the presence of DIDS and DNP in both systems. Furthermore, an increase in ellipticity values of Hb upon interaction with DIDS and DNP showed secondary structural changes of protein that determine to disrupt of hydrogen bonds and electrostatic interactions. Our results showed that the Hb destabilize in the presence of DIDS and DNP. Molecular modeling of the possible binding sites of DIDS and DNP in binary and ternary systems in Hb confirmed the experimental results.  相似文献   

19.
Carbendazim is a benzimidazole fungicide used to control the fungal invasion. However, its exposure might lead to potential health problems. The present study evaluates the interaction of carbendazim (CAR) with human serum albumin (HSA) which is an important drug carrier protein and plays a very crucial role in the transportation of small molecules. A number of biophysical techniques were employed to investigate the binding of CAR with HSA. The increased UV-absorption of HSA on titrating with CAR suggests the formation of HSA–CAR complex and it could be due to the exposure of aromatic residues. The fluorescence study confirmed that CAR quenches the fluorescence of HSA and showed the static mode of quenching. CAR (50 µM) quenches around 56.14% of the HSA fluorescence. The quenching constant, binding constant, number of binding site and free energy change was calculated by fluorescence quenching experiment. Competitive displacement assay showed Sudlow’s site I as the primary binding site of CAR on HSA. The synchronous fluorescence study revealed the perturbation in the microenvironment around tyrosine and tryptophan residues upon binding of CAR to HSA. The circular dichroism results suggested that the binding of CAR to HSA altered its secondary structure. Molecular docking experiment demonstrated the binding of CAR to Sudlow’s site I of HSA. Docking studies suggested that the hydrogen bonding, van der Waals and pi-alkyl are playing role in the interaction of CAR with HSA. The study confirmed the conformational changes within HSA upon binding of CAR.  相似文献   

20.
A broader exploitation of enzymes in organic synthesis can be achieved by increasing their tolerance toward organic solvents. In this study, the stability and activity of Baeyer–Villiger monooxygenases from Thermobifida fusca (PAMO) and Acinetobacter sp. (CHMO) in the presence of water miscible organic solvents were compared. PAMO was more stable than CHMO. The concentration of solvent (v/v) at which it halved its activity (C50) was 4‐ to 16‐fold higher than that observed for CHMO. For PAMO, the C50 varied from 16% to 55% of solvent and followed the destabilizing order methanol < ethanol < 1,4‐dioxane < acetonitrile < trifluoroethanol. In the case of CHMO, the maximal C50 was 7% with methanol and even lower with the other solvents. Therefore, methanol was the most tolerated solvent. In the case of PAMO, methanol induced a significant increase of enzyme activity (up to fivefold), which was optimal at 20% (v/v) solvent. Only minor spectral variations were observed with PAMO in 20% methanol, suggesting that the increase of activity observed in this condition is not due to marked conformational changes. Fluorescence and circular dichroism analyses showed that the lower stability of CHMO toward organic solvent correlates with a more pronounced destructive effect on its secondary and tertiary structure. A possible rationale for the higher stability of PAMO could be inferred from inspection of the PAMO and CHMO (two enzymes of similar size) structure, which revealed a higher (up to twofold) number of ionic bridges in PAMO with respect to CHMO. Biotechnol. Bioeng. 2011; 108:491–499. © 2010 Wiley Periodicals, Inc.  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