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.
Steady-state and dynamic fluorescence measurements have been performed on DAPI in solution and in complexes formed with a number of synthetic and natural polydeoxynucleotides. The decay of DAPI in buffer at pH 7 was decomposed using two exponentials having lifetime values of approximately 2.8 ns and 0.2 ns. The double exponential character of the decay was maintained over a large pH range from 3 to 9. At pH 1 the short component dominated, whereas at pH 12, only the long component was detectable. Two distinct spectra were associated with the two lifetime components; the short component was shifted to the red. The short lifetime component occurs in the presence of water. In water the excitation spectra depended on the emission wavelength and there was no viscosity dependence of the two forms. To explain these results we propose that there is a ground state conformer in which preferential solvation of the indole ring allows proton transfer in the excited state. DAPI complexed with polydeoxynucleotides retained most of the features of the decay of DAPI in solution. However, the complexes with fuly AT-containing polymers stabilized the longer lifetime form of DAPI because the stronger binding enhanced solvent shielding. A gradual increase of the short lifetime component, which monitors dye solvent exposure, was obtained as the AT content was decreased. For polyd(GC) the decay was similar to that of free DAPI.Abbreviations DAPI 4-6-diamidino-2-phenylindole - POPOP 1,4-bis(5-phenyl-2-oxazolyl)-benzene; 2,2-p-phenylene-bis(5-phenyloxazole) Financial support for this work was provided by a M.P.I. grant 1984, Roma, Italy for M.L.B. and NSF-PCM 84-03107 and PHS-IP41RR03155 for E.G.  相似文献   

2.
The protamine thynnine is an arginine-rich protein approximately 30 amino acids long with a tyrosine in the middle of its sequence. Its fluorescence decay kinetics can be described by a biexponential function with lifetimes of 0.52 and 2.1 ns, with almost equal preexponential factors. The fluorescence quencher CsCl does not affect the short lifetime but shifts the equilibrium between the long and short lifetime toward the short one and reduces the long lifetime. In nature, thynnine is found complexed with chromosomal DNA. In vitro complexes of thynnine with double-stranded (ds) DNA are stable at physiologic ionic strength but dissociate at high NaCl concentration. This dissociation can be monitored by steady-state fluorescence. From the salt concentration dependence of the dissociation of the complex of thynnine with ds-DNA 145 bp long, it can be concluded that only 4 of 21 possible full electrostatic bonds are involved in thynnine-DNA binding. In addition, the binding constant at 1M NaCl is of the order of 106, indicating a strong nonelectrostatic component in arginine-DNA binding.  相似文献   

3.
Both the iron-containing and the manganese-containing superoxide dismutases from Escherichia coli show diminished activity with increasing ionic strength, indicative of electrostatic facilitation of the catalyzed reaction. Since both enzymes bear a net negative charge at the assay pH, as does the substrate, this suggests a cationic locale in the active site region. Acetylation of the enzymes inverted their response to increasing ionic strength. It thus appears that lysine residues provide the observed electrostatic facilitation. A specific inhibition by large monovalent anions was observed with the iron-containing superoxide dismutase and was taken to indicate the presence of a cationic group, within a hydrophobic crevice, at the active site.  相似文献   

4.
Steady-state and dynamic fluorescence titrations show that: (a) the complex between beta-lactoglobulin (BLG) and 1-anilinonaphthalene-8-sulfonate (ANS) displays a heterogeneous equilibrium with large changes in the binding strength vs. pH and ion concentration; and (b) the fluorescence response of bound ANS reveals two separate lifetimes that suggest two different sites (or binding modes). While steady-state fluorescence titrations yield effective values of the binding constant and of the bound ANS quantum efficiency, it is shown that, by combining steady-state fluorescence and lifetime decay of ANS, it is possible to give quantitative estimates of the association constants for each site. When heading from the acid (pH approximately 2) to the native state (pH approximately 6) the main result is a very large reduction of the effective binding constant. This and the results of titrations vs. ionic strength suggest that electrostatic interactions are a major contribution to ANS binding to BLG.  相似文献   

5.
The static accessibility discrete charge algorithm for protein charge interactions is extended to the case of linear polyelectrolytes. In this model, the effective dielectric value between surface charge sites depends predominantly on the solvent ionic strength and the solvent accessibilities of the charge sites. This treatment accounts for the phenomena of specific ion binding in the context of a general electrostatic effect [Matthew and Richards (1982) Biochemistry 21 , 4989]. Specific ion sites are determined by locating areas of high electrostatic potential at the solvent interface of the macromolecule. At a given ionic strength the calculated potential at a site is taken to describe a binding constant and therefore the ion site occupancy. For a 20-base-pair fragment of B-DNA, net charge of ?40, 16 ion sites are indicated in the minor groove. The partial occupancy of each site increases from 0.2 to 0.5 as the ionic strength is increased from 0.01 to 0.50. Over the same range of ionic strength, the electrostatic free energy of this charge array is calculated to change from +0.6 to ?0.05 kcal/bp. Parallel behavior is predicted for A- and Z-DNA charge geometries. The most stable configuration, based on electrostatic criteria, at high ionic strength (I = 0.1–0.5) is that of Z-DNA. In this range, the ratio of “bound” sodium to phosphate is predicted to be less than 0.4.  相似文献   

6.
Fluorescence decay studies, obtained by multifrequency phase-modulation fluorometry, have been performed on DAPI in solution and complexed with natural and synthetic polydeoxynucleotides. DAPI decay at pH 7 was decomposed using two exponential components of 2.8 and 0.2 ns of lifetime values, respectively. The double exponential character of the decay was maintained over a large pH range. Phase- and modulation-resolved spectra, collected between 420 and 550 nm, have indicated at least two spectral components associated with the two lifetime values. This, plus the observation of the dependence of the emission spectrum on the excitation wavelength, suggests a lifetime heterogeneity originating from ground-state molecular conformers, partially affected by pH changes. DAPI complexed with natural polydeoxynucleotides retained most of the features of DAPI decay in solution, except for the value of the long lifetime component that was longer (approximately 4 ns) and the relative fractional fluorescence intensities of the two components that were inverted. AT polymers/DAPI complexes show single exponential decay. Solvent shielding when DAPI is bound to DNA changes the indole ring solvation and stabilizes the longer lifetime decay component. For poly(GC)/DAPI complex, the decay was similar to that of free DAPI in solution, proving the dependence on the polydeoxynucleotides sequence the different types of binding and the reliability of the fluorescence method to solve them.  相似文献   

7.
The fluorescence properties of the single tryptophanyl residue present in amicyanin from Thiobacillus versutus are very similar to those of azurin from Pseudomonas aeruginosa and other mononuclear blue copper proteins. The emission maximum is well structured and centered at 318 nm. The quantum yield is strongly affected by the presence of copper, the removal of which is accompanied by a more than sixfold increase in fluorescence, without change in shape. The fluorescence decay of holo-amicyanin is heterogeneous with a longer component of 5.7 ns and a shorter one of 0.7 ns accounting for 90% of the total emitting molecules. Copper-free amicyanin shows instead a single exponential decay (3.3 ns) of intrinsic fluorescence. This lifetime decreases as the temperature increases as does the longer lifetime component of holoamicyanin.  相似文献   

8.
The electrostatic potentials within the pore of the nicotinic acetylcholine receptor (nAChR) were determined using lanthanide-based diffusion-enhanced fluorescence energy transfer experiments. Freely diffusing Tb3+ -chelates of varying charge constituted a set of energy transfer donors to the acceptor, crystal violet, a noncompetitive antagonist of the nAChR. Energy transfer from a neutral Tb3+ -chelate to nAChR-bound crystal violet was reduced 95% relative to the energy transfer to free crystal violet. This result indicated that crystal violet was strongly shielded from solvent when bound to the nAChR. Comparison of energy transfer from positively and negatively charged chelates indicate negative electrostatic potentials of -25 mV in the channel, measured in low ionic strength, and -10 mV measured in physiological ionic strength. Debye-Hückel analyses of potentials determined at various ionic strengths were consistent with 1-2 negative charges within 8 A of the crystal violet binding site. To complement the energy transfer experiments, the influence of pH and ionic strength on the binding of [3H]phencyclidine were determined. The ionic strength dependence of binding affinity was consistent with -3.3 charges within 8 A of the binding site, according to Debye-Hückel analysis. The pH dependence of binding had an apparent pKa of 7.2, a value indicative of a potential near -170 mV if the titratable residues are constituted of aspartates and glutamates. It is concluded that long-range potentials are small and likely contribute little to selectivity or conductance whereas close interactions are more likely to contribute to electrostatic stabilization of ions and binding of noncompetitive antagonists within the channel.  相似文献   

9.
The mobilities of several fluorescent probes placed at different locations on calmodulin in the absence of Ca2+ have been found to depend upon the charge, ionic strength, and temperature. In general, the time decay of fluorescence anisotropy could be fitted with two rotational correlation times. The shorter of these reflects primarily the motion of the probe itself, while the longer corresponds to the motion of a major portion of the molecule. An increase in ionic strength or a decrease in net charge results in a decrease in the relative amplitude of the shorter correlation time, while an increase in temperature produces an increase in its amplitude. These results are consistent with, and suggest, that an increase in probe mobility accompanies an expansion of the calmodulin molecule under conditions of high electrostatic stress.  相似文献   

10.
8-Anilino-1-naphthalenesulfonic acid (ANS) is widely used as a probe for locating binding sites of proteins. To characterize the binding sites of tear lipocalin (TL), we studied ANS binding to apoTL by steady-state and time-resolved fluorescence. Deconvolution of ANS binding revealed that two lifetime components, 16.99 ns and 2.76 ns at pH 7.3, have dissociation constants of 0.58 μM and 5.7 μM, respectively. At pH 3.0, the lifetime components show decreased affinities with dissociation constants of 2.42 μM and ∼21 μM, respectively. Selective displacement of ANS molecules from the ANS-apoTL complex by stearic acid discriminates the internal and external binding sites. Dependence of the binding affinity on ionic strength under various conditions provides strong evidence that an electrostatic interaction is involved. Time-resolved fluorescence is a promising tool to segregate multiple binding sites of proteins.  相似文献   

11.
The contribution of electrostatic forces to the interaction of Na,K-ATPase with adenine nucleotides was investigated by studying the effect of ionic strength on nucleotide binding. At pH 7.0 and 20 degrees C, there was a qualitative correlation between the equilibrium dissociation constant (K(d)) values for ATP, ADP, and MgADP and their total charges. All K(d) values increased with increasing ionic strength. According to the Debye-Hückel theory, this suggests that the nucleotide binding site and its ligands have "effective" charges of opposite signs. However, quantitative analysis of the dependence on ionic strength shows that the product of the effective electrostatic charges on the ligand and the binding site is the same for all nucleotides, and is therefore independent of the total charge of the nucleotide. The data suggest that association of nucleotides with Na,K-ATPase is governed by a partial charge rather than the total charge of the nucleotide. This charge, interacting with positive charges on the protein, is probably the one corresponding to the alpha-phosphate of the nucleotide. Dissociation rate constants measured in complementary transient kinetic experiments were 13 s(-1) for ATP and 27 s(-1) for ADP, independent of the ionic strength in the range 0.1-0.5 M. This implies similar association rate constants for the two nucleotides (about 40 x 10(6) M(-1) s(-1) at I = 0.1 M). The results suggest that long-range Coulombic forces, affecting association rates, are not the main contributors to the observed differences in affinities, and that local interactions, affecting dissociation rates, may play an even greater role.  相似文献   

12.
J B Matthew  G I Hanania  F R Gurd 《Biochemistry》1979,18(10):1928-1936
The electrostatic treatment applied in the preceding paper in this issue [Matthew, J. B., Hanania, G.I.H., & Gurd, F.R.N. (1979) Biochemistry (preceding paper in this issue)] to the titration behavior of individual groups in human deoxyhemoglobin and oxyhemoglobin was applied to the computation of the alkaline Bohr effect at various values of ionic strength. The enhanced proton binding of deoxyhemoglobin in the pH range of 6--9 was accounted for at ionic strength 0.01 M by the effects of the unique charge distributions of ionizable groups in the two quaternary states. At ionic strength 0.10 M the effects of 2--4 bound anions had to be considered in addition in the deoxyhemoglobin charge configuration. At the higher ionic strength 10 groups per tetramer contributed to the Bohr effect, whereas 28 groups were contributory at the lower ionic strength. The ionic strength dependence of individual groups in the two tetrameric structures as well as in the alpha-chain monomer was explained in terms of the electrostatic treatment. This examination showed that the differences in electrostatic behavior of deoxy- and oxyhemoglobin follow from particular dissymmetries in their configurations with respect to charge and static solvent accessibility.  相似文献   

13.
14.
The binding of oligopeptides containing basic and aromatic residues to phospholipid vesicles has been studied by fluorescence spectroscopy. Tryptophan-containing peptide such as Lys-Trp-Lys or Lys-Trp(OMe) exhibit a shift of their fluorescence toward shorter wavelengths and an increased fluorescence quantum yield upon binding to phosphatidylinositol (PI) or phosphatidylserine (PS) vesicles. No binding was detected with phosphatidylcholine vesicles. The binding is strongly dependent on ionic strength and pH. Binding decreases when ionic strength increases indicating an important role of electrostatic interactions. The pH-dependence of binding reveals that the apparent pK of the terminal carboxyl group of Lys-Trp-Lys is raised by ~3 units upon binding to PI and PS vesicles. The binding of tyrosine-containing peptides to PI and PS vesicles is characterized by an increase in the fluorescence quantum yield of the peptide without any shift in fluorescence maximum. A natural nonapeptide from the myelin basic protein which contains one tryptophan residue binds to PI and PS vesicles at low pH when the acidic groups are neutralized. This binding is accompanied by a shift of the tryptophyl fluorescence toward shorter wavelengths together with an enhancement of the fluorescence quantum yield. Dissociation of the complex is achieved at high ionic strength. These results indicate that aromatic residues of oligopeptides bound to the phospholipid polar heads by electrostatic interactions become buried in a more hydrophobic environment in the vicinity of the aliphatic chains of the lipids.  相似文献   

15.
The contributions of electrostatic interactions to the binding stability of barnase and barstar were studied by the Poisson-Boltzmann model with three different protocols: a), the dielectric boundary specified as the van der Waals (vdW) surface of the protein along with a protein dielectric constant (epsilon (p)) of 4; b), the dielectric boundary specified as the molecular (i.e., solvent-exclusion (SE)) surface along with epsilon (p) = 4; and c), "SE + epsilon (p) = 20." The "vdW + epsilon (p) = 4" and "SE + epsilon (p) = 20" protocols predicted an overall electrostatic stabilization whereas the "SE + epsilon (p) = 4" protocol predicted an overall electrostatic destabilization. The "vdW + epsilon (p) = 4" protocol was most consistent with experiment. It quantitatively reproduced the observed effects of 17 mutations neutralizing charged residues lining the binding interface and the measured coupling energies of six charge pairs across the interface and reasonably rationalized the experimental ionic strength and pH dependences of the binding constant. In contrast, the "SE + epsilon (p) = 4" protocol predicted significantly larger coupling energies of charge pairs whereas the "SE + epsilon (p) = 20" protocol did not predict any pH dependence. This study calls for further scrutiny of the different Poisson-Boltzmann protocols and demonstrates potential danger in drawing conclusions on electrostatic contributions based on a particular calculation protocol.  相似文献   

16.
Single tryptophan-containing mutants of low adenylylation state Escherichia coli glutamine synthetase have been studied by frequency-domain fluorescence spectroscopy in the presence of various substrates and inhibitors. At pH 6.5, the Mn-bound wild-type enzyme (wild type has two tryptophans/subunit) and the mutant enzymes exhibit heterogeneous fluorescence decay kinetics; the individual tryptophans are adequately described by a triple exponential decay scheme. The recovered lifetime values are 5.9 ns, 2.6 ns, and 0.4 ns for Trp-57 and 5.8 ns, 2.3 ns, and 0.4 ns for Trp-158. These values are nearly identical to the previously reported results at pH 7.5 (Atkins, W.M., Stayton, P.S., & Villafranca, J.J., 1991, Biochemistry 30, 3406-3416). In addition, Trp-57 and Trp-158 both exhibit an ATP-induced increase in the relative fraction of the long lifetime component, whereas only Trp-57 is affected by this ligand at pH 7.5. The transition-state analogue L-methionine-(R,S)-sulfoximine (MSOX) causes a dramatic increase in the fractional intensity of the long lifetime component of Trp-158. This ligand has no effect on the W158S mutant protein and causes a small increase in the fractional intensity of the long lifetime component of the W158F mutant protein. Addition of glutamate to the ATP complex, which affords the gamma-glutamylphosphate-ADP complex, results in the presence of new lifetime components at 7, 3.2, and 0.5 ns for Trp-158, but has no effect on Trp-57. Similar results were obtained when ATP was added to the MSOX complex; Trp-57 exhibits heterogeneous fluorescence decay with lifetimes of 7, 3.5, and 0.8 ns. Decay kinetics of Trp-158 are best fit to a nearly homogeneous decay with a lifetime of 5.5 ns in the MSOX-ATP inactivated complex. These results provide a model for the sequence of structural and dynamic changes that take place at the Trp-57 loop and the central loop (Trp-158) during several intermediate stages of catalysis.  相似文献   

17.
The two tryptophan residues of ferredoxin from Halobacterium of the Dead Sea differ in their fluorescence characteristics. One of these tryptophan residues (class 1) absorbs more to the red and is thus probably in a more apolar environment than the other (class 2). Upon removal of the ferric ions, i.e., in the apoferredoxin, a 2.2-fold increase in the quantum yield of fluorescence is observed. A double exponential decay of the fluorescence is found for ferredoxin, reduced ferredoxin, as well as for the apoferredoxin. The longer decay time assumes a constant value of 6.9 ns in all three cases, indicating that it originates in a tryptophan residue which is not affected by changes in the Fe3+ binding site (class 2 tryptophan). The shorter decay component increases gradually from 0.55 ns in oxidized ferredoxin, through 0.80 ns in the reduced ferredoxin to 1.24 ns in the apoprotein. This decay component is thus assumed to be largely due to the second tryptophan residue of the protein (class 1) located close to the Fe3+ binding site. On the other hand, the relative decay amplitude of the class 2 tryptophan is doubled upon formation of apoferredoxin. It is concluded that the class 1 tryptophan is quenched by the active site ferric ions and that the class 2 tryptophan is partially exposed to a polar environment. Whereas class 1 tryptophan may be similar to the single nonfluorescent tryptophan of spinach ferredoxin, class 2 tryptophan is found in a peptide which is present only in halophilic ferredoxins. Conformational changes occur in the molecule upon removal—but not reduction—of the ferric ions, causing the environment of the class 2 tryptophan to become more hydrophobic. It is possible that the class 1 tryptophan is associated with the occurrence of a higher redox potential in this ferredoxin, when compared with chloroplast-type ferredoxins.  相似文献   

18.
Electrophoresis measurements on Micrococcus lysodeikticus have shown that the net surface charge density on the cell wall is constant at around -1.5 microC/cm2 for the pH range 4-8. This result has enabled a quantitative analysis to be made of how the electrostatic field associated with the negatively charged cell wall influences the ionic strength and pH dependency of the lytic activity of lysozyme towards M. lysodeikticus. A dominant effect is the creation of a local pH gradient at the cell wall, and at high ionic strengths the lytic activity is found to be controlled by an electrostatic force of attraction between the lysozyme molecule and the cell wall. As the ionic strength of the supporting electrolyte is decreased, however, an electrostatic force of repulsion becomes dominant and is associated with a negative charge carried by the lysozyme molecule, which could possibly be the ionized Asp-52 residue at the active site. This is considered to arise from the fact that at low ionic strengths the fine details of the heterogeneous charge distribution on the cell wall and lysozyme molecule are only partially screened by counter ions.  相似文献   

19.
Solubility of lysozyme chloride was determined in the absence of added salt and in the presence of 0.05-1.2 M NaCl, starting from isoionic lysozyme, which was then brought to pH values from 9 to 3 by addition of HCl. The main observation is the absence of a salting-in region whatever the pH studied. This is explained by a predominant electrostatic screening of the positively charged protein and/or by adsorption of chloride ions by the protein. The solubility increases with the protein net charge at low ionic strength, but the reverse is observed at high ionic strength. The solubility of lysozyme chloride seems to become independent of ionic strength at pH approximately 9.5, which is interpreted as a shift of the isoionic pH (10.8) to an isoelectric pH due to chloride binding. The crystallization at very low ionic strength, where lysozyme crystallizes at supersaturation values as low as 1.1, amplifies the effect of pH on protein solubility. Understanding the effect of the net charge and of ionic strength on protein-protein interactions is valuable not only for protein crystal growth but more generally also for the formation of protein-protein or protein-ligand complexes.  相似文献   

20.
Human interferon binds to a omega-carboxpentyl-agarose column at low ionic strength (0.15 M NaCl) and is still retained when the ionic strength is raised (to 1.0 M NaCl). The binding can be reversed, however, by ethylene glycol, indicating a hydrophobic interaction. The binding of human interferon to omega-aminohexyl-agarose is weak, even at a low ionic strength, and is probably exclusively electrostatic. This disparate binding behavior may be caused by the presence of a positive charge, adjacent to the hydrophobic binding site, on human interferon. The interaction of human interferon with omega-carboxypentyl-agarose is quite selective, inasmuch as the majority of proteins present in interferon preparations pass through the column unretained. Hydrophobic chromatography of human interferon may thus be useful in its purification.  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