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.
In this study the electrostatic and nonelectrostatic contributions to the binding free energy of a number of different protein-DNA recognition complexes are investigated. To determine the electrostatic effects in the protein-DNA association the Poisson-Boltzmann approach was applied. Overall the salt-dependent electrostatic free energy opposed binding in all protein-DNA complexes except one, and the salt-independent electrostatic contribution favored binding in more than half of the complexes. Further the salt-dependent electrostatic free energy increased with higher ionic concentrations and therefore complex association is stronger opposed at higher ionic concentrations. The hydrophobic effect in the protein-DNA complexes was determined from the buried accessible surface area and the surface tension. A majority of the complexes showed more polar than nonpolar buried accessible surface area. Interestingly the buried DNA-accessible surface area was preferentially hydrophilic, only in one complex a slightly more hydrophobic buried accessible surface area was observed. A quite sophisticated balance between several different free energy components seems to be responsible for determining the free energy of binding in protein-DNA systems.  相似文献   

2.
Hutchins RA  Crenshaw JM  Graves DE  Denny WA 《Biochemistry》2003,42(46):13754-13761
The DNA binding energetics of a series of analogues derived from the anticancer agent N-[2-(dimethylamino)ethyl]-9-aminoacridine-4-carboxamide (AAC) are investigated. The effects of substituent modification at the C5 position of the acridine chromophore on the interaction of AAC with DNA are determined using spectrophotometry and isothermal titration calorimetry (ITC). The binding affinity and binding free energy associated with the interaction of AAC with DNA are significantly enhanced upon substitution at the C5 position. Energetic profiles describing ligand-DNA complex formation obtained from ITC indicate that C5 substitution significantly enhances binding enthalpy relative to the parent AAC. In many cases, the enhanced binding enthalpies of the C5-substituted analogues correlate with anticancer activity. Because of the cationic character of AAC and its analogues, the DNA binding properties of these compounds are dependent on ionic strength. To quantitate the ionic contributions to complex formation, the observed binding free energy of each compound is parsed into its polyelectrolyte and nonelectrostatic components. Enhanced nonelectrostatic contributions to the overall binding free energies observed with C5-substituted analogues relative to the parent AAC suggest that C5 substituents play a critical role in directing both thermodynamic mechanisms associated with complex formation and molecular interactions between the ligand and its DNA binding site. These studies have demonstrated that substitution of AAC at the C5 position results in enhanced DNA binding affinity and energetics.  相似文献   

3.
Nidhi Singh  Arieh Warshel 《Proteins》2010,78(7):1705-1723
Calculating the absolute binding free energies is a challenging task. Reliable estimates of binding free energies should provide a guide for rational drug design. It should also provide us with deeper understanding of the correlation between protein structure and its function. Further applications may include identifying novel molecular scaffolds and optimizing lead compounds in computer‐aided drug design. Available options to evaluate the absolute binding free energies range from the rigorous but expensive free energy perturbation to the microscopic linear response approximation (LRA/β version) and related approaches including the linear interaction energy (LIE) to the more approximated and considerably faster scaled protein dipoles Langevin dipoles (PDLD/S‐LRA version) as well as the less rigorous molecular mechanics Poisson–Boltzmann/surface area (MM/PBSA) and generalized born/surface area (MM/GBSA) to the less accurate scoring functions. There is a need for an assessment of the performance of different approaches in terms of computer time and reliability. We present a comparative study of the LRA/β, the LIE, the PDLD/S‐LRA/β, and the more widely used MM/PBSA and assess their abilities to estimate the absolute binding energies. The LRA and LIE methods perform reasonably well but require specialized parameterization for the nonelectrostatic term. The PDLD/S‐LRA/β performs effectively without the need of reparameterization. Our assessment of the MM/PBSA is less optimistic. This approach appears to provide erroneous estimates of the absolute binding energies because of its incorrect entropies and the problematic treatment of electrostatic energies. Overall, the PDLD/S‐LRA/β appears to offer an appealing option for the final stages of massive screening approaches. Proteins 2010. © 2010 Wiley‐Liss, Inc.  相似文献   

4.
The DNA binding of iron(II) mixed-ligand complexes containing 1,10-phenanthroline(phen) and 4,7-diphenyl-1,10-phenanthroline(dip), [Fe(phen)(3)](2+), [Fe(phen)(2)(dip)](2+) and [Fe(phen)(dip)(2)](2+) has been characterized by spectrophotometric titration and melting temperature measurements. The salt concentration dependence of the binding constant has allowed us to dissect the DNA-binding constant and free energy change of each iron(II) complex into the nonelectrostatic and polyelectrolyte contributions. A comparison of the nonelectrostatic components in the binding free energy changes among iron(II) complexes has made it possible to rigorously evaluate the contribution of the ligand substituents to the DNA-binding event. The peripheral substitution of phen by two phenyl groups increases the nonelectrostatic binding constant of the iron(II) complex more than 20 times, which is equivalent to approximately 7.5 kJ mol(-1) of more favorable contribution to the DNA binding. In general, the iron(II) complexes studied have higher affinity towards the more facile A-T sequence than the G-C sequence. This preferential binding may be attributed to the steric effect induced by the ancillary part of the ligands in the course of DNA binding. The binding of disubstituted iron(II) complex to DNA is quite strong as reflected in the modest increase in the denaturation temperature (T(m)) of double helical DNA upon the interaction with the iron(II) complex.  相似文献   

5.
The structure of a 58 nucleotide ribosomal RNA fragment buries several phosphate groups of a hairpin loop within a large tertiary core. During refinement of an X-ray crystal structure containing this RNA, a potassium ion was found to be contacted by six oxygen atoms from the buried phosphate groups; the ion is contained completely within the solvent-accessible surface of the RNA. The electrostatic potential at the ion chelation site is unusually large, and more than compensates for the substantial energetic penalties associated with partial dehydration of the ion and displacement of delocalized ions. The very large predicted binding free energy, approximately -30 kcal/mol, implies that the site must be occupied for the RNA to fold. These findings agree with previous studies of the ion-dependent folding of tertiary structure in this RNA, which concluded that a monovalent ion was bound in a partially dehydrated environment where Mg2+ could not easily compete for binding. By compensating the unfavorable free energy of buried phosphate groups with a chelated ion, the RNA is able to create a larger and more complex tertiary fold than would be possible otherwise.  相似文献   

6.
Genheden S  Ryde U 《Proteins》2012,80(5):1326-1342
We have compared the predictions of ligand‐binding affinities from several methods based on end‐point molecular dynamics simulations and continuum solvation, that is, methods related to MM/PBSA (molecular mechanics combined with Poisson–Boltzmann and surface area solvation). Two continuum‐solvation models were considered, viz., the Poisson–Boltzmann (PB) and generalised Born (GB) approaches. The nonelectrostatic energies were also obtained in two different ways, viz., either from the sum of the bonded, van der Waals, nonpolar solvation energies, and entropy terms (as in MM/PBSA), or from the scaled protein–ligand van der Waals interaction energy (as in the linear interaction energy approach, LIE). Three different approaches to calculate electrostatic energies were tested, viz., the sum of electrostatic interaction energies and polar solvation energies, obtained either from a single simulation of the complex or from three independent simulations of the complex, the free protein, and the free ligand, or the linear‐response approximation (LRA). Moreover, we investigated the effect of scaling the electrostatic interactions by an effective internal dielectric constant of the protein (?int). All these methods were tested on the binding of seven biotin analogues to avidin and nine 3‐amidinobenzyl‐1H‐indole‐2‐carboxamide inhibitors to factor Xa. For avidin, the best results were obtained with a combination of the LIE nonelectrostatic energies with the MM+GB electrostatic energies from a single simulation, using ?int = 4. For fXa, standard MM/GBSA, based on one simulation and using ?int = 4–10 gave the best result. The optimum internal dielectric constant seems to be slightly higher with PB than with GB solvation. © Proteins 2012; © 2012 Wiley Periodicals, Inc.  相似文献   

7.
The interaction of the human adenovirus proteinase (AVP) with various DNAs was characterized. AVP requires two cofactors for maximal activity, the 11-amino acid residue peptide from the C-terminus of adenovirus precursor protein pVI (pVIc) and the viral DNA. DNA binding was monitored by changes in enzyme activity or by fluorescence anisotropy. The equilibrium dissociation constants for the binding of AVP and AVP-pVIc complexes to 12-mer double-stranded (ds) DNA were 63 and 2.9 nM, respectively. DNA binding was not sequence specific; the stoichiometry of binding was proportional to the length of the DNA. Three molecules of the AVP-pVIc complex bound to 18-mer dsDNA and six molecules to 36-mer dsDNA. When AVP-pVIc complexes bound to 12-mer dsDNA, two sodium ions were displaced from the DNA. A Delta of -4.6 kcal for the nonelectrostatic free energy of binding indicated that a substantial component of the binding free energy results from nonspecific interactions between the AVP-pVIc complex and DNA. The cofactors altered the interaction of the enzyme with the fluorogenic substrate (Leu-Arg-Gly-Gly-NH)2-rhodamine. In the absence of any cofactor, the Km was 94.8 microM and the kcat was 0.002 s(-1). In the presence of adenovirus DNA, the Km decreased 10-fold and the kcat increased 11-fold. In the presence of pVIc, the Km decreased 10-fold and the kcat increased 118-fold. With both cofactors present, the kcat/Km ratio increased 34000-fold, compared to that with AVP alone. Binding to DNA was coincident with stimulation of proteinase activity by DNA. Although other proteinases have been shown to bind to DNA, stimulation of proteinase activity by DNA is unprecedented. A model is presented suggesting that AVP moves along the viral DNA looking for precursor protein cleavage sites much like RNA polymerase moves along DNA looking for a promoter.  相似文献   

8.
9.
We showed recently that the high-salt transition of poly[d(G-C)]. poly[d(G-C)] between B-DNA and Z-DNA (at [NaCl] = 2.25 M or [MgCl(2)] = 0.7 M) can be ascribed to the lesser electrostatic free energy of the B form, due to better immersion of the phosphates in the solution. This property was incorporated in cylindrical DNA models that were analyzed by Poisson-Boltzmann theory. The results are insensitive to details of the models, and in fair agreement with experiment. In contrast, the Z form of the poly[d(G-m5C)] duplex is stabilized by very small concentrations of magnesium. We now show that this striking difference is accommodated quantitatively by the same electrostatic theory, without any adjustable parameter. The different responses to magnesium of the methylated and nonmethylated polymers do not come from stereospecific cation-DNA interactions: they stem from an experimentally derived, modest difference in the nonelectrostatic component of the free energy difference (or NFED) between the Z and B forms. The NFED is derived from circular DNA measurements. The differences between alkaline earth and transition metal ions are explained by weak coordination of the latter. The theory also explains the induction of the transition by micromolar concentrations of cobalt hexammine, again without specific binding or adjustable parameters. Hence, in the case of the B-Z transition as in others (e.g., the folding of tRNA and of ribozymes), the effect of multivalent cations on nucleic acid structure is mediated primarily by nonspecific ion-polyelectrolyte interactions. We propose this as a general rule for which convincing counter-examples are lacking.  相似文献   

10.
We studied the effect of methanol, ethanol, iso-propanol, and n-propanol on the reaction of hemoglobin with oxygen. The oxygen affinity was found to decrease with increasing alcohol concentration and alkyl group size; no detectable effect on Hill's constant was found. Difference spectroscopy indicated KR not to be affected by the presence of alcohols; the lowered affinity was then attributed to an altered equilibrium between T and R conformations of hemoglobin. The results have been analyzed in such a way as to allow separation of electrostatic contributions to free energy difference between the T and R states from nonelectrostatic ones. The nonelectrostatic term has been attributed to protein–solvent hydrophobic interactions. Values of hydrophobic free energy are in good agreement with analogous data estimated by correlating different results previously reported in the literature.  相似文献   

11.
Coordinate ion pair formation between EcoRI endonuclease and DNA   总被引:10,自引:0,他引:10  
The free energy of the binding reaction between EcoRI restriction endonuclease and a specific cognate dodecadeoxynucleotide (d(CGCGAATTCGCG)) has contributions from both electrostatic and nonelectrostatic components. These contributions were dissected by measuring the effects of varying salt concentration on the equilibrium binding constant and applying the thermodynamic analyses of Record et al. (Record, M. T., Jr., Lohman, T. M., and deHaseth, P. L. (1976) J. Mol. Biol. 107, 145-158). Endonuclease mutation S187 (Arg 187 to Ser) (Greene, P. J., Gupta, M., Boyer, H. W., Brown, W. E., and Rosenberg, J. M. (1981) J. Biol. Chem. 256, 2143-2153) did not significantly affect the nonelectrostatic component but did perturb the electrostatic contribution to the binding energy (we are numbering the amino acid residues according to the DNA sequence). The former was determined by extrapolating the linear portion of the salt dependence curve (0.125 to 0.25 M KCl) to 1 M ionic strength, with the same result for both wild type and S187 endonucleases at both pH 6.0 and 7.4 (-8.5 +/- 1.5 kcal/mol or greater than 50% of the total binding free energy). The slopes of these same curves yield estimates of eight ionic interactions between wild type endonuclease and the DNA at both pH values. By contrast, binding of EcoRI-S187 to dodecanucleotide involves six charge-charge interactions at pH 6.0. Only two ionic interactions are observed at pH 7.4. This was unexpected since gel permeation chromatography demonstrated that the recognition complex for both wild type and S187 proteins contains an enzyme dimer and a DNA duplex. EcoRI-S187 endonuclease retains wild type DNA sequence specificity, and the rate of the phosphodiester hydrolysis step is also unchanged. Thus, electrostatic interactions are functionally separable from sequence recognition and strand cleavage. Our results also establish that arginine 187 plays a key role in the electrostatic function and suggest that it might be located at the DNA-protein interface. The disproportionate loss of ion pairs at pH 7.4 can be rationalized by a model which suggests that six conformationally mobile ionic groups on the protein act in a coordinated manner during the interaction with DNA.  相似文献   

12.
Booth J  Brown T  Vadhia SJ  Lack O  Cummins WJ  Trent JO  Lane AN 《Biochemistry》2005,44(12):4710-4719
DNA duplexes are stabilized by aminopropynyl modification of pyrimidines at the 5 position. A combination of thermodynamic analyses as a function of ionic strength, NMR, and molecular modeling has been applied to determine the origin of the stabilization. UV melting studies of a dodecamer bearing one, two, or three nonadjacent modified dU and dC and of a single dU(8) in the Dickerson-Drew dodecamer revealed that the modifications are essentially additive in terms of T(m), DeltaG, and DeltaH, and there is little difference between dU and dC. The free energy change was parsed into electrostatic and nonelectrostatic components, which showed a significant contribution from charge interactions at physiological ionic strength but also a nonelectrostatic contribution that arises in part from hydration. NMR spectroscopy of the modified Dickerson-Drew dodecamer revealed that the conformation of the duplexes is not significantly altered by the modifications, though (31)P NMR shows that the positive charge may affect ionic interactions with the oxygen atoms of the neighboring phosphates. The modified duplex showed significant hydration in both major and minor grooves. The single strands were also analyzed by NMR, which showed evidence of significant stacking interactions in the modified oligonucleotide. Parsing the energy contribution has shown that electrostatics and hydration can produce substantial increases in thermodynamic stability without significant changes in the conformation of the duplex state. These considerations have significance for the design of oligonucleotides used for hybridization.  相似文献   

13.
Homeodomains are helix-turn-helix type DNA-binding domains that exhibit sequence-specific DNA binding by insertion of their "recognition" alpha helices into the major groove and a short N-terminal arm into the adjacent minor groove without inducing substantial distortion of the DNA. The stability and DNA binding of four representatives of this family, MATalpha2, engrailed, Antennapedia, and NK-2, and truncated forms of the last two lacking their N-terminal arms have been studied by a combination of optical and microcalorimetric methods at different temperatures and salt concentrations. It was found that the stability of the free homeodomains in solution is rather low and, surprisingly, is reduced by the presence of the N-terminal arm for the Antennapedia and NK-2 domains. Their stabilities depend significantly upon the presence of salt: strongly for NaCl but less so for NaF, demonstrating specific interactions with chloride ions. The enthalpies of association of the homeodomains with their cognate DNAs are negative, at 20 degrees C varying only between -12 and -26 kJ/mol for the intact homeodomains, and the entropies of association are positive; i.e., DNA binding is both enthalpy- and entropy-driven. Analysis of the salt dependence of the association constants showed that the electrostatic component of the Gibbs energy of association resulting from the entropy of mixing of released ions dominates the binding, being about twice the magnitude of the nonelectrostatic component that results from dehydration of the protein/DNA interface, van der Waals interactions, and hydrogen bonding. A comparison of the effects of NaCl/KCl with NaF showed that homeodomain binding results in a release not only of cations from the DNA phosphates but also of chloride ions specifically associated with the proteins. The binding of the basic N-terminal arms in the minor groove is entirely enthalpic with a negative heat capacity effect, i.e., is due to sequence-specific formation of hydrogen bonds and hydrophobic interactions rather than electrostatic contacts with the DNA phosphates.  相似文献   

14.
Specific RNA binding by Q beta coat protein   总被引:11,自引:0,他引:11  
The interaction between the bacteriophage Q beta coat protein and its specific binding site on Q beta genomic RNA was characterized by using a nitrocellulose filter binding assay. Q beta coat protein bound to a synthetic 29-nucleotide RNA hairpin with an association constant of 400 microM-1 at 4 degrees C, 0.2 M ionic strength, pH 6.0. Complex formation had a broad pH optimum centered around pH 6.0 and was favored by both enthalpy and entropy. The salt dependence of Ka revealed that four to five ion pairs may be formed in the complex although approximately 80% of the free energy of complex formation is contributed by nonelectrostatic interactions. Truncation experiments revealed that coat protein binding required only the presence of a hairpin with an eight base pair stem and a three-base loop. Analysis of the binding properties of hairpin variants showed that the sequence of the stem was not important for coat protein recognition and only one of the three loop residues was essential. A bulged adenosine present in the coat protein binding site was not required for coat protein binding. Q beta coat protein binding specific is therefore primarily achieved by the structure and not by the sequence of the operator.  相似文献   

15.
We have developed a non‐redundant protein–RNA binding benchmark dataset derived from the available protein–RNA structures in the Protein Database Bank. It consists of 73 complexes with measured binding affinity. The experimental conditions (pH and temperature) for binding affinity measurements are also listed in our dataset. This binding affinity dataset can be used to compare and develop protein–RNA scoring functions. The predicted binding free energy of the 73 complexes from three available scoring functions for protein–RNA docking has a low correlation with the binding Gibbs free energy calculated from Kd. © 2013 The Protein Society  相似文献   

16.
The predictions of the derivative of the electrostatic binding free energy of a biomolecular complex, ΔGel, with respect to the logarithm of the 1:1 salt concentration, d(ΔGel)/d(ln[NaCl]), SK, by the Poisson-Boltzmann equation, PBE, are very similar to those of the simpler Debye-Hückel equation, DHE, because the terms in the PBE's predictions of SK that depend on the details of the dielectric interface are small compared to the contributions from long-range electrostatic interactions. These facts allow one to obtain predictions of SK using a simplified charge model along with the DHE that are highly correlated with both the PBE and experimental binding data. The DHE-based model developed here, which was derived from the generalized Born model, explains the lack of correlation between SK and ΔGel in the presence of a dielectric discontinuity, which conflicts with the popular use of this supposed correlation to parse experimental binding free energies into electrostatic and nonelectrostatic components. Moreover, the DHE model also provides a clear justification for the correlations between SK and various empirical quantities, like the number of ion pairs, the ligand charge on the interface, the Coulomb binding free energy, and the product of the charges on the complex's components, but these correlations are weak, questioning their usefulness.  相似文献   

17.
The secondary structure of encapsidated MS2 genomic RNA poses an interesting RNA folding challenge. Cryoelectron microscopy has demonstrated that encapsidated MS2 RNA is well-ordered. Models of MS2 assembly suggest that the RNA hairpin-protein interactions and the appropriate placement of hairpins in the MS2 RNA secondary structure can guide the formation of the correct icosahedral particle. The RNA hairpin motif that is recognized by the MS2 capsid protein dimers, however, is energetically unfavorable, and thus free energy predictions are biased against this motif. Computer programs called Crumple, Sliding Windows, and Assembly provide useful tools for prediction of viral RNA secondary structures when the traditional assumptions of RNA structure prediction by free energy minimization may not apply. These methods allow incorporation of global features of the RNA fold and motifs that are difficult to include directly in minimum free energy predictions. For example, with MS2 RNA the experimental data from SELEX experiments, crystallography, and theoretical calculations of the path for the series of hairpins can be incorporated in the RNA structure prediction, and thus the influence of free energy considerations can be modulated. This approach thoroughly explores conformational space and generates an ensemble of secondary structures. The predictions from this new approach can test hypotheses and models of viral assembly and guide construction of complete three-dimensional models of virus particles.  相似文献   

18.
In vitro channel activity of the C-terminal colicin E1 channel polypeptide under conditions of variable electrostatic interaction with synthetic lipid membranes showed distinct maxima with respect to pH and membrane surface potential. The membrane binding energy was determined from fluorescence quenching of the intrinsic tryptophans of the channel polypeptide by liposomes containing N-trinitrophenyl-phosphatidylethanolamine. Maximum in vitro colicin channel activity correlated with an intermediate magnitude of the electrostatic interaction. For conditions associated with maximum activity (40% anionic lipid, I = 0.12 M, pH 4.0), the free energy of binding was delta G approximately -9 kcal/mol, with nonelectrostatic and electrostatic components, delta Gnel approximately -5 kcal/mol and delta Gel approximately -4 kcal/mol, and an effective binding charge of +7 at pH 4.0. Binding of the channel polypeptide to negative membranes at pH 8 is minimal, whereas initial binding at pH 4 followed by a shift to pH 8 causes only 3-10% reversal of binding, implying that it is kinetically trapped, probably by a hydrophobic interaction. It was inferred that membrane binding and insertion involves an initial electrostatic interaction responsible for concentration and binding to the membrane surface. This is followed by insertion into the bilayer driven by hydrophobic forces, which are countered in the case of excessive electrostatic binding.  相似文献   

19.
J Carey  O C Uhlenbeck 《Biochemistry》1983,22(11):2610-2615
A filter retention assay is used to examine the kinetic and equilibrium properties of the interaction between phage R17 coat protein and its 21-nucleotide RNA binding site. The kinetics of the reaction are consistent with the equilibrium association constant and indicate a diffusion-controlled reaction. The temperature dependence of Ka gives delta H = -19 kcal/mol. This large favorable delta H is partially offset by a delta S = -30 cal mol-1 deg-1 to give a delta G = -11 kcal/mol at 2 degrees C in 0.19 M salt. The binding reaction has a pH optimum centered around pH 8.5, but pH has no effect on delta H. While the interaction is insensitive to the type of monovalent cation, the affinity decreases with the lyotropic series among monovalent anions. The ionic strength dependence of Ka reveals that ionic contacts contribute to the interaction. Most of the binding free energy, however, is a result of nonelectrostatic interactions.  相似文献   

20.
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