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.
Isothermal titration calorimetry data for very low c (≡K[M]0) must normally be analyzed with the stoichiometry parameter n fixed — at its known value or at any reasonable value if the system is not well characterized. In the latter case, ΔH° (and hence n) can be estimated from the T-dependence of the binding constant K, using the van't Hoff (vH) relation. An alternative is global or simultaneous fitting of data at multiple temperatures. In this Note, global analysis of low-c data at two temperatures is shown to estimate ΔH° and n with double the precision of the vH method.  相似文献   

2.
Of all the techniques that are currently available to measure binding, isothermal titration calorimetry is the only one capable of measuring not only the magnitude of the binding affinity but also the magnitude of the two thermodynamic terms that define the binding affinity: the enthalpy (AH) and entropy (AS) changes. Recent advances in instrumentation have facilitated the development of experimental designs that permit the direct measurement of arbitrarily high binding affinities, the coupling of binding to protonation/deprotonation processes and the analysis of binding thermodynamics in terms of structural parameters. Because isothermal titration calorimetry has the capability to measure different energetic contributions to the binding affinity, it provides a unique bridge between computational and experimental analysis. As such, it is increasingly becoming an essential tool in molecular design.  相似文献   

3.
The thermodynamics of binding of the trivalent cations cobalt hexammine and spermidine to plasmid DNA was studied by isothermal titration calorimetry. Two stages were observed in the course of titration, the first attributed to cation binding and the second to DNA condensation. A standard calorimetric data analysis was extended by applying an electrostatic binding model, which accounted for most of the observed data. Both the binding and condensation reactions were entropically driven (TDeltaS approximately +10 kcal/mol cation) and enthalpically opposed (DeltaH approximately +1 kcal/mol cation). As predicted from their relative sizes, the binding constants of the cations were indistinguishable, but cobalt hexammine had a much greater DNA condensing capacity because it is more compact than spermidine. The dependence of both the free energy of cobalt hexammine binding and the critical cobalt hexammine concentration for DNA condensation on temperature and monovalent cation concentration followed the electrostatic model quite precisely. The heat capacity changes of both stages were positive, perhaps reflecting both the temperature dependence of the dielectric constant of water and the burial of polar surfaces. DNA condensation occurred when about 67 % of the DNA phosphate charge was neutralized by cobalt hexammine and 87 % by spermidine. During condensation, the remaining DNA charge was neutralized.  相似文献   

4.
A theoretical development in the evaluation of proton linkage in protein binding reactions by isothermal titration calorimetry (ITC) is presented. For a system in which binding is linked to protonation of an ionizable group on a protein, we show that by performing experiments as a function of pH in buffers with varying ionization enthalpy, one can determine the pK(a)'s of the group responsible for the proton linkage in the free and the liganded states, the protonation enthalpy for this group in these states, as well as the intrinsic energetics for ligand binding (delta H(o), delta S(o), and delta C(p)). Determination of intrinsic energetics in this fashion allows for comparison with energetics calculated empirically from structural information. It is shown that in addition to variation of the ligand binding constant with pH, the observed binding enthalpy and heat capacity change can undergo extreme deviations from their intrinsic values, depending upon pH and buffer conditions.  相似文献   

5.
Iron deposition within the iron storage protein ferritin involves a complex series of events consisting of Fe(2+) binding, transport, and oxidation at ferroxidase sites and mineralization of a hydrous ferric oxide core, the storage form of iron. In the present study, we have examined the thermodynamic properties of Fe(2+) binding to recombinant human H-chain apoferritin (HuHF) by isothermal titration calorimetry (ITC) in order to determine the location of the primary ferrous ion binding sites on the protein and the principal pathways by which the Fe(2+) travels to the dinuclear ferroxidase center prior to its oxidation to Fe(3+). Calorimetric titrations show that the ferroxidase center is the principal locus for Fe(2+) binding with weaker binding sites elsewhere on the protein and that one site of the ferroxidase center, likely the His65 containing A-site, preferentially binds Fe(2+). That only one site of the ferroxidase center is occupied by Fe(2+) implies that Fe(2+) oxidation to form diFe(III) species might occur in a stepwise fashion. In dilute anaerobic protein solution (3-5 microM), only 12 Fe(2+)/protein bind at pH 6.51 increasing to 24 Fe(2+)/protein at pH 7.04 and 7.5. Mutation of ferroxidase center residues (E62K+H65G) eliminates the binding of Fe(2+) to the center, a result confirming the importance of one or both Glu62 and His65 residues in Fe(2+) binding. The total Fe(2+) binding capacity of the protein is reduced in the 3-fold hydrophilic channel variant S14 (D131I+E134F), indicating that the primary avenue by which Fe(2+) gains access to the interior of ferritin is through these eight channels. The binding stoichiometry of the channel variant is one-third that of the recombinant wild-type H-chain ferritin whereas the enthalpy and association constant for Fe(2+) binding are similar for the two with an average values (DeltaH degrees = 7.82 kJ/mol, binding constant K = 1.48 x 10(5) M(-)(1) at pH 7.04). Since channel mutations do not completely prevent Fe(2+) binding to the ferroxidase center, iron gains access to the center in approximately one-third of the channel variant molecules by other pathways.  相似文献   

6.
We characterized the thermodynamics of binding reactions of nucleotides ADP and ATPgammaS (a nonhydrolyzable analog of ATP) to GroEL in a temperature range of 5 degrees C to 35 degrees C by isothermal titration calorimetry. Analysis with a noncooperative binding model has shown that the bindings of nucleotides are driven enthalpically with binding constants of 7x103 M-1 and 4x104 M-1 for ADP and ATPgammaS, respectively, at 26 degrees C and that the heat capacity change DeltaCp is about 100 cal/mol.K for both the nucleotides. The stoichiometries of binding were about 8 and 9 molecules for ADP and ATPgammaS, respectively, per GroEL tetradecamer at 5 degrees C, and both increased with temperature to reach about 14 (ADP) and 12 (ATPgammaS) for both nucleotides at 35 degrees C. The absence of initial increase of binding heat as well as Hill coefficient less than 1.2, which were obtained from the fitting to the model curve by assuming positive cooperativity, showed that there was virtually no positive cooperativity in the nucleotide bindings. Incorporating a difference in affinity for the nucleotide (ADP and ATPgammaS) between the two rings of GroEL into the noncooperative binding model improved the goodness of fitting and the difference in the affinity increased with decreasing temperature.  相似文献   

7.
Holdgate GA 《BioTechniques》2001,31(1):164-6, 168, 170 passim
Characterization of the thermodynamics of binding interactions is important in improving our understanding of bimolecular recognition and forms an essential part of the rational drug design process. Isothermal titration calorimetry (ITC) is rapidly becoming established as the method of choice for undertaking such studies. The power of ITC lies in its unique ability to measure binding reactions by the detection of the heat change during the binding interaction. Since heat changes occur during many physicochemical processes, ITC has a broad application, ranging from chemical and biochemical binding studies to more complex processes involving enthalpy changes, such as enzyme kinetics. Several features of ITC have facilitated its preferential use compared to other techniques that estimate affinity. It is a sensitive, rapid, and direct method with no requirement for chemical modification or immobilization. It is the only technique that directly measures enthalpy of binding and so eliminates the need for van't Hoff analysis, which can be time consuming and prone to uncertainty in parameter values. Although ITC has facilitated the measurement of the thermodynamics governing binding reactions, interpretation of these parameters in structural terms is still a major challenge.  相似文献   

8.
Talmard C  Bouzan A  Faller P 《Biochemistry》2007,46(47):13658-13666
Aggregation of the peptide amyloid-beta (Abeta) to amyloid plaques is a key event in Alzheimer's disease. According to the amyloid cascade hypothesis, Abeta aggregates are toxic to neurons via the production of reactive oxygen species and are hence directly involved in the cause of the disease. Zinc ions play an important role, because they are able to bind to Abeta and influence the aggregation properties. In the present work isothermal titration calorimetry and Zn sensors (zincon, Newport Green, and zinquin) were used to investigate the interaction of Zn with the full-length Abeta1-40 and Abeta1-42, as well as the truncated Abeta1-16 and Abeta1-28. The results suggest that Zn binding to Abeta induces a release of approximately 0.9 proton by the peptide. This correspond to the expected value upon Zn binding to the three histidines and indicates that further ligands are not deprotonated upon Zn binding. Such behavior is expected for carboxylates, but not the N-terminus. Moreover, the apparent dissociation constant (Kd,app) of Zn binding to all forms of Abeta is in the low micromolar range (1-20 microM) and rather independent of the aggregation state including soluble Abeta, Abeta fibrils, or Zn-induced Abeta aggregates. Finally, Zn in the soluble or aggregated Zn-Abeta form is well accessible for Zn chelators. The potential repercussions on metal chelation therapy are discussed.  相似文献   

9.
【目的】利用单针注射等温微量量热滴定法(SIITC)检测尿素水解菌对Hg~(2+)的耐受性。【方法】与传统的细胞生长曲线法和基于尿素水解的p H变化曲线相对比,应用SIITC研究耐盐尿素水解菌Staphylococcus succinus对不同浓度Hg~(2+)的耐受性,通过相关性分析验证SIITC检测细菌对重金属胁迫耐受性的可行性。【结果】S.succinus菌吸附水解尿素的最大热量交换速率和累积热量值随Hg~(2+)浓度升高而降低。最大热量交换速率由不含Hg~(2+)的-45μJ/s降至10μmol/L Hg~(2+)处理的-20μJ/s,吸收热量由-1 291.90μJ下降到-535.75μJ;不同浓度Hg~(2+)胁迫下S.succinus菌吸附水解尿素的反应热量值与溶液p H以及OD_(600)的相关系数r分别达到0.928和0.955。【结论】SIITC可用来检测Hg~(2+)对尿素水解菌S.succinus的毒性。  相似文献   

10.
The serine proteinase inhibitor antithrombin III (ATIII) is a key regulatory protein of intrinsic blood coagulation. ATIII attains its full biological activity only upon binding polysulfated oligosaccharides, such as heparin. A series of synthetic peptides have been prepared based on the proposed heparin binding regions of ATIII and their ability to bind heparin has been assessed by CD spectrometry, by isothermal titration calorimetry, and by the ability of the peptides to compete with ATIII for binding heparin in a factor Xa procoagulant enzyme assay. Peptide F123-G148, which encompasses both the purported high-affinity pentasaccharide binding region and an adjacent, C-terminally directed segment of ATIII, was found to bind heparin with good affinity, but amino-terminal truncations of this sequence, including L130-G148 and K136-G148 displayed attenuated heparin binding activities. In fact, K136-G148 appears to encompass only a low-affinity heparin binding site. In contrast, peptides based solely on the high-affinity binding site (K121-A134) displayed much higher affinities for heparin. By CD spectrometry, these high-affinity peptides are chiefly random coil in nature, but low microM concentrations of heparin induce significant alpha-helix conformation. K121-A134 also effectively competes with ATIII for binding heparin. Thus, through the use of synthetic peptides that encompass part, if not all, of the heparin binding site(s) within ATIII, we have further elucidated the structure-function relations of heparin-ATIII interactions.  相似文献   

11.
BackgroundSuccessful ITC experiments require conversion of cell reagent (titrand M) to product and production or consumption of heat. These conditions are quantified for 1:1 binding, M + X ⇔ MX.MethodsNonlinear least squares is used in error-propagation mode to predict the precisions with which the key quantities — binding constant K, reaction enthalpy ΔH°, and stoichiometry number n — can be estimated over a wide range of the dimensionless quantity that governs isotherm shape, c = K[M]0. The measurement precision σq is estimated from analysis of water–water blanks.ResultsWhen the product conversion exceeds 90%, the parameter relative standard errors are proportional to σq/qtot, where the total heat qtot ≈ ΔH° [M]0 V0. Specifically, σK/K × qtotq ≈ 25 for c = 10  3  10, ≈ 11 c1/3 for c = 10  104. For c > 1, n and ΔH° are more precise than K; this holds also at smaller c for the product n × ΔH° and for ΔH° when n can be held fixed. Use of as few as 10 titrant injections can outperform the customary 20–40 while also improving productivity.ConclusionThese principles are illustrated in experiment design using the program ITC-PLANNER15.General significanceSimple quantitative guidelines replace the “c rules” that have dominated the literature for decades. This article is part of a Special Issue entitled Microcalorimetry in the BioSciences — Principles and Applications, edited by Fadi Bou-Abdallah.  相似文献   

12.
The method of generalized least squares (GLS) is used to assess the variance function for isothermal titration calorimetry (ITC) data collected for the 1:1 complexation of Ba(2+) with 18-crown-6 ether. In the GLS method, the least squares (LS) residuals from the data fit are themselves fitted to a variance function, with iterative adjustment of the weighting function in the data analysis to produce consistency. The data are treated in a pooled fashion, providing 321 fitted residuals from 35 data sets in the final analysis. Heteroscedasticity (nonconstant variance) is clearly indicated. Data error terms proportional to q(i) and q(i)/v are well defined statistically, where q(i) is the heat from the ith injection of titrant and v is the injected volume. The statistical significance of the variance function parameters is confirmed through Monte Carlo calculations that mimic the actual data set. For the data in question, which fall mostly in the range of q(i)=100-2000 microcal, the contributions to the data variance from the terms in q(i)(2) typically exceed the background constant term for q(i)>300 microcal and v<10 microl. Conversely, this means that in reactions with q(i) much less than this, heteroscedasticity is not a significant problem. Accordingly, in such cases the standard unweighted fitting procedures provide reliable results for the key parameters, K and DeltaH(degrees) and their statistical errors. These results also support an important earlier finding: in most ITC work on 1:1 binding processes, the optimal number of injections is 7-10, which is a factor of 3 smaller than the current norm. For high-q reactions, where weighting is needed for optimal LS analysis, tips are given for using the weighting option in the commercial software commonly employed to process ITC data.  相似文献   

13.
Isothermal titration calorimetry (ITC) is an important technique used in quantitatively analyzing the global mechanism of protein-protein or protein-ligand interactions through thermodynamic measurements. Among different binding mechanisms, the parallel and ligand induced protein oligomerization mechanisms are technically difficult to analyze compared with a sequential binding mechanism. Here, we present a methodology implemented as a program "Open-ITC" that eliminates the need for exact analytical expressions for free ligand concentrations [L] and mole fractions of bound ligand θ that are required for the thermogram analysis. Adopting a genetic algorithm-based optimization, the thermodynamic parameters are determined, and its standard error is evaluated at the global minimum by calculating the Jacobian matrix. This approach yielded a statistically consistent result for a single-site and a two-site binding protein-ligand system. Further, a comparative simulation of a two-step sequential, a parallel, and a ligand induced oligomerization model revealed that their mechanistic differences are discernable in ITC thermograms, only if the first binding step is weaker compared with the second binding step (K(1) 相似文献   

14.
15.
16.
Yeast cytochrome c peroxidase (CcP) and horse metmyoglobin (Mb) bind HN3 with similar affinities at 25 degrees C. The pH-independent equilibrium association constants for formation of the CcP.HN3 and Mb.HN3 complexes are (1.05 +/- 0.06)x10(5) and (1.6 +/- 0.8)x10(5) M(-1), respectively. However, the thermodynamic parameters for formation of the two complexes are quite different. The DeltaH0 values for formation of CcP.HN3 and Mb.HN3 are -16.4 +/- 0.7 and -9.0 +/- 0.5 kcal/mol, respectively, and the Delta S0 values are -32 +/- 2 and -16 +/- 2 cal/deg mol, respectively. The proton associated with HN3 is retained in both protein complexes at low pH but dissociates with apparent pKA values of 5.5 +/- 0.2 and > or =8.2 for the Mb.HN3 and CcP.HN3 complexes, respectively. CcP and Mb differ significantly in their reactivity toward the azide anion, N3-. CcP binds N3- very weakly, if at all, and only an upper-limit of 18 +/-5 M(-1) for the pH-independent equilibrium association constant for the CcP.N3- complex can be determined. Mb binds N3- with an association constant of (1.8 +/- 0.1)x10(4) M(-1). The ratio of the equilibrium association constants for HN3 and N3- binding provides a discrimination factor between the neutral and charged forms of the ligand. The discrimination factor is greater than 5800 for CcP but only nine for Mb. Protonation of the distal histidines in the two proteins influences binding of HN3. Protonation of His-64 in Mb enhances HN3 binding due to a gating mechanism while protonation of His-52 in CcP decreases the affinity for HN3 due to loss of base-assisted association of the ligand to the heme iron.  相似文献   

17.
The peptides Asp-Ala-His-Lys (DAHK) and Gly-His-Lys (GHK) are naturally occurring Cu(II)-chelating motifs in human serum and cerebrospinal fluid. Here, the sensitive thermodynamic technique isothermal titration calorimetry was used to study the energetics of Cu(II) binding to DAHK and GHK peptides in the presence of the weaker ligand glycine as a competitor. DAHK and GHK bind Cu(II) predominantly in a 1:1 stoichiometry with conditional dissociation constants [i.e., at pH 7.4, in the absence of the competing chelators glycine and 2-(4-(2-hydroxyethyl)-1-piperazinyl)ethanesulfonic acid buffer] of 2.6 ± 0.4 × 10−14 M and 7.0 ± 1.0 × 10−14 M, respectively. Furthermore, the apparent ΔH values were measured and the number of protons released upon Cu(II) binding was determined by performing experiments in different buffers. This allowed us to determine the conditional ΔG, ΔH, and ΔS, i.e., corrected for the contributions of the weaker ligand glycine and the buffer at pH 7.4. We found that the entropic and enthalpic contributions to the Cu(II) binding to GHK and DAHK are distinct, with a enthalpic contribution for GHK. The thermodynamic parameters obtained correspond well to those in the literature obtained by other techniques, suggesting that the use of the weaker ligand glycine as a competitor in isothermal titration calorimetry provides accurate data for Cu(II) binding to high-affinity peptides, which cannot be accurately determined without the use of a competitor ligand.  相似文献   

18.
Interactions of bisANS and ANS to tubulin in the presence and absence of GTP were investigated, and the binding and thermodynamic parameters were determined using isothermal titration calorimetry. Like bisANS binding to tubulin, we observed a large number of lower affinity ANS binding sites (N1 = 1.3, K1 = 3.7 x 10(5) M(-1), N2 = 10.5, K2 = 7 x 10(4)/M(-1)) in addition to 1-2 higher affinity sites. Although the presence of GTP lowers the bisANS binding to both higher and lower affinity sites (N1 = 4.3, N2 = 11.7 in absence and N1 = 1.8, N2 = 3.6 in presence of GTP), the stoichiometries of both higher and lower affinity sites of ANS remain unaffected in the presence of GTP. BisANS-induced structural changes on tubulin were studied using site-specific proteolysis with trypsin and chymotrypsin. Digestion of both alpha and beta tubulin with trypsin and chymotrypsin, respectively, has been found to be very specific in presence of GTP. GTP has dramatic effects on lowering the extent of nonspecific digestion of beta tubulin with trypsin and stabilizing the intermediate bands produced from both alpha and beta. BisANS-treated tubulin is more susceptible to both trypsin and chymotrypsin digestion. At higher bisANS concentration (>20 microM) both alpha and beta tubulins are almost totally digested with enzymes, indicating bisANS-induced unfolding or destabilization of tubulin structure. Again, the addition of GTP has remarkable effect on lowering the bisANS-induced enhanced digestion of tubulin as well as stabilizing effect on intermediate bands. These results of isothermal titration calorimetry, proteolysis and the DTNB-kinetics data clearly established that the addition of GTP makes tubulin compact and rigid and hence the GTP-induced stabilization of tubulin structure. No such destabilization of tubulin structure has been noticed with ANS, although, like bisANS, ANS possesses a large number of lower affinity binding sites. On the basis of these results, we propose that the unique structure of bisANS, which in absence of GTP can bind tubulin as a bifunctional ligand (through its two ANS moieties), is responsible for the structural changes of tubulin.  相似文献   

19.
Synthetic peptides based on amino-acid residues 27-38 of human serum amyloid P component represent a novel type of heparin binders as they do not contain clusters of basic amino acids or other known features associated with protein or peptide heparin binding. Here, we characterize the binding using capillary electrophoresis (CE), surface plasmon resonance (SPR), and isothermal titration calorimetry (ITC). By CE, heparin-binding activity was readily apparent for both a regular peptide and a slightly N-terminally modified form, while a sequence-scrambled peptide had no measurable binding. Dissociation constants in the 1-15 microm range were estimated, but only a minor part of the binding isotherm was covered by the experiments. SPR measurements using immobilized peptides verified heparin binding, the range of the binding constants, and the reduced binding of the sequence-scrambled peptide. Structurally defined heparin oligosaccharides were used to establish that while the tetrasaccharide is too small to exhibit strong binding, little difference in binding strength is observed between hexa- and tetradeca-saccharides. These experiments also confirmed the almost complete lack of activity of the sequence-scrambled peptide. The amino-acid sequence-dependent binding and the importance of a disulfide bond in the peptide were verified by ITC, but the experimental conditions had to be modified because of peptide precipitation and ITC yielded significantly weaker binding constants than the other methods. While the precise function of the peptide in the intact protein remains unclear, the results confirm the specificity of the glycosaminoglycan interaction with regard to peptide sequence by applying two additional biophysical techniques and showing that the N-terminal part of the peptide may be modified without changing the heparin binding capabilities.  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