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 31 毫秒
1.
A G Kozlov  T M Lohman 《Biochemistry》1999,38(22):7388-7397
Isothermal titration calorimetry (ITC) was used to test the hypothesis that the relatively small enthalpy change (DeltaHobs) and large negative heat capacity change (DeltaCp,obs) observed for the binding of the Escherichia coli SSB protein to single-stranded (ss) oligodeoxyadenylates result from the temperature-dependent adenine base unstacking equilibrium that is thermodynamically coupled to binding. We have determined DeltaH1,obs for the binding of 1 mole of each of dT(pT)34, dC(pC)34, and dA(pA)34 to the SSB tetramer (20 mM NaCl at pH 8.1). For dT(pT)34 and dC(pC)34, we found large, negative values for DeltaH1,obs of -75 +/- 1 and -85 +/- 2 kcal/mol at 25 degrees C, with DeltaCp,obs values of -540 +/- 20 and -570 +/- 30 cal mol-1 K-1 (7-50 degrees C), respectively. However, for SSB-dA(pA)34 binding, DeltaH1,obs is considerably less negative (-14 +/- 1 kcal/mol at 25 degrees C), even becoming positive at temperatures below 13 degrees C, and DeltaCp,obs is nearly twice as large in magnitude (-1180 +/- 40 cal mol-1 K-1). These very different thermodynamic properties for SSB-dA(pA)34 binding appear to result from the fact that the bases in dA(pA)34 are more stacked at any temperature than are the bases in dC(pC)34 or dT(pT)34 and that the bases become unstacked within the SSB-ssDNA complexes. Therefore, the DeltaCp,obs for SSB-ssDNA binding has multiple contributions, a major one being the coupling to binding of a temperature-dependent conformational change in the ssDNA, although SSB binding to unstacked ssDNA still has an "intrinsic" negative DeltaCp,0. In general, such temperature-dependent changes in the conformational "end states" of interacting macromolecules can contribute significantly to both DeltaCp,obs and DeltaHobs.  相似文献   

2.
Little is known about the thermodynamic forces that drive the folding pathways of higher-order RNA structure. In this study, we employ calorimetric [isothermal titration calorimetry (ITC) and differential scanning calorimetry (DSC)] and spectroscopic (NMR and UV) methods to characterize the thermodynamics of the GAAA tetraloop-receptor interaction, utilizing a previously described bivalent construct. ITC studies indicate that the bivalent interaction is enthalpy driven and highly stable, with a binding constant (K(obs)) of 5.5x10(6) M(-1) and enthalpy (DeltaH(obs)(o)) of -33.8 kcal/mol at 45 degrees C in 20 mM KCl and 2 mM MgCl(2). Thus, we derive the DeltaH(obs)(o) for a single tetraloop-receptor interaction to be -16.9 kcal/mol at these conditions. UV absorbance data indicate that an increase in base stacking quality contributes to the enthalpy of complex formation. These highly favorable thermodynamics are consistent with the known critical role for the tetraloop-receptor motif in the folding of large RNAs. Additionally, a significant heat capacity change (DeltaC(p,obs)(o)) of -0.24 kcal mol(-1) K(-1) was determined by ITC. DSC and UV-monitored thermal denaturation experiments indicate that the bivalent tetraloop-receptor construct follows a minimally five-state unfolding pathway and suggest the observed DeltaC(p,obs)(o) for the interaction results from a temperature-dependent unbound receptor RNA structure.  相似文献   

3.
The peripheral subunit-binding domain (PSBD) of the dihydrolipoyl acetyltransferase (E2, EC 2.3.1.12) binds tightly but mutually exclusively to dihydrolipoyl dehydrogenase (E3, EC 1.8.1.4) and pyruvate decarboxylase (E1, EC 1.2.4.1) in the pyruvate dehydrogenase multienzyme complex of Bacillus stearothermophilus. Isothermal titration calorimetry (ITC) experiments demonstrated that the enthalpies of binding (DeltaH degrees ) of both E3 and E1 with the PSBD varied with salt concentration, temperature, pH, and buffer composition. There is little significant difference in the free energies of binding (DeltaG degrees = -12.6 kcal/mol for E3 and = -12.9 kcal/mol for E1 at pH 7.4 and 25 degrees C). However, the association with E3 was characterized by a small, unfavorable enthalpy change (DeltaH degrees = +2.2 kcal/mol) and a large, positive entropy change (TDeltaS degrees = +14.8 kcal/mol), whereas that with E1 was accompanied by a favorable enthalpy change (DeltaH degrees = -8.4 kcal/mol) and a less positive entropy change (TDeltaS degrees = +4.5 kcal/mol). Values of DeltaC(p) of -316 cal/molK and -470 cal/molK were obtained for the binding of E3 and E1, respectively. The value for E3 was not compatible with the DeltaC(p) calculated from the nonpolar surface area buried in the crystal structure of the E3-PSBD complex. In this instance, a large negative DeltaC(p) is not indicative of a classical hydrophobic interaction. In differential scanning calorimetry experiments, the midpoint melting temperature (T(m)) of E3 increased from 91 degrees C to 97.1 degrees C when it was bound to PSBD, and that of E1 increased from 65.2 degrees C to 70.0 degrees C. These high T(m) values eliminate unfolding as a major source of the anomalous DeltaC(p) effects at the temperatures (10-37 degrees C) used for the ITC experiments.  相似文献   

4.
Horn JR  Brandts JF  Murphy KP 《Biochemistry》2002,41(23):7501-7507
The complexity of binding reactions, including the linkage with other equilibria, is becoming increasingly apparent in biological processes such as signal transduction. Understanding these interactions requires obtaining thermodynamic profiles for each of the equilibria that occur in a binding event. Concern has been raised as to whether linked equilibria contribute differently to thermodynamics, such as DeltaH degrees and DeltaC(p), obtained from calorimetric and van't Hoff methods. We have previously shown that linked equilibria do not contribute differently to the van't Hoff and calorimetrically determined DeltaH degrees for processes such as linked folding or hydration. Here, examples of proton and ion linkage are examined. We show that there is no reason to expect the calorimetric and van't Hoff DeltaH degrees to be different, even without prior knowledge of the presence or absence of linked equilibria, as long as the system is permitted to equilibrate. However, it is possible to create experimental scenarios that result in and discrepancies. Furthermore, it is found that the presence of linked equilibria in all cases can result in "nonconventional" DeltaH degrees and DeltaC(p) profiles, making data analysis nontrivial.  相似文献   

5.
A technique based on fluorescence polarization (anisotropy) was used to measure the binding of antibodies to DNA under a variety of conditions. Fluorescein-labeled duplexes of 20 bp in length were employed as the standard because they are stable even at low ionic strength yet sufficiently short so that both arms of an IgG cannot bind to the same duplex. IgG Jel 274 binds duplexes in preference to single-stranded DNA; in 80 mM NaCl Kobs for (dG)20.(dC)20 is 4.1x10(7) M-1 compared with 6.4x10(5) M-1 for d(A5C10A5). There is little sequence specificity, but the interaction is very dependent on ionic strength. From plots of log Kobs against log[Na+] it was deduced that five or six ion pairs are involved in complex formation. At low ionic strength,Kobs is independent of temperature and complex formation is entropy driven with DeltaH degrees obs and DeltaC degrees p,obs both zero. In contrast, in 80 mM NaCl DeltaC degrees p,obs is -630 and -580 cal mol-1K-1 for [d(TG)]10.[d(CA)]10 and (dG)20.(dC)20 respectively. IgG Jel 241 also binds more tightly to duplexes than single-stranded DNA, but sequence preferences were apparent. The values for Kobs to [d(AT)]20 and [d(GC)]20 are 2.7x10(8) and 1.3x10(8) M-1 respectively compared with 5.7x10(6) M-1 for both (dA)20. (dT)20 and (dG)20.(dC)20. As with Jel 274, the binding of Jel 241 is very dependent on ionic strength and four or five ionic bonds are involved in complex formation with all the duplex DNAs which were tested. DeltaC degrees p,obs for Jel 241 binding to [d(AT)]20 was negative (-87 cal mol-1K-1) in 80 mM NaCl but was zero at high ionic strength (130 mM NaCl). Therefore, for duplex-specific DNA binding antibodies DeltaC degrees p,obs is dependent on [Na+] and a large negative value does not correlate with sequence-specific interactions.  相似文献   

6.
The alphabeta T cell receptor (TCR) is responsible for recognizing peptides bound and "presented" by major histocompatibility complex (MHC) molecules. We recently reported that at 25 degrees C the A6 TCR, which recognizes the Tax peptide presented by the class I MHC human leukocyte antigen-A*0201 (HLA-A2), binds with a weak DeltaH degrees , a favorable DeltaS degrees , and a moderately negative DeltaC(p). These observations were of interest given the unfavorable binding entropies and large heat capacity changes measured for many other TCR-ligand interactions, suggested to result from TCR conformational changes occurring upon binding. Here, we further investigated the A6-Tax/HLA-A2 interaction using titration calorimetry. We found that binding results in a pK(a) shift, complicating interpretation of measured binding thermodynamics. To better characterize the interaction, we measured binding as a function of pH, temperature, and buffer ionization enthalpy. A global analysis of the resulting data allowed determination of both the intrinsic binding thermodynamics separated from the influence of protonation as well as the thermodynamics associated with the pK(a) shift. Our results indicate that intrinsically, A6 binds Tax/HLA-A2 with a very weak DeltaH degrees , an even more favorable DeltaS degrees than previously thought, and a relatively large negative DeltaC(p). Comparison of these energetics with the makeup of the protein-protein interface suggests that conformational adjustments are required for binding, but these are more likely to be structural shifts, rather than disorder-to-order transitions. The thermodynamics of the pK(a) shift suggest protonation may be linked to an additional process such as ion binding.  相似文献   

7.
We showed that the alpha-CH(2) --> NH substitution in octanoyl-CoA alters the ground and transition state energies for the binding of the CoA ligands to medium-chain acyl-CoA dehydrogenase (MCAD), and such an effect is caused by a small electrostatic difference between the ligands. To ascertain the extent that the electrostatic contribution of the ligand structure and/or the enzyme site environment modulates the thermodynamics of the enzyme-ligand interaction, we undertook comparative microcalorimetric studies for the binding of 2-azaoctanoyl-CoA (alpha-CH(2) --> NH substituted octanoyl-CoA) and octenoyl-CoA to the wild-type and Glu-376 --> Gln mutant enzymes. The experimental data revealed that both enthalpy (DeltaH degrees ) and heat capacity changes (DeltaC(p) degrees ) for the binding of 2-azaoctanoyl-CoA (DeltaH degrees (298) = -21.7 +/- 0.8 kcal/mole, DeltaC(p) degrees = -0.627 +/- 0.04 kcal/mole/K) to the wild-type MCAD were more negative than those obtained for the binding of octenoyl-CoA (DeltaH degrees (298) = -17.2 +/- 1.6 kcal/mole, DeltaC(p) degrees = -0.526 +/- 0.03 kcal/mole/K). Of these, the decrease in the magnitude of DeltaC(p) degrees for the binding of 2-azaoctanoyl-CoA (vis-à-vis octenoyl-CoA) to the enzyme was unexpected, because the former ligand could be envisaged to be more polar than the latter. To our further surprise, the ligand-dependent discrimination in the above parameters was completely abolished on Glu-376 --> Gln mutation of the enzyme. Both DeltaH degrees and DeltaC(p) degrees values for the binding of 2-azaoctanoyl-CoA (DeltaH degrees (298) = -13.3 +/- 0.6 kcal/mole, DeltaC(p) degrees = -0.511 +/- 0.03 kcal/mole/K) to the E376Q mutant enzyme were found to be correspondingly identical to those obtained for the binding of octenoyl-CoA (DeltaH degrees (298) = -13.2 +/- 0.6 kcal/mole, DeltaC(p) degrees = -0.520 +/- 0.02 kcal/mole/K). However, in neither case could the experimentally determined DeltaC(p) degrees values be predicted on the basis of the changes in the water accessible surface areas of the enzyme and ligand species. Arguments are presented that the origin of the above thermodynamic differences lies in solvent reorganization and water-mediated electrostatic interaction between ligands and enzyme site groups, and such interactions are intrinsic to the molecular basis of the enzyme-ligand complementarity.  相似文献   

8.
Specific recognition of the mRNA 5' cap by eukaryotic initiation factor eIF4E is a rate-limiting step in the translation initiation. Fluorescence spectroscopy and high-sensitivity isothermal titration calorimetry were used to examine the thermodynamics of eIF4E binding to a cap-analogue, 7-methylGpppG. A van't Hoff plot revealed nonlinearity characterized by an unexpected, large positive molar heat capacity change (DeltaC(degree)(p) = +1.92 +/- 0.93 kJ.mol(-1).K(-1)), which was confirmed by direct ITC measurements (DeltaC(degree)(p) = +1.941 +/- 0.059 kJ.mol(-1).K(-1)). This unique result appears to come from an extensive additional hydration upon binding and charge-related interactions within the binding site. As a consequence of the positive DeltaC(degree)(p), the nature of the thermodynamic driving force changes with increasing temperature, from enthalpy-driven and entropy-opposed, through enthalpy- and entropy-driven in the range of biological temperatures, into entropy-driven and enthalpy-opposed. Comparison of the van't Hoff and calorimetric enthalpy values provided proof for the ligand protonation at N(1) upon binding, which is required for tight stabilization of the cap-eIF4E complex. Intramolecular self-stacking of the dinucleotide cap-analogue was analyzed to reveal the influence of this coupled process on the thermodynamic parameters of the eIF4E-mRNA 5' cap interaction. The temperature-dependent change in the conformation of 7-methylGpppG shifts significantly the intrinsic DeltaH(degree)(0) = -72.9 +/- 4.2 kJ.mol(-1) and DeltaS(degree)(0) = -116 +/- 58 J.mol(-1).K(-1) of binding to the less negative resultant values, by DeltaH(degree)(sst) = +9.76 +/- 1.15 kJ.mol(-1) and DeltaS(degree)(sst) = +24.8 +/- 2.1 J.mol(-1).K(-1) (at 293 K), while the corresponding DeltaC(degree)(p)(sst) = -0.0743 +/- 0.0083 kJ.mol(-1).K(-1) is negligible in comparison with the total DeltaC(degree)(p) .  相似文献   

9.
OCP1 and OCP2, the most abundant proteins in the cochlea, are putative subunits of an SCF E3 ubiquitin ligase. Previous work has demonstrated that they form a heterodimeric complex. The thermodynamic details of that interaction are herein examined by isothermal titration calorimetry. At 25 degrees C, addition of OCP1 to OCP2 yields an apparent association constant of 4.0 x 10(7) M(-1). Enthalpically-driven (DeltaH=-35.9 kcal/mol) and entropically unfavorable (-TDeltaS=25.5 kcal/mol), the reaction is evidently unaccompanied by protonation/deprotonation events. DeltaH is strongly dependent on temperature, with DeltaC(p)=-1.31 kcal mol(-1) K(-1). Addition of OCP2 to OCP1 produces a slightly less favorable DeltaH, presumably due to the requirement for dissociation of the OCP2 homodimer prior to OCP1 binding. The thermodynamic signature for OCP1/OCP2 complex formation is inconsistent with a rigid-body association and suggests that the reaction is accompanied by a substantial degree of folding.  相似文献   

10.
Grucza RA  Fütterer K  Chan AC  Waksman G 《Biochemistry》1999,38(16):5024-5033
The cytosolic tyrosine kinase Syk is recruited to immune cell receptors via interactions of its tandem-SH2 domain with tyrosine-phosphorylated sequences called immune receptor tyrosine activation motifs (ITAMs). We have characterized the binding of the tandem-SH2 domain of Syk (Syk-tSH2) to a dually phosphorylated peptide derived from the ITAM of the T cell receptor CD3-epsilon subunit. The CD3-epsilon peptide binds with an affinity of 18-81 nM at 150 mM NaCl over the 4.5-30 degrees C temperature range that was studied. The enthalpy of binding, DeltaH degrees obs, shows an unusual nonlinear dependence on temperature, suggesting the possibility of a temperature-dependent conformational equilibrium coupled to binding. This hypothesis was tested and confirmed by examining the temperature dependence of (1) the on-rate constant for binding and (2) the fluorescence of Syk-tSH2 and its CD3-epsilon peptide complex. The DeltaH degrees obs, Kobs, fluorescence, and kinetic data are all well described by a model incorporating the hypothesized conformational equilibrium. Circular dichroism spectra at various temperatures indicate that the conformational change is not due to a partial unfolding of the protein. We suggest that the conformational equilibrium enables Syk-tSH2 to exhibit flexibility in its binding modality, which may be necessitated by Syk's involvement in a wide variety of signal tranduction pathways.  相似文献   

11.
12.
RecA protein recognises two complementary DNA strands for homologous recombination. To gain insight into the molecular mechanism, the thermodynamic parameters of the DNA binding have been characterised by isothermal calorimetry. Specifically, conformational changes of protein and DNA were searched for by measuring variations in enthalpy change (DeltaH) with temperature (heat capacity change, DeltaC(p)). In the presence of the ATP analogue ATPgammaS, the DeltaH for the binding of the first DNA strand depends upon temperature (large DeltaC(p)) and the type of buffer, in a way that is consistent with the organisation of disordered parts and the protonation of RecA upon complex formation. In contrast, the binding of the second DNA strand occurs without any pronounced DeltaC(p), indicating the absence of further reorganisation of the RecA-DNA filament. In agreement with these findings, a significant change in the CD spectrum of RecA was observed only upon the binding of the first DNA strand. In the absence of nucleotide cofactor, the DeltaH of DNA binding is almost independent of temperature, indicating a requirement for ATP in the reorganisation of RecA. When the second DNA strand is complementary to the first, the DeltaH is larger than that for non-complementary DNA strand, but less than the DeltaH of the annealing of the complementary DNA without RecA. This small DeltaH could reflect a weak binding that may facilitate the dissociation of only partly complementary DNA and thus speed the search for complementary DNA. The DeltaH of binding DNA sequences displaying strong base-base stacking is small for both the first and second binding DNA strand, suggesting that the second is also stretched upon interaction with RecA. These results support the proposal that the RecA protein restructures DNA, preparing it for the recognition of a complementary second DNA strand, and that the recognition is due mainly to direct base-base contacts between DNA strands.  相似文献   

13.
Arnulphi C  Jin L  Tricerri MA  Jonas A 《Biochemistry》2004,43(38):12258-12264
The interaction of lipid-free apolipoprotein A-I (apoA-I) with small unilamellar vesicles (SUVs) of 1-palmitoyl-2-oleoylphosphatidylcholine (POPC) with and without free cholesterol (FC) was studied by isothermal titration calorimetry and circular dichroism spectroscopy. Parameters reported are the affinity constant (K(a)), the number of protein molecules bound per vesicle (n), enthalpy change (DeltaH degrees), entropy change (DeltaS degrees ), and the heat capacity change (DeltaC(p) degrees). The binding process of apoA-I to SUVs of POPC plus 0-20% (mole) FC was exothermic between 15 and 37 degrees C studied, accompanied by a small negative entropy change, making enthalpy the main driving force of the interaction. The presence of cholesterol in the vesicles increased the binding affinity and the alpha-helix content of apoA-I but lowered the number of apoA-I bound per vesicle and the enthalpy and entropy changes per bound apoA-I. Binding affinity and stoichiometry were essentially invariant of temperature for binding to SUVs of POPC/FC at a molar ratio of 6/1 at (2.8-4) x 10(6) M(-1) and 2.4 apoA-I molecules bound per vesicle or 1.4 x 10(2) phospholipids per bound apoA-I. A plot of DeltaH degrees against temperature displayed a linear behavior, from which the DeltaC(p) degrees per mole of bound apoA-I was calculated to be -2.73 kcal/(mol x K). These results suggested that binding of apoA-I to POPC vesicles is characterized by nonclassical hydrophobic interactions, with alpha-helix formation as the main driving force for the binding to cholesterol-containing vesicles. In addition, comparison to literature data on peptides suggested a cooperativity of the helices in apoA-I in lipid interaction.  相似文献   

14.
Sawas AH  Pentyala SN  Rebecchi MJ 《Biochemistry》2004,43(39):12675-12685
This study directly examines the enthalpic contributions to binding in aqueous solution of closely related anesthetic haloethers (desflurane, isoflurane, enflurane, and sevoflurane), a haloalkane (halothane), and an intravenous anesthetic (propofol) to bovine and human serum albumin (BSA and HSA) using isothermal titration calorimetry. Binding to serum albumin is exothermic, yielding enthalpies (DeltaH(obs)) of -3 to -6 kcal/mol for BSA with a rank order of apparent equilibrium association constants (K(a) values): desflurane > isoflurane approximately enflurane > halothane >or= sevoflurane, with the differences being largely ascribed to entropic contributions. Competition experiments indicate that volatile anesthetics, at low concentrations, share the same sites in albumin previously identified in crystallographic and photo-cross-linking studies. The magnitude of the observed DeltaH increased linearly with increased reaction temperature, reflecting negative changes in heat capacities (DeltaC(p)). These -DeltaC(p) values significantly exceed those calculated for burial of each anesthetic in a hydrophobic pocket. The enhanced stabilities of the albumin/anesthetic complexes and -DeltaC(p) are consistent with favorable solvent rearrangements that promote binding. This idea is supported by substitution of D(2)O for H(2)O that significantly reduces the favorable binding enthalpy observed for desflurane and isoflurane, with an opposing increase of DeltaS(obs). From these results, we infer that solvent restructuring, resulting from release of water weakly bound to anesthetic and anesthetic-binding sites, is a dominant and favorable contributor to the enthalpy and entropy of binding to proteins.  相似文献   

15.
Although bovine beta-lactoglobulin assumes a monomeric native structure at pH 3 in the absence of salt, the addition of salts stabilizes the dimer. Thermodynamics of the monomer-dimer equilibrium dependent on the salt concentration were studied by sedimentation equilibrium. The addition of NaCl, KCl, or guanidine hydrochloride below 1 M stabilized the dimer in a similar manner. On the other hand, NaClO(4) was more effective than other salts by about 20-fold, suggesting that anion binding is responsible for the salt-induced dimer formation, as observed for acid-unfolded proteins. The addition of guanidine hydrochloride at 5 M dissociated the dimer into monomers because of the denaturation of protein structure. In the presence of either NaCl or NaClO(4), the dimerization constant decreased with an increase in temperature, indicating that the enthalpy change (DeltaH(D)) of dimer formation is negative. The heat effect of the dimer formation was directly measured with an isothermal titration calorimeter by titrating the monomeric beta-lactoglobulin at pH 3.0 with NaClO(4). The net heat effects after subtraction of the heat of salt dilution, corresponding to DeltaH(D), were negative, and were consistent with those obtained by the sedimentation equilibrium. From the dependence of dimerization constant on temperature measured by sedimentation equilibrium, we estimated the DeltaH(D) value at 20 degrees C and the heat capacity change (DeltaC(p)) of dimer formation. In both NaCl and NaClO(4), the obtained DeltaC(p) value was negative, indicating the dominant role of burial of the hydrophobic surfaces upon dimer formation. The observed DeltaC(p) values were consistent with the calculated value from the X-ray dimeric structure using a method of accessible surface area. These results indicated that monomer-dimer equilibrium of beta-lactoglobulin at pH 3 is determined by a subtle balance of hydrophobic and electrostatic effects, which are modulated by the addition of salts or by changes in temperature.  相似文献   

16.
Kozlov AG  Lohman TM 《Biochemistry》2002,41(19):6032-6044
We have examined the kinetic mechanism for binding of the homotetrameric Escherichia coliSSB protein to single-stranded oligodeoxynucleotides [(dT)(70) and (dT)(35)] under conditions that favor the formation of a fully wrapped ssDNA complex in which all four subunits interact with DNA. Under these conditions, a so-called (SSB)(65) complex is formed in which either one molecule of (dT)(70) or two molecules of (dT)(35) bind per tetramer. Stopped-flow studies monitoring quenching of the intrinsic SSB Trp fluorescence were used to examine the initial binding step. To examine the kinetics of ssDNA wrapping, we used a single-stranded oligodeoxythymidylate, (dT)(66), that was labeled on its 3'-end with a fluorescent donor (Cy3) and on its 5'-end with a fluorescent acceptor (Cy5). Formation of the fully wrapped structure was accompanied by extensive fluorescence resonance energy transfer (FRET) from Cy3 to Cy5 since the two ends of (dT)(66) are in close proximity in the fully wrapped complex. Our results indicate that initial ssDNA binding to the tetramer is very rapid, with a bimolecular rate constant, k(1,app), of nearly 10(9) M(-1) s(-1) in the limit of low salt concentration (<0.2 M NaCl, pH 8.1, 25.0 degrees C), whereas the rate of dissociation is very low at all salt concentrations that were examined (20 mM to 2 M NaCl or NaBr). However, the rate of initial binding and the rate of formation of the fully wrapped complex are identical, indicating that the rate of wrapping of the ssDNA around the SSB tetramer is very rapid, with a lower limit rate of 700 s(-1). The implications of this rapid binding and wrapping reaction are discussed.  相似文献   

17.
The cardiac-specific Nkx2.5 homeodomain has been expressed as a 79-residue protein with the oxidizable Cys(56) replaced with Ser. The Nkx2.5 or Nkx2.5(C56S) homeodomain is 73% identical in sequence to and has the same NMR structure as the vnd (ventral nervous system defective)/NK-2 homeodomain of Drosophila when bound to the same specific DNA. The thermal unfolding of Nkx2.5(C56S) at pH 6.0 or 7.4 is a reversible, two-state process with unit cooperativity, as measured by differential scanning calorimetry (DSC) and far-UV circular dichroism. Adding 100 mM NaCl to Nkx2.5(C56S) at pH 7.4 increases T(m) from 44 to 54 +/- 0.2 degrees C and DeltaH from 34 to 45 +/- 2 kcal/mol (giving a DeltaC(p) of approximately 1.2 kcal K(-)(1) mol(-)(1) for homeodomain unfolding). DSC profiles of Nkx2.5 indicate fluctuating nativelike structures at <37 degrees C. Titrations of specific 18 bp DNA with Nkx2.5(C56S) in buffer at pH 7.4 with 100 mM NaCl yield binding constants of 2-6 x 10(8) M(-)(1) from 10 to 37 degrees C and a stoichiometry of 1:1 for homeodomain binding DNA, using isothermal titration calorimetry. The DNA binding reaction of Nkx2.5 is enthalpically controlled, and the temperature dependence of DeltaH gives a DeltaC(p) of -0.18 +/- 0.01 kcal K(-)(1) mol(-)(1). This corresponds to 648 +/- 36 A(2) of buried apolar surface upon Nkx2.5(C56S) binding duplex B-DNA. Thermodynamic parameters differ for Nkx2.5 and vnd/NK-2 homeodomains binding specific DNA. Unbound NK-2 is more flexible than Nkx2.5.  相似文献   

18.
Kozlov AG  Lohman TM 《Biochemistry》2002,41(39):11611-11627
The kinetic mechanism of transfer of the homotetrameric Escherichia coli SSB protein between ssDNA molecules was studied using stopped-flow experiments. Dissociation of SSB from the donor ssDNA was monitored after addition of a large excess of unlabeled acceptor ssDNA by using either SSB tryptophan fluorescence or the fluorescence of a ssDNA labeled with an extrinsic fluorophore [fluorescein (F) or Cy3]. The dominant pathway for SSB dissociation occurs by a "direct transfer" mechanism in which an intermediate composed of two DNA molecules bound to one SSB tetramer forms transiently prior to the release of the acceptor DNA. When an initial 1:1 SSB-ssDNA complex is formed with (dT)(70) in the fully wrapped (SSB)(65) mode so that all four SSB subunits are bound to (dT)(70), the formation of the ternary intermediate complex occurs slowly with an apparent bimolecular rate constant, k(2,app), ranging from 1.2 x 10(3) M(-1) s(-1) (0.2 M NaCl) to approximately 5.1 x 10(3) M(-1) s(-1) (0.4 M NaBr), and this rate limits the overall rate of the transfer reaction (pH 8.1, 25 degrees C). These rate constants are approximately 7 x 10(5)- and approximately 7 x 10(4)-fold lower, respectively, than those measured for binding of the same ssDNA to an unligated SSB tetramer to form a singly ligated complex. However, when an initial SSB-ssDNA complex is formed with (dT)(35) so that only two SSB subunits interact with the DNA in an (SSB)(35) complex, the formation of the ternary intermediate occurs much faster with a k(2,app) ranging from >6.3 x 10(7) M(-1) s(-1) (0.2 M NaCl) to 2.6 x 10(7) M(-1) s(-1) (0.4 M NaBr). For these experiments, the rate of dissociation of the donor ssDNA determines the overall rate of the transfer reaction. Hence, an SSB tetramer can be transferred from one ssDNA molecule to another without proceeding through a free protein intermediate, and the rate of transfer is determined by the availability of free DNA binding sites within the initial SSB-ssDNA donor complex. Such a mechanism may be used to recycle SSB tetramers between old and newly formed ssDNA regions during lagging strand DNA replication.  相似文献   

19.
T Wieprecht  M Beyermann  J Seelig 《Biochemistry》1999,38(32):10377-10387
Magainins are positively charged amphiphatic peptides which permeabilize cell membranes and display antimicrobial activity. They are usually thought to bind specifically to anionic lipids, and binding studies have been performed almost exclusively with negatively charged membranes. Here we demonstrate that binding of magainins to neutral membranes, a reaction which is difficult to assess with spectroscopic means, can be followed with high accuracy using isothermal titration calorimetry. The binding mechanism can be described by a surface partition equilibrium after correcting for electrostatic repulsion by means of the Gouy-Chapman theory. Unusual thermodynamic parameters are observed for the binding process. (i) The three magainin analogues that were investigated bind to neutral membranes with large exothermic reaction enthalpies DeltaH of -15 to -18 kcal/mol (at 30 degrees C). (ii) The reaction enthalpies increase with increasing temperature, leading to a large positive heat capacity DeltaC(p) of approximately 130 cal mol(-)(1) K(-)(1) (at 25 degrees C). (iii) The Gibbs free energies of binding DeltaG are between -6.4 and -8.6 kcal/mol, resulting in a large negative binding entropy DeltaS. The binding of magainin to small unilamellar vesicles is hence an enthalpy-driven reaction. The negative DeltaH and DeltaS and the large positive DeltaC(p) contradict the conventional understanding of the hydrophobic effect. CD experiments reveal that the membrane-bound fraction of magainin is approximately 80% helical at 8 degrees C, decreasing to approximately 60% at 45 degrees C. Since the random coil --> alpha-helix transition in aqueous solution is known to be an exothermic process, the same process occurring at the membrane surface is shown to account for up to 65% of the measured reaction enthalpy. In addition to membrane-facilitated helix formation, the second main driving force for membrane binding is the insertion of the nonpolar amino acid side chains into the lipid bilayer. It also contributes a negative DeltaH and follows the pattern for the nonclassical hydrophobic effect. Addition of cholesterol drastically reduces the extent of peptide binding and reveals an enthalpy-entropy compensation mechanism. Membrane permeability was measured with a dye assay and correlated with the extent of peptide binding. The level of dye efflux is linearly related to the amount of surface-bound peptide and can be traced back to a membrane perturbation effect.  相似文献   

20.
The mechanism of recognition of proteins and peptides by antibodies and the factors determining binding affinity and specificity are mediated by essentially the same features. However, additional effects of the usually unfolded and flexible solution structure of peptide ligands have to be considered. In an earlier study we designed and optimized six peptides (pepI to pepVI) mimicking the discontinuous binding site of interleukin-10 for the anti-interleukin-10 monoclonal antibody (mab) CB/RS/1. Three of them were selected for analysis of their solution conformation by circular dichroism measurements. The peptides differ in the content of alpha-helices and in the inducibility of helical secondary structures by trifluoroethanol. These properties, however, do not correlate with the binding affinity. PepVI, a 32-mer cyclic epitope mimic, has the highest affinity to mab CB/RS/1 identified to date. CD difference spectroscopy suggests an increase of the alpha-helix content of pepVI with complex formation. Binding of pepVI to mab CB/RS/1 is characterized by a large negative, favorable binding enthalpy and a smaller unfavorable loss of entropy (DeltaH degrees = -16.4 kcal x mol(-1), TDeltaS degrees = -6.9 kcal x mol(-1)) resulting in DeltaG degrees = -9.5 kcal x mol(-1) at 25 degrees C as determined by isothermal titration calorimetry. Binding of pepVI is enthalpically driven over the entire temperature range studied (10-35 degrees C). Complex formation is not accompanied by proton uptake or release. A negative heat capacity change DeltaC(p) of -0.354 kcal x mol(-1) x K(-1) was determined from the temperature dependence of DeltaH degrees. The selection of protein mimics with the observed thermodynamic properties is promoted by the applied identification and iterative optimization procedure.  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