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.
A model mechanism was developed for the binding of a rigid multisite protein with a randomly coiled multivalent ligand. Probabilities of the formation of chain loops between sites located at given distances at the protein were calculated by an extension of the concept of ring closure in coiled chain molecules. Expressions were derived for the dependence of overall equilibrium quantities, such as the binding constant between the protein and the ligand, on intrinsic parameters such as intrinsic binding constants, number of sites at the protein and their distances and on the chain length of the polymeric ligand. A pronounced chain length dependence of the overall binding constant was predicted even at chain lengths much longer than the size of the protein. Such a dependence was previously observed for the enzyme prolyl hydroxylase which acts on polymeric substrates like (ProProGly)n. This so far unexplained feature is quantitatively described by the model mechanism which is believed to be applicable to many other interactions of biological importance.  相似文献   

2.
We have mathematically analyzed ligand-induced monomerization and dimerization in a protein monomer-dimer equilibrium system, in which the monomer has one and the dimer two binding sites. These dimer sites have the same binding constants for the first ligand but may cooperatively interact when one of them is occupied by a ligand molecule. In this system, the apparent dimerization constant and the apparent molecular weight are functions of free ligand concentration, and depend on the intrinsic binding constants of the ligand molecule to the monomer and the dimer. The behavior of these functions is classified into 17 cases according to the values of the three intrinsic binding constants, and some calculated examples are shown graphically for selected parameters. The theory was also applied to D-amino acid oxidase [EC 1.4.3.3], a flavoprotein, and the pH dependence of the apparent dimerization constant and the apparent molecular weight in the presence of ligand, p-aminobenzoate, were studied theoretically using parameters obtained in our previous experiments (5).  相似文献   

3.
P J Hogg  P E Reilly  D J Winzor 《Biochemistry》1987,26(7):1867-1873
Theoretical consideration is given to the interaction of a bivalent ligand with particulate receptor sites, not only from the viewpoint of quantitatively describing the binding behavior but also from that of the kinetics of ligand release upon infinite dilution of a receptor-ligand mixture. In the latter regard, a general expression is derived that describes the time dependence of the amount of ligand bound as a function of two rate constants for the stepwise dissociation of cross-linked ligand-receptor complex and a thermodynamic parameter expressing the initial ratio of singly linked to doubly linked ligand-receptor complexes. An experimental study of the interaction between Sephadex and concanavalin A is then used to illustrate application of this recommended theoretical approach for characterizing the binding behavior and dissociation kinetics of a bivalent ligand for a system in which all ligand-receptor interactions may be described by a single intrinsic association constant. Published results on the interaction of phosphorylase b with butylagarose are also shown to comply with this simplest model of the bivalent ligand hypothesis; but those for the interaction between immunoglobulin G (IgG) dimers and Fc receptors require modification of the model by incorporation of different intrinsic association constants for the successive binding of receptor sites to the bivalent ligand. These results emphasize the need to consider ligand bivalency as a potential phenomenon in studies of interactions between protein ligands and particulate receptors and illustrate procedures by which the effects of ligand bivalency may be identified and characterized.  相似文献   

4.
In a previous paper we have reported on the structural perturbation of the erythrocyte membrane anion exchanger by a regular series of model amphiphiles, as shown by differential scanning calorimetry (Gruber, H.J. and Low, P.S., Biochim. Biophys. Acta, preceding article). Now the data are interpreted by a model in which the effects of amphiphile structure upon buffer-membrane partitioning are well separated from the dependence of the intrinsic potencies of membrane-bound amphiphiles upon amphiphile structure. The buffer-membrane partitioning situation was demonstrated to regularly change between extremes within a series of homologous amphiphiles, i.e. from a negligible to a predominant fraction of total amphiphile in the sample residing in the membrane. Based upon this demonstration a large number of reports on the chain length dependence of apparent potency could be reinterpreted in terms of chain length profiles of intrinsic potency, allowing for a comparison of the responses of various membrane proteins to homologous series of amphiphiles. The response patterns for chain length variation could be divided into three distinct classes: the intrinsic potency (i) can be independent of chain length over a very wide range of length, (ii) it can be rather independent up to a critical length where a sudden cut-off in potency occurs, or (iii) it can drop monotonically over a wide range of chain length. The intrinsic potency values of saturated fatty acids in destabilizing the anion exchanger were interpreted by very simple assumptions: only direct interactions between amphiphiles and target proteins and a simple amphiphile partition equilibrium between a pool of equivalent low affinity sites on the protein and the bulk lipid matrix. The observed monotonic decay of the intrinsic potency of saturated fatty acids with increasing chain length from C8 to C20 was translated into a constant increment of free energy by which each additional CH2 favors the transfer away from sites on the protein towards the bulk lipid matrix. Arguments were presented suggesting that the direct interaction between amphiphiles and target protein is completely nonspecific for alkyl chain length while the residual specificity for shorter over longer amphiphiles is due to the higher tendency of longer chains to preferentially bind in the bulk lipid matrix. Thus a completely new role of the lipid as a competitor, rather than a mediator, was postulated.  相似文献   

5.
Equilibrium binding of human alpha-thrombin to heparin was investigated at pH 7.4 as a function of thrombin and heparin concentrations, NaCl concentration, temperature, and heparin chain length with the extrinsic fluorescence probe, p-aminobenzamidine, or by quantitative affinity chromatography, in order to distinguish between sequence-specific and nonspecific electrostatic modes of binding. Analysis of binding data by a nonspecific binding model developed for protein-nucleic acid interactions, or by the discrete binding site model previously used to analyze the thrombin-heparin interaction, indicated that both models described the binding interaction equally well over the range of thrombin binding densities accessible to measurement. However, the strong dependence of the thrombin-heparin binding interaction on NaCl concentration, its minimal dependence on temperature, and the increase in apparent binding affinity with increasing heparin oligosaccharide chain length were best accounted for by a nonspecific electrostatic association of thrombin with 5 to 6 anionic residues contained in a 3-disaccharide binding site of heparin. This interaction was characterized by an intrinsic dissociation constant (KD,obs) of 6-10 microM at physiological ionic strength. Although the nonspecific binding model satisfactorily described the binding of thrombin to heparin chains ranging in size from 3 to approximately 13 disaccharides in terms of a single intrinsic KD,obs, deviations from this model were apparent with longer heparin chains (approximately 22 to approximately 35 disaccharides) from a progressive decrease in the intrinsic KD,obs of up to 4-fold. Sedimentation equilibrium analyses of thrombin-heparin complexes suggested a second weaker binding site on thrombin for heparin, which accounted for these deviations as well as the observed insolubility of thrombin-heparin complexes at high thrombin binding densities.  相似文献   

6.
A theory is formulated for an isoelectric focusing procedure which permits determination of intrinsic ligand-binding constants. The protein is first focused in the absence of ligand, after which ligand is added to the appropriate electrode compartment and then driven by the electric field into the focusing column where it complexes with the protein. The band of protein and its complexes moves to the constituent isoelectric point. An equation linearly relates the reciprocal of the overall distance moved to the reciprocal of the local concentration of ligand. The quotient of the intercept and slope gives the intrinsic binding constant. If the concentration of ligand in the electrode compartment is used in lieu of the local concentration, an apparent constant is obtained. Extrapolation of the apparent constant to infinite dilution of protein gives the intrinsic constant. For certain systems, conditions may be realized which give an apparent constant within 4% of the intrinsic constant.  相似文献   

7.
The kinetics of the interaction of bovine folate binding protein and folate at pH 7.4 and 5.0 were followed by measuring the changes of the intrinsic protein fluorescence intensity using the stopped-flow technique, which enables the study of reactions from the millisecond time-range. Our results immediately reject a simple one-step binding model, which requires a linear dependence of the observed rate constant on the concentration of the ligand. Thus, we are able to conclude that at pH 5.0 the interaction occurs in two steps and at pH 7.4 in three steps. Changes of fluorescence spectra at equilibrium were used to estimate the overall binding constants. Comparative studies on the binding of folate to human albumin are also reported.  相似文献   

8.
To investigate the functional sites on a protein and the prediction of binding sites (residues)in proteins, it is often required to identify the binding site residues at different distance threshold from protein three dimensional (3D)structures. For the study of a particular protein chain and its interaction with the ligand in complex form, researchers have to parse the output of different available tools or databases for finding binding-site residues. Here we have developed a tool for calculating amino acid contact distances in proteins at different distance threshold from the 3D-structure of the protein. For an input of protein 3D-structure, ContPro can quickly find all binding-site residues in the protein by calculating distances and also allows researchers to select the different distance threshold, protein chain and ligand of interest. Additionally, it can also parse the protein model (in case of multi model protein coordinate file)and the sequence of selected protein chain in Fasta format from the input 3D-structure. The developed tool will be useful for the identification and analysis of binding sites of proteins from 3D-structure at different distance thresholds. AVAILABILITY: IT CAN BE ACCESSED AT: http://procarb.org/contpro/  相似文献   

9.
A mathematical model is presented which describes the theoretical relationship between ligand concentration and physiological response for systems in which the response is dependent upon simultaneous occupancy of two receptor ligand-binding sites. The treatment considers both the possibility of intrinsic differences between the binding sites with regard to ligand affinity, as well as the possibility of mutually induced changes in affinity resulting from allosteric interactions. Unlike the Monod-Wyman-Changeux formulation for allosteric enzymes, the general model put forward here makes double occupancy an absolute requirement for enzymatic function. It is shown that such a model leads to the prediction of a curvilinear Hill plot from which one can obtain an explicit estimate of the degree of allosteric interaction between the two ligand binding sites as well as the Gibbs standard free energy change for the overall binding reaction. It is then shown that, in the specific instance of Na, K-ATPase-mediated K+ transport by the turkey erythrocyte, the configuration of the Hill curve describing the rate of ouabain-sensitive K+ transport as a function of external K+ concentration conforms closely to that predicted by the model described above. The results are of particular interest because they indicate a strongly cooperative interaction between the two K+ binding sites on the transport protein such that occupancy of one site results in an enhancement of the affinity of the other site for K+ by a minimum of 15- to 20-fold. Finally, we consider in detail a model of the Monod-Wyman- Changeux type in which, by contrast, both singly and doubly occupied forms of the enzyme are assumed to be catalytically active, and which we analogously extend to allow for the possibility of asymmetry between the two ligand binding sites. Although it is shown that the two models can not be differentiated from each other in the present experimental system, they yield virtually identical estimates for the degree of positive cooperativity between the two K+ binding sites.  相似文献   

10.
The theory for the salt dependence of the free energy, entropy, and enthalpy of a polyelectrolyte in the PB (PB) model is extended to treat the nonspecific salt dependence of polyelectrolyte–ligand binding reactions. The salt dependence of the binding constant (K) is given by the difference in osmotic pressure terms between the react ants and the products. For simple 1-1 salts it is shown that this treatment is equivalent to the general preferential interaction model for the salt dependence of binding [C. Anderson and M. Record (1993) Journal of Physical Chemistry, Vol. 97, pp. 7116–7126]. The salt dependence, entropy, and enthalpy are compared for the PB model and one specific form of the preferential interaction coefficient model that uses counterion condensation/limiting law (LL) behavior. The PB and LL models are applied to three ligand–polyelectrolyte systems with the same net ligand charge: a model sphere–cylinder binding reaction, a drug–DNA binding reaction, and a protein–DNA binding reaction. For the small ligands both the PB and limiting law models give (ln K vs. In [salt]) slopes close in magnitude to the net ligand charge. However, the enthalpy/entropy breakdown of the salt dependence is quite different. In the PB model there are considerable contributions from electrostatic enthalpy and dielectric (water reorientation) entropy, compared to the predominant ion cratic (release) entropy in the limiting law model. The relative contributions of these three terms in the PB model depends on the ligand: for the protein, ion release entropy is the smallest contribution to the salt dependence of binding. The effect of three approximations made in the LL model is examined: These approximations are (1) the ligand behaves ideally, (2) the preferential interaction coefficient of the polyelectrolyte is unchanged upon ligand binding, and (3) the polyelectrolyte preferential interaction coefficient is given by the limiting law/counterion-condensation value. Analysis of the PB model shows that assumptions 2 and 3 break down at finite salt concentrations. For the small ligands the effects on the slope cancel, however, giving net slopes that are similar in the PB and LL models, but with a different entropy/enthalpy breakdown. For the protein ligand the errors from assumptions 2 and 3 in the LL model do not cancel. In addition, the ligand no longer behaves ideally due to its complex structure and charge distribution. Thus for the protein the slope is no longer related simply to the net ligand charge, and the PB model gives a much larger slope than the LL model. Additionally, in the PB model most of the salt dependence of the protein binding comes from the change in ligand activity, i.e. from nonspecific anion effects, in contrast to the small ligand case. While the absolute binding is sensitive to polyelectrolyte length, little length effect is seen on the salt dependence for the small ligands at 0.1M salt, and for lengths > 60 Å. Almost no DNA length dependenceis seen in the salt dependence of the protein binding, since this is determined primarily by the protein, not the DNA. © 1995 John Wiley & Sons, Inc.  相似文献   

11.
Explicit expressions are derived which describe the binding of a univalent ligand to equivalent and independent sites on each state of an acceptor undergoing indefinite self-association that is governed by an isodesmic equilibrium constant KI. From considerations of systems in which the same site-binding constant kA applies to all acceptor-ligand interactions, the general forms of binding curves and Scatchard plots are deduced for situations in which binding sites are either created or lost at each monomer-monomer interface. Greater generality is then introduced into the model by allowing ligand interactions with polymeric acceptor states to be governed by a site-binding constant kp that differs in magnitude from that for monomeric acceptor kA. Finally, experimental results with the glutamate dehydrogenase-GTP and lysozyme-saccharide systems are used to illustrate ways in which the present quantitative expressions may be applied to the characterization of inteactions between a ligand and an indefinitely self-associating acceptor.  相似文献   

12.
M Takahashi  B Blazy  A Baudras 《Biochemistry》1980,19(22):5124-5130
The binding of adenosine cyclic 3',5'-monophosphate (cAMP) and guanosine cyclic 3',5'-monophosphate (cGMP) to the adenosine cyclic 3',5'-monophosphate receptor protein (CRP) from Escherichia coli was investigated by equilibrium dialysis at pH 8.0 and 20 degrees C at different ionic strengths (0.05--0.60 M). Both cAMP and cGMP bind to CRP with a negative cooperativity that is progressively changed to positive as the ionic strength is increased. The binding data were analyzed with an interactive model for two identical sites and site/site interactions with the interaction free energy--RT ln alpha, and the intrinsic binding constant K and cooperativity parameter alpha were computed. Double-label experiments showed that cGMP is strictly competitive with cAMP, and its binding parameters K and alpha are not very different from that for cAMP. Since two binding sites exist for each of the cyclic nucleotides in dimeric CRP and no change in the quaternary structure of the protein is observed on binding the ligands, it is proposed that the cooperativity originates in ligand/ligand interactions. When bound to double-stranded deoxyribonucleic acid (dsDNA), CRP binds cAMP more efficiently, and the cooperativity is positive even in conditions of low ionic strength where it is negative for the free protein. By contrast, cGMP binding properties remained unperturbed in dsDNA-bound CRP. Neither the intrinsic binding constant K nor the cooperativity parameter alpha was found to be very sensitive to changes of pH between 6.0 and 8.0 at 0.2 M ionic strength and 20 degrees C. For these conditions, the intrinsic free energy and entropy of binding of cAMP are delta H degree = -1.7 kcal . mol-1 and delta S degree = 15.6 eu, respectively.  相似文献   

13.
The location of certain amino acid sequences like repeats along the polypeptide chain is very important in the context of forming the overall shape of the protein molecule which in fact determines its function. In gram‐positive bacteria, fibronectin‐binding protein (FnBP) is one such repeat containing protein, and it is a cell wall‐attached protein responsible for various acute infections in human. Several studies on sequence, structure, and function of fibronectin‐binding regions of FnBPs were reported; however, no detailed study was carried out on the full‐length protein sequence. In the present study, we have made a thorough sequence and structure analysis on FnBP_A of Staphylococcus aureus and explored the presence of dual ligand‐binding ability of fibrinogen (fg)‐binding region and its molecular recognition processes. Multiple sequence alignment and protein‐protein docking analysis reveal the regions which are likely involved in dual ligand binding. Further analysis of docking of FnBP_A fg‐binding region and fn N‐terminal modules suggests that if the latter binds to the fg‐binding region of FnBP_A, it would inhibit the subsequent binding of fg because of steric hindrance. The sequence analysis further suggests that the abundance of disorder promoting residue glutamic acid and dual personality (both order/disorder promoting) residue threonine in tandem repeats of FnBP_A and B proteins possibly would help the molecule to undergo a conformational change while binding with fn by β‐zipper mechanism. The segment‐based power spectral analysis was carried out which helps to understand the distribution of hydrophobic residues along the sequence particularly in intrinsic disordered tandem repeats. The results presented here will help to understand the role of internal repeats and intrinsic disorder in the molecular recognition process of a pathogenic cell surface protein.  相似文献   

14.
A theory is formulated to provide guidelines for the quantitative interpretation of steady-state counterion electrophoretic patterns (T.-H. Ueng and F. Bronne, Arch. Biochem. Biophys. 197 (1979) 205) in terms of intrinsic ligand-binding constant and number of binding sites on the protein molecule. Briefly, the prescribed procedure calls for extrapolation of the steady-state binding constant to infinite dilution of protein to obtain a quantity which is the product of a readily evaluated kinetic factor and the intrinsic binding constant. On the other hand, extrapolation of the steady-state number of binding sites to infinite dilution can probably be dispensed with if determined at a reasonably low protein concentration.  相似文献   

15.
When the model protein bovine serum albumin (BSA) was dissolved in a concentrated aqueous solution of the multifunctional ligand L-malic acid, the solution was lyophilized, and the solid residue thoroughly washed with tetrahydrofuran to extract malic acid, then the resultant ("imprinted") protein was capable of binding 26.4 +/-0.9 mol equivalents of the ligand in anhydrous ethyl acetate. The nonimprinted BSA (i.e., that prepared in the same manner apart from the absence of malic acid) bound less then one-tenth of that amount under identical conditions. Furthermore, both imprinted and nonimprinted BSA exhibited little binding of L-malic acid in water. The imprinted BSA retained its "memory" for the ligand in ethyl acetate even after a prolonged incubation under vacuum; dissolution in water, however, eliminated the imprinted protein's binding capacity. The BSA imprinted with L-malic acid displayed affinity for this ligand not only in ethyl acetate but also in many other anhydrous solvents. It was found that the higher the solvent's propensity to form hydrogen bonds, the lower the protein-ligand binding in it, thus pointing to hydrogen bonds as the driving force of this binding. Studies with completely or partially cleaved BSA, with other globular proteins, glutathione, and poly(L-aspartic acid) revealed that the critical requirement for the imprintability is the presence of a sufficiently long polymeric chain. Moreover, many hydrogen-bond-forming macromolecules other than proteins, such as dextrans and their derivatives, partially hydrolyzed starch, and poly(methacrylic acid), also could be imprinted for subsequent binding in ethyl acetate. The mechanism of imprinting and binding inferred from these experiments involves a multipoint hydrogen bonding in water of each ligand molecule with two or more sites on the polymeric chain, thereby folding a segment of the latter into a cavity around the ligand; following lyophilization and extraction of the ligand, the cavities remain in organic solvents (but not in water) and give rise to ligand binding. This conclusion is supported by the results of binding of numerous malic acid analogs and related ligands to BSA imprinted with L-malic acid. Finally, BSA imprinted with malic acid was used as a selective adsorbent for a chromatographic separation of an equimolar mixture of maleic and acrylic acids in ethyl acetate.  相似文献   

16.
We describe the synthesis of N-trifluoroacetyl chito-oligosaccharides and their use as ligands to probe the binding sites of wheat-germ agglutinin, a lectin specific for N-acetylglucosamine. The binding is monitored using intrinsic protein fluorescence, which is due to tryptophan side-chains. We present arguments purporting to show the presence of a fluorophore close to each of the four sites. The binding of chito-oligosaccharides to wheat-germ agglutinin is complex and can only be approximately described by an independent and equivalent sites model. This model applies when the ligand concentration range is restricted to higher values. The possible role of ligand-mediated protein aggregation and of site inequivalence is discussed. We find that the affinity of trifluoroacetylated chito-oligosaccharides for wheat-germ agglutinin is higher than that of the N-acetylated parent compounds, the difference increasing with chain length. Our results are in agreement with a model of the binding site previously proposed by Clegg et al. (Biochemistry 22 (1983) 4797-4804).  相似文献   

17.
18.
Consideration is given to the interactions of a ligand with self-associating acceptor systems for which preferential binding is an ambiguous term in that ligand-mediated self-association does not necessarily imply a greater binding constant for polymeric acceptor--even in instances where binding sites are preserved in the self-association process. This dilemma is shown to arise in situations involving the binding of ligand to monomeric and polymeric forms of an acceptor that also coexist in equilibrium with inactive isomeric states. For example, the ten-fold increase in the measured dimerization constant for prothrombin Fragment 1 in the presence of a saturating concentration of Ca2+ ion may well reflect the existence of a 12% greater binding constant for the interaction of metal ion with dimeric acceptor. However, that result, as well as the detailed form of the sigmoidal binding curve, are also reasonably described by another extreme model in which the monomeric and dimeric forms of the acceptor possess equal affinities for Ca2+ ion. Likewise, the fact that the same experimental dimerization constant applies to prothrombin and its Ca(2+)-saturated complex does not preclude the possibility that the active form of dimeric zymogen exhibits a 12% greater affinity for metal ion. Numerical simulations have established that characterization of the dimerization behaviour as a function of free ligand concentration should allow greater discrimination between such models of the interplay between calcium binding and self-association of prothrombin and Fragment 1. Finally, by illustrating the likelihood that the disparity in self-association behaviour of prothrombin and Fragment 1 merely reflects minor differences in the relative magnitudes of isomerization constants and/or binding constants for monomeric and dimeric states of the two acceptors, the present investigation serves to allay concern about the validity of employing the proteolytic fragment as a model of the intact zymogen.  相似文献   

19.
The crystal structure of the kringle 2 domain of tissue plasminogen activator was determined and refined at a resolution of 2.43 A. The overall fold of the molecule is similar to that of prothrombin kringle 1 and plasminogen kringle 4; however, there are differences in the lysine binding pocket, and two looping regions, which include insertions in kringle 2, take on very different conformations. Based on a comparison of the overall structural homology between kringle 2 and kringle 4, a new sequence alignment for kringle domains is proposed that results in a division of kringle domains into two groups, consistent with their proposed evolutionary relation. The crystal structure shows a strong interaction between a lysine residue of one molecule and the lysine/fibrin binding pocket of a noncrystallographically related neighbor. This interaction represents a good model of a bound protein ligand and is the first such ligand that has been observed in a kringle binding pocket. The structure shows an intricate network of interactions both among the binding pocket residues and between binding pocket residues and the lysine ligand. A lysine side chain is identified as the positively charged group positioned to interact with the carboxylate of lysine and lysine analogue ligands. In addition, a chloride ion is located in the kringle-kringle interface and contributes to the observed interaction between kringle molecules.  相似文献   

20.
We have examined the salt and temperature dependences of the equilibrium binding of the Escherichia coli single strand binding (SSB) tetramer to a series of oligodeoxythymidylates, dT(pT)N-1, with N = 16, 28, 35, 56 and 70. Absolute binding isotherms were obtained, based on the quenching of the intrinsic protein fluorescence upon formation of the complexes. The shorter oligonucleotides, with N = 16, 28 and 35, bind to multiple sites on the SSB tetramer and negative co-operativity is observed among these binding sites. We have quantitatively analyzed these isotherms, using a statistical thermodynamic ("square") model to obtain the intrinsic binding constant KN, and the negative co-operativity constant, sigma N. For all oligonucleotides, we find that KN decreases significantly with increasing concentration of monovalent salt, indicating a large electrostatic component to the free energy of the interaction (e.g. delta log KN/delta log [NaBr] = -2.7, -4.6 and -7.1 for N = 16, 35 and 70, respectively), with contributions from both cations and anions. For oligonucleotides that span two or more subunits, there is a significant unfavorable contribution to the binding free energy for each intersubunit crossing, with an accompanying uptake of anions. Therefore, the extent of anion uptake increases as the number of intersubunit crossings increase. There is a strong temperature dependence for the intrinsic binding of dT(pT)15, such that delta Ho = -26(+/- 3) kcal/mol dT(pT)15. Negative co-operativity exists under all solution conditions tested, i.e. sigma N less than 1, and this is independent of anion concentration and type. However, the negative co-operativity constant does decrease with decreasing concentration of cation. The dependence of sigma 16 on Na+ concentration indicates that an average of one sodium ion is taken up as a result of the negative co-operativity between two dT(pT)15 binding sites. These data and the lack of a temperature dependence for sigma 16 suggest that the molecular basis for the negative co-operativity is predominantly electrostatic and may be due to the repulsion of regions of single-stranded DNA that are required to bind in close proximity on an individual SSB tetramer.  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