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.
J B Chaires 《Biopolymers》1985,24(2):403-419
Fluorescence and absorbance methods were used to study the interaction of daunomycin with calf-thymus DNA over a wide range of temperatures and NaCl concentrations. van't Hoff analysis provided estimates for the enthalpy of the binding reaction over the NaCl range of 0.05–1.0 M. Daunomycin binding is exothermic over this entire range, and the favorable binding free energy arises primarily from the large, negative enthalpy. Both the enthalpy change and entropy change are strong functions of ionic strength. Possible molecular contributions to the enthalpy and entropy are discussed, leading to the tentative conclusion that hydrogen-bonding interactions at the interacalation site are the primary contributors to the observed thermodynamic parameters. The dependence of the enthalpy on the ionic strength is well beyond the predictions of current polyelectrolyte theory and cannot be fully accounted for. The enthalpy and entropy changes observed compensate one another to produce relatively small free-energy changes over the range of solution conditions studied.  相似文献   

2.
T. G. Dewey 《Biopolymers》1990,29(14):1793-1799
A ligand binding model of counterion association in finite length polyelectrolytes is presented. This model introduces counterion condensation features into a binding formalism. It agrees well with the predictions of other finite length models and is consistent with experimental data on helix–coil melting transitions for short nucleic acid oligomers. This model uses a discrete charge distribution for the polyelectrolyte. An expression for the electrostatic self-energy of finite length polyelectrolytes is derived using the Euler–Maclaurin sum formula. This sum is shown to be accurate over a wide range of salt concentrations. This electrostatic term is used in an energy minimization analysis. The energy minimization is solved analytically using a Lagrange inversion formula. This general procedure gives a rapidly convergent series and requires no assumptions with regard to “limiting law” behavior. However, when used in the Manning minimization formalism [(1977) Biophysical Chemistry, 24 , 2086], the volume of the condensed phase becomes unrealistically large at low ionic strength. The ligand binding model does not have a condensed phase volume as a parameter. It provides a single expression that agrees both with Manning's theory and with the theory of Ramanathan and Woodbury [(1982) Journal of Chemical Physics 77 , 4133] under the respective conditions of these theories.  相似文献   

3.
4.
The protein turkey-heart cytochrome c is used as a model protein to study charge effects in sedimentation equilibrium experiments in three-component solutions. Data are given for the dependence of the apparent M (1–υ ρ) on ρ in solutions of KCl, RbCl, CsCl, and triethylamine hydrochloride. The results show the Donnan effect to have a significant influence on the apparent molecular weight, found by extrapolation of the data to a solution density of one. The apparent molecular weights are for protein at infinite dilution. A theoretical treatment is presented where the magnitude of this effect can be predicted accurately from the formal net charge of the protein as computed from the amino acid composition. The results are shown to be important in computing the preferential hydration of the protein in concentrated salt solutions. For such systems the Donnan effect should be subtracted from the total interaction coefficient for multicomponent system in order to obtain the preferential hydration.  相似文献   

5.
The thermodynamic parameters of the interaction of agonists and antagonists with heart and brain muscarinic receptors were determined. The binding of quinuclidinyl [3H]benzilate and the inhibition of quinuclidinyl benzilate (QNB) binding by agonists and antagonists were examined at temperatures between 2 degrees C and 27 degrees C. The density of specific binding sites and the relative proportions of high- and low-affinity binding components of drugs were unaffected by the temperature changes. The binding of atropine was entropy driven in brain and heart membranes. In contrast, net values of these thermodynamic parameters for QNB binding and for the high-affinity binding component of pirenzepine to brain membranes were decreased with the enhancement of the temperature. The low-affinity binding component of the agonists carbachol, oxotremorine and pilocarpine was enthalpy driven. Their high-affinity binding component was entropy driven at 2 degrees C and became enthalpy driven when the incubation temperature was increased. The guanine nucleotide Gpp[NH]p partly prevented the temperature-dependent decrease of net entropy and enthalpy values. Considering that the net changes of thermodynamic parameters are relevant of the interactions between the ligand, the receptor protein and the adjoining membranous molecules, a three-state conformational model is proposed for the muscarinic receptor protein. The receptor selectivity is reappreciated owing to these three states of the receptor protein and the different components of the muscarinic receptor complexes.  相似文献   

6.
The temperature dependence of preferential solvent interactions with ribonuclease A in aqueous solutions of 30% sorbitol, 0.6 M MgCl2, and 0.6 M MgSO4 at low pH (1.5 and 2.0) and high pH (5.5) has been investigated. This protein was stabilized by all three co-solvents, more so at low pH than high pH (expect 0.6 M MgCl2 at pH 5.5). The preferential hydration of protein in all three co-solvents was high at temperatures below 30 degrees C and decreased with a further increase in temperature (for 0.6 M MgCl2 at pH 5.5, this was not significant), indicating a greater thermodynamic instability at low temperature than at high temperature. The preferential hydration of denatured protein (low pH, high temperature) was always greater than that of native protein (high pH, high temperature). In 30% sorbitol, the interaction passed to preferential binding at 45% for native ribonuclease A and at 55 degrees C for the denatured protein. Availability of the temperature dependence of the variation with sorbitol concentration of the chemical potential of the protein, (delta mu(2)/delta m3)T,p,m2, permitted calculation of the corresponding enthalpy and entropy parameters. Combination with available data on sorbitol concentration dependence of this interaction parameter gave (approximate) values of the transfer enthalpy, delta H2,tr, and transfer entropy delta S2,tr. Transfer of ribonuclease A from water into 30% sorbitol is characterized by positive values of the transfer free energy, transfer enthalpy, transfer entropy, and transfer heat capacity. On denaturation, the transfer enthalpy becomes more positive. This increment, however, is small relative to both the enthalpy of unfolding in water and to the transfer enthalpy of the native protein from water a 30% sorbitol solution.  相似文献   

7.
The counterion density and the condensation region around DNA have been examined as functions of both ion size and added-salt concentration using Metropolis Monte Carlo (MC) and Poisson–Boltzmann (PB) methods. Two different definitions of the “bound” and “free” components of the electrolyte ion atmosphere were used to compare these approaches. First, calculation of the ion density in different spatial regions around the polyelectrolyte molecule indicates, in agreement with previous work, that the PB equation does not predict an invariance of the surface concentration of counterions as electrolyte is added to the system. Further, the PB equation underestimates the counterion concentration at the DNA surface, compared to the MC results, the difference being greatest in the grooves, where ionic concentrations are highest. If counterions within a fixed radius of the helical axis are considered to be bound, then the fraction of polyelectrolyte charge neutralized by counterions would be predicted to increase as the bulk electrolyte concentration increases. A second categorization—one in which monovalent cations in regions where the average electrostatic potential is ledd than ?kT are considered to be bound—provides an informative basis for comparison of MC and PB with each other and with counterion-condensation theory. By this criterion, PB calculations on the B from of DNA indicate that the amount of bound counterion charge per phosphate group is about .67 and is independent of salt concentration. A particularly provocative observatiob is that when this binding criterion is used, MC calculations quantitatively reproduce the bound fraction predicated by counterion-condensation theory for all-atom models of B-DNA and A-DNA as well as for charged cylindera of varying lineat charge densities. For example, for B-DNA and A-DNA, the fractions of phosphate groups neutralized by 2 Å hard sphere counterions are 0.768 and .817, respectively. For theoretical studies, the rediys enclosing the region in which the electrostatic potential is calculated studies, the radius enclosing the region in which the electrostatic potential is calculated to be less than ?kT is advocated s a more suitable binding or condensation radius that enclosing the fraction of counterions given by (1 – ξ?1). A comparsion of radii calculated using both of these definitions is presented. © 1994 John Wiley & Sons, Inc.  相似文献   

8.
Kinetics of thermal inactivation of acrylodan-labeled cAMP dependent protein kinase catalytic subunit, its binary complexes with ATP and peptide inhibitor PKI[5–24], respectively, and the ternary complex involving both of these ligands were studied at different temperatures (5–50 °C). The thermodynamic parameters ΔH and ΔS for ligand binding equilibria as well as for the allosteric interaction between the binding sites of these ligands were obtained by using the Van’t Hoff analysis. The results indicated that more inter- and intra-molecular non-covalent bonds were involved in ATP binding with the protein when compared to the peptide binding. Similarly, nucleotide and peptide binding steps were accompanied with different entropy effects, while almost no entropy change accompanied PKI[5–24] binding, suggesting that the protein flexibility was not affected in this case. Differently from the binary complex formation the ternary complex formation was accompanied by a significant entropy change and with intensive formation of new non-covalent interactions (ΔH). At the same time both ligand binding steps as well as the allosteric interaction between ligand binding sites could be described by a common entropy–enthalpy compensation plot, pointing to a similar mechanism of these phenomena. It was concluded that numerous weak interactions govern the allostery of cAMP dependent protein kinase catalytic subunit.  相似文献   

9.
Grand canonical Monte Carlo (GCMC) simulations are reported for aqueous solutions containing excess univalent salt (activities a +/- = 1.76-12.3 mM) and one of the following species: an octacationic rod-like ligand, L8+; a B-DNA oligomer with N phosphate charges (8 < or = N < or = 100); or a complex resulting from the binding of L8+ at the center of an N-mer (24 < or = N < or = 250). Simplified models of these multiply charged species are used in the GCMC simulations to predict the fundamental coulombic contributions to the following experimentally relevant properties: 1) the axial distance over which ligand binding affects local counterion concentrations at the surface of the N-mer; 2) the dependence on N of GCMC preferential interaction coefficients, gamma 32MC identical to delta C3/delta C2l a +/-, T, where C3 and C2 are, respectively, the molar concentrations of salt and the multiply charged species (ligand, N-mer or complex); and 3) the dependence on N of SaKobs identical to d in Kobs/d in a +/- = delta (magnitude of ZJ + 2 gamma 32J), where Kobs is the equilibrium concentration quotient for the binding of L8+ to the center of an N-mer and delta denotes the stoichiometric combination of terms, each of which pertains to a reactant or product J having magnitude of ZJ charges. The participation of electrolyte ions in the ligand binding interaction is quantified by the magnitude of SaKobs, which reflects the net (stoichiometrically weighted) difference in the extent of thermodynamic binding of salt ions to the products and reactants. Results obtained here from GCMC simulations yield a picture of the salient molecular consequences of binding a cationic ligand, as well as thermodynamic predictions whose applicability can be tested experimentally. Formation of the central complex is predicted to cause a dramatic reduction in the surface counterion (e.g., Na+) concentration over a region including but extending well beyond the location of the ligand binding site. For binding a cationic ligand, SaKobs is predicted to be negative, indicating net electrolyte ion release in the binding process. At small enough N, -SaKobs is predicted to decrease strongly toward zero with decreasing N. At intermediate N, -SaKobs appears to exceed its limiting value as N-->infinity.  相似文献   

10.
A theory of polyelectrolyte effects on site-binding equilibria is generalized to multivalent ligands, multivalent supporting salt, intercalation, and multiple-site exclusion. The theory, which contains no adjustable parameters, except the number of sites excluded by a bound ligand, gives the dependence of the equilibrium constant on the binding fraction and the salt concentration. The theory is compared with prior experimental data for the dissociation of poly(acrylic acid), the binding of magnesium to polyphosphate, and the binding of ethidium and actionomycin D to DNA. The theory predicts the binding fraction dependence of the dissociation constant of poly(acrylic acid) well. The theory predicts the binding fraction dependence of the association constant of the binding of Mg2+ to polyphosphate well, if either one or two phosphates are bound by a magnesium ion. We conclude that polyelectrolyte effects on drug-DNA equilibria must be substantial. It follows that an incorrect estimate of the number of sites excluded by a bound drug molecule (because of its size or some other nonpolyelectrolyte effect) can be obtained from binding data if polyelectrolyte effects are ignored. The estimate is also within the context of, and subject to the validity of, the model used to describe the nonpolyelectrolyte contribution to binding. Our results suggest that, subject to these conditions, the anticooperativity of the binding of ethidium to DNA might be explained solely in terms of polyelectrolyte effects, and without reference to multiple-site exclusion, if sequence-specificity effects can be safely ignored. Our results also suggest that as few as two base pairs might be excluded by an actinomycin molecule. The theory gives fairly good agreement for the salt-concentration dependence of the association constant of all of the systems studied, including the complex of the neutral drug actinomycin with DNA.  相似文献   

11.
Closed-form expression have been derived for the polyelectrolyte contribution to the colligative properties of solutions containing rodlike polyions in the presence of excess added salt. The derivations are based on: the conventional Poisson-Boltzmann equation for cylindrical symmetry; the thermodynamics of the cell model developed by Marcus [J. Chem. Phys. 23 , 1057–1068 (1955)]; and an equation derived from the cylindrical Poisson-Boltzmann cell model by Anderson and Record [Biophys. Chem. 11 , 353–360 (1980)]. Subject to the inherent limitations of the Poisson-Boltzmann approximation [Fixman (1979) J. Chem. Phys. 70 , 4995–5005], the resulting expressions are nevertheless applicable outside the “limit of infinite dilution.” They conform over a range of salt concentrations to the limiting laws deduced by Manning from the hypothesis of counterion condensation [J. Chem. Phys. 51 , 924–933 (1969)]. This hypothesis is found to be compatible with the Poisson-Boltzmann cell model but is not required in the derivation of the thermodynamic coefficients presented here. It is demonstrated that the magnitude of the polyion axial charge density plays a critical role in determining the low-salt limiting forms of the colligative properties obtained from the Poisson-Boltzmann equation, in close analogy with Manning's model.  相似文献   

12.
Sharrow SD  Novotny MV  Stone MJ 《Biochemistry》2003,42(20):6302-6309
The mouse pheromone 2-sec-butyl-4,5-dihydrothiazole (SBT) binds to an occluded, nonpolar cavity in the mouse major urinary protein-I (MUP-I). The thermodynamics of this interaction have been characterized using isothermal titration calorimetry (ITC). MUP-I-SBT binding is accompanied by a large favorable enthalpy change (DeltaH = -11.2 kcal/mol at 25 degrees C), an unfavorable entropy change (-TDeltaS = 2.8 kcal/mol at 25 degrees C), and a negative heat capacity change [DeltaC(p)() = -165 cal/(mol K)]. Thermodynamic analysis of binding between MUP-I and several 2-alkyl-4,5-dihydrothiazole ligands indicated that the alkyl chain contributes more favorably to the enthalpy and less favorably to the entropy of binding than would be expected on the basis of the hydrophobic desolvation of short-chain alcohols. However, solvent transfer experiments indicated that desolvation of SBT is accompanied by a net unfavorable change in enthalpy (DeltaH = +1.0 kcal/mol) and favorable change in entropy (-TDeltaS = -1.8 kcal/mol). These results are discussed in terms of the possible physical origins of the binding thermodynamics, including (1) hydrophobic desolvation of both the protein and the ligand, (2) formation of a buried water-mediated hydrogen bond network between the protein and ligand, (3) formation of strong van der Waals interactions, and (4) changes in the structure, dynamics, and/or hydration of the protein upon binding.  相似文献   

13.
Numerical calculations, using Poisson-Boltzmann (PB) and counterion condensation (CC) polyelectrolyte theories, of the electrostatic free energy difference, DeltaGel, between single-stranded (coil) and double-helical DNA have been performed for solutions of NaDNA + NaCl with and without added MgCl2. Calculations have been made for conditions relevant to systems where experimental values of helix coil transition temperature (Tm) and other thermodynamic quantities have been measured. Comparison with experimental data has been possible by invoking values of Tm for solutions containing NaCl salt only. Resulting theoretical values of enthalpy, entropy, and heat capacity (for NaCl salt-containing solutions) and of Tm as a function of NaCl concentration in NaCl + MgCl2 solutions have thus been obtained. Qualitative and, to a large extent, quantitative reproduction of the experimental Tm, DeltaHm, DeltaSm, and DeltaCp values have been found from the results of polyelectrolyte theories. However, the quantitative resemblance of experimental data is considerably better for PB theory as compared to the CC model. Furthermore, some rather implausible qualitative conclusions are obtained within the CC results for DNA melting in NaCl + MgCl2 solutions. Our results argue in favor of the Poisson-Boltzmann theory, as compared to the counterion condensation theory.  相似文献   

14.
Isoforms A (LGB‐A) and B (LGB‐B) of bovine lactoglobulin, the milk protein, differ in positions 64 (D?G) and 118 (V?A). Interactions of LGB‐A and LGB‐B with sodium dodecyl sulfate (SDS), dodecyltrimethylammonium chloride (DTAC) and lauric acid (LA), 12‐carbon ligands possessing differently charged polar groups, were investigated using isothermal titration calorimetry and X‐ray crystallography, to study the proton linkage phenomenon and to distinguish between effects related to different isoforms and different ligand properties. The determined values of ΔS and ΔH revealed that for all ligands, binding is entropically driven. The contribution from enthalpy change is lower and shows strong dependence on type of buffer that indicates proton release from the protein varying with protein isoform and ligand type and involvement of LA and Asp64 (in isoform A) in this process. The ligand affinities for both isoforms were arranged in the same order, DTAC < LA < SDS, and were systematically lower for variant B. The entropy change of the complexation process was always higher for isoform A, but these values were compensated by changes in enthalpy, resulting in almost identical ΔG for complexes of both isoforms. The determined crystal structures showed that substitution in positions 64 and 118 did not influence the overall structure of LGB complexes. The chemical character of the ligand polar group did not affect the position of its aliphatic chain in protein β‐barrel, indicating a major role of hydrophobic interactions in ligand binding that prevailed even with the repulsion between positively charged DTAC and lysine residues located at binding site entrance. Copyright © 2013 John Wiley & Sons, Ltd.  相似文献   

15.
T Arakawa  R Bhat  S N Timasheff 《Biochemistry》1990,29(7):1914-1923
The correlation between protein solubility and the preferential interactions of proteins with solvent components was critically examined with aqueous MgCl2 as the solvent system. Preferential interaction and solubility measurements with three proteins, beta-lactoglobulin, bovine serum albumin, and lysozyme, resulted in similar patterns of interaction. At acid pH (pH 2-3) and lower salt concentrations (less than 2 M), the proteins were preferentially hydrated, while at higher salt concentrations, the interaction was either that of preferential salt binding or low salt exclusion. At pH 4.5-5, all three proteins exhibited either very low preferential hydration or preferential binding of MgCl2. These results were analyzed in terms of the balance between salt binding and salt exclusion attributed to the increase in the surface tension of water by salts, which is invariant with conditions. It was shown that the increase in salt binding at high salt concentration is a reflection of mass action, while its decrease at acid pH is due to the electrostatic repulsion between Mg2+ ions and the high net positive charge on the protein. The preferential interaction pattern was paralleled by the variation of protein solubility with solvent conditions. Calculation of the transfer free energies from water to the salt solutions for proteins in solution and in the precipitate showed dependencies on salt concentration. This indicates that the nature of interactions between proteins and solvent components is the same in solution and in the solid state, which implies no change in protein structure during precipitation. Analysis of the transfer free energies and preferential interaction parameter in terms of the salting-in, salting-out, and weak ion binding contributions has led to the conclusions that, when the weak ion binding contribution is small, the predominant protein-salt interaction must be that of preferential salt exclusion most probably caused by the increase of the surface tension of water by addition of the salt. A necessary consequence of this is salting-out of the protein, if the protein structure is to remain unaltered.  相似文献   

16.
17.
Under the free energy landscape theory, both the protein-folding and protein–ligand binding processes are driven by the decrease in total Gibbs free energy of the protein-solvent or protein–ligand-solvent system, which involves the non-complementary changes between the entropy and enthalpy, ultimately leading to a global free energy minimization of these thermodynamic systems (Ji & Liu, 2011; Liu et al., 2012; Yang, Ji & Liu, 2012). In the case of protein folding, the lowering of the system free energy coupled with the gradual reduction in conformational degree of freedom of the folding intermediates determines that the shape of the free energy landscape for protein folding must be funnel-like (Dill & Chan, 1997), rather than non-funneled shapes (Ben-Naim, 2012). In the funnel-like free energy landscape, protein folding can be viewed as going down the hill via multiple parallel routes from a vast majority of individual non-native states on surface outside the funnel to the native states located around the bottom of the funnel. The first stage of folding, i.e. the rapid hydrophobic collapse process, is driven by the solvent entropy maximization. Concretely, the water molecules squeeze and sequestrate the hydrophobic amino acid side chains within the interior of the folding intermediates while exposing the polar and electrostatically charged side chains on the intermediate surface so as to minimize the solvent-accessible surface area of the solute and thus, the minimal contacts between the folding intermediates and the water molecules. This will maximize the entropy of the solvent, thus contributing substantially to lowering of the system free energy due to an absolute advantage of the solvent in both quantity and mass (Yang, Ji & Liu, 2012). The resulting molten globule states (Ohgushi & Wada, 1983), within which a few transient secondary structural components and tertiary contacts have been formed but many native contacts or close residue–residue interactions has yet to form, need to be further sculptured into the native states. This is a relatively slow “bottleneck” process because the competitive interactions between protein residues within the folding intermediates and between residues and water molecules may repeat many rounds to accumulate a large enough number of stable noncovalent bonds capable of counteracting the conformational entropy loss of the intermediates, thus putting this bottleneck stage under the enthalpy control (i.e. negative enthalpy change), contributing further to the lowering of the system free energy. Although the protein–ligand association occurs around the rugged bottom of the free energy landscape, the exclusion of water from the binding interfaces and the formation of noncovalent bonds between the two partners can still lower the system free energy. In conjunction with the loss of the rotational and translational degrees of freedom of the two partners as well as the loss of the conformational entropy of the protein, these processes could merge, downwards expand, and further narrow the free energy wells within which the protein–ligand binding process takes place, thereby making them look like a funnel, which we term the binding funnel. In this funnel, the free energy downhill process follows a similar paradigm to the protein-folding process. For example, if the initial collisions/contacts occur between the properly complementary interfaces of the protein and ligand, a large amount of water molecules (which usually form a water network around the solute surface) will be displaced to suit the need for maximizing the solvent entropy. This process is similar to that of the hydrophobic collapse during protein folding, resulting in a loosely associated protein–ligand complex that needs also to be further adapted into a tight complex, i.e. the second step which is mainly driven by the negative enthalpy change through intermolecular competitive interactions to gradually accumulate the noncovalent bonds and ultimately, to stabilize the complex at a tightly bound state. Taken together, we conclude that whether in the protein-folding or in the protein–ligand binding process, both the entropy-driven first step and the enthalpy-driven second step contribute to the lowering of the system free energy, resulting in the funnel-like folding or binding free energy landscape.  相似文献   

18.
We employ Monte Carlo simulations to investigate the interaction between an adsorbing linear flexible cationic polyelectrolyte and a ternary mixed fluid membrane containing neutral (phosphatidylcholine, PC), monovalent (phosphatidylserine, PS), and multivalent (phosphatidylinositol, PIP2) anionic lipids. We systematically explore the influences of polyelectrolyte chain length, polyelectrolyte charge density, polyelectrolyte total charge amount, and salt solution ionic strength on the static and dynamic properties of different anionic lipid species. Our results show that the multivalent PIP2 lipids dominate the polyelectrolyte–membrane interaction and competitively inhibit polyelectrolyte–PS binding. When the total charge amount of the polyelectrolyte is less than that of the local oppositely charged PIP2 lipids, the polyelectrolyte can drag the bound multivalent lipids to diffuse on the membrane, but cannot interact with the PS lipids. Under this condition, the diffusion behaviors of the polyelectrolyte closely follow the prediction of the Rouse model, and the polyelectrolyte chain properties determine the adsorption amount, concentration gradients, and hierarchical mobility of the bound PIP2 lipids. However, when the total charge amount of the polyelectrolyte is larger than that of the local PIP2 lipids, the polyelectrolyte further binds the PS lipids around the polyelectrolyte–PIP2 complex to achieve local electrical neutrality. In this condition, parts of the polyelectrolyte desorb from the membrane and show faster mobility, and the bound PS presents much faster mobility than the segregated PIP2. This work provides an explanation for heterogeneity formation in different anionic lipids induced by polyelectrolyte adsorption.  相似文献   

19.
Thermodynamic analysis is an effective tool in screening of lead-compounds for development of potential drug candidates. In most cases, a ligand achieve high affinity and specificity to a target protein by means of both favorable enthalpy and entropy terms, which can be reflected in binding profiles of Isothermal Titration Calorimetry (ITC). A favorable enthalpy change suggests the contribution of noncovalent contacts such as hydrogen bonding and van der Waals interaction between a ligand and its target protein. In general, optimization of binding enthalpy is more difficult than that of entropies in ligand-design; therefore, it is desirable to choose firstly a lead-compound based on its binding enthalpic gain. In this paper, we demonstrate the utility of thermodynamic approach to ligand screening using anti-ciguatoxin antibody 10C9 as a model of a target protein which possesses a large hydrophobic pocket. As a result of this screening, we have identified three compounds that could bind to the antigen-binding pocket of 10C9 with a few kcal/mol of favorable binding enthalpy. Comparison of their structure with the proper antigen ciguatoxin CTX3C revealed that 10C9 rigorously identifies their cyclic structure and a characteristic hydroxyl group. ITC measurement might be useful and powerful for a rational ligand screening and the optimization of the ligand; the enthalpic gain is an effective index for ligand-design studies.  相似文献   

20.
Strong anion exchange chromatography has frequently been employed as a viral clearance step during downstream processing of biological therapeutics. When challenged with viruses having only slightly acidic isoelectric points, the performance of the anion exchange operation becomes highly dependent on the buffer salt concentration, with the virus log reduction value (LRV) dropping dramatically in buffers with 50–150 mM salt. In this work, a series of anion exchange membrane adsorbers utilizing alternative ligand chemistries instead of the traditional quaternary amine (Q) ligand have been developed that overcome this limitation. Four different ligands (agmatine, tris‐2‐aminoethyl amine, polyhexamethylene biguanide, and polyethyleneimine) achieved >5 LRV of bacteriophage ΦX174 (pI ~ 6.7) at pH 7.5 and up to 150 mM salt, compared to 0 LRV for the Q ligand. By evaluating structural derivatives of the successful ligands, three factors were identified that contributed to ligand salt tolerance: ligand net charge, ligand immobilization density on the membrane, and molecular structure of the ligand‐binding group. Based on the results of this study, membrane adsorbers that incorporate alternative ligands provide a more robust and salt tolerant viral clearance‐processing step compared to traditional strong anion exchange membrane adsorbers. Biotechnol. Bioeng. 2009;103: 920–929. © 2009 Wiley Periodicals, Inc.  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