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 375 毫秒
1.
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   

2.
A detailed analysis of the DNA-binding sites of 26 proteins is presented using data from the Nucleic Acid Database (NDB) and the Protein Data Bank (PDB). Chemical and physical properties of the protein-DNA interface, such as polarity, size, shape, and packing, were analysed. The DNA-binding sites shared common features, comprising many discontinuous sequence segments forming hydrophilic surfaces capable of direct and water-mediated hydrogen bonds. These interface sites were compared to those of protein-protein binding sites, revealing them to be more polar, with many more intermolecular hydrogen bonds and buried water molecules than the protein-protein interface sites. By looking at the number and positioning of protein residue-DNA base interactions in a series of interaction footprints, three modes of DNA binding were identified (single-headed, double-headed and enveloping). Six of the eight enzymes in the data set bound in the enveloping mode, with the protein presenting a large interface area effectively wrapped around the DNA.A comparison of structural parameters of the DNA revealed that some values for the bound DNA (including twist, slide and roll) were intermediate of those observed for the unbound B-DNA and A-DNA. The distortion of bound DNA was evaluated by calculating a root-mean-square deviation on fitting to a canonical B-DNA structure. Major distortions were commonly caused by specific kinks in the DNA sequence, some resulting in the overall bending of the helix. The helix bending affected the dimensions of the grooves in the DNA, allowing the binding of protein elements that would otherwise be unable to make contact. From this structural analysis a preliminary set of rules that govern the bending of the DNA in protein-DNA complexes, are proposed.  相似文献   

3.
PDZ domains are widespread protein modules that commonly recognize C-terminal sequences of target proteins and help to organize macromolecular signaling complexes. These sequences usually bind in an extended conformation to relatively shallow grooves formed between a beta-strand and an alpha-helix in the corresponding PDZ domains. Because of this binding mode, many PDZ domains recognize primarily the C-terminal and the antepenultimate side-chains of the target protein, which commonly conform to motifs that have been categorized into different classes. However, an increasing number of PDZ domains have been found to exhibit unusual specificities. These include the PDZ domain of RIMs, which are large multidomain proteins that regulate neurotransmitter release and help to organize presynaptic active zones. The RIM PDZ domain binds to the C-terminal sequence of ELKS with a unique specificity that involves each of the four ELKS C-terminal residues. To elucidate the structural basis for this specificity, we have determined the 3D structure in solution of an RIM/ELKS C-terminal peptide complex using NMR spectroscopy. The structure shows that the RIM PDZ domain contains an unusually deep and narrow peptide-binding groove with an exquisite shape complementarity to the four ELKS C-terminal residues in their bound conformation. This groove is formed, in part, by a set of side-chains that is conserved selectively in RIM PDZ domains and that hence determines, at least in part, their unique specificity.  相似文献   

4.
The Oxytricha nova telomere end binding protein (OnTEBP) recognizes, binds and protects the single-stranded 3'-terminal DNA extension found at the ends of macronuclear chromosomes. The structure of this complex shows that the single strand GGGGTTTTGGGG DNA binds in a deep cleft between the two protein subunits of OnTEBP, adopting a non-helical and irregular conformation. In extending the resolution limit of this structure to 1.86 A, we were surprised to find a G-quartet linked dimer of the GGGGTTTTGGGG DNA also packing within the crystal lattice and interacting with the telomere end binding protein. The G-quartet DNA exhibits the same structure and topology as previously observed in solution by NMR with diagonally crossing d(TTTT) loops at either end of the four-stranded helix. Additionally, the crystal structure reveals clearly visible Na(+), and specific patterns of bound water molecules in the four non-equivalent grooves. Although the G-quartet:protein contact surfaces are modest and might simply represent crystal packing interactions, it is interesting to speculate that the two types of telomeric DNA-protein interactions observed here might both be important in telomere biology.  相似文献   

5.
Protein interactions are often accompanied by significant changes in conformation. We have analyzed the relationships between protein structures and the conformational changes they undergo upon binding. Based upon this, we introduce a simple measure, the relative solvent accessible surface area, which can be used to predict the magnitude of binding-induced conformational changes from the structures of either monomeric proteins or bound subunits. Applying this to a large set of protein complexes suggests that large conformational changes upon binding are common. In addition, we observe considerable enrichment of intrinsically disordered sequences in proteins predicted to undergo large conformational changes. Finally, we demonstrate that the relative solvent accessible surface area of monomeric proteins can be used as a simple proxy for protein flexibility. This reveals a powerful connection between the flexibility of unbound proteins and their binding-induced conformational changes, consistent with the conformational selection model of molecular recognition.  相似文献   

6.
Hydration of protein cavities influences protein stability, dynamics, and function. Protein active sites usually contain water molecules that, upon ligand binding, are either displaced into bulk solvent or retained to mediate protein–ligand interactions. The contribution of water molecules to ligand binding must be accounted for to compute accurate values of binding affinities. This requires estimation of the extent of hydration of the binding site. However, it is often difficult to identify the water molecules involved in the binding process when ligands bind on the surface of a protein. Cytochrome P450cam is, therefore, an ideal model system because its substrate binds in a buried active site, displacing partially disordered solvent, and the protein is well characterized experimentally. We calculated the free energy differences for having five to eight water molecules in the active site cavity of the unliganded enzyme from molecular dynamics simulations by thermodynamic integration employing a three-stage perturbation scheme. The computed free energy differences between the hydration states are small (within 12 kJ mol−1) but distinct. Consistent with the crystallographic determination and studies employing hydrostatic pressure, we calculated that, although ten water molecules could in principle occupy the volume of the active site, occupation by five to six water molecules is thermodynamically most favorable. Proteins 32:381–396, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

7.
8.
Protein folding and the transfer of hydrocarbons from a dilute aqueous solution to the pure liquid phase are thermodynamically similar in that both processes remove nonpolar surface from water and both are accompanied by anomalously large negative heat capacity changes. On the basis of a limited set of published surface areas, we previously proposed that heat capacity changes (delta C degrees p) for the transfer of hydrocarbons from water to the pure liquid phase and for the folding of globular proteins exhibit the same proportionality to the reduction in water-accessible nonpolar surface area (delta Anp) [Spolar, R.S., Ha, J.H., & Record, M.T., Jr. (1989) Proc. Natl. Acad. Sci. U.S.A. 86, 8382-8385]. The consequence of this proposal is that the experimental delta C degrees p for protein folding can be used to obtain estimates of delta Anp and of the contribution to the stability of the folded state from removal of a nonpolar surface from water. In this paper, a rigorous molecular surface area algorithm [Richmond, T.J. (1984) J. Mol. Biol. 178, 63-89] is applied to obtain self-consistent values of the water-accessible nonpolar surface areas of the native and completely denatured states of the entire set of globular proteins for which both crystal structures and delta C degrees p of folding have been determined and for the set of liquid and liquefiable hydrocarbons for which delta C degrees p of transfer are known. Both processes (hydrocarbon transfer and protein folding) exhibit the same direct proportionality between delta C degrees p and delta Anp. We conclude that the large negative heat capacity changes observed in protein folding and other self-assembly processes involving proteins provide a quantitative measure of the reduction in the water-accessible nonpolar surface area and of the contribution of the hydrophobic effect to the stability of the native state and to protein assembly.  相似文献   

9.
The X-ray structure of the periplasmic ribose receptor (binding protein) of Escherichia coli (RBP) was solved at 3 A resolution by the method of multiple isomorphous replacement. Alternating cycles of refitting and refinement have resulted in a model structure with an R-factor of 18.7% for 27,526 reflections from 7.5 to 1.7 A resolution (96% of the data). The model contains 2228 non-hydrogen atoms, including all 271 residues of the amino acid sequence, 220 solvent atoms and beta-D-ribose. The protein consists of two highly similar structural domains, each of which is composed of a core of parallel beta-sheet flanked on both sides by alpha-helices. The two domains are related to each other by an almost perfect 2-fold axis of rotation, with the C termini of the beta-strands of each sheet pointing toward the center of the molecule. Three short stretches of amino acid chain (from symmetrically related portions of the protein) link these two domains, and presumably act as a hinge to allow relative movement of the domains in functionally important conformational changes. Two water molecules are also an intrinsic part of the hinge, allowing crucial flexibility in the structure. The ligand beta-D-ribose (in the pyranose form) is bound between the domains, held by interactions with side-chains of the interior loops. The binding site is precisely tailored, with a combination of hydrogen bonding, hydrophobic and steric effects giving rise to tight binding (0.1 microM for ribose) and high specificity. Four out of seven binding-site residues are charged (2 each of aspartate and arginine) and contribute two hydrogen bonds each. The remaining hydrogen bonds are contributed by asparagine and glutamine residues. Three phenylalanine residues supply the hydrophobic component, packing against both faces of the sugar molecule. The arrangement of these hydrogen bonding and hydrophobic residues results in an enclosed binding site with the exact shape of the allowed sugar molecules; in the process of binding, the ligand loses all of its surface-accessible area. The sites of two mutations that affect the rate of folding of the ribose receptor are shown to be located near small cavities in the wild-type protein. The cavities thus allow the incorporation of the larger residues in the mutant proteins. Since these alterations would seriously affect the ability of the protein to build the first portion of the hydrophobic core in the first domain, it is proposed that this process is the rate-limiting step in folding of the ribose receptor.  相似文献   

10.
Understanding the mechanisms of the interaction between a protein surface and its outer molecular environment is of primary relevance for the rational design of new drugs and engineered proteins. Protein surface accessibility is emerging as a new dimension of Structural Biology, since NMR methods have been developed to follow how molecules, even those different from physiological ligands, preferentially approach specific regions of the protein surface. Hen egg-white lysozyme, a paradigmatic example of the state of the art of protein structure and dynamics, has been selected as a model system to study protein surface accessibility. Bound water and soluble spin-labels have been used to investigate the interaction of this enzyme, both free and bound to the inhibitor (NAG)(3), with its molecular environment. No tightly bound water molecules were found inside the enzyme active site, which, conversely, appeared as the most exposed to visits from the soluble paramagnetic probe TEMPOL. From the presented set of data, an integrated view of lysozyme surface accessibility towards water and TEMPOL molecules is obtained.  相似文献   

11.
This paper reports on the use of molecular dynamics (MD) simulation to investigate the coupling effects of wettability, surface roughness and interfacial nanobubbles (INBs) on wall–fluid interfaces. The fluid properties close to the wall–fluid interface, such as potential energy, density, diffusion coefficients of fluid molecules and effective slip length are simulated. In the cases without surface nanobubbles, regions with lower potential energy have a higher probability of hosting water molecules. The local translational and rotational diffusion coefficients of water within the cavities are strongly influenced by wettability but largely unaffected by hydrodynamic effects. In cases where INBs exist, variations in wettability result in distinctly different argon morphologies. Argon nanobubbles form a convex shape on Wenzel-like interfaces but a shallow concave shape on Cassie-like interfaces. The phenomenon of water molecules invading grooves tends to occur on Wenzel-like interfaces; however, this depends largely on the morphology of the grooves. The high mobility and high density of argon molecules indicate that the state of the argon molecules within the grooves may require further investigation. Our results also show that the effective slip length is significantly influenced by wall–fluid wettability as well as the morphology of INBs.  相似文献   

12.
We refine a protein model that reproduces fundamental aspects of protein thermodynamics. The model exhibits two transitions, hot and cold unfolding. The number of relevant parameters is reduced to three: (1) binding energy of folding relative to the orientational energy of bound water, (2) ratio of degrees of freedom between the folded and unfolded protein chain, and (3) the number of water molecules that can access the hydrophobic parts of the protein interior upon unfolding. By increasing the number of water molecules in the model, the separation between the two peaks in the heat capacity curve comes closer, which is more consistent with experimental data. In the end we show that if we, as a speculative assumption, assign only two distinct energy levels for the bound water molecules, better correspondence with experiments can be obtained.  相似文献   

13.
The buried surface area (BSA), which measures the size of the interface in a protein–protein complex may differ from the accessible surface area (ASA) lost upon association (which we call DSA), if conformation changes take place. To evaluate the DSA, we measure the ASA of the interface atoms in the bound and unbound states of the components of 144 protein–protein complexes taken from the Protein–Protein Interaction Affinity Database of Kastritis et al. (2011). We observe differences exceeding 20%, and a systematic bias in the distribution. On average, the ASA calculated in the bound state of the components is 3.3% greater than in their unbound state, and the BSA, 7% greater than the DSA. The bias is observed even in complexes where the conformation changes are small. An examination of the bound and unbound structures points to a possible origin: local movements optimize contacts with the other component at the cost of internal contacts, and presumably also the binding free energy.  相似文献   

14.
We have performed a multivariate logistic regression analysis to establish a statistical correlation between the structural properties of water molecules in the binding site of a free protein crystal structure, with the probability of observing the water molecules in the same location in the crystal structure of the ligand-complexed form. The temperature B-factor, the solvent-contact surface area, the total hydrogen bond energy and the number of protein-water contacts were found to discriminate between bound and displaceable water molecules in the best regression functions obtained. These functions may be used to identify those bound water molecules that should be included in structure-based drug design and ligand docking algorithms.FIGURE The binding site ( thin sticks) of penicillopepsin (3app) with its crystallographically determined water molecules ( spheres) and superimposed ligand (in thick sticks, from complexed structure 1ppk). Water molecules sterically displaced by the ligand upon complexation are shown in cyan. Bound water molecules are shown in blue. Displaced water molecules are shown in yellow. Water molecules removed from the analysis due to a lack of hydrogen bonds to the protein are shown in white. WaterScore correctly predicted waters in blue as Probability=1 to remain bound and waters in yellow as Probability<1x10(-20) to remain bound.  相似文献   

15.
16.
Protein adsorption is a source of variability in the release profiles of therapeutic proteins from biodegradable microspheres. We employ optical reflectometry and total internal reflection fluorescence to explore the extent and kinetics of ribonuclease A (RNase A) adsorption to spin-cast films of poly(lactide-co-glycolide) (PLG) and, in particular, to determine how covalent grafting of polyethylene glycol (PEG) to RNase A affects adsorption. Adsorption kinetics on PLG surfaces are surface-limited for RNase A but transport-limited for unconjugated PEG homopolymers and for PEG-modified RNase A, indicating that PEG anchors the conjugates to the surface during the transport-limited regime. PEG modification of RNase A decreases the total number of adsorbed molecules per unit area but increases the areal surface coverage because the grafted PEG chains exclude additional surface area. Total internal reflection fluorescence-based exchange measurements show that there is no exchange between adsorbed and solution-phase protein molecules. This indicates an unusually tenacious adsorption. Streaming current measurements indicate that the zeta potential of the PLG surface becomes increasingly negative as the film is exposed to water for several weeks, as expected. Aging of the PLG surface results in increased adsorption of unmodified RNase A but decreased adsorption of unconjugated PEG homopolymers and of PEG-RNase A conjugates, relative to the extent of adsorption on freshly prepared PLG surfaces. Adsorption results correlate well with an increase in the rate, total extent and preservation of bioactivity of RNase A released from PLG microspheres for the PEG-modified version of RNase A.  相似文献   

17.
A range of studies were carried out to investigate the underlying reason for differences in dynamic binding capacities observed with various antibodies and Fc-fusion proteins during Protein A chromatography. Dynamic binding capacities were determined for these biomolecules using different protein A stationary phase materials. SEC was carried out to determine the relative sizes of the antibodies and fusion proteins. Pore diffusivities and static binding capacities were also determined on these Protein A resin materials. Trends in the dynamic binding capacities for these molecules did not correlate with differences in pore diffusion coefficients as might be expected for a mass transfer limited system. Instead, dynamic binding capacities were seen to follow the same trends as the static binding capacities and the apparent size of the molecules. Differences in static binding capacities were attributed to be due to differences in steric factor between the molecules. Solution binding stoichiometry studies were employed to estimate intra-Protein A steric effects while binding to the various domains within a Protein A ligand. In addition, steric hindrance was also found to exist between adjacent immobilized Protein A ligands on the chromatographic surface. The combination of intra and inter Protein A steric hindrances can explain differences in binding capacities observed between various antibody and Fc fusion proteins. The effect of Protein A ligand density on these supports was also examined and the results indicate that increasing Protein A ligand density leads to a situation of diminishing returns for binding capacity due to increased steric hindrance on the resin surface. The results presented in this paper show that steric hindrances can dominate over mass transfer effects in causing capacity variation between different molecules on the same stationary phase. This can lead to the development of more cost-efficient chromatographic stationary phases as well as provide information during the selection of Protein A media for preparative purification of monoclonal antibodies and Fc fusion proteins.  相似文献   

18.
Protein disulfide isomerases comprise a large family of enzymes responsible for catalyzing the proper oxidation and folding of newly synthesized proteins in the endoplasmic reticulum (ER). Protein disulfide isomerase-related (PDIR) protein (also known as PDIA5) is a specialized member that participates in the folding of α1-antitrypsin and N-linked glycoproteins. Here, the crystal structure of the non-catalytic domain of PDIR was determined to 1.5 Å resolution. The structure adopts a thioredoxin-like fold stabilized by a structural disulfide bridge with a positively charged binding surface for interactions with the ER chaperones, calreticulin and ERp72. Crystal contacts between molecules potentially mimic the interactions of PDIR with misfolded substrate proteins. The results suggest that the non-catalytic domain of PDIR plays a key role in the recognition of protein partners and substrates.  相似文献   

19.
Local hydrophobic collapse of the polypeptide chain and transient long-range interactions in unfolded states of apomyoglobin appear to occur in regions of the amino acid sequence which, upon folding, bury an above-average area of hydrophobic surface. To explore the role of these interactions in protein folding, we prepared and characterized apomyoglobins with compensating point mutations designed to change the average buried surface area in local regions of the sequence, while conserving as much as possible the constitution of the hydrophobic core. The behavior of the mutants in quench-flow experiments to determine the folding pathway was exactly as predicted by the changes in the buried surface area parameter calculated from the amino acid sequence. In addition, spin label experiments with acid-unfolded mutant apomyoglobin showed that the transient long-range contacts that occur in the wild-type protein are abolished in the mutant, while new contacts are observed between areas that now have above-average buried surface area. We conclude that specific groupings of amino acid side-chains, which can be predicted from the sequence, are responsible for early hydrophobic interactions in the first phase of folding in apomyoglobin, and that these early interactions determine the subsequent course of the folding process.  相似文献   

20.
Kríz Z  Otyepka M  Bártová I  Koca J 《Proteins》2004,55(2):258-274
The interactions between the protein and the solvent were analyzed, and protein regions with a high density of water molecules, as well as structural water molecules, were determined by using molecular dynamics (MD) simulations. A number of water molecules that were in contact with the protein for the whole trajectory were determined. Their interaction energies and hydrogen bonds with protein residues were analyzed. Altogether, 39, 27, 49, and 32 water molecules bound to the protein were found for trajectories of the free CDK2, CDK2/ATP, CDK2/roscovitine, and CDK2/isopentenyladenine complexes, respectively. Positions of observed water molecules were compared with X-ray crystallography data. Special attention was paid to water molecules in the active site of the enzyme, and especially to the deep pocket, where the N9 roscovitine side-chain is buried. Exchange of active-site water molecules with bulk water through the tunnel from the pocket was observed. In the CDK2/isopentenyladenine complex simulation, two water molecules that arrange interaction between the inhibitor and the enzyme via an H-bond were observed. Two stable water molecules in the trajectory of the free CDK2 were found that occupy the same position as the nitrogens N3 and N9 of the isopentenyladenine or N1 and N6 nitrogens of the adenosine triphosphate (ATP). The positions of structural water molecules were compared with the positions of substrate polar groups and crystallographic water molecules found in the Brookhaven Protein Data Bank for various CDK2 complexes. It was concluded that tracing tightly bound water molecules may substantially help in designing new inhibitors.  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