首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
In Part 1 of this study (Weinbaum et al., 1988) a short time model has been proposed to describe the initial time dependent leakage of macromolecules at short distances (5 microns or less) from the exit of a transient open junction which the authors have hypothesized as a characteristic feature of endothelial cells in the process of turnover (Weinbaum et al., 1985). This open junction pathway has also been proposed (Weinbaum et al., 1988) to be the primary ultrastructural correlate of the 20 nm diameter large pore suggested by Renkin et al. (1977) using the predictions of cylindrical pore theory. The short time model in (Weinbaum et al., 1988), however, has major limitations in that it neglects the interaction between leakage sites, macromolecular entry through other pathways, the finite thickness of the vessel wall and the curvature of the cell perimeter. The longer time model developed herein will attempt to describe each of these features and also present an improved model and analytic solution for the steady state flux and uptake. In the previous steady state model developed by Weinbaum et al. (1985) the effect of the resistance of the transient open junctions and the non-isotropic diffusion in the underlying tissue due to the internal elastic lamina (IEL) were both neglected. New solutions are first presented which describe the effect of these important model refinements on the steady state macromolecular permeability of the major arteries. Time dependent solutions are then presented to predict the transient longer time labeling following the introduction of tracer macromolecules of varying size. These solutions and the corresponding short time solutions in Weinbaum et al. (1988) are the first solutions to our knowledge to describe the difficult time-dependent boundary value problem to determine how the channel exit concentration and flux at a leaky junction vary with time. This is accomplished by casting the boundary value problem in the form of an integral equation for the unknown flux at the cleft exit and then solving this problem using a specially designed numerical technique. The theoretical predictions are used to interpret the behavior of the localized leaks to HRP and albumin that have been reported in Stemerman et al. (1986) and our own recent experiments (Lin et al., 1988).  相似文献   

2.
Cytochrome c2 (cyt) is the mobile electron donor to the reaction center (RC) in photosynthetic bacteria. The electrostatic interactions involved in the dynamics of docking of cyt onto the RC were examined by double mutant studies of the rates of electron transfer between six modified Rhodobacter sphaeroides RCs in which negatively charged acid residues were replaced with Lys and five modified Rhodobacter capsulatus Cyt c2 molecules in which positively charged Lys residues were replaced with Glu. We measured the second-order rate constant, k2, for electron transfer from the reduced cyt to the oxidized primary donor on the RC, which reflects the energy of the transition state for the formation of the active electron transfer complex. Strong interactions were found between Lys C99 and Asp M184/Glu M95, and between Lys C54 and Asp L261/Asp L257. The interacting residues were found to be located close to each other in the recently determined crystal structure of the cyt-RC complex [Axelrod, H., et al. (2002) J. Mol. Biol. (in press)]. The interaction energies were approximately inversely proportional to the distances between charges. These results support earlier suggestions [Tetreault, M., et al. (2001) Biochemistry 40, 8452-8462] that the structure of the transition state in solution resembles the structure of the cyt-RC complex in the cocrystal and indicate that specific electrostatic interactions facilitate docking of the cyt onto the RC in a configuration optimized for both binding and electron transfer. The specific interaction between Asp M184 and Lys C99 may help to nucleate short-range hydrophobic contacts.  相似文献   

3.
The structure of the antigen-binding fragment (Fab) of an anti-p-azophenylarsonate monoclonal antibody, 36-71, bearing a major cross-reactive idiotype of A/J mice has been refined to an R factor of 24.8% at a resolution of 1.85 A. The previously solved partial structure of this Fab at a resolution of 2.9 A (Rose et al., 1990) was used as an initial model for refinement against the high-resolution data. The complex with hapten has been modeled by docking the small-molecule crystal structure of phenylarsonic acid into the structure of the native Fab on the basis of a low-resolution electron density map of the complex. In this model, residue Arg-96 in the light chain and residues Asn-35, Trp-47, and Ser-99 in the heavy chain contact the arsonate moiety of the hapten; an additional bond is found between the arsonate group and a tightly bound water molecule. The phenyl moiety of the hapten packs against two tyrosine side chains at positions 50 and 106 in the heavy chain. Residue Arg-96 in the light chain had been implicated as involved in hapten binding on the basis of previous experiments, and indeed, this residue appears to play a crucial role in this model. Experiments employing site-directed mutagenesis directly support this conclusion. The heavy-chain complementarity-determining regions have novel conformations not previously observed in immunoglobulins except for the recently solved anti-p-azophenylarsonate Fab R 19.9 (Lascombe et al., 1989).  相似文献   

4.
van Dijk AD  Fushman D  Bonvin AM 《Proteins》2005,60(3):367-381
When classical, Nuclear Overhauser Effect (NOE)-based approaches fail, it is possible, given high-resolution structures of the free molecules, to model the structure of a complex in solution based solely on chemical shift perturbation (CSP) data in combination with orientational restraints from residual dipolar couplings (RDCs) when available. RDCs can be incorporated into the docking following various strategies: as direct restraints and/or as intermolecular intervector projection angle restraints (Meiler et al., J Biomol NMR 2000;16:245-252). The advantage of the latter for docking is that they directly define the relative orientation of the molecules. A combined protocol in which RDCs are first introduced as intervector projection angle restraints and at a later stage as direct restraints is shown here to give the best performance. This approach, implemented in our information-driven docking approach HADDOCK (Dominguez et al., J Am Chem Soc 2003;125:1731-1737), is used to determine the solution structure of the Lys48-linked di-ubiquitin, for which chemical shift mapping, RDCs, and (15)N-relaxation data have been previously obtained (Varadan et al., J Mol Biol 2002;324:637-647). The resulting structures, derived from CSP and RDC data, are cross-validated using (15)N-relaxation data. The solution structure differs from the crystal structure by a 20 degrees rotation of the two ubiquitin units relative to each other.  相似文献   

5.
In this paper, we present our results from a docking study of the title compounds with the DNA/topoisomerase I complex based on the recently published X-ray crystal structure of the topotecan/DNA/topoisomerase I ternary cleavable complex (Staker, B.L., et al. PNAS 2002, 99, 15387) using the Autodock program. Simple intermolecular docking energies (E(dock)) correlate well with in vitro DNA cleavage data suggesting that the binding mode from the crystal structure is a reasonable binding mode for these compounds.  相似文献   

6.
In the tightly regulated glycogenolysis cascade, the breakdown of glycogen to glucose‐1‐phosphate, phosphorylase kinase (PhK) plays a key role in regulating the activity of glycogen phosphorylase. PhK is a 1.3 MDa hexadecamer, with four copies each of four different subunits (α, β, γ and δ), making the study of its structure challenging. Using hydrogen‐deuterium exchange, we have analyzed the regulatory β subunit and the catalytic γ subunit in the context of the intact non‐activated PhK complex to study the structure of these subunits and identify regions of surface exposure. Our data suggest that within the non‐activated complex the γ subunit assumes an activated conformation and are consistent with a previous docking model of the β subunit within the cryoelectron microscopy envelope of PhK.  相似文献   

7.
We present a general approach to the design, docking, and virtual screening of multiple combinatorial libraries against a family of proteins. The method consists of three main stages: docking the scaffold, selecting the best substituents at each site of diversity, and comparing the resultant molecules within and between the libraries. The core "divide-and-conquer" algorithm for side-chain selection, developed from an earlier version (Sun et al., J Comp Aided Mol Design 1998;12:597-604), provides a way to explore large lists of substituents with linear rather than combinatorial time dependence. We have applied our method to three combinatorial libraries and three serine proteases: trypsin, chymotrypsin, and elastase. We show that the scaffold docking procedure, in conjunction with a novel vector-based orientation filter, reproduces crystallographic binding modes. In addition, the free-energy-based scoring procedure (Zou et al., J Am Chem Soc 1999;121:8033-8043) is able to reproduce experimental binding data for P1 mutants of macromolecular protease inhibitors. Finally, we show that our method discriminates between a peptide library and virtual libraries built on benzodiazepine and tetrahydroisoquinolinone scaffolds. Implications of the docking results for library design are explored.  相似文献   

8.
Camelids produce functional "heavy chain" antibodies which are devoid of light chains and CH1 domains [Hamers-Casterman, C., et al. (1993) Nature 363, 446-448]. It has been shown that the variable domains of these heavy chain antibodies (the V(HH) fragments) are functional at or after exposure to high temperatures, in contrast to conventional antibodies [Linden van der, R. H. J., et al. (1999) Biochim. Biophys. Acta 1431, 37-44]. For a detailed understanding of the higher thermostability of these V(HH) fragments, knowledge of their structure and conformational dynamics is required. As a first step toward this goal, we report here the essentially complete (1)H and (15)N NMR backbone resonance assignments of a llama V(HH) antibody fragment, and an extensive analysis of the structure at higher temperatures. The H-D exchange NMR data at 300 K indicate that the framework of the llama V(HH) fragment is highly protected with a DeltaG(ex) of >5.4 kcal/mol, while more flexibility is observed for surface residues, particularly in the loops and the two outer strands (residues 4-7, 10-13, and 58-60) of the beta-sheet. The CD data indicate a reversible, two-state unfolding mechanism with a melting transition at 333 K and a DeltaH(m) of 56 kcal/mol. H-D exchange studies using NMR and ESI-MS show that below 313 K exchange occurs through local unfolding events whereas above 333 K exchange mainly occurs through global unfolding. The lack of a stable core at high temperatures, observed for V(HH) fragments, has also been observed for conventional antibody fragments. The main distinction between the llama V(HH) fragment and conventional antibody fragments is the reversibility of the thermal unfolding process, explaining its retained functionality after exposure to high temperatures.  相似文献   

9.
Histone tail modifications control many nuclear processes by dictating the dynamic exchange of regulatory proteins on chromatin. Here we report novel insights into histone H3 tail structure in complex with the double PHD finger (DPF) of the lysine acetyltransferase MOZ/MYST3/KAT6A. In addition to sampling H3 and H4 modification status, we show that the DPF cooperates with the MYST domain to promote H3K9 and H3K14 acetylation, although not if H3K4 is trimethylated. Four crystal structures of an extended DPF alone and in complex with unmodified or acetylated forms of the H3 tail reveal the molecular basis of crosstalk between H3K4me3 and H3K14ac. We show for the first time that MOZ DPF induces α-helical conformation of H3K4-T11, revealing a unique mode of H3 recognition. The helical structure facilitates sampling of H3K4 methylation status, and proffers H3K9 and other residues for modification. Additionally, we show that a conserved double glycine hinge flanking the H3 tail helix is required for a conformational change enabling docking of H3K14ac with the DPF. In summary, our data provide the first observations of extensive helical structure in a histone tail, revealing the inherent ability of the H3 tail to adopt alternate conformations in complex with chromatin regulators.  相似文献   

10.
It has been shown before by (31)P NMR that Ras bound to the nonhydrolyzable GTP analogue guanosine 5'-O-(beta, gamma-imidotriphosphate) (GppNHp) exists in two conformations which are rapidly interconverting with a rate constant of 3200 s-1 at 30 degrees C [Geyer, M., et al. (1996) Biochemistry 35, 10308-10320]. Here we show that Ran complexed with GTP also exists in two conformational states, 1 and 2, which can be directly inferred from the occurrence of two (31)P NMR resonance lines for the gamma-phosphate group of bound GTP. The exchange between the two states is slow on the NMR time scale with a value of <200 s-1 at 5 degrees C for the corresponding first-order rate constants. In wild-type Ran, the equilibrium constant K' between the two states is 0.7 at 278 K, is different for various mutants, and is strongly dependent on the temperature. The standard enthalpy DeltaH degrees and the standard entropy DeltaS degrees for the conformational transitions determined from the NMR spectra are as follows: DeltaH degrees = 37 kJ mol-1 and DeltaS degrees = 130 J mol-1 K-1 for wild-type Ran.GTP. In complex with the Ran-binding protein RanBP1, one of the Ran.GTP conformations (state 2) is stabilized. The interaction of Ran with the guanine nucleotide exchange factor protein RCC1 was also studied by (31)P NMR spectroscopy. In the presence of nucleotide, the ternary complex of Ran.nucleotide.RCC1, an intermediate in the guanine nucleotide exchange reaction, could be observed. A model for the conformational transition of Ran.GTP is proposed where the two states observed are caused by the structural flexibility of the effector loop of Ran; in solution, state 2 resembles the GTP-bound form found in the crystal structure of the Ran-RanBP complex.  相似文献   

11.
We report the backbone dynamics of monomeric phospholamban in dodecylphosphocholine micelles using (1)H/(15)N heteronuclear NMR spectroscopy. Phospholamban is a 52-amino acid membrane protein that regulates Ca-ATPase in cardiac muscle. Phospholamban comprises three structural domains: a transmembrane domain from residues 22 to 52, a connecting loop from 17 to 21, and a cytoplasmic domain from 1 to 16 that is organized in an "L"-shaped structure where the transmembrane and the cytoplasmic domain form an angle of approximately 80 degrees (Zamoon et al., 2003; Mascioni et al., 2002). T(1), T(2), and (1)H/(15)N nuclear Overhauser effect values measured for the amide backbone resonances were interpreted using the model-free approach of Lipari and Szabo. The results point to the existence of four dynamic domains, revealing the overall plasticity of the cytoplasmic helix, the flexible loop, and part of the transmembrane domain (residues 22-30). In addition, using Carr-Purcell-Meiboom-Gill-based experiments, we have characterized phospholamban dynamics in the micros-ms timescale. We found that the majority of the residues in the cytoplasmic domain, the flexible loop, and the first ten residues of the transmembrane domain undergo dynamics in the micros-ms range, whereas minimal dynamics were detected for the transmembrane domain. Hydrogen/deuterium exchange factors measured at different temperatures support the existence of slow motion in both the loop and the cytoplasmic helix. We propose that these dynamic properties are critical factors in the biomolecular recognition of phospholamban by Ca-ATPase and other interacting proteins such as protein kinase A and protein phosphatase 1.  相似文献   

12.
Using BIACORE SPR, we have examined the mechanism of temperature effects on the binding kinetics of two closely related antibody Fabs (H10 and H26) which recognize coincident epitopes on hen egg-white lysozyme (HEL), and whose association and dissociation kinetics are best described by the two-step conformational change model which we interpret as molecular encounter and docking. Time-course series data obtained at a series of six temperatures (6, 10, 15, 25, 30 and 37 degrees C) showed that temperature differentially affects the rate constants of the encounter and docking steps. Docking is more temperature-sensitive than the encounter step, and energetically less favorable at higher temperatures. At elevated temperatures, the time required for docking is longer and the apparent increase in off-rate reflects the greater proportion of the molecules failing to dock and remaining in the less stable encounter state. As a consequence, distribution of free energy change between the encounter and docking steps is altered. At physiological temperature (37 degrees C) the docking step of the H26 complex is energetically unfavorable and most complexes essentially do not dock. There is a significant decrease in total free energy change of the H26 complex at higher temperatures. Elevated temperature changes the rate-limiting step of H26--HEL association from the encounter to the docking step, but not that of H10--HEL. Our results indicate that the mechanism by which elevated temperature reduces the affinities of antigen--antibody complexes is to decrease the net docking rate, and/or stability of the docked complex; at higher temperatures, a smaller proportion of the complexes actually anneal to a more stable docked state. This mechanism may have broad applicability to other receptor--ligand complexes.  相似文献   

13.
Laser-induced temperature jump experiments were used for testing the rates of thermoinduced conformational transitions of reaction center (RC) complexes in chromatophores of Chromatium minutissimum. The thermoinduced transition of the macromolecular RC complex to a state providing effective electron transport from the multiheme cytochrome c to the photoactive bacteriochlorophyll dimer within the temperature range 220–280 K accounts for tens of seconds with activation energy 0.166 eV/molecule. The rate of the thermoinduced transition in the cytochrome–RC complex was found to be three orders of magnitude slower than the rate of similar thermoinduced transition of the electron transfer reaction from the primary to secondary quinone acceptors studied in the preceding work (Chamorovsky et al. in Eur Biophys J 32:537–543, 2003). Parameters of thermoinduced activation of the electron transfer from the multiheme cytochrome c to the photoactive bacteriochlorophyll dimer are discussed in terms of cytochrome c docking onto the RC.  相似文献   

14.
15.
MOTIVATION: Protein-protein docking algorithms typically generate large numbers of possible complex structures with only a few of them resembling the native structure. Recently (Duan et al., Protein Sci, 14:316-218, 2005), it was observed that the surface density of conserved residue positions is high at the interface regions of interacting protein surfaces, except for antibody-antigen complexes, where a lesser number of conserved positions than average is observed at the interface regions. Using this observation, we identified putative interacting regions on the surface of interacting partners and significantly improved docking results by assigning top ranks to near-native complex structures. In this paper, we combine the residue conservation information with a widely used shape complementarity algorithm to generate candidate complex structures with a higher percentage of near-native structures (hits). What is new in this work is that the conservation information is used early in the generation stage and not only in the ranking stage of the docking algorithm. This results in a significantly larger number of generated hits and an improved predictive ability in identifying the native structure of protein-protein complexes. RESULTS: We report on results from 48 well-characterized protein complexes, which have enough residue conservation information from the same 59 benchmark complexes used in our previous work. We compute conservation indices of residue positions on the surfaces of interacting proteins using available homologous sequences from UNIPROT and calculate the solvent accessible surface area. We combine this information with shape-complementarity scores to generate candidate protein-protein complex structures. When compared with pure shape-complementarity algorithms, performed by FTDock, our method results in significantly more hits, with the improvement being over 100% in many instances. We demonstrate that residue conservation information is useful not only in refinement and scoring of docking solutions, but also helpful in enrichment of near-native-structures during the generation of candidate geometries of complex structures.  相似文献   

16.
A challenge in protein-protein docking is to account for the conformational changes in the monomers that occur upon binding. The RosettaDock method, which incorporates sidechain flexibility but keeps the backbone fixed, was found in previous CAPRI rounds (4 and 5) to generate docking models with atomic accuracy, provided that conformational changes were mainly restricted to protein sidechains. In the recent rounds of CAPRI (6-12), large backbone conformational changes occur upon binding for several target complexes. To address these challenges, we explicitly introduced backbone flexibility in our modeling procedures by combining rigid-body docking with protein structure prediction techniques such as modeling variable loops and building homology models. Encouragingly, using this approach we were able to correctly predict a significant backbone conformational change of an interface loop for Target 20 (12 A rmsd between those in the unbound monomer and complex structures), but accounting for backbone flexibility in protein-protein docking is still very challenging because of the significantly larger conformational space, which must be surveyed. Motivated by these CAPRI challenges, we have made progress in reformulating RosettaDock using a "fold-tree" representation, which provides a general framework for treating a wide variety of flexible-backbone docking problems.  相似文献   

17.
In this issue of Molecular Cell, Sampath et al. show a lysine methylase exhibits substrate promiscuity and variability in degree of product methylation (Sampath et al., 2007). Two lysines are found to be automethylated in G9a, and one is H3K9-like and can establish a docking site for HP1 chromodomain.  相似文献   

18.
The class C subset of vacuolar protein sorting (Vps) proteins (Vps11, Vps18, Vps16 and Vps33) assembles into a vacuole/prevacuole-associated complex. Here we demonstrate that the class C-Vps complex contains two additional proteins, Vps39 and Vps41. The COOH-terminal 148 amino acids of Vps39 direct its association with the class C-Vps complex by binding to Vps11. A previous study has shown that a large protein complex containing Vps39 and Vps41 functions as a downstream effector of the active, GTP-bound form of Ypt7, a rab GTPase required for the fusion of vesicular intermediates with the vacuole (Price, A., D. Seals, W. Wickner, and C. Ungermann. 2000. J. Cell Biol. 148:1231-1238). Here we present data that indicate that this complex also functions to stimulate nucleotide exchange on Ypt7. We show that Vps39 directly binds the GDP-bound and nucleotide-free forms of Ypt7 and that purified Vps39 stimulates nucleotide exchange on Ypt7. We propose that the class C-Vps complex both promotes Vps39-dependent nucleotide exchange on Ypt7 and, based on the work of Price et al., acts as a Ypt7 effector that tethers transport vesicles to the vacuole. Thus, the class C-Vps complex directs multiple reactions during the docking and fusion of vesicles with the vacuole, each of which contributes to the overall specificity and efficiency of this transport process.  相似文献   

19.
We investigate the extent to which the conformational fluctuations of proteins in solution reflect the conformational changes that they undergo when they form binary protein-protein complexes. To do this, we study a set of 41 proteins that form such complexes and whose three-dimensional structures are known, both bound in the complex and unbound. We carry out molecular dynamics simulations of each protein, starting from the unbound structure, and analyze the resulting conformational fluctuations in trajectories of 5 ns in length, comparing with the structure in the complex. It is found that fluctuations take some parts of the molecules into regions of conformational space close to the bound state (or give information about it), but at no point in the simulation does each protein as whole sample the complete bound state. Subsequent use of conformations from a clustered MD ensemble in rigid-body docking is nevertheless partially successful when compared to docking the unbound conformations, as long as the unbound conformations are themselves included with the MD conformations and the whole globally rescored. For one key example where sub-domain motion is present, a ribonuclease inhibitor, principal components analysis of the MD was applied and was also able to produce conformations for docking that gave enhanced results compared to the unbound. The most significant finding is that core interface residues show a tendency to be less mobile (by size of fluctuation or entropy) than the rest of the surface even when the other binding partner is absent, and conversely the peripheral interface residues are more mobile. This surprising result, consistent across up to 40 of the 41 proteins, suggests different roles for these regions in protein recognition and binding, and suggests ways that docking algorithms could be improved by treating these regions differently in the docking process.  相似文献   

20.
The 1H NMR spectrum of the tetradeoxynucleotide d(TpCpGpA) was examined as a function of temperature, pH, and concentration. At pH 7 and above the solution conformation for this oligodeoxynucleotide appears to be a mixture of random coil and Watson-Crick duplex. At 25 degrees C, a pH titration of d(TpCpGpA) shows that distinct conformational changes occur as the pH is lowered below 7.0. These conformational changes are reversible upon readjusting the pH to neutrality, indicating the presence of a pH-dependent set of conformational equilibria. At 25 degrees C, the various conformational states in the mixture are in rapid exchange on the NMR time scale. Examination of the titration curve shows the presence of distinct conformational states at pH greater than 7, and between pH 4 and pH 5. At pH less than 4, a third conformational state is present. When the pH titration is repeated at 5 degrees C, the conformational equilibria are in slow exchange on the NMR time scale; distinct signals from each conformational state are observable. The stable conformational state present between pH 4 and pH 5 represents an ordered conformation of d(TpCpGpA) which dissociates to a less ordered structure upon raising the temperature. This ordered conformation does not result from an intramolecular rearrangement, as is shown by by spectra obtained by varying oligodeoxynucleotide concentration at constant pH. The ordered conformation differs from the Watson-Crick helix, as is shown from nuclear Overhauser enhancement experiments, as well as chemical shift data. An ordered conformation for d(TpCpGpA) was previously reported [Reid, D. G., Salisbury, S. A., Brown, T., & Williams, D. H. (1985) Biochemistry 24, 4325-4332].(ABSTRACT TRUNCATED AT 250 WORDS)  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