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.
The cyclic 2,18-bridged biliverdin (2) is excreted in rat bile without reduction to the corresponding bilirubin. Conformational analysis, employing an optimized Monte Carlo method and a mixed Monte Carlo/stochastic dynamics, reveals that biliverdin IXalpha (1) and the cyclic analogue 2 adopt 'lock washer' conformations, stabilized by the presence of intramolecular hydrogen bonds between N23...H22N and, to a lesser extent, between N23...H24N. Although 2 is very similar in overall shape to 1, the former adopts a 'locked lock washer' conformation unable to undergo fluctuations, thus possibly hampering a proper recognition by biliverdin reductase.  相似文献   

2.
We have developed a computational method of protein design to detect amino acid sequences that are adaptable to given main-chain coordinates of a protein. In this method, the selection of amino acid types employs a Metropolis Monte Carlo method with a scoring function in conjunction with the approximation of free energies computed from 3D structures. To compute the scoring function, a side-chain prediction using another Metropolis Monte Carlo method was performed to select structurally suitable side-chain conformations from a side-chain library. In total, two layers of Monte Carlo procedures were performed, first to select amino acid types (1st layer Monte Carlo) and then to predict side-chain conformations (2nd layers Monte Carlo). We applied this method to sequence design for the entire sequence on the SH3 domain, Protein G, and BPTI. The predicted sequences were similar to those of the wild-type proteins. We compared the results of the predictions with and without the 2nd layer Monte Carlo method. The results revealed that the two-layer Monte Carlo method produced better sequence similarity to the wild-type proteins than the one-layer method. Finally, we applied this method to neuraminidase of influenza virus. The results were consistent with the sequences identified from the isolated viruses.  相似文献   

3.
N.m.r., i.r. and optical spectra of model compounds were recorded. These were to help in elucidating the structures of the phenylazo derivatives of bilirubin conjugates isolated from human bile. Model compounds included commercial and human bile bilirubin, mesobilirubin, bilirubin dimethyl ester, dimethoxybilirubin dimethyl ester and the corresponding phenylazo derivatives. The phenylazo derivative of vinylneoxanthobilirubinic acid was also investigated. All compounds were of the type IXα, and no other isomer could be detected with the spectroscopic methods employed. The compounds crystallize as the lactams, except for dimethoxybilirubin dimethyl ester and its phenylazo derivative, which are held in the lactim ether configuration. With all other compounds no tautomeric forms other than the lactams could be detected, although small proportions of bilirubin must exist as the lactim. Bilirubin does not form a betaine, a structure that has been proposed by von Dobeneck & Brunner (1965) to explain the bathochromic shift of its optical spectrum as compared with the expected position of the absorption maximum at 420nm. However, this shift to 453nm can be explained on the basis of internal hydrogen bonds occurring between the carboxylic protons and the pyrrole rings of bilirubin, as proposed by Fog & Jellum (1963), and new evidence for such a bonding has been accumulated. The bilirubin sulphate described by Watson (1958), which is formed by treatment of bilirubin with concentrated sulphuric acid and acetic anhydride, was also investigated. The main product of this reaction was isolated as its phenylazo derivative, and was shown to be 3,18-di(ethylidene sulphate)-2,7,13,17-tetramethylbiladiene-ac-8,12-dipropionic acid. The reaction leading to this compound is an addition of sulphuric acid to the vinyl side chains of bilirubin according to Markownikoff's rule.  相似文献   

4.
To speed up dose calculation, an analytical pencil-beam method has been developed to calculate the mean radial dose distributions due to secondary electrons that are set in motion by light ions in water. For comparison, radial dose profiles calculated using a Monte Carlo technique have also been determined. An accurate comparison of the resulting radial dose profiles of the Bragg peak for (1)H(+), (4)He(2+) and (6)Li(3+) ions has been performed. The double differential cross sections for secondary electron production were calculated using the continuous distorted wave-eikonal initial state method (CDW-EIS). For the secondary electrons that are generated, the radial dose distribution for the analytical case is based on the generalized Gaussian pencil-beam method and the central axis depth-dose distributions are calculated using the Monte Carlo code PENELOPE. In the Monte Carlo case, the PENELOPE code was used to calculate the whole radial dose profile based on CDW data. The present pencil-beam and Monte Carlo calculations agree well at all radii. A radial dose profile that is shallower at small radii and steeper at large radii than the conventional 1/r(2) is clearly seen with both the Monte Carlo and pencil-beam methods. As expected, since the projectile velocities are the same, the dose profiles of Bragg-peak ions of 0.5 MeV (1)H(+), 2 MeV (4)He(2+) and 3 MeV (6)Li(3+) are almost the same, with about 30% more delta electrons in the sub keV range from (4)He(2+)and (6)Li(3+) compared to (1)H(+). A similar behavior is also seen for 1 MeV (1)H(+), 4 MeV (4)He(2+) and 6 MeV (6)Li(3+), all classically expected to have the same secondary electron cross sections. The results are promising and indicate a fast and accurate way of calculating the mean radial dose profile.  相似文献   

5.
We propose a novel Metropolis Monte Carlo procedure for protein modeling and analyze the influence of hydrogen bonding on the distribution of polyalanine conformations. We use an atomistic model of the polyalanine chain with rigid and planar polypeptide bonds, and elastic alpha carbon valence geometry. We adopt a simplified energy function in which only hard-sphere repulsion and hydrogen bonding interactions between the atoms are considered. Our Metropolis Monte Carlo procedure utilizes local crankshaft moves and is combined with parallel tempering to exhaustively sample the conformations of 16-mer polyalanine. We confirm that Flory's isolated-pair hypothesis (the steric independence between the dihedral angles of individual amino acids) does not hold true in long polypeptide chains. In addition to 3(10)- and alpha-helices, we identify a kink stabilized by 2 hydrogen bonds with a shared acceptor as a common structural motif. Varying the strength of hydrogen bonds, we induce the helix-coil transition in the model polypeptide chain. We compare the propensities for various hydrogen bonding patterns and determine the degree of cooperativity of hydrogen bond formation in terms of the Hill coefficient. The observed helix-coil transition is also quantified according to Zimm-Bragg theory.  相似文献   

6.
Abstract

This paper continues our Monte Carlo simulation study of liquid hydrogen chloride [1]. The importance of non-additive interactions is carefully analyzed. Computed atom pair correlation functions are compared to neutron scattering experiments [2]. A difference algorithm (“Δ—algorithm”) is developed, which makes non-additive Monte Carlo simulations practicable. We also report an implementation of this algorithm on a transputer network, taking advantage of the inherent parallelism of the Δ — algorithm.  相似文献   

7.
In order to investigate the Conformational change of the α-aminoisobutyric acid (Aib) containing peptide by the D /L replacement of an amino acid residue, single crystals of two diastereomers, Dnp-L -Val-Aib-Gly-L -Leu-pNA (L -L isomer) and Dnp-D -Val-Aib-Gly-L -Leu-pNA (D -L isomer), were prepared from aqueous methanol solutions as CH3OH and CH3OH · H2O solvates, respectively, and were analyzed by the x-ray diffraction method. Molecular conformation of L -L isomer adopts consecutive two different types of β-turns, a type II′ β-turn bent at Aib-Gly, and a type III β-turn bent at Gly-Leu, stabilized by two intramolecular (Leu) NH …? O?C (Val) and (pNA) NH …? O?C(Aib) hydrogen bonds. In contrast, these two intramolecular hydrogen bonds lead the D -L isomer to a distorted 310-helix conformation consisting of consecutive two type-III β-turn of Aib-Gly-Leu sequence. The most significant structural difference between these diastereomers is the mutual orientation between the Dnp and pNA chromophores. While the extensive stacking of both the chromophores is intramolecularly formed for the folded conformation of L -L isomer, they are oriented toward an opposite direction in the open conformation of D -L isomer and are intermolecularly stacked with each other. The large separation between these diastereomers observed in the chromatography is discussed in the relation with their Conformational differences. © 1993 John Wiley & Sons, Inc.  相似文献   

8.
Alvarez M  Jover A  Carrazana J  Meijide F  Soto VH  Tato JV 《Steroids》2007,72(6-7):535-544
The crystal structures of chenodeoxycholic acid (CDCA), ursodeoxycholic acid (7beta isomer of CDCA) and their other two epimers (3beta,7alpha- and 3beta,7beta-isomers) have been resolved. The four isomers were recrystallized from p-xylene. CDCA crystal is hexagonal P6(5) while the crystals of the other three isomers are orthorhombic (P2(1)2(1)2(1) space group). Only the 3beta,7beta isomer forms an inclusion complex with the solvent with a 1:1 stoichiometry. In all cases, the three hydrogen bond sites (the two hydroxy groups, O3-H and O7-H, and the carboxylic acid group of the side chain, O24bO24a-H) simultaneously act as hydrogen bond donors and acceptors. By considering that O24a is always donor and O24b is always acceptor, the hydrogen bond sequences can be understood on the basis of the interaction between the two hydroxy groups. However the comparison between the four compounds is complicated by the existence of two molecules in the asymmetric unit in the UDCA crystal resulting in that the same hydrogen bond site (for instance O3) can be donor towards two different acceptors (either O7 or O24b). As in the case of the four isomers of deoxycholic acid (Steroids 2004, 69, 379), the other three isomers present a donor-->acceptor sequence, which is O7-->O3 when O3-H is beta and O3-->O7 when O3-H is alpha. The spatial orientation of the carboxylic acid of the side chain is referred to two almost perpendicular planes (defined by (1) the carbon atoms C1/C6-C17/C20 and by (2) the methyl groups C18-C19 and the two carbon atoms to which they are linked, C10 and C13, respectively). Only the side chain of CDCA evidences a positive deviation towards the hydrophobic beta side of the molecule.  相似文献   

9.
We propose an algorithmic strategy for improving the efficiency of Monte Carlo searches for the low-energy states of proteins. Our strategy is motivated by a model of how proteins alter their shapes. In our model, when proteins fold under physiological conditions, their backbone dihedral angles change synchronously in groups of four or more to avoid steric clashes and respect the kinematic conservation laws. They wriggle; they do not thrash. We describe a simple algorithm that can be used to incorporate wriggling in Monte Carlo simulations of protein folding. We have tested this wriggling algorithm against a code in which the dihedral angles are varied independently (thrashing). Our standard of success is the average root-mean-square distance (rmsd) between the alpha-carbons of the folding protein and those of its native structure. After 100,000 Monte Carlo sweeps, the relative decrease in the mean rmsd, as one switches from thrashing to wriggling, rises from 11% for the protein 3LZM with 164 amino acids (aa) to 40% for the protein 1A1S with 313 aa and 47% for the protein 16PK with 415 aa. These results suggest that wriggling is useful and that its utility increases with the size of the protein. One may implement wriggling on a parallel computer or a computer farm.  相似文献   

10.
A range of N-glycosyl-thiophene-2-carboxamides, including a 6H-thieno[2,3-c]pyridin-7-one and a bivalent compound, have been synthesised and assayed for their effects on DNA synthesis in bovine aortic endothelial cells or on the growth of synoviocytes. Per-O-acetylated analogues of the glycoconjugates were significantly more effective inhibitors when compared to their corresponding non-acetylated analogues, indicating that the lower potency observed for hydroxylated derivatives is due to less efficient transport of these compounds across the cell membrane. Thiophene-2-carboxamide was inactive as an inhibitor of bFGF induced proliferation, confirming the requirement of the carbohydrate residue for the observed biological properties. Glucose, mannose, galactose and 2-amino-2-deoxy-glucose analogues were active as were a variety of substituted thiophene derivatives; the 6H-thieno[2,3-c]pyridin-7-one conjugate was inactive. Conformational analysis of the title compounds was investigated. X-ray crystal structural analysis of four N-glucosyl-thiophene-2-carboxamides showed that the pyranose rings adopted the expected 4C1 conformations and that Z-anti structures were predominant (H1-C1-N-H anomeric torsion angle varied from -168.2 degrees to -175.0 degrees ) and that the carbonyl oxygen and sulfur of the thiophene adopted an s-cis conformation in three of the isomers. In a crystal structure of a 3-alkynyl derivative, the hydrogen atom of the NH group was directed toward the acetylene group. The distance between the hydrogen atom and acetylene carbons and angles between nitrogen, hydrogen and carbon atoms were consistent with hydrogen bonding and this was supported by IR and NMR spectroscopic studies. The geometries of thiophene-2-carboxamides were explored by density functional theory (DFT) and M?ller-Plesset (MP2) calculations and the s-cis conformer of thiophene-2-carboxamide was found to be more stable than its s-trans isomer by 0.83 kcal mol(-1). The s-cis conformer of 3-ethynyl-thiophene-2-carboxamide was 5.32 kcal mol(-1) more stable than the s-trans isomer. The larger stabilisation for the s-cis conformer in the 3-alkynyl derivatives is explained to be due to a moderate hydrogen bonding interaction between the alkyne and NH group.  相似文献   

11.
GABA(A) receptors are ligand-gated chloride channels composed of five homologous subunits that specifically recognize one another and assemble around an aqueous pore. To identify domains responsible for the specificity of subunit association, we constructed C-terminal truncated gamma(2) subunits, as well as mutated and chimeric fragments. From their ability to interfere with alpha(1)beta(3)gamma(2) receptor assembly and to associate with full-length subunits, we concluded that amino acid sequences gamma(2)-(91-104) and gamma(2)-(83-90) form the sites mediating assembly with alpha(1) and beta(3) subunits, respectively. Neural network-based secondary structure prediction, Monte Carlo optimization, and hydrophobicity analysis led to the conclusion that these sites also form the intersubunit contacts in the completely assembled receptor and provided important information on the benzodiazepine-binding site and structure of GABA(A) receptors.  相似文献   

12.
The structures of the complexes with alpha-lytic protease of both phosphorus stereoisomers of N-[(2S)-2-[[[(1R)-1-[N-[(tert-butyloxycarbonyl)-L-alanyl-L-alanyl- L-prolyl]amino]-2-methylpropyl]-phenoxyphosphinyl]oxy]propanoyl]- L-alanine methyl ester, an analogue of the peptide Boc-Ala-Ala-Pro-Val-Ala-Ala where Val is replaced with an analogous phosphonate phenyl ester and the subsequent Ala is replaced with lactate, have been determined to high resolution (1.9 A) by X-ray crystallography. Both stereoisomers inactivate the enzyme but differ by a factor of 2 in the second-order rate constant for inactivation [Sampson, N. S., & Bartlett, P. A. (1991) Biochemistry (preceding paper in this issue)]. One isomer (B) forms a tetrahedral adduct in which the phosphonate phenyl ester is displaced by the active site serine (S195) and interacts with the enzyme across seven substrate recognition sites that span both sides of the scissile bond. Seven hydrogen bonds are formed with the enzyme, and 510 A2 of hydrophobic surface area is buried when the inhibitor interacts with the enzyme. Although two hydrogen bonds are gained by incorporation of two residues on the C-terminal side of the scissile bond into the inhibitor, there is very little adjustment in the structure of the enzyme in this region. Surprisingly, the active site histidine (H57) does not interact with the phosphonate, apparently because the phosphonate lacks negative charge in or near the oxyanion hole, and instead, the side chain rotates out of the active site cleft and hydrogen bonds with solvent. The other isomer (A) forms a mixture of two different tetrahedral adducts in the active site, both covalently bonded to Ser 195. One adduct, at approximately 58% occupancy, is exactly the same in structure as the complex formed with isomer B, and the other adduct, at 42% occupancy, has lost the two residues C-terminal to the scissile bond by hydrolysis. In the lower occupancy structure, His 57 does not rotate out of the active site and forms a hydrogen bond with the phosphonate oxygen instead. The structures of both complexes were insensitive to pH. As very little change in structure accompanies the histidine rotation, the complex with isomer B provides an excellent mimic for the structure of the transition state (or high-energy reaction intermediate) that spans both sides of the scissile bond.  相似文献   

13.
The specificity of horse liver alcohol dehydrogenase for cyclohexanol and its 3-methyl derivatives was investigated by stopped-flow and initial velocity kinetic studies. The (1S,3S)-3-methylcyclohexanol was 7 times more reactive (V/Km) than cyclohexanol, whereas the (1R,3R)-3-methylcyclohexanol was at least 1000 times less reactive than its enantiomer. Computer simulation of the transient reaction of NAD+ and the cyclohexanols catalyzed by the enzyme suggests that the rate of transfer of hydrogen from the alcohol to NAD+ is increased with the 1S,3S isomer. Modeling of the three-dimensional structure of the ternary complex of the enzyme suggests that the 1S,3S isomer should only bind in a productive, reactive mode, whereas the 1R,3R isomer would bind predominantly in a nonproductive, inhibitory mode.  相似文献   

14.
Zhan L  Chen JZ  Liu WK 《Proteins》2007,66(2):436-443
Using a newly developed Monte Carlo global optimization method called basin paving, we have performed an ab initio computation for the structure of Trp-cage based on the ECEPP/3 force field in vacuo. The lowest energy minimum has been located. Its corresponding configuration is comparable to the native structure of Trp-cage (PDB code 1L2Y) with a backbone root mean square deviation of 2.24 A.  相似文献   

15.
The substrate specificity of the different molecular forms of biliverdin reductase (bilirubin:NAD(P)+ oxidoreductase, EC 1.3.1.24) using biliverdin XIII alpha, XIII beta and XIII gamma was examined. It was found that molecular form 1 (the major form in normal rat liver) reduced biliverdin XIII alpha at a much higher rate than the other two isomers. Molecular form 2 (the minor form) reduced isomers XIII alpha and XIII beta at similar rates, while molecular form 3 (the major form induced by CoCl2 treatment) reduced the XIII beta isomer at a slightly higher rate than the XIII alpha isomer. Molecular forms 2 and 3, both reduced isomer XIII gamma more slowly than they reduced the XIII alpha and XIII beta isomers. These results are similar to those obtained previously using biliverdins IX alpha, IX beta and IX gamma, suggesting that biliverdin reductase specificity is related to the type of the isomer rather than to the series (IX or XIII) of the isomer.  相似文献   

16.
Energy-optimized structure of antifreeze protein and its binding mechanism.   总被引:7,自引:0,他引:7  
A combination of Monte Carlo simulated annealing and energy minimization was utilized to determine the conformation of the antifreeze protein from the fish winter flounder. It was found from the energy-optimized structure that the hydroxyl groups of its four threonine residues, i.e. Thr2, Thr13, Thr24, Thr35, are aligned on almost the same line parallel to the helix axis and separated successively by 16.1, 16.0 and 16.2 A, respectively, very close to the 16.6 A repeat spacing along [0112] in ice. Based on such a space match, a zipper-like model is proposed to elucidate the binding mechanism of the antifreeze protein to ice crystals. According to the current model, the antifreeze protein may bind to an ice nucleation structure in a zipper-like fashion through hydrogen bonding of the hydroxyl groups of these four Thr residues to the oxygen atoms along the [0112] direction in ice lattice, subsequently stopping or retarding the growth of ice pyramidal planes so as to depress the freeze point. The calculated results and the binding mechanism thus derived accord with recent experimental observations. The mechanistic implications derived from such a special antifreeze molecule might be generally applied to elucidate the structure-function relationship of other antifreeze proteins with the following two common features: (1) recurrence of a Thr residue (or any other polar amino acid residue whose side-chain can form a hydrogen bond with water) in an 11-amino-acid period along the sequence concerned; and (2) a high percentage of Ala residue component therein. Further experiments are suggested to test the ice binding model.  相似文献   

17.
The interaction between the nucleic acid bases and solvent molecules has an important effect in various biochemical processes. We have calculated total energy and free energy of the solvation of DNA bases in water by Monte Carlo simulation. Adenine, guanine, cytosine, and thymine were first optimized in the gas phase and then placed in a cubic box of water. We have used the TIP3 model for water and OPLS for the nucleic acid bases. The canonical (T, V, N) ensemble at 25 degrees C and Metropolis sampling technique have been used. Good agreement with other available computational data was obtained. Radial distribution functions of water around each site of adenine, guanine, cytosine, and thymine have been computed and the results have shown the ability of the sites for hydrogen bonding and other interactions. The computations have shown that guanine has the highest value of solvation free energy and N7 and N6 in adenine and guanine, N3 in cytosine, and N3 and O4 in thymine have the largest radial distribution function. Monte Carlo simulation has also been performed using the CHARMM program under the same conditions, and the results of two procedures are compared.  相似文献   

18.
The interaction between the nucleic acid bases and solvent molecules has an important effect in various biochemical processes. We have calculated total energy and free energy of the solvation of DNA bases in water by Monte Carlo simulation. Adenine, guanine, cytosine, and thymine were first optimized in the gas phase and then placed in a cubic box of water. We have used the TIP3 model for water and OPLS for the nucleic acid bases. The canonical (T, V, N) ensemble at 25°C and Metropolis sampling technique have been used. Good agreement with other available computational data was obtained. Radial distribution functions of water around each site of adenine, guanine, cytosine, and thymine have been computed and the results have shown the ability of the sites for hydrogen bonding and other interactions. The computations have shown that guanine has the highest value of solvation free energy and N7 and N6 in adenine and guanine, N3 in cytosine, and N3 and O4 in thymine have the largest radial distribution function. Monte Carlo simulation has also been performed using the CHARMM program under the same conditions, and the results of two procedures are compared.  相似文献   

19.
Modeling of protein loops by simulated annealing.   总被引:6,自引:5,他引:1       下载免费PDF全文
A method is presented to model loops of protein to be used in homology modeling of proteins. This method employs the ESAP program of Higo et al. (Higo, J., Collura, V., & Garnier, J., 1992, Biopolymers 32, 33-43) and is based on a fast Monte Carlo simulation and a simulated annealing algorithm. The method is tested on different loops or peptide segments from immunoglobulin, bovine pancreatic trypsin inhibitor, and bovine trypsin. The predicted structure is obtained from the ensemble average of the coordinates of the Monte Carlo simulation at 300 K, which exhibits the lowest internal energy. The starting conformation of the loop prior to modeling is chosen to be completely extended, and a closing harmonic potential is applied to N, CA, C, and O atoms of the terminal residues. A rigid geometry potential of Robson and Platt (1986, J. Mol. Biol. 188, 259-281) with a united atom representation is used. This we demonstrate to yield a loop structure with good hydrogen bonding and torsion angles in the allowed regions of the Ramachandran map. The average accuracy of the modeling evaluated on the eight modeled loops is 1 A root mean square deviation (rmsd) for the backbone atoms and 2.3 A rmsd for all heavy atoms.  相似文献   

20.
Aspartic acid protease model peptides Z-Phe-Asp(COOH)-Thr-Gly-Ser-Ala-NHCy (1) and AdCO-Asp(COOH)-Val-Gly-NHBzl (3), and their aspartate anions (NEt4)[Z-Phe-Asp(COO-)-Thr-Gly-Ser-Ala-NHCy] (2) and (NEt4)[AdCO-Asp(COO-)-Val-Gly-NHBzl] (4), having an invariant primary sequence of the Asp-X(Thr,Ser)-Gly fragment, were synthesized and characterized by 1H-NMR, CD, and infrared (IR) spectroscopies. NMR structure analyses indicate that the Asp O(delta) atoms of the aspartate peptide 2 are intramolecularly hydrogen-bonded with Gly, Ser, Ala NH, and Ser OH, supporting the rigid beta-turn-like conformation in acetonitrile solution. The tripeptide in the aspartic acid 3 forms an inverse gamma-turn structure, which is converted to a beta-turn-like conformation because of the formation of the intramolecular NH . . . O- hydrogen bonds with the Asp O(delta) in 4. Such a conformational change is not detected between dipeptides AdCO-Asp(COOH)-Va-NHAd (5) and (NEt4)[AdCO-Asp(COO-)-Val-NHAd] (6). The pK(a) value of side-chain carboxylic acid (5.0) for 3 exhibits a lower shift (0.3 unit) from that of 5 in aqueous polyethyleneglycol lauryl ether micellar solution. NMR structure analyses for 3 in an aqueous micellar solution indicate that the preorganized turn structure, which readily forms the NH . . . O- hydrogen bonds, lowers the pK(a) value and that resulting hydrogen bonds stabilize the rigid conformation in the aspartate anion state. We found that the formation of the NH . . . O- hydrogen bonds involved in the hairpin turn is correlated with the protonation and deprotonation state of the Asp side chain in the conserved amino acid fragments.  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