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 20 毫秒
1.
R. Chauvin  H. B. Kagan 《Chirality》1991,3(4):242-253
With the early aim of quantifying steric consequences of chirality, efforts to define a nonempirical steric parameter of chemical groups are reported. Steric hindrance of a reacting center by any acyclic saturated R group has been characterized by a geometric “axial steric parameter”: the solid angle of R. When the group is a “symmetric top substituent” (i.e., when all the terminal atoms are equivalent), the solid angle matches the solid angle of a cone envelope of R. The definition of this cone is compared with Tolman's definition of a ligand cone in organometallic complexes. The chemical significance of this parameter is shown by an excellent correlation with the Dubois' experimental steric parameter Es. Modeling steric repulsion by the cone of R, and correcting solid angles for conformational effects, only 3 empirical coefficients are needed to calculate 33 values of Es with less than 10% error. The cone model is suggested to be relevant within the limits of random and independent free rotations about all the bonds in the C–R group. A separation between “axial cone steric hindrance” and other steric effects is proposed. The basic model and the corrections proposed allow the conformational features of esters to be discussed.  相似文献   

2.
The mode of binding of the irreversible thermolysin inhibitor ClCH2CO-DL-(N-OH)Leu-OCH3 [Rasnick, D., & Powers, J.C. (1978) Biochemistry 17, 4363-4369] has been determined by X-ray crystallography at a resolution of 2.3 A and the structure of the covalent complex refined to give a crystallographic residual of 17.0%. This is the first such structural study of an active-site-directed covalent complex of a zinc protease. As anticipated by Rasnick and Powers, the inhibitor alkylates Glu-143 in the thermolysin active site, and the hydroxamic acid moiety coordinates the zinc ion. The formation of the covalent complex is associated with a significant shift in a segment of the polypeptide backbone in the vicinity of the active site. This conformational adjustment appears to be necessary to relieve steric hindrance which would otherwise prevent alkylation of Glu-143. It is suggested that this steric hindrance, which occurs for thermolysin but would not be expected for carboxypeptidase A, accounts for the previously inexplicable difference in reactivity of these two metalloproteases toward N-haloacetyl amino acids. The relevance of this steric hindrance to the mechanism of catalysis is discussed. In agreement with previous results [Kester, W. R., & Matthews, B. W. (1977) Biochemistry 16, 2506-2516], it appears that steric hindrance prevents the direct attack of Glu-143 on the carbonyl carbon of an extended substrate, therefore ruling out the anhydride pathway in thermolysin-catalyzed hydrolysis of polypeptide substrates and their ester analogues.  相似文献   

3.
The non-covalent interaction between a series of N-phosphoryl dipeptides (or methyl esters) (DPP) and protein was studied by ESI-MS and UV–vis spectrometer. The function of different groups in DPP and binding sites of protein were investigated. The results revealed that hydroxyl and aromatic ring in DPP were both important group for the interaction, and aromatic ring had double functions on the interaction. In addition, the molecular size, flexibility and steric hindrance showed obvious effects on the interaction, while, the chirality, sequence and length of carbon chains (changing 1–2C) of amino acid residue in DPP showed little effects on the interaction under the experimental conditions. Phosphoryl oligopeptides having extended structure, good molecular flexibility and smaller spatial hindrance could contract the protein conformation in solution. The aromatic, basic, acid and amide amino acid residues of protein may be the main binding sites and contributed to the survival of the complexes.  相似文献   

4.
Inhibition of tryptophanase-catalyzed decomposition of S-(o-nitrophenyl)-L-cysteine by a variety of amino acids has been investigated. For amino acids similar to the natural substrate and for those having minimal steric requirements for the side chain, the linear correlation exists between-RTlnKi and side chain hydrophobicity. L-ornithine and L-arginine are anomalously potent inhibitors taking into account low hydrophobicity of their side chains. This can be explained by an interaction between a positively charged group of the side chain of L-arginine or L-ornithine and a nucleophilic group of the active site. The comparison of affinity of tryptophanase for L-phenylalanine and L-homophenylalanine indicates that there is a special locus in the active site where aromatic groups are bound and oriented approximately parallel to the cofactor plane experiencing no steric hindrance. For a large number of amino acids the rates of the enzymic alpha-proton exchange in 2H2O are comparable with the rate of the reaction with L-tryptophan. Very low rate of alpha-proton exchange observed with L-alanine is an exception.  相似文献   

5.
Chen S 《Amino acids》2004,26(3):291-298
Summary. A variety of -amino acids are enantioresolved for the first time on naphthylethylcarbamate--cyclodextrin bonded phases (i.e., SN- and RN--CD) using the acetonitrile-based mobile phase after their pre-column derivatization with phenyl isothiocyanate in alkaline medium. The resolution is better obtained on RN--CD phase and fails to reproduce if the amino acid is N-benzoylated or N-carbobenzyloxylated under the same chromatographic conditions. The enhanced resolution is believed to be due to the re-location of the hydrogen receptor site from sulfur to nitrogen on the isothiocyanyl fragment of derivatizing reagent, which in turn changes the enantioselectivity. Also, the sulfur atom is larger in size and subject to steric hindrance more significantly in comparison with oxygen. The carboxyl group of amino acid is essential toward a satisfactory resolution. The position of the amino group on the backbone affects the resolution as well. Finally, the resolution is either not observed or unsatisfactory in the reversed- or normal phase mode for most of the amino acids examined in this study.  相似文献   

6.
To estimate the steric distance between the bitter taste determinant sites in peptides, some cyclic dipeptides, amino acid anilides, amino acid cyclohexylamides, and benzoyl amino acids were synthesized and their tastes were evaluated. The diketopiperazine ring of cyclic dipeptides acted as a bitter taste determinant site due to its hydrophobicity. The steric distance between 2 sites was estimated as 4.1 Å from the molecule models of cyclic dipeptides composed of typical amino acids in the bitter peptides. Due to the hypothesis of two bitter taste determinant sites, which bind with the bitter taste receptor via a “binding unit” and a “stimulating unit,” a mechanism for the bitterness in peptides was postulated.  相似文献   

7.
Summary A direct synthesis of a series ofN-SES amino acids is described.N-SES Ala has been further utilized in the synthesis of a perhydro-1,4-diazepin-2-one.  相似文献   

8.
A new and convenient stereocontrolled synthesis of the optically pure (S)-α-methyl,α-amino acids 6(ad) that exploits the chiral synthon 1,4-N,N-[(S)-1-phenylethyl]-piperazine-2,5-dione (1) is described. The (S)-1-phenylethyl group, bonded to each of the N-atoms of the 2,5-diketopiperazine, acts as a chiral inductor in the first alkylation, while the steric hindrance appears to be the determining factor of stereocontrol in third and forth alkylation.  相似文献   

9.
The population of a strain ofEscherichia coli that was resistant to nalidixic acid and streptomycin declined rapidly in samples of sterile and nonsterile Cayuga Lake water and reached an undetectable level in nonsterile water at 24 and 72 hours when counted on eosin-methylene blue (EMB) agar and half-strength trypticase soy agar (TSA), respectively. In sterile lake water amended with 10g amino acids per ml or 0.1 M phosphate,E. coli multiplied exponentially for more than 24 hours. The addition ofRhizobium leguminosarum biovarphaseoli to unamended sterile lake water prevented the decline ofE. coli, and its addition to amended sterile lake water preventedE. coli multiplication. The cell density of this strain ofE. coli declined in the first 8 hours after its introduction into an inorganic salts solution, but the bacterium then grew extensively. This increase in abundance was not observed in the presence ofR. phaseoli, andE. coli counts on half-strength TSA remained unchanged between 8 hours and 6 days. When counted on EMB agar, the abundance of the antibiotic-resistant strain ofE. coli and a strain not selected for resistance increased in solutions containing phosphate and amino acids but declined in the presence of high densities ofR. phaseoli. Many of the cells of the antibiotic-resistantE. coli strain failed to grow on antibiotic-amended EMB agar after introduction of the organism into nonsterile or sterile lake water or into an inorganic salts solution containingR. phaseoli, although colonies appeared on TSA. The data suggest thatE. coli cells grown on rich media suffer a shock when introduced into lake water because of low hypotonicity, the indigenous competing flora, or both. This shock is prevented by either phosphate buffer or by amino acids at low concentration. The shocked bacteria formed colonies on half-strength TSA. Depending on environmental conditions, the presence of a second organism either has no effect or results in an increase or decrease inE. coli numbers.  相似文献   

10.
Stig Allenmark 《Chirality》1993,5(5):295-299
In order to get further insight into the mechanism by which bovine serum albumin (BSA) discriminates between enantiomers of organic acids, some radioisotopically labeled, nonaromatic carboxylic acids were studied under varying mobile phase conditions. It was found for a series of N-alkanoyl-DL -[3H]leucines that the D -enantiomers were much more strongly retained and that the composition of the mobile phase could be adjusted to give very large (α > 20) enantiomeric separation factors. The elution order was consistent with what has been found earlier for other N-acyl derivatives as well as for N-arylcarbamoyl derivatives of simple aliphatic amino acids. A marked increase in the hydrophobic interaction of the D -enantiomers with the chiral phase was found upon a lowering of the mobile phase strength, conditions under which the L -form was only slightly influenced. These and other results are consistent with a chiral recognition model by which inclusion of the compound in a hydrophobic chiral cavity of BSA with simultaneous charge interaction is assumed to take place and whereby discrimination is determined by the steric bulk and orientation of the α-substituent. © 1993 Wiley-Liss, Inc.  相似文献   

11.
To test the hypothesis that substrate-induced steric compression between His 57 and Asp 102 at the active site of chymotrypsin can increase the basicity of His 57, we have synthesized thecis- andtrans-isomers of 2-bromo-3-(N-tritylimidazole)-2-propenoic acid and 2-chloro-3-(N-tritylimidazole)-2-propenoic acid and compared selected properties with those ofcis-andtrans-urocanic acids. Thecis-isomers display low field1H NMR signals at 17 ppm in dimethylsulfoxide, similar tocis-urocanic acid; whereas thetrans-isomers do not show strong hydrogen bonds. Increasing the size of the C2 substituent (H < Cl < Br) in thecis-isomers increases the pKaof the imidazolium group from 6.78 for H to 7.81 and 9.10 for Cl and Br, respectively; whereas the pKas of thetransisomers are all 6.0 ± 0.1. The results indicate that thecis-urocanic acid derivatives with large substituents at C2 act as proton sponges in water, and they support the concept that steric compression in the catalytic triad of chymotrypsin can increase the basicity of His 57.  相似文献   

12.
A chitinase gene (pCHi58) encoding a 58 kDa chitinase was isolated from theSerratia marcescens KCTC 2172 cosmid library. The chitinase gene consisted of a 1686 bp open reading frame that encoded 562 amino acids.Escherichia coil harboring the pChi58 gene secreted a 58 kDa chitinase into the culture supernatant. The 58 kDa chitinase was purified using a chitin affinity column and mono-S column. A nucleotide andN-terminal amino acid sequence analysis showed that the 58 kDa chitinase had a leader peptide consisting of 23 amino acids which was cleaved prior to the 24th alanine. The 58 KDa chitinase exhibited a 98% similarity to that ofS. marcescens QMB 1466 in its nuclotide sequence. The chitinolytic patterns of the 58 kDa chitinase released N,N′-diacetyl chitobiose (NAG2) as the major hydrolysis end-product with a trace amount ofN-acetylglucosamine. When a 4-methylumbellyferyl-N-acetylglucosamin monomer, dimmer, and tetramer were used as substrates, the 58 kDa chitinase did not digest the 4-Mu-NAG monomer (analogue of NAG2), thereby indicating that the 58 kDa chitinase was likely an endochitinase. The optimum reaction temperature and pH of the enzyme were 50°C and 5.0, respectively.  相似文献   

13.
ATP-sensitive K+ (KATP) channels are gated by intracellular ATP, proton and phospholipids. The pore-forming Kir6.2 subunit has all essential machineries for channel gating by these ligands. It is known that channel gating involves the inner helix bundle of crossing in which a phenylalanine residue (Phe168) is found in the TM2 at the narrowest region of the ion-conduction pathway in the Kir6.2. Here we present evidence that Phe168-Kir6.2 functions as an ATP- and proton-activated gate via steric hindrance and hydrophobic interactions. Site-specific mutations of Phe168 to a small amino acid resulted in losses of the ATP- and proton-dependent gating, whereas the channel gating was well maintained after mutation to a bulky tryptophan, supporting the steric hindrance effect. The steric hindrance effect, though necessary, was insufficient for the gating, as mutating Phe168 to a bulky hydrophilic residue severely compromised the channel gating. Single-channel kinetics of the F168W mutant resembled the wild-type channel. Small residues increased Popen, and displayed long-lasting closures and long-lasting openings. Kinetic modeling showed that these resulted from stabilization of the channel to open and long-lived closed states, suggesting that a bulky and hydrophobic residue may lower the energy barrier for the switch between channel openings and closures. Thus, it is likely that the Phe168 acts as not only a steric hindrance gate but also potentially a facilitator of gating transitions in the Kir6.2 channel.  相似文献   

14.
A series of model peptides containing α-trifluoromethyl-substituted amino acids in five different positions relative to the predominant cleavage site of the serine protease α-chymotrypsin was synthesized by solution methods to investigate the influence of α-Tfm substitution on the proteolytic stability of peptides. Proteolysis studies demonstrated absolute stability of peptides substituted in the P1 position and still considerable proteolytic stability for peptides substituted at the P2 and P′2 positions compared with the corresponding unsubstituted model peptide. Comparison with peptides containing the fluorine-free disubstituted amino acid α-aminoisobutyric acid allowed to separate electronic from steric effects. Furthermore, the absolute configuration of the α-Tfm-substituted amino acid was found to exert considerable effects on the proteolytic stability, especially in P′1 substituted peptides. Investigations of this phenomenon using empirical force field calculations revealed that in the (S,R,S)-diasteromer the steric constraints exhibited by the α-Tfm group can be outweighed by an advantageous interaction of the fluorine atoms with the serine side chain of the enzyme. In contrast, a favourable interaction between substrate and enzyme is impossible for the (S,S,S)-diastereomer. © 1997 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
A. Hutinec  A. Ziogas  A. Rieker 《Amino acids》1996,11(3-4):345-366
Summary Several non-natural phenolic amino acids have been synthesized.t-Butylated tyrosine and thyroxine derivatives, on one-electron oxidation, give persistent radicals which can be used as positional and/or spin labels for amino acids. Two-electron oxidation ofN-protected tyrosines leads to spirolactones, useful active esters for peptide coupling.  相似文献   

16.
Conformational constraints of amino acid side chains in alpha-helices   总被引:3,自引:0,他引:3  
L Piela  G Nemethy  H A Scheraga 《Biopolymers》1987,26(8):1273-1286
The conformational freedom of amino acid side chains is strongly reduced when the side chains occur on an α-helix. A quantitative evaluation of this freedom has been carried out by means of conformational energy computations for all naturally occurring amino acids and for α-aminobutyric acid when they are placed in the middle of a right-handed poly(L-alanine) α-helix. One of the three possible rotameric states for rotation around the Cα ? Cβ bond (viz. g+) is excluded completely on the helix because of steric hindrance, and the relative populations of the other two rotamers (t and g?) are altered because of steric interactions and the reduction of hydrogen-bonding possibilities. The computed tendencies of the changes in distributions of rotamers, on going from an ensemble of all backbone conformations to the α-helix, agree with the observed tendencies in proteins. Minimum-energy side-chain conformations in an α-helix have been tabulated for use in conformational energy computations on polypeptides.  相似文献   

17.
We employed the first principles computational method MembStruk and homology modeling techniques to predict the 3D structures of the human phenylthiocarbamide (PTC) taste receptor. This protein is a seven-transmembrane-domain G protein-coupled receptor that exists in two main forms worldwide, designated taster and nontaster, which differ from each other at three amino-acid positions. 3D models were generated with and without structural similarity comparison to bovine rhodopsin. We used computational tools (HierDock and ScanBindSite) to generate models of the receptor bound to PTC ligand to estimate binding sites and binding energies. In these models, PTC binds at a site distant from the variant amino acids, and PTC binding energy was equivalent for both the taster and nontaster forms of the protein. These models suggest that the inability of humans to taste PTC is due to a failure of G protein activation rather than decreased binding affinity of the receptor for PTC. Amino-acid substitutions in the sixth and seventh transmembrane domains of the nontaster form of the protein may produce increased steric hindrance between these two α-helices and reduce the motion of the sixth helix required for G protein activation.  相似文献   

18.
Gas-liquid chromatography (GLC) and enzymatic assays were developed for quantitating the imino acids, alanopine and strombine, alternate products of anaerobic glycolysis (replacing lactate) in the tissues of many marine invertebrates. For GLC analysis, d-strombine (2-methyliminodiacetic acid) and meso-alanopine (2,2′-iminodipropionic acid) were chromatographeo as N-trifluoroacetyl isobutyl esters. Modifications of techniques used for GLC analysis of amino acids were required to overcome steric hindrance in the acylation reaction caused by the presence of imino, rather than amino, groups. Both imino acids were separated from each other and from all amino acids by GLC. Detection limit of the technique was 0.05 μg imino acid. Enzymatic determination of imino acids made use of the alanopine-specific alanopine dehydrogenase (ADH) purified from the periwinkle, Littorina littorea, and the strombine/alanopine utilizing strombine dehydrogenase (SDH) from the clam, Mercenaria mercenaria, with assay conditions: 300 mm hydrazine buffer, pH 9.0, 5 mm NAD, and 0.3 unit ADH or 1.0 unit SDH. Enzymatic determinations of mixtures of alanopine and strombine in tissue samples required a dual analysis using both enzymes. Production of alanopine and strombine during anoxic stress in two species of marine molluscs was quantitated.  相似文献   

19.
Enantiomeric separations of N‐phthaloyl (N‐PHT), N‐tetrachlorophthaloyl (N‐TCPHT), and N‐naphthaloyl (N‐NPHT) α‐amino acids and their esters were examined on several kinds of polysaccharide‐derived chiral stationary phases (CSPs). Resolution capability of CSPs was greater Chiralcel OF than the others for N‐PHT and N‐NPHT α‐amino acids and their esters. In N‐TCPHT α‐amino acids and their esters, good enantioselectivities showed Chiralcel OG for N‐TCPHT α‐amino acids, Chiralpak AD for N‐TCPHT α‐amino acid methyl esters, and Chiralcel OD for N‐TCPHT α‐amino acid ethyl esters, respectively. From the results of liquid chromatography and computational chemistry, it is concluded that l ‐form is preferred and more retained with electrostatic interaction in case of interaction between N‐PHT α‐amino acid derivatives and Chiralcel OF, N‐TCPHT α‐amino acid derivatives and Chiralcel OD, and N‐NPHT α‐amino acid derivatives and Chiracel OF. On the other hand, d ‐form is preferred and more retained with van der Waals interaction in case of interaction between N‐TCPHT α‐amino acid ester derivatives and Chiralcel OG and Chiralpak AD. Chirality 24:1037–1046, 2012. © 2012 Wiley Periodicals, Inc.  相似文献   

20.
The 655 bp cytochrome c oxidase subunit I barcode region of single specimens of 388 species of fishes (four Holocephali, 61 Elasmobranchii and 323 Actinopterygii) was examined. All but two (Urolophus cruciatus and Urolophus sufflavus) showed different cox1 nucleotide sequences (99.5% species discrimination); the two that could not be resolved are suspected to hybridize. Most of the power of cox1 nucleotide sequence analysis for species identification comes from the degenerate nature of the genetic code and the highly variable nature of the third codon position of amino acids. Variation at the third codon position is bimodally distributed, and the more variable mode is dominated by amino acids with four or six codons, while the less variable mode is dominated by amino acids with two codons. The ratio of nonsynonymous to synomymous changes is much less than one, indicating that this gene is subject to strong purifying selection. Consequently, cox1 amino acid sequence diversity is much less than nucleotide sequence diversity and has very poor species resolution power. Fourteen of the 16 amino acid residues recognized as having important functions in the region of cox1 sequenced were completely conserved over all 388 species (and the bovine cox1 sequence), with one fish species varying at one of these sites, and three fish at another site. No significant differences in amino acid conservation were observed between residues in helices, strands and turns. Patterns of nucleotide and amino acid variability were very similar between elasmobranchs and actinopterygians.  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