首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The synthetic antithrombin-binding heparin pentasaccharide and a full-length heparin of approximately 26 saccharides containing this specific sequence have been compared with respect to their interactions with antithrombin and their ability to promote inhibition and substrate reactions of antithrombin with thrombin and factor Xa. The aim of these studies was to elucidate the pentasaccharide contribution to heparin's accelerating effect on antithrombin-proteinase reactions. Pentasaccharide and full-length heparins bound antithrombin with comparable high affinities (KD values of 36 +/- 11 and 10 +/- 3 nM, respectively, at I 0.15) and induced highly similar protein fluorescence, ultraviolet and circular dichroism changes in the inhibitor. Stopped-flow fluorescence kinetic studies of the heparin binding interactions at I 0.15 were consistent with a two-step binding process for both heparins, involving an initial weak encounter complex interaction formed with similar affinities (KD 20-30 microM), followed by an inhibitor conformational change with indistinguishable forward rate constants of 520-700 s-1 but dissimilar reverse rate constants of approximately 1 s-1 for the pentasaccharide and approximately 0.2 s-1 for the full-length heparin. Second order rate constants for antithrombin reactions with thrombin and factor Xa were maximally enhanced by the pentasaccharide only 1.7-fold for thrombin, but a substantial 270-fold for factor Xa, in an ionic strength-independent manner at saturating oligosaccharide. In contrast, the full-length heparin produced large ionic strength-dependent enhancements in second order rate constants for both antithrombin reactions of 4,300-fold for thrombin and 580-fold for factor Xa at I 0.15. These enhancements were resolvable into a nonionic component ascribable to the pentasaccharide and an ionic component responsible for the additional rate increase of the larger heparin. Stoichiometric titrations of thrombin and factor Xa inactivation by antithrombin, as well as sodium dodecyl sulfate-polyacrylamide gel electrophoresis of the products of these reactions, indicated that pentasaccharide and full-length heparins similarly promoted the formation of proteolytically modified inhibitor during the inactivation of factor Xa by antithrombin, whereas only the full-length heparin was effective in promoting this substrate reaction of antithrombin during the reaction with thrombin.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

2.
Thrombospondin (Tsp), a protein secreted by activated platelets, forms disulfide-linked complexes with thrombin [K. J. Danishefsky, R. J. Alexander and T. C. Detwiler (1984) Biochemistry 23, 4984]. Thiols and disulfide bonds of Tsp were analyzed, and a search was made for other Tsp covalent complexes. Platelets in 1 mM EDTA were activated with ionophore A23187, and the secreted proteins were analyzed by gel electrophoresis in sodium dodecyl sulfate. One millimolar dithioerythritol (DTE) decreased the electrophoretic mobility of Tsp, indicating reduction of an intrachain disulfide bond; Ca2+ prevented this effect. Electrophoresis of single-chain Tsp prepared with 50 mM DTE in either EDTA or Ca2+ also revealed a Ca2+-stabilized intrachain disulfide bond. Ca2+ prevented the retention of Tsp on an activated thiol-Sepharose column, indicating protection of a thiol by Ca2+. Incubation at 37 degrees C for 60 min resulted in complexes with apparent mass much greater than 500 kDa. Formation of complexes was prevented by N-ethylmaleimide, by a temperature less than 25 degrees C, and by Ca2+ or Mg2+. From pH 6 to 9, complexes formed better at lower pH. Two-dimensional (nonreduced/reduced) electrophoresis revealed Tsp but no other constituents of the complexes. With 10 nM thrombin, complexes formed faster and included thrombin; Ca2+ only partially inhibited. The complex was very susceptible to dissociation by low concentrations (2.5 mM) of DTE. It is concluded that Tsp has a reactive thiol and an intrachain disulfide bond that are protected by Ca2+. When these groups are unprotected, there is intermolecular thiol-disulfide exchange.  相似文献   

3.
4.
Cholinesterases (ChE), use a Glu-His-Ser catalytic triad to enhance the nucleophilicity of the catalytic serine. It has been shown that serine proteases, which employ an Asp-His-Ser catalytic triad for optimal catalytic efficiency, decrease the hydrogen bonding distance between the Asp-His pair to form a short, strong hydrogen bond (SSHB) upon binding mechanism-based inhibitors, which form tetrahedral Ser-adducts, analogous to the tetrahedral intermediates in catalysis, or at low pH when the histidine is protonated [Cassidy, C. S., Lin, J., Frey, P. A. (1997) Biochemistry 36, 4576-4584]. Two types of mechanism-based inhibitors were bound to pure equine butyrylcholinesterase (BChE), a 364 kDa homotetramer, and the complexes were studied by (1)H NMR at 600 MHz and 25-37 degrees C. The downfield region of the (1)H NMR spectrum of free BChE at pH 7.5 showed a broad, weak, deshielded resonance with a chemical shift, delta = 16.1 ppm, ascribed to a small amount of the histidine-protonated form. Upon addition of a 3-fold excess of diethyl 4-nitrophenyl phosphate (paraoxon) and subsequent dealkylation, the broad 16.1 ppm resonance increased in intensity 4.7-fold, and yielded a D/H fractionation factor phi = 0.72+/-0.10 consistent with a SSHB between Glu and His of the catalytic triad. From an empirical correlation of delta with hydrogen-bond length in small crystalline compounds, the length of this SSBH is 2.64+/-0.04 A, in agreement with the length of 2.62+/-0.02 A independently obtained from phi. The addition of a 3-fold excess of m-(N,N, N-trimethylammonio)trifluoroacetophenone to BChE yielded no signal at 16.1 ppm, and a 640 Hz broad, highly deshielded proton resonance with a chemical shift delta = 18.1 ppm and a D/H fractionation factor phi = 0.63+/-0.10, also consistent with a SSHB. The length of this SSHB is calculated to be 2.62+/-0.04 A from delta and 2.59+/-0.03 A from phi. These NMR-derived distances agree with those found in the X-ray structures of the homologous acetylcholinesterase complexed with the same mechanism-based inhibitors, 2.60+/-0.22 and 2.66+/-0.28 A. However, the order of magnitude greater precision of the NMR-derived distances establish the presence of SSHBs. We suggest that ChEs achieve their remarkable catalytic power in ester hydrolysis, in part, due to the formation of a SSHB between Glu and His of the catalytic triad.  相似文献   

5.
Juneja J  Udgaonkar JB 《Biochemistry》2002,41(8):2641-2654
The unfolding of ribonuclease A was studied in 5.2 M guanidine hydrochloride at pH 8 and 10 degrees C using multiple optical probes, native-state hydrogen exchange (HX), and pulse labeling by hydrogen exchange. First, native-state HX studies were used to demonstrate that the protein exists in two slowly interconverting forms under equilibrium native conditions: a predominant exchange-incompetent N form and an alternative ensemble of conformations, N(I), in which some amide hydrogens are fully exposed to exchange. Pulsed HX studies indicated that, during unfolding, the rates of exposure to exchange with solvent protons were similar for all backbone NH probe protons. It is shown that two parallel routes of unfolding are available to the predominant N conformation as soon as it encounters strong unfolding conditions. A fraction of molecules appears to rapidly form N(I) on one route. On the other route an exchange-incompetent intermediate state ensemble, I(U)(2), is formed. The kinetics of unfolding measured by far-UV circular dichroism (CD) were faster than those measured by near-UV CD and intrinsic tyrosine fluorescence of the protein. The logarithms of the rate constants of the unfolding reaction measured by all three optical probes also showed a nonlinear dependence on GdnHCl concentration. All of the data suggest that N(I) and I(U)(2) are nativelike in their secondary and tertiary structures. While N(I) unfolds directly to the fully exchange-competent unfolded state (U), I(U)(2) forms another intermediate I(U)(3) which then unfolds to U. I(U)(3) is devoid of all native alpha-helical secondary structure and has only 30% of the tertiary interactions still intact. Since the rates of global unfolding measured by near-UV CD and fluorescence agree well with the rates of exposure determined for all of the backbone NH probe protons, it appears that the rate-limiting step for the unfolding of RNase A is the dissolution of the entire native tertiary structure and penetration of water into the hydrophobic core.  相似文献   

6.
Summary The relaxation rates of the multiple-quantum coherence for the amide hydrogen of Gly13 in ras p21·GDP were determined in the presence and absence of 17O labeling in the -phosphate of GDP. No significant difference could be observed between labeled and unlabeled samples, in spite of the fact that the hydrogen bond from the amide group of Gly13 to the -phosphate is shorter than is typical, based on its chemical shift. For macromolecules in which an oxygen atom is the acceptor of a hydrogen bond, dipolar coupling between 17O and hydrogen is unlikely to be observable, except for extremely short H-bonds.Abbreviations p21 21-kD protein product of the human H-ras gene - GMPPCP guanylyl [,-methylene]diphosphate - HPLC high-performance liquid chromatography - CSA chemical shift anisotropy - HMQC heteronuclear multiple-quantum coherence  相似文献   

7.
Spies MA  Toney MD 《Biochemistry》2003,42(17):5099-5107
Alanine racemase catalyzes the pyridoxal phosphate-dependent interconversion of the D- and L-isomers of alanine. Previous studies have shown that the enzyme employs a two-base mechanism in which Lys39 and Tyr265 are the acid/base catalysts. It is thus possible that stereoisomerization of the external aldimine intermediates occurs through a concerted double proton transfer without the existence of a distinct carbanionic intermediate. This possibility was tested by the application of multiple kinetic isotope effect (KIE) methodology to alanine racemase. The mutual dependence of primary substrate and solvent deuterium KIEs has been measured using equilibrium perturbation-type experiments. The conceptually straightforward measurement of the substrate KIE in H(2)O is complemented with a less intuitive protium washout perturbation-type measurement in D(2)O. The primary substrate KIE in the D --> L direction at 25 degrees C is reduced from 1.297 in H(2)O to 1.176 in D(2)O, while in the L --> D direction it is reduced from 1.877 in H(2)O to 1.824 in D(2)O. Similar reductions are also observed at 65 degrees C, the temperature to which the Bacillus stearothermophilus enzyme is adapted. These data strongly support a stepwise racemization of stereoisomeric aldimine intermediates in which a substrate-based carbanion is an obligatory intermediate. The ionizations observed in k(cat)/K(M) pH profiles have been definitively assigned based on the DeltaH(ion) values of the observed pK(a)'s with alanine and on the pH dependence of k(cat)/K(M) for the alternative substrate serine. The acidic pK(a) in the bell-shaped curve is due to the phenolic hydroxyl of Tyr265, which must be unprotonated for reaction with either isomer of alanine. The basic pK(a) is due to the substrate amino group, which must be protonated to react with Tyr265-unprotonated enzyme. A detailed reaction mechanism incorporating these results is proposed.  相似文献   

8.
Heparin enhances strongly the mitogenic properties of human acidic fibroblast growth factor (h-aFGF) on hamster fibroblast (CCL 39) or bovine lens epithelial cells (BEL). We report here that a synthetic heparin pentasaccharide with high affinity for antithrombin III has the same effect as heparin at about the same concentration. Thus a pentasaccharidic sequence may represent the shortest heparin structure which interacts with h-aFGF.  相似文献   

9.
Partitioning of the ketimine (or ketimine + quinonoid) intermediate(s) in the mitochondrial aspartate aminotransferase reactions was investigated by following the rates of loss of 18O from carbonyl-18O-enriched alpha-ketoglutarate together with the rate of L-glutamate formation. The ratio of these rate constants was found to equal 1 at 10 degrees C, implying that the above intermediate(s) face(s) equal barriers with respect to the forward and reverse reactions. This partition ratio of 1 together with that measured from the alpha-amino acid side of the reaction [Julin, D.A., Wiesinger, H., Toney, M. D., & Kirsch, J.F. (1989) Biochemistry (preceding paper in this issue)] suggests that the rate constant for exchange of alpha-ketoglutarate-derived H2(18)O from the ketimine (or ketimine + quinonoid) form(s) of the enzyme with solvent is comparable with that for kcat.  相似文献   

10.
The base-pair opening kinetics of the self-complementary oligomer d(CGCGAATTCGCG)(2) has been derived from NMR measurements of the imino proton exchange. In general, it has previously been found that imino proton exchange in duplex DNA is limited by the proton-transfer step from the open state and that the dependence of the exchange times on the inverse concentration of an added exchange catalyst is linear. In the present study, a curvature is observed for, in particular, the innermost AT base pair with both ammonia and trimethylamine (TMA) as exchange catalysts. The two catalysts act on the same open states, but the accessibility of TMA is reduced by a factor of 2-3 compared to ammonia. Assuming that ammonia accesses the imino proton equally in the open state of the base pair and in the mononucleoside, the curvature is consistent with 7-9% of the openings ending in open states with lifetimes of about 1 micros while the bulk of open-state lifetimes fall in the nanosecond range. A curvature is also found for the exchange times of the imino protons in the A-tract sequence CGCA(8)CGC/GCGT(8)GCG. This curvature becomes increasingly pronounced from the 5'-end toward the center of the tract and hereby seems to be correlated with the contraction of the minor groove. Thus, while the base-pair lifetimes deduced from the present study are in accordance with previous measurements, a substantial fraction of the open states formed by the central AT-base pairs in the two oligomers exhibits microsecond lifetimes in contrast to previous estimates in the nanosecond range. These findings may be of relevance for the way sequence specific recognition is accomplished by proteins and ligands.  相似文献   

11.
Experimental evidence for a persistent spore bank in Sphagnum   总被引:1,自引:0,他引:1  
Spore capsules of four Sphagnum species were buried at different depths in peat on a bog. Spore viability was determined after 0, 1, 2 and 3 yr. Viability generally declined with time, but viable spores were still found at all depths after 3 yr. The light-coloured spores of S. balticum and S. tenellum retained their viability better than the darker spores of S. fuscum and S. lindbergii . Survival was highest under wet but aerobic conditions, but was also high under humid or periodically desiccated conditions. By contrast, most spores stored under wet, anaerobic conditions died within 2–3 yr. These results, and predictions from them, are not consistent with earlier results for spores of long-lived and dominant bryophytes, or for seeds of phanerogams of undisturbed wetlands and forests. There was no correlation between spore size and longevity across species, but the small spores from small capsules of S. balticum and S. tenellum generally showed higher viability than those from the medium-sized and large capsules of the same species. This suggests a positive intraspecific relationship between longevity and dispersal distance. There was an indication of conditional dormancy, controlled by weather, in Sphagnum spores. The experiments indicate that Sphagnum spores can form a long-term persistent spore bank under suitable conditions, with a half-life of between 1 and 20 yr (mean across species of 2.6 and 5.0 yr at two depths studied), and with potential values in individual spore capsules of several decades, or even of centuries. Sphagnum spores kept refrigerated showed 15–35% viable spores after 13 yr. The capacity to form a persistent spore bank that can be activated whenever favourable conditions occur might help explain the wide geographical distribution of many Sphagnum species in the boreal and temperate zones, where they have managed to colonize almost every suitable patch of acidic, nutrient-poor wetland.  相似文献   

12.
Dimensions of a toroidal globule appearing at the compactization of DNA macro-molecule in a poor low-molecular solvent are estimated. Diagram of DNA molecule state was analysed in variables: contour length-concentration of the precipitating agent. Possible existence of the tricritical point in the joint of the regions of coil, isotropic globular and liquid crystalline globular states was suggested.  相似文献   

13.
Barrick D  Dahlquist FW 《Proteins》2000,39(4):278-290
The structural role of a side-chain to side-chain protein hydrogen bond is examined using trans-substitution of the proximal histidine of myoglobin with methylimidazoles (Barrick, Biochemistry 1994;33:6546-6554). Modification of the chemical structure of exogenous ligands allows this hydrogen bond to be disrupted. Comparison of the crystal structures of H93G myoglobin complexed 4-methylimidazole (4meimd; methylation at carbon 4) and 1-methylimidazole (1meimd; methylation at the adjacent nitrogen, preventing hydrogen bonding between the imidazole ligand and the protein) shows that the polypeptide, heme, and methylimidazole orientations are the same within error. For 4meimd there appear to be major and minor conformations corresponding to different tautomeric states of the ligand. Conformational heterogeneity is also seen in the hyperfine-shifted region of the NMR spectrum of 4meimd complexed with high-spin H93G deoxyMb. The major conformation of the 4meimd ligand and the 1meimd ligand, as seen in the respective crystal structures, are quite similar except that the proximal ligand NH-to-Ser92-OH hydrogen bond is eliminated in the 1meimd complex, and instead the proximal ligand CH is adjacent to the Ser92-OH. Thus, this system provides a means to eliminate the Mb proximal hydrogen bond in a chemically and structurally conservative way.  相似文献   

14.
A previous comprehensive analysis of the pH dependence of native-state amide hydrogen (NH) exchange in turkey ovomucoid third domain (OMTKY3) yielded apparent opening and closing rate constants (k(op) and k(cl)) at 14 NH groups involved in global conformational changes. This analysis has been extended to 18 additional slowly exchanging NH groups. Quench-flow experiments were performed to monitor NH exchange in native OMTKY3 from neutral to very alkaline pH ( approximately 12) conditions. Above pH 10 the mechanism of exchange switched from one governed by a rapid equilibrium preceding the chemistry of exchange (i.e. EX2 exchange), to one where exchange was limited by the rate of opening (i.e. EX1 exchange). Kinetics of solvent exposure are now known for nearly all backbone NH groups in native OMTKY3, yielding rate constants that span five orders of magnitude, 0.004 to 200 s(-1).  相似文献   

15.
16.
Interaction of the peptide bond with solvent water: a vapor phase analysis   总被引:1,自引:0,他引:1  
R Wolfenden 《Biochemistry》1978,17(1):201-204
A dynamic technique, using radioactivity as a means of detection, makes it possible to measure the partial pressures of highly polar compounds in dilute aqueous solution. The results can be expressed in terms of the dimensionless distribution coefficient for transfer of a compound from dilute aqueous solution to the vapor phase. For acetic acid this coefficient is 1.1 X 10(-5), for acetamide 7.6 X 10(-8), for N-methylacetamide 4.1 X 10(-8), and for N,N-dimethylacetamide 5.4 X 10(-7). Thus acetamide is much more strongly solvated than the uncharged acetic acid molecule. The results suggest: (1) that the peptide bond represents an extreme among uncharged functional groups in the degree to which it is stabilized by solvent water; (2) that the very great hydrophilic character of the peptide bond may be associated mainly with hydrogen bonding of the solvent to the carbonyl oxygen atom (rather than the N-H group); and (3) that the observed equilibria of biosynthesis and hydrolysis of peptide bonds in aqueous solution are largely determined by differences between reactants and products in their free energies of solvation. It is anticipated that where "bound" water is found in proteins, it will often be found to be associated with peptide bonds, and will tend to be associated with the C-O group rather than with the N-H group.  相似文献   

17.
The serine protease subtilisin BPN' is a useful catalyst for peptide synthesis when dissolved in high concentrations of a water-miscible organic co-solvent such as N,N-dimethylformamide (DMF). However, in 50% DMF, the k(cat) for amide hydrolysis is two orders of magnitude lower than in aqueous solution. Surprisingly, the k(cat) for ester hydrolysis is unchanged in 50% DMF. To explain this alteration in activity, the structure of subtilisin 8397+1 was determined in 20, 35, and 50% (v/v) DMF to 1.8 A resolution. In 50% DMF, the imidazole ring of His64, the central residue of the catalytic triad, has rotated approximately 180 degrees around the Cbeta-Cgamma bond. Two new water molecules in the active site stabilize the rotated conformation. This rotation places His64 in an unfavorable geometry to interact with the other members of the catalytic triad, Ser221 and Asp32. NMR experiments confirm that the characteristic resonance due to the low barrier hydrogen bond between the His64 and Asp32 is absent in 50% DMF. These experiments provide a clear structural basis for the change in activity of serine proteases in organic co-solvents.  相似文献   

18.
The structures of the tetrasaccharide (β-D-glucuronic acid)1→4 (N-sulfate-3,6-di-O-sulfate-α-D-glucosamine)1→4(2-O-sulfate-α-L-iduronic acid)1→4(N-sulfate-6-O-sulfate-D-glucosamine) and of the pentasaccharide (N-sulfate-6-O-sulfate-α-D-glucosamine)1→4(β-D-glucuronic acid)1→4(N- sulfate-3,6-di-O-sulfate-α-D-glucosamine)1→4(2-O-sulfate-α-L-iduronic acid)1→4(N-sulfate-6-O-sulfate-D-glucosamine), both prepared for the first time, by chemical synthesis from D-glucose and D-glucosamine, have been confirmed by nuclear magnetic resonance. The synthetic tetrasaccharide neither binds to AT-III nor induces anti-factor Xa activity enhancement of this inhibitor. In contrast, the synthetic pentasaccharide strongly binds to AT-III (Ka : 7.106M?1) forming an equimolar complex and also enhances the AT-III inhibitory activity towards factor Xa. These results confirm that the synthetic pentasaccharide with the above structure corresponds to the actual minimal sequence required in heparin for binding to AT-III.  相似文献   

19.
During the preparation of multi-component model membranes, a primary consideration is that compositional homogeneity should prevail throughout the suspension. Some conventional sample preparation methods pass the lipid mixture through an intermediary, solvent-free state. This is an ordered, solid state and may favor the demixing of membrane components. A new preparative method has been developed which is specifically designed to avoid this intermediary state. This novel strategy is called rapid solvent exchange (RSE) and entails the direct transfer of lipid mixtures between organic solvent and aqueous buffer. RSE liposomes require no more than a minute to prepare and manifest considerable entrapment volumes with a high fraction of external surface area. In phospholipid/cholesterol mixtures of high cholesterol content, suspensions prepared by more conventional methods reveal evidence of artifactual demixing, whereas samples prepared by rapid solvent exchange do not. The principles which may lead to artifactual demixing during conventional sample preparation are discussed.  相似文献   

20.
The crystal structures of Klebsiella pneumoniae pullulanase and its complex with glucose (G1), maltose (G2), isomaltose (isoG2), maltotriose (G3), or maltotetraose (G4), have been refined at around 1.7-1.9A resolution by using a synchrotron radiation source at SPring-8. The refined models contained 920-1052 amino acid residues, 942-1212 water molecules, four or five calcium ions, and the bound sugar moieties. The enzyme is composed of five domains (N1, N2, N3, A, and C). The N1 domain was clearly visible only in the structure of the complex with G3 or G4. The N1 and N2 domains are characteristic of pullulanase, while the N3, A, and C domains have weak similarity with those of Pseudomonas isoamylase. The N1 domain was found to be a new type of carbohydrate-binding domain with one calcium site (CBM41). One G1 bound at subsite -2, while two G2 bound at -1 approximately -2 and +2 approximately +1, two G3, -1 approximately -3 and +2 approximately 0', and two G4, -1 approximately -4 and +2 approximately -1'. The two bound G3 and G4 molecules in the active cleft are almost parallel and interact with each other. The subsites -1 approximately -4 and +1 approximately +2, including catalytic residues Glu706 and Asp677, are conserved between pullulanase and alpha-amylase, indicating that pullulanase strongly recognizes branched point and branched sugar residues, while subsites 0' and -1', which recognize the non-reducing end of main-chain alpha-1,4 glucan, are specific to pullulanase and isoamylase. The comparison suggested that the conformational difference around the active cleft, together with the domain organization, determines the different substrate specificities between pullulanase and isoamylase.  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