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 present study deals with the conformation in solution of two heparin octasaccharides containing the pentasaccharide sequence GlcN(NAc,6S)-GlcA-GlcN(NS,3,6S)-IdoA(2S)-GlcN(NS,6S) [AGA*IA; where GlcN(NAc,6S) is N-acetylated, 6-O-sulfated alpha-D-glucosamine, GlcN(NS,3,6S) is N,3,6-O-trisulfated alpha-D-glucosamine and IdoA(2S) is 2-O-sulfated IdoA (alpha-L-iduronic acid)] located at different positions in the heparin chain and focuses on establishing geometries of IdoA residues (IdoA(2S) and IdoA) both inside and outside the AGA*IA sequence. AGA*IA constitutes the active site for AT (antithrombin) and is essential for the expression of high anticoagulant and antithrombotic activities. Analysis of NMR parameters [NOEs (nuclear Overhauser effects), transferred NOEs and coupling constants] for the two octasaccharides indicated that between the 1C4 and 2S0 conformations present in dynamic equilibrium in the free state for the IdoA(2S) residue within AGA*IA, AT selects the 2S0 form, as previously shown [Hricovini, Guerrini, Bisio, Torri, Petitou and Casu (2001) Biochem. J. 359, 265-272]. Notably, the 2S0 conformation is also adopted by the non-sulfated IdoA residue preceding AGA*IA that, in the absence of AT, adopts predominantly the 1C4 form. These results further support the concept that heparin-binding proteins influence the conformational equilibrium of iduronic acid residues that are directly or indirectly involved in binding and select one of their equi-energetic conformations for best fitting in the complex. The complete reversal of an iduronic acid conformation preferred in the free state is also demonstrated for the first time. Preliminary docking studies provided information on the octasaccharide binding location agreeing most closely with the experimental data. These results suggest a possible biological role for the non-sulfated IdoA residue preceding AGA*IA, previously thought not to influence the AT-binding properties of the pentasaccharide. Thus, for each AT binding sequence longer than AGA*IA, the interactions with the protein could differ and give to each heparin fragment a specific biological response.  相似文献   

2.
A complete study of the dynamics of two synthetic heparin-like hexasaccharides, D-GlcNHSO3-6-SO4-alpha-(1-->4)-L-IdoA-2-SO4-alpha-(1-->4)-D-GlcNHSO3-6-SO4-alpha-(1-->4)-L-IdoA-2-SO4-alpha-(1-->4)-D-GlcNHSO3-6-SO4-alpha-(1-->4)-L-IdoA-2-SO4-alpha-1-->iPr (1) and -->4)-L-IdoA-2-SO4-alpha-(1-->4)-D-GlcNHAc-6-SO4-alpha-(1-->4)-L-IdoA-alpha-(1-->4)-D-GlcNHSO3-alpha-(1-->4)-L-IdoA-2-SO4-alpha-1-->iPr (2), has been performed using 13C-nuclear magnetic resonance (NMR) relaxation parameters, T1, T2, and heteronuclear nuclear Overhauser effect (NOEs). Compound 1 is constituted from sequences corresponding to the major polysaccharide heparin region, while compound 2 contains a sequence never found in natural heparin. They differ from each other only in sulphation patterns, and are capable of stimulating fibroblast growth factors (FGFs)-1 induced mitogenesis. Both oligosaccharides exhibit a remarkable anisotropic overall motion in solution as revealed by their anisotropic ratios (tau /tau||), 4.0 and 3.0 respectively. This is a characteristic behaviour of natural glycosaminoglycans (GAG) which has also been observed for the antithrombin (AT) binding pentasaccharide D-GlcNHSO3-6-SO4-alpha-(1-->4)-D-GlcA-beta-(1-->4)-D-GlcNHSO3-(3,6-SO4)-alpha-(1-->4)-L-IdoA-2-SO4-alpha-(1-->4)-D-GlcNHSO3-6-SO4-alpha-1-->Me (3) (Hricovíni, M., Guerrini, M., Torri, G., Piani, S., and Ungarelli, F. (1995) Conformational analysis of heparin epoxide in aqueous solution. An NMR relaxation study. Carbohydr. Res., 277, 11-23). The motional properties observed for 1 and 2 provide additional support to the suitability of these compounds as heparin models in agreement with previous structural (de Paz, J.L., Angulo, J., Lassaletta, J.M., Nieto, P.M., Redondo-Horcajo, M., Lozano, R.M., Jiménez-Gallego, G., and Martín-Lomas, M. (2001) The activation of fibroblast growth factors by heparin: synthesis, structure and biological activity of heparin-like oligosaccharides. Chembiochem, 2, 673-685; Ojeda, R., Angulo, J., Nieto, P.M., and Martin-Lomas. M. (2002) The activation of fibroblast growth factors by heparin: synthesis and structural study of rationally modified heparin-like oligosaccharides. Can. J. Chem,. 80, 917-936; Lucas, R., Angulo, J., Nieto, P.M., and Martin-Lomas, M. (2003) Synthesis and structural studies of two new heparin-like hexasaccharides. Org. Biomol. Chem., 1, 2253-2266) and biological data (Angulo, J., Ojeda, R., de Paz, J.L., Lucas, R., Nieto, P.M., Lozano, R.M., Redondo-Horcajo, M., Giménez-Gallego, G., and Martín-Lomas, M. (2004) The activation of fibroblast growth factors (FGFs) by glycosaminoglycans: influence of the sulphation pattern on the biological activity of FGF-1. Chembiochem, 5, 55-61). Fast internal motions observed for the less sulphated compound 2, as compared with 1, may be related to their different behavior in stimulating FGF1-induced mitogenic activity.  相似文献   

3.
Biosynthesis of heparin. Availability of glucosaminyl 3-O-sulfation sites   总被引:1,自引:0,他引:1  
Heparin preparations isolated from pig intestinal mucosa and from bovine lung were fractionated with regard to affinity for antithrombin. The resulting fractions, with high (HA) or low (LA) affinity for the proteinase inhibitor, were analyzed by 13C NMR or by identification of di- and tetrasaccharides obtained through deaminative cleavage with nitrous acid. Structural differences between corresponding HA and LA fractions were essentially restricted to minor constituents, in particular 3-O-sulfated glucosamine units that occurred (1 or 2 residues/chain) in all HA preparations but were scarce or absent in LA heparin. The HA fractions also consistently showed higher contents of nonsulfated iduronic acid and, to a lesser extent, N-acetylated glucosamine units than the LA fractions. The two tetrasaccharide sequences, -IdoA-GlcNAc(6-OSO3)-GlcA-GlcNSO3- and -IdoA-GlcNAc(6-OSO3)-GlcA-GlcNSO3(6-OSO3)- , recently implicated as part of the acceptor site for glucosaminyl 3-O-sulfate groups (Kusche, M., B?ckstr?m, G., Riesenfeld, J., Petitou, M., Choay, J., and Lindahl, U. (1988) J. Biol. Chem. 263, 15474-15484), were identified in mucosal LA heparin; it was calculated that the preparation contained approximately one potential acceptor site/polysaccharide chain. Yet this material did not yield any labeled HA components on incubation with adenosine 3'-phosphate 5'-phospho-[35S]sulfate in the presence of glucosaminyl 3-O-sulfotransferase, solubilized from a mouse mastocytoma microsomal fraction. The failure to incorporate any 3-O-sulfate groups could conceivably be explained by the occurrence of a D-glucuronic rather than L-iduronic acid unit linked at the reducing ends of the above tetrasaccharide sequences. Alternatively, 3-O-sulfation may be restricted by other, as yet unidentified, inhibitory structural elements that are preferentially expressed in polysaccharide sequences selected for the generation of LA heparin.  相似文献   

4.
A synthetic tetradecapeptide having the sequence of the region of the antithrombin chain amino-terminal to the reactive bond, i.e. comprising residues P1 to P14, was shown to form a tight equimolar complex with antithrombin. A similar complex has previously been demonstrated between alpha 1-proteinase inhibitor and the analogous peptide of this inhibitor (Schulze, A. J., Baumann, U., Knof, S., Jaeger, E., Huber, R. and Laurell, C.-B. (1990) Eur. J. Biochem. 194, 51-56). The antithrombin-peptide complex had a conformation similar to that of reactive bond-cleaved antithrombin and, like the cleaved inhibitor, also had a higher conformational stability and lower heparin affinity than intact antithrombin. These properties suggest that the peptide bound to intact antithrombin at the same site that the P1 to P14 segment of the inhibitor occupies in reactive-bond-cleaved antithrombin, i.e. was incorporated as a sixth strand in the middle of the major beta-sheet, the A sheet. The extent of complex formation was reduced in the presence of heparin with high affinity for antithrombin, which is consistent with heparin binding and peptide incorporation being linked. Antithrombin in the complex with the tetradecapeptide had lost its ability to inactivate thrombin, but the reactive bond of the inhibitor was cleaved as in a normal substrate. These observations suggest a model, analogous to that proposed for alpha 1-proteinase inhibitor (Engh, R.A., Wright, H.T., and Huber, R. (1990) Protein Eng. 3, 469-477) for the structure of intact antithrombin, in which the A sheet contains only five strands and the P1 to P14 segment of the chain forms part of an exposed loop of the protein. The results further support a reaction model for serpins in which partial insertion of this loop into the A sheet is required for trapping of a proteinase in a stable complex, and complete insertion is responsible for the conformational change accompanying cleavage of the reactive bond of the inhibitor.  相似文献   

5.
The three-dimensional structure of an N-terminal fragment comprising the first 51 amino acids from human plasma vitronectin, the somatomedin B (SMB) domain, has been determined by two-dimensional NMR approaches. An average structure was calculated, representing the overall fold from a set of 20 minimized structures. The core residues (18-41) overlay with a root mean square deviation of 2.29 +/- 0.62 A. The N- and C-terminal segments exhibit higher root mean square deviations, reflecting more flexibility in solution and/or fewer long-range NOEs for these regions. Residues 26-30 form a unique single-turn alpha-helix, the locus where plasminogen activator inhibitor type-1 (PAI-1) is bound. This structure of this helix is highly homologous with that of a recombinant SMB domain solved in a co-crystal with PAI-1 (Zhou, A., Huntington, J. A., Pannu, N. S., Carrell, R. W., and Read, R. J. (2003) Nat. Struct. Biol. 10, 541-544), although the remainder of the structure differs. Significantly, the pattern of disulfide cross-links observed in this material isolated from human plasma is altogether different from the disulfides proposed for recombinant forms. The NMR structure reveals the relative orientation of binding sites for cell surface receptors, including an integrin-binding site at residues 45-47, which was disordered and did not diffract in the co-crystal, and a site for the urokinase receptor, which overlaps with the PAI-1-binding site.  相似文献   

6.
The antithrombotic activity of low molecular weight heparins (LMWHs) is largely associated with the antithrombin (AT)-binding pentasaccharide sequence AGA(*)IA (GlcN(NAc/NS,6S)-GlcA-GlcN(NS,3,6S)-IdoUA(2S)-GlcN(NS,6S)). The location of the AGA(*)IA sequences along the LMWH chains is also expected to influence binding to AT. This study was aimed at investigating the role of the structure and molecular conformation of different disaccharide extensions on both sides of the AGA(*)IA sequence in modulating the affinity for AT. Four high purity octasaccharides isolated by size exclusion chromatography, high pressure liquid chromatography, and AT-affinity chromatography from the LMWH enoxaparin were selected for the study. All the four octasaccharides terminate at their nonreducing end with 4,5-unsaturated uronic acid residues (DeltaU). In two octasaccharides, AGA(*)IA was elongated at the reducing end by units IdoUA(2S)-GlcN(NS,6S) (OCTA-1) or IdoUA-GlcN(NAc,6S) (OCTA-2). In the other two octasaccharides (OCTA-3 and OCTA-4), AGA(*)IA was elongated at the nonreducing side by units GlcN(NS,6S)-IdoUA and GlcN(NS,6S)-GlcA, respectively. Extensions increased the affinity for AT of octasaccharides with respect to pentasaccharide AGA(*)IA, as also confirmed by fluorescence titration. Two-dimensional NMR and docking studies clearly indicated that, although elongation of the AGA(*)IA sequence does not substantially modify the bound conformation of the AGA(*)IA segment, extensions promote additional contacts with the protein. It should be noted that, as not previously reported, the unusual GlcA residue that precedes the AGA(*)IA sequence in OCTA-4 induced an unexpected 1 order of magnitude increase in the affinity to AT with respect to its IdoUA-containing homolog OCTA-3. Such a residue was found to orientate its two hydroxyl groups at close distance to residues of the protein. Besides the well established ionic interactions, nonionic interactions may thus contribute to strengthen oligosaccharide-AT complexes.  相似文献   

7.
The O-specific polysaccharide chains (O-antigens) of the lipopolysaccharides (LPSs) of Proteus mirabilis O48 and Proteus vulgaris O21 were found to have tetrasaccharide and pentasaccharide repeating units, respectively, interlinked by a glycosidic phosphate. Polysaccharides and an oligosaccharide were derived from the LPSs by various degradation procedures and studied by 1H and 13C NMR spectroscopy, including 2D COSY, TOCSY, NOESY, H-detected 1H,13C and 1H,31P HMQC experiments. The following related structures of the repeating units of the O-antigens were established (top: Proteus mirabilis O48; bottom: Proteus vulgaris O21) The O-specific polysaccharide of P. vulgaris O21 has the same structure as that of Hafnia allvei 744 and PCM 1194 [Petersson C., Jachymek, W., Klonowska, A., Lugowski, C., Niedziela, T. & Kenne, L. (1997) Eur. J. Biochem., 245, 668-675], except that the GlcN residue carries the N-acetyl rather than the N-[(R)-3-hydroxybutyryl] group. Serological investigations confirmed the close relatedness of the Proteus and Hafnia O-antigens studied.  相似文献   

8.
Two-dimensional (2D) 1H-NMR spectra of porcine-brain natriuretic peptide (pBNP) have been recorded at 300 MHz and 400 MHz. Peak assignments have been made and the combined information from chemical shifts, coupling constants, temperature coefficients and NOEs have been used to determine the conformational properties of pBNP in (C2H3)2SO. Overall the peptide appears to be flexible, with the possibility of some beta-type structure near the C terminus. Some of the assignments and deduced structural features in the current study differ from those in a recent report by Inooka et al. [Inooka, H., Kikuchi, T., Endo, S., Ishibashi, Y., Wakimasu, M. and Mizuta, E. (1990) Eur. J. Biochem. 193, 127-134] which may indicate the sensitivity of the structure of this peptide to differences in solution conditions.  相似文献   

9.
1H and 31P NMR spectroscopies have been used to examine the effects of limited proteolysis with subtilisin Carlsberg on the global conformation of ovalbumin and on the local environment of phosphoserine 344, a residue two positions removed from the site of proteolysis. Such limited proteolysis has been shown to result in excision of a hexapeptide from the region of the protein that, in other serine protease inhibitors (serpins), contains the reactive center. Based on the structure of the related serpin alpha 1-antitrypsin, it has been predicted that phosphoserine 344 should undergo a large change in environment upon proteolysis of ovalbumin (L?bermann, H., Tokuoka, R., Deisenhofer, J., and Huber, R. (1984) J. Mol. Biol. 177, 531-550). Proteolysis of ovalbumin produces a small upfield shift (0.15 ppm) of the 31P resonance of phosphoserine 344. In addition, the pKa of phosphoserine 344 is raised by 0.1 pH unit. At pH 8.5, phosphoserine 344 in cleaved ovalbumin (plakalbumin) is as accessible to hydrolysis by Escherichia coli alkaline phosphatase as it is in native ovalbumin. 1H NMR shows that dephosphorylation of serine 344 has an imperceptible effect on the protein's conformation. Similarly, little effect on conformation is seen by 1H NMR upon proteolysis of ovalbumin. These findings suggest that ovalbumin does not undergo a marked conformational change analogous to that inferred for the related members of the serpin superfamily, alpha 1-antitrypsin and antithrombin III, nor do the residues close to the site of proteolysis appear to change environment from that of an exposed loop to a buried strand of beta-sheet. These findings are not consistent with the hypothesis of Carrell and Owen ((1985) Nature 317, 730-732) for the role of the exposed loop in serpins of directly facilitating conformational change upon cleavage of the loop. Instead, it is proposed that cleavage of the exposed loop alters the solvent accessibility of residues formerly covered by the loop and that this provides the thermodynamic impetus for conformational change, perhaps by disruption of a salt bridge crucial to the integrity of the native structure.  相似文献   

10.
The high-resolution X-ray crystal structure of staphylococcal nuclease (SNase) suggests that the guanidinium groups of Arg 35 and Arg 87 participate as electrophilic catalysts in the attack of water on the substrate phosphodiester. Both arginine residues have been replaced with "conservative" lysine residues so that both the importance of these residues in catalysis and the effect of changes in electrostatic interactions on active site conformation can be assessed. The catalytic efficiencies of R35K and R87K are decreased by factors of 10(4) and 10(5) relative to wild-type SNase, with R87K showing a very significant reduction in its affinity for both DNA substrate and the competitive inhibitor thymidine 3',5'-bisphosphate (pdTp). The thermal denaturation behavior of both mutant enzymes differs from that of wild type both in the absence and in the presence of the active site ligands Ca2+ and pdTp. Both the 1H NMR chemical shifts and interresidue nuclear Overhauser effects (NOEs) of residues previously assigned to be in the hydrophobic core of SNase are altered in R35K and R87K. These observations, similar to those recently reported by our laboratories for substitutions for Glu 43 [Hibler, D. W., Stolowich, N. J., Reynolds, M. A., Gerlt, J. A., Wilde, J. A., & Bolton, P. H. (1987) Biochemistry 26, 6278; Wilde, J. A., Bolton, P. H., Dell'Acqua, M., Hibler, D. W., Pourmotabbed, T., & Gerlt, J. A. (1988) Biochemistry 27, 4127], suggest that lysine substitutions are not conservative in SNase and disrupt the conformation of the active site.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

11.
A A Yee  J D O'Neil 《Biochemistry》1992,31(12):3135-3143
An alamethicin, secreted by the fungus Trichoderma viride and containing a glutamine at position 18 instead of the usual glutamic acid, has been uniformly labeled with 15N and purified by HPLC. The extent of 15N incorporation at individual backbone and side-chain sites was found to vary from 85% to 92%, as measured by spin-echo difference spectroscopy. The proton NMR spectrum of the peptide dissolved in methanol was assigned using correlation spectroscopies and nuclear Overhauser enhancements (NOE) measured in the rotating frame. The 15N resonances were assigned by the 2D 1H-15N correlation via heteronuclear multiple-quantum coherence experiment. NOEs and 3JNHC alpha H coupling constants strongly suggest that, in methanol, from Aib-3 to Gly-11, the peptide adopts a predominantly helical conformation, in agreement with previous 1H NMR studies [Esposito, G., Carver, J.A, Boyd, J., & Campbell, I.D. (1987) Biochemistry 26, 1043-1050; Banerjee, U., Tsui, F.-P., Balasubramanian, T.N., Marshall, G.R., & Chan, S I. (1983) J. Mol. Biol. 165, 757-775]. The conformation of the carboxyl terminus (12-20) is less well determined, partly because the amino acid composition reduces the number of NOEs and coupling constants which can be determined by 1H NMR spectroscopy. The 3JNHC alpha H in the C-terminus suggest the possibility of conformational averaging at Leu-12, Val-15, and Gln-19, an interpretation which is supported by a recent molecular dynamics simulation of the peptide [Fraternalli, F. (1990) Biopolymers 30, 1083-1099].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
A heparin binding, cell adhesion promoting domain, termed peptide F-9, from the B1 chain of human laminin, residues 641 to 660, i.e. RYVVLPRPVCFEKGMNYTVR, has been investigated by 1H NMR (500 MHz) spectroscopy and CD spectropolarimetry. While small linear peptides in water solution normally exist in a number of fluctuating conformational states, CD data analysis of peptide F9 indicates the existence of some preferred average structural populations consisting of about 30% beta-sheet, 22% beta-turn, and 6% alpha-helix. NMR structural analysis supports this observation and indicates specific sequences of preferred structural populations. Evidence for these is indicated by the presence of dNN nuclear Overhauser effect (NOE) populations and attenuated or absent d alpha N NOEs at short mixing times (0.1 s), 3J alpha N coupling constants of 5 and 10 Hz, and chemical shifts significantly removed from random coil positions. The NH2-terminal VVL sequence primarily exists in an extended chain conformation by virtue of large d alpha N NOEs and 9-10 Hz 3J alpha N coupling constants. Residues C10-N16 have turn-like or helix character with a run of dNN and d beta N NOEs and attenuated d alpha N NOEs. These midchain reversals include the lysine and asparagine residues proposed to be involved in heparin binding and N-glycosylation, respectively, to laminin peptide F-9.  相似文献   

13.
Resonances from the main repeating unit of heparan, →4)-β- -GlcA-(1→4)-- -GlcNAc-(1→, have been assigned by using a sample of the capsular polysaccharide of E. coli K5. Comparison of the spectra of heparan sulphate samples before and after O- and/or N-desulphation, with re-N-acetylation or re-N-sulphation, allowed assignment of some of the H-1 doublets in terms of sequence effects. Chemical shifts for H-1 of unsulphated uronic acid residues are influenced by 6-sulphation of the nearest neighbor GlcN on the reducing side; those of GlcN residues vary according to whether they have IdoA or GlcA as the nearest neighbour on the reducing side. The H-1 doublets due to residues in the binding sequence for antithrombin have been assigned by comparison of the spectra of heparins having high and low affinities for immobilised antithrombin.  相似文献   

14.
The lipopolysaccharide (LPS) of Chlamydia trachomatis L2 was isolated from tissue culture-grown elementary bodies using a modified phenol/water procedure followed by extraction with phenol/chloroform/light petroleum. From a total of 5 x 10(4) cm2 of infected monolayers, 22.3 mg of LPS were obtained. Compositional analysis indicated the presence of 3-deoxy-D-manno-oct-2-ulopyranosonic acid (Kdo), GlcN, phosphorus, and fatty acids in a molar ratio of 2.8:2:2.1:4.5. Matrix-assisted laser-desorption ionization mass spectrometry performed on the de-O-acylated LPS gave a major molecular ion peak at m/z 1781.1 corresponding to a molecule of 3 Kdo, 2 GlcN, 2 phosphates, and two 3-hydroxyeicosanoic acid residues. The structure of deacylated LPS obtained after successive treatment with hydrazine and potassium hydroxide was determined by 600 MHz NMR spectroscopy as Kdoalpha2-->8Kdoalpha2-->4Kdoalpha2-->6D-GlcpNbeta1 -->6D-GlcpNalpha 1,4'-bisphosphate. These data, together with those published recently on the acylation pattern of chlamydial lipid A (Qureshi, N., Kaltashov, I., Walker, K., Doroshenko, V., Cotter, R. J., Takayama, K, Sievert, T. R., Rice, P. A., Lin, J.-S. L., and Golenbock, D. T. (1997) J. Biol. Chem. 272, 10594-10600) allow us to present for the first time the complete structure of a major molecular species of a chlamydial LPS.  相似文献   

15.
NMR spectra of human lymphotactin (hLtn), obtained under various solution conditions, have revealed that the protein undergoes a major conformational rearrangement dependent on temperature and salt concentration. At high salt (200 mm NaCl) and low temperature (10 degrees C), hLtn adopts a chemokine-like fold, which consists of a three-stranded antiparallel beta-sheet and a C-terminal alpha-helix (Kulo?lu, E. S., McCaslin, D. R., Kitabwalla, M., Pauza, C. D., Markley, J. L., and Volkman, B. F. (2001) Biochemistry 40, 12486-12496). We have used NMR spectroscopy, sedimentation equilibrium, and intrinsic fluorescence to monitor the reversible conformational change undergone by hLtn as a function of temperature and ionic strength. We have used two-, three- and four-dimensional NMR spectroscopy of isotopically enriched protein samples to determine structural properties of the conformational state stabilized at 45 degrees C and 0 mm NaCl. Patterns of NOEs and (1)H(alpha) and (13)C chemical shifts show that hLtn rearranges under these conditions to form a four-stranded, antiparallel beta-sheet with a pattern of hydrogen bonding that is completely different from that of the chemokine fold stabilized at 10 degrees C and 200 mm NaCl. The C-terminal alpha-helix observed at 10 degrees C and 200 mm NaCl, which is conserved in other chemokines, is absent at 45 degrees C and no salt, and the last 38 residues of the protein are completely disordered, as indicated by heteronuclear (15)N-(1)H NOEs. Temperature dependence of the tryptophan fluorescence of hLtn in low and high salt confirmed that the chemokine conformation is stabilized by increased ionic strength. Sedimentation equilibrium analytical ultracentrifugation showed that hLtn at 40 degrees C in the presence of 100 mm NaCl exists mainly as a dimer. Under near physiological conditions of temperature, pH, and ionic strength, both the chemokine-like and non-chemokine-like conformations of hLtn are significantly populated. The functional relevance of this structural interconversion remains to be elucidated.  相似文献   

16.
A complex between the headpiece amino-terminal residues 1-56 of lac repressor (HP56) and an 11-bp lac operator fragment was studied by 1H NMR. The sequence specific assignment of the exchangeable and non-exchangeable protons has been accomplished. Several protons have favourable chemical shifts in the complex, therefore new intraprotein NOEs could be found that had not been unambigously identified in the free protein. By comparison, most of these intraprotein NOEs are also present in the spectra of the free headpiece but some are different. Furthermore, several new proteins DNA NOEs could be identified. The NOE between the side-chain amide protons of Gln18 and C5H of C7 confirms the specific contact between these residues which was proposed from genetic experiments [Ebright, R. M. (1985) J. Biomol. Struct. & Dyn. 3, 281-297]. The implications of the new data for the interaction between the lac repressor headpiece and its operator are discussed.  相似文献   

17.
The 18.1-kDa protein IIIGlc from Escherichia coli acts as both a phosphocarrier protein in the phosphoenolpyruvate:glycose phosphotransferase system (PTS) and as a signal-transducing protein with respect to the uptake of non-PTS sugars. Phosphorylation of IIIGlc at the N epsilon (N3) position of His-90 was effected through a regeneration system that included MgCl2, DTT, excess PEP, and catalytic amounts of Enzyme I and HPr. NH, 15N, and 13C alpha signal assignments for P-IIIGlc were made through comparison of 15N-1H correlation spectra (HSQC) of uniformly 15N-labeled preparations of phosphorylated and unphosphorylated protein and through analysis of three-dimensional triple-resonance HNCA spectra of P-IIIGlc uniformly labeled with both 15N and 13C. Backbone and side-chain 1H and 13C beta signals were assigned using 3D heteronuclear HCCH-COSY and HCCH-TOCSY spectra of P-IIIGlc. Using this approach, the assignments were made without reference to nuclear Overhauser effect data or assumptions regarding protein structure. The majority of NH, 15N, H alpha, and 13C alpha chemical shifts measured for P-IIIGlc were identical to those obtained for the unphosphorylated protein [Pelton, J. G., Torchia, D. A., Meadow, N. D., Wong, C.-Y., & Roseman, S. (1991) Biochemistry 30, 10043]. Those signals that exhibited shifts corresponded to residues within four segments (1) Leu-87-Gly-100, (2) Val-36-Val-46, (3) His-75-Ser-78, and (4) Ala-131-Val-138. These four segments are in close proximity to the active site residues His-75 and His-90 in the unphosphorylated protein [Worthylake, D., Meadow, N. D., Roseman, S., Liao, D., Hertzberg, O., & Remington, S.J. (1991) Proc. Natl. Acad. Sci. U.S.A. 88, 10382], and the chemical shift data provide strong evidence that if any structural changes accompany phosphorylation, they are confined to residues in these four segments. This conclusion is confirmed by comparing NOEs observed in 3D 15N/13C NOESY-HMQC spectra of the two forms of the protein. No NOE differences are seen for residues having the same chemical shifts in IIIGlc and P-IIIGlc. Furthermore, with the exception of residues Ala-76, Asp-94, and Val-96, the NOEs of residues (in the four segments) which exhibited chemical shift differences also had the same NOEs in IIIGlc and P-IIIGlc. In the case of residues Ala-76, Asp-94, and Val-96, minor differences in NOEs, corresponding to interproton distances changes of less than 1.5 A, were observed.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

18.
19.
Heparin has been shown to exhibit lower affinity for the antithrombin-thrombin complex than for antithrombin alone (Carlstrom, A.-S., Lieden, K., and Bjork, I. (1977) Thromb. Res. 11, 785-797), suggesting that structural alterations in antithrombin may accompany its reaction with thrombin. The hydroxy-nitrobenzyl (HNB) group attached to a unique tryptophan has been used in the present study as an extrinsic probe for localization of conformational changes to the heparin-binding region within antithrombin III using immunochemical and spectral techniques. Site-specific modification of tryptophan-49 in antithrombin with the hydroxynitrobenzyl reagent blocks heparin binding to the protein and provides a chemical label in the heparin-binding region of the protein (Blackburn, M. N., Smith, R. L., Carson, J., and Sibley, C. C. (1984) J. Biol. Chem. 259, 939-941). Antibodies specific for the hydroxynitrobenzyl hapten, which bind to HNB-tryptophan-49 in antithrombin, were used to detect a change in conformation in the region of tryptophan-49 which occurs upon thrombin binding to antithrombin. This thrombin-induced structural change was also apparent from spectral perturbations which were detected with the environmentally sensitive HNB moiety. Thus, the HNB group was used as an immunochemical probe as well as a spectral reporter group to provide insight into an allosteric mechanism of control in the catalytic role of heparin. The thrombin-promoted alteration of the structure in the heparin-binding region is presumably responsible for recycling of heparin, allowing it to catalyze further reactions between antithrombin and thrombin.  相似文献   

20.
Solution structure of the region 4 of sigma(70) subunit of Escherichia coli RNA polymerase, whose 4.2 subregion is involved in specific recognition of the -35 element of cognate promoters, has not been yet studied. Using multinuclear NMR spectroscopy, we have assigned recently all the backbone and aliphatic side-chain (13)C resonances for a recombinant His(6)-tagged protein containing the whole region 4 and a part of region 3.2 of sigma(70) in aqueous solution at pH 2.8 (Poznański, J., Zhukov, I., Bolewska, K., and Wierzchowski, K. L. (2001) J. Biomol. NMR 20, 181-2). The protein proved to be sufficiently soluble and did not aggregate only in the protonated state. In this paper, the structure and dynamics of this state at pH 2.8 have been extensively examined using CD and NMR spectroscopy. Both analysis of CD spectra and NMR observables (secondary chemical shifts of the (13)Calpha, (13)CO, and (1)Halpha nuclei and of vicinal (3)J(HNH)(alpha) coupling constants) indicated that a significant amount of helical structure remained in the protonated protein. The amount of this structure increased upon deprotonation of carboxylic amino acids, as shown by pH titration CD experiments. 2,2,2-Trifluoroethanol induced an even more extensive build up of this structure. Distribution along the protein sequence of the secondary shifts and (3)J(HNH)(alpha) couplings demonstrated partition of the helical secondary structure into three helices located similarly as in the crystal structures of the homologous region 4 of the sigma(A) subunit of Thermus aquaticus RNA polymerase (Campbell, E. A., Muzzin, O., Chlenov, M., Sun, J. L., Olson, A., Weinman, O., Trester-Zedlitz, M. L., and Darst, S. A. (2002) Mol. Cell 9, 527-39) and sigma(70) of the Thermus thermophilus RNA polymerase (Vassylyev, D. G., Sekine, S., Laptenko, O., Lee, J., Vassylyeva, M. N., Borukhov, S., and Yokoyama, S. (2002) Nature 417, 712-9.). Spectral density analysis of NMR relaxation parameters, R(1) and R(2), and [(1)H]-(15)N heteronuclear NOEs indicated that backbone fluctuations in the whole region embracing the three helices and intervening nonhelical sequences are severely restricted on the nanosecond time scale as compared with the N- and C-terminal protein segments. Inspection of the side-chain contacts stabilizing the crystal structures well explains the observed folding and solution properties of sigma(70)(4) protein in its protonated state.  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