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 718 毫秒
1.
Heparin and low-molecular-weight heparins (LMWHs) are anticoagulant drugs that mainly inhibit the coagulation cascade by indirectly interacting with factor Xa and factor IIa (thrombin). Inhibition of factor Xa by antithrombin (AT) requires the activation of AT by specific pentasaccharide sequences containing 3-O-sulfated glucosamine. Activated AT also inhibits thrombin by forming a stable ternary complex of AT, thrombin, and a polysaccharide (requires at least an 18-mer/octadeca-mer polysaccharide). The full structure of any naturally occurring octadecasaccharide sequence has yet to be determined. In the context of the development of LMWH biosimilars, structural data on such important biological mediators could be helpful for better understanding and regulatory handling of these drugs. Here we present the isolation and identification of an octadecasaccharide with very high anti-factor Xa activity (∼3 times higher than USP [U.S. Pharmacopeia] heparin). The octadecasaccharide was purified using five sequential chromatographic methods with orthogonal specificity, including gel permeation, AT affinity, strong anion exchange, and ion-pair chromatography. The structure of the octadecasaccharide was determined by controlled enzymatic sequencing and nuclear magnetic resonance (NMR). The isolated octadecasaccharide contained three consecutive AT-binding sites and was tested in coagulation assays to determine its biological activity. The isolation of this octadecasaccharide provides new insights into the modulation of thrombin activity.  相似文献   

2.
Oligosaccharides of well-defined molecular size were prepared from heparin by nitrous acid depolymerization, affinity chromatography on immobilized antithrombin III (see footnote on Nomenclature) and gel chromatography on Sephadex G-50. High affinity (for antithrombin III) octa-, deca-, dodeca-, tetradeca-, hexadeca- and octadeca-saccharides were prepared, as well as oligosaccharides of larger size than octadecasaccharide. The inhibition of Factor Xa by antithrombin III was greatly accelerated by all of these oligosaccharides, the specific anti-Factor Xa activity being invariably greater than 1300 units/mumol. The anti-Factor Xa activity of the decasaccharide was not significantly decreased in the presence of platelet factor 4, even at high platelet factor 4/oligosaccharide ratios. Measurable but incomplete neutralization of the anti-Factor Xa activities of the tetradeca- and hexadeca-saccharides was observed, and complete neutralization of octadeca- and larger oligo-saccharides was achieved with excess platelet factor 4. The octa-, deca-, dodeca-, tetradeca- and hexadeca-saccharides had negligible effect on the inhibition of thrombin by antithrombin III, whereas specific anti-thrombin activity was expressed by the octadeca-saccharide and by the larger oligosaccharides. An octadecasaccharide is therefore the smallest heparin fragment (prepared by nitrous acid depolymerization) that can accelerate thrombin inhibition by antithrombin III. The anti-thrombin activities of the octadecasaccharide and larger oligosaccharides were more readily neutralized by platelet factor 4 than were their anti-Factor Xa activities. These findings are compatible with two alternative mechanisms for the action of platelet factor 4, both involving the binding of the protein molecule adjacent to the antithrombin III-binding site. Such binding results in either steric interference with the formation of antithrombin III-proteinase complexes or in displacement of the antithrombin III molecule from the heparin chain.  相似文献   

3.
The therapeutic anticoagulant action of heparin is mediated by the ability of a multifunctional octadecasaccharide region of the molecule to bind to and differentially alter the conformational integrity of antithrombin, and the sugar sequence of the primary binding domain is known. Low ultraviolet circular dichroism spectroscopy of heparin-derived anticoagulant octa-, deca-, dodeca-, tetradeca-, and octadecasaccharides has been useful in elucidating the nature of the sugars that are contained within the second functional domain of the octadecasaccharide. The difference between the spectra of the molar ellipticity of the above sequential oligosaccharides was taken to be the CD spectrum of the corresponding additional disaccharide unit(s). Optical models of the component disaccharides of heparin were derived from CD spectra of heparins having a high degree of sulfation, synthetic glycosides of N-acetylglucosamine, and glycosides of component uronic acids. These were sufficiently distinct in magnitude and spectral position to warrant interpretation of the experimental difference CD spectra. The uronic acids of the disaccharides between deca- and octamer, dodeca- and decamer, and tetradeca- and dodecamer were thereby ascribed to sulfated iduronate, unsulfated iduronate, and glucuronate residues, respectively, while those of the tetrasaccharide between the octadeca- and tetradecasaccharide were tentatively assigned to sulfated iduronate moieties. Interpretation of the difference CD spectra on the basis of the optical models was less certain in regard to the amino sugar components. It appears that the amino sugar derivative between the dodeca- and decamer was N-acetylglucosamine, while the other disaccharides of the octa- to octadecasaccharide probably contained the N-sulfated derivative. A speculative disaccharide sequence drawn from these data indicates that relatively less strongly anionic disaccharides, having nonsulfated uronic acid moieties and N-acetylglucosamine, were flanked by trisulfated disaccharide units, constituting a structural element similar to that which contains the primary binding domain of the anticoagulant.  相似文献   

4.
Oligosaccharides (10-20 monosaccharide units) with high affinity for antithrombin, as well as larger high-affinity heparin fractions (having relative molecular masses between 6,000 and 21,500), all markedly accelerated the inhibition of Factor Xa by antithrombin. Moreover, all high-affinity oligosaccharides and heparins enhanced, to a similar extent, the amount of free proteolytically modified antithrombin cleaved at the reactive bond by Factor Xa. In contrast, a minimum high-affinity heparin size of approximately 18 monosaccharide units was required to significantly accelerate the inactivation of thrombin by antithrombin and to enhance the production of modified antithrombin by this enzyme. All high-affinity fractions studied had similar affinities for antithrombin, as determined by fluorescence titrations. In competition experiments, binary complexes of antithrombin with octadecasaccharide or larger high-affinity heparins, but not with smaller oligosaccharides, displaced inactivated 125I-thrombin from matrix-linked low-affinity heparin. Moreover, similar binary complexes with 3H-labeled octadecasaccharide or larger chains, but not with smaller oligosaccharides, were capable of binding to matrix-linked inactivated thrombin. These results indicate that simultaneous binding of antithrombin and thrombin to high-affinity heparin is a prerequisite to the acceleration of the antithrombin-thrombin reaction and that the minimum heparin sequence capable of binding both proteins comprises approximately 18 monosaccharide units. Similar complex formation apparently is not required for the acceleration of the antithrombin-Factor Xa reaction.  相似文献   

5.
Sclerostin (SOST) is a glycoprotein having many important functions in the regulation of bone formation as a key negative regulator of Wnt signaling in bone. Surface plasmon resonance (SPR), which allows for a direct quantitative analysis of the label-free molecular interactions in real-time, has been widely used for the biophysical characterization of glycosaminoglycan (GAG)-protein interactions. In the present study, we report kinetics, structural analysis and the effects of physiological conditions (e.g., salt concentrations, Ca2+ and Zn2+concentrations) on the interactions between GAGs and recombinant human (rh) and recombinant mouse (rm) SOST using SPR. SPR results revealed that both SOSTs bind heparin with high affinity (rhSOST-heparin, KD~36 nM and rmSOST-heparin, KD~77 nM) and the shortest oligosaccharide of heparin that effectively competes with full size heparin for SOST binding is octadecasaccharide (18mer). This heparin binding protein also interacts with other highly sulfated GAGs including, disulfated-dermatan sulfate and chondroitin sulfate E. In addition, liquid chromatography-mass spectrometry was used to characterize the structure of sulfated GAGs that bound to SOST.  相似文献   

6.
Green fluorescent proteins (GFPs) and their derivatives are widely used as markers to visualize cells, protein localizations in in vitro and in vivo studies. The use of GFP fusion protein for visualization is generally thought to have negligible effects on cellular function. However, a number of reports suggest that the use of GFP may impact the biological activity of these proteins. Heparin is a glycosaminoglycan (GAG) that interacts with a number of proteins mediating diverse patho-physiological processes. In the heparin-based interactome studies, heparin-binding proteins are often prepared as GFP fusion proteins. In this report, we use surface plasmon resonance (SPR) spectroscopy to study the impact of the GFP tagging on the binding interaction between heparin and a heparin-binding protein, the Roundabout homolog 1 (Robo1). SPR reveals that heparin binds with higher affinity to Robo1 than GFP-tagged Robo1 and through a different kinetic mechanism. A conformational change is observed in the heparin-Robo1 interaction, but not in the heparin-Robo1-GFP interaction. Furthermore the GFP-tagged Robo1 requires a shorter (hexasaccharide) than the tag-free Robo1 (octadecasaccharide). These data demonstrate that GFP tagging can reduce the binding affinity of Robo1 to heparin and hinder heparin binding-induced Robo1 conformation change.  相似文献   

7.
Different interactions between particles of Escherichia coli capsule bacteriophage 29 and its receptor, the E. coli serotype 29 capsular polysaccharide have been studied. The inactivation of phage 29 (8 x 10(3) PFU/ml) by isolated host capsular glycan was found to be physiologically insignificant (50% inactivation dose equals 100 mug after 1 h at 37 C). No adsorption (less than 2 x 10(4) PFU/mug) of the viruses to K29 polysaccharide-coated erythroyctes (at 0 or 37 C) was observed either. The phage particles were, however, found to catalyze the hydrolysis of beta-D-glucosido-(1leads to 3)-D-glucuronic acid bonds (arrow) in the receptor polymer, leading, ultimately, to the formation of a mixture of K29 hexasaccharide (one repeating unit), dodecasaccharide, and octadecasaccharide: (see article). Testing derivatives of K29 polysaccharide, as well as 82 heterologous bacterial (mainly Enteriobactericeae) capsular glycans, the viral glycanase was found to be highly specific; in accordance with the host range of phage 29, only one enzymatic cross-reaction (with the Klebsiella K31 polysaccharide) was observed. These and previous results, as well as the electron optical findings of M. E. Bayer and H. Thurow (submitted for publication), are discussed in terms of a unifying mechanism of phage 29-host capsule interaction. We propose that the viruses penetrate the capsules by means of their spike-associated glycanase activity, which leads them along capsular polysaccharide strands to membrane-cell wall adhesions where ejection of the viral genomes occurs.  相似文献   

8.
Heparin and HS (heparan sulfate) exert their wide range of biological activities by interacting with extracellular protein ligands. Among these important protein ligands are various angiogenic growth factors and cytokines. HS binding to VEGF (vascular endothelial growth factor) regulates multiple aspects of vascular development and function through its specific interaction with HS. Many studies have focused on HS-derived or HS-mimicking structures for the characterization of VEGF165 interaction with HS. Using a heparinase 1-prepared small library of heparin-derived oligosaccharides ranging from hexasaccharide to octadecasaccharide, we systematically investigated the heparin-specific structural features required for VEGF binding. We report the apparent affinities for the association between the heparin-derived oligosaccharides with both VEGF165 and VEGF55, a peptide construct encompassing exclusively the heparin-binding domain of VEGF165. An octasaccharide was the minimum size of oligosaccharide within the library to efficiently bind to both forms of VEGF and a tetradecasaccharide displayed an effective binding affinity to VEGF165 comparable to unfractionated heparin. The range of relative apparent binding affinities among VEGF and the panel of heparin-derived oligosaccharides demonstrate that the VEGF binding affinity likely depends on the specific structural features of these oligosaccharides, including their degree of sulfation, sugar-ring stereochemistry and conformation. Notably, the unique 3-O-sulfo group found within the specific antithrombin binding site of heparin is not required for VEGF165 binding. These findings afford new insight into the inherent kinetics and affinities for VEGF association with heparin and heparin-derived oligosaccharides with key residue-specific modifications and may potentially benefit the future design of oligosaccharide-based anti-angiogenesis drugs.  相似文献   

9.
Thrombin-inhibitory activity of whale heparin oligosaccharides   总被引:1,自引:0,他引:1  
Whale heparin was partially digested with a purified heparinase and the oligosaccharide fractions with 8-20 monosaccharide units were isolated from the digest by gel filtration on Sephadex G-50, followed by affinity chromatography on a column of antithrombin III immobilized on Sepharose 4B. A marked difference in the inhibitory activity for thrombin in the presence of antithrombin III was observed between the high-affinity fractions for antithrombin III of octasaccharide approximately hexadecasaccharide and those of octadecasaccharide approximately eicosasaccharide. The disaccharide compositions of these hexadeca-, octadeca-, and eicosasaccharides were analyzed by high-performance liquid chromatography after digestion with a mixture of purified heparitinases 1 and 2 and heparinase. The analytical data indicated that the proportions of trisulfated disaccharide (IdUA(2S)alpha 1----4GlcNS(6S)) and disulfated disaccharide (UA1----4GlcNS(6S)) increased with the manifestation of high thrombin-inhibitory activity, while that of monosulfated disaccharide (UA1----4GlcNS) decreased. The present observations, together with those so far reported, suggest that the presence of the former structural elements, specifically IdUA(2S)alpha 1----4GlcNS(6S), as well as the antithrombin III-binding pentasaccharide at the proper positions in the molecules of whale heparin oligosaccharides is essential for the manifestation of high inhibitory activity for thrombin in the presence of antithrombin III. The structural bases for the manifestation of the anticoagulant activity of whale and porcine heparins and their oligosaccharides are also discussed.  相似文献   

10.
The minimum concentrations of heparin, dermatan sulfate, hirudin, and D-Phe-Pro-ArgCH2Cl required to delay the onset of prothrombin activation in contact-activated plasma also prolong the lag phases associated with both factor X and factor V activation. Heparin and dermatan sulfate prolong the lag phases associated with the activation of the three proteins by catalyzing the inhibition of endogenously generated thrombin. Thrombin usually activates factor V and factor VIII during coagulation. The smallest fragment of heparin able to catalyze thrombin inhibition by antithrombin III is an octadecasaccharide with high affinity for antithrombin III. In contrast, a dermatan sulfate hexasaccharide with high affinity for heparin cofactor II can catalyze thrombin inhibition by heparin cofactor II. A highly sulfated bis(lactobionic acid amide), LW10082 (Mr 2288), which catalyzes thrombin inhibition by heparin cofactor II and has both antithrombotic and anticoagulant activities, has been synthesized. In this study, we determined how the minimum concentration of LW10082 required to delay the onset of intrinsic prothrombin activation achieved this effect. We demonstrate that, like heparin and dermatan sulfate, LW10082 delays the onset of intrinsic prothrombin activation by prolonging the lag phase associated with both factor X and factor V activation. In addition, LW10082 is approximately 25% as effective as heparin and 10 times as effective as dermatan sulfate in its ability to delay the onset of prothrombin activation. The strong anticoagulant action of LW10082 is consistent with previous reports which show that the degree of sulfation is an important parameter for the catalytic effectiveness of sulfated polysaccharides on thrombin inhibition.  相似文献   

11.
Vacuolar-type rotary H(+)-ATPase/synthase (V(o)V(1)) from Thermus thermophilus, composed of nine subunits, A, B, D, F, C, E, G, I, and L, has been reconstituted from individually isolated V(1) (A(3)B(3)D(1)F(1)) and V(o) (C(1)E(2)G(2)I(1)L(12)) subcomplexes in vitro. A(3)B(3)D and A(3)B(3) also reconstituted with V(o), resulting in a holoenzyme-like complexes. However, A(3)B(3)D-V(o) and A(3)B(3)-V(o) did not show ATP synthesis and dicyclohexylcarbodiimide-sensitive ATPase activity. The reconstitution process was monitored in real time by fluorescence resonance energy transfer (FRET) between an acceptor dye attached to subunit F or D in V(1) or A(3)B(3)D and a donor dye attached to subunit C in V(o). The estimated dissociation constants K(d) for V(o)V(1) and A(3)B(3)D-V(o) were ~0.3 and ~1 nm at 25 °C, respectively. These results suggest that the A(3)B(3) domain tightly associated with the two EG peripheral stalks of V(o), even in the absence of the central shaft subunits. In addition, F subunit is essential for coupling of ATP hydrolysis and proton translocation and has a key role in the stability of whole complex. However, the contribution of the F subunit to the association of A(3)B(3) with V(o) is much lower than that of the EG peripheral stalks.  相似文献   

12.
Sulfide oxidation at halo-alkaline conditions in a fed-batch bioreactor   总被引:1,自引:0,他引:1  
A biotechnological process is described to remove hydrogen sulfide (H(2)S) from high-pressure natural gas and sour gases produced in the petrochemical industry. The process operates at halo-alkaline conditions and combines an aerobic sulfide-oxidizing reactor with an anaerobic sulfate (SO(4) (2-)) and thiosulfate (S(2)O(3) (2-)) reducing reactor. The feasibility of biological H(2)S oxidation at pH around 10 and total sodium concentration of 2 mol L(-1) was studied in gas-lift bioreactors, using halo-alkaliphilic sulfur-oxidizing bacteria (HA-SOB). Reactor operation at different oxygen to sulfide (O(2):H(2)S) supply ratios resulted in a stable low redox potential that was directly related with the polysulfide (S(x) (2-)) and total sulfide concentration in the bioreactor. Selectivity for SO(4) (2-) formation decreased with increasing S(x) (2-) and total sulfide concentrations. At total sulfide concentrations above 0.25 mmol L(-1), selectivity for SO(4) (2-) formation approached zero and the end products of H(2)S oxidation were elemental sulfur (S(0)) and S(2)O(3) (2-). Maximum selectivity for S(0) formation (83.3+/-0.7%) during stable reactor operation was obtained at a molar O(2):H(2)S supply ratio of 0.65. Under these conditions, intermediary S(x) (2-) plays a major role in the process. Instead of dissolved sulfide (HS(-)), S(x) (2-) seemed to be the most important electron donor for HA-SOB under S(0) producing conditions. In addition, abiotic oxidation of S(x) (2-) was the main cause of undesirable formation of S(2)O(3) (2-). The observed biomass growth yield under SO(4) (2-) producing conditions was 0.86 g N mol(-1) H(2)S. When selectivity for SO(4) (2-) formation was below 5%, almost no biomass growth was observed.  相似文献   

13.
Leaf intrinsic water-use efficiency (WUE), the ratio of photosynthetic rate to stomatal conductance (A/g(s) ), is a key plant trait linking terrestrial carbon and water cycles. A rapid, integrative proxy for A/g(s) is of benefit to crop breeding programmes aiming to improve WUE, but also for ecologists interested in plant carbon-water balance in natural systems. We hypothesize that the carbon isotope composition of leaf-respired CO(2) (δ(13) C(Rl) ), two hours after leaves are transferred to the dark, records photosynthetic carbon isotope discrimination and so provides a proxy for A/g(s) . To test this hypothesis, δ(13) C(Rl) was measured in four barley cultivars grown in the field at two levels of water availability and compared to leaf-level gas exchange (the ratio of leaf intercellular to ambient CO(2) partial pressure, C(i) /C(a) , and A/g(s) ). Leaf-respired CO(2) was more (13) C-depleted in plants grown at higher water availability, varied between days as environmental conditions changed, and was significantly different between cultivars. A strong relationship between δ(13) C(Rl) and δ(13) C of sucrose was observed. δ(13) C(Rl) was converted into apparent photosynthetic discrimination (Δ(13) C(Rl) ) revealing strong relationships between Δ(13) C(Rl) and C(i) /C(a) and A/g(s) during the vegetative stage of growth. We therefore conclude that δ(13) C(Rl) may provide a rapid, integrative proxy for A/g(s) in barley.  相似文献   

14.
Two new cyclic peptides, named sajaponicin C (1) and sajaponicin D (2), were isolated from the whole plants of Sagina japonica (Caryophyllaceae). Their structures were determined as cyclo(Pro(2)-Leu(2)-Tyr-Leu(1)-Phe(1)-Pro(3)-Phe(2)-Pro(1)) (1) and cyclo(Pro(1)-Pro(2)-Pro(3)-Pro(4)-Phe(1)-Gly-Thr-Ser-Phe(2)-Ile-Tyr) (2) on the basis of spectroscopic data, especially by two-dimensional (2D) NMR techniques.  相似文献   

15.
Classical benzodiazepines, for example diazepam, interact with alpha(x)beta(2)gamma(2) GABA(A) receptors, x = 1, 2, 3, 5. Little is known about effects of alpha subunits on the structure of the binding pocket. We studied here the interaction of the covalently reacting diazepam analog 7-Isothiocyanato-5-phenyl-1,3-dihydro-2H-1,4-benzodiazepin-2-one (NCS compound) with alpha(1)H101Cbeta(2)gamma(2) and with receptors containing the homologous mutation, alpha(2)H101Cbeta(2)gamma(2), alpha(3)H126Cbeta(2)gamma(2) and alpha(5)H105Cbeta(2)gamma(2). This comparison was extended to alpha(6)R100Cbeta(2)gamma(2) receptors as this mutation conveys to these receptors high affinity towards classical benzodiazepines. The interaction was studied at the ligand binding level and at the functional level using electrophysiological techniques. Results indicate that the geometry of alpha(6)R100Cbeta(2)gamma(2) enables best interaction with NCS compound, followed by alpha(3)H126Cbeta(2)gamma(2), alpha(1)H101Cbeta(2)gamma(2) and alpha(2)H101Cbeta(2)gamma(2), while alpha(5)H105Cbeta(2)gamma(2) receptors show little interaction. Our results allow conclusions about the relative apposition of alpha(1)H101 and homologous positions in alpha(2), alpha(3), alpha(5) and alpha(6) with the position occupied by -Cl in diazepam. During this study we found evidence for the presence of a novel site for benzodiazepines that prevents modulation of GABA(A) receptors via the classical benzodiazepine site. The novel site potentially contributes to the high degree of safety to some of these drugs. Our results indicate that this site may be located at the alpha/beta subunit interface pseudo-symmetrically to the site for classical benzodiazepines located at the alpha/gamma interface.  相似文献   

16.
The three-dimensional structure of a potent SSTR3-selective analogue of somatostatin, cyclo(3-14)H-Cys(3)-Phe(6)-Tyr(7)-D-Agl(8)(N(beta) Me, 2-naphthoyl)-Lys(9)-Thr(10)-Phe(11)-Cys(14)-OH (des-AA(1, 2, 4, 5, 12, 13)[Tyr(7), D-Agl(8)(N(beta) Me, 2-naphthoyl)]-SRIF) (peptide 1) has been determined by (1)H NMR in water and molecular dynamics (MD) simulations. The peptide exists in two conformational isomers differing mainly by the cis/trans isomerization of the side chain in residue 8. The structure of 1 is compared with the consensus structural motifs of other somatostatin analogues that bind predominantly to SSTR1, SSTR2/SSTR5 and SSTR4 receptors, and to the 3D structure of a non-selective SRIF analogue, cyclo(3-14)H-Cys(3)-Phe(6)-Tyr(7)-D-2Nal(8)-Lys(9)-Thr(10)-Phe(11)-Cys(14)-OH (des-AA(1, 2, 4, 5, 12, 13)[Tyr(7), D-2Nal(8)]-SRIF) (peptide 2). The structural determinant factors that could explain selectivity of peptide 1 for SSTR3 receptors are discussed.  相似文献   

17.
The synthesis of racemic and enantiomerically pure 3-butanamidoquinuclidines ((+/-)-Bu, (R)-Bu and (S)-Bu), (1-3) and 3-benzamidoquinuclidines ((+/-)-Bz, (R)-Bz, and (S)-Bz), (4-6) is described. The N-quaternary derivatives, N-benzyl-3-butanamidoquinuclidinium bromides ((+/-)-BnlBu, (R)-BnlBu and (S)-BnlBu), (7-9) and N-benzyl-3-benzamidoquinuclidinium bromides ((+/-)-BnlBz, (R)-BnlBz and (S)-BnlBz), (10-12) were subsequently synthesized. The interaction of the four enantiomerically pure quaternary derivatives with horse serum butyrylcholinesterase (BChE) was tested. All tested compounds inhibited the enzyme. The best inhibitior of the enzyme was (S)-BnlBz with a K(i) = 3.7 microM. The inhibitor potency decreases in order (S)-BnlBz > (R)-BnlBz > (R)-BnlBu > (S)-BnlBu.  相似文献   

18.
血管钠肽抑制异丙肾上腺素增强的大鼠心肌细胞钙瞬变   总被引:2,自引:0,他引:2  
Guo HT  Zhu MZ  Zhang RH  Bi H  Zhang B  Zhang HF  Yu J  Lu SY  Pei JM 《生理学报》2004,56(3):335-340
采用光谱荧光法研究血管钠肽(vasonatrin peptide,VNP)对心肌细胞内钙瞬变的作用及其机制,观察钠尿肽鸟苷酸环化酶(guanylate cyclase,GC)受体的特异性阻断剂(HS-142-1)、8-溴-环磷酸鸟苷(8-Br-cGMP)和镁蓝(methylene blue,MB)对心肌细胞内钙瞬变的影响。结果显示,异丙肾上腺素(isoproterenol,Iso)(10~(-10)~10~(-6)mol/L)可剂量依赖性地引起心肌细胞内钙瞬变增强,相对于对照组分别增强(13±8)%(P>0.05)、(26±13)%(P<0.05)、(66±10)%(P<0.01)、(150±10)%(P<0.01)和(300±25)%(P<0.01)。此效应可被β肾上腺素受体阻断剂普萘洛尔(10~(-6)mol/L)所阻断。VNP(10~(-10)~10~(-6)mol/L)可剂量依赖性地抑制Iso(10~(-8)mol/L)引起的心肌细胞内钙瞬变幅值的升高,相对于Iso(10~(-8)mol/L)分别减弱(99±3)%(P>0.05)、(96±2)%(P<0.05)、(84±6)%(P<0.01)、(66±3)%(P<0.01)和(62±3)%(P<0.01)。8-Br-cGMP(10~(-7)~10~(-3)mol/L)也可剂量依赖性地抑制Iso(10~(-8)mol/L)引起心肌细胞内钙瞬变的增强。HS-142-1(2×10~(-5)mol/L)使VNP的作用几乎完全消失。MB是GC的抑制剂,10~(-5)mol/L MB不但使VNP的作用完全消失,而且增强Iso对心肌细胞内钙瞬变的效应。VNP和HS-142-1本身对心肌细胞内钙瞬变无显著影响。而MB使心  相似文献   

19.
beta(2)-Microglobulin (beta(2)m) is non-covalently linked to the major histocompatibility complex (MHC) class I heavy chain and interacts with CD8 and Ly49 receptors. Murine MHC class I heavy chains can bind human beta(2)m (hbeta(2)m) and peptide, and such hybrid molecules are often used in structural and functional studies. The replacement of mouse beta(2)m (mbeta(2)m) with hbeta(2)m has several functional consequences for MHC class I complex stability and specificity, but the structural basis for this is presently unknown. To investigate the impact of species-specific beta(2)m subunits on MHC class I conformation, we provide a crystallographic comparison of H-2D(b) in complex with LCMV-derived gp33 peptide and either hbeta(2)m or mbeta(2)m. The conformation of the gp33 peptide is not affected by the beta(2)m species. Comparison of the interface between beta(2)m and the alpha(1)alpha(2) domains of the heavy chain in these two crystal structures reveals a marked increase in both polarity and number of hydrogen bonds between hbeta(2)m and the alpha(1)alpha(2) domains of H-2D(b). We propose that the positioning of two hydrogen bond rich regions at the hbeta(2)m/alpha(1)alpha(2) interface plays a central role in the increased overall stability and peptide exchange capacity in the H-2D(b)/hbeta(2)m complex. These two regions act as bridges, holding and stabilizing the underside of the alpha(1) and alpha(2) helices, enabling a prolonged peptide-receptive conformation of the peptide binding cleft. Furthermore, analysis of H-2D(b) in complex with either mbeta(2)m or hbeta(2)m provides a structural explanation for the differential binding of H-2D(b)/hbeta(2)m to both Ly49A and Ly49C. Our comparative structural study emphasizes the importance of beta(2)m residues at positions 3, 6 and 29 for binding to Ly49A and suggests that sterical hindrance by residue K6 on hbeta(2)m impairs the recognition of Ly49C by H-2D(b)/gp33/hbeta(2)m. Finally, comparison of the two H-2D(b) crystal structures implies that the beta(2)m species may affect the strength of TCR recognition by affecting CD8 binding.  相似文献   

20.
Ammonium D-glucarate, NH(4)(C(6)H(9)O(8)) [ammonium D-saccharate, NH(4)-SAC], has been synthesized, and its crystal structure solved by single-crystal X-ray diffraction methods. NH(4)-SAC crystallizes in the monoclinic space group P2(1) (#4) with cell parameters a = 4.8350(4) Angstroms, b = 11.0477(8) Angstroms, c = 16.7268(12) Angstroms, beta = 90.973(1) degrees, V = 894.34(12) Angstroms(3), Z = 3. The structure was refined by full-matrix least-squares on F(2) yielding final R-values (all data) R1 = 0.0353 and R(w)2 = 0.0870. The structure consists of alternating (NH(4))(+) and (C(6)H(11)O(6))(-) layers parallel to the bc plane. An extended network of N-H...O(SAC) and O(SAC)-H...O(SAC) hydrogen bonds provide the 3-D connectivity. The aqueous solubility (S(w)) has been shown to be pH independent at ambient conditions within the range 4.5 < pH < 10 with S(w) = 2.19 M/L, whose value is about a factor of two lower than that of the ammonium isosaccharate analogue.  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