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.
《Biophysical journal》2021,120(20):4399-4417
We used computational methods to analyze the mechanism of actin filament nucleation. We assumed a pathway where monomers form dimers, trimers, and tetramers that then elongate to form filaments but also considered other pathways. We aimed to identify the rate constants for these reactions that best fit experimental measurements of polymerization time courses. The analysis showed that the formation of dimers and trimers is unfavorable because the association reactions are orders of magnitude slower than estimated in previous work rather than because of rapid dissociation of dimers and trimers. The 95% confidence intervals calculated for the four rate constants spanned no more than one order of magnitude. Slow nucleation reactions are consistent with published high-resolution structures of actin filaments and molecular dynamics simulations of filament ends. One explanation for slow dimer formation, which we support with computational analysis, is that actin monomers are in a conformational equilibrium with a dominant conformation that cannot participate in the nucleation steps.  相似文献   

2.
通过分子表面的计算考察了三类金属硫蛋白(大鼠金属硫蛋白亚型Ⅱ,兔肝金属硫蛋白亚型Ⅰ和Ⅱ)二聚体短聚体中组成单元之间的疏水性相互作用。计算结果表明二聚体和三聚体中各组成单元之间均可以形成较好的几何匹配。对于二聚体而言,单体和单体之间存在一定的疏水性相互作用,但作用力 三聚体中,单体和二聚体之间的疏水残基能通过好的空间匹配形成很强的疏水性相互作用。对于这三种金属硫蛋白,二聚体中单体和单体之间的疏水性相  相似文献   

3.
J B Dietrich 《FEBS letters》1986,201(2):311-314
Chemical crosslinking was used for a direct analysis of the different forms of large tumor (T) antigen, the simian virus 40 A gene product. The first subclass, sedimenting at 14-16S, is composed of monomeric to tetrameric units, whereas the second, sedimenting at 5-6S, only contains dimers and monomers of T. The occurrence of oligomeric structures of T in solution which are higher than dimers suggests the possibility of direct binding of such trimers or tetramers to the origin of replication of the viral DNA as an alternative to the formation of these structures by aggregation of bound dimers or monomers after their sliding along the DNA.  相似文献   

4.
Stirring strongly enhanced irreversible inactivation and aggregation of lysozyme being studied as a model enzyme. From 0 to 740 rpm (equivalent to impeller tip speeds from 0 to 0.77 m s–1), the inactivation kinetic constant was proportional to the power imparted by the impeller. Collisions between inactive and native molecules induced inactivation of the latter and led to lysozyme aggregation. These fractal aggregates of lysozyme were made of monomers, dimers and trimers.  相似文献   

5.
Tu L  Deutsch C 《Biophysical journal》1999,76(4):2004-2017
Voltage-gated K+ channels are tetrameric, but how the four subunits assemble is not known. We analyzed inactivation kinetics and peak current levels elicited for a variety of wild-type and mutant Kv1.3 subunits, expressed singly, in combination, and as tandem constructs, to show that 1) the dominant pathway involves a dimerization of dimers, and 2) dimer-dimer interaction may involve interaction sites that differ from those involved in monomer-monomer association. Moreover, using nondenaturing gel electrophoresis, we detected dimers and tetramers, but not trimers, in the translation reaction of Kv1.3 monomers.  相似文献   

6.
Escherichia coli murein was specifically labeled with [14C]diaminopimelic acid in the mutant strains W7 (dap lysA) and BUG6. Pulse-labeled heat-denatured E. coli cells were digested with 2 mg of egg-white lysozyme per ml to degrade the murein completely and free any lipoprotein-bound muropeptide trimers, dimers, and monomers. Pulse-chase experiments showed that the relative percentage of trimers and dimers found in the newly synthesized murein increased somewhat with time at the expense of monomers. The increase in cross-links indicated that the radioactive monomers served as acceptors in multisite transpeptidations occurring after the labeling period. The content of nonreducing monomers (C7 and C8) remained unaltered, indicating that the oligosaccharide chain length did not change with time. A gradual conversion of the reducing disaccharide tetrapeptide monomer to its tripeptide analog occurred during chasing. Braun lipoprotein was linked to about 2% of the murein subunits within 30 s of the incorporation of subunits into insoluble murein, and after one-half a generation of chase, lipoprotein-associated muropeptides had approached the maximum (16% of the total murein subunits). The distribution of muropeptides was similar in lipoprotein-linked and lipoprotein-free murein, showing that the enzyme that links Braun lipoprotein to murein does not discriminate between monomers, dimers, and trimers. No evidence for a chasable, soluble polymer of murein was found in our experiments. Hence, our data support the idea that new murein is incorporated directly into the sacculus without first existing as a soluble intermediate.  相似文献   

7.
The montmorillonite-catalyzed reactions of the 5′-phosphorimidazolides of D, L-adenosine (D, L-ImpA) (Figure 1a. N = A, R = H) and D, L-uridine (Figure 1a., N = U, R = H) yields oligomers that were as long as 7 mers and 6 mers, respectively. The reactions of dilute solutions of D-ImpA and D-ImpU under the same conditions gave oligomers as long as 9 and 8 mers respectively. This demonstrated that oligomer formation is only partially inhibited by incorporation of both the D- and L-enantiomers. The structures of the dimers, trimers and tetramer fractions formed from D, L-ImpA was investigated by selective enzymatic hydrolysis, comparison with authentic samples and mass spectrometry. Homochiral products were present in greater amounts than would be expected if theoretical amounts of each were formed. The ratio of the proportion of homochiral products to that of the amount of each expected for the dimers (cyclic and linear), trimers and tetramers, was 1.3, 1.6, and 2.1, respectively. In the D, L-ImpU reaction homochiral products did not predominate with ratios of dimers (cyclic and linear), trimers and tetramers 0.8, 0.44, and 1.4, respectively. The proportions of cyclic dimers in the dimer fraction were 52–66% with D, L-ImpA and 44–69% with D, L-ImpU. No cyclic dimers were formed in the absence of montmorillonite. The differences in the reaction products of D, L-ImpA and D, L-ImpU are likely to be due to the difference in the orientations of the activated monomers when bound to the catalytic sites on montmorillonite. The consequences of the selectivity of montmorillonite as a prebiotic catalyst are discussed.  相似文献   

8.
Homodimeric triosephosphate isomerases from Trypanosoma cruzi (TcTIM) and Trypanosoma brucei (TbTIM) have markedly similar catalytic properties and 3-D structures; their overall amino acid sequence identity is 68% and 85% in their interface residues. Nonetheless, active dimer formation from guanidinium chloride unfolded monomers is faster and more efficient in TcTIM than in TbTIM. The enzymes thus provide a unique opportunity for exploring the factors that control the formation of active dimers. The kinetics of reactivation at different protein concentrations showed that the process involved three reactions: monomer folding, association of folded monomers, and a transition from inactive to active dimers. The rate constants of the reactions indicated that, at relatively low protein concentrations, the rate-limiting step of reactivation was the association reaction; at high protein concentrations the transition of inactive to active dimers was rate limiting. The rates of the latter two reactions were higher in TcTIM than in TbTIM. Studies with a mutant of TcTIM that had the interface residues of TbTIM showed that the association rate constant was similar to that of TbTIM. However, the rate of the transition from inactive to active dimers was close to that of TcTIM; thus, this transition depends on the noninterfacial portion of the enzymes. When unfolded monomers of TcTIM and TbTIM were allowed to reactivate together, TcTIM, the hybrid, and TbTIM were formed in a proportion of 1:0.9:0.2. This distribution suggests that, in the hybrid, the characteristics of the TcTIM monomers influence the properties of TbTIM monomers.  相似文献   

9.
10.
The exposure of the flavivirus tick-borne encephalitis (TBE) virus to an acidic pH is necessary for virus-induced membrane fusion and leads to a quantitative and irreversible conversion of the envelope protein E dimers to trimers. To study the structural requirements for this oligomeric rearrangement, the effect of low-pH treatment on the oligomeric state of different isolated forms of protein E was investigated. Full-length E dimers obtained by solubilization of virus with the detergent Triton X-100 formed trimers at low pH, whereas truncated E dimers lacking the stem-anchor region underwent a reversible dissociation into monomers without forming trimers. These data suggest that the low-pH-induced rearrangement in virions is a two-step process involving a reversible dissociation of the E dimers followed by an irreversible formation of trimers, a process which requires the stem-anchor portion of the protein. This region contains potential amphipathic alpha-helical and conserved structural elements whose interactions may contribute to the rearrangements which initiate the fusion process.  相似文献   

11.
The major covalently linked multimolecular D fragments found in plasmic digests of factor XIIIa cross-linked fibrin formed under physiological pH and ionic strength conditions consist of D dimers, D trimers, and D tetramers. These fragments are linked by epsilon-amino-gamma-glutamyllysine bonds in the carboxy-terminal regions of their gamma chains, which had originated in the cross-linked fibrin as gamma dimers, gamma trimers, and gamma tetramers, respectively. In this study, factors affecting the degree and rate of formation of these three classes of cross-linked gamma chains were determined by analyzing the D-fragment content of plasmic digests of cross-linked fibrin that had been sampled after all gamma-chain monomers had been consumed in the cross-linking process. D trimers and D tetramers, expressed as a proportion of the total D-fragment content, both increased at the expense of the D-dimer population as a function of increasing factor XIII concentration, the time of cross-linking, or the CaCl2 concentration. Their levels decreased as the ionic strength was raised by NaCl addition. However, the ionic strength effect could be reversed by concomitantly raising the CaCl2 concentration. Digests of clots prepared from recalcified fresh citrated plasma also contained each type of cross-linked D fragment, and the proportion of D trimers and D tetramers in the digest increased with increasing clot incubation time. These results indicate that gamma-trimer and gamma-tetramer formation is a dynamic physiological process.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

12.
The serine protease inhibitor (serpin) family can readily form long-chain polymers by a process that underlies a variety of diseases. We show here that monomers of plasma serpins α1-antitrypsin and antithrombin are stable on incubation with the rate-limiting step in their polymerisation being the formation of the initial dimer. Once formed, the dimers readily interlink to form tetramers and can bind monomers to form trimers and longer oligomers. Cleavage of the only exposed reactive loop, in unit I of the dimers, prevents their interlinkage, but these cleaved dimers can still link to monomers. The rapid binding by the cleaved dimers of a peptide specific to the lower half of β-sheet A of the molecule indicates the ready opening of this β-sheet in unit II of the dimers. The failure of the cleaved dimers to bind peptide-complexed monomers, together with the relative inaccessibility of the P14 hinge residue in the oligomers, is evidence that partial insertion of the reactive loop into its own A-sheet is required for polymer formation. We propose that serpin dimers initiate and propagate polymerisation by having one exposed loop with an optimal conformation as a β-strand donor and a readily opened β-sheet as an acceptor. The sequential reformation of these activated β-interfaces as the oligomer extends, molecule by molecule, provides a model for the fibril and amyloid formation of conformational diseases in general as well as for the infectivity of prion encephalopathies.  相似文献   

13.
A homohexameric molecule of Escherichia coli pyrophosphatase is arranged as a dimer of trimers, with an active site present in each of its six monomers. Earlier we reported that substitution of His(136) and His(140) in the intertrimeric subunit interface splits the molecule into active trimers (Velichko, I. S., Mikalahti, K., Kasho, V. N., Dudarenkov, V. Y., Hyyti?, T., Goldman, A., Cooperman, B. S., Lahti, R., and Baykov, A. A. (1998) Biochemistry 37, 734-740). Here we demonstrate that additional substitutions of Tyr(77) and Gln(80) in the intratrimeric interface give rise to moderately active dimers or virtually inactive monomers, depending on pH, temperature, and Mg(2+) concentration. Successive dissociation of the hexamer into trimers, dimers, and monomers progressively decreases the catalytic efficiency (by 10(6)-fold in total), and conversion of a trimer into dimer decreases the affinity of one of the essential Mg(2+)-binding sites/monomer. Disruptive substitutions predominantly in the intratrimeric interface stabilize the intertrimeric interface and vice versa, suggesting that the optimal intratrimeric interaction is not compatible with the optimal intertrimeric interaction. Because of the resulting "conformational strain," hexameric wild-type structure appears to be preformed to bind substrate. A hexameric triple variant substituted at Tyr(77), Gln(80), and His(136) exhibits positive cooperativity in catalysis, consistent with this model.  相似文献   

14.
The state of oligomerization of the tumor suppressor p53 is an important factor in its various biological functions. It has a well-defined tetramerization domain, and the protein exists as monomers, dimers and tetramers in equilibrium. The dissociation constants between oligomeric forms are so low that they are at the limits of measurement by conventional methods in vitro. Here, we have used the high sensitivity of single-molecule methods to measure the equilibria and kinetics of oligomerization of full-length p53 and its isolated tetramerization domain, p53tet, at physiological temperature, pH and ionic strength using fluorescence correlation spectroscopy (FCS) in vitro. The dissociation constant at 37 °C for tetramers dissociating into dimers for full-length p53 was 50 ± 7 nM, and the corresponding value for dimers into monomers was 0.55 ± 0.08 nM. The half-lives for the two processes were 20 and 50 min, respectively. The equivalent quantities for p53tet were 150 ± 10 nM, 1.0 ± 0.14 nM, 2.5 ± 0.4 min and 13 ± 2 min. The data suggest that unligated p53 in unstressed cells should be predominantly dimeric. Single-molecule FCS is a useful procedure for measuring dissociation equilibria, kinetics and aggregation at extreme sensitivity.  相似文献   

15.
Mild nuclease digestion of rat liver chromatin generates particles with sedimentation coefficients of about 33S, 60S, and 90S (in 50 mM NaCl). The kinetics of appearance and disappearance of these particles with progressive digestion suggest that they are produced by cleavage from a higher order repeat structure, the 33S particle representing the monomer. At an intermediate stage of digestion, about 75 % of the nuclear chromatin can be recovered as monomers to trimers of this higher order structure. Sedimentation profiles indicate that monomer particles containing 7–8 nucleosomes occur at the highest frequency. The DNA fragments in monomers have a size corresponding to hepta- and octanucleosomes, and those in dimers have a size corresponding to chains of sixteen nucleosomes. The higher order repeat structure is only stable between 30 and 200 mM NaCl; the particles unfold below 30 and above 200 mM NaCl. When examined by electron microscopy, monomers and dimers appear as compact globular structures. Relaxation by lowering the salt concentration results in the appearance of polynucleosomes with a chain length of eight beads in the monomer and sixteen in the dimer particle. These results indicate that the unit particle of the higher order repeat structure of rat liver chromatin contains eight nucleosomes.  相似文献   

16.
The urea-induced inactivation and dissociation of catalytically active hexamer of glutamate dehydrogenase (L-glutamate-NAD(P)-oxidoreductase, EC 1.4.1.3) from bovine liver were studied using radioactive phosphopyridoxyl derivative of the enzyme immobilized on cyanogen bromide-activated Sepharose CL-4B. It is shown that at neutral pH (7.0-7.8) urea causes dissociation of glutamate dehydrogenase to directly yield catalytically inactive immobilized monomers rather than hexamer's stable fragments at the same time. At pH 8.9 or 5.6 the urea-induced is accompanied by the formation of conformationally stable immobilized dimers or trimers, respectively. The trimers are catalytically active, whereas the dimers did not exhibit any enzymatic activity. The data obtained led to suggestion that the hexamer consists of three either equivalent dimers (3 alpha 2) or of two equivalent trimers (2 alpha 3).  相似文献   

17.
The vertebrate sodium (Nav) channel is composed of an ion-conducting α subunit and associated β subunits. Here, we report the crystal structure of the human β3 subunit immunoglobulin (Ig) domain, a functionally important component of Nav channels in neurons and cardiomyocytes. Surprisingly, we found that the β3 subunit Ig domain assembles as a trimer in the crystal asymmetric unit. Analytical ultracentrifugation confirmed the presence of Ig domain monomers, dimers, and trimers in free solution, and atomic force microscopy imaging also detected full-length β3 subunit monomers, dimers, and trimers. Mutation of a cysteine residue critical for maintaining the trimer interface destabilized both dimers and trimers. Using fluorescence photoactivated localization microscopy, we detected full-length β3 subunit trimers on the plasma membrane of transfected HEK293 cells. We further show that β3 subunits can bind to more than one site on the Nav 1.5 α subunit and induce the formation of α subunit oligomers, including trimers. Our results suggest a new and unexpected role for the β3 subunits in Nav channel cross-linking and provide new structural insights into some pathological Nav channel mutations.  相似文献   

18.
Garai K  Baban B  Frieden C 《Biochemistry》2011,50(13):2550-2558
The apolipoprotein apoE plays a key role in cholesterol and lipid metabolism. There are three isoforms of this protein, one of which, apoE4, is the major risk factor for Alzheimer's disease. At micromolar concentrations all lipid-free apoE isoforms exist primarily as monomers, dimers, and tetramers. However, the molecular weight form of apoE that binds to lipid has not been clearly defined. We have examined the role of self-association of apoE with respect to interactions with phospholipids. Measurements of the time dependence of turbidity clearance of small unilamellar vesicles of dimyristoyl-sn-glycero-3-phosphocholine (DMPC) upon addition of apoE show that higher molecular weight oligomers bind poorly if at all. The kinetic data can be described by a reaction model in which tetramers and dimers of apoE must first dissociate to monomers which then bind to the liposome surface in a fast and reversible manner. A slow but not readily reversible conformational conversion of the monomer then occurs. Prior knowledge of the rate constants for the association-dissociation process allows us to determine the rate constant of the conformational conversion. This rate constant is isoform dependent and appears to correlate with the stability of the apoE isoforms with the rate of dissociation of the apoE oligomers to monomers being the rate-limiting process for lipidation. Differences in the lipidation kinetics between the apoE isoforms arise from their differences in the self-association behavior leading to the conclusion that self-association behavior may influence biological functions of apoE in an isoform-dependent manner.  相似文献   

19.
The envelope glycoprotein of human immunodeficiency virus type 1 (HIV-1) is thought to exist on the virion surface as a trimer of non-covalently associated gp120/gp41 molecules. We expressed trimeric envelope glycoprotein from three primary, macrophage tropic HIV-1 isolates in baby hamster kidney cells and analyzed the furin-mediated cleavage, stability, and receptor binding properties of the oligomers. The envelope glycoprotein was secreted in a soluble form deleted of its transmembrane anchor and the intracytoplasmic domain (gp140). A mixture of trimers, dimers, and monomers of gp140 as well as monomeric gp120 was detected on polyacrylamide gels. Analysis by sucrose gradient centrifugation revealed that trimers and dimers were essentially composed of uncleaved gp140, whereas most of the gp120 was found in the monomeric fraction. To analyze the effect of the cleavage of gp140 to gp120/Delta41 on trimerization, we co-expressed the furin protease along with gp140. Surprisingly, furin expression changed the subcellular localization of the envelope glycoprotein, which became in majority sequestered in the major furin compartment, the trans-Golgi network, as judged by confocal laser microscopy. The envelope glycoprotein secreted from furin-co-expressing cells was almost completely cleaved to gp120 and Deltagp41, but gp120 was found exclusively in the monomeric fraction, with a few residual oligomers being composed of uncleaved gp140. Secreted uncleaved gp140 trimers were purified to homogeneity and analyzed for their capacity to interact with cellular receptors CD4 and CC chemokine receptor 5 (CCR5). Receptor binding was analyzed on CD4- and CCR5-expressing cells as well as on peripheral blood mononuclear cells. Trimers showed greatly reduced binding to CD4 as compared with monomers. Neither monomers nor trimers bound directly to CCR5. In conclusion, our results show that the cleaved form of the envelope glycoprotein does not form stable trimers, suggesting that gp120/gp41 oligomers on the virion surface might be stabilized by a yet to be identified mechanism and that the virion might attach to CD4 via a monomeric form of gp120. These results are relevant to the development of an envelope-based vaccine against AIDS.  相似文献   

20.
Kwon C  Lee S  Jung S 《Carbohydrate research》2011,(14):4071-2314
Low-molecular-weight (LMW) succinoglycans (monomers, dimers, and trimers) were isolated from Sinorhizobium meliloti 1021 and have been firstly investigated by matrix-assisted laser desorption/ionization time-of-flight mass spectrometry (MALDI-TOF MS) using 2,4,6-trihydroxyacetophenone (THAP) as an optimal matrix in the negative ion mode. The main fractions of LMW succinoglycans contain molecules assembled of octasaccharide subunits. MALDI-TOF mass spectra of the LMW succinoglycan monomers, the dimers, and the trimers showed the daughter ions resulting from the losses of the terminal galactose residues at the reducing ends, clearly indicating that the galactosyl linkages are more labile than the other glucosyl linkages. Furthermore, the losses of the acetyl groups as substituents rather than the succinyl and pyruvyl ester linkages by prompt fragmentation primarily occurred during MALDI-TOF analysis, suggesting the greater instability of acetyl linkages compared to pyruvyl and succinyl linkages.  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