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 0 毫秒
1.
Using a coarse-grained lipid and peptide model, we show that the free energy stabilization of amyloid-β in heterogeneous lipid membranes is predicted to have a dependence on asymmetric distributions of cholesterol compositions across the membrane leaflets. We find that a highly asymmetric cholesterol distribution that is depleted on the exofacial leaflet but enhanced on the cytofacial leaflet of the model lipid membrane thermodynamically favors membrane retention of a fully embedded Aβ peptide. However, in the case of cholesterol redistribution that increases concentration of cholesterol on the exofacial layer, typical of aging or Alzheimer’s disease, the free energy favors peptide extrusion of the highly reactive N-terminus into the extracellular space that may be vulnerable to aggregation, oligomerization, or deleterious oxidative reactivity.  相似文献   

2.
Using homonuclear 1H NOESY spectra, with chemical shifts, 3JHNHα scalar couplings, residual dipolar couplings, and 1H-15N NOEs, we have optimized and validated the conformational ensembles of the amyloid-β 1–40 (Aβ40) and amyloid-β 1–42 (Aβ42) peptides generated by molecular dynamics simulations. We find that both peptides have a diverse set of secondary structure elements including turns, helices, and antiparallel and parallel β-strands. The most significant difference in the structural ensembles of the two peptides is the type of β-hairpins and β-strands they populate. We find that Aβ42 forms a major antiparallel β-hairpin involving the central hydrophobic cluster residues (16–21) with residues 29–36, compatible with known amyloid fibril forming regions, whereas Aβ40 forms an alternative but less populated antiparallel β-hairpin between the central hydrophobic cluster and residues 9–13, that sometimes forms a β-sheet by association with residues 35–37. Furthermore, we show that the two additional C-terminal residues of Aβ42, in particular Ile-41, directly control the differences in the β-strand content found between the Aβ40 and Aβ42 structural ensembles. Integrating the experimental and theoretical evidence accumulated over the last decade, it is now possible to present monomeric structural ensembles of Aβ40 and Aβ42 consistent with available information that produce a plausible molecular basis for why Aβ42 exhibits greater fibrillization rates than Aβ40.  相似文献   

3.
Although the oligomers formed by Aβ peptides appear to be the primary cytotoxic species in Alzheimer's disease, detailed information about their structures appears to be lacking. In this article, we use exhaustive replica exchange molecular dynamics and an implicit solvent united-atom model to study the structural properties of Aβ monomers, dimers, and tetramers. Our analysis suggests that the conformational ensembles of Aβ dimers and tetramers are very similar, but sharply distinct from those sampled by the monomers. The key conformational difference between monomers and oligomers is the formation of β-structure in the oligomers occurring together with the loss of intrapeptide interactions and helix structure. Our simulations indicate that, independent of oligomer order, the Aβ aggregation interface is largely confined to the sequence region 10-23, which forms the bulk of interpeptide interactions. We show that the fractions of β structure computed in our simulations and measured experimentally are in good agreement.  相似文献   

4.
Five new lignan racemates, (±)-tatarinoid D–H (15), and three known analogues (68) were isolated from the rhizomes of Acorus tatarinowii. Their structures were established by 1D, 2D-NMR, HR-MS, IR and optical spectral data. Among them, 1 and 68 can alleviate the cognitive deterioration of Aβ transgenic flies.  相似文献   

5.
Fibril deposit formation of amyloid β-protein (Aβ) in the brain is a hallmark of Alzheimer's disease (AD). Increasing evidence suggests that toxicity is linked to diffusible Aβ oligomers, which have been found in soluble brain extracts of AD patients, rather than to insoluble fibers. Here we report a study of the toxicity of two distinct forms of recombinant Aβ small oligomers and fibrillar aggregates to simulate the action of diffusible Aβ oligomers and amyloid plaques on neuronal cells. Different techniques, including dynamic light scattering, fluorescence, and scanning electron microscopy, have been used to characterize the two forms of Aβ. Under similar conditions and comparable incubation times in neuroblastoma LAN5 cell cultures, oligomeric species obtained from Aβ peptide are more toxic than fibrillar aggregates. Both oligomers and aggregates are able to induce neurodegeneration by apoptosis activation, as demonstrated by TUNEL assay and Hoechst staining assays. Moreover, we show that aggregates induce apoptosis by caspase 8 activation (extrinsic pathway), whereas oligomers induce apoptosis principally by caspase 9 activation (intrinsic pathway). These results are confirmed by cytochrome c release, almost exclusively detected in the cytosolic fraction of LAN5 cells treated with oligomers. These findings indicate an active and direct interaction between oligomers and the cellular membrane, and are consistent with internalization of the oligomeric species into the cytosol.  相似文献   

6.
7.
Human serum albumin (HSA) is not only a fatty acid and drug carrier protein, it is also a potent inhibitor of Aβ self-association in plasma. However, the mechanism underlying the inhibition of Aβ fibrillization by HSA is still not fully understood. We therefore investigated the Aβ-HSA system using a combined experimental strategy based on saturation transfer difference (STD) NMR and intrinsic albumin fluorescence experiments on three Aβ peptides with different aggregation propensities (i.e., Aβ(12-28), Aβ(1-40), and Aβ(1-42)). Our data consistently show that albumin selectively binds to cross-β-structured Aβ oligomers as opposed to Aβ monomers. The HSA/Aβ oligomer complexes have KD values in the micromolar to submicromolar range and compete with the further addition of Aβ monomers to the Aβ assemblies, thus inhibiting fibril growth (“monomer competitor” model). Other putative mechanisms, according to which albumin acts as a “monomer stabilizer” or a “dissociation catalyst”, are not supported by our data, thus resolving previous discrepancies in the literature regarding Aβ-HSA interactions. In addition, the model and the experimental approaches proposed here are anticipated to have broad relevance for the characterization of other systems that involve amyloidogenic peptides and oligomerization inhibitors.  相似文献   

8.
PGLa and magainin 2 (MAG2) are amphiphilic antimicrobial peptides from frog skin with known synergistic activity. The orientation of the two helices in membranes was studied using solid-state 15N-NMR, for each peptide alone and for a 1:1 mixture of the peptides, in a range of different lipid systems. Two types of orientational behavior emerged. 1), In lipids with negative spontaneous curvature, both peptides remain flat on the membrane surface, when assessed both alone and in a 1:1 mixture. 2), In lipids with positive spontaneous curvature, PGLa alone assumes a tilted orientation but inserts into the bilayer in a transmembrane alignment in the presence of MAG2, whereas MAG2 stays on the surface or gets only slightly tilted, when observed both alone and in the presence of PGLa. The behavior of PGLa alone is identical to that of another antimicrobial peptide, MSI-103, in the same lipid systems, indicating that the curvature-dependent helix orientation is a general feature of membrane-bound peptides and also influences their synergistic intermolecular interactions.The two antimicrobial peptides PGLa and magainin 2 (MAG2) from the African frog Xenopus laevis, which are active against Gram-positive and Gram-negative bacteria, show intriguing synergistic effects that are not yet well understood (1). Structural insights into this synergy may help in the development of a new antibiotic-combination therapy. Both peptides are known to form α-helices when bound to lipid bilayers (2–4). The orientation of such α-helices in a membrane can be readily determined from the 15N-NMR chemical shift in oriented lipid bilayers that are aligned with the sample normal parallel to the external magnetic field (5). If the 15N chemical shift is ∼90 ppm, the peptide lies flat on the membrane surface (in the so-called S-state). On the other hand, when the 15N chemical shift is ∼200 ppm, the peptide is fully inserted (the I-state) in a transmembrane alignment. For intermediate orientations, where the peptide is tilted (the T-state) with an angle of typically 30°–60° relative to the membrane normal, intermediate chemical shifts are expected, but the exact tilt angle cannot be determined from a single label in such cases (6). Using several selectively 2H- or 19F-labeled peptide analogs, more exact orientations can be obtained, since both the tilt and the azimuthal angles can be measured with high accuracy, and valuable information about dynamics also can be deduced (3,7–10).The orientation of PGLa and MAG2 in membranes has been extensively studied with solid-state NMR, and some clues about the synergistic mechanism have been observed. Notably, in DMPC/DMPG membranes, it has been shown that the peptides on their own are in the S-state or T-state, but when the peptides are mixed in a 1:1 molar ratio, PGLa changes to the I-state, whereas MAG2 stays on the membrane surface (1,11,12). Thus, in the mixed system, transmembrane pores, which would not be spontaneously formed by each peptide on its own, appear to be stable, which could be the basis for synergy. On the other hand, it was also reported that in POPC/POPG there is no change in orientation when PGLa and MAG2 are mixed, as both peptides stay always in the S-state (12). This observation was attributed to the greater hydrophobic thickness of the POPC/POPG bilayer, compared to the DMPC/DMPG bilayer, suggesting that the PGLa helix is so short that it can only insert into thin DMPC/DMPG membranes. However, we have recently shown for MSI-103, a designer-made antimicrobial peptide based on the PGLa sequence, that the orientation determined by 2H-NMR depends not on the bilayer thickness but rather on the intrinsic spontaneous curvature of the lipids (13). Accordingly, an insertion of PGLa and MAG2 into POPC/POPG should be prevented by the pronounced negative spontaneous curvature induced by the unsaturated acyl chains. However, a simple comparison of only two lipid systems does not yield an answer as to which of these two hypotheses is correct. Therefore, we have now collected data over a wide range of lipid systems, with systematic variations of acyl chain lengths (to address bilayer thickness) as well as chain saturations (to address lipid curvature) (see Table S1 in the Supporting Material). In this way, we found unambiguously that lipid curvature is also the decisive factor in the insertion of PGLa/MAG2.Here, 15N-NMR spectra of singly labeled peptides were recorded, and for each liquid-crystalline lipid system we prepared four oriented samples: 15N-MAG2 alone, 15N-MAG2 with PGLa, 15N-PGLa with MAG2, and 15N-PGLa alone. The total peptide/lipid molar ratio (P/L) was 1:50. 15N-NMR spectra are shown in Fig. 1, and the chemical shifts are listed in Fig. 2. The quality of each oriented sample was checked with 31P-NMR (see Fig. S1). Our previous detailed 2H- and 19F-NMR analysis of PGLa in DMPC and DMPC/DMPG showed that the helix realigns depending on the peptide concentration. Namely, at low concentration, PGLa is in an S-state with a tilt angle of ∼98° (3), but above a threshold concentration around P/L = 1:100, it flips into a tilted T-state with a tilt angle of ∼125° (9,14). In the presence of MAG2, it was found that PGLa inserts almost upright in an I-state with a tilt angle of ∼158° (11). The present 15N-NMR study confirms that in DMPC/DMPG (3:1), PGLa is in the I-state when mixed with MAG2, as indicated by the 15N chemical shift of 205 ppm. PGLa alone at P/L = 1:50 has a chemical shift of 116 ppm, which corresponds to a tilted orientation, as expected. MAG2 alone is found to be in the S-state (91 ppm), but when it is mixed with PGLa its signal moves to 105 ppm, indicating a small change in the alignment. MAG2 is, however, clearly not inserted like PGLa.Open in a separate windowFigure 115N-NMR spectra of 15N-labeled PGLa or MAG2, alone or in a synergistic 1:1 mixture with the other peptide, in differently oriented lipids. Powder spectra are shown in the top row. Red, green, and blue lines indicate chemical shifts associated with the S-, T-, and I-state orientations, respectively.Open in a separate windowFigure 2Schematic overview of the orientation of PGLa (red), MAG2 (green), and MSI-103 (orange (13)) in different lipids. The corresponding 15N-NMR chemical shifts (in ppm) of the spectra in Fig. 1 are indicated beneath each peptide.In thin DLPC bilayers (12 carbon atoms in the chains) we see a behavior similar to that in DMPC (14 carbons), even though the exact chemical shifts are slightly different. PGLa alone is in the T-state, but in combination with MAG2, it flips into the I-state. MAG2, on the other hand, stays in the S-state with and without PGLa. In DPPC bilayers (16 carbons) also, the behavior is similar. PGLa alone is in the T-state but flips into the I-state in the presence of MAG2. MAG2 is slightly more tilted than in DMPC, but it never reaches the I-state.In contrast, in unsaturated lipids, both peptides are always in the S-state, both alone and in the presence of the synergistic partner. In POPC/POPG (9:1), the 15N chemical shifts of both PGLa and MAG2, alone and in the 1:1 mixture, are between 84 and 89 ppm, clearly indicating a flat alignment on the bilayer surface. As there are no changes in chemical shift with or without the other peptide, this could indicate that there are no interactions between them, in contrast to the situation in saturated lipids. Also, in thin DMoPC bilayers (with 14 carbon atoms and a double bond), the chemical shifts of all samples show that both peptides remain always in the S-state, whether alone or mixed.These results clearly demonstrate that the hydrophobic membrane thickness is not a critical factor for the insertion of PGLa in the presence of MAG2. In DMoPC (thinner than DMPC), there is no insertion, whereas in DPPC (thicker than POPC) insertion occurs. On the other hand, the results fully support the lipid-curvature hypothesis, which states that peptides remain on the surface in membranes composed of lipids with a negative spontaneous curvature, but are more easily tilted or inserted when the lipids have a positive spontaneous curvature (13).In a special lipid mixture, POPE/POPG/TOCL (72:23:5), often used to mimic the composition of the inner membrane of Escherichia coli (15), the result is practically the same as in POPC/POPG (9:1). Also here, chemical shifts of ∼84 ppm indicate that PGLa and MAG2 are always in the S-state, both alone and as a mixture. This behavior is in accordance with the curvature hypothesis, since PE and CL both have a strong negative curvature. On the other hand, when lyso-MPC is added to DMPC to increase the positive curvature, the chemical shift of MAG2 increases to 117 ppm, indicating a more tilted orientation in the membrane with enhanced curvature compared to DMPC or DMPC/DMPG, both with and without PGLa. PGLa alone gives a somewhat larger chemical shift but stays in the T-state, whereas PGLa together with MAG2 flips again into the I-state.We can now compare the results presented here with those from our previous study on the related peptide MSI-103 (13) to find strong correlations. Fig. 2 gives an overview of all results, illustrating the peptide orientations in the different lipid systems. PGLa on its own behaves just like MSI-103 and assumes the same S-state or T-state in the same systems, in full accordance with the lipid-curvature hypothesis. MAG2 alone behaves similarly but seems to have a higher concentration threshold to flip from the S-state to the T-state. In DMPC and DMPC/DMPG, where PGLa is already in the T-state, MAG2 is still in the S-state at P/L = 1:50. However, at P/L = 1:10 (Fig. S2), MAG2 has also reached the T-state. Since MAG2 is charged at both termini, whereas PGLa and MSI-103 are amidated and thus uncharged on the C terminus, it is indeed expected that MAG2 should not start to tilt as easily as PGLa or MSI-103. The polar sector of MAG2 is also larger (Fig. S3).When PGLa and MAG2 are mixed 1:1, their behavior correlates well with that of the individual peptides. In systems where PGLa and MSI-103 are in the S-state, the mixture of PGLa and MAG2 also remains in the S-state. Only when PGLa alone prefers the T-state does it get fully pushed into the I-state by the presence of MAG2. Thus, the model of MAG2-assisted insertion of PGLa proposed previously (12), which suggested that MAG2 would facilitate a thinning of the membrane such that PGLa would be able to insert into it, cannot be correct. We can instead conclude that only lipid systems that encourage peptide insertion per se show the MAG2-induced I-state of PGLa. The relationship between lipid shape and the tendency of peptides to insert into the membrane, as previously discussed (13), is illustrated in Fig. S4. Interestingly, common bacterial lipids like PE and CL have a negative spontaneous curvature and should thus not support peptide insertion and stable pores. However, pores could still be transient in native membranes, or other components like membrane proteins could influence the overall spontaneous curvature.In conclusion, we propose several criteria that encourage a peptide to insert from the surface-bound S-state more deeply into the membrane (i.e., into a T-state or I-state): 1), positive lipid spontaneous curvature, which is enhanced by large headgroups and ordered lipid chains (due to saturation, but also found at low temperatures close to the gel-to-liquid-crystalline phase transition); 2), a narrow polar sector and uncharged termini of the peptide; and 3), the presence of another peptide. The other peptide might have an indirect effect by changing the membrane properties via crowding. However, for PGLa/MAG2, a distinct synergistic activity has been demonstrated, indicating more specific interactions between these two peptides. The present 15N-NMR analysis shows that the two partner peptides are not aligned side-by-side as a dimer. Further solid-state NMR distance measurements will be required to clarify their detailed mode of assembly.  相似文献   

9.
Chromatographic, mass spectrometric and spectroscopic evidence has been obtained for four β-diketones occurring in the leaf waxes of some members of  相似文献   

10.
The effect of nonionic detergents of the n-alkyl-β-D-glucopyranoside class on the ordering of lipid bilayers and the dynamics of membrane-embedded peptides were investigated with 2H- and 31P-NMR. 1,2-dipalmitoyl-sn-glycero-3-phosphocholine was selectively deuterated at methylene segments C-2, C-7, and C-16 of the two fatty acyl chains. Two trans-membrane helices, WALP-19 and glycophorin A71-98, were synthesized with Ala-d3 in the central region of the α-helix. n-Alkyl-β-D-glucopyranosides with alkyl chains with 6, 7, 8, and 10 carbon atoms were added at increasing concentrations to the lipid membrane. The bilayer structure is retained up to a detergent/lipid molar ratio of 1:1. The insertion of the detergents leads to a selective disordering of the lipids. The headgroup region remains largely unaffected; the fatty acyl chain segments parallel to the detergent alkyl chain are only modestly disordered (10-20%), whereas lipid segments beyond the methyl terminus of the detergent show a decrease of up to 50%. The change in the bilayer order profile corresponds to an increase in bilayer entropy. Insertion of detergents into the lipid bilayers is completely entropy-driven. The entropy change accompanying lipid disorder is equivalent in magnitude to the hydrophobic effect. Ala-d3 deuterated WALP-19 and GlycA71-97 were incorporated into bilayers of 1,2-dimyristoyl-sn-glycero-3-phosphocholine at a peptide/lipid molar ratio of 1:100 and measured above the 1,2-dimyristoyl-sn-glycero-3-phosphocholine gel/liquid-crystal phase transition. Well-resolved 2H-NMR quadrupole splittings were observed for the two trans-membrane helices, revealing a rapid rotation of the CD3 methyl rotor superimposed on an additional rotation of the whole peptide around the bilayer normal. The presence of detergent fluidizes the membrane and produces magnetic alignment of bilayer domains but does not produce essential changes in the peptide conformation or dynamics.  相似文献   

11.
The β2-adrenergic receptor is an important member of the G-protein-coupled receptor (GPCR) superfamily, whose stability and function are modulated by membrane cholesterol. The recent high-resolution crystal structure of the β2-adrenergic receptor revealed the presence of possible cholesterol-binding sites in the receptor. However, the functional relevance of cholesterol binding to the receptor remains unexplored. We used MARTINI coarse-grained molecular-dynamics simulations to explore dimerization of the β2-adrenergic receptor in lipid bilayers containing cholesterol. A novel (to our knowledge) aspect of our results is that receptor dimerization is modulated by membrane cholesterol. We show that cholesterol binds to transmembrane helix IV, and cholesterol occupancy at this site restricts its involvement at the dimer interface. With increasing cholesterol concentration, an increased presence of transmembrane helices I and II, but a reduced presence of transmembrane helix IV, is observed at the dimer interface. To our knowledge, this study is one of the first to explore the correlation between cholesterol occupancy and GPCR organization. Our results indicate that dimer plasticity is relevant not just as an organizational principle but also as a subtle regulatory principle for GPCR function. We believe these results constitute an important step toward designing better drugs for GPCR dimer targets.  相似文献   

12.
It is thought that the pathological cascade in Alzheimer's disease is initiated by the formation of amyloid-β (Aβ) peptide complexes on cell membranes. However, there is considerable debate about the nature of these complexes and the type of solution-phase Aβ aggregates that may contribute to their formation. Also, it is yet to be shown that Aβ attaches strongly to living cell membranes, and that this can happen at low, physiologically relevant Aβ concentrations. Here, we simultaneously measure the aggregate size and fluorescence lifetime of fluorescently labeled Aβ1-40 on and above the membrane of cultured PC12 cells at near-physiological concentrations. We find that at 350 nM Aβ concentration, large (>>10 nm average hydrodynamic radius) assemblies of codiffusing, membrane-attached Aβ molecules appear on the cell membrane together with a near-monomeric species. When the extracellular concentration is 150 nM, the membrane contains only the smaller species, but with a similar degree of attachment. At both concentrations, the extracellular solution contains only small (∼2.3 nm average hydrodynamic radius) Aβ oligomers or monomers. We conclude that at near-physiological concentrations only the small oligomeric Aβ species are relevant, they are capable of attaching to the cell membrane, and they assemble in situ to form much larger complexes.  相似文献   

13.
Protein dynamics in cells may be different from those in dilute solutions in vitro, because the environment in cells is highly concentrated with other macromolecules. This volume exclusion because of macromolecular crowding is predicted to affect both equilibrium and kinetic processes involving protein conformational changes. To quantify macromolecular crowding effects on protein folding mechanisms, we investigated the folding energy landscape of an α/β protein, apoflavodoxin, in the presence of inert macromolecular crowding agents, using in silico and in vitro approaches. By means of coarse-grained molecular simulations and topology-based potential interactions, we probed the effects of increased volume fractions of crowding agents (ϕc) as well as of crowding agent geometry (sphere or spherocylinder) at high ϕc. Parallel kinetic folding experiments with purified Desulfovibro desulfuricans apoflavodoxin in vitro were performed in the presence of Ficoll (sphere) and Dextran (spherocylinder) synthetic crowding agents. In conclusion, we identified the in silico crowding conditions that best enhance protein stability, and discovered that upon manipulation of the crowding conditions, folding routes experiencing topological frustrations can be either enhanced or relieved. Our test-tube experiments confirmed that apoflavodoxin''s time-resolved folding path is modulated by crowding agent geometry. Macromolecular crowding effects may be a tool for the manipulation of protein-folding and function in living cells.  相似文献   

14.
15.
Amyloid-β (Aβ) peptides and other amyloidogenic proteins can form a wide range of soluble oligomers of varied morphologies at the early aggregation stage, and some of these oligomers are biologically relevant to the pathogenesis of Alzheimer's disease. Spherical micelle-like oligomers have been often observed for many different types of amyloids. Here, we report a hybrid computational approach to systematically construct, search, optimize, and rank soluble micelle-like Aβ25-35 structures with different side-chain packings at the atomic level. Simulations reveal for the first time, to our knowledge, that two Aβ micelles with antiparallel peptide organization and distinct surface hydrophobicity display high structural stability. Stable micelles experience a slow secondary structural transition from turn to α-helix. Energetic analysis coupled with computational mutagenesis reveals that van der Waals and solvation energies play a more pronounced role in stabilizing the micelles, whereas the electrostatic energies present a stable but minor energetic contribution to peptide assemblies. Modeled Aβ micelles with shapes and dimensions similar to those of experimentally derived spherical structures also provide detailed information about the roles of structural dynamics and transition in the formation of amyloid fibrils. The strong binding affinity of our micelles to antibodies implies that micelles may be a biologically relevant species.  相似文献   

16.
Deletion of the β-bulge trigger-loop results in both a switch in the preferred folding route, from the functional loop packing folding route to barrel closure, as well as conversion of the agonist activity of IL-1β into antagonist activity. Conversely, circular permutations of IL-1β conserve the functional folding route as well as the agonist activity. These two extremes in the folding-functional interplay beg the question of whether mutations in IL-1β would result in changes in the populations of heterogeneous folding routes and the signaling activity. A series of topologically equivalent water-mediated β-strand bridging interactions within the pseudosymmetric β-trefoil fold of IL-1β highlight the backbone water interactions that stabilize the secondary and tertiary structure of the protein. Additionally, conserved aromatic residues lining the central cavity appear to be essential for both stability and folding. Here, we probe these protein backbone-water molecule and side chain-side chain interactions and the role they play in the folding mechanism of this geometrically stressed molecule. We used folding simulations with structure-based models, as well as a series of folding kinetic experiments to examine the effects of the F42W core mutation on the folding landscape of IL-1β. This mutation alters water-mediated backbone interactions essential for maintaining the trefoil fold. Our results clearly indicate that this perturbation in the primary structure alters a structural water interaction and consequently modulates the population of folding routes accessed during folding and signaling activity.  相似文献   

17.
α/β-Galactoside α2,3-sialyltransferase produced by Photobacterium phosphoreum JT-ISH-467 is a unique enzyme that catalyzes the transfer of N-acetylneuraminic acid residue from cytidine monophosphate N-acetylneuraminic acid to acceptor carbohydrate groups. The enzyme recognizes both mono- and di-saccharides as acceptor substrates, and can transfer Neu5Ac to both α-galactoside and β-galactoside, efficiently. To elucidate the structural basis for the broad acceptor substrate specificity, we determined the crystal structure of the α2,3-sialyltransferase in complex with CMP. The overall structure belongs to the glycosyltransferase-B structural group. We could model a reasonable active conformation structure based on the crystal structure. The predicted structure suggested that the broad substrate specificity could be attributed to the wider entrance of the acceptor substrate binding site.  相似文献   

18.
19.
Aβ17-42 (so-called p3) amyloid is detected in vivo in the brains of individuals with Alzheimer's disease or Down's syndrome. We investigated the polymorphism of Aβ17-42 oligomers based on experimental data from steady-state NMR measurements, electron microscopy, two-dimensional hydrogen exchange, and mutational studies, using all-atom molecular-dynamics simulation with explicit solvent. We assessed the structural stability and the populations. Our results suggest that conformational differences in the U-turn of Aβ17-42 lead to polymorphism in β-sheet registration and retention of an ordered β-strand organization at the termini. Further, although the parallel Aβ17-42 oligomer organization is the most stable of the conformers investigated here, different antiparallel Aβ17-42 organizations are also stable and compete with the parallel architectures, presenting a polymorphic population. In this study we propose that 1), the U-turn conformation is the primary factor leading to polymorphism in the assembly of Aβ17-42 oligomers, and is also coupled to oligomer growth; and 2), both parallel Aβ17-42 oligomers and an assembly of Aβ17-42 oligomers that includes both parallel and antiparallel organizations contribute to amyloid fibril formation. Finally, since a U-turn motif generally appears in amyloids formed by full proteins or long fragments, and since to date these have been shown to exist only in parallel architectures, our results apply to a broad range of oligomers and fibrils.  相似文献   

20.
Rotation of the γ subunit of the F1-ATPase plays an essential role in energy transduction by F1-ATPase. Hydrolysis of an ATP molecule induces a 120° step rotation that consists of an 80° substep and 40° substep. ATP binding together with ADP release causes the first 80° step rotation. Thus, nucleotide binding is very important for rotation and energy transduction by F1-ATPase. In this study, we introduced a βY341W mutation as an optical probe for nucleotide binding to catalytic sites, and a βE190Q mutation that suppresses the hydrolysis of nucleoside triphosphate (NTP). Using a mutant monomeric βY341W subunit and a mutant α3β3γ subcomplex containing the βY341W mutation with or without an additional βE190Q mutation, we examined the binding of various NTPs (i.e., ATP, GTP, and ITP) and nucleoside diphosphates (NDPs, i.e., ADP, GDP, and IDP). The affinity (1/Kd) of the nucleotides for the isolated β subunit and third catalytic site in the subcomplex was in the order ATP/ADP > GTP/GDP > ITP/IDP. We performed van’t Hoff analyses to obtain the thermodynamic parameters of nucleotide binding. For the isolated β subunit, NDPs and NTPs with the same base moiety exhibited similar ΔH0 and ΔG0 values at 25°C. The binding of nucleotides with different bases to the isolated β subunit resulted in different entropy changes. Interestingly, NDP binding to the α3β(Y341W)3γ subcomplex had similar Kd and ΔG0 values as binding to the isolated β(Y341W) subunit, but the contributions of the enthalpy term and the entropy term were very different. We discuss these results in terms of the change in the tightness of the subunit packing, which reduces the excluded volume between subunits and increases water entropy.  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