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.
It is proposed that staphylococcal delta-lysin, a membrane-damaging peptide of 26 residues adopts an α-helical rod-like configuration with separate hydrophilic and hydrophobic faces. Association of six such monomers in a cell membrane may result in the formation of a transmembrane “pore” lined by the hydrophilic faces of the monomers.  相似文献   

2.
Abstract

The amyloid β-peptides (Aßs) are the main protein components of amyloid deposits in Alzheimer's disease (AD). Detailed knowledge of the structure and assembly dynamics of Aß is important for the development of properly targeted AD therapeutics. So far, the process of the monomeric Aß assembling into oligomeric fibrils and the mechanism underlying the aggregation process remain unclear. In this study, several molecular dynamics simulations were conducted to investigate the aggregation behaviors of the Aß(17–42) oligomers associated with various numbers of monomers (dimer, trimer, tetramer, and pentamer). Our results showed that the structural stability of the Aß(17–42) oligomers increases with increasing the number of monomer. We further demonstrated that the native hydrophobic contacts are positive correlated with the ß-sheet contents, indicating that hydrophobic interaction plays an important role in maintaining the structural stability of the Aß(17–42) oligomers, particularly for those associated with more monomers. Our results also showed that the stability of the C-terminal hydrophobic segment 2 (residues 30–42) is higher than that of the N-terminal hydrophobic segment 1 (residues 17–21), suggesting that hydrophobic segment 2 may act as the nucleation site for aggregation. We further identified that Met35 residue initiates the hydrophobic interactions and that the intermolecular contact pairs, Gly33-Gly33 and Gly37-Gly37, form a stable “molecular notch”, which may mediate the packing of the ß-sheet involving many other hydrophobic residues during the early stage of amyloid-like fibril formation.  相似文献   

3.
Many protein architectures exhibit evidence of internal rotational symmetry postulated to be the result of gene duplication/fusion events involving a primordial polypeptide motif. A common feature of such structures is a domain‐swapped arrangement at the interface of the N‐ and C‐termini motifs and postulated to provide cooperative interactions that promote folding and stability. De novo designed symmetric protein architectures have demonstrated an ability to accommodate circular permutation of the N‐ and C‐termini in the overall architecture; however, the folding requirement of the primordial motif is poorly understood, and tolerance to circular permutation is essentially unknown. The β‐trefoil protein fold is a threefold‐symmetric architecture where the repeating ~42‐mer “trefoil‐fold” motif assembles via a domain‐swapped arrangement. The trefoil‐fold structure in isolation exposes considerable hydrophobic area that is otherwise buried in the intact β‐trefoil trimeric assembly. The trefoil‐fold sequence is not predicted to adopt the trefoil‐fold architecture in ab initio folding studies; rather, the predicted fold is closely related to a compact “blade” motif from the β‐propeller architecture. Expression of a trefoil‐fold sequence and circular permutants shows that only the wild‐type N‐terminal motif definition yields an intact β‐trefoil trimeric assembly, while permutants yield monomers. The results elucidate the folding requirements of the primordial trefoil‐fold motif, and also suggest that this motif may sample a compact conformation that limits hydrophobic residue exposure, contains key trefoil‐fold structural features, but is more structurally homologous to a β‐propeller blade motif.  相似文献   

4.
The first representatives of proteins of the macroglobulin family appeared 500–700 million years ago. At present representatives of this family have been revealed in crustaceans, molluscs, fish, amphibians, reptiles, ticks, insects, birds, and mammals, the macroglobulin family in blood of some species being represented simultaneously by several proteins that have different molecular weight and partly duplicate functions of each other. In different species, they are present as monomers, dimers, or tetramers. The distinguishing feature of each protein subunit is the presence of a “trap” with cyclic thioether on the bottom and of a sufficiently large hydrophobic area. All representatives are able to form complexes with different regulatory substances through covalent or hydrophobic bonds, which allows them to perform a wide range of regulatory functions. The ancient origin, evolutionary conservatism, widespread presence, and a diversity of regulatory functions permit proteins of the macroglobulin family to be considered as the main regulatory biomolecules of organism fluid media.  相似文献   

5.
The importance of diadinoxanthin (Ddx) de‐epoxidation in the short‐term modulation of the temperature effect on photosynthetic membranes of the diatom Phaeodactylum tricornutum was demonstrated by electron paramagnetic resonance (EPR), Laurdan fluorescence spectroscopy, and high‐performance liquid chromatography. The 5‐SASL spin probe employed for the EPR measurements and Laurdan provided information about the membrane area close to the polar head groups of the membrane lipids, whereas with the 16‐SASL spin probe, the hydrophobic core, where the fatty acid residues are located, was probed. The obtained results indicate that Ddx de‐epoxidation induces a two component mechanism in the short‐term regulation of the membrane fluidity of diatom thylakoids during changing temperatures. One component has been termed the “dynamic effect” and the second the “stable effect” of Ddx de‐epoxidation. The “dynamic effect” includes changes of the membrane during the time course of de‐epoxidation whereas the “stable effect” is based on the rigidifying properties of Dtx. The combination of both effects results in a temporary increase of the rigidity of both peripheral and internal parts of the membrane whereas the persistent increase of the rigidity of the hydrophobic core of the membrane is solely based on the “stable effect.”  相似文献   

6.
Purified preparations of the recombinant b′x domain fragment of human protein‐disulphide isomerase (PDI), which are homogeneous by mass spectrometry and sodium dodecyl sulfate polyacrylamide gel electrophoresis, comprise more than one species when analyzed by ion‐exchange chromatography and nondenaturing polyacrylamide gel electrophoresis. These species were resolved and shown to be monomer and dimer by analytical ultracentrifugation and analytical size‐exclusion chromatography. Spectroscopic properties indicate that the monomeric species corresponds to the “capped” conformation observed in the x‐ray structure of the I272A mutant of b′x (Nguyen, Wallis, Howard, Haapalainen, Salo, Saaranen, Sidhu, Wierenga, Freedman, Ruddock, and Williamson, J Mol Biol 2008;383:1144‐1155) in which the x region binds to a hydrophobic patch on the surface of the b′ domain; conversely, the dimeric species has an “open” or “uncapped” conformation in which the x region does not bind to this surface. The larger bb′x fragment of human PDI shows very similar behavior to b′x and can be resolved into a capped monomeric species and an uncapped dimer. Preparations of recombinant b′ domain of human PDI and of the bb′ domain pair are found exclusively as dimers. Full‐length PDI is known to comprise a mixture of monomeric and dimeric species, whereas the isolated a , b , and a′ domains of PDI are found exclusively as monomers. These results show that the b′ domain of human PDI tends to form homodimers—both in isolation and in other contexts—and that this tendency is moderated by the adjacent x region, which can bind to a surface patch on the b′ domain.  相似文献   

7.
Ion channels from sheep cardiac mitoplast (inverted inner mitochondrial membrane vesicle) preparations were incorporated into voltage-clamped planar lipid bilayers. A low-conductance anion channel (~40 or ~85 pS in symmetric 300 or 550 mM choline Cl, respectively), characterized by the presence of two well-defined substates, at ~25 and ~50% of the fully open level, was studied in detail. The substate behavior was consistent with a multibarrelled channel containing four functionally coupled pores. At negative (cis-trans) membrane potentials, the putative portomers appeared to gate with substantial positive cooperativity, accounting for the apparent absence of a ~75% sublevel. At positive holding potentials, allosteric protomer interactions were more complicated, and the channel complex could be modeled as a dimer of dimers. The protochannels in one dimer (“dimer A”) appeared to open independently of each other, and with a relatively high probability, while the monomers comprising the second dimer (“dimer B”) were functionally coupled, could only open if both protomers in dimer A were open, and closed as soon as one of the monomers in dimer A shut. The channels also displayed Ca2+- (and Mg2+-) sensitive rectification related to bilayer lipid surface charge. By assuming that Ca2+ acted solely by screening surface charge, the membrane surface potential profile was used as a “microscopic ruler” to place one mouth of the channel within 10–11 Å of the bilayer surface.  相似文献   

8.
A library of 17 nanoparticles made of acrylate and methacrylate copolymers is prepared, characterized, and screened against six epitopes of adeno‐associated viruses (AAV)‐neutralizing antibodies to assess their affinity and specificity. Peptide epitopes are immobilized onto the surface of glass beads, packed in filtration microplates, and incubated with fluorescein‐labelled nanoparticles. Following intense washing, the affinity of nanoparticles to immobilized epitopes is assessed by measuring the fluorescence of captured nanoparticles. The results show that polar monomers, acrylic acid in particular, have a positive impact on polymer affinity towards all peptides used in this study. The presence of hydrophobic monomers, on other hand, has a negative impact on polymer binding. The composition of peptides used in this study has no noticeable impact on the affinity of synthesized nanoparticles. The affinity of nanoparticles with the highest affinity to peptide targets does not exceed millimolar level. Overall, it is found that the synthesized library showed modest affinity but lacked specificity, which should be further “tuned,” for example, by using molecular imprinting to achieve an acceptable level of affinity and specificity for practical application.  相似文献   

9.
Artificial ribonucleases of AnBCL series were synthesized by solid‐phase method. They consist of a hydrophobic alkyl radical A (n = 3–12 carbon atoms), an “RNA‐binding domain” B (bisquaternary salt of 1,4‐diazabicyclo[2.2.2]octane), a “catalytic domain” C (histidine residue) and a “linker” L that joins the domains B and C. The effect of the alkyl radical on the catalytic properties of the chemical catalyst was studied using three activated phosphate ester substrates: p‐nitrophenyl phosphate, bis‐p‐nitrophenyl phosphate, and thymidine‐3′‐p‐nitrophenyl phosphate.  相似文献   

10.
Abstract

Attempts at synthesis of “chimeric” oligonucleotide constructs (PO/PS-Oligos) possessing phosphate and P-stereodefined phosphorothioate internucleotide linkages via combined phosphoramidite/oxathiaphospholane methods were unsuccessful. Therefore, novel monomers for oxathiaphospholane method, namely 5′-O-DMT-deoxyribonucleoside 3′-O(2-oxo-.spiro-4.4-pentamethylene-1.3.2-oxathiaphospholane)s, were prepared and used together with their diastereomerically pure 2-thio analogues for the stereocontrolled synthesis of “chimeric” oligonucleotide constructs (PO/PS-Oligos).  相似文献   

11.
《Biophysical journal》2021,120(23):5309-5321
Gramicidin A (gA) is a hydrophobic pentadecapeptide readily incorporating into a planar bilayer lipid membrane (BLM), thereby inducing a large macroscopic current across the BLM. This current results from ion-channel formation due to head-to-head transbilayer dimerization of gA monomers with rapidly established monomer-dimer equilibrium. Any disturbance of the equilibrium, e.g., by sensitized photoinactivation of a portion of gA monomers, causes relaxation toward a new equilibrium state. According to previous studies, the characteristic relaxation time of the gA-mediated electric current decreases as the current increases upon elevating the gA concentration in the membrane. Here, we report data on the current relaxation kinetics for gA analogs with N-terminal valine replaced by glycine or tyrosine. Surprisingly, the relaxation time increased rather than decreased upon elevation of the total membrane conductance induced by these gA analogs, thus contradicting the classical kinetic scheme. We developed a general theoretical model that accounts for lateral interaction of monomers and dimers mediated by membrane elastic deformations. The modified kinetic scheme of the gramicidin dimerization predicts the reverse dependence of the relaxation time on membrane conductance for gA analogs, with a decreased dimerization constant that is in a good agreement with our experimental data. The equilibration process may be also modulated by incorporation of other peptides (“impurities”) into the lipid membrane.  相似文献   

12.
Elastin self-assembles from monomers into polymer networks that display elasticity and resilience. The first major step in assembly is a liquid–liquid phase separation known as coacervation. This process represents a continuum of stages from initial phase separation to early growth of droplets by coalescence and later “maturation” leading to fiber formation. Assembly of tropoelastin-rich globules is on pathway for fiber formation in vivo. However, little is known about these intermediates beyond their size distribution. Here we investigate the contribution of sequence and structural motifs from full-length tropoelastin and a set of elastin-like polypeptides to the maturation of coacervate assemblies, observing their growth, stability and interaction behavior, and polypeptide alignment within matured globules. We conclude that maturation is driven by surface properties, leading to stabilization of the interface between the hydrophobic interior and aqueous solvent, potentially through structural motifs, and discuss implications for droplet interactions in fiber formation.  相似文献   

13.
The “fuzzy oil drop” model assumes that the idealized hydrophobic core in a protein body can be described by a 3D Gauss function. The structure of the 1ICF protein (cathepsin), which participates in the proteolysis process and has cysteine-type peptidase activity, has been analyzed on the basis of the “fuzzy oil drop” model. The authors have determined the contribution of individual exon fragments to the creation of a common hydrophobic core and assessed the involvement of each chain in this process, depending on the number of complexed chains. Quantitative assessment of exons, chains, dimers, and the whole complex suggest that each of these units plays a different role in shaping the protein’s hydrophobic core.  相似文献   

14.
In this study, 1064 nonhomologous “unsplit”, “one-strand split” and “two-strand split” right-handed βαβ-units having standard α-helices and loops up to seven residues in length have been analyzed. It was found that the α-helices in these kinds of βαβ-units have different distributions of the hydrophobic and hydrophilic amino acid residues along the chain. In the unsplit βαβ-units, most α-helices have hydrophobic residues in positions N4-N7-N8-N11 or N6-N7-N10, where N1 is the first N-terminal residue. In the one-strand split βαβ-units, most α-helices have hydrophobic residues in positions N4-N7-N8-N11 and those in two-strand split βαβ-units in positions N4-N5-N8-N12. On the other hand, in all kinds of βαβ-units, there are commonly occurring hydrophobic stripes of type C4-C7-C8 at the C-terminal parts of the α-helices. As a rule, the C- and N-terminal hydrophobic stripes overlap and the extent of their overlapping determine the length of α-helices.  相似文献   

15.
Hue Sun Chan  Ken A. Dill 《Proteins》1996,24(3):335-344
Proteins fold to unique compact native structures. Perhaps other polymers could be designed to fold in similar ways. The chemical nature of the monomer “alphabet” determines the “energy matrix” of monomer interactions—which defines the folding code, the relationship between sequence and structure. We study two properties of energy matrices using two-dimensional lattice models: uniqueness, the number of sequences that fold to only one structure, and encodability, the number of folds that are unique lowest-energy structures of certain monomer sequences. For the simplest model folding code, involving binary sequences of H (hydrophobic) and P (polar) monomers, only a small fraction of sequences fold uniquely, and not all structures can be encoded. Adding strong repulsive interactions results in a folding code with more sequences folding uniquely and more designable folds. Some theories suggest that the quality of a folding code depends only on the number of letters in the monomer alphabet, but we find that the energy matrix itself can be at least as important as the size of the alphabet. Certain multi-letter codes, including some with 20 letters, may be less physical or protein-like than codes with smaller numbers of letters because they neglect correlations among inter-residue interactions, treat only maximally compact conformations, or add arbitrary energies to the energy matrix.  相似文献   

16.
Bothropstoxin I (BthTX-I) from the venom of Bothrops jararacussuis a myotoxic phospholipase A2 (PLA2) homologue which, although catalytically inactive due to an Asp49→Lys substitution, disrupts the integrity of lipid membranes by a Ca2+-independent mechanism. The crystal structures of two dimeric forms of BthTX-I which diffract X-rays to resolutions of 3.1 and 2.1 Å have been determined. The monomers in both structures are related by an almost perfect twofold axis of rotation and the dimer interfaces are defined by contacts between the N-terminal α-helical regions and the tips of the β-wings of partner monomers. Significant differences in the relative orientation of the monomers in the two crystal forms results in “open” and “closed” dimer conformations. Spectroscopic investigations of BthTX-I in solution have correlated these conformational differences with changes in the intrinsic fluorescence emission of the single tryptophan residues located at the dimer interface. The possible relevance of this structural transition in the Ca2+-independent membrane damaging activity is discussed. Proteins 30:442–454, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

17.
Homo-Thy hetero-oligomer probes composed of trans-4-hydroxy-L-proline based PNA-like (HypNA) monomers and phosphono PNA (pPNA) monom-ers demonstrated strong binding to complementary poly A+ RNA strands. We used a mixture of chimeric oligomers containing both “linear” and “clamping” PNA-analogues to develop an mRNA isolation procedure and demonstrate the improved recovery of RNA molecules with secondary structure at the 3′end as well as RNAs with short poly A tails.  相似文献   

18.
Abstract

Poly (Val-Gly-Gly-Leu-Gly), a polypeptide mimicking the physico-chemical properties of the glycine-rich regions of elastin, has been synthesized and studied both in solution and in the aggregated state. By comparison, also the conformation of different “monomeric” units has been investigated. The polymer showed increased disorder with respect to the “monomers”, the molecular conformation being accounted for by a more or less random collection of isolated β-turns. Nevertheless, in the solid state the polymer is able to adopt supramolecular structures reminiscent of those found for elastin.  相似文献   

19.
A model for the interfacial orientation and the mode of action of lipase is proposed. Lipase is oriented so that its active site is near the oil-water boundary. This orientation is achieved by oil-enzyme bonding at the “hydrophobic head” of the enzyme, a region free of electric charges and relatively resistant to unfolding. The measured KM is a complex constant including the dissociation constant of this oil-enzyme “complex”. The interfacial orientation of lipase is further aided by hydrophilic negative charges on the “back” of the enzyme and by a hydrophilic carbohydrate “tail”.It is suggested that similar hydrophobic heads and hydrophilic tails and asymmetric charge distributions establish the orientation of many enzymes which act at interfaces. Many phospholipases, for instance, appear to be charge-oriented, and the carbohydrate residues of ribonucleases and many other glycoproteins may be hydrophilic tails.Lipase is probably a serine enzyme with a catalytic center similar to that of chymotrypsin, but more hindered, perhaps owing to the presence of a leucine residue, and there is no binding of substrate lipid chains in the “active complex”. The substrate molecule is fixated on the enzyme in a two-dimensional orientation, because its leaving alkoxy group must be received by the serine hydroxyl hydrogen which is directed towards the imidazol ring of the reactive histidine through a hydrogen bond. The high turnover rate of lipolysis, 5 × 105/min, exceptional even for an enzyme, results from the extremely high substrate concentration near the active site, and from an almost complete extrusion of water because of the hydrophobicity of both the active site and the substrate. In addition, both substrate and enzyme, because of their polarity, are already so favorably positioned at the interface that the formation of the “active complex” requires only a proper two-dimensional alignment, perhaps with partial extraction of the substrate molecule from the lipid phase.  相似文献   

20.
It is an important goal of computational biology to correctly predict the association state of a protein based on its amino acid sequence and the structures of known homologues. We have pursued this goal on the example of anthranilate phosphoribosyltransferase (AnPRT), an enzyme that is involved in the biosynthesis of the amino acid tryptophan. Firstly, known crystal structures of naturally occurring homodimeric AnPRTs were analyzed using the Protein Interfaces, Surfaces, and Assemblies (PISA) service of the European Bioinformatics Institute (EBI). This led to the identification of two hydrophobic “hot spot” amino acids in the protein-protein interface that were predicted to be essential for self-association. Next, in a comprehensive multiple sequence alignment (MSA), naturally occurring AnPRT variants with hydrophilic or charged amino acids in place of hydrophobic residues in the two hot spot positions were identified. Representative variants were characterized in terms of thermal stability, enzymatic activity, and quaternary structure. We found that AnPRT variants with charged residues in both hot spot positions exist exclusively as monomers in solution. Variants with hydrophilic amino acids in one hot spot position occur in both forms, monomer and dimer. The results of the present study provide a detailed characterization of the determinants of the AnPRT monomer-dimer equilibrium and show that analysis of hot spots in combination with MSAs can be a valuable tool in prediction of protein quaternary structures.  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