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.
A 34-amino acid portion of the third domain of alpha-fetoprotein possesses antigrowth and anticancer activities. Three analogs of this sequence were chemically synthesized, in which the two cysteines of the original sequence were replaced by alanines, glycines or serines. The original cysteine and alanine peptides formed trimers at 0.20 g/L in pH 7.4 phosphate buffer, and the glycine and serine peptides formed dimers. Trimer preparations were more potent in inhibiting estrogen-induced growth in the mouse uterine assays than the two dimeric oligomers. Of salient importance is that the alanine peptide retained its trimeric form in solution much longer than the cysteine peptide. Antigrowth assays were performed starting with stock solutions at a peptide concentration of 0.20 g/L, because at very high peptide concentration (8.0 g/L) the peptides aggregated extensively. All the peptides, although differing in biological activity, had almost identical secondary structures. Unlike alpha-fetoprotein, the three peptides have low amounts of alpha-helix. Trifluoroethanol has the ability to convert peptides into a helical conformation when they have a propensity for that structure. At trifluoroethanol concentrations of 20% and higher, the alanine and glycine peptides were changed into highly helical structures.  相似文献   

2.
A 34-amino acid synthetic peptide was derived from the third domain of human alpha-fetoprotein, and the peptide was shown to inhibit estrogen-stimulated growth. Under certain conditions, however, the peptide lost growth-inhibitory activity. A biophysical study of the peptide was undertaken with a goal of obtaining completely reliable preparations. The peptide was studied using gel-filtration column chromatography as a function of peptide concentration and age of solution, and was found to exhibit complex aggregation behaviors. During the early period (0-3 h) after dissolving lyophilized peptide into pH 7.4 buffer, solutions were composed mostly of trimers. At higher peptide concentrations (> or = 3.0 g/L), the trimers aggregated extensively to a large aggregate (minimum size approximately 102 peptides). At 5.0-8.0 g/L, these large aggregates increased in size (up to approximately 146 peptides) until trimers were largely exhausted from solution. During the later times (>3 h) after sample preparation, the trimeric oligomer of the peptide dissociated slowly to form dimers for samples at 0.10-3.0 g/L. After their build-up, a very small number of dimers associated to form hexamers. Disulfide bonds stabilized the dimers as indicated by the conversion of dimers to trimers upon the addition of a reducing agent, and the failure of dimers to form in the presence of reducing agent. Reducing agent did not affect trimer or large aggregate formation. Trimers were found to be active in an assay monitoring inhibition of estrogen-stimulated growth, whereas dimers and large aggregates were inactive. The two cysteines in the peptide were modified to either S-methylcysteine or S-(2-aminoethyl)cysteine, and both derivatives showed significant growth-inhibition activity. A serine analog in which both cysteines were replaced had very different aggregation behavior than the cysteine peptide and lacked its growth inhibitory ability. Peptide aggregation is critically important in establishing the ability of the peptide to inhibit growth and have anticancer activity, but the state of its two cysteines is of little influence.  相似文献   

3.
《Biophysical journal》2022,121(5):852-861
We previously speculated that the synergistically enhanced antimicrobial activity of Magainin 2 and PGLa is related to membrane adhesion, fusion, and further membrane remodeling. Here we combined computer simulations with time-resolved in vitro fluorescence microscopy, cryoelectron microscopy, and small-angle X-ray scattering to interrogate such morphological and topological changes of vesicles at nanoscopic and microscopic length scales in real time. Coarse-grained simulations revealed formation of an elongated and bent fusion zone between vesicles in the presence of equimolar peptide mixtures. Vesicle adhesion and fusion were observed to occur within a few seconds by cryoelectron microscopy and corroborated by small-angle X-ray scattering measurements. The latter experiments indicated continued and time-extended structural remodeling for individual peptides or chemically linked peptide heterodimers but with different kinetics. Fluorescence microscopy further captured peptide-dependent adhesion, fusion, and occasional bursting of giant unilamellar vesicles a few seconds after peptide addition. The synergistic interactions between the peptides shorten the time response of vesicles and enhance membrane fusogenic and disruption properties of the equimolar mixture compared with the individual peptides.  相似文献   

4.
Informatics for protein identification by mass spectrometry   总被引:3,自引:0,他引:3  
High throughput protein analysis (i.e., proteomics) first became possible when sensitive peptide mass mapping techniques were developed, thereby allowing for the possibility of identifying and cataloging most 2D gel electrophoresis spots. Shortly thereafter a few groups pioneered the idea of identifying proteins by using peptide tandem mass spectra to search protein sequence databases. Hence, it became possible to identify proteins from very complex mixtures. One drawback to these latter techniques is that it is not entirely straightforward to make matches using tandem mass spectra of peptides that are modified or have sequences that differ slightly from what is present in the sequence database that is being searched. This has been part of the motivation behind automated de novo sequencing programs that attempt to derive a peptide sequence regardless of its presence in a sequence database. The sequence candidates thus generated are then subjected to homology-based database search programs (e.g., BLAST or FASTA). These homology search programs, however, were not developed with mass spectrometry in mind, and it became necessary to make minor modifications such that mass spectrometric ambiguities can be taken into account when comparing query and database sequences. Finally, this review will discuss the important issue of validating protein identifications. All of the search programs will produce a top ranked answer; however, only the credulous are willing to accept them carte blanche.  相似文献   

5.
Small-angle X-ray scattering data have been measured for rabbit skeletal muscle troponin C and its complexes with the venom peptides melittin and mastoparan as well as synthetic peptides based on regions of the troponin I sequence implicated in troponin C binding. At the neutral pH used in this study (pH 6.8), troponin C shows a tendency to form dimers in the presence of 4 mol equiv of Ca2+, but is monomeric in solution when 2 or less mol equiv of Ca2+ is present. The 4Ca2+.troponin C dimers dissociate upon binding melittin, mastoparan, and peptides based on residues 96-115, 1-30, and 1-40 in the troponin I sequence. This result suggests that the peptide-binding sites overlap with the regions of contact between troponin C molecules forming a dimer. Like the structurally homologous calcium-binding protein calmodulin, troponin C shows conformational flexibility upon binding different peptides. Upon binding melittin, troponin C contracts in a similar manner to calmodulin when it binds peptides known to form amphiphilic helices (e.g., melittin, mastoparan, or MLCK-I). In contrast, mastoparan binding to troponin C does not result in a contracted structure. The scattering data indicate troponin C also remains in an extended structure upon binding the inhibitory peptides having the same sequence as residues 96-115 in troponin I.  相似文献   

6.
The process of protein misfolding and self-assembly into various, polymorphic aggregates is associated with a number of important neurodegenerative diseases. Only recently, crystal structures of several short peptides have provided detailed structural insights into -sheet rich aggregates, known as amyloid fibrils. Knowledge about early events of the formation and interconversion of small oligomeric states, an inevitable step in the cascade of peptide self-assembly, however, remains still limited. We employ molecular dynamics simulations in explicit solvent to study the spontaneous aggregation process of steric zipper peptide segments from the tau protein and insulin in atomistic detail. Starting from separated chains with random conformations, we find a rapid formation of structurally heterogeneous, -sheet rich oligomers, emerging from multiple bimolecular association steps and diverse assembly pathways. Furthermore, our study provides evidence that aggregate intermediates as small as dimers can be kinetically trapped and thus affect the structural evolution of larger oligomers. Alternative aggregate structures are found for both peptide sequences in the different independent simulations, some of which feature characteristics of the known steric zipper conformation (e.g., -sheet bilayers with a dry interface). The final aggregates interconvert with topologically distinct oligomeric states exclusively via internal rearrangements. The peptide oligomerization was analyzed through the perspective of a minimal oligomer, i.e., the dimer. Thereby all observed multimeric aggregates can be consistently mapped onto a space of reduced dimensionality. This novel method of conformational mapping reveals heterogeneous association and reorganization dynamics that are governed by the characteristics of peptide sequence and oligomer size.  相似文献   

7.
A 24-amino acid peptide, Humanin (HN), is a novel peptide that protects neuronal cells in vitro and in vivo from Alzheimer's disease-related toxicities. We have shown before that the structures of HN and a 1000-fold more active analog, HNG, with a Ser14Gly mutation are largely disordered. During additional mutational analysis, a shorter 17-amino acid form, AGA-(C8R)HNG17, was accidentally discovered to have a 100-fold higher activity than HNG. Here we have characterized the structural properties of the AGA-(C8R)HNG17 analog by circular dichroism (CD) and sedimentation equilibrium analysis. First, the structure in water was characterized, since these peptides have been dissolved in water prior to biological analysis. The AGA-(C8R)HNG17 peptide exhibited extensive beta-sheet structure in water, completely different from the aqueous HN and HNG structures. The beta-sheet structure was converted to a disordered structure upon dilution into phosphate-buffered saline (PBS) at low peptide concentration (e.g., below 0.2mg/ml), which was similar to the structure of HN and HNG, observed under similar conditions. Sedimentation equilibrium analysis showed that the AGA-(C8R)HNG17 analog was essentially monomeric in PBS, while HNG showed extensive aggregation. Such aggregation of HNG was observed when the peptide was added to the serum-containing cell culture media. Thus, the mutations introduced into the AGA-(C8R)HNG17 analog generated a peptide different from the parent HNG and HN peptides in the self-association properties and hence the solubility, which most likely contributed to the increased biological activity of the AGA-(C8R)HNG17 analog.  相似文献   

8.
9.
R E Jacobs  S H White 《Biochemistry》1986,25(9):2605-2612
The interactions of several members of a homologous series of peptides with the phospholipid bilayer have been examined by using fluorescence and deuterium NMR spectroscopy, differential scanning calorimetry, and measurements of water-to-bilayer partition coefficients. 1,2-Dimyristoyl-sn-glycero-3-phosphocholine (DMPC) bilayers and tripeptides of the form Ala-X-Ala-O-tert-butyl are used as a model system to probe the influence of amino acid side-chain substitution on the insertion of peptides into membranes and the behavior of peptide/bilayer mixtures. Tripeptides with X = Gly, Ala, Phe, and Trp have been examined. All of the tripeptides are water soluble, and all partition into DMPC bilayer vesicles to some extent. The Gly-containing peptide is the least soluble and the Trp-containing peptide the most soluble in the bilayer. The extent of perturbation of the bilayer structure induced by the peptides parallels their bilayer solubility: the Gly and Ala peptides act as simple impurities while peptides containing bulky aromatic rings cause a phase separation. Changes in the fluorescence properties of the Trp analogue upon incorporation into the bilayer indicate that the Trp side chain is probably immersed in the hydrocarbon region of the bilayer. Peptides of this form should serve as easily modifiable model systems with which to examine details of how the bilayer environment affects peptide conformation, as well as how hydrophobic peptides affect the bilayer structure.  相似文献   

10.
Peptide libraries displayed by filamentous bacteriophage have proven a powerful tool for the discovery of novel peptide agonists, antagonists and epitope mimics. Most phage-displayed peptides are fused to the N terminus of either the minor coat protein, pIII, or the major coat protein, pVIII. We report here that peptides containing cysteine residues, displayed as N-terminal fusions to pVIII, can form disulfide-bridged homodimers on the phage coat. Phage clones were randomly selected from libraries containing one or two fixed Cys residues, and surveyed for the presence of peptide-pVIII homodimers by SDS-PAGE analysis that involved pretreatment of the phage with reducing or thiol-modifying agents. For all phage whose recombinant peptide contained a single Cys residue, a significant fraction of the peptide-pVIII molecules were displayed as dimers on the phage coat. The dimeric form was in greater abundance than the monomer in almost all cases in which both forms could be reliably observed. Occasionally, peptides containing two Cys residues also formed dimers. These results indicate that, for a given pVIII-displayed peptide bearing a single Cys residue, a significant fraction of the peptide (>40 %) will dimerize regardless of its sequence; however, sequence constraints probably determine whether all of the peptide will dimerize. Similarly, only occasionally do peptides bearing two Cys residues form intermolecular disulfide bridges instead of intramolecular ones; this indicates that sequence constraints may also determine dimerization versus cyclization. Sucrose-gradient analysis of membranes from cells expressing pVIII fused to a peptide containing a single Cys residue showed that dimeric pVIII is present in the cell prior to its assembly onto phage. A model of the peptide-pVIII homodimer is discussed in light of existing models of the structure and assembly of the phage coat. The unique secondary structures created by the covalent association of peptides on the phage surface suggest a role for homo- and heterodimeric peptide libraries as novel sources of bioactive peptides.  相似文献   

11.
The conformational states of two peptide sequences that bind to staphylococcal enterotoxin B are sampled by replica exchange molecular dynamic (REMD) simulations in explicit water. REMD simulations were treated with 52 replicas in the range of 280–501 K for both peptides. The conformational ensembles of both peptides are dominated by random coil, bend and turn structures with a small amount of helical structures for each temperature. In addition, while an insignificant presence of β-bridge structures were observed for both peptides, the β-sheet structure was observed only for peptide 3. The results obtained from simulations at 300 K are consistent with the experimental results obtained from circular dichroism spectroscopy. From the analysis of REMD results, we also calculated hydrophobic and hydrophilic solvent accessible surface areas for both peptides, and it was observed that the hydrophobic segments of the peptides tend to form bend or turn structures. Moreover, the free-energy landscapes of both peptides were obtained by principal component analysis to understand how the secondary structural properties change according to their complex space. From the free-energy analysis, we have found several minima for both peptides at decreased temperature. For these obvious minima of both peptides, it was observed that the random coil, bend and turn structures are still dominant and the helix, β-bridge or β-sheet structures can appear or disappear with respect to minima. On the other hand, when we compare the results of REMD with conventional MD simulations for these peptides, the configurations of peptide 3 might be trapped in energy minima during the conventional MD simulations. Hence, it can be said that the REMD simulations have provided a sufficiently high sampling efficiency.  相似文献   

12.
Jensen KJ  Brask J 《Biopolymers》2005,80(6):747-761
Monosaccharides and amino acids are fundamental building blocks in the assembly of nature's polymers. They have different structural aspects and, to a significant extent, different functional groups. Oligomerization gives rise to oligosaccharides and peptides, respectively. While carbohydrates and peptides can be found conjoined in nature, e.g., in glycopeptides, the aim of this review is the radical redesign of peptide structures using carbohydrates, particularly monosaccharides and cyclic oligosaccharides, to produce novel peptides, peptidomimetics, and abiotic proteins. These hybrid molecules, chimeras, have properties arising largely from the combination of structural characteristics of carbohydrates with the functional group diversity of peptides. This field includes de novo designed synthetic glycopeptides, sugar (carbohydrate) amino acids, carbohydrate scaffolds for nonpeptidal peptidomimetics of cyclic peptides, cyclodextrin functionalized peptides, and carboproteins, i.e., carbohydrate-based proteinmimetics. These successful applications demonstrate the general utility of carbohydrates in peptide and protein architecture.  相似文献   

13.
Polyglutamine (polyQ, a peptide) with an abnormal repeat length is the causative agent of polyQ diseases, such as Huntington’s disease. Although glutamine is a polar residue, polyQ peptides form insoluble aggregates in water, and the mechanism for this aggregation is still unclear. To elucidate the detailed mechanism for the nucleation and aggregation of polyQ peptides, replica exchange molecular dynamics simulations were performed for monomers and dimers of polyQ peptides with several chain lengths. Furthermore, to determine how the aggregation mechanism of polyQ differs from those of other peptides, we compared the results for polyQ with those of polyasparagine and polyleucine. The energy barrier between the monomeric and dimeric states of polyQ was found to be relatively low, and it was observed that polyQ dimers strongly favor the formation of antiparallel β-sheet structures. We also found a characteristic behavior of the monomeric polyQ peptide: a turn at the eighth residue is always present, even when the chain length is varied. We previously showed that a structure including more than two sets of β-turns is stable, so a long monomeric polyQ chain can act as an aggregation nucleus by forming several pairs of antiparallel β-sheet structures within a single chain. Since the aggregation of polyQ peptides has some features in common with an amyloid fibril, our results shed light on the mechanism for the aggregation of polyQ peptides as well as the mechanism for the formation of general amyloid fibrils, which cause the onset of amyloid diseases.  相似文献   

14.
beta-Amyloid peptides are the main protein components of neuritic plaques and may be important in the pathogenesis of Alzheimer's Disease. The determination of the structure of beta-amyloid fibrils poses a challenge because of the limited solubility of beta-amyloid peptides and the noncrystalline nature of fibrils formed from these peptides. In this paper, we describe several physicochemical approaches which have been used to examine fibrils and the fibrillogenesis of peptide models of beta-amyloid. Recent advances in solid state NMR, such as the DRAWS pulse sequence, have made this approach a particularly attractive one for peptides such as beta-amyloid, which are not yet amenable to high-resolution solution phase NMR and crystallography. The application of solid state NMR techniques has yielded information on a model peptide comprising residues 10-35 of human beta-amyloid and indicates that in fibrils, this peptide assumes a parallel beta-strand conformation, with all residues in exact register. In addition, we discuss the use of block copolymers of Abeta peptides and polyethylene glycol as probes for the pathways of fibrillogenesis. These methods can be combined with other new methods, such as high-resolution synchrotron X-ray diffraction and small angle neutron and X-ray scattering, to yield structural data of relevance not only to disease, but to the broader question of protein folding and self-assembly.  相似文献   

15.
Laminin is present on the apical and basolateral sides of epithelial cells of very early sea urchin blastulae. We investigated whether small laminin-peptides, known to have cell binding activities, alter the development of sea urchin embryos. The peptide YIGSR-NH2 (850 μM) and the peptide PA22-2 (5 μM), which contains the peptide sequence IKVAV (Tashiro et al., J. Biol. Chem. 264, 16174, 1989), typically blocked archenteron formation when added to the sea water soon after fertilization. At lower doses, the YIGSR peptide allowed invagination of the archenteron but blocked archenteron extension and differentiation and evagination of the feeding arms. The effect of YIGSR and PA22-2 peptides declined when added to progressively older stages until no effect was seen when added at the mesenchyme blastula stage (24 hours after fertilization). Control peptides GRGDS, YIGSE, and SHA22, a dodeca-peptide with a scrambled IKVAV sequence, had no effect on development. The YIGSK peptide containing a conserved amino acid modification had only a small effect on gastrulation. The results suggest that YIGSR and IKVAV peptides specifically disrupt cell/extracellular matrix interactions required for normal development of the archenteron and feeding arms. Our recent finding that YTGIR is at the cell binding site of the B1 chain of S. purpuratus laminin supports this conclusion. Evidently, laminin or other laminin-like molecules are among the many extracellular matrix components needed for the invagination and extension of the archenteron during the gastrulation movements of these embryos.  相似文献   

16.
Many capsid proteins have peptides that influence their assembly. In hepatitis B virus capsid protein, the peptide STLPETTVV, linking the shell-forming 'core' domain and the nucleic acid-binding 'protamine' domain, has such a role. We have studied its morphogenic properties by permuting its sequence, substituting it with an extraneous peptide, deleting it to directly fuse the core and protamine domains and assembling core domain dimers with added linker peptides. The peptide was found to be necessary for the assembly of protamine domain-containing capsids, although its size-determining effect tolerates some modifications. Although largely invisible in a capsid crystal structure, we could visualize linker peptides by cryo-EM difference imaging: they emerge on the inner surface and extend from the capsid protein dimer interface towards the adjacent symmetry axis. A closely sequence-similar peptide in cellobiose dehydrogenase, which has an extended conformation, offers a plausible prototype. We propose that linker peptides are attached to the capsid inner surface as hinged struts, forming a mobile array, an arrangement with implications for morphogenesis and the management of encapsidated nucleic acid.  相似文献   

17.
Modifying a linear peptide near each terminus with a fluorescent dye can make it able to signal its own binding to a protein. As originally described, the dye pair is composed of fluorescein and tetramethylrhodamine [Wei, A.-P., Blumenthal, D. K., and Herron, J. N. (1994) Anal. Chem. 66, 1500-1506]. This paper shows that it may also be two molecules of tetramethylrhodamine. In aqueous solution, mutual affinity of the dyes causes fluorescence-quenching contact between them. When the peptide is bound by an antibody or cleaved by a proteinase, or when acetonitrile is added, dye-to-dye contact decreases and fluorescence increases 3-15-fold. When five peptides of 4-20 amino acid residues were doubly modified with tetramethylrhodamine, each product had the absorption spectrum of a tetramethylrhodamine dimer. As the peptides were not known to have special conformational features, self-affinity of the dye appeared to be the main cause of dimerization. Disruption of the dye dimers by acetonitrile suggested that dimerization of the dye(s) in aqueous solution was largely an effect of hydrophobicity. Dye-tagged peptides were used in fluorometric assays for two peptide-protein interactions. First, a peptide from type II collagen recognized by a monoclonal antibody was derivatized with two different dye pairs. The monoclonal bound each modified peptide, disrupting dye-to-dye contact and increasing fluorescence up to 4-fold. Second, a phosphopeptide recognized by an SH2 domain was tagged with fluorescein and tetramethylrhodamine, and its binding to the SH2 domain was detected through fluorescence. Doubly dye-tagged peptides offer a direct, solution-phase assay for protein-peptide binding.  相似文献   

18.
The high-resolution three-dimensional structure of an antimicrobial peptide has implications for the mechanism of its antimicrobial activity, as the conformation of the peptide provides insights into the intermolecular interactions that govern the binding to its biological target. For many cationic antimicrobial peptides the negatively charged membranes surrounding the bacterial cell appear to be a main target. In contrast to what has been found for other classes of antimicrobial peptides, solution NMR studies have revealed that in spite of the wide diversity in the amino acid sequences of amphibian antimicrobial peptides (AAMPs), they all adopt amphipathic α-helical structures in the presence of membrane-mimetic micelles, bicelles or organic solvent mixtures. In some cases the amphipathic AAMP structures are directly membrane-perturbing (e.g. magainin, aurein and the rana-box peptides), in other instances the peptide spontaneously passes through the membrane and acts on intracellular targets (e.g. buforin). Armed with a high-resolution structure, it is possible to relate the peptide structure to other relevant biophysical and biological data to elucidate a mechanism of action. While many linear AAMPs have significant antimicrobial activity of their own, mixtures of peptides sometimes have vastly improved antibiotic effects. Thus, synergy among antimicrobial peptides is an avenue of research that has recently attracted considerable attention. While synergistic relationships between AAMPs are well described, it is becoming increasingly evident that analyzing the intermolecular interactions between these peptides will be essential for understanding the increased antimicrobial effect. NMR structure determination of hybrid peptides composed of known antimicrobial peptides can shed light on these intricate synergistic relationships. In this work, we present the first NMR solution structure of a hybrid peptide composed of magainin 2 and PGLa bound to SDS and DPC micelles. The hybrid peptide adopts a largely helical conformation and some information regarding the inter-helix organization of this molecule is reported. The solution structure of the micelle associated MG2-PGLa hybrid peptide highlights the importance of examining structural contributions to the synergistic relationships but it also demonstrates the limitations in the resolution of the currently used solution NMR techniques for probing such interactions. Future studies of antimicrobial peptide synergy will likely require stable isotope-labeling strategies, similar to those used in NMR studies of proteins.  相似文献   

19.
We have studied the self-association reactions of purified GDP-liganded tubulin into double rings and taxoid-induced microtubules, employing synchrotron time-resolved x-ray solution scattering. The experimental scattering profiles have been interpreted by reference to the known scattering profiles to 3 nm resolution and to the low-resolution structures of the tubulin dimer, tubulin double rings, and microtubules, and by comparison with oligomer models and model mixtures. The time courses of the scattering bands corresponding to the different structural features were monitored during the assembly reactions under varying biochemical conditions. GDP-tubulin essentially stays as a dimer at low Mg(2+) ion activity, in either the absence or presence of taxoid. Upon addition of the divalent cations, it associates into either double-ring aggregates or taxoid-induced microtubules by different pathways. Both processes have the formation of small linear (short protofilament-like) tubulin oligomers in common. Tubulin double-ring aggregate formation, which is shown by x-ray scattering to be favored in the GDP- versus the GTP-liganded protein, can actually block microtubule assembly. The tubulin self-association leading to double rings, as determined by sedimentation velocity, is endothermic. The formation of the double-ring aggregates from oligomers, which involves additional intermolecular contacts, is exothermic, as shown by x-ray and light scattering. Microtubule assembly can be initiated from GDP-tubulin dimers or oligomers. Under fast polymerization conditions, after a short lag time, open taxoid-induced microtubular sheets have been clearly detected (monitored by the central scattering and the maximum corresponding to the J(n) Bessel function), which slowly close into microtubules (monitored by the appearance of their characteristic J(0), J(3), and J (n) - (3) Bessel function maxima). This provides direct evidence for the bidimensional assembly of taxoid-induced microtubule polymers in solution and argues against helical growth. The rate of microtubule formation was increased by the same factors known to enhance taxoid-induced microtubule stability. The results suggest that taxoids induce the accretion of the existing Mg(2+)-induced GDP-tubulin oligomers, thus forming small bidimensional polymers that are necessary to nucleate the microtubular sheets, possibly by binding to or modifying the lateral interaction sites between tubulin dimers.  相似文献   

20.
Animals vaccinated with heat shock protein (HSP)--peptide complexes develop specific protective immunity against cancers from which the HSPs were originally isolated. This autologous specific immunity has been demonstrated using a number of HSP--peptide antigen complexes. A prototypical HSP-based cancer vaccine is the gp96--peptide antigen complex, which is currently undergoing human clinical trials. Here, we analyzed the structure of a recombinant wild-type and a mutant gp96 protein and their peptide complexes using a number of biophysical techniques. Gel filtration chromatography, dynamic light scattering, and equilibrium analytical ultracentrifugation demonstrated that both a wild-type gp96 and a gp96 mutant lacking a dimerization domain formed higher order structures. More detailed analysis using scanning transmission electron microscopy indicated that both the wild-type and dimerization deletion mutant gp96 protein were organized, unexpectedly, into large aggregates. Size distributions ranged from dimers to octamers and higher. Circular dichroism and intrinsic Trp fluorescence suggested that the gp96 dimerization domain deletion mutant protein was more compact than the wild-type gp96. A fluorescent peptide antigen was synthesized, and the peptide-binding properties of wild-type and the dimerization domain deletion mutant gp96 were studied. Fluorescence lifetime and anisotropy decay showed that the bound antigenic peptide was located in a hydrophobic pocket, with considerable free space for the rotation of the probe. Deletion of the dimerization domain affected the peptide-binding microenvironment, although peptide-binding affinity was reduced by only a small extent. Peptide--gp96 complexes were extremely stable, persisting for many days in the cold. The extraordinary stability of peptide--gp96 complexes and the plasticity of the peptide-binding pocket support the proposed relay of diverse peptides to MHC and/or other molecules via molecular recognition.  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