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.
Avbelj F  Baldwin RL 《Proteins》2006,63(2):283-289
The principle of group additivity is a standard feature of analyses of the energetics of protein folding, but it is known that it may not always be valid for the polar peptide group. The neighboring residue effect shows that group additivity is not strictly valid for a heteropeptide. We show here that group additivity fails seriously for peptide groups close to either peptide end, even for a homopeptide that has blocked end groups with no formal charges involved. The failure of group additivity is caused by the electrostatic character of the solvation of peptide polar groups and is illustrated with values of the electrostatic solvation free energy (ESF) calculated by DelPhi. Solvation free energies and enthalpies are known experimentally for monoamides and are often used to model the solvation of peptide groups, but ESF results show that monoamide values are very different from those of peptide groups. A main cause of the difference is that peptide solvation depends on the dipole-dipole interactions made between adjacent peptide groups, which vary with peptide conformation. Ligands that interact with the peptide backbone by an electrostatic mechanism could show a similar peptide end effect, and hydrogen exchange results from the literature confirm that exchange rates are position-dependent close to peptide ends.  相似文献   

2.
Bovine heart troponin T was hydrolyzed at the single cysteine residue. This procedure resulted in two peptides--a short N-terminal peptide (40-50 amino acid residues) and a long C-terminal peptide (240 amino acid residues). The C-terminal peptide was purified to homogeneity by ion-exchange chromatography; its properties were compared to those of intact troponin T. Data from circular dichroism spectroscopy suggest that the short N-terminal peptide cleavage was unaccompanied by any conspicuous changes in the secondary structure of the large C-terminal peptide of troponin T. Unlike intact troponin T, its C-terminal peptide can interact with troponin C in the presence of Ca2+. Data from affinity chromatography demonstrated that troponin I and tropomyosin more strongly interacted with native troponin T than with its C-terminal peptide. It is concluded that the short N-terminal peptide (40-50 residues) plays an essential role in cardiac troponin T interaction with troponin and tropomyosin components.  相似文献   

3.
The peptidyl transferase (PT) center of the ribosome catalyzes two nucleophilic reactions, peptide bond formation between aminoacylated tRNA substrates and, together with release factor, peptide release. Structure and function of the PT center are modulated by binding of aminoacyl-tRNA or release factor, thus providing the basis for the specificity of catalysis. Another way by which the function of the PT center is controlled is signaling from the peptide exit tunnel. The SecM nascent peptide induces ribosome stalling, presumably by inhibition of peptide bond formation. Similarly, the release factor-induced hydrolytic activity of the PT center can be suppressed by the TnaC nascent peptide contained in the exit tunnel. Thus, local and long-range conformational rearrangements can lead to changes in the reaction specificity and catalytic activity of the PT center.  相似文献   

4.
The hydrophobic region of the signal peptide of the OmpA protein of the Escherichia coli outer membrane was extensively altered in its hydrophobicity and predicted secondary structure by site-specific mutagenesis. The mutated signal peptides were fused to nuclease A from Staphylococcus aureus, and the function of the signal peptide was examined by measuring the rate of processing of the signal peptide. Six of the 12 mutated signal peptides in the nuclease hybrid were processed faster than the wild-type. In particular, the processing of the mutated signal peptide in which the alanine residue at position 9 was substituted with a valine residue was enhanced almost twofold over the processing of the wild-type signal peptide. In addition, the production of nuclease A fused with this mutated signal peptide also increased twofold. However, these effects were not observed when the mutated signal peptide was fused to TEM beta-lactamase. Analysis of the present mutations suggests that both overall hydrophobicity and distinct structural requirements in the hydrophobic region have important roles in signal peptide function.  相似文献   

5.
The simian immunodeficiency virus fusion peptide constitutes a 12-residue N-terminal segment of the gp32 protein that is involved in the fusion between the viral and cellular membranes, facilitating the penetration of the virus in the host cell. Simian immunodeficiency virus fusion peptide is a hydrophobic peptide that in Me(2)SO forms aggregates that contain beta-sheet pleated structures. When added to aqueous media the peptide forms large colloidal aggregates. In the presence of lipidic membranes, however, the peptide interacts with the membranes and causes small changes of the membrane electrostatic potential as shown by fluorescein phosphatidylethanolamine fluorescence. Thioflavin T fluorescence and Fourier transformed infrared spectroscopy measurements reveal that the interaction of the peptide with the membrane bilayer results in complete disassembly of the aggregates originating from an Me(2)SO stock solution. Above a lipid/peptide ratio of about 5, the membrane disaggregation and water precipitation processes become dependent on the absolute peptide concentration rather than on the lipid/peptide ratio. A schematic mechanism is proposed, which sheds light on how peptide-peptide interactions can be favored with respect to peptide-lipid interactions at various lipid/peptide ratios. These studies are augmented by the use of the fluorescent dye 1-(3-sulfonatopropyl)-4-[beta[2-(di-n-octylamino)-6-naphthyl]vinyl ] pyridinium betaine that shows the interaction of the peptide with the membranes has a clear effect on the magnitude of the so-called dipole potential that arises from dipolar groups located on the lipid molecules and oriented water molecules at the membrane-water interface. It is shown that the variation of the membrane dipole potential affects the extent of the membrane fusion caused by the peptide and implicates the dipolar properties of membranes in their fusion.  相似文献   

6.
High affinity binding sites for brain natriuretic peptide were characterized in the rat superior cervical ganglia by quantitative autoradiography. In addition, the peptide increased the formation of cyclic GMP in the ganglia in vitro. Brain natriuretic peptide displaced atrial natriuretic peptide from its binding sites. Our results suggest that brain natriuretic peptide and atrial natriuretic peptide may share physiologically active receptors in sympathetic ganglia. Brain natriuretic peptide may modulate the synaptic transmission in sympathetic ganglia, in addition or in conjunction with atrial natriuretic peptide.  相似文献   

7.
Previously it was reported that the 16-amino-acid (aa) C-terminal cytoplasmic tail of Moloney murine leukemia virus (MoMLV) transmembrane protein Pr15E is cleaved off during virus synthesis, yielding the mature, fusion active transmembrane protein p15E and the 16-aa peptide (R peptide or p2E). It remains to be elucidated how the R peptide impairs fusion activity of the uncleaved Pr15E. The R peptide from MoMLV was analyzed by Tricine-sodium dodecyl sulfate-polyacrylamide gel electrophoresis and immunostained with antiserum against the synthetic 16-aa R peptide. The R peptide resolved with an apparent molecular mass of 7 kDa and not the 4 kDa seen with the corresponding synthetic peptide. The 7-kDa R peptide was found to be membrane bound in MoMLV-infected NIH 3T3 cells, showing that cleavage of the 7-kDa R-peptide tail must occur before or during budding of progeny virions, in which only small amounts of the 7-kDa R peptide were found. The 7-kDa R peptide was palmitoylated since it could be labeled with [(3)H]palmitic acid, which explains its membrane association, slower migration on gels, and high sensitivity in immunoblotting. The present results are in contrast to previous findings showing equimolar amounts of R peptide and p15E in virions. The discrepancy, however, can be explained by the presence of nonpalmitoylated R peptide in virions, which were poorly detected by immunoblotting. A mechanistic model is proposed. The uncleaved R peptide can, due to its lipid modification, control the conformation of the ectodomain of the transmembrane protein and thereby govern membrane fusion.  相似文献   

8.
In order to investigate the role of residues inside and outside the peptide binding cleft of the L2 molecule in peptide presentation to cytotoxic T lymphocytes (CTL), we constructed a series of point mutations in the L d gene. We determined the effects of the mutations in the Ld molecule on the binding and recognition of an Ld-restricted CTL epitope derived from the nucleoprotein (NP) of the lymphocytic phoriomeningitis virus (LCMV). Each of the mutations within the Ld peptide binding cleft resulted in a complete loss of CTL recognition. Addition of the LCMV NP peptide to cells expressing these mutants did not increase surface Ld expression, suggesting that the mutations altered peptide binding. Mutations involving pockets D and E within the cleft affected LCMV peptide binding and recognition as drastically as those in pocket B, which was predicted to interact with a main anchor residue of the peptide. In striking contrast, the mutations located outside the cleft did not change either recognition or binding. These results demonstrate that the Ld residues in the peptide binding cleft are the main determinants dictating LCMV NP peptide binding, and that the residues in each of the pockets within the cleft play a role in this interaction. Surprisingly, one mutation outside the peptide binding cleft, T92S, abrogated CTL lysis of target cells treated with the LCMV NP peptide, but not virus-infected cells. These data show that this mutation selectively altered the presentation of the LCMV NP peptide introduced to the cell exogenously, but not endogenously. This implies that the pathway by which peptides associate with class I molecules within the cell differs from that of exogenous peptide binding.  相似文献   

9.
The conformation of the inactivating peptide of the Shaker B K+ channel (ShB peptide) and that of a noninactivating mutant (ShBL7E peptide) have been studied. Under all experimental conditions explored, the mutant peptide remains in a predominantly nonordered conformation. On the contrary, the inactivating ShB peptide has a great tendency to adopt a highly stable beta structure, particularly when challenged "in vitro" by anionic phospholipid vesicles. Because the putative peptide binding elements at the inner mouth of the channel comprise a ring of anionic residues and a hydrophobic pocket, we hypothesize that the conformational restrictions imposed on the ShB peptide by its interaction with the anionic lipid vesicles could partly imitate those imposed by the above ion channel elements. Thus, we propose that adoption of beta structure by the inactivating peptide may also occur during channel inactivation. Moreover, the difficulties encountered by the noninactivating ShBL7E peptide mutant to adopt beta structure and the observation that trypsin hydrolysis of the ShB peptide prevent both structure formation and channel inactivation lend further support to the hypothesis that adoption of beta structure by the inactivating peptide in a hydrophobic environment is important in determining channel blockade.  相似文献   

10.
Soybean 4-kDa peptide, a hormone-like peptide, is a ligand for the 43-kDa protein in legumes that functions as a protein kinase and controls cell proliferation and differentiation. As this peptide stimulates protein kinase activity, the interaction between the 4-kDa peptide (leginsulin) and the 43-kDa protein is considered important for signal transduction. However, the mechanism of interaction between the 4-kDa peptide and the 43-kDa protein is not clearly understood. We therefore investigated the binding mechanism between the 4-kDa peptide and the 43-kDa protein, by using gel-filtration chromatography and dot-blot immunoanalysis, and found that the 4-kDa peptide bound to the dimer form of the 43-kDa protein. Surface plasmon resonance analysis was then used to explore the interaction between the 4-kDa peptide and the 43-kDa protein. To identify the residues of the 4-kDa peptide involved in the interaction with the 43-kDa protein, alanine-scanning mutagenesis of the 4-kDa peptide was performed. The 4-kDa peptide-expression system in Escherichia coli, which has the ability to install disulfide bonds into the target protein in the cytoplasm, was employed to produce the 4-kDa peptide and its variants. Using mass spectrometry, the expressed peptides were confirmed as the oxidized forms of the native peptide. Surface plasmon resonance analysis showed that the C-terminal hydrophobic area of the 4-kDa peptide plays an important role in binding to the 43-kDa protein.  相似文献   

11.
The nonclassical major histocompatibility complex class I molecule HLA-E acts as a ligand for CD94/NKG2 receptors on the surface of natural killer cells and a subset of T cells. HLA-E presents closely related nonameric peptide epitopes derived from the highly conserved signal sequences of classical major histocompatibility complex class I molecules as well as HLA-G. Their generation requires cleavage of the signal sequence by signal peptidase followed by the intramembrane-cleaving aspartic protease, signal peptide peptidase. In this study, we have assessed the subsequent proteolytic requirements leading to generation of the nonameric HLA-E peptide epitopes. We show that proteasome activity is required for further processing of the peptide generated by signal peptide peptidase. This constitutes the first example of capture of a naturally derived short peptide by the proteasome, producing a class I peptide ligand.  相似文献   

12.
Parmar AS  Nunes AM  Baum J  Brodsky B 《Biopolymers》2012,97(10):795-806
Type XXV collagen, or collagen‐like amyloidogenic component, is a component of amyloid plaques, and recent studies suggest this collagen affects amyloid fibril elongation and has a genetic association with Alzheimer's disease. The relationship between the collagen triple helix and amyloid fibrils was investigated by studying peptide models, including a very stable triple helical peptide (Pro‐Hyp‐Gly)10, an amyloidogenic peptide GNNQQNY, and a hybrid peptide where the GNNQQNY sequence was incorporated between (GPO)n domains. Circular dichroism and nuclear magnetic resonance (NMR) spectroscopy showed the GNNQQNY peptide formed a random coil structure, whereas the hybrid peptide contained a central disordered GNNQQNY region transitioning to triple‐helical ends. Light scattering confirmed the GNNQQNY peptide had a high propensity to form amyloid fibrils, whereas amyloidogenesis was delayed in the hybrid peptide. NMR data suggested the triple‐helix constraints on the GNNQQNY sequence within the hybrid peptide may disfavor the conformational change necessary for aggregation. Independent addition of a triple‐helical peptide to the GNNQQNY peptide under aggregating conditions delayed nucleation and amyloid fibril growth. The inhibition of amyloid nucleation depended on the Gly‐Xaa‐Yaa sequence and required the triple‐helix conformation. The inhibitory effect of the collagen triple‐helix on an amyloidogenic sequence, when in the same molecule or when added separately, suggests Type XXV collagen, and possibly other collagens, may play a role in regulating amyloid fibril formation. © 2012 Wiley Periodicals, Inc. Biopolymers 97: 795–806, 2012.  相似文献   

13.
A R Merrill  F S Cohen  W A Cramer 《Biochemistry》1990,29(24):5829-5836
Acidic pH conditions required in vitro for membrane binding and activity of the channel-forming colicin E1 resulted in an increased susceptibility to proteases of the 178-residue thermolytic channel peptide, an increased accessibility to acrylamide of a fluorescence probe linked to cysteine-505 of the peptide, and an increased partition into nonionic detergent. The structural change in the peptide sensed by the fluorescence probe caused by a transition from pH 6.0 to 3.5 occurred in less than 1 s. The presence of low concentrations of detergents (0.001% SDS or 0.44% octyl beta-D-glucoside) or urea (0.2 M) at pH 6 or 4 also increased the susceptibility of the channel peptide to proteases. The increase in protease susceptibility and acrylamide accessibility at low pH, as well as partition of the peptide into nonionic detergent, suggested that acidic pH or the detergents might cause peptide unfolding. However, the hydrodynamic radius of the channel peptide at pH 6, 21-23 A, was not changed at pH 3.5 or by detergents or urea under conditions that increased the susceptibility of the peptide to protease. The activity of the channel peptide at pH 6 measured with liposomes and planar bilayers, which was a factor of 10(3)-10(4) smaller than that at pH 4, was increased by 2-4 orders of magnitude by 0.001% SDS or 0.44% octyl beta-D-glucoside, with an additional small increment of activity on planar bilayers caused by 0.01% SDS. A small increase in Stokes radius of the peptide in the presence of SDS could be detected that was approximately correlated with increased activity.  相似文献   

14.
Plant secretome comprises dozens of secreted proteins. However, little is known about the composition of the whole secreted peptide pools and the proteases responsible for the generation of the peptide pools. The majority of studies focus on target detection and characterization of specific plant peptide hormones. In this study, we performed a comprehensive analysis of the whole extracellular peptidome, using moss Physcomitrella patens as a model. Hundreds of modified and unmodified endogenous peptides that originated from functional and nonfunctional protein precursors were identified. The plant proteases responsible for shaping the pool of endogenous peptides were predicted. Salicylic acid (SA) influenced peptide production in the secretome. The proteasome activity was altered upon SA treatment, thereby influencing the composition of the peptide pools. These results shed more light on the role of proteases and posttranslational modification in the “active management” of the extracellular peptide pool in response to stress conditions. It also identifies a list of potential peptide hormones in the moss secretome for further analysis.  相似文献   

15.
The pheST operon codes for the two subunits of phenylalanyl-tRNA synthetase and it expression is controlled by attenuation in a way similar to many amino acid biosynthetic operons. The nucleotide sequence of the control regions of the operon indicates the presence of several open reading frames besides that of the leader peptide. One of these open reading frames, called the alternative leader peptide, starts at about the same place as the leader peptide and ends after the terminator of the attenuator. Another open reading frame, called the terminator peptide, starts after the terminator and covers about half the distance to pheS, the first structural gene of the operon. The present report shows that, in fact, the only open reading frame to be translated efficiently is the leader peptide itself. The alternative leader peptide and the terminator peptide are both translated at a negligible rate.  相似文献   

16.
In this study, a peptide–peptide co-administration therapy between hybrid peptide kla-TAT and cationic anticancer peptide HPRP-A1 was designed to increase the anticancer activity of the combination peptides through synergistic effect. kla is a pro-apoptotic peptide which could induce rapid cancer cell apoptosis by disruption the mitochondrial membrane when internalized the cells. To enhance more kla peptides pass through cell membrane, a double improvement strategy was designed by chemically conjugation with cell penetration peptide TAT as well as co-administration with cationic membrane active peptide HPRP-A1, and the double anticancer mechanism of the kla-TAT peptide and HPRP-A1 including membrane disruption and apoptosis induction was verified through in vitro experiments. The CompuSyn synergism/antagonism analysis showed that kla-TAT acted synergistically with HPRP-A1 against a non-small cell lung cancer (NSCLC) A549 cell line. The anticancer activities of the two peptides were dramatically increased by co-administration, under the mechanism of cell membrane disruption, caspase-dependent apoptosis induction, as well as cyclin-D1 down-regulation based G1 phase arrest. We believe that the synergic therapeutic strategy would be a meaningful method for the anticancer peptides used in cancer treatment.  相似文献   

17.
The MHC class Ib molecule Qa-1 binds specifically and predominantly to a single 9-aa peptide (AMAPRTLLL) derived from the leader sequence of many MHC class Ia proteins. This peptide is referred to as Qdm. In this study, we report the isolation and sequencing of a heat shock protein 60-derived peptide (GMKFDRGYI) from Qa-1. This peptide is the dominant peptide bound to Qa-1 in the absence of Qdm. A Qa-1-restricted CTL clone recognizes this heat shock protein 60 peptide, further verifying that it binds to Qa-1 and a peptide from the homologous Salmonella typhimurium protein GroEL (GMQFDRGYL). These observations have implications for how Qa-1 can influence NK cell and T cell effector function via the TCR and CD94/NKG2 family members, and how this effect can change under conditions that cause the peptides bound to Qa-1 to change.  相似文献   

18.
The ribosomal peptidyl transferase center is responsible for two fundamental reactions, peptide bond formation and nascent peptide release, during the elongation and termination phases of protein synthesis, respectively. We used in vitro genetics to investigate the functional importance of conserved 23S rRNA nucleotides located in the peptidyl transferase active site for transpeptidation and peptidyl-tRNA hydrolysis. While mutations at A2451, U2585, and C2063 (E. coli numbering) did not significantly affect either of the reactions, substitution of A2602 with C or its deletion abolished the ribosome ability to promote peptide release but had little effect on transpeptidation. This indicates that the mechanism of peptide release is distinct from that of peptide bond formation, with A2602 playing a critical role in peptide release during translation termination.  相似文献   

19.
The peptide NH(2)-DTEDQEDQVDPR-COOH is released during activation of protein C zymogen. We measured the effect of a synthetic peptide with an amino acid sequence similar to that of the natural peptide on platelets from healthy individuals using platelet aggregometry. We found that this synthetic peptide inhibits platelet aggregation induced by thrombin; furthermore, it diminishes mobilization of intraplatelet calcium. Molecular docking showed weak interaction between the synthetic peptide and thrombin. Our findings suggest that this synthetic peptide may interact with a receptor located on the platelet cell membrane.  相似文献   

20.
Both in vivo and in vitro studies suggest that the Glu(724)-Pro(760) (peptide C) region of the dihydropyridine receptor alpha1 II-III loop is important for excitation-contraction coupling, although its actual function has not yet been elucidated. According to our recent studies, peptide C inhibits Ca(2+) release induced by T-tubule depolarization or peptide A. Here we report that peptide C has Ca(2+)-dependent dual functions on the skeletal muscle ryanodine receptor. Thus, at above-threshold [Ca(2+)]s (> or =0.1 microm) peptide C blocked peptide A-induced activation of the ryanodine receptor (ryanodine binding and Ca(2+) release); peptide C also blocked T-tubule depolarization-induced Ca(2+) release. However, at sub-threshold [Ca(2+)]s (<0.1 microm), peptide C enhanced ryanodine binding and induced Ca(2+) release. If peptide A was present, together with peptide C, both peptides produced additive activation effects. Neither peptide A nor peptide C produced any appreciable effect on the cardiac muscle ryanodine receptor at both high (1.0 microm) and low (0.01 microm) Ca(2+) concentrations. These results suggest the possibility that the in vivo counterpart of peptide C retains both activating and blocking functions of the skeletal muscle-type excitation-contraction coupling.  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