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.
Despite the importance of Callitrichid primates in both biomedical and conservation research, practical and reliable immunoassays for the measurement of follicle-stimulating hormone (FSH) have not yet been described. A panel of monoclonal antibodies against specific peptide fragments within either the alpha or beta subunit of human FSH was evaluated for their ability to recognize FSH from Callitrichid and other New World primates. One of these, monoclonal antibody 46.3h6.b7 raised against human FSH, was selected due to its ability to recognize marmoset monkey FSH and its low crossreactivity with other gonadotrophins. The antibody formed the basis of an enzymeimmunoassay using a highly purified human urinary FSH (Metrodin®, Serono) preparation coupled to biotin as label and unmodified as standard. After 24 h incubation, antibody bound label was visualized by addition of streptavidin-peroxidase followed by the appropriate substrate. Parallelism was obtained between the standard and dilutions of pituitary extracts, urine and plasma from the common marmoset (Callithrix jacchus) as well as from two tamarin species (Saguinus fuscicollis and S. oedipus) and one squirrel monkey (Saimiri sciureus). Profiles of plasma and urinary FSH during the follicular phase are shown for two individual marmosets. The ability to measure FSH in Callitrichidae provides new opportunities for studies of the reproductive biology of these New World primate species. Am. J. Primatol. 41:179–193, 1997. © 1997 Wiley-Liss, Inc.  相似文献   

2.
Background Common marmoset monkeys (Callithrix jacchus) are readily used in biomedical research. However, superovulation for embryonic stem cell production and developmental research still remain difficult. Inexplicably, follicle‐stimulating hormone (FSH) as key player in superovulation has to be administered in extremely high dosages in this non‐human primate compared to human. Methods To evaluate whether marmoset FSH (cjFSH) is functionally more competent than its human homologue on the marmoset FSH receptor (cjFSHR), we established in vitro a homologous system characterizing homologous and recombinantly produced cjFSH. Results Upon stimulation of two cell lines stably expressing either the marmoset or the human FSH receptor (cj/hFSHR), respectively, the second messenger signaling of the activated receptor displayed no significant difference in ED50 values. Thermostability of cjFSH was significantly prolonged by roughly 20% on both FSHRs. Conclusion High FSH dosage in marmoset superovulation cannot be explained by enhanced biopotency of the natural animal’s gonadotropin.  相似文献   

3.
We have recently demonstrated that synthetic peptides corresponding to the third cytoplasmic (3i) loop (residues 533 to 555) and a region in the carboxy-terminal cytoplasmic tail (residues 645 to 653) of the rat testicular follicle-stimulating hormone receptor (FSHR) affected signal transduction in rat testis membranes and cultured rat Sertoli cells. In order to define more precisely the peptide domains involved, we synthesized truncated peptide amides corresponding to FSHR residues 551–555 (KIAKR) and 650–653 (RKSH), respectively. These two regions were chosen since they contained a minimal structural motif present in G protein activator regions of several other G protein-coupled receptors (i.e., B-X-X-B-B or B-B-X-B, B representing a basic amino acid). Neither peptide inhibited binding of FSH to testis membrane receptors. Each peptide significantly reduced FSH-stimulated estradiol biosynthesis by intact cultured rat Sertoli cells. The same results were obtained with streptolysin O-permeabilized Sertoli cells. No effect was noted on forskolin-induced steroidogenesis, indicating that the peptide effects were not due to interaction with adenylyl cyclase. Each peptide amide, however, induced concentration-dependent increases in guanine nucleotide exchange in rat testis membranes. Our results indicate that interaction of FSH receptor with its associated G protein may involve relatively restricted peptide sequences, and include residues 551–555 (KIAKR) in the third cytoplasmic loop, and residues 650–653 (RKSH) in the carboxy-terminal cytoplasmic tail of the FSH receptor.  相似文献   

4.
We are developing a rapid, time‐resolved method using laser‐activated cross‐linking to capture protein‐peptide interactions as a means to interrogate the interaction of serum proteins as delivery systems for peptides and other molecules. A model system was established to investigate the interactions between bovine serum albumin (BSA) and 2 peptides, the tridecapeptide budding‐yeast mating pheromone (α‐factor) and the decapeptide human gonadotropin‐releasing hormone (GnRH). Cross‐linking of α‐factor, using a biotinylated, photoactivatable p‐benzoyl‐L‐phenylalanine (Bpa)–modified analog, was energy‐dependent and achieved within seconds of laser irradiation. Protein blotting with an avidin probe was used to detect biotinylated species in the BSA‐peptide complex. The cross‐linked complex was trypsinized and then interrogated with nano‐LC–MS/MS to identify the peptide cross‐links. Cross‐linking was greatly facilitated by Bpa in the peptide, but some cross‐linking occurred at higher laser powers and high concentrations of a non‐Bpa–modified α‐factor. This was supported by experiments using GnRH, a peptide with sequence homology to α‐factor, which was likewise found to be cross‐linked to BSA by laser irradiation. Analysis of peptides in the mass spectra showed that the binding site for both α‐factor and GnRH was in the BSA pocket defined previously as the site for fatty acid binding. This model system validates the use of laser‐activation to facilitate cross‐linking of Bpa‐containing molecules to proteins. The rapid cross‐linking procedure and high performance of MS/MS to identify cross‐links provides a method to interrogate protein‐peptide interactions in a living cell in a time‐resolved manner.  相似文献   

5.
Cysteine‐rich secretory proteins (CRISPs) are mainly found in the mammalian male reproductive tract and reported to be involved at different stages of fertilization. CRISPs have been shown to interact with prostate secretory protein of 94 amino acids (PSP94) from diverse sources, and the binding of these evolutionarily conserved proteins across species is proposed to be of functional significance. Of the three mammalian CRISPs, PSP94–CRISP3 interaction is well characterized, and specific binding sites have been identified; whereas, CRISP2 has been shown to interact with PSP94 in vitro. Interestingly, human CRISP3 and CRISP2 proteins are closely related showing 71.4% identity. In this study, we identified CRISP2 as a potential binding protein of PSP94 from human sperm. Further, we generated antisera capable of specifically detecting CRISP2 and not CRISP3. In this direction, specific peptides corresponding to the least conserved ion channel regulatory region were synthesized, and polyclonal antibodies were generated against the peptide in rabbits. The binding characteristics of the anti‐CRISP2 peptide antibody were evaluated using competitive ELISA. Immunoblotting experiments also confirmed that the peptide was able to generate antibodies capable of detecting the mature CRISP2 protein present in human sperm lysate. Furthermore, this anti‐CRISP2 peptide antibody also detected the presence of native CRISP2 on sperm.Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

6.
The follicle stimulating hormone (FSH) beta-subunit cDNAs were cloned and sequenced for an old world primate, the rhesus monkey (Macaca mulatta), and two New World primates, the common marmoset (Callithrix jacchus) and pygmy marmoset (Cebuella pygmaea). The cDNA and predicted amino acid sequences of the rhesus monkey FSH beta-subunit were related most closely to the human FSH beta-subunit (> 96% identity). The common and pygmy marmosets have identical FSH beta-subunit cDNAs, whereas the marmoset FSH beta-subunit diverges from the rhesus and human molecules with less than 93% identity. These results have significance for the implementation of assisted reproductive technologies in the nonhuman primate as well as the evolution of genes encoding reproductive hormones.  相似文献   

7.
The β‐subunit of the human chorionic gonadotropin (hCG) hormone, which is believed to be related to certain types of cancer, contains three hairpin‐like fragments. To investigate the role of β‐hairpin formation in the early stages of the hCGβ folding, a 28‐residue peptide with the sequence RDVRFESIRLPGSPRGVNPVVSYAVALS, corresponding to the H3‐β hairpin fragment (residues 60–87) of the hCGβ subunit, was studied under various conditions using three optical spectroscopic methods: Fourier transform ir spectroscopy, electronic CD, and vibrational CD. Environmental conditions are critical factors for formation of secondary structure in this peptide. TFE : H2O mixed solvents induced helical formation. Formation of β‐structure in this peptide, which may be related to the native β‐hairpin formation in the intact hormone, was found to be induced only under conditions such as high concentration, high temperature, and the presence of nonmicellar sodium dodecyl sulfate concentrations. These findings support a protein folding mechanism for the hCGβ subunit in which an initial hydrophobic collapse, which increases intermolecular interactions in hCGβ, is needed to induce the H3‐β hairpin formation. © 1999 John Wiley & Sons, Inc. Biopoly 50: 413–423, 1999  相似文献   

8.
Because of the long‐term co‐evolution of TCR and MHC molecules, numerous nucleotide substitutions have accumulated within the domains of TCRβ genes. We previously found that nonsynonymous nucleotide substitutions occurred more frequently in complementarity determining region (CDR)β than in CDRα, even though only a limited number of common marmoset (Callithrix jacchus) and human T‐cell receptor β variable (TRBV) sequences were compared. This interesting finding raised the question of whether the increased selective pressure within CDRβ was species‐specific. In this study, we identified 21 TRBV region sequences from the common marmoset and performed comparative sequence analyses of the T‐cell receptor α variable (TRAV) and TRBV regions from human, chimpanzee, rhesus monkey, cotton‐top tamarin, Ma's night monkey, and common marmoset. The ratios of the number of nonsynonymous nucleotide substitutions per site (dN) to the dS values (dN/dS) were less than 1 within the framework regions (FRs) of TRAV and TRBV region sequences, suggesting that purifying selection is largely dominant within the FRs. In contrast, the dN values were statistically significantly greater for CDRβ than for CDRα only in New World monkeys. Also, increased dN/dS ratios (dN/dS>1) were observed within CDRβ between humans and New World monkeys and, interestingly, between New World monkeys, which share a relatively recent common ancestor. Moreover, phylogenetic analysis by maximum likelihood analysis provided firm evidence to support that positive selection occurred within CDRβ along New World monkey lineages. These results suggest that increased positive selection pressure within CDRβ is common in New World monkeys rather than being species‐specific. This study provides an intriguing insight into the co‐evolution of TCR and MHC molecules within primates. Am. J. Primatol. 73:1082–1092, 2011. © 2011 Wiley‐Liss, Inc.  相似文献   

9.
10.
XPF‐St7 (GLLSNVAGLLKQFAKGGVNAVLNPK) is an antimicrobial peptide isolated from Silurana tropicalis. We developed an α‐helical segment of XPF‐St7 termed as XPF2. Using the XPF2 as a framework, we increased the positive net charge of XPF2 by amino acid substitutions, and thus obtained two novel antimicrobial peptides XPF4 and XPF6. These were each fused with an ubiquitin tag and successfully expressed in Escherichia coli. This ubiquitin fusion system may present a viable alternative for industrial production of antimicrobial peptides. XPF4 and XPF6 showed much better overall antimicrobial activity against both Gram‐negative and Gram‐positive bacteria than XPF2. The therapeutic index of XPF4 and XPF6 was 5.6‐fold and 6.7‐fold of XPF2, respectively. Bacterial cell membrane permeabilization and genomic DNA interaction assays were utilized to explore the mechanism of action of XPF serial peptides. The results revealed that the target of these antimicrobial peptides was the bacterial cytoplasmic membrane. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

11.
Modulation of protein–protein interactions involved in the immune system by using small molecular mimics of the contact interfaces may lead to the blockage of the autoimmune response and the development of drugs for immunotherapy. The nonpolymorphic β‐regions, exposed to the microenvironment, of the modeled HLA‐DQ7, which is genetically linked to autoimmune diseases, were determined. Peptides 132–141 and 58–67, located at the β1 and β2 domains of HLA‐DQ7, respectively, were tested for their involvement in the interactions with CD4+ T lymphocytes. Linear, cyclic, and dimeric analogs that mimic the exposed surfaces of HLA‐DQ7 were designed and synthesized. Their immunosuppressory activities, found in the secondary, humoral immune response to sheep erythrocytes (SRBC) in mice in vitro, ranged from 11% to 53%. The significance of the total charge of the peptides, the pattern of the hydrogen bonding, and the presence of secondary structure were investigated in relation to the immunomodulatory effect of the peptides. Two dimeric analogs of the HLA‐DQ7 58–67 fragment, consisting of the two monomers covalently linked by a polyethylene glycol (PEG) spacer, able to mimic the superdimers, were also synthesized and studied. As the 58–67 segment is located at the β1 region of HLA‐DQ7, close to the major histocompatibility complex (MHC) groove, one may assume that the 58–67 peptide could accommodate the association between T‐cell receptor (TCR) and human leukocyte antigen (HLA) by activating a co‐stimulatory molecule of the TCR/HLA interaction. This hypothesis is supported by the confocal laser image of the fluorescein‐labeled 58–67 peptide and by the fact that it is an immunostimulator at low concentration. Copyright © 2009 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
Ghosalkar JD  Mahale SD 《Peptides》2006,27(11):2894-2900
The extracellular domain (ECD) of the follicle stimulating hormone receptor (FSHR) has been shown to be a major determinant of hormone selectivity. The N-terminal 9-30 region, the sequence of which is unique to FSHR, has been extensively studied earlier and has been proposed to be an FSHR neutralizing epitope. In this study antipeptide antibodies specific to the peptide 9-30 were generated and used for identifying a specific immunodominant region within it. Overlapping peptides corresponding to the regions 9-19, 15-25 and 20-30 were synthesized. The ability of the antipeptide antibodies to 9-30 of FSHR to bind to different peptides was checked. The results indicated that the antibodies mainly recognized the peptide 20-30 and not the other two overlapping peptides. Further, the effect of the peptide 20-30 on the binding of radiolabeled FSH to its receptor was monitored. This peptide showed FSH-binding inhibitory activity with an IC(50) value of 0.598 x 10(-4)M and was more effective than the peptide 9-30 itself. Binding kinetics revealed that the observed effect of the peptide 20-30 is due to mixed type of inhibitory mechanism. This is the smallest peptide from the rat FSHR sequence having ability to inhibit FSH binding to its receptor by more than 90%.  相似文献   

13.
Deposition of insoluble fibrillar aggregates of β‐amyloid (Aβ) peptides in the brain is a hallmark of Alzheimer's disease. Apart from forming fibrils, these peptides also exist as soluble aggregates. Fibrillar and a variety of nonfibrillar aggregates of Aβ have also been obtained in vitro. Hexafluoroisopropanol (HFIP) has been widely used to dissolve Aβ and other amyloidogenic peptides. In this study, we show that the dissolution of Aβ40, 42, and 43 in HFIP followed by drying results in highly ordered aggregates. Although α‐helical conformation is observed, it is not stable for prolonged periods. Drying after prolonged incubation of Aβ40, 42, and 43 peptides in HFIP leads to structural transition from α‐helical to β‐conformation. The peptides form short fibrous aggregates that further assemble giving rise to highly ordered ring‐like structures. Aβ16–22, a highly amyloidogenic peptide stretch from Aβ, also formed very similar rings when dissolved in HFIP and dried. HFIP could not induce α‐helical conformation in Aβ16–22, and rings were obtained from freshly dissolved peptide. The rings formed by Aβ40, 42, 43, and Aβ16–22 are composed of the peptides in β‐conformation and cause enhancement in thioflavin T fluorescence, suggesting that the molecular architecture of these structures is amyloid‐like. Our results clearly indicate that dissolution of Aβ40, 42 and 43 and the amyloidogenic fragment Aβ16–22 in HFIP results in the formation of annular amyloid‐like structures. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
An efficient peptide purification strategy is established, comprising the selective reaction of an N‐terminal N‐(methoxy)glycine residue of the peptide and isothiocyanato‐functionalized resins, and subsequent Edman degradation. These reactions take place in acidic media; in particular, the Edman degradation proceeds smoothly in media containing more than 50% trifluoroacetic acid (v/v). These acidic conditions offer increased solubility, making them advantageous for the purification of hydrophobic and aggregation‐prone peptides. The effectiveness of this method, together with scope and limitations, is demonstrated using model peptides and the practical purification of the loop region of the human dopamine D2 receptor long isoform (residues 240–272). Copyright © 2016 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
We report a series of synthetic anticancer heptapeptides (H‐KKWβ2,2WKK‐NH2) containing eight different central lipophilic β2,2‐amino acid building blocks, which have demonstrated high efficiency when used as scaffolds in small cationic antimicrobial peptides and peptidomimetics. The most potent peptides in the present study had IC50 values of 9–23 µm against human Burkitt's lymphoma and murine B‐cell lymphoma and were all nonhaemolytic (EC50 > 200 µm ). The most promising peptide 10e also demonstrated low toxicity against human embryonic lung fibroblast cells and peripheral blood mononuclear cells and exceptional proteolytic stability. Copyright © 2012 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

16.
α‐melanocyte stimulating hormone (α‐MSH) is a tridecapeptide fragment of pro‐opiomelanocortin (POMC) with broad effects on appetite, skin pigmentation, hormonal regulation, and potential roles in both inflammation and autoimmunity. The use of this peptide as an anti‐inflammatory agent is limited by its low selectivity between the melanocortin receptors, susceptibility to proteolytic degradation, and rapid clearance from circulation. A retro‐inverso (RI) sequence of α‐MSH was characterized for receptor activity and resistance to protease. This peptide demonstrated surprisingly high selectivity for binding the melanocortin receptor 1 (MC1R). However, RI‐α‐MSH exhibited a diminished binding affinity for MC1R compared to α‐MSH. Mapping of the residues critical for agonist activity, receptor binding, and selectivity by alanine scanning, identified the same critical core tetrapeptide required for the native peptide. Modest improvements in affinity were obtained by conservative changes employing non‐natural amino acids and substitution of the C‐terminal sequence with a portion of a MC1R ligand peptide previously identified by phage display. Recombination of these elements yielded a peptide with an identical Ki as α‐MSH at MC1R and a lower EC50 in Mel‐624 melanoma cells. A number of other structural modifications of the RI peptide were found to differ in effect from those reported for the L ‐form α‐MSH, suggesting a significantly altered interaction with the MC1R. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

17.
18.
Alzheimer's disease is a progressive neurodegenerative disease characterized by extracellular deposits of β‐amyloid (Aβ) plaques. Aggregation of the Aβ42 peptide leading to plaque formation is believed to play a central role in Alzheimer's disease pathogenesis. Anti‐Aβ monoclonal antibodies can reduce amyloid plaques and could possibly be used for immunotherapy. We have developed a monoclonal antibody C706, which recognizes the human Aβ peptide. Here we report the crystal structure of the antibody Fab fragment at 1.7 Å resolution. The structure was determined in two crystal forms, P21 and C2. Although the Fab was crystallized in the presence of Aβ16, no peptide was observed in the crystals. The antigen‐binding site is blocked by the hexahistidine tag of another Fab molecule in both crystal forms. The poly‐His peptide in an extended conformation occupies a crevice between the light and heavy chains of the variable domain. Two consecutive histidines (His4–His5) stack against tryptophan residues in the central pocket of the antigen‐binding surface. In addition, they form hydrogen bonds to the acidic residues at the bottom of the pocket. The mode of his‐tag binding by C706 resembles the Aβ recognition by antibodies PFA1 and WO2. All three antibodies recognize the same immunodominant B‐cell epitope of Aβ. By similarity, residues Phe–Arg–His of Aβ would be a major portion of the C706 epitope. Copyright © 2010 John Wiley & Sons, Ltd.  相似文献   

19.
Among the pathological hallmarks of Alzheimer's disease (AD) is the deposition of amyloid‐β (Aβ) peptides, primarily Aβ (1–40) and Aβ (1–42), in the brain as senile plaques. A large body of evidence suggests that cognitive decline and dementia in AD patients arise from the formation of various aggregated forms of Aβ, including oligomers, protofibrils and fibrils. Hence, there is increasing interest in designing molecular agents that can impede the aggregation process and that can lead to the development of therapeutically viable compounds. Here, we demonstrate the ability of the specifically designed α,β‐dehydroalanine (ΔAla)‐containing peptides P1 (K‐L‐V‐F‐ΔA‐I‐ΔA) and P2 (K‐F‐ΔA‐ΔA‐ΔA‐F) to inhibit Aβ (1–42) aggregation. The mechanism of interaction of the two peptides with Aβ (1–42) seemed to be different and distinct. Overall, the data reveal a novel application of ΔAla‐containing peptides as tools to disrupt Aβ aggregation that may lead to the development of anti‐amyloid therapies not only for AD but also for many other protein misfolding diseases. © 2009 Wiley Periodicals, Inc. Biopolymers 91: 456–465, 2009. This article was originally published online as an accepted preprint. The “Published Online” date corresponds to the preprint version. You can request a copy of the preprint by emailing the Biopolymers editorial office at biopolymers@wiley.com  相似文献   

20.
The relationship between the conformation and biological activity of the peptide allosteric modulator of the interleukin‐1 receptor 101.10 (D ‐Arg‐D ‐Tyr‐D ‐Thr‐D ‐Val‐D ‐Glu‐D ‐Leu‐D ‐Ala‐NH2) has been studied using (R)‐ and (S)‐Bgl residues. Twelve Bgl peptides were synthesized using (R)‐ and (S)‐cyclic sulfamidate reagents derived from L ‐ and D ‐aspartic acid in an optimized Fmoc‐compatible protocol for efficient lactam installment onto the supported peptide resin. Examination of these (R)‐ and (S)‐Bgl 101.10 analogs for their potential to inhibit IL‐1β‐induced thymocyte cell proliferation using a novel fluorescence assay revealed that certain analogs exhibited retained and improved potency relative to the parent peptide 101.10. In light of previous reports that Bgl residues may stabilize type II′β‐turn‐like conformations in peptides, CD spectroscopy was performed on selected compounds to identify secondary structure necessary for peptide biological activity. Results indicate that the presence of a fold about the central residues of the parent peptide may be important for activity. Copyright © 2011 European Peptide Society and John Wiley & Sons, Ltd.  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