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.
The aim of this study was to detect new polymorphisms in the bovine β‐casein (β‐CN) gene and to evaluate association of (new) β‐CN protein variants with milk production traits and milk protein composition. Screening of the β‐CN gene in genomic DNA from 72 Holstein Friesian (HF) bulls resulted in detection of 19 polymorphisms and revealed the presence of β‐CN protein variant I in the Dutch HF population. Studies of association of β‐CN protein variants with milk composition usually do not discriminate protein variant I from variant A2. Association of β‐CN protein variants with milk composition was studied in 1857 first‐lactation HF cows and showed that associations of protein variants A2 and I were quite different for several traits. β‐CN protein variant I was significantly associated with protein percentage and protein yield, and with αs1‐casein (αs1‐CN), αs2‐casein (αs2‐CN), κ‐casein (κ‐CN), α‐lactalbumin (α‐LA), β‐lactoglobulin (β‐LG), casein index and casein yield. Inferring β‐κ‐CN haplotypes showed that β‐CN protein variant I occurred only with κ‐CN variant B. Consequently, associations of β‐κ‐CN haplotype IB with protein percentage, κ‐CN, α‐LA, β‐LG and casein index are likely resulting from associations of κ‐CN protein variant B, while associations of β‐κ‐CN haplotype IB with αs1‐CN and αs2‐CN seem to be resulting from associations of β‐CN variant I.  相似文献   

2.
The multiphosphorylated tryptic peptide αs1‐casein(59–79) has been shown to be antigenic with anti‐casein antibodies. In an approach to determine the amino acyl residues critical for antibody binding we undertook an epitope analysis of the peptide using overlapping synthetic peptides. With αs1‐casein(59–79) as the adsorbed antigen in a competitive ELISA only two of five overlapping synthetic peptides at 1 mM significantly inhibited binding of the anti‐casein antibodies. Peptides Glu‐Ser(P)‐Ile‐Ser(P)‐Ser(P)‐Ser(P)‐Glu‐Glu and Ile‐Val‐Pro‐Asn‐Ser(P)‐Val‐Glu‐Glu inhibited antibody binding by 20.0±3.6% and 60.3±7.9%, respectively. The epitope of Glu63‐Ser(P)‐Ile‐Ser(P)‐Ser(P)‐Ser(P)‐Glu‐Glu70 was further localised to the phosphoseryl cluster as the peptide Ser(P)‐Ser(P)‐Ser(P) significantly inhibited binding of the anti‐casein antibodies to αs1‐casein(59–79) by 29.5±7.4%. Substitution of Ser(P)75 with Ser75 in the second inhibitory peptide Ile‐Val‐Pro‐Asn‐Ser(P)75‐Val‐Glu‐Glu also abolished inhibition of antibody binding to αs1‐casein (59–79) demonstrating that Ser(P)75 is also a critical residue for recognition by the antibodies. These data show that the phosphorylated residues in the cluster sequence ‐Ser(P)66‐Ser(P)‐Ser(P)68 and in the sequence ‐Pro73‐Asn‐Ser(P)‐Val‐Glu77‐ are critical for antibody binding to αs1‐casein(59–79) and further demonstrate that a highly phosphorylated segment of a protein can be antigenic. Copyright © 1999 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

3.
The β‐amyloid peptides (Aβ), Aβ1–40 and Aβ1–42, have been implicated in Alzheimer's disease (AD) pathology. Although Aβ1–42 is generally considered to be the pathological peptide in AD, both Aβ1–40 and Aβ1–42 have been used in a variety of experimental models without discrimination. Here we show that monomeric or oligomeric forms of the two Aβ peptides, when interact with the neuronal cation channel, α7 nicotinic acetylcholine receptors (α7nAChR), would result in distinct physiologic responses as measured by acetylcholine release and calcium influx experiments. While Aβ1–42 effectively attenuated these α7nAChR‐dependent physiology to an extent that was apparently irreversible, Aβ1–40 showed a lower inhibitory activity that could be restored upon washings with physiologic buffers or treatment with α7nAChR antagonists. Our data suggest a clear pharmacological distinction between Aβ1–40 and Aβ1–42. © 2003 Wiley Periodicals, Inc. J Neurobiol 55: 25–30, 2003  相似文献   

4.
J P Segrest 《FEBS letters》1976,69(1):111-115
The bovine αs2-, αs3-, αs4- and αs6-caseins [1] were isolated. The 4 proteins had the same amino-acid composition and C-terminal sequence, but different phosphorus contents. From a mixture of these proteins (designated as ‘αs2-complex’) and from αs3-casein a single and identical N-terminal sequence was obtained by Edman degradation. It seems therefore that the 4 proteins have the same peptide chain and only differ in their phosphorus content. For this reason we propose to modify the nomenclature of Annan and Manson [1] and to use in future the single term αs2 to designate the caseins which have been previously called αs2, αs3, αs4 and αs6 by these authors. The study of the primary structure of the peptide chain, which has confirmed these results, was undertaken on the S-carboxymethylated αs2-complex. From a cyanogen bromide digest and from a tryptic hydrolyzate of the αs2-complex, 5 and 25 peptides were obtained respectively, both sets of peptides accounting for the whole peptide chain. Examination of the tryptic peptides containing methionine combined with the N- and C-terminal sequences of the αs2-complex and some CNBr peptides, gave the order of the CNBr peptides, H.CN4CN2CN5CN1CN3.OH, which contain 4, 22, 115, 49 and 17 residues respectively. A partial sequence accounting for half of the peptide chain of the αs2-complex is given. This peptide chain is likely composed of 207 amino-acid residues  相似文献   

5.
This study designs a prediction model to differentiate pasteurized milk from heated extended shelf life (ESL) milk based on milk peptides. For this purpose, quantitative peptide profiles of a training set of commercial samples including pasteurized (n = 20), pasteurized‐ESL (n = 13), and heated‐ESL (n = 16) milk are recorded by matrix‐assisted laser desorption/ionization time‐of‐flight mass spectrometry (MALDI‐TOF‐MS). Seven peptides are selected as putative markers, and cutoff levels and performance measures of each marker are defined by receiver operating characteristic (ROC) analysis. The accuracy of these peptides in the training set range between 71% and 90%. A prediction model is established based on the combined cutoff levels and evaluated by an independent blind test set. The processing method of 19 out of 20 unknown milk samples is predicted correctly achieving 95% accuracy. Five peptides of the prediction model are identified as αS1‐casein182–199 (m/z 2014.0), αS1‐casein180–199 (m/z 2216.1), αS1‐casein1–24 (m/z 2910.6), β‐casein108–125 (m/z 2126.0), and β‐casein106–125 (m/z 2391.2) indicating thermal release and the action of plasmin and cathepsins. Thus, the present study demonstrates that the milk peptide profile reflects even minor differences in production parameters.  相似文献   

6.
There is a critical need for compounds that target cell surface integrin receptors for applications in cancer therapy and diagnosis. We used directed evolution to engineer the Ecballium elaterium trypsin inhibitor (EETI‐II), a knottin peptide from the squash family of protease inhibitors, as a new class of integrin‐binding agents. We generated yeast‐displayed libraries of EETI‐II by substituting its 6‐amino acid trypsin binding loop with 11‐amino acid loops containing the Arg‐Gly‐Asp integrin binding motif and randomized flanking residues. These libraries were screened in a high‐throughput manner by fluorescence‐activated cell sorting to identify mutants that bound to αvβ3 integrin. Select peptides were synthesized and were shown to compete for natural ligand binding to integrin receptors expressed on the surface of U87MG glioblastoma cells with half‐maximal inhibitory concentration values of 10–30 nM. Receptor specificity assays demonstrated that engineered knottin peptides bind to both αvβ3 and αvβ5 integrins with high affinity. Interestingly, we also discovered a peptide that binds with high affinity to αvβ3, αvβ5, and α5β1 integrins. This finding has important clinical implications because all three of these receptors can be coexpressed on tumors. In addition, we showed that engineered knottin peptides inhibit tumor cell adhesion to the extracellular matrix protein vitronectin, and in some cases fibronectin, depending on their integrin binding specificity. Collectively, these data validate EETI‐II as a scaffold for protein engineering, and highlight the development of unique integrin‐binding peptides with potential for translational applications in cancer. Proteins 2009. © 2009 Wiley‐Liss, Inc.  相似文献   

7.
The present work describes the synthesis and characterization of α/γ hybrid peptides, Boc‐Phe‐γ4‐Phe‐Val‐OMe, P1 ; Boc‐Ala‐γ4‐Phe‐Val‐OMe, P2 ; and Boc‐Leu‐γ4‐Phe‐Val‐OMe, P3 together with the formation of self‐assembled structures formed by these hybrid peptides in dimethyl sulfoxide (DMSO)/water (1:1). The self‐assembled structures were characterized by infrared (IR) spectroscopy, circular dichroism (CD), and scanning electron microscopy (SEM). Further, α/γ hybrid peptide self‐assembled structures were evaluated for antibacterial properties. Among all, the self‐assembled peptide P1 exhibited the antimicrobial activity against Escherichia coli and Klebsiella pneumoniae, while self‐assembled peptide P3 inhibited the biofilms of Salmonella typhimurium and Pseudomonas aeruginosa. In this study, we have shown the significance of self‐assembled structures formed from completely hydrophobic α/γ hybrid peptides in exploring the antibacterial properties together with biofilm inhibition.  相似文献   

8.
Three linear peptides incorporating d ‐Phe‐2‐Abz as the turn motif are reported. Peptide 1 , a hydrophobic β‐hairpin, served as a proof of principle for the design strategy with both NMR and CD spectra strongly suggesting a β‐hairpin conformation. Peptides 2 and 3, designed as amphipathic antimicrobials, exhibited broad spectrum antimicrobial activity, with potency in the nanomolar range against Staphylococcus aureus. Both compounds possess a high degree of selectivity, proving non‐haemolytic at concentrations 500 to 800 times higher than their respective minimal inhibitory concentrations (MICs) against S. aureus. Peptide 2 induced cell membrane and cell wall disintegration in both S. aureus and Pseudomonas aeruginosa as observed by transmission electron microscopy. Peptide 2 also demonstrated moderate antifungal activity against Candida albicans with an MIC of 50 μM. Synergism was observed with sub‐MIC levels of amphotericin B (AmB), leading to nanomolar MICs against C. albicans for peptide 2 . Based on circular dichroism spectra, both peptides 2 and 3 appear to exist as a mixture of conformers with the β‐hairpin as a minor conformer in aqueous solution, and a slight increase in hairpin population in 50% trifluoroethanol, which was more pronounced for peptide 3 . NMR spectra of peptide 2 in a 1:1 CD3CN/H2O mixture and 30 mM deuterated sodium dodecyl sulfate showed evidence of an extended backbone conformation of the β‐strand residues. However, inter‐strand rotating frame Overhauser effects (ROE) could not be detected and a loosely defined divergent hairpin structure resulted from ROE structure calculation in CD3CN/H2O. The loosely defined hairpin conformation is most likely a result of the electrostatic repulsions between cationic strand residues which also probably contribute towards maintaining low haemolytic activity.  相似文献   

9.
Self‐assembly of PAs composed of palmitic acid and several repeated heptad peptide sequences, C15H31CO‐(IEEYTKK)n‐NH2 (n = 1–4, represented by PA1–PA4), was investigated systematically. The secondary structures of the PAs were characterized by CD. PA3 and PA4 (n = 3 and 4, respectively) showed an α‐helical structure, whereas PA1 and PA2 (n = 1 and 2, respectively) did not display an α‐helical conformations under the tested conditions. The morphology of the self‐assembled peptides in aqueous medium was studied by transmission electron microscopy. As the number of heptad repeats in the PAs increased, the nanostructure of the self‐assembled peptides changed from nanofibers to nanovesicles. Changes of the secondary structures and the self‐assembly morphologies of PA3 and PA4 in aqueous medium with various cations were also studied. The critical micelle concentrations were determined using a pyrene fluorescence probe. In conclusion, this method may be used to design new peptide nanomaterials. Copyright © 2014 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

10.
Endo‐βN‐acetylglucosaminidase isolated from B. infantis ATCC 15697 (EndoBI‐1) is a novel enzyme that cleaves N‐N′‐diacetyl chitobiose moieties found in the N‐glycan core of high mannose, hybrid, and complex N‐glycans. These conjugated N‐glycans are recently shown as a new prebiotic source that stimulates the growth of a key infant gut microbe, Bifidobacterium longum subsp. Infantis. The effects of pH (4.45–8.45), temperature (27.5–77.5°C), reaction time (15–475 min), and enzyme/protein ratio (1:3,000–1:333) were evaluated on the release of N‐glycans from bovine colostrum whey by EndoBI‐1. A central composite design was used, including a two‐level factorial design (24) with four center points and eight axial points. In general, low pH values, longer reaction times, higher enzyme/protein ratio, and temperatures around 52°C resulted in the highest yield. The results demonstrated that bovine colostrum whey, considered to be a by/waste product, can be used as a glycan source with a yield of 20 mg N‐glycan/g total protein under optimal conditions for the ranges investigated. Importantly, these processing conditions are suitable to be incorporated into routine dairy processing activities, opening the door for an entirely new class of products (released bioactive glycans and glycan‐free milk). The new enzyme's activity was also compared with a commercially available enzyme, showing that EndoBI‐1 is more active on native proteins than PNGase F and can be efficiently used during pasteurization, streamlining its integration into existing processing strategies. © 2015 American Institute of Chemical Engineers Biotechnol. Prog., 31:1331–1339, 2015  相似文献   

11.
A single chiral cyclic α,α‐disubstituted amino acid, (3S,4S)‐1‐amino‐(3,4‐dimethoxy)cyclopentanecarboxylic acid [(S,S)‐Ac5cdOM], was placed at the N‐terminal or C‐terminal positions of achiral α‐aminoisobutyric acid (Aib) peptide segments. The IR and 1H NMR spectra indicated that the dominant conformations of two peptides Cbz‐[(S,S)‐Ac5cdOM]‐(Aib)4‐OEt ( 1) and Cbz‐(Aib)4‐[(S,S)‐Ac5cdOM]‐OMe (2) in solution were helical structures. X‐ray crystallographic analysis of 1 and 2 revealed that a left‐handed (M) 310‐helical structure was present in 1 and that a right‐handed (P) 310‐helical structure was present in 2 in their crystalline states. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

12.
The Wilms' tumor gene WT1 is overexpressed in various tumors, and the WT1 protein has been demonstrated to be an attractive target antigen for cancer immunotherapy. A WT1 protein‐derived 16‐mer peptide, WT1332 (KRYFKLSHLQMHSRKH), which was naturally generated through processing in cells and could elicit Th1‐type CD4+ helper T cell responses with an HLA‐DRB1*0405‐restriction has previously been identified by us. In the present study, it has been demonstrated that WT1332 can induce WT1332‐specific CD4+ T cell responses with the restriction of not only HLA‐DRB1*0405 but also HLA‐DRB1*1501, ‐DRB1*1502, or ‐DPB1*0901. These HLA class II‐restricted WT1332‐specific CD4+ T cell lines produced IFN‐γ but neither IL‐4 nor IL‐10 with WT1332 stimulation, thus showing a Th1‐type cytokine profile. Furthermore, HLA‐DRB1*1501 or ‐DRB1*1502‐restricted WT1332‐specific CD4+ T cell lines responded to WT1‐expressing transformed cells in an HLA‐DRB1‐restricted manner, which is consistent with our previous finding that WT1332 is a naturally processed peptide. These results indicate that the natural peptide, WT1332, is a promiscuous WT1‐specific helper epitope. WT1332 is expected to apply to cancer patients with various types of HLA class II as a WT1‐specific helper peptide in combination with HLA class I‐restricted WT1 peptides.  相似文献   

13.
We have recently reported that a ~19‐kDa polypeptide, rPK‐4, is a protein kinase Cs inhibitor that is 89% homologous to the 1171–1323 amino acid region of the 228‐kDa human pericentriolar material‐1 (PCM‐1) protein (Chakravarthy et al. 2012). We have now discovered that rPK‐4 binds oligomeric amyloid‐β peptide (Aβ)1‐42 with high affinity. Most importantly, a PCM‐1‐selective antibody co‐precipitated Aβ and amyloid β precursor protein (AβPP) from cerebral cortices and hippocampi from AD (Alzheimer's disease) transgenic mice that produce human AβPP and Aβ1‐42, suggesting that PCM‐1 may interact with amyloid precursor protein/Aβ in vivo. We have identified rPK‐4′s Aβ‐binding domain using a set of overlapping synthetic peptides. We have found with ELISA, dot‐blot, and polyacrylamide gel electrophoresis techniques that a ~ 5 kDa synthetic peptide, amyloid binding peptide (ABP)‐p4‐5 binds Aβ1‐42 at nM levels. Most importantly, ABP‐p4‐5, like rPK‐4, appears to preferentially bind Aβ1‐42 oligomers, believed to be the toxic AD‐drivers. As expected from these observations, ABP‐p4‐5 prevented Aβ1‐42 from killing human SH‐SY5Y neuroblastoma cells via apoptosis. These findings indicate that ABP‐p4‐5 is a possible candidate therapeutic for AD.  相似文献   

14.
We previously reported infiltration of immune-inflammatory cells in coronary arteries from cardiac allografts, associated with increased endothelial and smooth muscle cell fibronectin synthesis regulated by interleukin (IL)-1b?. We now investigate, using a porcine endothelial-smooth muscle cell co-culture system, whether IL-1b?-stimulated fibronectin production is functionally important in lymphocyte transendothelial migration. Lymphocytes were harvested from porcine peripheral blood and, in the unactivated state or following activation with phorbol myristic acetate (PMA) and IL-2, were characterized by fluorescence-activated cell sorter (FACS) analysis and added to a confluent endothelial monolayer on the upper chamber of a transwell system. Endothelial cells, as well as smooth muscle cells (in the bottom of the chamber), were stimulated with IL-1b?. Then transendothelial lymphocyte migration was determined in the presence of CS1 and RGD (fibronectin) peptides, blocking α4b?1 and α5b?1 integrin receptors on lymphocyte surfaces, respectively. A 55-70% inhibition of lymphocyte migration was observed when compared to control peptides. The combination of CS1 and RGD peptides did not significantly enhance the inhibitory effect of either peptide alone. A similar decrease in lymphocyte transendothelial migration toward smooth muscle cells was documented using a monoclonal antibody to cellular fibronectin. Furthermore, using smooth muscle cell conditioned medium; we reproduced the enhanced transendothelial lymphocyte migration as well as the inhibition with blocking peptides or fibronectin antibodies. Our data suggest that cytokine-mediated fibronectin synthesis in vascular cells recruits inflammatory cells through interactions of specific peptides with cell surface α4b?1 α5b?1 integrins. © 1995 Wiley-Liss, Inc.  相似文献   

15.
Fragments containing 5′ flanking regions of four bovine milk protein genes—alpha lactalbumin (bαLA), alpha S1 casein (bαS1CN), beta casein (bβCN), kappa casein (bkCN)—and mouse whey acidic protein (mWAP) gene were prepared by PCR and ligated to human growth hormone (hGH) gene. These recombinant DNAs were microinjected into rat embryos to produce transgenic rats, and the functions of the 5′ regions to direct secretion of hGH in the milk were tested. Although milk was obtained only in 5 of 19 mWAP/hGH rat lines, more than two-thirds of the rats carrying the other four DNAs produced milk. More than 80% of the lactated rats carrying bαLA/, bβCN/, and mWAP/hGH, and 33% of the laclated bαS1CN/hGH rats secreted detectable amounts of hGH (> 0.05 μg/ml) in the milk. In some rats, the hGH concentrations in the milk were comparable to or more than that of the corresponding milk protein in bovine milk. The ranges of hGH concentrations in the milk of bαLA/, bβCN/, bαS1CN/, and mWAP/hGH rats were 1.13–4,360 μg/ml, 0.11–10,900 μg/ml, 86.8–6,480 μg/ml, and 6.87–151 μg/ml, respectively. HGH was also detected in the sera of these rats, and some abnormalities of growth and reproduction were observed. All but one virgin mWAP/hGH rat secreted up to 0.0722 μg/ml of hGH in the serum, and more than half of them showed abnormal fat accumulations at their abdomen. None of the bαCN/hGH rats secreted detectable amount of hGH into their milk, whereas 8 of the 11 lines secreted hGH into their sera. For the production of hGH in transgenic rat milk, the 5′ region of bαS1CN was shown most suitable, because the bαS1CN/hGH rat secreted > 6,000 μg/ml of hGH into the milk and could be reproduced. © 1994 Wiley-Liss, Inc.  相似文献   

16.
The present study was designed to investigate the role of β‐amyloid (Aβ1‐42) in inducing neuronal pyroptosis and its mechanism. Mice cortical neurons (MCNs) were used in this study, LPS + Nigericin was used to induce pyroptosis in MCNs (positive control group), and Aβ1‐42 was used to interfere with MCNs. In addition, propidium iodide (PI) staining was used to examine cell permeability, lactate dehydrogenase (LDH) release assay was employed to detect cytotoxicity, immunofluorescence (IF) staining was used to investigate the expression level of the key protein GSDMD, Western blot was performed to detect the expression levels of key proteins, and enzyme‐linked immunosorbent assay (ELISA) was utilized to determine the expression levels of inflammatory factors in culture medium, including IL‐1β, IL‐18 and TNF‐α. Small interfering RNA (siRNA) was used to silence the mRNA expression of caspase‐1 and GSDMD, and Aβ1‐42 was used to induce pyroptosis, followed by investigation of the role of caspase‐1‐mediated GSDMD cleavage in pyroptosis. In addition, necrosulfonamide (NSA), an inhibitor of GSDMD oligomerization, was used for pre‐treatment, and Aβ1‐42 was subsequently used to observe the pyroptosis in MCNs. Finally, AAV9‐siRNA‐caspase‐1 was injected into the tail vein of APP/PS1 double transgenic mice (Alzheimer's disease mice) for caspase‐1 mRNA inhibition, followed by observation of behavioural changes in mice and measurement of the expression of inflammatory factors and pyroptosis‐related protein. As results, Aβ1‐42 could induce pyroptosis in MCNs, increase cell permeability and enhance LDH release, which were similar to the LPS + Nigericin‐induced pyroptosis. Meanwhile, the expression levels of cellular GSDMD and p30‐GSDMD were up‐regulated, the levels of NLRP3 inflammasome and GSDMD‐cleaved protein caspase‐1 were up‐regulated, and the levels of inflammatory factors in the medium were also up‐regulated. siRNA intervention in caspase‐1 or GSDMD inhibited Aβ1‐42‐induced pyroptosis, and NSA pre‐treatment also caused the similar inhibitory effects. The behavioural ability of Alzheimer's disease (AD) mice was relieved after the injection of AAV9‐siRNA‐caspase‐1, and the expression of pyroptosis‐related protein in the cortex and hippocampus was down‐regulated. In conclusion, Aβ1‐42 could induce pyroptosis by GSDMD protein, and NLRP3‐caspase‐1 signalling was an important signal to mediate GSDMD cleavage, which plays an important role in Aβ1‐42‐induced pyroptosis in neurons. Therefore, GSDMD is expected to be a novel therapeutic target for AD.  相似文献   

17.
The present work describes three novel nonpolar host peptide sequences that provide a ready assessment of the 310- and α-helix compatibilities of natural and unnatural amino acids at different positions of small- to medium-size peptides. The unpolar peptides containing Ala, Aib, and a C-terminal p-iodoanilide group were designed in such a way that the peptides could be rapidly assembled in a modular fashion, were highly soluble in solvent mixtures of triflouroethanol and H2O for CD- and two-dimensional (2D) nmr spectroscopic analyses, and showed excellent crystallinity suited for x-ray structure analysis. To validate our approach we synthesized 9-mer peptides 79a–96 (Table IV), 12-mer peptides 99–110c (Table V), and 10-mer peptides 120a–125d and 129–133 (Table VI and Scheme 8) incorporating a series of optically pure cyclic and open-chain (R)- and (S)-α,α-disubstituted glycines 1–10 (Figure 2). These amino acids are known to significantly modulate the conformations of small peptides. Based on x-ray structures of 9-mers 79a, 80, and 87 (Figures 4–7), 10-mers 124c, 131, and 132 (Figures 9–12), and 12-mer peptide 102b (Figure 13), CD spectra of all peptides recorded in acidic, neutral, and basic media and detailed 2D-nmr analyses of 9-mer peptide 86 and 12-mer 102b, several interesting conformational observations were made. Especially interesting results were obtained using the convex constraint CD analysis proposed by Fasman on 9-mer peptides 79a–d, 80, 81, 86, and 87, which allowed us to determine the relative content of 310- and α-helical conformations. These results were fully supported by the corresponding x-ray and 2D-nmr analyses. As a striking example we found that the (S)- and (R)-β-tetralin derived amino acids (R)- and (S)-1 show excellent α-helix stabilisation, more pronounced than Aib and Ala. These novel reference peptide sequences should help establish a scale for natural and unnatural amino acids concerning their intrinsic 310- and α-helix compatibilities at different positions of medium-sized peptides and thus improve our understanding in the folding processes of peptides. © 1997 John Wiley & Sons, Inc. Biopoly 42: 575–626, 1997  相似文献   

18.
The role aromatic amino acids play in the formation of amyloid is a subject of controversy. In an effort to clarify the contribution of aromaticity to the self‐assembly of human islet amyloid polypeptide (hIAPP)22‐29, peptide analogs containing electron donating groups (EDGs) or electron withdrawing groups (EWGs) as substituents on the aromatic ring of Phe‐23 at the para position have been synthesized and characterized using turbidity measurements in conjunction with Raman and fluorescence spectroscopy. Results indicate the incorporation of EDGs on the aromatic ring of Phe‐23 virtually abolish the ability of hIAPP22‐29 to form amyloid. Peptides containing EWGs were still capable of forming aggregates. These aggregates were found to be rich in β‐sheet secondary structure. Transmission electron microscopy images of the aggregates confirm the presence of amyloid fibrils. The observed difference in amyloidogenic propensity between peptides containing EDGs and those with EWGs appears not to be based on differences in peptide hydrophobicity. Fluorescence and Raman spectroscopic investigations reveal that the environment surrounding the aromatic ring becomes more hydrophobic and ordered upon aggregation. Furthermore, Raman measurements of peptide analogs containing EWGs, conclusively demonstrate a distinct downshift in the ? C?C? ring mode (ca. 1600 cm?1) upon aggregation that has previously been shown to be indicative of π‐stacking. While previous work has demonstrated that π‐stacking is not an absolute requirement for fibrillization, our findings indicate that Phe‐23 also contributes to fibril formation through π‐stacking interactions and that it is not only the hydrophobic nature of this residue that is relevant in the self‐assembly of hIAPP22‐29. © Proteins 2013. © 2012 Wiley Periodicals, Inc.  相似文献   

19.
Aggregation of Aβ peptides is a seminal event in Alzheimer's disease. Detailed understanding of the Aβ assembly process would facilitate the targeting and design of fibrillogenesis inhibitors. Here, conformational studies using FTIR spectroscopy are presented. As a model peptide, the 11–28 fragment of Aβ was used. This model peptide is known to contain the core region responsible for Aβ aggregation. The structural behavior of the peptide during aggregation provoked by the addition of water to Aβ(11–28) solution in hexafluoroisopropanol was compared with the properties of its variants corresponding to natural, clinically relevant mutants at positions 21–23 (A21G, E22K, E22G, E22Q and D23N). The results showed that the aggregation of the peptides proceeds via a helical intermediate, and it is possible that the formation of α‐helical structures is preceded by creation of 310‐helix/310‐turn structures. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

20.
Fibroblast growth factor‐2 (FGF2) plays a major role in angiogenesis. The pattern recognition receptor long‐pentraxin 3 (PTX3) inhibits the angiogenic activity of FGF2. To identify novel FGF2‐antagonistic peptide(s), four acetylated (Ac) synthetic peptides overlapping the FGF2‐binding region PTX3‐(97–110) were assessed for their FGF2‐binding capacity. Among them, the shortest pentapeptide Ac‐ARPCA‐NH2 (PTX3‐[100–104]) inhibits the interaction of FGF2 with PTX3 immobilized to a BIAcore sensorchip and suppresses FGF2‐dependent proliferation in endothelial cells, without affecting the activity of unrelated mitogens. Also, Ac‐ARPCA‐NH2 inhibits angiogenesis triggered by FGF2 or by tumorigenic FGF2‐overexpressing murine endothelial cells in chick and zebrafish embryos, respectively. Accordingly, the peptide hampers the binding of FGF2 to Chinese Hamster ovary cells overexpressing the tyrosine‐kinase FGF receptor‐1 (FGFR1) and to recombinant FGFR1 immobilized to a BIAcore sensorchip without affecting heparin interaction. In all the assays the mutated Ac‐ARPS A‐NH2 peptide was ineffective. In keeping with the observation that hydrophobic interactions dominate the interface between FGF2 and the FGF‐binding domain of the Ig‐like loop D2 of FGFR1, amino acid substitutions in Ac‐ARPCA‐NH2 and saturation transfer difference‐nuclear magnetic resonance analysis of its mode of interaction with FGF2 implicate the hydrophobic methyl groups of the pentapeptide in FGF2 binding. These results will provide the basis for the design of novel PTX3‐derived anti‐angiogenic FGF2 antagonists.  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