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 15 毫秒
1.
α-Amylases (EC 3.2.1.1) hydrolyze internal α-1,4-glucosidic linkages of starch and related glucans. Bacillus sp. AAH-31 produces an alkalophilic thermophilic α-amylase (AmyL) of higher molecular mass, 91 kDa, than typical bacterial α-amylases. In this study, the AmyL gene was cloned to determine its primary structure, and the recombinant enzyme, produced in Escherichia coli, was characterized. AmyL shows no hydrolytic activity towards pullulan, but the central region of AmyL (Gly395-Asp684) was similar to neopullulanase-like α-amylases. In contrast to known neopullulanase-like α-amylases, the N-terminal region (Gln29-Phe102) of AmyL was similar to carbohydrate-binding module family 20 (CBM20), which is involved in the binding of enzymes to starch granules. Recombinant AmyL showed more than 95% of its maximum activity in a pH range of 8.2–10.5, and was stable below 65 °C and from pH 6.4 to 11.9. The k cat values for soluble starch, γ-cyclodextrin, and maltotriose were 103 s?1, 67.6 s?1, and 5.33 s?1, respectively, and the K m values were 0.100 mg/mL, 0.348 mM, and 2.06 mM, respectively. Recombinant AmyL did not bind to starch granules. But the substitution of Trp45 and Trp84, conserved in site 1 of CBM20, with Ala reduced affinity to soluble starch, while the mutations did not affect affinity for oligosaccharides. Substitution of Trp61, conserved in site 2 of CBM20, with Ala enhanced hydrolytic activity towards soluble starch, indicating that site 2 of AmyL does not contribute to binding to soluble long-chain substrates.  相似文献   

2.
Amyloid-β (Aβ) peptides and other amyloidogenic proteins can form a wide range of soluble oligomers of varied morphologies at the early aggregation stage, and some of these oligomers are biologically relevant to the pathogenesis of Alzheimer's disease. Spherical micelle-like oligomers have been often observed for many different types of amyloids. Here, we report a hybrid computational approach to systematically construct, search, optimize, and rank soluble micelle-like Aβ25-35 structures with different side-chain packings at the atomic level. Simulations reveal for the first time, to our knowledge, that two Aβ micelles with antiparallel peptide organization and distinct surface hydrophobicity display high structural stability. Stable micelles experience a slow secondary structural transition from turn to α-helix. Energetic analysis coupled with computational mutagenesis reveals that van der Waals and solvation energies play a more pronounced role in stabilizing the micelles, whereas the electrostatic energies present a stable but minor energetic contribution to peptide assemblies. Modeled Aβ micelles with shapes and dimensions similar to those of experimentally derived spherical structures also provide detailed information about the roles of structural dynamics and transition in the formation of amyloid fibrils. The strong binding affinity of our micelles to antibodies implies that micelles may be a biologically relevant species.  相似文献   

3.
The transglucosylation action of buckwheat α-glucosidase on soluble starch, maltose maltotriose and maltotetraose are described and discussed. The transglucosylation products of soluble starch were isolated by carbon-Celite column chromatography and by paper chromatography. Among the products were found follows: nigerose, maltose, kojibiose and isomaltose as disaccharides and 2-α-isomaltosylglucose, 2-α-nigerosylglucose, nigerotriose. 3-α-maltosylglucose and maltotriose as trisaccharides.

Furthermore, the existence of 6-α-nigerosylglucose, 4-α-kojibiosylglucose, panose, isopanose and 3-α-isomaItosylglucose was suspected. A new trisaccharide, 2-α-nigerosylglucose which was obtained in a crystalline form (monohydrate) melted at 186~188°C and gave [α]D+ 178.3 (c = 0.6, in water).

These experimental results on the reaction products seem to indicate that the activated glucosyl group from the substrate (starch, in this case) is transferred to any position of C–2,3,4 or 6 of glucose released from the substrate and the same type of transglucosylation occurrs upon the non-reducing terminal of disaccharides just produced, which leads to the formation of various kinds of trisaccharide, tetrasaccharide etc. The synthesis of α-oligosaccharides from free glucose could not be detected by paper chromatography.  相似文献   

4.
We have mapped the epitopes to which two monoclonal antibodies against acetylcholinesterase (AChE) from Torpedo californica are directed. One antibody, 2C9, has equivalent affinity for both the 5.6S (amphiphilic) and 11S (hydrophilic) enzyme forms; the other, 4E7, recognizes only the amphiphilic form and has been shown previously to require an N-linked oligosaccharide residue on the protein. Isolation of cyanogen bromide peptides from the amphiphilic form and assay by a competition ELISA for 2C9 and by a direct binding ELISA for 4E7 identified the same peptide, residues 44–82, as containing epitopes against both antibodies. The epitope for 4E7 includes the oligosaccharide conjugated to Asp59, an N-linked glycosylation site not present in mouse AChE. A 20-amino-acid synthetic peptide, RFRRPEPKKPWSQVWNASTY, representing residues 44–63, was synthesized and found to inhibit completely 2C9 binding to 5.6S enzyme at molar concentrations comparable to those of the cyanogen bromide peptide. It was unreactive with 4E7. Fractionation of the synthetic peptide further localized the 2C9 epitope. Peptides RFRRPEPKKPW and KPWSGVWNASTY both reacted but less so than the entire synthetic peptide at equivalent molar concentrations, whereas the peptide RPEPKKPWSGVWNASTY was as effective as the larger synthetic peptide. The crystal structure of AChE shows the peptide to be on the surface of the molecule as part of a convex hairpin loop starting before the first α-helix.  相似文献   

5.
1H and 13C high-resolution nmr spectra of cationic, zwitterionic, and anionic forms of the peptides: H-Trp-(Pro)n-Tyr-OH, n = 0-5, and H-Trp-Pro-OCH3 were obtained in D2O solution. Analysis of Hα(Pro1), Hα(Trp), Cγ(Pro), Hε(Tyr), and Hδ(Trp) resonances provided evidence for the presence of two predominant backbone isomers: the all-trans one and another with the Trp-Pro peptide bond in cis conformation; the latter constituted about 0.8 molar fraction of the total peptide (n > 1) concentration. Relative content of these isomers varied in a characteristic way with the number of Pro residues and the ionization state of the peptides. The highest content of the cis (Trp-Pro) isomer, 0.74, was found in the anionic form of H-Trp-Pro-Tyr-OH; it decreased in the order of: anion ? zwitterion ≈ cation, and with the number of Pro residues to reach the value of 0.42 in the cationic form of H-Trp- (Pro)5-Tyr-OH. Isomerization equilibria about Pro-Pro bond(s) were found to be shifted far (?0.9) in favor of the trans conformation. Interpretation of the measured vicinal coupling constants Jα?β′ and Jα?β″ for CαH-CβH2 proton systems of Trp and Tyr side chains in terms of relative populations of g+, g?, and t staggered rotamers around the χ1 dihedral angle indicated that in all the peptides studied (a) rotation of Trp indole ring in cis (Trp-Pro) isomers is strongly restricted, and (b) rotation of Tyr phenol ring is relatively free. The most preferred χ1 rotamer of Trp (0.8-0.9 molar fraction) was assigned as the t one on the basis of a large value of the vicinal coupling constant between the high-field Hβ and carbonyl carbon atoms of Trp, estimated for the cis (Pro1) form of H-Trp-Pro-Tyr-OH from a 1H, 13C correlated spectroscopy 1H detected multiple quantum experiment. This indicates that cis ? trans equilibrium in the Trp-Pro fragment is governed by nonbonding interactions between the pyrrolidine (Pro) and indole (Trp) rings. A molecular model of the terminal cis Trp-Pro dipeptide fragment is proposed, based on the presented nmr data and the results of our molecular mechanics modeling of low-energy conformers of the peptides, reported elsewhere. © 1993 John Wiley & Sons, Inc.  相似文献   

6.
Human α- and β-enolases are highly homologous enzymes, difficult to differentiate immunologically. In this work, we describe production, purification and properties of anti-α- and anti-β-enolase polyclonal antibodies. To raise antibodies, rabbits were injected with enolase isoenzymes that were purified from human kidney (α-enolase) and skeletal muscle (β-enolase). Selective anti-α- and anti-β-enolase antibodies were obtained by affinity chromatography on either α- or β-enolase-Sepharose columns. On Western blots, antibodies directed against human β-enolase, did not react with human α-isoenzyme, but recognized pig and rat β-enolase. To determine what makes these antibodies selective bioinformatic tools were used to predict conformational epitopes for both enolase isoenzymes. Three predicted epitopes were mapped to the same regions in both α- and β-enolase. Peptides corresponding to predicted epitopes were synthesized and tested against purified antibodies. One of the pin-attached peptides representing α-enolase epitope (the C-terminal portion of the epitope 3 - S262PDDPSRYISPDQ273) reacted with anti-α-enolase, while the other also derived from the α-enolase sequence (epitope 2 - N193VIKEKYGKDATN205) was recognized by anti-β-enolase antibodies. Interestingly, neither anti-α- nor anti-β-antibody reacted with a peptide corresponding to the epitope 2 in β-enolase (G194VIKAKYGKDATN206). Further analysis showed that substitution of E197 with A in α-enolase epitope 2 peptide lead to 70% loss of immunological activity, while replacement of A198 with E in peptide representing β-enolase epitope 2, caused 67% increase in immunological activity. Our results suggest that E197 is essential for preserving immunologically active conformation in epitope 2 peptidic homolog, while it is not crucial for this epitope's antigenic activity in native β-enolase.  相似文献   

7.
The structures of the first and the second transmembrane segment of the bovine mitochondrial oxoglutarate carrier (OGC) were studied by circular dichroism (CD) and nuclear magnetic resonance (NMR) spectroscopies. Peptides 21–46 and 78–108 of its primary sequence were synthesized and structurally characterized in membrane-mimetic environments. CD data showed that at high concentrations of TFE (>50%) and SDS (>2%) both peptides assume α-helical structures, whereas in more hydrophilic environments only peptide 78–108 has a helical structure. 1H-NMR spectra of the two peptides in TFE/water and SDS were fully assigned, and the secondary structures of the peptides were obtained from nuclear Overhauser effects, 3JαH-NH coupling constants and αH chemical shifts. The three-dimensional solution structures of the peptides in TFE/water were generated by distance geometry calculations. A well-defined α-helix was found in the region K24-V39 of peptide 21–46 and in the region A86–F106 of peptide 78–108. We cannot exclude that in intact OGC the extension of these helices is longer. The helix of peptide 21–46 is essentially hydrophobic, whereas that of peptide 78–108 is predominantly hydrophilic.  相似文献   

8.
The interaction of α-helical peptides with lipid bilayers is central to our understanding of the physicochemical principles of biological membrane organization and stability. Mutations that alter the position or orientation of an α-helix within a membrane, or that change the probability that the α-helix will insert into the membrane, can alter a range of membrane protein functions. We describe a comparative coarse-grained molecular dynamics simulation methodology, based on self-assembly of a lipid bilayer in the presence of an α-helical peptide, which allows us to model membrane transmembrane helix insertion. We validate this methodology against available experimental data for synthetic model peptides (WALP23 and LS3). Simulation-based estimates of apparent free energies of insertion into a bilayer of cystic fibrosis transmembrane regulator-derived helices correlate well with published data for translocon-mediated insertion. Comparison of values of the apparent free energy of insertion from self-assembly simulations with those from coarse-grained molecular dynamics potentials of mean force for model peptides, and with translocon-mediated insertion of cystic fibrosis transmembrane regulator-derived peptides suggests a nonequilibrium model of helix insertion into bilayers.  相似文献   

9.
A convenient detection of sugar chains in Taka-amylase A (TAA) was done by using 40 μg of enzyme, where a decrease in the UV absorption of NaIO4 during the periodate oxidation reaction was monitored. The periodate-oxidized sugar chain was labeled with a fluorescent reagent, N-1-ethylenediaminonaphthalene (EDAN), by incubation at pH 9.5 and 30°C for 1 h. The excess EDAN was removed by either quenching with o-phthaladehyde or Bio-Gel P-2 gel adsorption. Among the peptide fragments prepared from the EDAN-labeled TAA, a fluorescent peptide corresponding to the sugar chain was distinguished by the ODS column. These results suggest that periodate oxidation and subsequent fluorescent labeling were useful for the sensitive analysis of various glycoprotein samples.  相似文献   

10.
Affinity chromatography using immobilized S-protein was used for the screening of affinity peptide ligands from two soluble peptide libraries. Peptide library I consisted of octamers with glycine (G) at both termini of each peptide, i.e. GXXXXXXG. The six center positions were constructed using random sequences of six L-amino acids (Y, N, F, E, V, and L). Peptide library II also consisted of octamers but with glycine and valine (V) at both termini of each peptide (GVZZZZVG). The four variable center positions of peptide library II were random sequences of 18 L-amino acids. Peptides that were retained specifically on the immobilized S-protein column were eluted by 2% acetic acid. The peptides in the acid eluate were further separated using reversed-phase HPLC. Each separated peptide fraction was collected and the peptide sequences deconvoluted by mass spectrometry (MS/MS). The screenings of peptide libraries I and II resulted in 12 and 7 affinity peptides, respectively. Eight out of the twelve peptides from peptide library I contained the clear consensus sequence NFEV. Peptide library II resulted in affinity peptides with the sequences GVNFEVVG, GVNFTVVG and GVFFEL(I)VG. The advantages and limitations of affinity chromatography in peptide library screening are discussed.  相似文献   

11.
Two HLA-B27 subtypes, B*2702 and B*2705, both associated with ankylosing spondylitis, were tested for binding affinity with a panel of polyalanine model nonapeptides carrying Arg at position 2 (P2) and a series of different amino acids at position 9 (P9). The alpha chains were isolated from BTB(B*2705), C1R/B*2702 (a B*2702 transfectant cell line) and from the NW(B*2702) cell line that has a peculiar peptide presentation behavior. Peptide binding was measured by the HLA alpha chain refolding assay. The results obtained show that: 1) Peptides with basic residues (Arg and Lys) and also aliphatic (Leu) and aromatic (Phe and Tyr) peptides at P9 have a similar high affinity in the binding to B*2705; 2) B*2702 binds well to P9 aliphatic and aromatic peptides but only very weakly to P9 basic peptides. Since both B*2702 and B*2705 are associated with AS the presumed arthritogenic peptide is hypothesized to have an aromatic or aliphatic residue at position 9. Peptides with basic residues in this position would be excluded as candidates because of their low binding affinity with B*2702.  相似文献   

12.
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   

13.
Cyclomaltodextrin glucanotransferase was purified from B. circulans C31 through two successive steps of starch and Biogel column chromatography. The enzyme was purified up to 90-fold with a 30% yield. Its molecular weight was around 103,000. The purified enzyme converted 28% of the soluble starch to β-cyclodextrin at pH 7.0 and a substrate concentration of 5%. The optimum pH for the enzyme was found to be 5.5. The optimum temperature was 60°C. The enzyme optimum was stable from pH 5.5~9.0 and up to 50°C.  相似文献   

14.
One of the most important problems in vaccine development consists in understanding receptor–ligand interactions between Class II Major Histocompatibility Complex molecules (MHC II) and antigenic peptides involved in inducing an appropriate immune response. In this study, we used X-ray crystallography structural data provided by the HLA-DRβ1*0301–CLIP peptide interaction to compare native non-immunogenic and specifically-modified immunogenic peptides derived from the malarial SALSA protein, by analyzing molecular electrostatic potential surfaces on the most important regions of the peptide binding groove (Pockets 1, 4, 6 and 9). Important differences were found on the electrostatic potential induced by these peptides, particularly in MHC II conserved residues: Qα9, Sα53, Nα62, Nα69, Yβ30, Yβ60, Wβ61, Qβ70, Kβ71 and Vβ86, the same ones involved in establishing hydrogen bonds between Class II molecule-peptide and the recognition by T cell receptor, it correlating well with the change in their immunological properties.The results clearly suggest that modifications done on the electrostatic potential of these amino acids could favor the induction of different immune responses and therefore, their identification could allow modifying peptides a priori and in silico, so as to render them into immunogenic and protection-inducers and hence suitable components of a chemically-synthesized, multi-antigenic, minimal subunit based vaccine.  相似文献   

15.
Global internal standard technology (GIST) is being developed for the quantification of all primary structure and post-translational variants of proteins in a proteome. This paper is directed at an analysis of phosphorylation, primarily of serine and threonine. Quantification was achieved by acylation of primary amino groups in peptide cleavage fragments of proteins with isotopically coded derivatizing agents. Peptides from controls were globally coded with an isotopically "light" form of the reagent while those from experimental samples were coded with a "heavy" form of the reagent. The two types coding reagents used in this work were N-hydroxyl succinimide derivatives of acetate and 4-trimethylammoniumbutyrate. Heavy isotope forms were produced by deuteration of methyl groups. Subsequent to coding and mixing, the two samples were passed through a Ga(III) immobilized metal affinity chromatography (IMAC) column and the selected peptide fraction was further resolved by reversed-phase chromatography (RPC) and analyzed by mass spectrometry (MS). Relative differences in phosphopeptide concentration between samples were derived from isotope ratio measurements of the peptide isoforms observed in mass spectra. The method was validated with model peptides.  相似文献   

16.
Abstract

The peptide (35 residues) corresponding to the basic subdomain (bSD) of c-Jun (residues 252–281) and its fragments NP (N-terminal peptide, 1–19) and CP (C-terminal peptide, 1635) were synthesized in stepwise solid-phase using the tert-butyloxycarbonyl/benzyl strategy. In a previous paper, we have shown that during its binding to the DNA site CRE (cAMP- responsive element) the bSD structure was converted into α-helix from an initial random coil conformation [Krebs, D., Dahmani, B., El Antri, S., Monnot, M., Convert, O., Mauffret, O., Troalen, F. & Fermandjian, S. Eur. J. Biochem. 231, 370–380 (1995)]. Our results suggested both a high flexibility and a helical potential in bSD, these two properties seeming crucial for the accommodation of the basic subdomain of c-Jun to its specific DNA targets. In this work, we assessed the conformational variability of bSD through the study of the secondary structures of its NP and CP fragments in trifluoroethanol (TFE)/2H2O mixtures, using Fourier transform infrared (FTIR) spectroscopy. The IR results were critically analyzed in light of our previously reported circular dichroism (CD) and NMR data [Krebs, D., Dahmani, B., Monnot, M., Mauffret, O., Troalen, F. & Fermandjian, S. Eur. J. Biochem. 235, 699–712 (1996)]. Upon addition of TFE, the relative areas of the seven components of the amide I band (1700–1620 cm?1) reflected the conversion of a large amount of random coil conformation into α-helix for the two fragments and bSD. This effect was accompanied by more subtle variations of the less populated structures, in agreement with the results of CD and NMR experiments. The IR results stipulated the conservation of the parent bSD secondary structures in both fragments; however, NP and CP peptides did not display similar random-to-α-helix stabilization pattern upon additions of TFE to aqueous solutions. The profile from CD signal at 222 nm was found sigmoidal for NP and almost linear for CP, while that corresponding to the parent peptide bSD was just in between those of its fragments. Thus, the present study confirms the high flexibility and helix propensity of the c-Jun basic subdomain and suggests that the N- and C-terminal parts of the peptide do not follow the same random-to-helix conversion profile during their complexation with DNA.  相似文献   

17.
Being different from anti-phosphotyrosine antibodies, anti-phosphoserine- or anti-phosphothreonine-specific antibodies with high affinity for the detection of serine/threonine kinase substrates are not readily available. Therefore, chemical modification methods were developed for the detection of phosphoserine or threonine in the screening of protein kinase substrates based on β-elimination and Michael addition. We have developed a biotin-based detection probe for identification of the phosphorylated serine or threonine residue. A biotin derivative induced a color reaction using alkaline phosphate-conjugated streptavidin that amplified the signal. It was effective for the detection and separation of the target peptide on the resin. The detection probe was successfully used in identifying PKA substrates from peptide libraries on resin beads. The peptide library was prepared as a ladder-type, such that the active peptides on the colored resin beads were readily sequenced with the truncated peptide fragments by MALDI-TOF/MS analysis after releasing the peptides from the resin bead through photolysis.  相似文献   

18.
Lysine oligopeptides. Preparation by ion-exchange chromatography   总被引:3,自引:0,他引:3  
The preparation of L -lysine peptides (Lysn, n = 2–14) from polyL -lysine is described. Fractionation by ion-exchange column chromatography of poly-L -lysine hydrolysates on a preparative scale resulted in 0.2–1.0 g quantities of individual members of the poly-L -lysine series. The peptides isolated proved to be analytically pure and the optical configuration was fully retained, as demonstrated by complete enzymic digestion. Peptides higher than n = 14 were also prepared. They consisted of oligolysine groups of narrow and accurately determined size distribution. Potentiometric titrations were used both to characterize the products and to demonstrate the characteristic dependence of the dissociation constants on size of the peptide.  相似文献   

19.
Conformational energy computations have been carried out on the N-acetyl-N′-methylamide of 5-hydroxytryptophan (5OH-Trp) using ECEPP/3. As observed with tryptophan (Trp), the most preferred conformation about theC α ?C β bond of the side chain isg + ort. This preference is reduced to only thet conformational state when 5-hydroxyTrp is in the middle of a right-handed poly(l-alanine)α-helix. A similar result has been obtained with Trp [Pielaet al. (1987),Biopolymers 1987, 1273–1286]. These results suggest that replacement of Trp by its analog 5-hydroxyTrp may be tolerated in anα-helix. To test this hypothesis, we have replaced Trp by 5OH-Trp in the fifth helices of two functionally active mutants of the N-terminal domain of the bacteriophage λ repressor. Computations on the packing of these helices have shown that no significant structural changes result from the replacement of Trp by 5OH-Trp. The DNA-binding activity of these mutants, as assessed indirectly through geometrical parameters, is also unaltered.  相似文献   

20.
A simple purification procedure for raw starch-adsorbable and -digesting amylases (RSAs) was devised. The method depended on an affinity column, which was prepared by mixing raw corn starch and Hyflo Super-Cel. RSAs were specifically adsorbed on the matrix, and eluted with a buffer containing 1% β-cyclodextrin. This column could be used to purify RSAs from Streptomyces thermo-cyaneoviolaceus and a recombinant strain of E. coli.  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