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 urea and heat-induced unfolding-refolding behaviours of chicken egg white ovomucoid and its four fragments representing domains I, II + III, I + II and III were systematically investigated in 0.06 M sodium phosphate buffer (pH 7.0) by difference spectral measurements. The effect of temperature on ovomucoid and its fragments was also studied in 0.05 M sodium acetate buffer (pH 5.0) and in presence of 2 M urea at pH 7.0. Intrinsic viscosity data showed that ovomucoid and its different fragments did not lose any significant amount of their structure under mild acidic conditions (pH 4.6). Difference spectral results showed extensive disruption of the native structure by urea or temperature. Isothermal transitions showed single-step for domain I, domain I + II and domain III, and two-step having one stable intermediate, for ovomucoid and its fragment representing domain II + III. However, the presence of intermediate was not detected when the transitions were studied with temperature at pH 7.0. Strikingly, the single-step thermal transitions of ovomucoid and its fragment representing domain II + III, became two-step when measured either at pH 5.0 or in presence of 2 M urea at pH 7.0. Analysis of the equilibrium data on urea and heat denaturation showed that the second transition observed with ovomucoid or domain II + III represent the unfolding of domain III. The kinetic results of ovomucoid and its fragments indicate that the protein unfolds with three kinetic phases. A comparison of three rate constants for the unfolding of intact ovomucoid with that of its various fragments revealed that domain I, II and III of the protein correspond to the three kinetic phases having rate constants 0.456, 0.120 and 0.054 min-1, respectively. These data have led us to conclude: (i) the unusual stability of ovomucoid towards various denaturants, including temperature, is due to its domain III, (ii) initiation of the folding of the ovomucoid molecule starts from its NH2-terminal region which probably provides the nucleation site for the formation of the subsequent structure and (iii) domains I and II have greater mutual recognition between them as compared to the recognition either of them have with domain III.  相似文献   

2.
The primary structure of ovomucoid shows considerable sequence homology at three contiguous regions which form structural domains I, II and III. In order to see whether or not the three domains fold similarly and acquire similar overall native conformation/shape, two fragments A and C were obtained by controlled peptic digestion of ovomucoid. The two fragments were investigated for their chemical composition, molecular weight, anti-tryptic activity, hydrodynamic behaviour, optical properties and acid denaturation. Results on molecular weight, amino acid composition and inhibitory acitivity show that the fragments A and C correspond respectively to domain I-II and domain III. Optical data suggested more exposure of tyrosine residues in the fragments than in the intact molecule. Domain III exists in a compact and globular conformation under native conditions whereas domain I-II and ovomucoid appear to possess asymmetric conformation. Results on acid denaturation show that the process is thermodynamically reversible and that inter-domain separation probably precedes denaturation of domains during acidification of ovomucoid.  相似文献   

3.
Ovomucoid domains: preparation and physico-chemical characterization   总被引:1,自引:0,他引:1  
Four fragments of ovomucoid representing its individual domains and their different combinations were prepared by peptic and cyanogen bromide cleavages of the protein. The fragments corresponding to domains I + II, II + III, I and III of the parent ovomucoid molecule, were found to be homogeneous by gel filtration and polyacrylamide gel electrophoresis in presence and absence of SDS. Various physico-chemical properties of these proteins, such as molecular weight, NH2- and COOH-terminal amino acid residues, sugar content, isoionic pH, specific extinction coefficient, fluorescence emission spectra, intrinsic viscosity, frictional coefficient, Stokes radius, diffusion coefficient and geometrical mean radius were determined. Analysis of the results on trypsin inhibitory activity of ovomucoid and its different fragments suggested that only domain II is involved in the antitryptic activity of the inhibitor. Optical characteristics of these fragments indicate that they are devoid of tryptophan residues. The hydrodynamic properties suggest that intact ovomucoid and two of its fragments (domain I + II and domain II + III) are significantly different from those of typical globular proteins and are asymmetric in nature. However, the shape of the two remaining fragments representing domains I and III of the intact protein appeared to be compact and globular. Furthermore, domain II of ovomucoid has been suggested to primarily contribute towards the apparent asymmetry in the intact protein.  相似文献   

4.
The toxic fragment of Bacillus thuringiensis crystal proteins consists of three distinct structural domains. There is evidence that domain I is involved in pore formation and that domain II is involved in receptor binding and specificity. It has been found that, in some cases, domain III is also important in determining specificity. Furthermore, involvement of domain III in binding has also been reported recently. To investigate the role of toxin domains in the diamondback moth (Plutella xylostella), we used hybrid toxins with domain III substitutions among Cry1C, Cry1E, and Cry1Ab. Neither Cry1E nor G27 (a hybrid with domains I and II from Cry1E and domain III from Cry1C) was toxic, whereas Cry1C and F26 (the reciprocal hybrid) were equally toxic. H04 (a hybrid with domains I and II from Cry1Ab and domain III from Cry1C) showed toxicity that was of a similar level as that of Cry1Ab and significantly higher than that of Cry1C. Binding assays with 125I-Cry1C showed that Cry1C and F26 competed for the same binding sites on midgut membrane vesicles, whereas Cry1E, G27, and H04 did not bind to these sites. Our results show that, in contrast to findings in other insects for the toxins and hybrids used here, toxin specificity as well as specificity of binding to membrane vesicles in the diamondback moth is mediated by domain II (and/or I) and not by domain III.  相似文献   

5.
Interactions among the three structural domains of Bacillus thuringiensis Cry1 toxins were investigated by functional analysis of chimeric proteins. Hybrid genes were prepared by exchanging the regions coding for either domain I or domain III among Cry1Ab, Cry1Ac, Cry1C, and Cry1E. The activity of the purified trypsin-activated chimeric toxins was evaluated by testing their effects on the viability and plasma membrane permeability of Sf9 cells. Among the parental toxins, only Cry1C was active against these cells and only chimeras possessing domain II from Cry1C were functional. Combination of domain I from Cry1E with domains II and III from Cry1C, however, resulted in an inactive toxin, indicating that domain II from an active toxin is necessary, but not sufficient, for activity. Pores formed by chimeric toxins in which domain I was from Cry1Ab or Cry1Ac were slightly smaller than those formed by toxins in which domain I was from Cry1C. The properties of the pores formed by the chimeras are therefore likely to result from an interaction between domain I and domain II or III. Domain III appears to modulate the activity of the chimeric toxins: combination of domain III from Cry1Ab with domains I and II of Cry1C gave a protein which was more strongly active than Cry1C.  相似文献   

6.
The carbohydrate of ovomucoid was analyzed for components I, II, III and IV which were, fractionated by CMC-column chromatography. The total hexose content and the molar ratio of d-mannose to d-galactose (4:1) were identical in each component, but the d-glucosamine and sialic acid contents were found to be higher in components I and II (both are trypsin inhibitors) compared with components III and IV (both are apo-proteins of flavomucoid). The amino acid composition of each component of ovomucoid varied considerably. There were remarkable differences in the amino acid composition between components I and II, both had an antitryptic activity. The N-terminal amino acid of components I and II was alanine and in the case of components III and IV, threonine was found on the N-terminal. The free carboxylic residue of sialic acid was found to be responsible for the negative charge of ovomucoid, and its electrophoretic heterogeneity was reaffirmed by paper electrophoresis. It is evident from the ultracentrifugal analysis that the four components of ovomucoid have a similar molecular size.  相似文献   

7.
A non-phenolic β-O-4 lignin substructure model, 4-ethoxy-3-methoxyphenylglycerol-β-syringaldehyde ether (I), was metabolized by a ligninolytic culture of Coriolus versicolor. Based on the identification of the metabolic products (II~XI), the following reactions were found to occur in the culture; a) oxidation (III) and reduction (II) at the benzyl (Cα′) position of the substrate (I), b) β-ether cleavage to give arylglycerols (IV, V), and c) Cα-Cβ cleavage of the arylglycerols and/or arylglycerol moiety of the substrate (I). In addition, β-deoxy diol (VI) and γ-formylglycerol (VII) were obtained as degradation products from substrate (I).  相似文献   

8.
Four components of ovomucoid were digested exhaustively and four kinds of glycopeptide corresponding to the four components were separated by gel filtration. Each glycopeptide was shown to be homogenious by paper chromatography and paper electrophoresis. Molar ratios of carbohydrate components of these glycopeptides varied to some extent but the amino acid compositions of these glycopeptides were essentially identical with each other with the exception of alanine. Aspartic acid and threonine were predominant amino acids in the all glycopeptides. It is most likely that the modes of linkages between polysaccharide and protein in individual ovomucoid I, II, III and IV are essentially the same, and that the carbohydrate moiety is linked to the protein via asparaginyl residue or the hydroxyl group of threonine, although the possibility of the linkages to glutamine and serine can not be excluded.  相似文献   

9.
New types of azidoaryl analogs of GTP: γ-(4-azido)anilide of GTP (I), γ-(N-(4-azidobenzyl)-N-methyl)amide of GTP (II) and of GDP: β-(4-azido)anilide of GDP (III), β-(N-(4-azidobenzyl)-N-methyl)amide of GDP (IV) have been synthesized by treatment of the nucleotide in aqueous solution with N-cyclohexyl-N′-β-(4-methylmorpholinium)- ethylcarbodiimidep-toluene sulfonate and the respective amine. The analog of GTP bearing at the γ-phosphate an alkylating 2-chloroethylamino group: γ-(4-N-(2-chloroethyl)-N-methylaminobenzyl)amide of GTP (V) was prepared by the method described previously for the preparation of the analog of ATP (Knorre, D.G., Kurbatov, V.A. and Samukov, V.V. (1976) FEBS Lett. 70, 105–108). Azidoaryl analogs of GTP and GDP as well as the chloroethylaminoaryl analog of GTP compete with GDP in the formation of the binary complex EF-Tu·GDP with the respective Ki values 3.9·10?7 M (I), 2.9·10?8 M (II), 6.9·10?7 M (III), 5.0·10?7 M (IV) and 3.8·10?8 M (V) relative to GDP. constants of the complexes of the radioactively-labeled GTP analogs I, II and V with elongation factor Tu were calculated to be 8.5·10?6 M, 3.4·10?7 M and 4.6·10?8 M, respectively, or approx. 1740-, 70- and 9-times greater than that of GDP. GTP analogs I, II and V were found to substitute GTP in the stimulation of EF-Tu-dependent binding of aminoacyl-tRNA to the ribosome-mRNA complex.  相似文献   

10.
The Clostridium botulinum type C 16S progenitor toxin contains a neurotoxin and several nontoxic components, designated nontoxic nonhemagglutinin (HA), HA1 (HA-33), HA2 (HA-17), HA3a (HA-22-23), and HA3b (HA-53). The HA3b subcomponent seems to play an important role cooperatively with HA1 in the internalization of the toxin by gastrointestinal epithelial cells via binding of these subcomponents to specific oligosaccharides. In this study, we investigated the sugar-binding specificity of the HA3b subcomponent using recombinant protein fused to glutathione S-transferase and determined the three-dimensional structure of the HA3a-HA3b complex based on X-ray crystallography. The crystal structure was determined at a resolution of 2.6 Å. HA3b contains three domains, domains I to III, and the structure of domain I resembles HA3a. In crystal packing, three HA3a-HA3b molecules are assembled to form a three-leaved propeller-like structure. The three HA3b domain I and three HA3a alternate, forming a trimer of dimers. In a database search, no proteins with high structural homology to any of the domains (Z score > 10) were found. Especially, HA3a and HA3b domain I, mainly composed of β-sheets, reveal a unique fold. In binding assays, HA3b bound sialic acid with high affinity, but did not bind galactose, N-acetylgalactosamine, or N-acetylglucosamine. The electron density of liganded N-acetylneuraminic acid was determined by crystal soaking. In the sugar-complex structure, the N-acetylneuraminic acid-binding site was located in the cleft formed between domains II and III of HA3b. This report provides the first determination of the three-dimensional structure of the HA3a-HA3b complex and its sialic acid binding site. Our results will provide useful information for elucidating the mechanism of assembly of the C16S toxin and for understanding the interactions with oligosaccharides on epithelial cells and internalization of the botulinum toxin complex.  相似文献   

11.
The cDNA encoding the 182 amino acid long precursor stellacyanin from Cucumis sativus was isolated and characterized. The protein precursor consists of four sequence domains: I, a 23 amino acid hydrophobic N-terminal signal peptide with features characteristic of secretory proteins; II, a 109 amino acid copper-binding domain; III, a 26 amino acid hydroxyproline- and serine-rich peptide characteristic of motifs found in the extension family, extracellular structural glycoproteins found in plant cell walls; and IV, a 22 amino acid hydrophobic extension. Maturation of the protein involves posttranslational processing of domains I and IV. The copper-binding domain (domain II), which shares high sequence identity with other stellacyanins, has been expressed without its carbohydrate attachment sites, refolded from the Escherichia coli inclusion bodies, purified, and characterized by electronic absorption, EPR, ESEEM, and RR spectroscopy. Its spectroscopic properties are nearly identical to those of stellacyanin from the Japanese lacquer tree Rhus vernicifera, the most extensively studied and best characterized stellacyanin, indicating that this domain folds correctly, even in the absence of its carbohydrate moiety. The presence of a hydroxyproline- and serine-rich domain III suggests that stellacyanin may have a function other than that of a diffusible electron transfer protein, conceivably participating in redox reactions localized at the plant cell wall, which are known to occur in response to wounding or infection of the plant.  相似文献   

12.
In order to test our hypothesis that Bacillus thuringiensis delta-endotoxin Cry1Ca domain III functions as a determinant of specificity for Spodoptera exigua, regardless of the origins of domains I and II, we have constructed by cloning and in vivo recombination a collection of hybrid proteins containing domains I and II of various Cry1 toxins combined with domain III of Cry1Ca. Cry1Ab, Cry1Ac, Cry1Ba, Cry1Ea, and Cry1Fa all become more active against S. exigua when their domain III is replaced by (part of) that of Cry1Ca. This result shows that domain III of Cry1Ca is an important and versatile determinant of S. exigua specificity. The toxicity of the hybrids varied by a factor of 40, indicating that domain I and/or II modulate the activity as well. Cry1Da-Cry1Ca hybrids were an exception in that they were not significantly active against S. exigua or Manduca sexta, whereas both parental proteins were highly toxic. Incidentally, in a Cry1Ba-Cry1Ca hybrid, Cry1Ca domain III can also strongly increase toxicity for M. sexta.  相似文献   

13.
Six xylan-hydrolyzing enzymes have been isolated from the preparations Celloviridin G20x and Xybeten-Xyl, obtained earlier based on the strain 1 Trichoderma longibrachiatum (Trichoderma reesei) TW-1. The enzymes isolated were represented by three xylanases (XYLs), XYL I (20 kDa, pI 5.5), XYL II (21 kDa, pI 9.5), XYL III (30 kDa, pI 9.1); endoglucanase I (EG I), an enzyme exhibiting xylanase activity (57 kDa, pI 4.6); and two exodepolymerases, β-xylosidase (β-XYL; 80 kDa, pI 4.5) and α-L-arabinofuranosidase I (α-L-AF I; 55 kDa, pI 7.4). The substrate specificity of the enzymes isolated was determined. XYL II exhibited maximum specific xylanase activity (190 U/mg). The content of the enzymes in the preparation was assessed. Maximum contributions to the total xylanase activities of preparations Celloviridin G20x and Xybeten-Xyl were made by EG I and XYL II, respectively. Effects of temperature and pH on the enzyme activities, their stabilities under various conditions, and the kinetics of exhaustive hydrolysis of glucuronoxylan and arabinoxylan were studied. Combinations of endodepolymerases (XYL I, XYL II, XYL III, or EG I) and exodepolymerases (α-L-AF I or β-XYL) produced synergistic effects on arabinoxylan cleavage. The reverse was the case when endodepolymerases, such as XYL I or EG I, were combined with α-L-AF I.  相似文献   

14.
The conformational features of a chemically synthesized 23-residue glycopeptide construct (II) carrying Gal-beta-(1,3)-alpha-GalNAc and its deglycosylated counterpart (I; Gal: galactose; GalNAc: N-acetyl galactosamine) derived from the C-terminal domain of human salivary mucin (MUC7) were investigated using CD spectroscopy as well as molecular dynamic simulation studies. The corresponding deglycosylated peptide (I) was essentially used to compare and study the influence of the sugar moiety on peptide backbone conformation. CD measurements in aqueous medium revealed that the apopeptide (I) contains significant populations of beta-strand conformation while the glycopeptide (II) possess, partly, helical structure. This transition in the secondary structure upon glycosylation from beta-strand to helical conformation clearly demonstrates that the carbohydrate moiety exerts significant influence on the peptide backbone. On the other hand, upon titrating structure stabilizing organic cosolvent, trifluoroethanol (TFE), both the peptides showed pronounced helical structure. However, the propensity for helical structure formation is less pronounced in glycopeptide compared to apopeptide suggesting that the bulky carbohydrate moiety possibly posing steric hindrance to the formation of TFE-induced secondary structure in II. Energy-minimized molecular model for the glycopeptide revealed that the preferred helix conformation in aqueous medium appears to be stabilized by the hydrogen-bonded salt bridge like interaction between carbohydrate --OH and Lys-10 side--N(+)H(3) group. Size exclusion chromatographic analysis of both (glyco)peptides I and II showed an apparent Kd of 2.3 and 0.52 microM, respectively, indicating that glycopeptide (II) has greater tendency for self-association. Due to high amphipathic character as well as due to the presence of a leucine zipper motif ( approximately LLYMKNLL approximately ), which is known to increase the stability at the coiled-coil interface via hydrophobic interactions, we propose therefore that, this domain could be one of the key elements involved in the self-association of intact MUC7 in vivo. Profound conformational effects governed by glycosylation exemplified herein could have implications in determining structure-function relationships of mucin glycoproteins.  相似文献   

15.
The third domain of Japanese quail ovomucoid, a Kazal type inhibitor, has been crystallized and its crystal structure determined at 2.5 Å resolution using multiple isomorphous replacement techniques. The asymmetric unit contains four molecules. In the crystal the molecules are arranged in two slightly different octamers with approximate D4 symmetry. The molecules are held together mainly by interactions of the N-terminal residues, which form a novel secondary structural element, a β-channel.The molecule is globular with approximate dimensions 35 Å × 27 Å × 19 Å. The secondary structural elements are a double-stranded anti-parallel β-sheet of residues Pro22 to Gly32 and an α-helix from Asn33 to Ser44. The reactive site Lys18-Asp19 is located in an exposed loop. It is close to Asn33 at the N terminus of the helical segment. The polypeptide chain folding of ovomucoid bears some resemblance to other inhibitors in the existence of an anti-parallel double strand following the reactive site loop.  相似文献   

16.
Immunochemical studies on thermal denaturation of ovomucoid   总被引:1,自引:0,他引:1  
The thermal denaturation of ovomucoid was investigated by immunochemical methods, namely immunoprecipitation analyses and antibody-Sepharose 4B column chromatography. In the immunoprecipitation analyses, heated ovomucoid (90 degrees C, 90 min, pH 7.2) required about twice the antigen addition of the native protein to approach maximal precipitation with specific antibody, and the maximal immunoprecipitation was decreased to 80% of that by native ovomucoid. However, heated protein inhibited the binding of antibody with native ovomucoid, and 100% inhibition was attained at about 4-times the antigen addition necessary for the native protein. Heated ovomucoid (100 degrees C, 120 min) showed little immunoprecipitation and inhibition. To ovomucoid antigenicity was diminished more slowly than the trypsin inhibitory activity by heating, e.g., heated ovomucoid (90 degrees C, 120 min) retained more than 30% of the antigenicity but little trypsin inhibitory activity. By passing through the immunoaffinity column, heated ovomucoid (90 degrees C, 90 min) was separated into two fractions, either with (fraction II) or without (fraction I) antigenicity. Fraction II contained smaller fractions of ordered secondary structure than native ovomucoid, and trypsin inhibitory activity of fraction II was only 24% of the native one. These results indicated that thermally denatured ovomucoid was heterogeneous regarding the conformational damage caused by heating, and the structure around some antigenic sites in an ovomucoid molecule was retained even after the backbone conformation was partially destroyed and trypsin inhibitory activity was lost.  相似文献   

17.
Streptococcus gordonii is a primary colonizer and is involved in the formation of dental plaque. This bacterium expresses several surface proteins. One of them is the adhesin SspB, which is a member of the Antigen I/II family of proteins. SspB is a large multi-domain protein that has interactions with surface molecules on other bacteria and on host cells, and is thus a key factor in the formation of biofilms. Here, we report the crystal structure of a truncated form of the SspB C-terminal domain, solved by single-wavelength anomalous dispersion to 1.5 Å resolution. The structure represents the first of a C-terminal domain from a streptococcal Antigen I/II protein and is comprised of two structurally related β-sandwich domains, C2 and C3, both with a Ca2+ bound in equivalent positions. In each of the domains, a covalent isopeptide bond is observed between a lysine and an asparagine, a feature that is believed to be a common stabilization mechanism in Gram-positive surface proteins. S. gordonii biofilms contain attachment sites for the periodontal pathogen Porphyromonas gingivalis and the SspB C-terminal domain has been shown to have one such recognition motif, the SspB adherence region. The motif protrudes from the protein, and serves as a handle for attachment. The structure suggests several additional putative binding surfaces, and other binding clefts may be created when the full-length protein is folded.  相似文献   

18.
Knockdown resistance (kdr) mutations in the voltage-gated sodium channel (VGSC) of mosquitoes confer resistance to insecticides. Although insecticide resistance has been suspected to be widespread in the natural population of Aedes aegypti in Myanmar, only limited information is currently available. The overall prevalence and distribution of kdr mutations was analyzed in Ae. aegypti from Mandalay areas, Myanmar. Sequence analysis of the VGSC in Ae. aegypti from Myanmar revealed amino acid mutations at 13 and 11 positions in domains II and III of VGSC, respectively. High frequencies of S989P (68.6%), V1016G (73.5%), and F1534C (40.1%) were found in domains II and III. T1520I was also found, but the frequency was low (8.1%). The frequency of S989P/V1016G was high (55.0%), and the frequencies of V1016G/F1534C and S989P/V1016G/F1534C were also high at 30.1% and 23.5%, respectively. Novel mutations in domain II (L963Q, M976I, V977A, M994T, L995F, V996M/A, D998N, V999A, N1013D, and F1020S) and domain III (K1514R, Y1523H, V1529A, F1534L, F1537S, V1546A, F1551S, G1581D, and K1584R) were also identified. These results collectively suggest that high frequencies of kdr mutations were identified in Myanmar Ae. aegypti, indicating a high level of insecticide resistance.  相似文献   

19.
Hen ovomucoid was chemically deglycosylated by treatment with trifluoromethanesulfonic acid at 0 degrees C for 60 min. About 75 mol% of the carbohydrate moiety was removed from the glycoprotein without changing its amino acid composition, and its trypsin inhibitory activity and immunoreactivity with specific antibodies remained unchanged. The deglycosylated ovomucoid was inactivated and degraded easily by an excess amount of trypsin, whereas the native glycoprotein was not. Furthermore, the biological and immunological activities of the deglycosylated ovomucoid were lowered by heat treatment more easily than those of the native ovomucoid. These results suggest that the carbohydrate moiety of ovomucoid contributes to the stability of the ovomucoid molecule against tryptic hydrolysis and heat denaturation.  相似文献   

20.
Cell-free extracts from leaves of Tanacetum vulgare synthesised geraniol and nerol (3,7-dimethylocta-trans-2-ene-1-ol and its cis isomer) in up to 11·9 and 2·4% total yields from IPP-[4-14C] and MVA-[2-14C] respectively. Optimum preparations were obtained from plant material just before the onset of flowering. The ratio of the monoterpenols varied 28-fold for different preparations under conditions where these products or their phosphate esters were not interconverted. Similar extracts incorporated α-terpineol-[14C] and terpinen-4-ol-[14C] (p-menth-1-en-8- and -4-ol respectively) in 0·05 to 2·2% yields into a compound tentatively identified as isothujone (trans-thujan-3-one), and preparations from flowerheads converted IPP-[4-14C] in 2·7% yield into geranyl and neryl β-d-glucosides. Inhibitors of IPP-isomerase had little effect on the incorporation of IPP into the monoterpenols in cell-free systems from which endogenous compounds of low molecular-weight had been removed. The inference that a pool of protein-bonded DMAPP or its biogenetic equivalent was present was supported by the demonstration that geraniol and nerol biosynthesised in the absence of the inhibitors were predominantly (65 to 100%) labelled in the moiety derived from IPP.  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