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.
Aim: To study genetic diversity of Chromobacterium haemolyticum isolates recovered from a natural tropical lake. Methods and Results: A set of 31 isolates were recovered from a bacterial freshwater community by conventional plating methods and subjected to genetic and phenotypic characterization. The 16S ribosomal RNA (rRNA) gene phylogeny revealed that the isolates were related most closely with C. haemolyticum. In addition to the molecular data, our isolates exhibited strong β‐haemolytic activity, were nonviolacein producers and utilized i‐inositol, d ‐mannitol and d ‐sorbitol in contrast with the other known chromobacteria. Evaluation of the genetic diversity in the 16S rRNA gene, tRNA intergenic spacers (tDNA) and 16S‐23S internal transcribed spacers (ITS) unveiled different levels of genetic heterogeneity in the population, which were also observed with repetitive extragenic palindromic (rep)‐PCR genomic fingerprinting using the BOX‐AR1 primer. tDNA‐ and ITS‐PCR analyses were partially congruent with the 16S rRNA gene phylogeny. The isolates exhibited high resistance to β‐lactamic antibiotics. Conclusion: The population genetic heterogeneity was revealed by 16S rRNA gene sequence, ITS and BOX‐PCR analysis. Significance and Impact of the Study: This study provides for the first time an insight into the genetic diversity of phylogenetically close isolates to C. haemolyticum species.  相似文献   

2.
The crystal structure of a yeast hypothetical protein with sequence similarity to CN hydrolases has been determined to 2.4 A resolution by the multiwavelength anomalous dispersion (MAD) method. The protein folds as a four-layer alphabetabetaalpha sandwich and exists as a dimer in the crystal and in solution. It was selected in a structural genomics project as representative of CN hydrolases at a time when no structures had been determined for members of this family. Structures for two other members of the family have since been reported and the three proteins have similar topology and dimerization modes, which are distinct from those of other alphabetabetaalpha proteins whose structures are known. The dimers form an unusual eight-layer alphabetabetaalpha:alphabetabetaalpha structure. Although the precise enzymatic reactions catalyzed by the yeast protein are not known, considerable information about the active site may be deduced from conserved sequence motifs, comparative biochemical information, and comparison with known structures of hydrolase active sites. As with serine hydrolases, the active-site nucleophile (cysteine in this case) is positioned on a nucleophile elbow.  相似文献   

3.
Yersinia enterocolitica is a Gram‐negative bacterium that causes yersiniosis, a zoonotic disease affecting the gastrointestinal tract of humans, cattle, and pigs, among others. The lipopolysaccharide of Y. enterocolitica O:8 contains an unusual sugar, 6‐deoxy‐d ‐gulose, which requires four enzymes for its biosynthesis. Here, we describe a combined structural and functional investigation of WbcA, which catalyzes the third step in the pathway, namely an epimerization about the C‐3′ carbon of a CDP‐linked sugar. The structure of WbcA was determined to 1.75‐Å resolution, and the model was refined to an overall R‐factor of 19.5%. The fold of WbcA places it into the well‐defined cupin superfamily of sugar epimerases. Typically, these enzymes contain both a conserved histidine and a tyrosine residue that play key roles in catalysis. On the basis of amino acid sequence alignments, it was anticipated that the “conserved” tyrosine had been replaced with a cysteine residue in WbcA (Cys 133), and indeed this was the case. However, what was not anticipated was the fact that another tyrosine residue (Tyr 50) situated on a neighboring β‐strand moved into the active site. Site‐directed mutant proteins were subsequently constructed and their kinetic properties analyzed to address the roles of Cys 133 and Tyr 50 in WbcA catalysis. This study emphasizes the continuing need to experimentally verify assumptions that are based solely on bioinformatics approaches.  相似文献   

4.
N‐formylated sugars have been observed on the O‐antigens of such pathogenic Gram‐negative bacteria as Campylobacter jejuni and Francisella tularensis. Until recently, however, little was known regarding the overall molecular architectures of the N‐formyltransferases that are required for the biosynthesis of these unusual sugars. Here we demonstrate that the protein encoded by the wbtj gene from F. tularensis is an N‐formyltransferase that functions on dTDP‐4‐amino‐4,6‐dideoxy‐d ‐glucose as its substrate. The enzyme, hereafter referred to as WbtJ, demonstrates a strict requirement for N10‐formyltetrahydrofolate as its carbon source. In addition to the kinetic analysis, the three‐dimensional structure of the enzyme was solved in the presence of dTDP‐sugar ligands to a nominal resolution of 2.1 Å. Each subunit of the dimeric enzyme is dominated by a “core” domain defined by Met 1 to Ser 185. This core motif harbors the active site residues. Following the core domain, the last 56 residues fold into two α‐helices and a β‐hairpin motif. The hairpin motif is responsible primarily for the subunit:subunit interface, which is characterized by a rather hydrophobic pocket. From the study presented here, it is now known that WbtJ functions on C‐4′ amino sugars. Another enzyme recently investigated in the laboratory, WlaRD, formylates only C‐3′ amino sugars. Strikingly, the quaternary structures of WbtJ and WlaRD are remarkably different. In addition, there are several significant variations in the side chains that line their active site pockets, which may be important for substrate specificity. Details concerning the kinetic and structural properties of WbtJ are presented.  相似文献   

5.
Pantoea ananatis is a Gram‐negative bacterium first recognized in 1928 as the causative agent of pineapple rot in the Philippines. Since then various strains of the organism have been implicated in the devastation of agriculturally important crops. Some strains, however, have been shown to function as non‐pathogenic plant growth promoting organisms. To date, the factors that determine pathogenicity or lack thereof between the various strains are not well understood. All P. ananatis strains contain lipopolysaccharides, which differ with respect to the identities of their associated sugars. Given our research interest on the presence of the unusual sugar, 4‐formamido‐4,6‐dideoxy‐d ‐glucose, found on the lipopolysaccharides of Campylobacter jejuni and Francisella tularensis, we were curious as to whether other bacteria have the appropriate biosynthetic machinery to produce these unique carbohydrates. Four enzymes are typically required for their biosynthesis: a thymidylyltransferase, a 4,6‐dehydratase, an aminotransferase, and an N‐formyltransferase. Here, we report that the gene SAMN03097714_1080 from the P. ananatis strain NFR11 does, indeed, encode for an N‐formyltransferase, hereafter referred to as PA1080c. Our kinetic analysis demonstrates that PA1080c displays classical Michaelis–Menten kinetics with dTDP‐4‐amino‐4,6‐dideoxy‐d ‐glucose as the substrate and N10‐formyltetrahydrofolate as the carbon source. In addition, the X‐ray structure of PA1080c, determined to 1.7 Å resolution, shows that the enzyme adopts the molecular architecture observed for other sugar N‐formyltransferases. Analysis of the P. ananatis NFR11 genome suggests that the three other enzymes necessary for N‐formylated sugar biosynthesis are also present. Intriguingly, those strains of P. ananatis that are non‐pathogenic apparently do not contain these genes.  相似文献   

6.
7.
Unusual di- and trideoxysugars are often found on the O-antigens of Gram-negative bacteria, on the S-layers of Gram-positive bacteria, and on various natural products. One such sugar is 3-acetamido-3,6-dideoxy-d-glucose. A key step in its biosynthesis, catalyzed by a 3,4-ketoisomerase, is the conversion of thymidine diphosphate (dTDP)−4-keto-6-deoxyglucose to dTDP-3-keto-6-deoxyglucose. Here we report an X-ray analysis of a 3,4-ketoisomerase from Thermoanaerobacterium thermosaccharolyticum. For this investigation, the wild-type enzyme, referred to as QdtA, was crystallized in the presence of dTDP and its structure solved to 2.0-Å resolution. The dimeric enzyme adopts a three-dimensional architecture that is characteristic for proteins belonging to the cupin superfamily. In order to trap the dTDP-4-keto-6-deoxyglucose substrate into the active site, a mutant protein, H51N, was subsequently constructed, and the structure of this protein in complex with the dTDP–sugar ligand was solved to 1.9-Å resolution. Taken together, the structures suggest that His 51 serves as a catalytic base, that Tyr 37 likely functions as a catalytic acid, and that His 53 provides a proton shuttle between the C-3′ hydroxyl and the C-4′ keto group of the hexose. This study reports the first three-dimensional structure of a 3,4-ketoisomerase in complex with its dTDP–sugar substrate and thus sheds new molecular insight into this fascinating class of enzymes.  相似文献   

8.
A salt‐tolerant esterase, designated H9Est, was identified from a metagenomic library of the Karuola glacier. H9Est gene comprised 1071 bp and encoded a polypeptide of 357 amino acids with a molecular mass of 40 kDa. Sequence analysis revealed that H9Est belonged to the family IV of bacterial lypolitic enzyme. H9Est was overexpressed in Escherichia coli and the purified enzyme showed hydrolytic activity towards p‐nitrophenyl esters with carbon chain from 2 to 8. The optimal esterase activity was at 40°C and pH 8.0 and the enzyme retained its activity towards some miscible organic solvents such as polyethylene glycol. A three‐dimensional model of H9Est revealed that S200, D294, and H324 formed the H9Est catalytic triad. Circular Dichroism spectra and molecular dynamic simulation indicated that the esterase had a wide denaturation temperature range and flexible loops that would be beneficial for H9Est performance at low temperatures while retaining heat‐resistant features. © 2015 American Institute of Chemical Engineers Biotechnol. Prog., 31:890–899, 2015  相似文献   

9.
A combinatorial peptide library contains an enormous combination of amino acid sequences and drug candidates, but an effective screening strategy to identify a variety of bioactive peptides has yet to be established. In this article, a random hexapeptide library was screened to identify novel peptide ligands for a 5-oxo-ETE receptor (OXER), which is a G-protein-coupled receptor for bioactive lipids, by using an OXER-Gi1alpha fusion protein. We successfully identified 2 hexapeptides-Ac-HMQLYF-NH2 and Ac-HMWLYF-NH(2)-that exhibited agonistic activity. Although the corresponding affinities were relatively low (EC50 values of 146 and 6.7 microM, respectively), the activities were confirmed by other independent cell-based assay methods, namely, intracellular calcium mobilization and cell chemotaxis. This study demonstrates that a combinatorial peptide library may be screened using a [35S]GTPgammaS binding assay with G-protein-coupled receptor (GPCR)-Galpha fusion proteins, in general, and that of peptide ligands can be obtained even for nonpeptide receptors.  相似文献   

10.
Poland D 《Biopolymers》2004,73(2):216-228
In this article we give a new method for the calculation of DNA melting profiles. Based on the matrix formulation of the DNA partition function, the method relies for its efficiency on the fact that the required matrices are very sparse, essentially reducing matrix multiplication to vector multiplication and thus making the computer time required to treat a DNA molecule containing N base pairs proportional to N(2). A key ingredient in the method is the result that multiplication by the inverse matrix can also be reduced to vector multiplication. The task of calculating the melting profile for the entire genome is further reduced by treating regions of the molecule between helix-plateaus, thus breaking the molecule up into independent parts that can each be treated individually. The method is easily modified to incorporate changes in the assignment of statistical weights to the different structural features of DNA. We illustrate the method using the genome of Haemophilus influenzae.  相似文献   

11.
12.
13.
This study describes the identification and structural characterization of Sus scrofa statherin. HPLC–electrospray ionization mass spectrometry analysis on pig parotid secretory granule extracts evidenced a peptide with a molecular mass value of 5381.1 ± 0.6 Da and its truncated form, devoid of the C‐terminal Ala residue, with a molecular mass value of 5310.1 ± 0.6 Da. The complete sequence of pig statherin gene was determined by sequencing the full‐length cDNA obtained by rapid amplification of cDNA ends. The gene is 549 base pairs long and contains an open reading frame of 185 nucleotides, encoding a 42‐amino acid secretory polypeptide with a signal peptide of 19 residues. This sequence presents some typical features of the four statherins characterized till now, showing the highest degree of amino acid identity with bovine (57%) and human statherin (39%). Pig statherin is mono‐phoshorylated on Ser‐3, while primate statherins already characterized are di‐phosphorylated on Ser‐2 and Ser‐3. This difference, probably connected to the Asp‐4 → Glu substitution, suggests the involvement of the Golgi‐casein kinase, which strictly recognizes the SX(E/pS) consensus sequence. Copyright © 2010 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

14.
A novel conotoxin named lt6c, an O‐superfamily conotoxin, was identified from the cDNA library of venom duct of Conus litteratus. The full‐length cDNA contains an open reading frame encoding a predicted 22‐residue signal peptide, a 22‐residue proregion and a mature peptide of 28 amino acids. The signal peptide sequence of lt6c is highly conserved in O‐superfamily conotoxins and the mature peptide consists of six cysteines arranged in the pattern of C? C? CC? C? C that is defined the O‐superfamily of conotoxins. The mature peptide fused with thioredoxin, 6‐His tag, and a Factor Xa cleavage site was successfully expressed in Escherichia coli. About 12 mg lt6c was purified from 1L culture. Under whole‐cell patch‐clamp mode, lt6c inhibited sodium currents on adult rat dorsal root ganglion neurons. Therefore, lt6c is a novel O‐superfamily conotoxin that is able to block sodium channels. Copyright © 2008 European Peptide Society and John Wiley & Sons, Ltd.  相似文献   

15.
16.
Balanophora involucrata Hook .f. & Thomson (Balanophoraceae) is a parasite plant often growing on the roots of leguminous plants. The whole herb has been used medicinally for the treatment of irregular menstruation, cough, hemoptysis, traumatic injury and bleeding, dizziness and gastralgia in Yunnan Province, China. The 2,2‐diphenyl‐2‐picrylhydrazyl (DPPH) assay on the 60% aq. acetone extract of the fresh whole plant of B. involucrata showed considerable radical‐scavenging activity (SC50 15.3 μg/ml). Further purification on the extract led to the isolation of one new phenolic glycoside, sieboldin‐3′‐ketocarboxylic acid ( 1 ), and one new cyanogenic glycoside, proacacipetalin 6′‐Oβ‐D ‐glucopyranoside ( 2 ), together with 26 known compounds including three 4″‐O‐galloyl and 2″,3″‐O‐(S)‐hexahydroxydiphenoyl (HHDP) derivatives of dihydrochalcone glucosides, seven hydrolyzable tannins, and alkane glycosides. The cyanogenic compound isolated from the Balanophoraceae family for the first time might be a signal molecule between B. involucrata and its hosts. The free‐radical‐scavenging activity of the isolated compounds was also examined by DPPH assay.  相似文献   

17.
The high-molecular-weight glutenin subunits (HMW-GS) of wheat gluten in their native form are incorporated into an intermolecularly disulfide-linked, polymeric system that gives rise to the elasticity of wheat flour doughs. These protein subunits range in molecular weight from about 70 K-90 K and are made up of small N-terminal and C-terminal domains and a large central domain that consists of repeating sequences rich in glutamine, proline, and glycine. The cysteines involved in forming intra- and intermolecular disulfide bonds are found in, or close to, the N- and C-terminal domains. A model has been proposed in which the repeating sequence domain of the HMW-GS forms a rod-like beta-spiral with length near 50 nm and diameter near 2 nm. We have sought to examine this model by using noncontact atomic force microscopy (NCAFM) to image a hybrid HMW-GS in which the N-terminal domain of subunit Dy10 has replaced the N-terminal domain of subunit Dx5. This hybrid subunit, coded by a transgene overexpressed in transgenic wheat, has the unusual characteristic of forming, in vivo, not only polymeric forms, but also a monomer in which a single disulfide bond links the C-terminal domain to the N-terminal domain, replacing the two intermolecular disulfide bonds normally formed by the corresponding cysteine side chains. No such monomeric subunits have been observed in normal wheat lines, only polymeric forms. NCAFM of the native, unreduced 93 K monomer showed fibrils of varying lengths but a length of about 110 nm was particularly noticeable whereas the reduced form showed rod-like structures with a length of about 300 nm or greater. The 110 nm fibrils may represent the length of the disulfide-linked monomer, in which case they would not be in accord with the beta-spiral model, but would favor a more extended conformation for the polypeptide chain, possibly polyproline II.  相似文献   

18.
Aims: To characterize the kinetics of growth, sugar uptake and xylitol production in batch and fed‐batch cultures for a xylitol assimilation‐deficient strain of Candida tropicalis isolated via chemical mutagenesis. Methods and Results: Chemical mutagenesis using nitrosoguanidine led to the isolation of the xylitol‐assimilation deficient strain C. tropicalis SS2. Shake‐flask fermentations with this mutant showed a sixfold higher xylitol yield than the parent strain in medium containing 25 g l?1 glucose and 25 g l?1 xylose. With 20 g l?1 glycerol, replacing glucose for cell growth, and various concentrations of xylose, the studies indicated that the mutant strain resulted in xylitol yields from xylose close to theoretical. Under fully aerobic conditions, fed‐batch fermentation with repeated addition of glycerol and xylose resulted in 3·3 g l?1 h?1 xylitol volumetric productivity with the final concentration of 220 g l?1 and overall yield of 0·93 g g?1 xylitol. Conclusions: The xylitol assimilation‐deficient mutant isolated in this study showed the potential for high xylitol yield and volumetric productivity under aerobic conditions. In the evaluation of glycerol as an alternative low‐cost nonfermentable carbon source, high biomass and xylitol yields under aerobic conditions were achieved; however, the increase in initial xylose concentrations resulted in a reduction in biomass yield based on glycerol consumption. This may be a consequence of the role of an active transport system in the yeast requiring increasing energy for xylose uptake and possible xylitol secretion, with little or no energy available from xylose metabolism. Significance and Impact of the Study: The study confirms the advantage of using a xylitol assimilation‐deficient yeast under aerobic conditions for xylitol production with glycerol as a primary carbon source. It illustrates the potential of using the xylose stream in a biomass‐based bio‐refinery for the production of xylitol with further cost reductions resulting from using glycerol for yeast growth and energy production.  相似文献   

19.
20.
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