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.
The isolate RNP4 obtained from a long-term tannery waste contaminated soil was characterized and presumptively identified as Pseudomonas sp. The strain RNP4 tolerated concentrations up to 450 mg Cr6+/L on a Luria-Bartani (LB) agar medium and reduced a substantial amount of Cr6+ to Cr3+ in the LB liquid medium. The ability of performing multifarious activities in tandem suggested the uniqueness of isolate RNP4. The strain produced a substantial amount of indole acetic acid (IAA) in tryptophan-supplemented medium. The strain also exhibited the production of siderophore and solubilization of phosphorus in mineral salt medium and SRS1 medium, respectively. Concurrent production of IAA and siderophore and the solubilization of phosphorus revealed its plant growth promotion potential. Furthermore, the strain was able to promote the growth of black gram, Indian mustard, and pearl millet in the presence of Cr6+. Thus, the innate capability of this novel isolate for parallel bioremediation and plant growth promotion has significance in the management of environmental and agricultural problems.  相似文献   

2.
Studies were conducted to identify a 64-kD thylakoid membrane protein of unknown function. The protein was extracted from chloroplast thylakoids under low ionic strength conditions and purified to homogeneity by preparative sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Four peptides generated from the proteolytic cleavage of the wheat 64-kD protein were sequenced and found to be identical to internal sequences of the chloroplast-coupling factor (CF1) α-subunit. Antibodies for the 64-kD protein also recognized the α-subunit of CF1. Both the 64-kD protein and the 61-kD CF1 α-subunit were present in the monocots barley (Hordeum vulgare), maize (Zea mays), oat (Avena sativa), and wheat (Triticum aestivum); but the dicots pea (Pisum sativum), soybean (Glycine max Merr.), and spinach (Spinacia oleracea) contained only a single polypeptide corresponding to the CF1 α-subunit. The 64-kD protein accumulated in response to high irradiance (1000 μmol photons m−2 s−1) and declined in response to low irradiance (80 μmol photons m−2 s−1) treatments. Thus, the 64-kD protein was identified as an irradiance-dependent isoform of the CF1 α-subunit found only in monocots. Analysis of purified CF1 complexes showed that the 64-kD protein represented up to 15% of the total CF1 α-subunit.  相似文献   

3.
An α-galactosidase gene from Pyrococcus furiosus was identified, cloned and functionally expressed in Escherichia coli. It is the first α-galactosidase from a hyperthermophilic archaeon described to date. The gene encodes a unique amino acid sequence compared to other α-galactosidases. Highest homology was found with α-amylases classified in family 57 of glycoside hydrolases. The 364 amino acid protein had a calculated mass of 41.6 kDa. The recombinant α-galactosidase specifically catalyzed the hydrolysis of para-nitrophenyl-α-galactopyranoside, and to some extent that of melibiose and raffinose. The enzyme proved to be an extremely thermo-active and thermostable α-galactosidase with a temperature optimum of 115°C and a half-life time of 15 hours at 100°C. The pH optimum is between 5.0 and 5.5. Sequence analysis showed four conserved carboxylic residues. Site-directed mutagenesis was applied to identify the potential catalytic residues. Glu117Ala showed decreased enzyme activity, which could be rescued by the addition of azide or formate. It is concluded that glutamate 117 is the catalytic nucleophile, whereas the acid/base catalyst remains to be identified.  相似文献   

4.
The phylum Apicomplexa comprises obligate intracellular parasites that infect vertebrates. All invasive forms of Apicomplexa possess an apical complex, a unique assembly of organelles localized to the anterior end of the cell and involved in host cell invasion. Previously, we generated a chicken monoclonal antibody (mAb), 6D-12-G10, with specificity for an antigen located in the apical cytoskeleton of Eimeria acervulina sporozoites. This antigen was highly conserved among Apicomplexan parasites, including other Eimeria spp., Toxoplasma, Neospora, and Cryptosporidium. In the present study, we identified the apical cytoskeletal antigen of Cryptosporidium parvum (C. parvum) and further characterized this antigen in C. parvum to assess its potential as a target molecule against cryptosporidiosis. Indirect immunofluorescence demonstrated that the reactivity of 6D-12-G10 with C. parvum sporozoites was similar to those of anti-β- and anti-γ-tubulins antibodies. Immunoelectron microscopy with the 6D-12-G10 mAb detected the antigen both on the sporozoite surface and underneath the inner membrane at the apical region of zoites. The 6D-12-G10 mAb significantly inhibited in vitro host cell invasion by C. parvum. MALDI-TOF/MS and LC-MS/MS analysis of tryptic peptides revealed that the mAb 6D-12-G10 target antigen was elongation factor-1α (EF-1α). These results indicate that C. parvum EF-1α plays an essential role in mediating host cell entry by the parasite and, as such, could be a candidate vaccine antigen against cryptosporidiosis.  相似文献   

5.
Lesch–Nyhan syndrome is caused by a deficiency of hypoxanthine phosphoribosyltransferase (HPRT) encoded by HPRT1. About 20% of patients have a deletion of HPRT1 and large deletions of HPRT1 are not always fully characterized at the molecular level. Here, we report on a case of Lesch–Nyhan syndrome with a 33-kb deletion involving exon 1 of HPRT1. This novel mutation is caused by a nonhomologous recombination between different classes of interspersed repetitive DNA.  相似文献   

6.
Classic and novel protein kinase C (PKC) isozymes contain two zinc finger motifs, designated “C1a” and “C1b” domains, which constitute the recognition modules for the second messenger diacylglycerol (DAG) or the phorbol esters. However, the individual contributions of these tandem C1 domains to PKC function and, reciprocally, the influence of protein context on their function remain uncertain. In the present study, we prepared PKCδ constructs in which the individual C1a and C1b domains were deleted, swapped, or substituted for one another to explore these issues. As isolated fragments, both the δC1a and δC1b domains potently bound phorbol esters, but the binding of [3H]phorbol 12,13-dibutyrate ([3H]PDBu) by the δC1a domain depended much more on the presence of phosphatidylserine than did that of the δC1b domain. In intact PKCδ, the δC1b domain played the dominant role in [3H]PDBu binding, membrane translocation, and down-regulation. A contribution from the δC1a domain was nonetheless evident, as shown by retention of [3H]PDBu binding at reduced affinity, by increased [3H]PDBu affinity upon expression of a second δC1a domain substituting for the δC1b domain, and by loss of persistent plasma membrane translocation for PKCδ expressing only the δC1b domain, but its contribution was less than predicted from the activity of the isolated domain. Switching the position of the δC1b domain to the normal position of the δC1a domain (or vice versa) had no apparent effect on the response to phorbol esters, suggesting that the specific position of the C1 domain within PKCδ was not the primary determinant of its activity.One of the essential steps for protein kinase C (PKC)2 activation is its translocation from the cytosol to the membranes. For conventional (α, βI, βII, and γ) and novel (δ, ε, η, and θ) PKCs, this translocation is driven by interaction with the lipophilic second messenger sn-1,2-diacylglycerol (DAG), generated from phosphatidylinositol 4,5-bisphosphate upon the activation of receptor-coupled phospholipase C or indirectly from phosphatidylcholine via phospholipase D (1). A pair of zinc finger structures in the regulatory domain of the PKCs, the “C1” domains, are responsible for the recognition of the DAG signal. The DAG-C1 domain-membrane interaction is coupled to a conformational change in PKC, both causing the release of the pseudosubstrate domain from the catalytic site to activate the enzyme and triggering the translocation to the membrane (2). By regulating access to substrates, PKC translocation complements the intrinsic enzymatic specificity of PKC to determine its substrate profile.The C1 domain is a highly conserved cysteine-rich motif (∼50 amino acids), which was first identified in PKC as the interaction site for DAG or phorbol esters (3). It possesses a globular structure with a hydrophilic binding cleft at one end surrounded by hydrophobic residues. Binding of DAG or phorbol esters to the C1 domain caps the hydrophilic cleft and forms a continuous hydrophobic surface favoring the interaction or penetration of the C1 domain into the membrane (4). In addition to the novel and classic PKCs, six other families of proteins have also been identified, some of whose members possess DAG/phorbol ester-responsive C1 domains. These are the protein kinase D (5), the chimaerin (6), the munc-13 (7), the RasGRP (guanyl nucleotide exchange factors for Ras and Rap1) (8), the DAG kinase (9), and the recently characterized MRCK (myotonic dystrophy kinase-related Cdc42-binding kinase) families (10). Of these C1 domain-containing proteins, the PKCs have been studied most extensively and are important therapeutic targets (11). Among the drug candidates in clinical trials that target PKC, a number such as bryostatin 1 and PEP005 are directed at the C1 domains of PKC rather than at its catalytic site.Both the classic and novel PKCs contain in their N-terminal regulatory region tandem C1 domains, C1a and C1b, which bind DAG/phorbol ester (12). Multiple studies have sought to define the respective roles of these two C1 domains in PKC regulation, but the issue remains unclear. Initial in vitro binding measurements with conventional PKCs suggested that 1 mol of phorbol ester bound per mole of PKC (13-15). On the other hand, Stubbs et al., using a fluorescent phorbol ester analog, reported that PKCα bound two ligands per PKC (16). Further, site-directed mutagenesis of the C1a and C1b domains of intact PKCα indicated that the C1a and C1b domains played equivalent roles for membrane translocation in response to phorbol 12-myristate 13-acetate (PMA) and (-)octylindolactam V (17). Likewise, deletion studies indicated that the C1a and C1b domains of PKCγ bound PDBu equally with high potency (3, 18). Using a functional assay with PKCα expression in yeast, Shieh et al. (19) deleted individual C1 domains and reported that C1a and C1b were both functional and equivalent upon stimulation by PMA, with either deletion causing a similar reduction in potency of response, whereas for mezerein the response depended essentially on the C1a domain, with much weaker response if only the C1b domain was present. Using isolated C1 domains, Irie et al. (20) suggested that the C1a domain of PKCα but not those of PKCβ or PKCγ bound [3H]PDBu preferentially; different ligands showed a generally similar pattern but with different extents of selectivity. Using synthesized dimeric bisphorbols, Newton''s group reported (21) that, although both C1 domains of PKCβII are oriented for potential membrane interaction, only one C1 domain bound ligand in a physiological context.In the case of novel PKCs, many studies have been performed on PKCδ to study the equivalency of the twin C1 domains. The P11G point mutation of the C1a domain, which caused a 300-fold loss of binding potency in the isolated domain (22), had little effect on the phorbol ester-dependent translocation of PKCδ in NIH3T3 cells, whereas the same mutation of the C1b caused a 20-fold shift in phorbol ester potency for inducing translocation, suggesting a major role of the C1b domain for phorbol ester binding (23). A secondary role for the C1a domain was suggested, however, because mutation in the C1a domain as well as the C1b domain caused a further 7-fold shift in potency. Using the same mutations in the C1a and C1b domains, Bögi et al. (24) found that the binding selectivity for the C1a and C1b domains of PKCδ appeared to be ligand-dependent. Whereas PMA and the indole alkaloids indolactam and octylindolactam were selectively dependent on the C1b domain, selectivity was not observed for mezerein, the 12-deoxyphorbol 13-monoesters prostratin and 12-deoxyphorbol 13-phenylacetate, and the macrocyclic lactone bryostatin 1 (24). In in vitro studies using isolated C1a and C1b domains of PKCδ, Cho''s group (25) described that the two C1 domains had opposite affinities for DAG and phorbol ester; i.e. the C1a domain showed high affinity for DAG and the C1b domain showed high affinity for phorbol ester. No such difference in selectivity was observed by Irie et al. (20).PKC has emerged as a promising therapeutic target both for cancer and for other conditions, such as diabetic retinopathy or macular degeneration (26-30). Kinase inhibitors represent one promising approach for targeting PKC, and enzastaurin, an inhibitor with moderate selectivity for PKCβ relative to other PKC isoforms (but still with activity on some other non-PKC kinases) is currently in multiple clinical trials. An alternative strategy for drug development has been to target the regulatory C1 domains of PKC. Strong proof of principle for this approach is provided by multiple natural products, e.g. bryostatin 1 and PEP005, which are likewise in clinical trials and which are directed at the C1 domains. A potential advantage of this approach is the lesser number of homologous targets, <30 DAG-sensitive C1 domains compared with over 500 kinases, as well as further opportunities for specificity provided by the diversity of lipid environments, which form a half-site for ligand binding to the C1 domain. Because different PKC isoforms may induce antagonistic activities, inhibition of one isoform may be functionally equivalent to activation of an antagonistic isoform (31).Along with the benzolactams (20, 32), the DAG lactones have provided a powerful synthetic platform for manipulating ligand: C1 domain interactions (31). For example, the DAG lactone derivative 130C037 displayed marked selectivity among the recombinant C1a and C1b domains of PKCα and PKCδ as well as substantial selectivity for RasGRP relative to PKCα (33). Likewise, we have shown that a modified DAG lactone (dioxolanones) can afford an additional point of contact in ligand binding to the C1b domain of PKCδ (34). Such studies provide clear examples that ligand-C1 domain interactions can be manipulated to yield novel patterns of recognition. Further selectivity might be gained with bivalent compounds, exploiting the spacing and individual characteristics of the C1a and C1b domains (35). A better understanding of the differential roles of the two C1 domains in PKC regulation is critical for the rational development of such compounds. In this study, by molecularly manipulating the C1a or C1b domains in intact PKCδ, we find that both the C1a and C1b domains play important roles in PKCδ regulation. The C1b domain is predominant for ligand binding and for membrane translocation of the whole PKCδ molecule. The C1a domain of intact PKCδ plays only a secondary role in ligand binding but stabilizes the PKCδ molecule at the plasma membrane for downstream signaling. In addition, we show that the effect of the individual C1 domains of PKCδ does not critically depend on their position within the regulatory domain.  相似文献   

7.
Current anti-inflammatory strategies for the treatment of pulmonary disease in cystic fibrosis (CF) are limited; thus, there is continued interest in identifying additional molecular targets for therapeutic intervention. Given the emerging role of sphingolipids (SLs) in various respiratory disorders, including CF, drugs that selectively target the enzymes associated with SL metabolism are under development. Miglustat, a well-characterized iminosugar-based inhibitor of β-glucosidase 2 (GBA2), has shown promise in CF treatment because it reduces the inflammatory response to infection by P. aeruginosa and restores F508del-CFTR chloride channel activity. This study aimed to probe the molecular basis for the anti-inflammatory activity of miglustat by examining specifically the role of GBA2 following the infection of CF bronchial epithelial cells by P. aeruginosa. We also report the anti-inflammatory activity of another potent inhibitor of GBA2 activity, namely N-(5-adamantane-1-yl-methoxy)pentyl)-deoxynojirimycin (Genz-529648). In CF bronchial cells, inhibition of GBA2 by miglustat or Genz-529648 significantly reduced the induction of IL-8 mRNA levels and protein release following infection by P. aeruginosa. Hence, the present data demonstrate that the anti-inflammatory effects of miglustat and Genz-529648 are likely exerted through inhibition of GBA2.  相似文献   

8.
Aggregatibacter actinomycetemcomitans is a gram-negative opportunistic oral pathogen. It is frequently associated with subgingival biofilms of both chronic and aggressive periodontitis, and the diseased sites of the periodontium exhibit increased levels of the proinflammatory mediator interleukin (IL)-1β. Some bacterial species can alter their physiological properties as a result of sensing IL-1β. We have recently shown that this cytokine localizes to the cytoplasm of A. actinomycetemcomitans in co-cultures with organotypic gingival mucosa. However, current knowledge about the mechanism underlying bacterial IL-1β sensing is still limited. In this study, we characterized the interaction of A. actinomycetemcomitans total membrane protein with IL-1β through electrophoretic mobility shift assays. The interacting protein, which we have designated bacterial interleukin receptor I (BilRI), was identified through mass spectrometry and was found to be Pasteurellaceae specific. Based on the results obtained using protein function prediction tools, this protein localizes to the outer membrane and contains a typical lipoprotein signal sequence. All six tested biofilm cultures of clinical A. actinomycetemcomitans strains expressed the protein according to phage display-derived antibody detection. Moreover, proteinase K treatment of whole A. actinomycetemcomitans cells eliminated BilRI forms that were outer membrane specific, as determined through immunoblotting. The protein was overexpressed in Escherichia coli in both the outer membrane-associated form and a soluble cytoplasmic form. When assessed using flow cytometry, the BilRI-overexpressing E. coli cells were observed to bind 2.5 times more biotinylated-IL-1β than the control cells, as detected with avidin-FITC. Overexpression of BilRI did not cause binding of a biotinylated negative control protein. In a microplate assay, soluble BilRI bound to IL-1β, but this binding was not specific, as a control protein for IL-1β also interacted with BilRI. Our findings suggest that A. actinomycetemcomitans expresses an IL-1β-binding surface-exposed lipoprotein that may be part of the bacterial IL-1β-sensing system.  相似文献   

9.
A membrane preparation from tobacco (Nicotiana tabacum L.) cells contains at least one enzyme that is capable of transferring the methyl group from S-adenosyl-methionine (SAM) to the C6 carboxyl of homogalacturonan present in the membranes. This enzyme is named homogalacturonan-methyltransferase (HGA-MT) to distinguish it from methyltransferases that catalyze methyletherification of the pectic polysaccharides rhamnogalacturonan I or rhamnogalacturonan II. A trichloroacetic acid precipitation assay was used to measure HGA-MT activity, because published procedures to recover pectic polysaccharides via ethanol or chloroform:methanol precipitation lead to high and variable background radioactivity in the product pellet. Attempts to reduce the incorporation of the 14C-methyl group from SAM into pectin by the addition of the alternative methyl donor 5-methyltetrahydrofolate were unsuccessful, supporting the role of SAM as the authentic methyl donor for HGA-MT. The pH optimum for HGA-MT in membranes was 7.8, the apparent Michaelis constant for SAM was 38 μm, and the maximum initial velocity was 0.81 pkat mg−1 protein. At least 59% of the radiolabeled product was judged to be methylesterified homogalacturonan, based on the release of radioactivity from the product after a mild base treatment and via enzymatic hydrolysis by a purified pectin methylesterase. The released radioactivity eluted with a retention time identical to that of methanol upon fractionation over an organic acid column. Cleavage of the radiolabeled product by endopolygalacturonase into fragments that migrated as small oligomers of HGA during thin-layer chromatography, and the fact that HGA-MT activity in the membranes is stimulated by uridine 5′-diphosphate galacturonic acid, a substrate for HGA synthesis, confirms that the bulk of the product recovered from tobacco membranes incubated with SAM is methylesterified HGA.  相似文献   

10.
The selective degradation of many short一lived or abnormal Proteins in eukaryotic cells 15 carried out by the ubiquitin system.In this Pathway,Proteins are targeted for degradation by covalent ligation to ubiquitin,a highly conserved 76一resid…  相似文献   

11.
12.
Shen–Zhi–Ling (SZL) is a Chinese medicine formulated from a Kai–Xin–San decoction that is commonly used to treat depression caused by dual deficiencies in the heart and spleen. However, the underlying mechanisms remain unclear. We investigated biological changes in depression patients (DPs) exhibiting antidepressant responses to SZL treatment using proteomic techniques. We performed label-free quantitative proteomic analysis and liquid chromatography–tandem mass spectrometry to discover and examine altered proteins involved in depression and antidepressant treatment. Serum samples were collected from DPs, DPs who underwent 8 weeks of SZL treatment and healthy controls (HCs). The proteins that differed among the three groups were further validated by Western blot analysis. By performing multivariate analyses, we identified 12 potential serum biomarkers that were differentially expressed among the HC, DP, and SZL groups. We then confirmed the significant changes in alpha-1-antitrypsin, von Willebrand factors, apolipoprotein C-III, and alpha-2-macroglobulin among the three groups by performing Western blot analysis, which supported the proteomic results. Profiling the proteomic changes in DPs treated with SZL could improve our understanding of the pathways involved in SZL responses, such as alterations in platelet activation, inflammatory regulation, and lipid metabolism. Future studies involving larger patient cohorts are necessary to draw more definitive conclusions.  相似文献   

13.

Objective

Mutations associated with HIV drug resistance have been extensively characterized at the HIV-1 polymerase domain, but more studies have verified that mutations outside of the polymerase domain also results in resistance to antiviral drugs. In this study, mutations were identified in 354 patients experiencing antiretroviral therapy (ART) failure and in 97 naïve-therapy patients. Mutations whose impact on antiviral drugs was unknown were verified by phenotypic testing.

Methods

Pol sequences of HIV subtype B obtained from patients experiencing ART failure and from naïve-therapy patients were analyzed for mutations distinct between two groups. Mutations that occurred at a significantly higher frequency in the ART failure than the naïve-therapy group were submitted to the Stanford HIV Drug Resistance Database (SHDB) to analyze the correlation between HIV mutations and drug resistance. For mutations whose impact on the antiviral drug response is unknown, the site-directed mutagenesis approach was applied to construct plasmids containing the screened mutations. 50% inhibitory concentration (IC50) to AZT, EFV and NVP was measured to determine the response of the genetically constructed viruses to antiviral drugs.

Results

7 mutations at 6 positions of the RT region, D123E, V292I, K366R, T369A, T369V, A371V and I375V, occurred more frequently in the ART failure group than the naïve-therapy group. Phenotypic characterization of these HIV mutants revealed that constructed viruses with mutations A371V and T369V exhibited dual resistance to AZT and EFV respectively, whereas the other 5 mutations showed weak resistance. Although the impact of the other six mutations on response to NVP was minimal, mutation T369V could enhance resistance to NVP.

Conclusions

This study demonstrated that mutations at the RT C-terminal in subtype B′ could result in resistance to RT inhibitors if the mutations occurred alone, but that some mutations could promote susceptibility to antiviral drugs.  相似文献   

14.
15.
16.
GSH metabolism in yeast is carried out by the γ-glutamyl cycle as well as by the DUG complex. One of the last steps in the γ-glutamyl cycle is the cleavage of Cys-Gly by a peptidase to the constitutent amino acids. Saccharomyces cerevisiae extracts carry Cys-Gly dipeptidase activity, but the corresponding gene has not yet been identified. We describe the isolation and characterization of a novel Cys-Gly dipeptidase, encoded by the DUG1 gene. Dug1p had previously been identified as part of the Dug1p-Dug2p-Dug3p complex that operates as an alternate GSH degradation pathway and has also been suggested to function as a possible di- or tripeptidase based on genetic studies. We show here that Dug1p is a homodimer that can also function in a Dug2-Dug3-independent manner as a dipeptidase with high specificity for Cys-Gly and no activity toward tri- or tetrapeptides in vitro. This activity requires zinc or manganese ions. Yeast cells lacking Dug1p (dug1Δ) accumulate Cys-Gly. Unlike all other Cys-Gly peptidases, which are members of the metallopeptidase M17, M19, or M1 families, Dug1p is the first to belong to the M20A family. We also show that the Dug1p Schizosaccharomyces pombe orthologue functions as the exclusive Cys-Gly peptidase in this organism. The human orthologue CNDP2 also displays Cys-Gly peptidase activity, as seen by complementation of the dug1Δ mutant and by biochemical characterization, which revealed a high substrate specificity and affinity for Cys-Gly. The results indicate that the Dug1p family represents a novel class of Cys-Gly dipeptidases.GSH is a thiol-containing tripeptide (l-γ-glutamyl-l-cysteinyl-glycine) present in almost all eukaryotes (barring a few protozoa) and in a few prokaryotes (1). In the cell, glutathione exists in reduced (GSH) and oxidized (GSSG) forms. Its abundance (in the millimolar range), a relatively low redox potential (-240 mV), and a high stability conferred by the unusual peptidase-resistant γ-glutamyl bond are three of the properties endowing GSH with the attribute of an important cellular redox buffer. GSH also contributes to the scavenging of free radicals and peroxides, the chelation of heavy metals, such as cadmium, the detoxification of xenobiotics, the transport of amino acids, and the regulation of enzyme activities through glutathionylation and serves as a source of sulfur and nitrogen under starvation conditions (2, 3). GSH metabolism is carried out by the γ-glutamyl cycle, which coordinates its biosynthesis, transport, and degradation. The six-step cycle is schematically depicted in Fig. 1 (2).Open in a separate windowFIGURE 1.γ-Glutamyl cycle of glutathione metabolism. γ-Glutamylcysteine synthetase and GSH synthetase carry out the first two steps in glutathione biosynthesis. γ-glutamyltranspeptidase, γ-glutamylcyclotransferase, 5-oxoprolinase, and Cys-Gly dipeptidase are involved in glutathione catabolism. Activities responsible for γ-glutamylcyclotransferase and 5-oxoprolinase have not been detected in S. cerevisiae.In Saccharomyces cerevisiae, γ-glutamyl cyclotransferase and 5-oxoprolinase activities have not been detected, which has led to the suggestion of the presence of an incomplete, truncated form of the γ-glutamyl cycle (4) made of γ-glutamyl transpeptidase (γGT)4 and Cys-Gly dipeptidase and only serving a GSH catabolic function. Although γGT and Cys-Gly dipeptidase activities were detected in S. cerevisiae cell extracts, only the γGT gene (ECM38) has been identified so far. Cys-Gly dipeptidase activity has been identified in humans (5, 6), rats (710), pigs (11, 12), Escherichia coli (13, 14), and other organisms (15, 16), and most of them belong to the M17 or the M1 and M19 metallopeptidases gene families (17).S. cerevisiae has an alternative γGT-independent GSH degradation pathway (18) made of the Dug1p, Dug2p, and Dug3p proteins that function together as a complex. Dug1p also seem to carry nonspecific di- and tripeptidase activity, based on genetic studies (19).We show here that Dug1p is a highly specific Cys-Gly dipeptidase, as is its Schizosaccharomyces pombe homologue. We also show that the mammalian orthologue of DUG1, CNDP2, can complement the defective utilization of Cys-Gly as sulfur source of an S. cerevisiae strain lacking DUG1 (dug1Δ). Moreover, CNDP2 has Cys-Gly dipeptidase activity in vitro, with a strong preference for Cys-Gly over all other dipeptides tested. CNDP2 and its homologue CNDP1 are members of the metallopeptidases M20A family and have been known to carry carnosine (β-alanyl-histidine) and carnosine-like (homocarnosine and anserine) peptidase activity (20, 21). This study thus reveals that the metallopeptidase M20A family represents a novel Cys-Gly peptidase family, since only members of the M19, M1, and M17 family were known to carry this function.  相似文献   

17.

Background

Xanthomonas is a large genus of plant-associated and plant-pathogenic bacteria. Collectively, members cause diseases on over 392 plant species. Individually, they exhibit marked host- and tissue-specificity. The determinants of this specificity are unknown.

Methodology/Principal Findings

To assess potential contributions to host- and tissue-specificity, pathogenesis-associated gene clusters were compared across genomes of eight Xanthomonas strains representing vascular or non-vascular pathogens of rice, brassicas, pepper and tomato, and citrus. The gum cluster for extracellular polysaccharide is conserved except for gumN and sequences downstream. The xcs and xps clusters for type II secretion are conserved, except in the rice pathogens, in which xcs is missing. In the otherwise conserved hrp cluster, sequences flanking the core genes for type III secretion vary with respect to insertion sequence element and putative effector gene content. Variation at the rpf (regulation of pathogenicity factors) cluster is more pronounced, though genes with established functional relevance are conserved. A cluster for synthesis of lipopolysaccharide varies highly, suggesting multiple horizontal gene transfers and reassortments, but this variation does not correlate with host- or tissue-specificity. Phylogenetic trees based on amino acid alignments of gum, xps, xcs, hrp, and rpf cluster products generally reflect strain phylogeny. However, amino acid residues at four positions correlate with tissue specificity, revealing hpaA and xpsD as candidate determinants. Examination of genome sequences of xanthomonads Xylella fastidiosa and Stenotrophomonas maltophilia revealed that the hrp, gum, and xcs clusters are recent acquisitions in the Xanthomonas lineage.

Conclusions/Significance

Our results provide insight into the ancestral Xanthomonas genome and indicate that differentiation with respect to host- and tissue-specificity involved not major modifications or wholesale exchange of clusters, but subtle changes in a small number of genes or in non-coding sequences, and/or differences outside the clusters, potentially among regulatory targets or secretory substrates.  相似文献   

18.
19.
20.
Myrosinase (EC 3.2.1.147) catalyzes cleavage of glucosinolates, which consist of a thioglucoside moiety linked to amino acid-derived side chains. Myrosinase activity and expression profiles were investigated together with glucosinolate contents in Capparis ovata (caper) in order to characterize the glucosinolate–myrosinase system. The desulfoglucosinolates—glucocapparin, glucoiberin, progoitrin, epiprogoitrin, sinigrin, gluconapin, glucosinalbin, and glucobrassicin—were extracted and quantified from leaves, seeds, flowers, flower buds, and young shoots. The major desulfoglucosinolate was glucocapparin, which accumulated to values of 39.35 ± 0.09 and 25.56 ± 0.11 μmol g−1 dry weight in seed and leaf extracts, respectively. Myrosinase has high activity in caper seeds, leaves, flowers, and flower bud tissues having the highest total activities in seed extracts (79.23 ± 0.18 U). However, specific activities were the highest in flower bud extracts (200.44 ± 0.09 U mg−1 protein). The myrosinase protein migrated as a single band with a molecular weight of 65 kDa on sodium dodecyl sulfate polyacrylamide gel electrophoresis and on Western blots probed with the myrosinase-specific 3D7 antibodies. Native gel electrophoresis revealed two putative myrosinase isoenzymes in seeds, leaves, and flower tissues. The caper homolog of the Arabidopsis thaliana TGG1 gene was differentially expressed in seeds, leaves, flowers, and flower buds with the highest expression levels in leaves and flower bud tissues.  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